This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
(r,e,z) k0(t>(r, 6,z) = 0, b (r,d + 2n,z), p(r,6, z, t) + m$e(r, 0, z, t) [mSjhco] [mSjhco] (x,y,z) dh(x,y) dz dx dx dy dy = -V20(x,y,z).V2/z(x,y), (R,0) = D{6), dR (5.188a,b) A general solution to Eqs. (5.188) may be written as (Guenther and Lee, 1996, Chapter 8.2): N \ and 20 = 2*2 gives 1 / g3 \ on the LHS of the CKDFSBC above was identical to the linear velocity potential (x,z,t) 2 + dt dz 3 1,-\ , dB 2 2 e— Te Vc&(x,z,0«V |V (x,y,Z,t) (x,y,z,t)dri(x,y,t) + dt V dz dqi(t) and only measurements along the constant unit vectors Pi (t) and P2 (t) are required to determine if Ws (Ty) and WS(TY) intersect transversely (vide., Eqs. (6.197)). Wiggins (1988, Fig. 4.1.6) illustrates the intersection of Ws(Mya) and Wu(Mya) with an homoclinic plane that is spanned by N^> from Eq. (6.207) and the constant unit vectors in the Pi (?) and the P2W directions denoted as Pi (t) and Pi{t), respectively. It is not possible to sketch an analogous figure for the perturbed dissipative system a > 0 and y > 0 because the dimensionality of this system is n = l,m = 2,i = 2 (vide., Wiggins, R 2n x Mm x Tl, 1988, p. 362). Melnikov Integral (a > 0, y > 0) The distance between W\Tya(Q, 0)) and Wu(Tyoi(0,0)) at any point 3> e H may be computed from (Wiggins, 1988, (4.1.47), p. 365) 3/ 3$ — = -T7+ — / + — / - ^ - = 0 , Dt Ot dx ox ay dy dz (x) = 0, +oo [ dz
d
d 3r
iKn\(l,(r,0,z) \
= 0,
(5.71d)
mo,z,t)
exp—i(cot + v) = v '"r ' ' dt 0 <9 <2n, -h < z < 0,
(5.71e)
??(r,0,O = R e j - ^ ( r , 0 ' z ) — [ e x p - / M + v)] \ioo(j)(r,0,z) \ Re < exp —? (a)f + v) \, b
0 < # < 27r,
(5.71f)
z = 0,
P(r,0,z,O = {p(r,0,z,O} + p s (z)
(5.71g)
= Re \-p(/>(r,d,z) — [exp -i(cot + v)]\ - pgz = Re{icop(t>(r,6,z)exp—i(cot + v)} - pgz,
(5.71h)
4>(r,e,z) =
(5.7H)
(5.71J)
where ko = co2/g is the deep-water wave number, and where the dimensional Bernoulli constant Q(t) in Eq. (5.69h) has been absorbed into the scalar velocity potential <&(r,6,z,t).
210
Waves and Wave Forces on Coastal and Ocean Structures
The linearized (e = kmAj « 0 and kh = 0(1)) WMBVP in Eqs. (5.71) are prescribed in a semi-infinite annulus in the lower half space: —h
(5.72a)
where £/(•) is the Heaviside step function (Chapter 2.2.2), and where the following two boundary conditions will quantify the dimensionless coefficients a and/3: [mSj/(A/h)]X(z/h [mSj/(A/h)]x(z/h
= -l+d/h
+ Ab/h + A/h) =m Sj,
= - 1 + d/h + Ab/h) = Sb,
(5.72b) (5.72c)
that may be solved simultaneously for dimensionless coefficients a, fi to obtain a = (l-Sb/mSj),
0 = A/h-a(d/h
+ Ab/h + A/h-
1) (5.72d,e)
where mSj is the mth azimuthal mode of the dimensional wavemaker stroke in the y th degree-of-freedom that is measured at an arbitrary elevation above the wave flume bottom at z = — h + d + Ab + A in Fig. 5.11. The wavemaker shape function x (z/h) defined by Eqs. (5.72) is a generic expression that represents both piston and hinged circular cylindrical wavemakers. In order to connect with the radiated wave potential for the motions of large Lagrangian solid bodies that are reviewed in Chapter 8, the circular wavemaker displacement %{6, z, t) in Eqs. (5.68f,g) may be related to the Lagrangian motions of a circular cylindrical AM or PM wavemaker. Specifically, the horizontal translation displacement of a full-draft circular AM or
211
Wavemaker Theories
PM piston wavemaker with mSj measured at the SWL at z = 0 is given by Eqs. (5.72) when Sb = mSj,
_ "
l§1 m
a = a=d
= Ab = 0,
p = A/h = l,
Jcosm0sin(<wf + v)l \ sin(cot + v + m6) J '
(5.73a) (5. 7 3 b )
where the unit normal scalar component n\ = \e\\ = 1 is the amplitude of the direction cosine ei = ex that is illustrated in Fig. 5.11 (cf. Chapter 8.1), and where | £i | is the amplitude of the Lagrangian solid body horizontal displacement in Eq. (5.1a) for 7 = 1. Similarly, the rotational displacement of a full-draft circular cylindrical AM or PM hinged wavemaker with m Sj measured at the SWL at z = 0 and with the bottom of the circular cylindrical wavemaker fixed to the wave basin horizontal bottom at z = — h is given by Eqs. (5.72) when Sb =a = d = Ab = 0, ^0,z/h,t)
= ^-J-{z + h)nl
p = A/h = a = l,
cosm#sin(a>£ + »)i sin(a)/ + v + m9)
))1
_ , 0 , / [cosm^ sm.(a)t + v) ~ ' 5 | " 5 ' sin(eot + v + m6)
(5.73c) (5.73d) where the pseudo-direction cosine n'5 = (z + h)n\ is the pseudo-direction cosine in the ?2 direction that has the dimensions of [Length] (cf. Chapter 8.1), and where | fis | is the amplitude of the Lagrangian solid body rotational displacement in Eq. (5.1b) for j = 5. A solution to the Laplace's equation (5.71a) for
(5.74)
Substituting Eq. (5.74) into Eq. (5.71a) and then dividing by Eq. (5.74) gives d2R/dr2 R
+
(l/r)dR/dr R
+
(l/r2)d2Q/d62 (3
_ ~
d2Z/dZ2 _ 2 Z ~ ~K"'
212
Waves and Wave Forces on Coastal and Ocean Structures
where K2 is a separation constant. This equation may then separated into the following three systems of second-order ordinary differential equations (ODE's) with boundary conditions: d2Z , - ^ - K2nZ = 0,
(5.75a)
—- - k0Z = 0, z = 0, dz dZ — = 0, z = -h, dz d2& , 2 + m 2 0 = 0, d6
(5.75b) (5.75c)
0(60 = 0(6» + 2TT) periodicity
1 d ( dR\
lim
/
,
£-iKn\R
?\
(5.76a,b)
R
= 0,
(5.77b)
*/\Knr\-++oo
where ko = <*>2/g is the deep-water wave number. The inhomogeneous KWMBC Eq. (5.71e) has been omitted from Eqs. (5.77) because the full potential given by Eq. (5.70) is required at r = b because the azimuthal 9, vertical z and temporal t dependencies of an AM and a PM circular cylindrical wavemaker motions in Eqs. (5.68f,g) make it an inappropriate boundary condition for an ODE in R(r) in Eqs. (5.77). The boundary value problem for Z(z) in Eqs. (5.75) is a well-posed Sturm-Liouville problem (vide., Chapter 2.6); and, consequently, the eigenseries for Eqs. (5.75) must be computed first because Eq. (5.74) is a product solution in R(r) • @(0) • Z(z). The boundary value problem in Eqs. (5.75) is identical to the planar wavemaker boundary value problem in Eqs. (5.15d,e,f) in Sec. 2; and, correspondently, the solution may be expressed by the same orthonormal eigenseries in Eq. (5.30a) that are given by coshKnh(l +z/h) n Vn(Kn,z/h) = -± '—, « = 1,2, , (5.78a) where K\ = k — 2n/X for n = 1 and where Kn = iKn for n > 2 so that cosKnh(l + z/h) ^ _ „„, _ *n(Kn,z/h) = — '—, n>2, (5.78b) Nn
213
Wavemaker Theories
and where the dimensionless normalizing constants Nn are N2 =
-o
cosh1 Knh{\
+z/h)d(z/h)
l
2kh + sinh 2kh 4kh ' 2Knh + sin2/c„/j
n > 2 and integer
4/c„h
provided that the eigenvalues k for n computed from &o/i =
(5.78c)
1
(5.78d)
1 and K„ for n > 2 and integer are
khtanhkh, n= 1 —Knh tan Knh, n > 2,
(5.78e) (5.78f)
where &o = &>2/ is the deep-water wave number. The solutions to Eq. (5.76a) that are periodic in 6 for either an AM or a PM circular cylindrical wavemaker and that are functions of the azimuthal mode m are given by t®(0)=an
cosmn0 } exp —i
,
\ sin mO exp imd
m > 0 and integer.
(5.79a) (5.79b)
Equation (5.79a) represents an AM circular wavemaker and Eq. (5.79b) represents a PM circular wavemaker. Substituting Eqs. (5.79) into Eq. (5.74), and then substituting Eqs. (5.74,5.68f,g and 5.68h,i) into the KWMBC Eq. (5.71e); and then equating the coefficients of cos mO and sinm# requires that bm = 0 so that the solutions to Eqs. (5.76) for am = 1 are functions of the azimuthal mode number m that may be denoted as t®(fi)
= MMP)(rn9)
=
cos mO exp —imO
m > 0 and integer,
(5.79c) (5.79d)
where Eq. (5.79c) represents an AM circular cylindrical wavemaker and Eq. (5.79d) represents a PM circular cylindrical wavemaker. Solutions to the final remaining ODE Eq. (5.77a) with the mode index m = an integer constant must satisfy d2R dr2
\dR ldR r dr
/ , ,\ R (iKnr)2 + m2)-^ = 0
(5.80)
214
Waves and Wave Forces on Coastal and Ocean Structures
that is a form of Bessel equation (2.42) in Chapter 2.4.3 or Hildebrand (1976, Eq. (93), p. 146) with complex-valued solutions that are Bessel functions of the third kind or Hankel functions of the first HJ„ (Knr) and second Hm\Knr) kind that may be defined by Eqs (2.49) in Chapter 2.4.3: H£\Knr)
= Jm(Knr) + iYm{Knr),
(5.81a)
H%\Knr)
= Jm(Knr) - iYm(Knr),
(5.81b)
so that the solutions to Eq. (5.80) are a linear combination of Eqs. (5.81) that is given by R(r) = AmnH%\Knr)
+ BmnH%\Knr).
(5.82)
Asymptotically, the solutions from Eqs. (5.81) behave for large Knr ->• oo like (vide., Eqs. (2.64a,b) in Chapter 2.4.3) iw
"• W n\
H hK r)
° -
/
)
2
(
~ 17E?exp'' /
2
i
(
~ JlFEJ
exp
t
mit\
(*"• - 4 - - ) • i
t
(5 83a
' >
mit\
"' (*"'" 4 " - ) '
<5 83b)
'
Substituting Eqs. (5.83) into Eq. (5.82) and then Eq. (5.82) into the KRBC in Eq. (5.71d) requires that Bmn = 0 for n > 1. The significance of Eq. (5.81a) being a complex-valued function is that, in contrast to the added mass coefficient ixjj in Eq. (5.56a) for a planar wavemaker in Sec. 2, the added mass coefficient /JLJJ for a circular wavemaker is proportional to both the propagating and the evanescent eigenmodes. For the evanescent eigenmodes n > 2, the Hankel function of the first kind with imaginary argument Kn = iicn becomes (cf. Eq. (2.54) in Chapter 2.4.3 or Hildebrand, Eq. (97), 1976, p. 147) H$> (iKnr) = Jm (iK„r) + iYm (iKnr) = - r ( m + 1 ) K m (K„r),
(5.84)
where Km (•) is the Modified Bessel (or Kelvin) function of the second kind of order m. The Kelvin function symbol Km(«) with an argument should not be confused with the eigenvalue symbol without an argument Kn. Asymptotically for large K„r -> oo, the modified Bessel (Kelvin) function behave like (vide., Eq. (2.63b) in Chapter 2.4.3 or Abramowitz and Stegun, 1965,
Wavemaker
215
Theories
Chapter 9)
Km(/c„r) ~ /
Am2 - 1 (4m2 - l)(4m 2 - 9) 1 +— + SKnr 2!(8^„r)2
it
2Knr
exp -Knr • +
(4m2 - l)(4m 2 - 9)(4m2 - 25) 3!(8/c„r)3
+
(5.85) that are independent of the mode number m only when the first term (= unity) in the curly brackets {•} is included in the expansion by Eq. (5.85); and that satisfy the radiation condition in Eq. (5.7Id) when Kn = iicn for n > 2 in Eq. (5.71d). Substituting Eqs. (5.78a, 5.79c,d and 5.81a) into Eq. (5.74), the solution 4>(r, 0,z) = R(r)-m®(9)-Z(z) may be compactly expressed bythe following orthonormal eigenseries: oo m
= ^2cmn^n(Kn,
z/h)H%\Knr)MA{P)(m6),
(5.86a,b)
n=\
where the subscript m has been prepended to the velocity potential symbol to denote that it is a function of the azimuthal mode numm(p(r,9,z) ber m; where MA(m9) in Eq. (5.86a) represents an AM circular cylindrical wavemaker for Eq. (5.79c); where Mp(m9) in Eq. (5.86b) represents a PM circular cylindrical wavemaker for Eq. (5.79d); where ^"i = k; and where Eqs. (5.78b and 5.84) are to be substituted into Eqs. (5.86) when Kn = iKn for n > 2 and integer. Alternatively, because the vertical boundary conditions in z given by Eqs. (5.75b,c) are identical to the vertical boundary conditions in z given by Eqs. (5.71b,c), the Laplace's equation (5.68a) may be transformed from a PDE to an ODE by assuming the following orthonormal eigenseries that satisfies exactly the boundary conditions in Eqs. (5.71b,c) and the periodic 6 dependency of Eqs. (5.79a,b): oo A m4>(r,9,z) = YJ mnfmnir)^n
{Kn,z/h) MA{P)(m9).
(5.87a,b)
n=l
Substituting Eqs. (5.87) into the Laplace Eq. (5.68a) gives the following ODE for/ m „(r): /m [f,.„„ ,2 " Jmn iI«dj -2 ,„„2l 2l Jmn Jm n K 8m mnn h \{Knr) - m l —J- = 0, (5.88) dr2 + --—
216
Waves and Wave Forces on Coastal and Ocean Structures
that is of the form given by Eq. (2.42) in Chapter 2.4.3 or by Hildebrand (1976, Eq. (127), p. 152) with coefficients given by a = s = l,b — r = 0, c = —m2, and d = K2) and that has a solution that is given by Eq. (5.82). In order to evaluate the coefficients Cmn in the eigenseries in Eq. (5.86) as well as to compute the forces, moments, added mass and radiation damping coefficients for a circular wavemaker, the following dimensionless generic integral expression will be applied repeatedly (cf. Eq. (5.37) in Sec. 2 with b = a) that will replace integral calculus with algebraic substitution in computations:
/„
(a,p,a,d,Kn) -a/h
= f
[a(z/h) +
p]yn(Kn,z/h)d(z/h)
J-l+d/h
Knh I I
J sinhK n d — Kna (1
J sinhK n h
a (Knh)2Nn — 11 — - j cosh Knh + cosh Knd -\
/„
1(1 (Knh)Nn IV
h)
) sinh K„h — sinh Knd \ , i'
(5.89a)
(a,p,a,d,Kn)
-I
-a/h
[a(z/h) +
p]Vn(Kn,z/h)d(z/h)
l+d/h
K„h I I — — 1 sin K„d — K„a (1 — — 1 sin Knh a (Knh)2Nn
+ (1
P
1(1
+ (tc h)N n
J cos Knh — cos Knd
J sin Knh — sin Knd \,
n > 2 and integer, (5.89b)
n
that is dimensionless for dimensionless values of a and P (cf. Eqs. (5.9d,e) in Sec. 2).
217
Wavemaker Theories
The inhomogeneous KWMBC Eq. (5.71e) is given by the real part of: Re<
exp-i(cot + v) = or
T> Re
v-
|
^
at ,i. / ^
ndHm'(Knr)
E»=i Cm„V„(Kn,z/h)
\, I
r = b,
mSj(v
—
=
—^^zlh^
xM.A(P)(m6) exp —«'(&>/ + v) > . (5.90) The radial derivative of the Hankel function dHm (Knr)/dr in Eq. (5.90) is a dimensional derivative with the dimensions of [Length] ~ l . In order to remove the dimensions from this radial derivative and to replace it with a dimensionless derivative multiplied by a dimensional variable, define a dimensionless variable as Kn = Knr, (5.91a) l)
and a dimensionless derivative £„ (•) of the Hankel function by
d±.)
d(.)d$n d(.) \ k \ p .. — = ——— = Kn—— = . [ £„(•), d$n dr d$n {IK„\ drr
n > 2 and integer,
d{.) = -e„(-), dt,n
(5.91b) ) _. ' (5.91c) (5.9id)
so that the dimensions of the eigenseries coefficient Cmn in Eq. (5.90) may now be determined by the dimensions of the wavemaker variables because the implied dimensions of [Length] -1 of the derivatives of the Hankel functions have been replaced by a dimensionless derivative multiplied by a dimensional variable in Eqs. (5.91b,c), viz. KnX„(»). Multiplying both sides of Eq. (5.90) by a member of the set ^>i{Ki,z/h) and integrating over the dimensionless interval of orthogonality — 1 < z/h < 0 yields the following generic algebraic formula for the coefficient Cmn: mSjhcoIn(a,P,a,d,Kn) Cmn = —77— jr. , KnA Xn(Hml)(Knb))
n > 1 and integer,
(5.92a)
where the dimensionless derivatives of the Hankel function of the first kind H%\.) may be computed by adding Eq. (2.67a) and Eq. (2.68a) in
218
Waves and Wave Forces on Coastal and Ocean Structures
Chapter 2.4.3 for p = 0 (where p ^ mass density in Eq. (2.68a)) and obtaining <£n(Zm($n)) =
—
= ; r { Z m _ i ( f n ) — Z m + i (£„)},
dt,n Zmitn)
2
= JmUn), YmiU),
*«(&), #£}(fn),
(5.92b)
where Zm±\ (f„) = //^ (f„) and where the dimensionless derivative Xn (•) is defined in Eq. (5.91d); and where Eq. (5.92a) reduces to the following generic algebraic formula for the coefficient Cm\ for the propagating eigenmode for n = 1 and K\ = k: mSjhcoh(a,P,a,d,k) Cm\ =
—
*A
777
.
(5.92c)
Xi(HJnl\kb))
Substituting the radial derivative d(»)/dr of Eq. (5.84) and Eq. (5.91d) for the derivative Xn(») in the denominator on the RHS of Eq. (5.92a) gives the following generic algebraic formula for the coefficients Cmn for the evanescent eigenmodes ^>n{Kn,z/h) for n > 2, integer where Kn = iicn: -1T1 \ ' " - 2 a n d i n t e S e r - (5.92d) KnA ^i-im+Dj Xn(Km(K„b)) The following coefficient in the denominator of Eq. (5.92d): Cmn =
^.j—(m+l)
n will cancel the same coefficient in the numerator of Eq. (5.84) when Eq. (5.84) and Eq. (5.92d) are substituted for the eigenseries coefficient Cmn in Eqs. (5.86). The dimensions of Cmn in Eqs. (5.92c,d) may now readily be seen to be [Length]2/[Time] by comparison with the dimensions of the variables [mSjha>]. Substituting Eqs. (5.92c,d) into Eqs. (5.86), the real part gives «,< a
.\
r c u I D f X^ In(a,P,a,d,
Kn)
/
z\
Hm1}(Knr)
Kn—\
x MA(p)(m9) exp —i(cot + v)\,
(5.93a,b)
£n(.) = £n(H%\Knb)), where MA(>II6) in Eq. (5.93a) represents an AM circular cylindrical wavemaker for Eq. (5.79c) and Mp(m9) in Eq. (5.93b) represents a PM circular cylindrical wavemaker for Eq. (5.79d). The total potential in
219
Wavemaker Theories
Eqs. (5.93a,b) may be separated into propagating m<$p(r, 9,z, t) and evanescent m^eir, 9,z, t) eigenmodes according to m
h (a, P, a, d, k) kA
= -Re
Viilcz/h)
—
H£] (kr)
£X (H£> (kb)) + Km (K„r)
oo I„(a,P, a, d, K„) J2 H>n{Kn,z/h) »=2
K A
»
. £n
x M.A(P)(mO) exp —i(cot + v)
\Km(Knb))
(5.94a,b)
where Eq. (5.94a) represents MA(«) an AM circular cylindrical wavemaker by Eq. (5.79c); where Eq. (5.94b) represents Mp(«) a PM circular cylindrical wavemaker by Eq. (5.79d); where the algebraic notation <£,„{•) is the dimensionless derivative of (•) with respect to the dimensionless argument f„ = | ^ | fc that is defined in Eq. (5.9Id) so that all of the terms inside the curly brackets {•} in Eqs. (5.94a,b) are dimensionless; and where the eigenvalues in Eqs. (5.94a,b) are computed from ikhtanhkh, n=1 2 koh = co h/g = \ Knh \&nKnh, n > 2 and integer.
(5.94c) (5.94d)
5.3.1. Determination of Wave Amplitude from Wavemaker Motion The derivation of the circular cylindrical wavemaker gain function m Sj jmAj in Eq. (5.107) below follows the derivation for a 2D planar wavemaker in Sec. 2. The asymptotic behavior of the modified Bessel (or Kelvin) function Km (•) of the second kind of order m is a function of the order m and is given by Eq. (5.85). Figure 5.12 illustrates the dependency of Km(«) on the order m for modes m — 0,1,2,3,4, oo; and the asymptotic behavior in Eq. (5.85) for Ko(fn) for mode m = 0 only that is denoted by the symbol oo. Because the asymptotic behavior illustrated in Fig. 5.12 depends on the mode m, it is
220
Waves and Wave Forces on Coastal and Ocean Structures
10.00
Fig. 5.12. Dependency of modified Bessel function K m (f„) on order m (the symbol oo is the asymptotic form for Ko(f n ) for order m = 0).
not possible to specify a minimum distance from the wavemaker equilibrium boundary at r = b at which the evanescent eigenvalues are less than 1 % of their value at the circular wavemaker boundary as was done in Sec. 2. Instead, the determination of the wave amplitude below simply states that the wave field is computed far away from the wavemaker; and it is to be understood that the distance^ar away must be computed uniquely for each radial mode m for either AM or PM circular wavemakers. The total orthonormal eigenseries expansion in Eqs. (5.93) for the radiated velocity potential for a circular cylindrical wavemaker oscillating in the jth degree-of-freedom m<£>j(r,0,z,t) (vide., Chapter 8.1) may be separated into propagating m <£>pj (r,9,z,t) and evanescent m <$ej (r, 9,z,t) eigenmodes according to m$j(r,6,z,t)
= m<S>pj{r,0,z,t)+
m$ej(r,9,z,t)
7l(a,0, a, d, k)Vi(k, z/h)H$\kr) kA £x{H^\kb))
= [mSjhco] Re *
n=2
/ „ ( « , p, a, d, Kn) Vn(Kn, KnA
z/h)Km(Knr) Xn(Km(Knb))
x M.A(P)(md) exp —i(a>t + v) (5.95a,b)
221
Wavemaker Theories
Far away (depending on the azimuthal mode m in Eq. (5.85)) from the circular cylindrical wavemaker equilibrium boundary at r = b at a distance r = RQQ, say, the free surface profile mrij(Roo>Q>z,t) may be obtained from the mth azimuthal mode of the asymptotic form of the propagating potential m®pj(Roo,9,z,t) generated by the j'th mode of the wavemaker oscillation. The asymptotic form of the mth azimuthal mode of the propagating wave potential m<£>pj(r,9,z,t) evaluated at r —> Roo that is generated by the jth mode of wavemaker oscillation m St is given by m<&Pj(r,0,z,t)
« Re{ m 0 w -(r,0,z)exp-i(<wf + v)}r^Rx
,
(5.96a,b)
where from Eq. (5.83a) g
ty\{k,z/h)
2
(
n
mjT\
e x p ( ' r - 4 " -Y)MMP)ime). m*PJ(r,e,z) = ^ • - ^ ^ r V ^
(5.96c,d) Equation (5.96c) represents an AM circular cylindrical wavemaker from Eq. (5.79c) and Eq. (5.96d) represents a PM circular cylindrical wavemaker from Eq. (5.79d). The free surface profile mi]j (r, 0, z, t) may be computed by substituting Eqs. (5.96) into Eq. (5.71f) to obtain for r —> /?oo mr]j(r,e,t)
= mAj J ——Re \MA(P)(mO)
expi (kr - —
cot - vj J .
(5.96e,f) Average Rate of Work Done by a Circular Wavemaker on a Fluid Column For a dimensionless time x = t/T, a dimensionless time-average rate of work (m W)T (or power (mP)T) done by a circular wavemaker oscillating in the y'th degree-of-freedom (or mode of oscillation) on a fluid column is defined by (mPy)r = (mWj)T = (mFj.mqj)T,
r = b,
(5.97a,b,c)
where the hydrodynamic fluid pressure force is computed from the hydrodynamic pressure component in Eq. (5.7lg) according to d(z/h)
demPj(b,6,z/h,x)hj,
m$/r,0,z,T) = -Re{v m
,
(5.98a) (5.98b)
222
Waves and Wave Forces on Coastal and Ocean Structures
where J
mPj(r,6,z/h,r)=Re\-
—^-
,r = - ,
(5.98c,d)
and T+ l
/
(.)dx.
(5.98e)
Substituting Eqs. (5.98) into Eq. (5.97c) gives -pbh \m"j/T
—
rr+1 pr+l
/ Jz
rO
dx I
pin
d(z/h) /
J-I
dm
x
d6Re
JO
(5.99) r=b)
The negative sign in Eq. (5.99) is because the horizontal, radial unit normal nr = er points out of the circular cylindrical wavemaker boundary and into the fluid as illustrated in Fig. 5.11. The real part of the propagating components m $pj (r,0, z/h, x) from Eqs. (5.95) are i
mSjhco
Ii(a,P,a,d,k)' kA
*i(*,z/A)
£\(Jm(kb)) + £](Ym(kb))
Jm(kr)£i(Jm(kb)) -Ym(kr)£x(Ym{kb))
®pj(r,9,z/h,x) m^-pj X
+
' Jm{kr)£x{Ym{kb)) -Ym(kr)£x(Jm(kb))
cos m6 cos(2nx + v) cos(2;rr + v + mO) cos m6 sin(27TT + v) sin(2;rT + v +m8)
(5.100a,b) Equation (5.100a) represents an AM circular wavemaker by Eq. (5.79c), Eq. (5.100a) represents a PM circular wavemaker by Eq. (5.79d), and
223
Wavemaker Theories the real parts of the evanescent components m®ej{r,9,z,r)
i
of Eq. (5.100) are
„ , ^In{u,P,a,d,Kn) Km(A:„r) h , Sj hco 2 ^ — *« (*" >z/ ) KnA Xn(Km(Knb)) n=2 cosmO cos(27rr + v)
=
COS(2TTT + v +
m0)
(5.100c,d) where Eq. (5.100c) represents an AM circular wavemaker by Eq. (5.79c), where Eq. (5.100d) represents a PM circular wavemaker by Eq. (5.79d). The square of the dimensionless derivative «£ j (•) in Eqs. (5.100a,b) is defined by
(5101)
^-{mf-
where the derivatives of the Bessel functions may be computed by adding Eq. (2.67a) and Eq. (2.68a) in Chapter 2.4.3.2 and substituting into Eq. (5.92b) forZm±i(f„) = /m(f„). The dynamic component of the total pressure mpj(r,9,z/h,t) that is required for the hydrodynamic pressure force in Eq. (5.98a) may be computed by substituting Eqs. (5.94a,b) into Eq. (5.98c), then taking the real components to obtain mpj(r,6,z/h,r)
= mppj(r,0,z/h,T) =
+
P_R \dm^>pj(r,e,z/h,r)
~ T
e
\
3T
mpej(r,e,z/h,r)
dm<$>ej(r,e,z/h,T)\ 3r
J ' (5.102a,b)
where the real parts of the propagating components mpPj (r, 0, z/h, t) are from Eqs. (5.100a,b) I\(a,fi,a,d,k)
Vi(k,z/h) kA. JL\(Jm{kb)) + £\{Ym{kb)) J (kr)£i(J (kb)) m m cos mO sin(2;rr + v) iPPj(r,0,z/h,T)-+ Ym(kr)X1(Ym(kb)) sin(27TT + v + m9) (kr)£x(Ym(kb)) cos mO cos(27rr + v) -Ym{kr)£x{Jm{kb)) cos(27TT + v + m9) (5.102c,d) Equation (5.102c) represents an AM circular cylindrical wavemaker by pSjhay
224
Waves and Wave Forces on Coastal and Ocean Structures
Eq. (5.79c); Eq. (5.102d) represents a PM circular cylindrical wavemaker by Eq. (5.79d) and the real parts of the evanescent components mpej (r, 0, z, t) are from Eqs. (5.100c,d)
pSjhco2J2
I„(a,P,a,d,Kn) KnA
Vn(Kn,Z/h)
£2n(Km(Knb)) cos m6 sin(27TT + v) xKm(Knr) sin(27TT + v + m9) (5.102e,f) where Eq. (5.102e) represents an AM circular cylindrical wavemaker by Eq. (5.79c); where Eq. (5.102f) represents a PM circular cylindrical wavemaker by Eq. (5.79d) and where the square of the derivative of the Kelvin function £2(») in Eqs. (5.102e,f) may be computed by adding Eq. (2.67b) and Eq. (2.68a) in Chapter 2.4.3.2 and substituting into Eq. (5.92b). Note that in contrast to the propagating component of the dynamic component of the total pressure in the 2D planar wavemaker in Eq. (5.49g) that the propagating component of the dynamic component mppj(r,9,z/h,x) of the total pressure mpj (r,0,z/h, x) in the circular cylindrical wavemaker in Eq. (5.102d) is proportional to both a sin(») and a cos(») term in the dimensionless time x. The radial component of the horizontal velocity murj(r,9,z/h,r) in the total fluid velocity vector Eq. (5.98b) may be computed from the radial derivative of Eqs. (5.94a,b), taking the real parts, and then separating into propagating mUrpj (r,6,z/h, T) and evanescent murej (r, 0, z/h, r) components according to mpej{r,e,z/h,x)
(r,6,z/h,T)
=•
n=2
= mUrpj(r,Q,z/h,x)+
murej(r,6,z/h,r)
dm<&pj(r,0,z/h,T) dr
= Re
dm<&ej(r,8,z/h,T)] dr
(5.103a,b) where the real part of the propagating component m"rp, (r, 0, z/h, x) is
Hi)
h(ot,P,a,d,k)Vi(k,z/h) £2AJm(kb)) + £2AYm{kb))
£l(Jm(.kr))Xi(Jm(kb)) + £x{Ym{kr))£x{Ym{kb))
u„j(r,6,z/h,T)
+
£x{Jm{kr))£x{Ym{kb)) £i(Ym{kr))£i{Jm{kb))
cos m6 COS(27TT + v)
cos(27TT + v + m9) cos mO sin(2;rr + v) sin(2?rT + v + mO)
(5.103c,d)
225
Wavemaker Theories
Equation (5.103c) represents an AM circular cylindrical wavemaker by Eq. (5.79c); Eq. (5.103d) represents a PM circular cylindrical wavemaker by Eq. (5.79d) and the real part of the evanescent component murej (r, 9, z/h, r) is
iSjCo[ — I Yl irej{r,6,z/h,x)
=
In(<x,P,a,d,Kn)W„(Kn,z/h)
£n(Km(Knr))
COS md COS(27TT + V)
£n(Km(Knb))
COS(27TT + V
+m0)
(5.103e,f) The average rate of power required by a circular cylindrical wavemaker may now be computed by substituting Eqs. (5.102c-f) into Eq. (5.98a) and the radial velocity components from Eqs. (5.103e,f) into Eq. (5.97c) according to \m"j)z — {m*j1 "
mljlxt •T+l
=-bh
dx \
r=b d
(r)
dempj(b,9,z/h,r)murj(b,e,z/h,t) mppj(b,6,z/h,r)
+
mPej{b,6,z/h,x)_
[mUrpj (b, 6, z/h, T) + murej (b, 0, z/h, r ) ] .
(5.104)
At the equilibrium boundary of a circular cylindrical wavemaker at r = b, the wavemaker velocity is truncated over the total water depth by the Heaviside step functions in the wavemaker shape function in Eq. (5.72a) so that the real parts of the propagating components of the fluid velocity urpj (b, 9, z/h, r) at r = bm in Eqs. (5.103c,d) reduce to
,upj(b,9,z/h,T)
= mSjCo I — 1
h{a,P,a,d,k)^>\(k,z/h)
cosm9 cos(27TT + v) COS(27TT + v +
m9)
U(z/h + 1 - d/h) -U(z/h + a/h) _ (5.105a,b)
226
Waves and Wave Forces on Coastal and Ocean Structures
where £/(•) is the Heaviside step function (vide., Chapter 2.2.2) and the real part of the evanescent components urej(b,0,z/h,r) at r = bm in Eqs. (5.103e,f) reduce to / h \ °° r ) =mSjCo I — j ^
mUrej(b,6,z/h,
In(a,j8,a,d,/c„)*I>n(/c„,z/h)
n=2
U(z/h + 1 - d/h) -U(z/h+a/h) _
cos m9 cos(2^r + v) cos(2^r + v + mO)
(5.105c,d) so that rr+l
m Ay (b,6,z/h, T) [„a rpJ (6,0> z/fc, r) + mMre7. (b,0,z/h, r)] dr (5.105e)
= 0 by periodicity in the dimensionless time r . Similarly, {m?j)r
= -bhf
mPpj(b,9,z/h,T)murej(b,0,z,T)d(z/h)
= 0
(5.105f)
by orthogonality of the orthonormal eigenmodes ^n(Kn,z/h) in the dimensionless interval of orthogonality —1 < z//z < 0. Finally, Eq. (5.104) reduces to (mYj)z = -bh
/
/
mppj(b,9,z/h,r)murpj(b,e,z/h,r)ded(-jdr,
(5.105g) At the equilibrium position r = b, the depth integral may be taken over the total interval of orthogonality — 1 < z/h < 0 because murpj(b,9,z/h, r ) is given by Eq. (5.105a,b) at the equilibrium position r = b and mppj(b,0,z/h,t) is given by Eqs. (5.102c,d) so that Eq. (5.105g) reduces to I\{a,fi,a,d,k) k~A
pmSjh(i/ mppj(b,e,z/h,r)
=
Jm(kb)Xx{Jm{kb))
+
Ym(kb)£l(Ym(kb))
-W[Jm(kb),Ym(kb)]
*l(*,z/A)
£\{Jm{kb))
+
X\(Ym(kb))
cos m6 sin(2jrr + v) sin(27TT + v +m0) cos mO COS(2TTT + v) COS(2;TT + v +m0)
(5.105h,i) where W[Ji(kb),Yl(kb)] = 2/nkb is the Wronskian from Eq. (2.73) in Chapter 2.4.3; where <£^(») is defined in Eq. (5.101); where £„(•) is defined
227
Wavemaker Theories
inEq. (5.91d); and where the derivatives of Jm(kb), Ym(kb) zndKm(kb) may be computed by substituting Eqs. (2.67 and 2.68) from Chapter 2.4.3 into Eq. (5.92b). Substituting Eqs. (5.105a,b) and Eqs. (5.105h,i) into Eq. (5.105g) yields the following formulas for the average power generated by a circular cylindrical wavemaker: 2 u 3 ,.,3 CO ,Sjh
lP,->T =
£\{Jm{kb))
+ £.\{Ym{kb))
Ii(a,/3,a,d,k) k~A
(5.105j,k) where the average power generated by an AM circular cylindrical wavemaker in Eq. (5.105J) is one-half the power generated by a PM circular cylindrical wavemaker in Eq. (5.105k). The average power that is generated by a circular cylindrical wavemaker computed by Eq. (5.105g) implies that all of the average power generated by the wavemaker goes to generate the propagating eigenmode ^>\{k,z/h) in m<£>pj(r,9,z/h,r) and that the evanescent eigenmodes ^n(icn,z/h) in consume no average power. To verify this, the average m<&ej(r,0,z/h,T) rate of energy flux ( m E 7 ) T in a radiated linear wave may be computed from the mth azimuthal mode of the asymptotic form of the propagating potential in Eqs. (5.96a,b) generated by the jthmode ofwavemaker m®pj(Roo,0,z,t) oscillation. The average rate of energy flux in the propagating wave profile Eqs. (5.96e,f) may be computed from Eq. (5.49m) in circular cylindrical coordinates with r —>• R^ by (fflEj')r = (mFj»mqj)T;r
= 7?oo dm<&pj(r, 0, z/h,
= pbh I
dx J d ( | ) J d6Rc
x)
dx
K
dmQpj(r,G,z/h,T) dr
(5.106a,b)
where pam4>Pj(Roo,0,z/h,r) a dx
=P8
*i(*,z/A) " J *l(*,0) V ^ ^ o o /, n rmt - 2 n x — v) cos m6 cos (kr ^ 1 2 / it mix — v — mO) cos [kr 2itx V 4 2 (5.106c,d)
228
Waves and Wave Forces on Coastal and Ocean Structures
and where dm^pj(r,e,z/h,x)
£*i(*,z/A) •m™j
9r
co *i(fc,0)
cos mO cos (&r
y ^^/?oo IT
2
cos [kr
V
mn
\ 2JTT — v I
/
2TZX — v — m9 )
/-!/•->.*„ (5.106e,f) Substituting Eqs. (5.106c-f) into Eqs. (5.106a,b) yields the following formulae for the average rate of energy flux in a propagating wave when r -> RQQ: {m^j/i
Yi
4
2
gh
YmAJh
cotf(k,0)
kh
-CG
(5.106g) (5.106h)
where y = pg. The average rate of energy flux in a propagating wave generated by an AM circular cylindrical wavemaker in Eq. (5.106g) is one-half the power in the average rate of energy flux in a propagating wave generated by a PM circular cylindrical wavemaker in Eq. (5.106h); and the group velocity CQ is defined in Eq. (4.60d) in Chapter 4.5. Equating the average rate of energy flux in a propagating wave {mEj)r from Eqs. (5.106g,h) to the average power generated by a circular cylindrical watermaker ( m Pj) T from Eqs. (5.105j,k) gives the following dimensionless gain function for a circular cylindrical wavemaker mSj/mAj (cf. Eq. (5.49o)): kA\
mAj
k0h)
^X\{Jm(kb))
+
Xi(Ym(kb))
Vi(k,0)Ii(a,P,a,d,k)
(5.107)
The circular wavemaker gain function in Eq. (5.107) is valid for both AM and PM circular cylindrical wavemakers even though the average power required to generate PM spiral waves is twice the average power required to generate AM waves! The gain function Eq. (5.107) is a function of the azimuthal mode m because of the dependencies of the square of the Bessel function derivatives £\ (•) that may be computed by Eq. (5.92b); and is a function of the y'th mode of wavemaker motion because of the dependencies of the dimensionless cylindrical generic algebraic formula l\ (•) given by Eq. (5.89a). The circular cylindrical wavemaker gain function in Eq. (5.107) is a type of frequency domain transfer function that may be applied to simulate waves digitally in a wave channel/basin (vide., Chapter 9.6).
229
Wavemaker Theories
Full-Depth Circular AM or PM Piston Wavemaker For a full depth piston wavemaker, the generic wavemaker shape function variables in Eqs. (5.89) and Eqs. (5.72) are Sb = mSj,
a = a = d = Ab = 0, ft = A/h = 1,
smhkh sin Knh /i (0,1,0,0,*) = TTTTTT-.WO, l,0,0,Kn) = " , n>2, (kh)Ni (Knh)Nn so that Eq. (5.107) reduces to 2kh m 1+ £\{Jm{kb) + £\(Ym{kb)) S, ( kh \ 1 \k0hj 2 sinh 2kh m Aj kh koh
nJXi(Jm(kb))
+ £\(Ym(kb)),
(5.108a)
where n is defined in Eq. (4.60b) in Chapter 4.5; where the deep-water wave number ko = co2/g and where <£^(«) is defined in Eq. (5.101). For an azimuthal mode m = 1, the circular AM or PM piston wavemaker gain function in Eq. (5.108a) is illustrated in Fig. 5.13a for a cylinder radius to water depth ratios b/h = 0.1, 0.5, 1.0 and 5.0. The relatively high stroke values required when relatively small cylinder radius to water depth ratios b/h are applied in relatively shallow water depths where h/X < 1/25 or h/Xo < 0.01 makes a circular AM or PM piston wavemaker difficult to employ in shallow water wave experiments unless a relatively large cylinder radius to water depth ratio b/h is specified. For example, the relatively large circular wave basin shown in Fig. 5.1 at the O.H. Hinsdale-Wave Research Laboratory located at Oregon State University is a pulsating segmented circular wavemaker that has a cylinder radius to water depth ratio b/h of approximately 5/1. Figure 5.13a illustrates that the stroke requirements for this relatively large piston wavemaker cylinder radius to water depth ratio are quite reasonable. Full-Depth Circular AM or PM Hinged Wavemaker For a full depth hinged wavemaker, the generic shape function variables in Eqs. (5.89) and Eqs. (5.72) are Sb = a = d = Ab = 0, / l ( l , 1,0,0,*) = j n
1 n n
^
In(l, 1,0,0, K„) =
j5 = A/h = a = 1,
1 + (koh — l)cosh&/z (kh)2Ni
(1 - K0h)coSKnh
——y—
- I
,
n>2
Waves and Wave Forces on Coastal and Ocean Structures
230
(b) 1000.0 3
Fig. 5.13. (a) Gain function for full-draft circular AM and PM piston wavemakers for azimuthal mode m = 1. (b) Gain function for full-draft circular AM and PM hinged wavemakers for azimuthal mode m = 1.
so that Eq. (5.107) reduces to
m** j
2kh {khf 1 1+ k()h 2 sinh 2kh ikh)2 koh
sinh kh^/XfUmikb))
siahkhj £i(Jm(kb))
+
£\{Ym(kb))
1 + (koh — l)coshA;/i +
£\{Ym(kb))
1 + (koh — l)cosh£&
(5.108b)
where n is defined in Eq. (4.60b) in Chapter 4.5; where the deep-water wave number ko = a)2/g and where £?n (•) is defined in Eq. (5.101). For an azimuthal mode m = 1, the circular AM or PM hinged wavemaker gain function in Eq. (5.108b) is illustrated in Fig. 5.13b for cylinder radius to water depth ratios b/h = 0.1, 0.5, 1.0 and 5.0. Again, the relatively high stroke values required for relatively small cylinder radius to water depth ratios b/h in relatively shallow water depths where h/k < 1/25 or h/Xo < 0.01 makes a circular AM or PM hinged wavemaker difficult to employ in shallow water wave experiments unless a relatively large cylinder radius to water depth ratio b/h is specified as is illustrated in Fig. 5.13b for b/h = 5.0. Dean and Dalrymple (1991) plot the inverse of the wavemaker gain functions Eqs. (5.108) for ratios of cylinder radius to water depth b/h = 0.1, 0.5 and 1.0. The values in their Fig. 6.6 p. 183 for the inverse of Eq. (5.108a) for a circular PM piston wavemaker for azimuthal mode m = 1 appear to be
231
Wavemaker Theories
equivalent to the inverse of the values of Eq. (5.108a) if the value of the leftmost b/h contour is changed from 0.1 to 1.0; but the values in their Fig. 6.7 p. 184 for the inverse of Eq. (5.108b) for a circular PM hinged wavemaker for azimuthal mode m = 1 appear to be somewhat larger than the inverse values of Eq. (5.108b) even though their formula may be shown to be equivalent to the inverse of Eq. (5.108b).
5.3.2. Hydrodynamic Pressure Force and Moment (Added Mass and Radiation Damping) The total force/moment induced by the fluid on the vertical circular cylindrical wavemaker shown in Fig. 5.11 may be estimated by integrating the hydrodynamic pressure over the wetted surface of the wavemaker QS; i.e.,
(_"f(-J\ = - / fmP».0.z/h.-l)(J^)4S.
< 5109a >
\mM{z))
(5.109b)
J sJ
°
Vr-x»;
where r = t/T from Eq. (5.98d) and where the unit normal n to the vertical circular cylindrical wavemaker surface oS in Fig. 5.11 points out of the vertical cylinder and into the fluid and is given by n = n\e\ + «2?2 = cos#3i + sin#e2-
(5.109c)
The moment arm r' about the point of rotation that is located at z = — h + d is r' = h(z/h + 1 - d/h)h,
(5.109(1)
so that the pseudo-direction cosines (vide., Chapter 8.1) are r' x n = (z - h + rf)(—«2«i + «i?2) = n'4e\ + n'5e2 = h(z/h-
l+d/h)(-sia6ei
+cos6e2),
(5.109e)
where n'4 and n'5 are pseudo-direction cosines (vide., Chapter 8.1) and where the differential surface area of the vertical circular cylindrical wavemaker is dS = hbd6d(z/h).
(5.109f)
Force: For the generic circular cylindrical wavemaker shown in Fig. 5.11, the hydrodynamic pressure force on the wavemaker may be computed from
232
Waves and Wave Forces on Coastal and Ocean
Structures
Eq. (5.109a) by
„F(T)
= -
pbh
r-/
—a/h
mppj(b,6,z/h,r)
d(z/h)
+
•l+d/h
mPej(b,0,z/h,x) (5.110)
where mppj{») and mPej(») are defined by Eqs. (5.102c-f). Substituting Eqs. (5.102c-f) for the hydrodynamic pressure components into Eq. (5.110) gives for the mth azimuthal mode mFj(x)
= -mFjC
cos(2jrt + v) -
mFjS
sin(2^r + v)
= —m!Fj cos(27TT + v — cij),
(5.111a)
j = 1,2,
(5.111b)
where m^jC
.
COSCCj =
mFjS
sin at;
m* i
(5.111c,d)
7 = 1,2,
m?j
where the directional modes are j = 1 for e\ and j — 2 for e%. Circular AM Piston Wavemaker Substituting Eqs. (5.102c,e) for the hydrodynamic pressures into Eq. (5.110) and noting that the pressure azimuthal mode is orthogonal to the unit normal n from Eq. (5.109c) so that the integral over 6 gives 8m\Tt only for the cosm0 pressure mode, the resulting force from Eq. (5.111a) is only in the e\ x -direction and the components are given by
lf(0,P,a,d,k)
\F\c = pnbh a>
W[Mkb),Yi(kb)] £\{Ji{kb))
kA
+
£\{Yx(kb))
[\S\oo\, (5.111e)
' \Fis = pnbh
Ji(kb)Xi(Ji(kb)) + Yl(kb)Xi(Yl(kb))
lf(0,p,a,d,k)' k~A °° n=2
£\{JX{kb)) + £](Yi(kb)). lZ(0,P,a,d,Kn) Ki(y»fc) «n&
[-iSiof],
£n(Ki(Knb)) (5.111f)
where W[J\{kb),Y\(kb)} = 2/nkb is the Wronskian (Eq. (2.73) in 2 Chapter 2.4.3.4); where £ n{») is defined in Eq. (5.101) and where <£„(•) is defined in Eq. (5.9Id). There are no ?2 force components for an AM circular wavemaker because of the orthogonality of «2 = sin# and cos m9 in the interval of orthogonality 0 < 0 < 2n.
233
Wavemaker Theories
Added Mass / t n and Radiation Damping X\\ Coefficients: In an effort to connect with the radiation BVP for large Lagrangian solid bodies in Chapter 8, the added mass and radiation damping coefficients for a vertical circular cylindrical AM piston wavemaker may be obtained from Eqs. (5.111e,f) by noting first that the Lagrangian translational velocity and acceleration components for azimuthal mode m = 1 and for j = 1 in the e\ x-direction may be computed from the real parts of Eq. (5.1a) for j = 1 and for |£i| = 1S1 according to sin(27rT + v), (5.111g,h) where the overdots (•) denote ordinary temporal differentiation by the chain rule for r by Eq. (5.98d). The force components in Eq. (5.11 la) for j = m = 1 may be rewritten in terms of the Lagrangian translational velocity \X\ (T) and acceleration I X I ( T ) components in Eqs. (5.111g,h) as JXI(T)
=
I5I&;COS(2^T
+ v),
iX\(r)
lFi(r) = -(lMn)iXi(r) -
=—iSico
(iA.n)iXi(r),
(5.1 Hi)
Substituting Eqs. (5.111e,f and 5.111g,h) into Eq. (5.111a) and then equating cos(«) and sin(») terms yields iA.ii = pnbh co
lf(0,P,a,d,k)
W[J!(kb),Yi(kb)] £2AJi(kb)) + £](Yx(kb)) J\(kb)£i(Ji(kb)) + Yl(kb)Xi(Yi(kb)) £\{Mkb)) XiiYiikb))
'lf(0,p,a,d,k) lMll =
pnbh
~ n=2
I*(0,p,a,d,K„) f«A
(5-lHj)
(5.111k)
Ki(y„£>) A(Ki(«f„*))
where W[Ji(kb),Yi(kb)] = 2/nkb is the Wronskian (Eq. (2.73) in 2 Chapter 2.4.3.4); where £ n(») is defined in Eq. (5.101) and where <£„(•) is defined in Eq. (5.9Id). The dimensions of the added mass coefficient IJU-H are [Mass] and the dimensions of the radiation damping coefficient \k\\ are [Mass]/[Time]. Note that in contrast to the added mass coefficient for a 2D planar piston wavemaker given byEq.(5.56c) that both the propagating m
234
Waves and Wave Forces on Coastal and Ocean Structures
Circular PM Piston Wavemaker There are both e\ and £2 force vector components for a circular PM piston wavemaker. This difference from the AM piston wavemaker is a consequence of the azimuthal 9 dependency in the phase that yields both cos m6 and sin mO terms in the pressures and velocities of a PM piston wavemaker. Expanding the cos (•) and sin (•) terms in Eqs. (5.102d,f) by the angle sum identities in Eqs. (2.3 la and 2.32a) in Chapter 2.4.1 with +y = —imd yields the following: cos(27rr + v + m6) = cos(27TT + v)cosm# — sin(2^r + v)sinm# sin(27TT + v + m9) = sin(27rr + v)cosra# + COS(27TT + v)smm0 (5.1111) Consequently, substituting Eqs. (5.1111) and the unit normal «i from Eq. (5.109c) into Eq. (5.110) and then integrating over the azimuthal interval of orthogonality 0 < 8 < 2n yields both e\ and ?2 force components. Forces: For the direction mode j = l,the scalar force components in Eq. (5.111a) for azimuthal mode m = 1 are given by 1F2S = —prtbh a)
If(0,P,a,d,k) kA
W[Ji(kb),Yi(kb)] £\{Ji{kb))
+ XJiYiikb))
hSM, (5.111m)
JiikfyXiiMkb)) + Yi(kb)£i(Yi(kb))
lf(0,p,a,d,k)
JL\{Jx{kb)) + £\{Yx{kb)) J
\Fic = —pxbh
~
Ijj(0,p,a,d,Kn)
n=2
Kn&
hSiCO2],
Ki(«Bfe)
£a(£l(K„b))
(5.111n) where W[J\{kb),Y\{kb)} = 2/nkb is the Wronskian (Eq. (2.73) in Chapter 2.4.3.4), where £%(•) is defined in Eq. (5.101); and where £„(•) is defined in Eq. (5.91d).
235
Wavemaker Theories
Added Mass 2/^22 and Radiation Damping 2X22 Coefficients: In an effort to connect to the radiation BVP for large Lagrangian solid bodies in Chapter 8, there are both ei and ?2 added mass \pujj and radiation damping \kjj coefficients for a circular PM piston wavemaker. ei Component Coefficients: The \F\c and \F\$ force components are identical to Eqs. (5.111e,f), and the added mass \JJL\\ and radiation damping 1A.11 coefficients are identical to Eqs.(5.111j,k). ?2 Component Coefficients: The 1 FJC and 1 FJS force components may be computed from Eqs. (5.102d,f) by noting first that the Lagrangian translational velocity and acceleration components for azimuthal mode m = 1 may be computed from the real parts of Eq. (5.1a) by the temporal derivative of Eq. (5.68k) for a = 1 — St,/mS2 = 0 and A /h = 1 as 1X2(7) = -iS 2 ^ 2 sin(27T7 + v + 6), (5.111o,p) so that Eq. (5.111a) may be rewritten in terms of the Lagrangian translational velocity 1X2(7) and acceleration 1X2(7) components in Eqs. (5.111o,p) as 1X2(7) =1 52ft;cos(2^r + v + 6),
1^2(7) = -iju.22 1-^2(7) -
1A22 1X2(7),
(5-lllq)
where the overdots (•) denote ordinary temporal differentiation by the chain rule for 7 defined by (5.98d). Equating Eq. (5.111q) to Eqs. (5.11m,n) yields 1A.22 = pnbh
co
lf(0,P,a,d,k)
W[Ji(kb),Yi(kb)] X\{h(kb)) + X\{Yx(kb))
kA
(5.111r) Ji(kb)Xi(Ji(kb)) + Yl(kb)Xx(Yl(kb))
lf(0,p,a,d,k) kA 1M22 =
X](Ji(kb)) + XJiY^kb)) j
pnbh
+E
l£(0,P,a,d,Kn) KnA
Xn(±X(Knb)) (5.111s)
236
Waves and Wave Forces on Coastal and Ocean Structures
where W[Ji(kb),Yi(kb)] = 2/nkb is the Wronskian (Eq. (2.73) in Chapter 2.4.3.4); where £2n{») is defined in Eq. (5.101); and where «£„(•) is defined in Eq. (5.98d). The dimensions of the added mass coefficient 1/^22 are [Mass] and the dimensions of the radiation damping coefficient 1A22 are [Mass]/[Time]. Note that in contrast to the added mass coefficient for a 2D planar wavemaker given by Eq. (5.56c) that both the propagating m^pi and the evanescent m <$>e2 components of the velocity potential contribute to the added mass coefficient 1 \x22 for a circular AM piston wavemaker. Moments: The notation convention for the numerical subscripts for angular momentum variables are illustrated in Fig. 5.2 and defined in Chapter 8. For the generic vertical circular cylindrical wavemaker shown in Fig. 5.11, the hydrodynamic pressure moment on the wavemaker may be computed from r2n mM(r)
= -bh /
r—a/h
d0 I
Jo
d(z/h)
J-i+d/h
mPpj(b,6,z/h,r) r' x «, + mPej(b,0,z/h,t)
(5.112) where the pseudo-direction cosines f' x n are defined by Eq. (5.109e) (vide., Chapter 8.1). Substituting Eqs. (5.102c-f) for the hydrodynamic pressures and Eq. (5.109e) for the pseudo-direction cosines r' x n into Eq. (5.112) give mMj{x)
= —mMjc cos(27TT + v) — mMjs sin(27rr + v)
(5.113a)
= -mMj COS(2JTT + v — dj), j = 4,5,
(5.113b)
where the j subscript notation is illustrated in Fig. 5.2; and where mMjc
cos a 7 -= —-f-,
.
mMjs
sinay = — - j - ,
mMj
.
j =4,5.
...
(5.113c,d)
mMj
The parameters in the dimensionless equation of a straight line in Eq. (5.72) must now be modified in order to fix the point of rotation of a circular wavemaker at z = — h + d so that Sb = Ab = 0,
a = 1,
0 = 1- d/h,
(5.113e,f,g)
and the moment arm in Eq. (5.109d) may be written as r> = h[a(z/h) + j8]e3.
(5.109d')
237
Wavemaker Theories
Circular AM Hinged Wavemaker The direction cosine «2 = sin# in the e\ component of the pseudo-direction cosines in Eq. (5.109e) is orthogonal to cos mO in a circular AM hinged wavemaker in the interval of orthogonality 0 <9< 2TT . Consequently, substituting Eqs. (5.102c,e) for the hydrodynamic pressures into Eq. (5.112) will give only t?2 components for azimuthal mode m = 1 and for 7 = 5 ; and they are given by \M$c = pytbh co
lf(\,P,a,d,k)
W[h{kb),Y\(kb)] X](Ji(kb)) + £\(Yi(kb))
kh
iSico _h(\ -d/h)
(5.114a) Ji(kb)£i(J\(kb)) +Yi(kb)£l(Yl(kb))
lt(l,P,a,d,k)' kh
iXJUdkb)) + XJ(Yi(kb)) j
1M55 = pnbhA +
~
I*(l,P,a,d,Kn) K
n=2
_Xn(tl(Knb))_
nh
-iS\co-2 1 (5 114b)
Mi-d/k)y
-
where W[Ji(kb),Yi(kb)] = l/nkb is the Wronskian (Eq. (2.73) in Chapter 2.4.3), where £%(•) is defined in Eq. (5.101), and where <£„(•) is defined in Eq. (5.91d). Added Mass /tss and Radiation Damping A.55 Coefficients: In an effort to connect to the radiation BVP for large Lagrangian solid bodies in Chapter 8.1, the added mass and radiation damping coefficients for a circular AM hinged wavemaker may be obtained from Eqs. (5.114a,b) by first noting that the Lagrangian rotational velocity and acceleration components for azimuthal mode m = 1 may be computed from the real parts of Eq. (5.1b) for j = 5 by the temporal derivatives ofEq. (5.68f) with a = I, a = Sb = A& = 0 and f$ = 1 — d/h in Eqs. (5.72d,e) according to 1
05(T)
i€>5(r) = — 1Q5C02 sin(2^r + v), (5.114c,d)
= i^5
where 1^5 = ^
=
A
1S1
h{\-d/h)
,
(5.114e)
238
Waves and Wave Forces on Coastal and Ocean Structures
so that Eq. (5.113a) may be rewritten in terms of the Lagrangian rotational velocity i@5(r) and acceleration I@5(T) components in Eqs. (5.114c-e) as IM5(T) =
-/A55
I©5(T) -
A 55
I@5(T),
(5.114f)
where the overdots (•) denote ordinary temporal differentiation by the chain rule for t defined by Eq. (5.98d). Substituting Eqs. (5.114c,d)intoEq. (5.114f) and substituting Eqs. (5.114a,b and 5.114e) into Eq. (5.113a) and equating cos(») and sin(») terms yields 1A.55 = pnbh
to
'l?(l,P,a,d,k)' kh
W[Ji(kb),Yi(kb)] JC^Aikb)) + J^iYiikb)) (5.114g) Ji(kb)£i(Ji(kb)) + Yl(kb)XdYi(kb))
If(l,P,a,d,k) kh 11^55 =
.J^UiVcb)) + AWW
pnbh
+E n=2
I%{\,P,a,d,Kn) Knh
J
Ki(*„fr) £n(£l(Knb)) (5.114h)
where W[Ji(kb),Yi(kb)] = 2/nkb is the Wronskian (Eq. (2.73) in Chapter 2.4.3.4), where «C^(«) is defined in Eq. (5.101), and where £„(•) is defined in Eq. (5.91d). The dimensions of the added mass coefficient 1/X55 are [Mass»Length2] and the dimensions of the radiation damping coefficient 1 A.55 are [Mass«Length2]/[Time]. Note that in contrast to the added mass coefficient for a 2D planar piston wavemaker given by Eq. (5.56c) that both the propagating m $>pi and the evanescent m <J>ei components of the velocity potential contribute to the added mass coefficient 1 /X55 for a circular AM hinged wavemaker. Circular PM Hinged Wavemaker Expanding the phases of the trigonometric functions cos(») and sin(«) in the hydrodynamic pressures in Eqs. (5.102d,f) for a circular PM hinged wavemaker by the angle sum identities given by Eq. (2.31a) and Eq. (2.32a) with —iv = m9 in Chapter 2.4.1 and then substituting these expansions into Eq. (5.112) with the pseudo-direction cosines from Eq. (5.109e) will give both e\ and ?2 moment components for azimuthal mode m = 1 and rotational modes j =4 and 5. Expanding the phase of the trigonometric function sin(«) for the real part of the circular wavemaker displacement in Eq. (5.68k) by the angle sum identities and computing the ordinary temporal derivative
239
Wavemaker Theories
of Eq. (5.68k) gives the following wavemaker velocity for azimuthal mode m = 1 and rotational modes j' = 4 and 5: l©(T) = ie 4 (T)5i + i©5(T)22 — ) (COS(27TT + v)cos6ei — sin(2^r + v)sin0e2}. A / (5.115a,b) ~e\ Components: The e\ component of the pseudo-direction cosine in Eq. (5.109e) is n'Ae\ = -h I - - 1 + - ) sin9ei
=
_A a
[ (f)+/ 3 ] s i n 0 2 i»
(5.115c)
that has the dimensions of [Length] (vide., Chapter 8.1). The \M$c and 1M45 components of a circular PM hinged wavemaker are \M^c = pitbh co
~I*a,P,a,d,k)' kh
W[7i(ta),Yi(*fr)] JL\{h{kb)) + JL\{Yx{kb))
lS2co
h(\
(5.115d)
-d/h)]' lf(l,P,a,d,k) kh
1M4S =
pixbh
+n=2£ lS2co 2
Ji(kb)Xi(Ji(kb)) + Yi(kb)£{(Yi(kb))
J l£[(Ji(kb)) + £2i(Yi(kb))l I*(\,P,a,d,K„) Kifafr) Knh XniKdKnb))
-\
(5.115e)
where W[Ji(kb),Yi(kb)] = 2/nkb is the Wronskian (Eq. (2.73) in Chapter 2.4.3.4), where <£%(•) is defined in Eq. (5.101), and where <£«(•) is defined in Eq. (5.9Id). Added Mass 1/1,44 and Radiation Damping 1X44 Coefficients: In an effort to connect with the radiation BVP for large solid Lagrangian bodies in Chapter 8, the added mass i/i.44 and radiation damping coefficient 1A44 may be obtained from Eqs. (5.115d,e) so that Eq. (5.113a) may be rewritten in
240
Waves and Wave Forces on Coastal and Ocean Structures
terms of the Lagrangian rotational velocity i04(r) and acceleration I@4(T) components as l M 4 ( r ) = -1//44 l © 4 ( t ) - 1A.44 i@4(t),
(5.115f)
where the overdots (•) denote ordinary temporal differentiation by the chain rule for T defined by Eq. (5.98d). Equating Eq. (5.115f) to Eq. (5.113a) and substituting Eqs. (5.115d,e) yields
1^.44 = pnbh a>
lf(l,P,a,d,k) kh
W[Mkb),Yi(kb)] £\{Jx{kb)) + £\{Yx(kb)) (5.115g)
1/^44 =
pnbh
I*(l,P,a,d,k) kh
+ £ «=2
Ji(kb)Xi(Ji(kb)) + Yl(kb)Xi(Yl(kb))
l£](Mkb)) + ^(Ydkb)) J
I^(\,P,a,d,Kn) Knh
XniKiiKnb))
(5.115h) where W[Ji(kb),Yi(kb)] = 2/nkb is the Wronskian (Eq. (2.73) in Chapter 2.4.3.4), where <>C^(») is defined in Eq. (5.101), and where °C„(») is defined in Eq. (5.91d). The dimensions of the added mass coefficient 1/^44 are [Mass»Length2] and the dimensions of the radiation damping coefficient 1A.44 are [Mass»Length2]/[Time]. Note that in contrast to the added mass coefficient for a 2D planar wavemaker given by Eq. (5.56c) that both the propagating m <&p 1 and the evanescent m <£>e \ components of the velocity potential contribute to the added mass 11x44 coefficient for a circular PM hinged wavemaker. £2 Components: The £?2 component of the pseudo-direction cosine in Eq. (5.109e) is n'5e2 = h(z/h - 1 + d/h) cos#«?2
(5.115i)
Wavemaker Theories
241
that has the dimensions of [Length]. The \Msc and 1M55 components of a circular PM hinged wavemaker are
I?(l,P,a,d,k)
\Msc = pnbh co
kh
W[Jx{kb),Yx{kb)-\ £\(Ji(kb))
+
£\(Yi(kb))
iSico _h{\
\M$s = pnbh
(5.115J)
-d/h) Ji{kb)£i(Ji(kb)) + Yl(kb)Xi(Yl(kb)) _X]Ui(kb)) + £\(.Yi(.kb))_ l£(l,P,a,d,icn) Knh A (Kifo.fr))
If(l,P,a,d,k) kh
+£
l-W h{\ -d/h) (5.115k)
where W[Ji(kb),Yi(kb)] = 2/nkb is the Wronskian (Eq. (2.73) in Chapter 2.4.3.4), where £%(•) is defined in Eq. (5.101), and where <£„(•) is defined in Eq. (5.9Id). Added Mass 1^55 and Radiation Damping 1X55 Coefficients: In an effort to connect with the radiation B VP for large Lagrangian solid bodies in Chapter 8, the added mass iyu-55 and radiation damping coefficient 1A55 may be obtained from Eqs. (5.115j,k) so that Eq. (5.113a) may be rewritten in terms of the Lagrangian rotational velocity I @ S ( T ) and acceleration 16)5(1) components as lM 5 (T) = -1/X55 1© 4 (T) - lA.55 1© 4 (T),
(5.1151)
where the overdots (•) denote ordinary temporal differentiation by the chain rule for r defined by Eq. (5.98d). Equating Eq. (5.1151) to Eq. (5.113a) and substituting Eqs. (5.115j,k) yields 1A.55 = pjrbh
co
If(l,p,a,d,k) kh
W[Ji(kb),Yi(kb)] £\{Ji(kb))
+
£\(Y\{kb))
(5.115m)
242
Waves and Wave Forces on Coastal and Ocean Structures
Ji(kb)Xi(Ji(kb)) + Y1(kb)Xi(Yl(kb))
I?(l,P,a,d,k) 1/^55 =
kh
PTtbh +
IXJ(Mkb)) + XJiYiikb)) _\ °° lJ{\,P,a,d,Kn) Kl(Knb) n=2
Knh
XniK^Knb)) (5.115n)
where W[Ji(kb),Y\(kb)] = 2/jtkb is the Wronskian (Eq. (2.73) in Chapter 2.4.3.4), where <£2(.) is defined in Eq. (5.101), and where .£„(•) is defined in Eq. (5.91d). The dimensions of the added mass coefficient 1/X55 are [Mass.Length2] and the dimensions of the radiation damping coefficient 1X55 are [Mass»Length2 ] / [Time]. Note that in contrast to the added mass coefficient for a 2D planar wavemaker given by Eq. (5.56c) that both the evanescent m
5.4. Double-Actuated Wavemaker Ploeg and Funke (1980) identify TYPE E double-actuated wavemakers that include the two wavemakers that are illustrated in Figs. 5.3a,b. Doubleactuated wavemakers offer a unique opportunity to generate laboratory waves by combinations of both piston and hinged actuators. A theory for the dimensionless transfer functions (i.e. gain functions in Fig. 5.8 and Eq. (5.49o) in Sec. 2) are reviewed in Chapter 9.9. A review of operating double-actuated wavemakers by digital computers with digital-to-analog cards has been given by Hudspeth et al. (1994). These transfer functions are not unique because the ratio of the amplitudes of each actuator and the relative phase angle between the two actuators are arbitrary. Comparisons with monochromatic laboratory waves generated by the Type II 2D planar wavemaker illustrated in Fig. 5.3b for a variety of piston amplitude S to hinged amplitude S ratios measured in the wave channel at the Centres de Estudios de Puertos y Costas (CEPYC) that is one of the laboratories administered by the Centra de Estudios de Disenos y Experimentos (CEDEX) by the Ministerio de Fomento in Madrid, Spain confirmed the theory and illuminated the importance of selecting the amplitude ratios for optimum wave generation (Hudspeth et al, 1994). The theory for computing digital algorithms for each actuator in double-actuated wavemakers that is reviewed below corrects the errors in Hudspeth et al. (1994).
243
Wavemaker Theories
The nonlinear dimensional WMBVP (denoted by tildes (•)) that is defined by Eqs. (5.4) may be scaled by the following dimensionless variables (without tildes (;)): velocity and pressure:
geometry: time:
(x, z, b, d, h) — k(x, z, b, d, h),
[^ ~ (t, T) = y gk(t, T),
~~ & x = cot, co— —== y/gl
wavemaker and free-surface displacements: t ^ $ = ^ ^ >
^ * = -=>
A 5+ 5 wavemaker power and energy flux wavemaker per unit width:
where A is the dimensional amplitude of a linear wave, k = 2n/k is the dimensional propagating wave number, A, is the dimensional wavelength, g is the dimensional gravitational constant, 5(5) = dimensional amplitude of the piston (hinged) actuator for the Type II planar wavemaker in Fig. 5.3b in Sec. 2, and T = the dimensional wave period = period of wavemaker oscillation, and co = 2TT/T = radian wave frequency. The dimensionless 2D fluid motion is computed from the negative gradient of a dimensionless scalar velocity potential 4>(x, z, r) according to q(x,z,r) = u(x,z,r)ex + w(x,z,r)ez = -VQ>(x,z,r),
(5.116a)
244
Waves and Wave Forces on Coastal and Ocean Structures
where the two-dimensional vector gradient operator is V(.) = — (•)** + — (.)ez. (5.116b) ox az The dimensionless dynamic pressure p{x,z,x) may be computed from p{x,z,x)
k0h 3<J> = A/-~F—, V kh dt
x > 0,
-h < z < 0.
(5.117)
where the dimensional deep-water wave number ko = co2/g. The dimensionless scalar velocity potential <$>(x,z, r) is a solution to the following dimensionless linear WMBVP: VzcD = 0, 3
(5.118a)
x > 0, -h < z < 0, x > 0,
z
=
(5.118b)
-h,
d ]
(5.118c) 3$ 3x
Ikoh kh
S+S 3r :
x = 0, - / i < z < 0
(5.118d)
Asx —>• +ooinFig.5.3binSec.2,aradiationconditionisalsorequiredthat will admit only right progressing waves or bounded evanescent eigenmodes by Eq. (5.4f) in Sec. 2. The dimensionless free-surface elevation t](x, r) may be computed from k0hd<& T](x,x) = t / Y ~ - — , V ^ 3r
x >0, z
0
(5.118e)
Because the homogeneous boundary conditions given by Eqs. (5.118b,c) on the horizontal boundaries at z = — h and 0, respectively, are identical to the boundary conditions for the 2D planar wavemaker given by Eqs. (5.7b,c) in Sec. 2, the orthonormal eigenseries solutions to this well-posed SturmLiouville problem (vide., Chapter 2.6) are identical to the dimensionless orthonormal eigenseries given by Eqs. (5.30) in Sec. 2; viz. yn{Kn,z/h)
coshKnha+z/h) =
,
n
12
(5.119a)
245
Wavemaker Theories
where the dimensionless normalizing constant Nn is cosh2 Knh(l + z/h)d(z/h)
2h + sinh Ih Ah 2icnh + sm2icnh
=
4Z„h
1, n>2,
(5.119b,c) provided that K\ = k for n = l,K„ = ik„ for n > 2, and that the dimensional eigenvalues are computed from [ kh tanh kh,
koh =
n=1
I — icnhtanknh,
n > 2,
(5.119d) (5.119e)
where the dimensionless deep-water wave number is koh — u>2h/g. The dimensionless displacement of the generic Type II double-actuated planar wavemaker shown in Fig. 5.3b is given by Eqs. (5.9a'-e') as ? ( | , T ) - Re iB~lX (•)
(-)
( ^ ) exp -i(r
+ ( v) } ,
(5.120a)
(A)
where the notation (•) = (•)[(•)] represents a piston [hinged] actuator, where the phase angle v in Eq. (5.120a) is introduced for data analyses by finite Fourier transform (FFT) algorithms (vide., Chapter 9.2) and where B = \-b/h-d/h.
(5.120b)
The dimensionless equation of a straight line x(z/h) is given by
*(£)=[-(!)+'
U
, z h
d +1 h
<"K
(5.120c)
where U (•) is the Heaviside step function from Eq. (2.1) in Chapter 2.2.2, and where a =
P S+S
h
(5.120d,e) ,5 + 5/
The dimensionless values for b and d are positive when located inside the dimensionless still-water interval of orthogonality — 1 < z/h < 0, and are negative otherwise. A piston (hinged) only wavemaker may be recovered from
246
Waves and Wave Forces on Coastal and Ocean Structures
Eqs. (5.120) when S = 0 (S = 0) or, equivalently, when a = 0.0 (= 1.0), and a full-depth draft wavemaker when b/h = d/h = 0.0. The following dimensionless coefficients for an orthonormal eigenseries expansion of a scaled velocity potential solution to the dimensionless linear WMBVP given by Eqs (5.118) for the wavemaker shown in Fig. 5.3b may be obtained from the real part of the orthonormal eigenseries solution given by Eq. (5.38) in Sec 2; viz.
(kn, j-) exp
. Knx
(V
, (5.121a)
[n=\
()
, l&oh IS + S\ (•) / b d ~~\ - ^ \ Illa,p,-,-,kh),n V kh \ A J V h h )
Ci=iB-\-L-
()
, / kh \
k0h / S + S \ (•) /
= l (5.121b)
b d _ ~\ (5.121c)
provided that the scaled propagating eigenvalue is ^ i = k, that the evanescent eigenvalues are Kn — +ik~n for n > 2, and that the eigenvalues are computed from Eqs. (5.119d,e). The dimensionless generic algebraic formu-
(•)
las for In(a,l3,b/h,d/h, given by
-
Knh) are identical to Eqs. (5.37) in Sec. 2 and are
r-b/h
In(a,p,b,d,Kn)=
/
[a(z/h) +
P]Vn(Kn,z/h)d(z/h)
J-l+d/h a -
_\2
(Knh)
Nn
Knh | 1 - - ) sinh Knd - Knb (1 - - ) sinh Knh 1
) cosh Knh + cosh Knd h.
P
+ (Knh)Nn
1
) sinh Knh — sinh Knd \, (5.122a)
247
Wavemaker Theories
In{u,P,b,d,Kn)
-L
-b/h
[a(z/h) +
p]Vn(icn,z/h)d(z/h)
l+d/h
knh ( 1 — - J sin i
a
{Znh)2Nn
P
+ (Hnh)Nn
+ 11 1
h
I cosicnh — cosk n d
sin knh — smknd
n > 2 and integer. (5.122b) Hudspeth and Borgman (1979) demonstrate applications of dimensionless transfer functions to drive wavemakers that are controlled by digital computers through DA converter cards instead of by analog function generators. For a single-actuated wavemaker in a 2D wave channel these transfer functions may be derived by equating the average dimensionless power per unit width of the wavemaker (P)T to the average dimensionless energy flux per unit width of the propagating wave (E)T relatively far away from the wavemaker (x > 3h, say, Sec. 2) according to /<0* (5.123a) = (E) T where the average dimensionless power per unit width of the wavemaker (P) r at x = 0 may be computed from a scaled Eq. (5.49i) in Sec. 2 and the average dimensionless energy flux per unit width of the propagating wave (E) T relatively far away from the wavemaker (x > 3h, say) may be computed from a scaled Eq. (5.49m) in Sec. 2. Substituting Eqs. (5.121) into Eqs. (5.116b and 5.117), then into a scaled Eq. (5.49i) in Sec. 2 and then applying the orthogonality conditions in the scaled Eqs. (5.49j,k) in Sec. 2 yields
h-K%T
s+s
(•).
la,P,b,d,kh)
(5.123b)
Substituting a scaled Eq. (5.47) in Sec. 2 into Eq. (5.49m) in Sec. 2 yields
248
Waves and Wave Forces on Coastal and Ocean Structures
where n is defined in Eq. (4.60b) in Chapter 4.5 and where the dimensionless temporal averaging operator («) r in Eqs. (5.123a-c) is defined by 1
(•>r =
/-27T
— 27T JO
(•)dr.
(5.123d)
Equating Eq. (5.123b) to Eq. (5.123c) gives the following dimensionless transfer function: ~-|2
S+S
(5.123e) kahl\
(a,fi,b,d,kh)
that may be shown to be equivalent to the inverse of the dimensionless 2-D planar wavemaker gain function (5.49o) in Sec. 2. The dimensionless transfer functions for piston and hinged only wavemakers may be computed from Eqs. (5.123e) according to 2
= *o =
— sz —
(5.124a)
k0hlf(o,p,b,d,khY n
(5.124b)
(a,P,b,d,ich) For a generic Type II double-actuated wavemaker, the total average dimen(•)
sionless power per unit width of the wavemaker (P) T is equal to the average dimensionless energy flux per unit width of the propagating eigenmode (E) T and may be computed at x = 0 according to (•) (P) T = lj
(p + p)x(u + u)d(lyi
=(E)r:
(5.125a)
where the dimensionless dynamic pressures p and p and horizontal velocities u and u generated by the piston and hinged actuators, respectively, may be linearly decomposed into propagating and evanescent components at x = 0 according to p(z/h,t)
= pp(z/h,t)
+ pe(z/h,t),
u(z/h,t) = up(z/h,t)
+
ue(z/h,t). (5.125b,c)
249
Wavemaker Theories
Substituting Eqs. (5.121 and 5.122) into Eqs. (5.116a and 5.117) and then into Eq. (5.125a) gives the following dimensionless transfer functions for the piston and hinged actuators, respectively: S2 °0
(5.126a)
cos(i) — v) +
c2
^0
(5.126b)
l+2(|S)cos(v
VsSo Figure 5.8 in Sec. 2 illustrates that 5o «* 25o for full-draft wavemakers (b/h = d/h = 0) in water of finite depth (koh < n). When 0 = v and 5/5 ^ 1 in Eqs. (5.126), the dimensionless transfer function for the piston actuator S/A will decrease by 67% from the values shown in Fig. 5.8, while the dimensionless transfer function for the hinged actuator S/A will decrease by 33% from the values shown in Fig. 5.8 in order to generate waves having the same dimensional wave amplitudes A and wave periods f. Consequently, relatively larger actuator amplitude ratios 5/5 are required for the dimensionless transfer function in Eq. (5.126a) for the piston actuator S/A to decrease from the values illustrated in Fig. 5.8. In contrast, the dimensionless transfer function in Eq. (5.126b) for the hinged actuator S/A will vary significantly from the values illustrated in Fig. 5.8 for amplitudes ratios 5/5 that are only slightly different from unity. A calibration of Eqs. (5.126) for a variety of amplitude ratios and relative phase angles was conducted in the 40 m (L) x 6.5 m (W) x 1.5 m (D) wave channel at the Centres de Estudios de Puertos y Costas (CEPYC) that is one of the laboratories administered by the Centro de Estudios de Disefios y Experimentos (CEDEX) by the Ministerio de Fomento, in Madrid, Spain (Hudspeth et ai, 1994). The average power required by either a piston or a hinged wavemaker is identical because all of the average power generates the propagating wave in accordance with Eq. (5.123a) and is the same for both actuators. However,
Waves and Wave Forces on Coastal and Ocean Structures
250
near the wavemaker the evanescent pressure pe (and, correspondingly, the local water surface displacement rj(r)) is important and is different for each type of actuator. The amplitude of the local dimensionless dynamic evanescent pressure at the intersection of the still-water-level at z = 0 with the equilibrium location of the wavemaker at x = 0 may be computed from the real part of Eq. (5.117) according to (•) Pe
. ( • )
CoJ2CnVn(0)\
= Re
(5.127)
n=2
(•)
where C„ is defined in Eq. (5.121c). Figure 5.14 compares the ratio REV of the amplitude of the evanescent pressure for a piston wavemaker \pe\ to the amplitude of the evanescent pressure for a hinged wavemaker \pe | evaluated at the intersection of the SWL at z — 0 with the equilibrium location of the wavemaker at x = 0. Twenty evanescent eigenmodes (n = 2 1 ) were computed to compute the dimensionless evanescent pressure amplitudes. In Fig. 5.14, the piston to hinged evanescent pressure amplitude ratio RE V is very small and negative in water of finite depth koh < 1 where the piston evanescent pressure amplitude \pe\ is relatively small compared to the hinged evanescent pressure amplitude | pe |. In contrast, the hinged evanescent pressure amplitude | pe | becomes both positive and smaller than the piston evanescent pressure amplitude \pe\ in relatively deep-water k^h > 1. This difference is significant near the wavemaker, and illuminates the advantage of double-actuated wavemakers for generating bi-modal wave spectra in laboratory wave channels by applying the piston-actuator to generate the low frequency components and the hinged-actuator to generate the high frequency components (vide., Chapter 9.6 for further details). 14.191fl-
••
i
: ;;;:::
u.i• - -
0
•
"
•
! •
T •
e. D i.
IK
•
A
0
••••>••>•••<•!•>
f
i !->"H4i->
ir
i
9-
\
I-
i
K
I-
-4 J"
0.01
it
i i' T TTi IT
0.1
Kh
• W
I !
V••••!
\-rrtm 10
Fig. 5.14. Ratio of piston \pe\ to hinge \pe\ evanescent pressure amplitudes (n = 21).
251
Wavemaker Theories
5.5. Directional Wavemaker Directional wavemakers are vertically segmented wavemakers that undulate sinuously and, consequently, are also called snake wavemakers. They are applied in large rectangular wave basins like the 160 ft (L) x 87ft(W) x 6.6ft (D) tsunami wave basin shown in Fig. 5.1 at the O.H. Hinsdale-Wave Research Laboratory (OHH-WRL) located at Oregon State University that is part of the George E. Brown, Jr. Network for Earthquake Engineering Simulation (NEES). Segmented directional wavemakers may be driven either in the middle of each vertical segment or at the joint between vertical segments. Because of these two methods of generation, parasitic waves are formed along the wavemaker due to either the discontinuity of the wavemaker surface (middle segment driven) or of the derivative of the wavemaker surface (joint driven). A review of model testing facilities is given by Chakrabarti (1994, Chapter 4.0). The directional wavemaker theory that is reviewed is derived from the mild slope equation (Booij, 1978; Jonsson and Skovgaard, 1979; Radder, 1979; Smith and Sprinks, 1975; inter alios) that was applied by Dalrymple (1989). The local evanescent eigenseries in the vertical z direction (vide., Eq. (5.30) n > 2 in Sec. 2) and in the transverse y direction are neglected in applications of the mild slope equation. The origin of a Cartesian coordinate system is located at the intersection of the still water level (SWL) and the equilibrium location of a segmented wavemaker on the longitudinal centerline of the rectangular wave basin that is illustrated in Fig. 5.15. The vertical z axis is positive up from the SWL. The dimensional wave basin depth h(x,y) is assumed to vary slowly in both the wave basin longitudinal x and in the transverse y directions as
SOW)
\\.N.\\X\\N\X\\.N.XWW\.N.
IB
centerline
/ / / / S S / J J J ; ; J / ; ; ; S / / S
Fig. 5.15. Definition sketch for rectangular directional wave basin.
252
Waves and Wave Forces on Coastal and Ocean Structures
illustrated in Fig. 5.15 (Dalrymple, 1989 assumed only a longitudinal variation in x). The kinematic and dynamic wave fields may be computed from a 3-dimensional scalar velocity potential <&(x,y,z,t). The 3-dimensional fluid velocity q(x, y, z, t) may be computed from the negative gradient vector operator of the scalar velocity potential (x, y, z, t) by q{x,y,z,t)
= u(x,y,z,t)ex
+ v(x,y,z,t)ey
+
w{x,y,z,t)ez
= -V 3
(5.128a)
where the 3D gradient vector operator in Cartesian coordinates (vide., Eq. (2.10a) in Chapter 2.2.7) is V 3 (.) = -T- ex + ^7- ey + -^rez. (5.128b) dx dy dz The dimensional linear fluid dynamic pressure field p(x,y,z,t) may be computed from the dimensional linearized Bernoulli equation (4.1 If) in Chapter 4.3 by d<&(x,y,z,t) p(x,y,z,t) = p '/' '. (5.128c) at A dimensional scalar spatial velocity potential >(x, y, z) may be defined by the real part of <&(x,y,z,t) =Re{)(x,y,z)exp-i(cot
+ v)},
(5.129)
where co = 2n/T, where T is the dimensional wave period, and where the phase angle v in Eq. (5.129) is introduced for data analyses by FFT algorithms (vide., Chapter 9.2). The dimensional WMBVP for directional waves is given by VJ4>(x,y,z) = 0, d
— =0, dy
x > 0 , -B
ko4> = 0,
x > 0, -B < y < +B, z = 0,
x > 0 , y = ±B, lim x -»+oo
-h(x,y)
a dx
~ ~ \hh tanh kh, koh = \ \-KnhtanKnh,
-h(x,y)
(/>(x,y,z) = 0,
(5.130a) (5.130b) (5.130c) (5.130d)
n= \
(5.130e)
n > 2,
(5.130f)
253
Wavemaker Theories
where the deep-water wave number ko = co2/g, and where the negative sign in the KRBC in (5.130d) is selected from the KRBC Eq. (5.4f) in Sec. 2 because of the negative sign in the temporal dependency in the dimensional scalar spatial velocity potential (x, y, z, t) in Eq. (5.129). The BBC for a mildly longitudinal sloping bottom may be computed from a Stokes material surface for the bottom boundary given by W(x,y,z) DW(x,y,z) Dt
= z + h(x,y) = 0, __ d
z =
-h(x,y),
V2(«) = k £ +hy-) (•)•
(5.130e,f)
The unknown dimensional free-surface displacement rj(x,y,t) may be computed from the dimensional linearized Bernoulli equation for p(x, y, 0, t) = 0 at z = 0 according to
{
— i(o 1 $>(x,y,0,t)\,
x>0,
-B
z = 0.
(5.130g) The KWMBC may be computed from the instantaneous wavemaker displacement £(y, z, t) from its equilibrium position at x = 0 that is assumed to be strictly periodic in time with period T = liz/co according to Hy,z,t)
= Re ill^-[AU(y,a)][AU(z,b,d)]
exp-i(cot + v) , (5.130h)
where the 2D wavemaker motion U(y, z) is assumed to be given by ^ ' z > = 1 Y7T-)
r
OOxUA*o)>
(5-1301)
254
Waves and Wave Forces on Coastal and Ocean Structures
where S is the wavemaker stroke, T (y) is a generic distribution function in the transverse y direction, x(z/hxo) is given by the generic vertical shape function Eqs (5.9) in Sec. 2, hXo = h(x — 0, y) = local water depth at the wavemaker equilibrium position, and the Heaviside step functions At/(») in Eq. (2.1) from Chapter 2.2.2 are defined as AU(y,a) = U(y + a_) - U(y - a+), AU(z,b,d)
= U(z + h-d)-
U(z + b),
(5.130J) (5.130k)
where a± denote the (possibly) non-symmetric transverse orientation of the directional wavemaker in the transverse y direction in Fig. 5.15, and where the segmented wavemaker distribution in the vertical z direction is illustrated in Fig. 5.3a in Sec. 2. The linearized KWMBC (vide., Eq. (5.7e) in Sec. 2) is given by d
d%(y,z,t) dt
x = 0, —a- < y < +a+, -hXo(0,y)
(5.1301)
The CKDFSBC in Eq. (5.130b) and the BBC in Eq. (5.130c) are homogeneous so that the scalar spatial velocity potential 4>(x,y,z) in Eq. (5.129) may be expanded in the following set of orthogonormal eigenfunctions from Eqs. (5.30) in Sec. 2:
i-Y]tn(x,y)Tn(Kn,z/h), f
Tn{Kn,z/h)
=
*i(k,z/h)
,
n=1
(5.131b)
*l(M) I *„(*„, z/A);
«>2
(5.131c)
khi tanh khf, hhxo =
(5.131a)
-Knh£tanhKnhi;
n= \
(5.131d)
n>2
(5.131e)
where the deep-water wave number ko = u>2/g; and where hn(xi,yi) = local wave basin water depth. The orthonormal eigenfunctions in Eqs. (5.131) are strictly applicable only for constant depth wave basins; but, they may be
255
Wavemaker Theories
applied to wave basins with slowly varying depths if Eqs. (5.131) are interpreted as being evaluated only locally over relatively small horizontal length scales (e.g. one or two wavelengths X) where the depth may be considered to be locally equal to a constant by a Taylor series expansion of the depth (vide., Eq. (2.18) in Chapter 2.3.3), i.e.
h{xt +8x,yi
+ 8y) = ht(x =xi,y dh + dy
= yt) +
dh dx
(Sx) x=xe
(fry)+ •,-,• y=yi
hiix =xi,y
= yt) +
0(8x,8y).
Substituting Eq. (5.131a) into the 3D Laplace equation (5.130a) demonstrates that f(x,y) must be a solution to the two-dimensional Helmholtz equation (Hildebrand, 1976, Chapter. 8.6 or Morse and Feshbach, 1953, Chapter. 5.1), i.e.
viU^,y) + K^n(x,y) = o, x>o, -B
2
a
(5.l3if)
2
v 2 z (.) = dx2' dy2
(5.131g)
(•)•
However, because Dalrymple (1989) applied the mild slope equation, the mild slope equation will also be applied here to derive an algorithm for directional wave basins. The mild slope equation may now be derived from Eqs .(5.130 and 5.131). Multiplying Eq. (5.130a) by Eq. (5.131b) and integrating over the depth — h < z < 0 yields
£
Txik,z/h)V2(t>{x,y,z)dz
-f J-h
Ti(k,z/h)
\d2
j
g~2
n 2
^
x
+ V 2 0(x,y,z)
dz = 0.
(5.132a)
256
Waves and Wave Forces on Coastal and Ocean Structures
Integrating twice by parts the first term in the integrand on the RHS of Eq. (5.132a) gives
dz2
J-h
z=0
[
Ti(k, z/h)
Jz=-h
9 2 Ti(*,zA) + j-h
\Tx{k,z/h) +
(v 2 0(x,y,z).V 2 Mx,y))]
=
d(p(x,y,z) — (ktanhkh)
2=0
[TI(*,z/h) (V 2 0(JC,y,z).v 2 /*(x, y))]
(5.132b)
+ *2ATi(*.zAW(*.:y.z)
after substituting the CKDFSBC fromEq. (5.130b). Substituting Eq. (5.132b) into Eq. (5.132a) gives the following:
f
{rl(k,z/h)V2cP(x,y,z)
k2rl(k,z/h)(p(x,y,z)}dz
+
= -[rl(k,z/h)V2
=
,.
(5.132c)
\z=—h
The mild slope equation may be obtained from the terms in curly brackets {•} on the LHS of Eq. (5.132c) by first substituting Eq. (5.131a) for n = 1 only into Eq. (5.132c), then noting that
V 2 Ti(fc,z//0 =
' ' oh
'V2h{x,y),
(5.132(1)
257
Wavemaker Theories
and that Tl(k,z/h)V2(t>(x,y,z)
•^2h(x,y)
i-ri(k,z/h)V2U(x,y),ri(k,z/h)]»V2h(x,y)
= co
CO
T2(k,z/h)V^(x,y).V2h(x,y) dTi(k,z/h) + Ti(k,z/h)S(x,y) ' dh
'\V2h(x,y)\2
,
(5.132e)
In order to facilitate integrating by parts the first term in the integrand on the LHS of Eq. (5.132c), note that + ${x,y)dTx{-k,Z,h)V2h(x,y)\
V2
dh
J
fTi(*,z//i)V 2 ?(*,;y) ,„3Ti(fc, Z //i)~ y( -. +2 — V £(x,y) • V2h(x,y) 2 dh 2 (d T {k,z/h)^ x VJ
+
2\
V|/j(x,y)
dh
,
J
(5.132f,g) Substituting Eq. (5.132g) into the first term in the integrand on the LHS of Eq. (5.132c) and rearranging gives
f
Tl(k,z/h)V2$(x,y)+2
Tx(k,z/h)
xV2S(x,y).V2h(x,y)
dTi(k,z/h) dh + fc2Ti(£,z//*K(*,)>)
dz
J-h
- \r2(k,z/hm$(x,y)
• W2h(x,y)\ {d2rx{k,z/h) •\V2Kx,y)\ dh2
f
-f J-h
S(x,y)ri(k,z/h) \ +
dz.
2
Yh
V h{x,y)J\ (5.132h)
The first two terms in the integrand on the LHS and the first term on the RHS of Eq. (5.132h) may be replaced with a single integral by Leibnitz's rule (vide.,
258
Waves and Wave Forces on Coastal and Ocean Structures
Chapter 2.2.11 or Hildebrand, 1976, Eq. (92) in Chapter 7.9) according to
i:
V2«/
Tf(k,z/h)V2^x,y)dz
-h
T1(£,Z//O[T1(£)Z//OV22£(*,)0
j
dTi(k,z/h) "J \2h(x,y) dh •[-*?(*,z/A)V2?(*,;y) • V2h(x,y) +2
• V2!(x,y) dz
. (5.132i)
z=-h
Substituting Eq. (5.132i) into Eq. (5.132h) gives
V2«
f r*(k,z/h)V2i;{x,y)dz
+ k-
;
/ rf(k, z/h)$(x,y)dz J-h
J-h r Ti(fe,z/ft)C(jf,y)
dTi(k,z/h) ;,.
?(x,y)Ti(fc,z//i) ./-ft
-* , |V 2 /i(x,y)| 2
/d2r1(k1z/h)_ dh2
z=—h
\Vlh{x,y)\
<*z
V + —^—vf/K*,3oyj (5.133)
In order to proceed further in deriving the mild slope equation, the nonlinear terms on the RHS of Eq. (5.133) must first be eliminated by scaling the equation with nondimensional variables. This scaling in order to remove nonlinear product terms will, in effect, also make the BBC Eq. (5.130e) a homogeneous boundary condition and the propagating orthonormal eigenfunction in Eq. (5.131b) exact. The following nondimensional variables (denoted by tildes (i)) may be applied to scale Eq. (5.133):
(x,y) = k(x,y),
(z,h) =
(z,h) «x0
(5.134a,b,c) •<-$•
259
Wavemaker Theories
where A is a dimensional wave amplitude. Substituting Eqs (5.134) into RHS of Eq. (5.133) gives (kA)(khX0)
Tx{kh,~z/h)i;{x,y)
dTi(kh,Z/h)* dh
\V2Kx,y)\' ~z=-h
-|V 2 /i(x,y)r 2 dh dz, Ux,ym(kh,z/h) 2 8Ti(kh,z/h) V h(x,y) J-h V+ /J dh that is O(kAkhX0), where kA < 1 and khXQ — 0(1), and, consequently, both terms in curly brackets {•} in Eq. (5.132h) may be neglected. Because the RHS of the KBBC in Eq. (5.130e) is 0(kAkhXo), this boundary condition is homogenous and not inhomogeneous when terms of O (kAkhX0) are neglected. The mild slope equation may now be obtained from a homogeneous Eq. (5.133) by integrating over the local wave basin water depth where
'f-
+ (kA)(khX0)
J T2(k, z/h)d{z/h)
=h
Ni
CO
2
2gk
cosh kh
1 +
2kh sinh 2kh
CCG
(5.135)
where C = co/k is the wave celeity, CG is the group velocity (vide., Eqs. (4.60), Chapter 4.5), and the dimensional homogeneous Eq. (5.133) reduces to V 2 .(CC G V 2 f(x,v)) +k2CCGS(x,y)
= 0.
(5.136)
If the product C C G is a constant, Eq. (5.136) becomes the 2D Helmholtz equation (5.13If). Alternatively, Radder (1979) made the following scale transformation: Q = CC0, f = -|, and a dimensional homogeneous Eq. (5.133) also transforms to the following 2D Helmholtz equation: V|Z + K2Z = 0, K2
co2 n Q
^iQ 2<2
\V2QV (22)2
2kh
H =
1+ 2 sinh 2kh
260
Waves and Wave Forces on Coastal and Ocean Structures
For arbitrary non-constant wave basin depths h (x, y), Eq. (5.13 6) may only be solved numerically. However, for ID longitudinal wave basin depths h(x), there are several possibilities for solving Eq. (5.136). One possibility for the ID short-wavelength approximations is the Wentzel, Kramers, Brillouin and Jeffreys (WKBJ) method (Carrier et al., 1966, Chapter 6-7 or Morse and Feshbach, 1953, Chapter 9.3). Another possibility for ID short-wavelength approximations is the algorithm derived by Dalrymple (1989). The WKBJ approximation is reviewed because the WKBJ method is so well-documented in the physics of optics and acoustics, and because the x mode component of the Dalrymple algorithm may be shown to be equivalent to the WKBJ approximation. Because the dimensional boundary conditions defined by Eqs. (5.130c,d,h) are prescribed on boundaries that are constant values of the coordinates (x, y), a solution to the mild slope equation (5.136) may be computed by the method ofseparation of variables (Carrier and Pearson, 1968, Chapter 6; Hildebrand, 1978, Chapter 9.3; Ince, 1956, Chapter 9.41) according to ax,y)
= X{x)Y{y).
(5.137)
Substituting Eq. (5.137) into Eq. (5.136) and then dividing Eq. (5.136) by CCGX(x)Y(y) gives
CCGX(x)
Y(y)
The first term is a function of x only and the second term depends on y only. Consequently, each of these two terms must be equal to constants and the following two systems of ordinary differential equations (ODE's) result: d (
d2Y(y)
dX(x)\ CCG
dx V
dx J
CCGX(x)
2
n + &n =
dy Y(y) fj^, say, a constant.
(5.138a,b)
261
Wavemaker Theories
The two ordinary differential equations (ODE's) from Eqs. (5.138) and associated boundary conditions are for X(x) + (K2 - ix2m)C(x)CG(x)X(x)
£ • (c(x)CG(x)^^\
= 0,
(5.139a)
x =0 dXix)
-Y(y)Vn(Kn,z/h) dx
= -i-U(y,z),
\-B
X(x) = 0, and for Y(y) = Ym(y) d2Ym(y) A2 +• n-~m mYm(y) = 0, dy2
k2 > /j,2m,
-B < y < +B,
(5.139c)
(5.140a)
dYm(y)
= 0, y = ±B, (5.140b,c) dy where the explicit x dependency of the wave celerity C(x) and wave group velocity CG(X) has been added to the notation. Note that the ordinary differential equation (5.140a) and the two homogeneous boundary conditions in Eqs. (5.140b,c) for Y(y) that are prescribed at the constant boundaries y = ±B form a well-posed Sturm-Liouville problem (vide., Chapter 2.6; Morse and Feshbach, 1953, Chapter 6.3; Hildebrand, 1976, Chapter 5.7; or Benton, 1990, Chapter 6.6). Because the only non-trivial solution to the well-posed Sturm-Liouville problem in Eqs. (5.140) are eigenfunctions, the solutions to Eqs. (5.140) must be computed first. The solutions Ym(y) to Eq. (5.140a) for Mm > ° a r e Ym(y) = am cosi^my + bm sin fimy. (5.141a) Substituting Eq. (5.141a) into the two homogeneous boundary conditions Eqs. (5.140b,c) yields (Stakgold, 1979, Chapter 1.2) —am sin ixm B + bm cos ixm B = 0,
(5.141b)
am sin ixmB + bm cos ixmB = 0,
(5.141c)
DET{am,bm} = 0.
(5.141d)
where DET{», •} is the determinant of Eqs. (5.141b,c).
262
Waves and Wave Forces on Coastal and Ocean Structures
The eigenvalue solutions to Eq. (5.14Id) may be computed from sm2ixmB = 0 mn Hm = -—, m > 0 and integer. 2B Solving for bm from Eq. (5.141b) gives bm = cLm tan/i m B,
(5.141e)
m >0
so that the eigenseries from Eq. (5.141a) reduce to Y((im,y/B)=^2Amcos[fimB(y/B-l)],
m > 0 and integer, (5.141f)
m=0
provided that the eigenvalues ixm — rmt/2B for m > 0 and integer. The orthogonal eigenseries in Eq. (5.14If) may be normalized in order to obtain an orthonormal eigenseries according to cos amB(y IB — 1) ^ - ~ -, (5.141g) Mm where the non-dimensional normalizing constant Mm for /xm = mn/lB for m > 0 and integer is given by r+\ M2m= / cos2 iimB{ylB - 1) d(y/B) Em(jim,yB)
=
= 1 + Swo, where, 8mo is the Kronecker delta function (vide., Eq. (2.2) in Chapter 2.2.3). The orthonormal eigenfunction in Eq. (5.141g) in the nondimensional interval of orthogonality —1 < y/B < 1 reduce Eq. (5.14If) to ^ f c o ls [ 2 i ( y / B - l )ll ] Y(jim, y/B) = J^ Ami \ " ' = J^Am
S m (/x m , y/B),
(5.141h)
m=0
provided that the eigenvalues fim — mn/lB for m > 0 and integer. Comparing the orthonormal eigenseries in Eq. (5.141h) with Dalrymples's algorithm (1989, Eq. (7), Sec. 2.1) illuminates one of the differences between these two ID short-wavelength approximations. Dalrymple obtained two sets of eigenvalues for each of the two transverse modes in y instead of the single
263
Wavemaker Theories
set in Eq. (5.141e). The transverse y mode orthonormal eigenseries from Eq. (5.141h) and the single set of eigenvalues in Eq. (5.141e) /xm = mn/lB for m > 0 and integer are applied in this WKBJ approximation. The WKBJ approximation follows Carrier et al. (1966, Chapter 6-7). For Qmn = y/K-l - / i ^ , where K2 = k2 > \x2m is required for the propagating eigenmode only, the solution to Eq. (5.139a) is posited for X(x) = Xmn (x) as (Carrier et al. (1966, Chapter 6-7)): Qmnd%, k2 > ix2m.
Xmn(x) = Dmn(x) exp i
(5.142)
The restriction on the relative magnitudes of the eigenvalues that k2 > p?m for the propagating eigenmode requires that the transverse eigenseries in 3 m (/x m ,y/B) be truncated at a cut-off frequency \XM at some upper limit, M say. Substituting Eq. (5.142) into Eq. (5.139a) and separating the real and imaginary parts, respectively, yields the following two equations (Carrier et al., 1966, Chapter 6-7): d2Dmn(x) Z
•,
(dDmn(x)/dx)2
9
2
+ Qmn<\ - Q mn)Dmn(x) = l"
ax
^
'
,
(5.143a)
umn\X)
.n , ^dDmn(x) ld(C(x)CG(x)) fciAiu\ C(x)CG(x)— hDmn(x) = 0, (5.143b) ax 2 ax where Eq. (5.143b) has been substituted in order to eliminate the C(X)CG(X) terms in Eq. (5.143a). The coefficient Dmn (x) may be computed by integrating the first-order ODE in Eq. (5.143b) and obtaining C(x = 0)CG(x = 0) Dmn(.x) = \r t \ ' (5.143c) n( n(
y
C(x)CG(x)
where C(x = 0)CG(X — 0) is the constant of integration that is evaluated at the equilibrium position of the wavemaker at x = 0. For a horizontal wave basin bottom, h(x) is a constant and Eq. (5.143c) is equal to unity. Substituting the transverse y mode eigenseries Eq. (5.141h) and the WKBJ solution Eq. (5.142) with Eq. (5.143c) into Eq. (5.137) gives
A , S(x,y) = Z^A™J m=0
(rx
lc(x = o)cG(x = 0) C(x)C
(x)
s
m(Mm,.y/fl)exp; I /
\ Qmnd$\
'
Qmn = jKl - / 4
and
K2 =k2>fJ2mf
m
<M
(5 M 4 )
264
Waves and Wave Forces on Coastal and Ocean Structures
The WKBJ x mode solution in Eq. (5.142) may now be compared to the x mode solution in the Dalrymple algorithm (Dalrymple, 1989, Eq. (14), Sec. 2.1). Dalrymple assumed that the x component of the propagating wave potential solution to the homogeneous second-order mild slope equation that is propagating in the +x direction may be determined from the solution to the following inhomogeneous first-order ODE: dX+ - ijk2 - a2mX+ = Fix), dx where am is a member of one of the two sets of transverse y eigenvalues computed by Dalrymple (1989, Sec. 2.1), and where the forcing function F(x) is to be determined by substituting this assumed solution for X+ into the second-order mild slope equation (Dalrymple, 1989, Eq. (8), Sec. 2.1). The particular solution for X+ may also be computed for a specified F(x) forcing from an integrating factor pm(x) (vide., Chapter 2.5.2) given by
'if
pm (x) = exp - / ( /
Jk2 - a^ d%
The solution for X+ for the inhomogeneous solution from the integrating factor pm(x) is 4. 1 fx D(x) + X = — - / pmG)F($)d$ = Pm(x) Pm(x) J = D(x)expi(j Jk2 - a2md^\, that demonstrates the equivalence between the Dalrymple x mode algorithm (1989) and the WKBJ approximation for the propagating eigenmode. The wavemaker velocity potential 3> (x, y, z, t) is now given by substituting Eq. (5.144) and Eq. (5.131a) into Eq. (5.129) and obtaining <$(x,y,z,t) - Re {(/)(x,y, z) exp - i(cot + v)} = Re i ^ J ] f (*,y)T„(tf B ,z//0 e x p - i{a>t + v) I n=\ i = Re
^
E
E
m=0n=l
A
mnJ
xTn(Kn,z/h)expi
V
nl
.„
, ,
C(x)CG(x)
am{llm,y/B)
(Jx Qmnd%) exp —i(cot -i(cot ++ v) (5.145)
265
Wavemaker Theories
The unknown coefficients Amn may now be computed in a best leastsquares sense from the inhomogeneous kinematic wavemaker boundary condition Eq. (5.139b) according to
dXn(x) Ym(y/B)Vn(Kn,z/h) dx
koS = -i r(y)x(z/he), (A/hxo)
x=0 -B < y < +B -hxo(0,y)
where A is defined in Fig. 5.3a in Sec. 2. Substituting Eq. (5.142) for Xmn(x) with Dmn(0) equal to unity, Eq. (5.141h) for Y(/xm,y/B) and Eq. (5.131c) for ^n{Kn,z/h) into the LHS of Eq. (5.146a) and neglecting dDmn(x)/dx reduces Eq. (5.146a) to M
^2^2AmnQmnEm(iim,y/B)tyn(Kn,z/h)
= -
m=0n=l
kS r(y)x(z/hX0); (A//z*0)
x =0 -B
-Vn(Kn,0)In(a,IS,b,d,k)
(A/hxo) r+a+/B
x / J-a-/B
r(y/B)Em(iim,y/B)d(y/B),
(5.146c)
266
Waves and Wave Forces on Coastal and Ocean Structures
where /„(•) is defined in Eqs. (5.37) in Sec. 2 where h = hxo, where the Heaviside step functions £/(•) in Eqs. (5.130j,k) have changed the limits of integration for both y and z in Eq. (5.146c), and where the dimensions of Amn = [Length]. Full-Draft Piston Snake Wavemaker The dimensionless transverse y component of the snake displacement for a full-draft (b = d = 0) piston (a = 0and/3 = 1) wavemaker may be expressed by the following Fourier expansion in the y coordinate:
r(qj,y/B)=
J^ j=
Cjexpi[qjB(y/B
+ Vj)],
(5.147)
-OQ
where qj is the transverse wave number that may also be a member of the transverse eigenvalue set [x,m and where Vj is an arbitrary phase angle for each transverse component qj. Dalrymple (1989, Eq. (20), Sec. 2.2) assumed that the transverse y displacement of a paddle wavemaker is given by the real part of exp/(A.oy — cot), where A.o = ksinO that is also the symbol for a member of one of the two sets of his transverse eigenvalues, viz. kn = nit IB, and it is not clear if both eigenvalues given by the same symbol A.o are equivalent. Because the transverse wave number qj in Eq. (5.147) may or may not also be a member of the transverse eigenvalue set \±m, two separate symbols are applied in order to avoid a possible confusion between qj and the transverse y eigenvalue /im if they are not equal. For a full-draft piston snake wavemaker, substituting Eq. (5.147) into the dimensionless integral on the RHS in Eq. (5.146c) yields r+a+/B cmj = / J-a^/B
r{qj,y/B)am{nm,y/B)d{y/B)
^a+,a-{j,m)
+i la+, a2
mB(q]-Li m)(l+8mo)
(P8WS) ff-^8-
(vfb)ms [Y—) — 1 soo (f<\gfb) soo ( — ) soo (— ) soo (».^)uis (fag fb) ms ( ) ui LV v) uiui ' v ' ' \uut/ ' \(fL) _ L ] uis (vfb)soo (fagfb) soo ( — ) soo LV 27/ i « « J
'
V
7
(w'f)v'Dy
= i£l«l7
Vi£M//
J
0} qzluul —t"r' J0J aonpsa (o'q8f I'S) 's^a usq; '» = + » = - » jBqj os o = ^ = * suijjajuao aq; moqB msBq SABAY aq; ui X^BOUJSUIUIXS pajuauo si J35[BUI9ABA\ 95[BUS aqjji
(38frrS) [ [ ( • ^ - ^ ) f f ^ ] ™ [ ( ^ + ^)ff- r *]™-* [ ( • ' * - 4 ) a-'*] soo [ ( A + i ) ^ ] soo
[0 + £)i&i] SO3 ff f "H-
[ [ ( • ' « - ^ ) f f - f * ] ^ [ ( ^ + £)ff-'*]«>°[0 + ^)lL] , ™*" l Z+ )g/'*]ms [(f r t -A) f f .^]soo[(^ [('—•%)« f »]«»[(i-i) J s o o f f / V n
[('«-^)«'»M('-^)
uiui
ms null— (iu'f)
~V '+£> T
9I9qA\ pUB
(qstrs) [(fc _ i ) f f f t ] soo [(A + £ ) * ' * ] n i s [(•'«- i ) f f . ^ ] m s [ ( A + £)ff.'*]soo ) L ^ + [{U-45)gfb]u]s[(4-D + -a-v)afi>]u]S+)
+
°>
^ ™\™a»P+
{ [(U - i ) ff ft] soo [ ( * + i ) ff ft] soo ) U I + + J ^ ™ [(fa - i ) gfb]ms [(X - A ) Jjj] s o o f f . ^ -
UWZ+
(u/'.O -0
qi/iLiu Z.9Z
fD
a
= wrl JOJ 3J9qA\ SBUOBifj a3yvui3aVjH
268
Waves and Wave Forces on Coastal and Ocean Structures
and to Am n Ia
/mn\ . , , r cos I —- I cos(qjd) sin [qjBvj) sin LI 2 + sin V T / s i n ( ^ ' a ) c o s VHBVJ) c o s ["2" ( ^ ) \ /mjt\ rmn /a\-\ . . vmii
*KjB \
"" Q] I
s
IT/cosIT L T 1B J JSln w B y i) sm(^-fl)
. /mn\
+ Sm
. rmn /a \n
VTJ Sln LT \B)\
C0S
,
N
Wfi^') cos (^ a )
(5.148e) Finally, if the snake wavemaker spans the entire width of the wave basin so that a± — ±B, then (5.148a) rrnt/2B+to1)] sinEq. [qjBivj - reduces 1)] + ( -for l ) m/zsin m =[qjB(Vj 4qjB m -i {sin [^fi(v7- - 1)] + ( - l ) cos [^-B(v; + !)]}_ c
mj —
([qjB-(m7T)2)2)(l+Sm0) (5.148f)
5.6. Sloshing Waves in a 2D Wave Channel A long rectangular wave channel with a horizontal flat bottom, two rigid vertical side walls and a wavemaker may generate either 2D, long-crested progressive waves that dissipate on a sloping terminal beach or two types of transverse waves are possible, viz. (i) Sloshing waves that are excited directly in wave channels by transverse motion of the wavemaker (Barnard et al., 1977, Kit et al., 1987, and Shemer and Kit, 1988), and (ii) Cross waves that are excited parametrically in wave channels by the progressive waves at a subharmonic of the wavemaker frequency and are reviewed in Chapter 6.8 (Bowline et al., 1999 and Hudspeth et al, 2005). The linearized dimensional WMBVP for 3D sloshing waves in a wave channel is identical to Eqs. (5.7) for planar 2D wavemakers except for the kinematic wavemaker boundary condition (KWMBC) in Eq. (5.7e) at x = 0 and an additional kinematic boundary condition on the side walls of the 2D wave channel at y = ±B in Fig. 5.16. The kinematic and dynamic wave fields may be computed from a dimensional 3D scalar velocity potential 0 (x, y, z, t).
269
Wavemaker Theories
IB
\z,w,ez x,u,ex i
Fig. 5.16. Definition sketch for a 2D sloshing wave channel.
The fluid velocity q(x,y,z,t) may be computed from the negative gradient operator applied to a scalar velocity potential by q(x, y, z, t) = u(x, y, z, t)ex + v(x,y,z, t)ey + w(x, y, z, t)ez = -VQ>{x,y,z,t),
(5.149a)
where the 3D gradient operator from Eq. (2.10a) in Chapter 2.2.7 is - , 3W. *{•)* *{•)* V(») = -z—ex + -z—ey + —— ez. ox
ay
The dimensional fluid dynamic pressure field p(x,y,z,t) from p(x,y,z,t)
= p
(5.149b)
az
may be computed
d<&(x,y,z,t) dt
A dimensional scalar velocity potential &(x,y,z,t) real part of <&(x,y,z,t) = Re{(j>(x,y,z)exp-i(cot
(5.150) may be defined by the + v)},
(5.151)
where co = 2n/T, where T is the dimensional wave period, and where the phase angle vinEq.(5.151)is introduced for data analyses by FFT algorithms (vide., Chapter 9.2). The linearized dimensional WMBVP for a spatial velocity
270
Waves and Wave Forces on Coastal and Ocean Structures
potential
= 0,
x>0,-B
-h
(5.152a)
d
^o^ =
3x
i^H
| kh tanh ^ / j , —Hnh tan icnh,
4>(x,y,z) = 0,
(5.152e)
n= 1
(5.152f)
n > 2,
(5.152g)
where the deep-water wave number &o = co2/g and where the negative sign in the kinematic radiation boundary condition (KRBC) in Eq. (5.152e) is applied from the KRBC Eq. (5.4f) because of the negative sign in the temporal dependency in the dimensional scalar velocity potential in Eq. (5.151). The unknown dimensional free-surface displacement r](x, y, t) may be determined from the dimensional linearized Bernoulli equation for p(x, y, 0, t) = 0 at z = 0 according to V(x,y,t) = Re —ico-Q(x,y,0,t)
,
x>0,
-B
z = 0. (5.152h)
Kinematic wavemaker boundary condition (KWMBC) A Stokes material surface W(x, y, z, t) for a 3D sloshing wavemaker may be expressed as W(x,y,z,t)
=x-£(y,z,t)
= 0,
(5.153a)
where the instantaneous wavemaker displacement £(y,z, t) from its mean position at x = 0 is assumed to be strictly periodic in time with period T = 2n/co and may be expressed by
M(y,z) £(y,z,0 = Re i co exp — i (cot + v)
(5.153b)
271
Wavemaker Theories
Retaining only the linear terms (cf. Eq. (5.7e) in Sec. 2) of the Stokes material derivative by Eq. (2.13a) in Chapter 2.2.10 of Eq. (5.153a) gives a linearized KWMBC as DW{x,y,z,t) = 0 Dt d(x,y,z,t) my,z,t) , x = 0, -B < y < +B, -h < z < 0. dx dt (5.153c) The wave channel boundaries in Fig. 5.16 are separable in Cartesian coordinates so that a solution to the WMBVP in Eqs. (5.152) may be obtained by the method ofseparation of variables (Carrier and Pearson, 1968, Chapter 6; Hildebrand, 1976, Chapter 9.3; Ince, 1956, Chapter. 9.41). The homogeneous boundary conditions prescribed by Eqs. (5.152b,c) are identical to the homogeneous boundary conditions from Eqs. (5.7b,c) in Sec. 2 so that a solution to the WMBVP Eqs. (5.152) may be expanded in the same dimensionless orthonormal eigenseries in Eqs. (5.30) in Sec. 2; viz. ^(Kn,z/h)=C°ShK"f+Z/h\
n = 1,2, . , . , . , . ,
where the nondimensional normalizing constant N„ is 2kh + sinh 2kh 0 4kh ' Ni: = / cosh2z Knh(l + z/h)d(z/h) = 2Knh + sin2/c„/i 4K„JI
(5.154a)
n=1 n > 2,
(5.154b,c) provided that K\ = kforn = I, Kn = i\cn for n > 2, and that the eigenvalues are computed from koh = kh tanh kh =—Knh tan Knh, n>2 (5.154d,e) where k0h = co2h/g. Solutions to the WMBVP in Eqs. (5.152) that satisfy Eqs. (5.152a,b,c) exactly are the following orthonormal eigenseries expansions: ®(x,y,z,t)
= Re \j2Jjrn(x,y)^„(Kn,z/h)exp-i(.eot
V(x,y,t) = Re\^2^n(x,y)exp-i(cot
+ v) 1 ,
+ v) i , (5.155a)
(5.155b)
272
Waves and Wave Forces on Coastal and Ocean Structures
where cf)(x,y,z) = fn{x,y)^n{Kn,z/h) Eqs. (5.155) into Eq. (5.152f) yields Sn(x,y) = fn(x,y)
=i
in Eq. (5.151). Substituting
co ~-i-ir (x,y)^ (K ,0), n n n 8 8 $n(x,y)
a>*n(Kn,oy
(5.156a) (5.156b)
where f„ (x, y) is sometimes referred to as a displacement potential. The scalar potential in Eq. (5.155a) may now be expressed from Eqs. (5.156) as */• ^ p \\^-8>-f ^n(Kn,z/h) Q(.x,y,z,t) = Re ] > i-$„(x,y)
inTi
m
1 exp-i(otf + v) | .
Vn(Kn,0)
J
(5.157) The WMB VP in Eqs. (5.152) may now be expressed in terms of a displacement potential f„(x, y) by substituting Eq. (5.157) into Eqs. (5.152) and obtaining d2S„(x,y) d2S„(x,y) H —2 dx2 dy
2
\- Knt;„(x,y) = 0,
(K ,z h) E dS (x,y)ty dx V (K ,0) n
n
n
n
n
.co ~ g
lim \^--iKn\l;n(x,y) ^+oo [ dx J Hn(x,y) dy
= 0,
x > 0,
j = ±B,
= 0, -h
(5.158c) (5.158d,e)
where Kn are computed from Eqs. (5.154d,e) and where Eq. (5.158a) is the 2D Helmholtz equation (Hildebrand, 1976, Chapter. 8.6, or Morse and Feshbach, 1953, Chapter 5.1). Because the boundary conditions defined by Eqs. (5.158b-e) are prescribed on boundaries that are constant values of (x,y), a solution to the WMB VP given by Eqs. (5.158) may be computed by the method ofseparation
273
Wavemaker Theories
of variables (Carrier and Pearson, 1968, Chapter 7; Hildebrand, 1976, Chapter. 9.3; Ince, 1956, Chapter. 9.41) according to S„(x,y) =
(5.159)
X(x)Y(y).
Substituting Eq. (5.159) into Eq. (5.158a) and then dividing Eq. (5.158a) by Eq. (5.159) gives d2X(x)/dx2
, d2Y(y)/dy2
+
X(x)
2
+ K„ = U.
Y(y)
The first term is a function of x only and the second term is a function of y only. Consequently, each of these two terms must be equal to a constant that gives the following two ordinary differential equations (ODE's): d2X(x)/dx2 —
+ K22 =
2 d2Y(y)/dy y ' J
= ix2 ,
« , „ , „ say, a constant. (5.160a,b)
The two ODE's from Eqs. (5.160) and associated boundary conditions are d2X{x) + {Kln - ^ ) X ( i ) = 0. dx2
J2<m dx rw
n=\
Vn(Kn,z/h) *n(Kn,0)
.a)* = i-U(y,z/h),
(5.161a) x = 0, B
Hrn^ J £ - i^Kl - /i2 } XW = 0,
(5.161c)
and dlY{y) , 2 2f±+li mY(y) = 0, dy dY(y) = 0, dy
-B
(5.162a) (5.162b,c)
Note that the ODE in Eq. (5.162a) and the two homogeneous boundary conditions in Eqs. (5.162b,c) for Y(y) that are prescribed at the constant boundaries y = ±B form a well-posed Sturm-Liouville problem (cf. Chapter 2.6; Morse
274
Waves and Wave Forces on Coastal and Ocean Structures
and Feshbach, 1953, Chapter 6.3; Hildebrand, 1976, Chapter 5.7, or Benton, 1990, Chapter 6.6). Because the only non-trivial solutions to the well-posed Sturm-Liouville problem in Eqs. (5.162) are eigenfunctions that require eigenvalues, the solutions to Eqs. (5.162) must be computed first. The solutions to Eq. (5.162a) for //.^ > 0 are Ym(y) = «m cos/xmy + bm sin[i m y.
(5.163a)
Substituting Eq. (5.163a) into the two homogeneous boundary conditions in Eqs. (5.162b,c) yields —am sin pumB + bm cos ixm B = 0,
y = +B
am sin fimB + bm cos /xmB = 0, DET{am,bm}
y = —B
= 0.
(5.163b) (5.163c) (5.163d)
where DET{», •} is the determinant of Eqs. (5.163b,c). The eigenvalue solutions to Eq. (5.163d) may be computed from sin2/xmfi = 0 mrc /xm = -—, m > 0 and integer. 2B Solving for bm from Eq. (5.163b) gives bm
(5.163e)
=amtmfj,mB,
so that the eigenfunction in Eq. (5.163a) may be expanded in the following eigenseries: Ym(y/B) = ^2
A
m
cos
[VmB(y/B - 1)],
m > 0 and integer, (5.163f)
m=0
provided that the eigenvalues fim = rrnt/lB for m > 0 and integer. The orthogonal eigenseries in Eq. (5.163 f) may be normalized in order to obtain a dimensionless orthonormal eigenseries according to -YW /m cos ^mB(y/B-I) «i«x TmGu,m,y/,B) = — , (5.163g) Mm where the nondimensional normalizing constant Mm for /xm = mn/2B for m > 0 and integer is given by M2m = J cos2 nmB(y/B - \)d{y/B) -l
= 1 + <>m0,
Wavemaker Theories
275
where 8mo is the Kronecker delta function (vide., Eq. (2.2) in Chapter 2.2.3). The orthogonal eigenseries in Eq. (5.163f) reduces in the nondimensional interval of orthogonality —1 < y/B < 1 to the following orthonormal eigenseries:
= ^2 AmTm(ixm,y/B),
m > 0 and integer,
(5.163h)
m=0
provided that the eigenvalues /i m = m?T/2B for m > 0 and integer. The ODE in Eq. (5.161a) may now be expressed as —^-Q2mnX{x)
= 0,
x>0,
(5.164a)
where Qmn = Vm-
K
n>
m > 0,
n > 1,
and integer.
(5.164b)
Following the solution procedure for the ODE for X{x) given in Sec. 2, the general solutions to Eqs. (5.164) for all positive and negative values of the separation constants (^mn (including zero) depend on the magnitudes of K„ relative to the magnitudes of iim. The three possible solutions to the ODE Eq. (5.164a) are of the three forms given by Eqs. (5.18) in Sec. 2. First, consider the relative magnitude ofthe propagating eigenvalue K\ =k relative to the magnitudes of the transverse eigenvalues ixm, i.e.: (1) n = 1 and K\ = k (propagating eigenmode): Qmi = vfin ~ k2 a n d there are the following three possibilities: (i) k > \xm : Qm\ = i^/k2 — \j^ = iPm\ , say; and the solutions to Eq. (5.164a) are of the form X(x) = C3 expiP m i;c +C4exp— iPm\x.
(5.165a)
Substituting Eq. (5.165a) into the radiation condition Eq. (5.161c) with J K2 — ix2^ — yjk2 — ^ consequently, Xmi(x) = c3mi expiPmix,
= P m l requires that 04 = 0, and, k > /xm, m < M,
(5.165b)
where M is the maximum integer value for the index m in order for Hm < k.
276
Waves and Wave Forces on Coastal and Ocean Structures
(ii) k = ixm : Qml the form
= 0, and the solutions to Eq. (5.164a) are of Xm\{x) = c\x + c2.
(5.166)
Substituting Eq. (5.166) into the radiation condition Eq. (5.161c) with 2ml = Oyieldsci = c2 = 0, and there are no solutions to Eq. (5.164a) for 2ml = 0. (iii) k < fim : Qm\ = VM-m — k2 = am\ > 0 where the index m > M that is the maximum value inEq.(5.165b), and the radiation condition in Eq. (5.161c) must be modified to
\fx+^
lim { — + J ii2m - k\ X(x) = 0 lim I — + S m i | X(x) = 0, m > M. (5.167a) x^+oo [dx J The solutions to Eq. (5.164a) for Qm\ = E m i > 0 and for m > M are of the form Xm\{x) = csexp — S m ix + C6exp + H m ix, m > M. (5.167b) Substituting Eq. (5.167b) into the radiation condition in Eq. (5.161c) with— i-s/k2 — ix2^ = + S m i > 0 and form > M requires that ce = 0, and, consequently, Xm\(x) = c5mi e x p - S m i x ,
k < ixm,
m>M.
(5.167c)
Summary (1) Q2mn > 0 and for Kn = Kx= k; ,. v Xm\(x)=\
fc 3 m iexp/P m ix,
m<M (5.168a)
[c5miexp-amix,
m>M
M
Xi(x) = ] P c 3 m i e x p i/»„,!*+ m=0
^
c5miexp-Smix,
(5.168b)
m=M+l
provided that Pm\ = JK2 - ii2m = Jk2 - ix2m, k> fim, ami=J(i2-k2,
k < \Lmt
m>M
+ \,
m<M,
(5.168c) (5.168d)
277
Wavemaker Theories
where k$h =
co2h
g rmt \jim = , IB
= khtanhkh,
(5.168e)
m > 0 and integer
(5.168f)
(2) n > 2 and Kn — IKU (evanescent eigenmodes): Qmn = Mm + Kn > 0 f° r a ^ m an(^ «• The solutions to Eq. (5.164a) for Qmn > 0 for n > 2 and with the modified radiation condition Eq. (5.167a) are of the form of Eqs. (5.167) with S m i = Qmn- Substituting solutions of the form in Eq. (5.167b) into the modified radiation condition in Eq. (5.167a) where Em\ — Qmn for n > 2 that is now given by
HmTO{£ + y^7^}x(*) = 0, lim I — + Qmn | X(x) = 0, x->-+oo [ dx \
m > 0 , n > 2,
(5.169)
requires that ce = 0, as before in (1) iii above, for m > 0 and n > 2, and Eq. (5.167c) reduces to Xmn(x) = CSmntxp-QmnX,
m > 0,
H > 2,
(5.170)
where S m i = <2m« = 4vfyr±*ifor m > 0 andn > 2. Adding Eq. (5.170) to the solutions in Eqs. (5.168b) gives the following solution to the ODE (5.164a): M
X(x) =
y~] Cm\ expiPmix + m=0
Y2 m=M+\
c
m\ exp - S m i x •
(5.171)
m=0n=2
Substituting Eq. (5.163h) and Eq. (5.171) into Eq. (5.159) and then substituting Eq. (5.159) into Eq. (5.157) gives the following orthonormal eigenseries
278
Waves and Wave Forces on Coastal and Ocean Structures
solution to the WMBVP in Eqs. (5.152): Q(x,y,z,t)
= L CmiTm(y/B) m=0
Re
+
i—
u>
Z.
expiPmix *l(fe,0)
C„,iT m (y/g)
m=Af+l
exp-amix
exp —i(otf + v)
*l(^,0)
LcmnTm(y/B)————exp-Qmnx
+ 22 m=0«=2
*«(*>0)
(5.172a) where Tm(/xm,y/B) =
c o s i i m B ( y / B - 1)
(5.172b)
M„
M,£ = 1 + 5 m 0 ,
*„(^„,zA) = AK
(5.172c)
cosh/«:„/i(l + z / A ) Nn 2kh + sinh 2kh
(5.172d) '
4kh ' 2icnh + sin2icnh 4Knh
n = 1 (5.172e,f)
« > 2,
where ^.1
=
TJKI
-
V-m = V R l
am\ = JjA.
~ Vm>
- k2,
k
>
Mm,
k < iim,
m
<
M,
(5.172g)
m>M,
(5.172h)
n>2
(5.172i,j)
provided that co2h koh = — =
khta.nh.kh = —Knh tan Knh;
rrnt Mm = 2 ^ - ,
m > 0.
(5.172k)
The eigenseries coefficients Cmn may be determined in a best least squares sense from the KWMBC by substituting Eq. (5.171) for X(x) andEq. (5.163h)
Wavemaker Theories
279
for Y(y) into Eq. (5.161b) and obtaining 1
M
*i(k,z/h) *l(*,0)
m=0
VxiKz/h)
2^ SmiCmi*H*.^ m=M+\ _VAY^/o r ^n{Kn,z/h) *n(Kn,0) m=0 n=2
rm(jim,y/B)
CO
= i — U(y,z),
x — 0.
(5.173a,b,c)
Multiplying each side of Eqs. (5.173) by a member of both orthonormal eigenseries from Eqs. (5.172b,d) and integrating over both dimensionless intervals of orthogonality (viz. — 1 < y/B < + 1 and — 1 < z/h < 0) yields Cm\ —
ir/>©/>(*)"(>•>(*••*) x T , \V"n>
B)>
m < M,
(5.174a)
f+l
.coVi(Kn,0) Cm\ = ~i
^Mtf/GM^W^)
xTm(^m,-|), f+l
.coVn(Kn,0)
T
slmn (
(5.174b)
j(y
/., '©/..'(IW'-X*--*)
—i
8
m>M+\,
L\
m > 0, n > 2.
(5.174c)
The first three transverse eigenmodes are illustrated in Fig. 5.17.
5.7. Conformal and Domain Mapping of WMBVP Conformal and domain mappings are applications of complex variables to solve 2D boundary value problems (Morse and Feshbach, 1953; Carrier et al, 1966; Milne-Thompson, 1968; inter alios). Conformal mapping is an angle preserving transformation that will compute exact nonlinear solutions for surface gravity waves of constant form that may be treated as a steady flow
280
Waves and Wave Forces on Coastal and Ocean Structures
Fig. 5.17. First three transverse eigenmodes in a 2D wave channel.
following a Galilean transformation (vide., Chapter 6.1) from an inertial fixed coordinate system to a non-inertial moving coordinate system (Nekrasov, 1921 and 1922; Milne-Thompson, 1968, Chapter 14.70, or Wehausen and Laitone, 1960, Sec. 32). A periodic nonlinear surface gravity wave may be mapped into an annulus and the BVP solved numerically. An integral transformation for 2D surface gravity waves is given by Cokelet (1977). Domain mapping (Joseph, 1973) is a transformation of the wavemaker geometry into a fixed computational domain where a solution for the time-dependent, non-linear, free surface may be computed efficiently (DeSilva et al, 1996).
5.7.1. Conformal
Mapping
Conformal mapping of the wavemaker boundary value problem (WMBVP) provides a global solution that accurately accounts for the singular behavior at all irregular points (Tanaka, 1988). The irregular points in the physical wavemaker domain are transformed into both weak and strong singular kernels in a Fredholm integral equation (Garabedian, 1964, Chapter 9.3 or Morse and Feshbach, 1953, Chapter 8). The two irregular points for a full-draft wave maker on the WMBVP boundary that are located at: (1) the intersection between the free-surface and the wavemaker boundary, and (2) the intersection between the horizontal bottom and the wavemaker boundary. These two irregular points exhibit integrable weakly singular kernels. The far-field radiation boundary exhibits a strongly singular kernel and significantly affects the solution. The irregular frequencies (John, 1949 and 1950) are included in the global solution by the Fredholm alternative (Garabedian, 1964, Chapter 9.3 and Guenther and Lee, 1996, Chapter 7.7). General theories for treating singular solutions near
281
Wavemaker Theories
irregular points are given by Lewy (1946) and by Stoker (1947). Kravtchenko (1954) appears to have been the first to apply the theory of analytic functions that is generalized by Lewy (1950) to obtain two logarithmic solutions near the two irregular points. These two logarithmic solutions do not satisfy completely all of the boundary data because they are only local solutions that are valid near the irregular points, and he was not able to quantify all of the coefficients in his solutions. The numerical algorithm of Lin (1984) analyzes only the irregular point at the free-surface in the time domain for an impulsively generated full-draft piston wavemaker. A theory for the planar WMBVP is derived in Sec. 2. Tanaka (1988) computes a global solution for the WMBVP by a conformal mapping of the physical wavemaker boundary to a unit disk that accounts for the motion of an inviscid fluid near irregular points that are illustrated in Fig. 5.18. A numerical solution to Laplace's equation in a transformed unit disk may be computed from a Fredholm integral equation (Garabedian, 1964, Chapter 9.3 or Morse and Feshbach, 1953, Chapter 8). The WMBVP defined by Eqs. (5.7) in Sec. 2 is transformed into complex-valued analytical functions where the complex-valued coordinates are defined as z = x + iy.
(5.175)
where the x-y plane now replaces the x-z plane in Sec. 2. The coordinates for the semi-infinite wave channel that is illustrated in Fig. 5.3a in Sec. 2 must now be transformed to the complex-valued coordinates by Eq. (5.175). The dimensional fluid velocity may be computed from the negative gradient of a dimensional scalar velocity potential <J>(x, y, t) according to q(x,y,t)
= u(x,y,t)ex
+ v(x,y,t)ey
= -V<J>(x,y,0,
(5.176a)
where the 2D gradient vector operator is V(.) = exd(.)/dx
+ eyd(.)/dy
(5.176b)
where ex and ey are the unit vectors in the x and y coordinates, respectively. The dimensional linear dynamic pressure field p(x,y,t) may be computed from the unsteady Bernoulli equation (4.1 If) in Chapter 4.3 according to p(x,y,t)
=p
d$>(x,y,t) at
.
(5.176c)
282
Waves and Wave Forces on Coastal and Ocean Structures
02(B 2 ) Fig. 5.18. Combinations of Irregular (I) and Regular (R) boundary points Pi and P2 between Smooth (S) and Critical (C) boundary Bi and B2 intersections for domain D in the WMBVP (Tanaka, 1988).
A dimensional scalar radiated velocity potential 4>(x,y) may be defined by the real part of <&(x,y,t) = Re{(p(x,y)exp-i(cot
+ v)},
(5.176d)
where the Re{«} is the real part of {•}, co = 2n/T, T = the wave period and v = an arbitrary phase angle. The dimensional linearized WMBVP correct to
283
Wavemaker Theories
0(e = kA) and for kh = 0(1) is: V24>(x,y) = 0, 0 < x < +oo, -h < y < 0, dtp(x,y) = 0, dy dcp(x,y) -k04>(x,y) = 0, dy d(p(x,y) exp — i(cot + v) = dx
(5.177a)
0 <x + oo, y = -h, 0<x<+oo,y da(y/h,t) , at
(5.177b)
= 0,
(5.177c)
x = 0, — h < y < 0, (5.177d)
r](x,t) = Re
lim \-^-~iKn}(t>(x,y) x^.+oo [dx J —ico ] (/>(x,0)exp-i(cot + v)\,
I 8
I
= 0, x>0,y
(5.177e) = 0,
(5.177f)
J kh tanh JkA,
n= 1
(5.177g)
-Knhtanicnh, n>2, (5.177h) 2 where the deep-water wave number ko = co /g, where the homogeneous boundary conditions in Eqs. (5.177b,c) are the homogeneous Neumann and Robin boundary conditions in Table 2.3 in Chapter 2.5.1, respectively (Guenther and Lee, 1996, Chapter 8.1), and where the negative sign in the kinematic radiation boundary condition (KRBC) in Eq. (5.177e) is a consequence of the negative sign in the temporal phase dependency in Eq. (5.176d). The instantaneous dimensional wavemaker displacement H (y/h, t) from its mean position x = 0 is assumed to be strictly periodic in time with period T = 2TT/CO and may be expressed by Eq. (5.8) in Sec. 2 by E (y/h, t) = Re \ i \ ; ; ^ | x (y/h) exp -i (cot + v)
l(A/h)\
.1
+ v), (5.177g) fA/u.\X(y/h)sm(cot _(A/h)] X (y/h) = [a(y/h) + $] [U(y/h + 1 - d/h) - U(y/h + b/h)], (5.177h) where the dimensionless specified shape function x (y/h) for the Type I wavemaker that is shown in Fig. 5.3a in Sec. 2 is valid for either a double-articulated
284
Waves and Wave Forces on Coastal and Ocean Structures
piston or hinged wavemaker of variable draft (vide., Eqs. (5.9) in Sec. 2) and where £/(»)=Heaviside step function in Eq. (2.1) in Chapter 2.2.2. There are both Irregular (I) and Regular (R) points at the intersections between the Smooth (S) and Critical (C) boundaries B i and B2 in the WMB VP as illustrated in Fig. 5.18 where these two boundary intersection points are identified as Pi and P2, respectively. The classification of the boundary points Pi and P2 in Fig. 5.18 depends on: 1) the boundary conditions 0,(Py), and 2) the continuity of the boundaries B m and their derivatives where i, j and m — 1 or 2. A conformal mapping of the semi-infinite wave channel strip in the lower half x-y plane yields a Fredholm integral equation (Garabedian, 1964, Chapter 9.3 or Morse and Feshbach, 1953, Chapter 8.4) where these critical points may be transformed to singular points that are integrable over a smooth continuous mapped boundary. Tanaka (1988) and Averbeck (1993) applied two different conformal mappings to transform the semi-infinite strip in the lower half x-y plane to a unit disk. Conformal Mapping to a Unit Disk (Tanaka, 1988) The two conformal mappings applied by Tanaka (1988) are: (i) the physical z-plane to a semi-circle in the upper Z-plane, (ii) a semi-circle in the upper Z-plane to a unit disk in the <2-plane. Global numerical solutions are computed for both the linear and the fully nonlinear WMBVPs. z-plane ->• Z-plane Mapping: The first conformal mapping is illustrated in Fig. 5.19 and maps the semiinfinite strip in the lower half-plane in the physical z-plane into the upper half-plane in the Z-plane by the following Schwarz-Christoffel transformation (Carrier et al., 1966 or Milne-Thompson, 1968): dz
C\
dZ
JTT\<JZ~^\'
(5.178a)
that may be integrated according to z=
fzz(Z)
= d In \z + VZ - ll + C2.
(5.178b)
285
Wavemaker Theories
( z- plane)
y/////////////////M
f////////////////////& iY t
( Z-plane)
®
®
©©
V/////,////////////^ Fig. 5.19. Mapping of the semi-infinite strip in lower half x-y plane in the physical z-plane to the upper half plane in the Z-plane (Tanaka, 1988).
Tanaka (1988) evaluates the two unknown coefficients C\ and C2 in Eq. (5.178b) by mapping: (i) the equilibrium intersection of the SWL and the vertical wavemaker boundary B in Fig. 5.19 at x = y = 0 in the z-plane to the point B' at X = — 1 on the real line in the Z-plane; and (ii) the intersection between the horizontal flat bottom and the vertical wavemaker boundary D in Fig. 5.19 at x = 0 and y = — h in the zplane to the point D' at X = + 1 on the real line in Z-plane as illustrated in Fig. 5.19. The inverse mapping function from Eq. (5.178b) of the semi-infinite strip in the lower half x-y plane in the physical z space to the upper half-plane in Z space is given by z = x + iy = fZz(Z)
= -^-Log [ - Z - VZ 2 - l] ,
(5.178c)
where the Log[«] in Eq. (5.178c) denotes the principal value of ln[»] in Eq. (5.178b) and the argument of the Log[«] in Eq. (5.178c) is —it < arg < n (Hildebrand, 1976, Chapter 10.3 or Milne-Thompson, 1968, Chapter 5.20).
286
Waves and Wave Forces on Coastal and Ocean Structures
The semi-infinite strip in the lower half x-y plane in the physical z-plane may also be mapped into the upper half-plane in the Z-plane by Z = X + iY = fzz(z/h)
= - cosh[7r z/h].
(5.178d)
The WMB VP in Eqs. (5.177) may be transformed to the upper half-plane in the Z-plane by Eq. (5.178d) and the scalar velocity potential
(5.178e)
Eq. (5.178d) is now
where X and Y are each solutions to a transformed Laplaces's equation Eq. (5.177a). Partial spatial derivatives are transformed to dir(X, Y) 3X(x, y) dx dx djr(X,Y)dX(x,y)
d
dx 3X
3v
dy
+
d^{X, Y) 8Y(x, y) 8Y dx djs{X,Y)dY{x,y) 8Y 3y
(5.178h) (5.178i)
The coordinate transformation given by the mapping function fzz(z) in Eq. (5.178d) with Eqs. (5.178f,g) is an analytic function that satisfies the Cauchy-Riemann equations (Carrier et al, 1966, Chapter 2.1; Hildebrand, 1976, Chapter 10.4; Milne-Thompson, 1968, Chapter 5.30; Morse and Feshbach, 1953, Chapter 4.2) dX _dY dx dy'
3X _ dy
37 dx
(5.178j,k)
The transformed WMBVP in Eqs. (5.177) is given by (Arfken, 1985, Chapter 6.7) as: dX2(x,y) V 0(*,3O = v V ( X , r ) dx2 2
=
dY2(x,y) V ir(X,Y) dx2 2
+
dX2{x,y)
(5.179a)
dy' +
dY2(x,y) dy2
= 0,
(5.179b)
because both X(x,y) and Y(x,y) satisfy Laplace's equation (Arfken, 1985, Chapter 6.7). The homogeneous Neumann Eq. (5.177b) and Robin
287
Wavemaker Theories
Eq. (5.177c) boundary conditions (vide., Table 2.3 in Chapter 2.5.1) transform to
df(X,Y) W(X,Y) dY
dY koh
+'
= 0,
(5.179c)
X > 1, Y = 0,
-.ir(X,Y) = 0,
nVW^T
X < - 1 , F = 0.
(5.179d)
The kinematic wavemaker boundary condition Eq. (5.177d) transforms to djr(X,Y) dY — cos
1
co
L(A/A)J "• Vl - X2 ixb
< X/h < cos
X(X/h),
ltd
(5.179e)
h
Y = 0 where the transformed dimensionless shape function x (X/h) is x(X/A) = x(y//« = -arcco8(-X) [/
X-cos
7Td
£/ X + cos
7T&
-7T < arccos(-X) < 0,
(5.179f)
where [/(•) is the Heaviside step function in Chapter 2.2.2. Some care is required in order to transform the radiation boundary condition Eq. (5.177e) to the Z-plane. The radiation boundary is a straight line that connects points A-E in the physical z-plane in Fig. 5.19; and this straight line transforms to a semicircle of radius R from the origin C that connects point A' to point E'in the Z-plane in Fig. 5.19. Consequently, a transformation to circular cylindrical coordinates [R, 0} may be defined as R2 = X2 + Y2 =
1
C0S, Y
T) + C 0 S h (^r)]' vizx-\
tan6> = — = tanh —
Y7iy~\
tan
—
(5.179g) (5.179h)
288
Waves and Wave Forces on Coastal and Ocean Structures
The radiation condition in the Z-plane for the transformed velocity potential in circular cylindrical coordinates \j/[R(X, Y),0(X, Y)] is lim
R^-oo
A(R,9)
v
l'
OK
VK
± B(R,0)
J
- ikf{R,6)
60
+ B(R, 9) whenO < 9 <
= 0,
It
(5.179i)
It
B{R, 9) w h e n - < 0 < it where A(R,9) = ^— J2(1 - R~2) + 4R~2 sin2 9 + 2^(1 - R~2) +4R~2 sin2 9, 2/z
B(R, 9) = —J -2(1 - R~2) - 4/?- 2 sin2 9 + 2Jv (1 - R~2) + 4R~2 sin2 9. 2« ' As a consequence of the Jacobian of the transformations by Eqs. (5.178f,g), the points B at x = y = 0 and D at x = 0 and y = — h in the semi-infinite strip in the physical z-plane in Fig. 5.19 are now critical points in Fig. 5.18 at B' and D' in the Z-plane. Z-plane -> Unit Disk Mapping The upper half-plane in the Z-plane may be mapped into a unit disk in the <2-plane as illustrated in Fig. 5.20 by the following bilinear transformation (Carrier etal, 1966 or Milne-Thompson, 1968):
G = l + '? = /zfi(Z) Z-Z0 = exp(i#0)
(5.180a)
7*
where the superscript asterisk * denotes a complex conjugate value. Tanaka (1988) evaluates the constant #o in Eq. (5.180a) by equating the point P' = Z = i on the iY axis in Fig. 5.20 to the point E" = A" = exp i ± it on the unit disk; and Eq. (5.180a) reduces to
g = £ + i? = /zfi(Z) =
i-Z T
i+Z
(5.180b)
289
Wavemaker Theories
Z-plane ©•
®
v/////?v////;v///// x
®
©
g> ©
lane
I
Fig. 5.20. Mapping of the upper half-plane in the Z-plane to the unit disk in the g-plane (Tanaka, 1988).
that maps the upper half-plane in the Z-plane onto the circumference of the unit disk in the £>-plane- The inverse mapping function for Eq. (5.180b) is
.n-Q
Z = X + iY = fQZ(Q)
=i
1+ 0
(5.180c)
The mapping function coordinates in Eq. (5.180b) are 1-X^ £ = X2 + (F + 1)2'
C =
2X X + (F + 1) 2 ' 2
(5.180d,e)
that may be transformed into circular cylindrical coordinates for the unit disk in Fig. 5.20 by
2
,2 ,,2 §
+
^
(X' + -
Y*-iy-4X*
[X2 + (F + 1) 2 ] 2
m = arctan ( j J = arctan ( 1 _
x 2
'
_y2 j•
(5.181a) (5.181b)
290
Waves and Wave Forces on Coastal and Ocean Structures
The transformation by Eqs. (5.181) transforms the following: (1) the vertical wavemaker partition D-B in Fig. 5.19 to the circumference of the semi-circle D"-B" in the first and fourth quadrants in Fig. 5.20, (2) the free-surface B-A in Fig. 5.19 to the circumference of the quartercircle B"-A" in the third quadrant in Fig. 5.20, (3) the horizontal bottom D-E in Fig. 5.19 to the circumference of the quarter-circle D"-E" in the second quadrant in Fig. 5.20, (4) the vertical boundary at infinity A-E in Fig. 5.19 to a point at A"-E" at ± n in Fig. 5.20. The transformation of ± oo to a point on the circumference of the unit disk in the (2-plane results in substantial numerical complications in the evaluation of the transformed radiation condition Eq. (5.177e) (Tanaka, 1988). A transformed scalar velocity potential
= =
d&(r,m)dr dr 3X d*(r,m)dr dr dY
d&(r,m)dm
,
dm dX d®(r,m)dm dm dY '
« 1 0 ^ (5.181c) ' 181d> K• )
v
The transformed WMBVP is now given by V 2 $(r,ra-) = 0, ^ ^ = 0 , dr d<&{r,m) dr
koh
(5.182a)
r = l, -<m<„, 2
(5.182b)
$>{r,m)
V27T V
d$>(r,m) dr
r = l, -n < m < +n,
-
cos m (cos m + 1)
0,
it r = 1, — TT < m < 2 (5.182c)
w X(ar) c o s m cos l(A/h) Ti -Jl V ( nr + 1)'
r = 1, -7L<m<JL,
(5.182d)
291
Wavemaker Theories
where the transformed shape function / (m) is ~, x , y 1 / sin|nr| [tf(n7+0fc)-£/(nr-0d)], X O ) = X ( T- = - arccos \h n \ cos |JZT| + 1 sinnr (5.182e) <0, —7r < arccos cos nr + 1 and where -2cos(f)
#b = arctan
COS"
(?) + !.
2cos(f)
#d = arctan
(5.182f,g) In order to map the radiation condition Eq. (5.179i) from the Z-plane to the unit disk in the <2-plane, the mapping function [R, 0} —> [r, m} is (R2 -
2 r
=
\)2+4R2cos2
[R2 +
n
\+2RsinO]
m = arctan
2R cos6' \-R2
(5.182h,i)
and the mapping of the radiation condition Eq. (5.179i) to the point r ->• 1 and m —> JT± on the unit disk in the <2-plane yields lim
C(r, m)
r->l
9
m—>— j r :
where
(5.182J)
C(r,m) -
- (1 — r2)Ei(r,m)cosm ± (1m++r21)E2(r,m) r2 — 2r cos
sinnr (5.182k)
2\ ^ /l -r2 E\ (r, m) sin nr ± I I Ejir, m) cos m
Z>(r,nr) =
[m
r2 — 2r cos m + 1 (5.1821)
292
—J
Waves and Wave Forces on Coastal and Ocean Structures
I - 2 r cosm + j(r2
- l) 2 + 4 r 2 c o s 2 m\ [r2 - 2r cosm + l] + (r 2 - l) 2 , (5.182m)
E2{r,m) = J
\lr cos m + y/(r2 - l) 2 + 4 r 2 c o s 2 nrl [r 2 - 2r cos nx + l] - (r 2 - l) 2 .
(5.182n) The ± exponent for m = n ± in Eq. (5.182j) is to be interpreted as m -> — n+ in the second quadrant from D^-E" in Fig. 5.20; and as zu —> —n~ in the third quadrant from B"-A". The lengthy calculations that are required to evaluate these transformations are given by Tanaka (1988). Fredholm Integral Solution Tanaka (1988) computed numerically a solution to the transformed WMBVP (5.182) from Green's second identity Eq. (2.123c) in Chapter 2.6 (Guenther and Lee, 1989, Chapter 8.3 or Hildebrand, 1976, Chapter 6.14) that is given by the following Fredholm integral (Chapter 8.6; Morse and Feshbach, 1953, Chapter 8; Courant and Hilbert, 1966, Chapter III; Garabedian, 1964, Chapter 10; Stakgold, 1979, Chapter 6.2):
r;i -L
~ dG(r,r;m,m) Q(r,m) — dr
„ \-G(r,r;m,m)
3
where the fundamental Green's function G(r, r; m, m) in 2-dimensions is G(r, f; m,m) = - In [p(r, r; m, m)] , p2(r,r; m,m) - (£ - | ) 2 + (? - I)2
(5.183b)
= (r cos m — r cos m) + (r sin m — r sin m) = r2 -2rrcos(m
-m)+r2.
(5.183c)
Wavemaker Theories
Fig. 5.21. Nodal points on the unit disk in the g-plane and the corresponding nodal points on the wavemaker in the physical z-plane (Tanaka, 1988).
Tanaka (1988) computes numerically solutions to the Fredholm integral equation (5.183a) by discretizing the unit disk in the Q-plane as illustrated in Fig. 5.21 along with the corresponding discretized nodal points in the physical z-plane. The nodal numbering sequence begins with #1 near the bottom on the vertical boundary at x ->• +oo in the physical z-plane or, correspondingly, at the indented contour near ± n on the unit disk in the (2-plane. The numerical details regarding the evaluation of the Fredholm integral equation Eq. (5.183a) at the two weakly singular irregular points at B and D in the physical z-plane in Fig. 5.19 and the strongly singular point at ±n that is a mapping of the vertical line A-E at +oo in the physical z-plane in Fig. 5.19 are tedious but familiar to numerical analysts and are given by Tanaka (1988). Global numerical solutions may be computed for both the linear and the nonlinear WMBVPs. The numerical difficulties associated with the strongly singular point at ± re
294
Waves and Wave Forces on Coastal and Ocean Structures
are the result of the points chosen to quantify the transformation coefficients C\ and C2 in Eq. (5.178b) and provided the motivation for Averbeck (1993) to consider a different mapping algorithm from the physical z-plane to the unit disk. Conformal Mapping to the Unit Disk (Averbeck, 1993) The physical rectangular wavemaker geometry illustrated in Fig. 5.22 is mapped to the unit disk by two successive conformal transformations. The WMBVP defined by Eqs. (5.177) remains unchanged except for the wavemaker displacement x (y/h) that is now given by X(y/h)
= [a(y/h) + 0] [U(y/h + 1 - b0/h) - U(y/h + a0/h)],
(5.184)
where a, ft are constants and £/(•) is the Heaviside step function in Chapter 2.2.2. In order to obtain a transformation to a Jacobian elliptic function (Milne-Thompson, 1965, Chapter 16), Averbeck (1993) first rotated counterclockwise and then shifted the rectangular wavemaker strip a-b-c-d in the physical lower half-plane in Fig. 5.22 to the w-plane as illustrated in Fig. 5.23. The horizontal length of the finite-length rectangular wavemaker strip was set to ^oo = 3h as illustrated in Fig. 5.23(i). The 90° rotation to the z' plane in Fig. 5.23(h) is given by z'
= x' + iy' = iz = -y + ix.
(5.185a)
-y X
V7
*"*'J.
Suriaee EC
n
Wavemakar
Lower
Wall d> x = 0 EC
^
1
radiation BC
= 5(ox(v/A) Bottom . GC
= 0
'
Fig. 5.22. Linear WMBVP with the six critical boundary points at a-a^-b-bo-c-d (Averbeck, 1993).
Wavemaker Theories
*
295
iy'
z-plane
z'-plane
7>h
-h
3/i
(I) z"-plane
ti>" 3A
-.5*
(Hi)
.5h
Fig. 5.23. Rotation and translation of the physical wavemaker rectangular strip in the z-plane to the w-plane (Averbeck, 1993).
The following horizontal shift to the left that is illustrated in Fig. 5.23(iii) is required in order to center symmetrically the rectangular h x 3/z wavemaker strip and to map it to the upper half-plane: h h (5.185b) J - - + ix. = -y 2 2 In order to map the physical z-plane to the upper half Z-plane as a Jacobian elliptic function, the rotated and translated strip must have the dimensions of -K < § < K and of 0 < f < K' where K = h/2 and K' = 3h = 6K. These dimensions require a coordinate amplification of 2K „ w h 2K ., ,, = —(x" + iy") h IK ( h (5.185c) + ix = ~h\-y-2 The Schwarz-Christoffel transformation (Carrier et ah, 1966, Chapter 4.4) maps the irregular boundary points at the vertices of the horizontal flat bottom at b and c in the physical z plane in Fig. 5.22 to the reflections of the irregular z — x + iy = w
296
Waves and Wave Forces on Coastal and Ocean Structures
boundary points at the vertices of the horizontal SWL at a and d in Fig. 5.22, respectively, according to dw =
CkdZ , V(a - Z)(b - Z)(c - Z)(d - Z)
(5.185d)
where C is the arbitrary constant to be determined. The mapping in Eq. (5.185d) may be interpreted visually as opening a "folded fan"; i.e. the rotated rectangular wavemaker a ->• b —>• c —> d in Fig. 5.22 into a semi-circle with the perimeter of the wavemaker boundary on the following: (1) the free-surface (a-d), (2) the wavemaker partition (a-b), (3) the horizontal bottom (b-c) being mapped as the base of the opened fan along the real axis in the Z-plane and the vertical boundary at infinity (c-d) mapped to the circumference of the semi-circle of the opened fan in the Z-plane as shown in Fig. 5.23 The reflected symmetry illustrated in Fig. 5.23(iv) requires that b = —a and c = —d. The following change of variables Z = aZ,
dZ = adZ,
K
= a/c,
C= c
modifies Eq. (5.185d) to the following Jacobian elliptic integral (Churchill et al, 1976, Chapter 93; Churchill, 1960; Chapter 93, or Milne-Thompson, 1965, Chapter 16):
w
= f [(l-Z2)(l-^z2)j
1/2
' =sn'1[Z,K],
(5.185e)
Jo
where SH[Z,K] is the Jacobian elliptic function of modulus K or sine amplitude function (Carrier et al, 1966, Chapter 4.4 or Whittaker and Watson, 1927, Chapter 22). A special value ofEq. (5.185e) when Z = 1 may be defined as k = sn-l[l,K].
(5.185f)
297
Wavemaker Theories
Substituting Eqs. (5.185c,f) into the inverse of Eq. (5.185e) yields the following mapping of the rectangle with dimensions {xi, %2\ y\,yi\ = {0,3h; 0, h}: Z = X + iY
fxH)dn2 y +
dn
l)>K}
- [-f( 2K +i
2Kx
2Kx -' ,k
{y + i)'K\dn[-T{y dn2 2
-'
+
[- -Tr(y+h2)>K.
^)'K\sn[^r'k] \cny 2Kx
~2Kx -'
—
,k
-,k
(5.185g) where sn[», •] in the copolar half-period trio in Eq. (5.185g) is defined in Eq. (5.185e)andcn[», •] and dn[», •] are defined by (Milne-Thompson, 1965, Chapter 16) cn2[; •] = 1 - sn2[; . ] ,
dn2[; .] = 1 - K2sn2[; . ] .
(5.185h,i)
The mapping of the wavemaker rectangle in the physical z-plane to a semicircle in the upper half-plane of the Z-plane is illustrated in Fig. 5.24. The modulus K of the Jacobian elliptic functions is a function of the ratio of adjacent sides of the rectangular wavemaker where this ratio is equal to three in the z-plane that is illustrated in Fig. 5.23(i). The next mapping from the upper halfplane in the Z-plane to the unit disk in the g-plane is illustrated in Fig. 5.25. z-plane
iL///////, V
////////
/ft/?'ay0bBb/?
Fig. 5.24. Mapping of the wavemaker rectangle in the z-plane to a semi-circle in the upper half-plane in the Z-plane (Averbeck, 1993).
298
Waves and Wave Forces on Coastal and Ocean Structures
Z-plane
'd a a„
bj
Fig. 5.25. Mapping of the wavemaker semi-circle in the upper half-plane in the Z-plane to the unit disk in the 2-plane (Averbeck, 1993).
The mapping from the upper half-plane in the Z-plane to the unit disk in the g-plane may be obtained from Averbeck (1993) by (1 - a2) - Z 2 (l + a)2 + 2iZ(l - a)2 (l-a2) + Z2(l+a)2 (5.186a) where — 1 < a < +1 and may be adjusted in order to expand the singularity at the far-field radiation boundary that had complicated the numerical solutions computed by Tanaka (1988) and that was the motivation for Averbeck (1993) to try a different conformal mapping. Transforming the unit disk mapping Eq. (5.186a) to circular cylindrical coordinates results in i - Z - a(i + Z) Q = i + Z-a(i-Z)
2
2
R (X,Y)
(1
Z\2z q) - 27(1 - a1) + {X1 + Yz)(\ + Z\2 a)
= (1 - a)2 + 27(1 - a2) + (X2 + Y2)(l + a)2'
9(X,Y) = arctan
2X(l-oe2) ( l - a ) 2 - ( X 2 + 72)(l+a)2
(5.186b) (5.186c)
that may be transformed into very lengthy and complicated functions of the copolar trio of Jacobian elliptic functions (Milne-Thompson, 1965, Chapter 16) by substituting the real and imaginary parts of Eq. (5.185f) into Eqs. (5.186b,c) (Averbeck, 1993, Eqs. (3.11 and 3.12)). The parameter a in Eqs. (5.186) maps a point to the unit disk that allows an expansion of the far-field point x -* oo that reduces the numerical complications that Tanaka (1988) experienced in evaluating the radiation condition Eq. (5.182J) at the point {r,m} = {l,±7r}. Expanding Eq. (5.186c) and
Wavemaker
299
Theories
solving the resulting quadratic equation for a gives tan 9 + sec 9 a =
(5.186d)
cos 9
WMBVP Transformed to a Unit Disk in the Q-plane The WMBVP boundary conditions Eqs. (5.177b to e) may be transformed to a unit disk according to d
R = I,
s/2k0h(l
d(R,9) dR 4K
-
(5.187a)
eb<e<ec, a2)
(2a + ( l + a 2 ) c o s 6 » ) x ( 4 a + K(l - a)2(0-2a2) cos9 - 1))
R= h
1/2
(5.187b)
< v <
d(/)(R,e) X (Y (R, 0)) (1 - a2) sec Q V « + a2 - (l - a) tan
(^\) 1/2'
(2a + (1 + a2)cos6>) 2
4K ( 1 + a + (a - a) tan ( - J J R = l,
d
z2
x | ( 1 + a»\2 ) - K ( 1 - a ) \ tta, „n22 ( (5.187c)
J
a0
(l-a2)sec(-
ikh AK r
(2a + (1 + a2) cos 6)
1/2
= 0,
2 2 2 x ((l+a) -/c(l-a) tan Q))
R = h
9C < 6 < it and — n < 9 < 0d.
(5.187d)
300
Waves and Wave Forces on Coastal and Ocean Structures
The transformed boundary conditions Eqs. (5.187) are mapped to the arcs on the perimeter of the unit disk as illustrated in Fig. 5.26. The transformed WMBVP may be compactly written as 1 3 v cf>(R,e) = ~RJR z
kd(f>(R,8)\
R-
dR
, 1 d2cP(R,6) R2 B62
\0
=Q
\-n
<0 < +7T,
d
cKR,e) = J2R" n=0
cos n0 + bn sin n6
(5.189)
1 +<5n0
where <5,7 is the Rronecker delta function from Chapter 2.2.3. Substituting Eq. (5.189) into the generic boundary conditions on each of the six arcs on the perimeter of the unit disk illustrated in Fig. 5.26, multiplying each of these six boundary conditions by a member of the set of the orthogonal series in Eq. (5.189), and integrating over the interval of orthogonality — n < 0 < + TZ
Fig. 5.26. Transformed boundary conditions mapped to arcs on the perimeter of the unit disk (Averbeck, 1993).
301
Wavemaker Theories
yields the following matrix equation for each of the coefficients an and bn: AB = H.
(5.190)
The numerical details are lengthy and may be found in Averbeck (1993). 5.7.2. Domain
Mapping
Domain mapping of the WMBVP by DeSilva et al. (1996) follows the theory by Joseph (1973). The fully nonlinear WMBVP with surface tension is computed numerically without the approximate perturbation methods reviewed in Chapter 6.3. The physical fluid domain for the fully nonlinear WMBVP is mapped to a fixed computational fluid domain; and the discretized coupled free surface boundary conditions are computed by an implicit Crank-Nicholson (C-N) method (Carnahan et al, 1965, Chapter 7.12 or Glowinski, 1984, Chapter III.3.3). At each iteration of the C-N method, the potential field is computed by the conjugate gradient method (Atkinson, 1989, Chapter 8.9 or Glowinski, 1984, Chapter VII). This review corrects the errors in DeSilva et al. (1996). The fully nonlinear WMBVP with dimensional surface tension T for the wave channel geometry shown in Fig. 5.27 is given by Q(x,y,t)
= -
dx
3v
®(x,y,t) (5.191a)
= -V2$(x,y,0, P(x,y,t) p
d
1 V <E>(jc,y,0 2 2
V22<S>{x,y,t) = A2<S>(x,y,t) = 0,
dt>(x,y,t) dt
1 V2®(x,y,t) 2
(5.191b)
gy,
P ~ ^ - r(x'°' \a(y/h,t) <x
d2F(x,t) dx2
1 + ar(jc,o
3/2
(5.191c)
+ sr(*, o = o,
dx
(5.191d)
302
Waves and Wave Forces on Coastal and Ocean Structures
d*(x,y,t) dy
dr(x,t)3Q(x,y,t) dx dx E(T(x,t),t)<x
d<S>{x,y,t) = 0, dy 3<£(x,y,Q = 0, dx
3<6(jc,y,0 dx
|
3r(x,Q^Q dt y=
T(x,t),
S(0,/) < x < L, y = 0,
(5.191f)
x = L, 0
(5.191g)
< T(L,t),
3S(y/A,0 , 3E(y/h,t) 3? 3j
I
(5.191e)
x =
3
a(y/h,t),
(5.191h)
0
where the instantaneous dimensional wavemaker displacement a(y/h,t) from its mean equilibrium position x = 0 is assumed to be strictly periodic in time with period T = 2n/co and is given by Eq. (5.177g). The initial conditions are T(x,0) = H(x) dT(x,t) dt
= 0
3(x,0) < x < L.
(5.191i,j,k)
<&(jc,r(jc,o),o) = o Domain Mapping to a Fixed Rectangle
(0,0) Fig. 5.27. Physical wavemaker fluid domain (DeSilva et al., 1996).
303
Wavemaker Theories
The physical fluid domain shown in Fig. 5.27 may be mapped into a dimensionless fixed rectangle of dimensions 0 < £ < 1 by 0 < £ < 1 by the transforms t —x
? _
?
Z'
—
y
T=cot,
y(£,r) =
~ r(jc,o'
h
(5.192a-d)
'
and the dimensional dependent variables and dimensional surface tension f may be scaled to the following dimensionless dependent variables that are denoted by lower case symbols and surface tension that is denoted by a carat (•): q(%, f, T) = .(b ,
,
Aco
,
p{g, f, T) =
7^-5-, pA2co2
E(y/h,t)
*(x,y,t) Aha>
~
(5.192e,f) T pALha)2' (5.192g,h,i)
where A = the wave amplitude, S = the wavemaker stroke, p = fluid mass density and co = 2n/T is the radian wave frequency. Because f is a function of both x and y in Eq. (5.192b), transforming partial derivatives with respect to x must be done with some care. The transformed partial derivatives with respect to x and y are 9W dx
1 '_3_ L
$ dy d (•),
3(.) 3(.) 3y " yfc 3C '
" 32 3 (.) 1 3? 2 2 dx " L2 2
+1
3 (.)
y 1 3 (.)
3y 2
y2/z2 3f 2
2
(5.192J) (5.192k)
2f 3y 3 2 U dy\2 3 2 " y d$ dl-dt; + \y d% ) H2 (•), 2 /3y\2 32y 3 _ y ^ / 3§ 2 _ 3£
(5.1921)
2
(5.192m)
304
Waves and Wave Forces on Coastal and Ocean Structures
The dimensionless transformed WMBVP in Eqs. (5.191) in the fixed {£, f} mapped domain is
3
g
q{S,S,x) = -- L |_3£
P(£,?,r) =
3f
2
3/(^,T) 9
y(?,T) 1 3
9f
0(£,?,r),
/3/(g,r)\ 2 y(f,r) yd. r) V 9? / 1 y2(|)r)
3 2 y(?,r)~ 2
0 < | < 1,
3 9?
9^ /|2
(5.193a)
(5.193b)
&0A J '
3T
2? 3y(g,r) 3 2 V(?,T) 3? 3£3?
3§
2
- (<*r ) j
32 3f 2
*(£,?,r)=0, (5.193c)
0 < ? < 1,
30(£,?,T) 3T
+
30(?,?,T)
?
9?
y(£,t)
1A A 2LL
2
+
2
9£
2
1+
h\
3? = 0,
3?
7(^r) 3^2
L
2-i
'S^.C.T)
y(?,t)/i f3
3y(§,T)30(f,?,T)'
3 2
fdyi^r)^ ' / 2
+ -Ly(^T) k0A
3? 0<£<1,
f = l,
(5.193d)
305
Wavemaker Theories
30(£,?,r)
+
(y£,x)h\
3?
V
3y(g,r)
A
30(£,f,T)
Y{$,T)h2\(dy($,T) L2 3£
= 0,
3£ y(?,T) 0<§<1,
3>(£,?,T)
3£
3§
? = 1,
3£ (5.193e)
0,
0 < £ < 1, ? = 0,
(5.193f)
= 0,
§ = 1, 0 < f < 1 ,
(5.193g)
30(^,C,T) 3? S L 3S(f,T) Ah 3T §=0,
+
• £ 9y(g,T) 30(£,?,T) y(?,r) 3£ SAL 1 3E(f,T) 3? + AHy2(|,T) 3r J (5.193h)
0<1,
where the deep-water wave number ko = co2/g. The initial conditions Eqs. (5.191i,j,k) transform to y(£,0)
= H(|)
3y(g,Q) = 0
0 < f < 1.
(5.193ij,k)
3T
0(1,1,0) = 0 The nonlinear WMBVP in Eqs. (5.191) in the {x,y} physical domain with a moving time- and space-dependent nonlinear free-surface boundary r](x,t) is mapped to a nonlinear WMBVP by Eqs. (5.193) in a fixed {£, f}
306
Waves and Wave Forces on Coastal and Ocean Structures
mapped domain. DeSilva et al. (1996) discretized the rectangular domain and solved the matrix equations by the conjugate gradient method (Atkinson, 1989, Chapter 8.9 or Kershaw, 1978). Details of the computational scheme and several numerical examples may be found in DeSilva et al. (1996).
5.8. Problems 5.1
For the piston wavemaker shown, compute the following: (a) Formulas for the orthonormal eigenfunction coefficients C\ and C2(b) Formula for the average power (P)T per unit width of the wavemaker. (c) Formula for the wavemaker gain function S/H, where S is twice the wavemaker amplitude. (d) Formulas for the amplitudes of the dynamic pressure from orthonormal eigenfunctions
;/ ^__ r -\J2- T
I EZZ
I
h/3
V\ h/3 h/3
77777777777777777777777, Problem 5.1
5.2
For the hinged wavemaker shown that is hinged at mid depth at z = —h/2, compute the following: (a) The wavemaker shape function / (z). (b) The ratio of the amplitude of the dynamic wave pressure from the propagating eigenmode
Wavemaker Theories
307
V< S J > v / * * / *
•
/
X
2
/ —
V
\ \A
_
/• * X
V
,
' i ..'
//
* » • • • • • • • • • • • • •
Problem 5.2
the first evanescent eigenmode 2 at the wavemaker equilibrium position x = z = 0 for a 3 sec wave in 10 feet of water, (c) The wavemaker gain function S/A where A = H/2.
Chapter 6
Nonlinear Wave Theories
6.1. Introduction Nonlinear surface gravity waves may be analyzed by many analytical and numerical methods. The modest review given here is limited to perturbation and semi-numerical methods. In Sec. 2, the classical Stokes theory of the method of successive approximations is reviewed for its historical perspective. In Sec. 3, the Lindstedt-Poincare perturbation method is reviewed to 4th order in the perturbation parameter e = kA. There are two methods that may be applied in the Lindstedt-Poincare perturbation method to suppress resonant forcing at the higher perturbation orders. One is to expand the wave celerity C in a perturbation expansion and the second is to expand the radian wave frequency co in a perturbation expansion. Both methods are applied in this review. The wave celerity C is expanded in the traditional Stokes wave expansion and the radian wave frequency co is expanded in the weakly nonlinear planar wavemaker theory in Sec. 7. Included in the Lindstedt-Poincare perturbation method review is the derivation of Stokes drift in Sec. 6.3.1 where both Eulerian and Lagrangian derivations are reviewed as well as Stokes drift in a 2D wave channel. In Sec. 4, the modern method of multiple scales (MMS) is reviewed to the 3rd order of approximation. In Sec. 5, the semi-numerical stream function theory is reviewed. In Sec. 6, progressive wave breaking is reviewed. In Sec. 7, a weakly nonlinear planar wavemaker theory is reviewed to only 2nd order where the complete 2nd order theory accurately computes the Stokes drift in a closed wave channel. In Sec. 8, an instability in a wave channel is reviewed by the generalized Melnikov method (GMM) and by the Liapunov characteristic exponents. These chaotic instabilities generated by 2D planar wavemakers are called cross-waves and are a strongly nonlinear phenomena 309
310
Waves and Wave Forces on Coastal and Ocean Structures
that may be evaluated by the GMM and by the largest Liapunov exponent that are adopted from analyses by the Floquet theory for the parametric excitation of non-linear dynamical systems. Because the GMM computes Smale horseshoes from the recurring transverse intersections of the stable and unstable manifolds about a separatrix, the WMBVP for chaotic cross-wave instabilities is reviewed by constructing a Lagrangian density for the WMBVP; and then obtaining a Hamiltonian by the Legendre transform from which the fixed points may be computed that are connected by a separatrix that is required for applications of the GMM. The WMBVP that is reviewed in Sec. 8 includes both surface tension and dissipation.
6.2. Classical Stokes: The Method of Successive Approximations Motivated by the experiments and conclusions presented by Mr. Russell in the Report of the Fourteenth Meeting of the British Association for the Advancement of Science, Stokes (1847) developed the classical solution for nonlinear surface gravity waves in order to prove that the speed of nonlinear surface gravity waves (or the velocity of propagation) depends on the amplitude of the wave by applying the method of successive approximations that is now commonly referred to as Stokes perturbation (Dean and Dalrymple, 1991, Chapter 11.2). However, the perturbation technique was developed by Poincare and Van Zeipel (Nayfeh, 1981) more than 50 years after Stokes published his solution. The method of successive approximations will lead to results similar to perturbation, but it is more tedious to apply than perturbation and it has no method for suppressing resonant forcing (viz., perturbing either the frequency co = 2n f or the wave celerity C). Stokes applied a 2D coordinate system with the vertical y axis positive down from the still-water-level (SWL); and recovered the fluid velocity vector from a scalar velocity potential (f>{x, y, t) by positive spatial derivatives according to {u(x,y,t),v(x,y,t)}
=
dd){x,y,t) d
where the ordinary derivative notation d/dx was applied instead of the contemporary partial derivative notation d/dx and d/dy. Stokes proposed the
311
Nonlinear Wave Theories
following boundary value problem (BVP) for a scalar velocity potential 4>{x,y,t) (Stokes, 1847, p. 200):
dcf> p\fd
fl
2
+
dx
fd^2
f± = 0j
(6.1)
(62)
dy2
— =0, when y = h, (6.3) dy dp d
+ Ct,
where c is the wave celerity and C is a dimensional constant. This Galilean transformation removed the explicit dependency on time t and introduced two definitions for wave celerity c (Stokes, 1847, Sec. 3, pp. 202-3). Temporal derivatives in this moving coordinate system were replaced by spatial derivatives according to dp
dp
d(p
d(f>
dt dxm dt dxm where xm = x—ct is the horizontal coordinate in the non-inertial moving axes. Stokes (1847, Eq. (5), p. 200) observed that there is a dimensional similarity
312
Waves and Wave Forces on Coastal and Ocean Structures
in the DFSBC between the Bernoulli constant C and the free surface term gt]; and he replaced the Bernoulli constant with the product C = —gk where k is another constant with the dimension of [Length]. Stokes demonstrated that the correct scaling of the Bernoulli constant C that is based on physics is the free surface amplitude a and not the finite water depth h (Dean and Dalrymple, 1991, Sec. 11.2.2, p. 297). The pressure from Eq. (6.1) and the dynamic free surface boundary condition in Eq. (6.4), respectively, were transformed to d(f) ly
(6.5)
+^
= o, when p = 0. (6.6) dy dy Stokes then applied (but did not state so explicitly) a KFSBC that may be computed from the Stokes material derivative (vide., Chapter 2.2.10) of a Stokes material surface given by F = y — 77 = 0 according to DF Dt
=
D(y - V) Dt
dx
) dx
dy
when y = rj.
Stokes did not employ the Greek letter r\ to represent the free-surface; rather he continued to employ the letter y to represent both the vertical coordinate and the free-surface and noted in a footnote (Stokes, 1847, p. 201) that "The context will always shew in which sense y is employed". For clarity here, however, the Greek letter r\ will be substituted for the letter y employed by Stokes to represent the free-surface. The unknown free-surface r\ was eliminated from Eq. (6.5) by the KFSBC to obtain the following CKDFSBC: dcp d(/) d2
(6.7)
that is exact. At this point, Stokes applied the following short-hand notation for spatial differentiation: ^ =
d ^-=
313
Nonlinear Wave Theories
This notation in the equations that results from the method of successive approximations applied by Stokes is very tedious! Stokes assumed that the scalar velocity potential (j>{x,y) was a series in sin(mx) only and that the free surface rj(x) was a phased-locked series in cos(rax) only that traveled without change of form (Stokes, 1847, Sec. 13, p. 212). Because Stokes had no perturbation parameter e, say, to estimate the convergence of his solution obtained by the method of successive approximations, he evaluated the ratio of the modulus of the second-order correction I2/7I of the water surface elevation to the modulus of the first-order estimate \\i]\ of the water surface elevation in shallow-water where mh
ma(3 + 5 smh2mh + 2 sinh4ra/i) 2
2 sinh m/z sinh 2mh
3>ma 4(m/i)
3
3 4
where m = 2TT/X is the wave number, X is the wavelength, a is the wave amplitude, h is the water depth and the Stokes parameter S is ymny that is a ratio of a measure ofthe free-surface nonlinearities given by ma to a measure of the shallow-water nonlinearities given by (ra/i)3. It may be of value to compare the Stokes parameter S in Eq. (6.8a) with the relatively larger Ursell parameter U (Ursell, 1953b, Eq. (14)) that is defined by
where "/70 is the maximum vertical amplitude and X is a horizontal characteristic length" (Ursell, 1953b, p. 686). For a strictly periodic wave that travels without a change of form, the Stokes parameter 5" may be related to the Ursell parameter U (if the characteristic length scale X = wave length) by U = (2n)2 (—) S.
(6.8c)
Stokes (1847, Sec. 3, pp. 202-3) gave two definitions for wave celerity c (or velocity of propagation in his vocabulary). The first definition was that the
314
Waves and Wave Forces on Coastal and Ocean Structures
mean horizontal water particle velocity is zero at every point in the fluid. If this definition for wave celerity is defined as Ui=u-c,
(6.9)
and if the spatial average (•) in the non-inertial moving coordinate system is given by 1
/-+A./2
((•)> = T /
(•)<**, -A/2
then {U\
=u-c),
(Ui) = (u) - (c), U\ = -c,
(6.10)
where (u) = 0. The second definition for wave celerity was that the average mass flux between two vertical planes perpendicular to the horizontal x axis is zero. If the mean mass flux is defined by (m) = (7—— / \h + r]J_h
udz), I
and this second wave celerity c* is defined by Ui = u — c*, then the spatial averaging (•) in the non-inertial moving coordinate system for wave celerity in a moving coordinate system gives
(^/IH^/>f<^£/4 pUi = (m) + pc*. Equating the spatial average of the two moving coordinate celerities t/, gives the relation between the two celerity definitions as pc + {m) = pc*.
(6.11)
315
Nonlinear Wave Theories
The inviscid mean mass flux (m) is a weakly nonlinear quantity that may be computed from linear wave theory and is called the Stokes drift (Mei, 1989). Stokes made the following classical understatement about proceeding to higher orders of approximation applying the method of successive approximations (Stokes, 1847, Sec. 12, p. 210): "There is no difficulty in proceeding to the higher orders ofapproximation, except what arises from the length of the formulae". In applying the Lindstedt-Poincare perturbation method to this boundary value problem (BVP), Fenton (1985, p. 220) found that the "length of the formulae" were 44 and 83 terms at the fourth and fifth perturbation orders of approximation, respectively! In order to prove that the wave celerity c depended on the wave amplitude a, Stokes (1847, Sec. 12, Eq. (25), p. 210) reduced the dimensional CKDFSBC for deep-water conditions at the third order of approximation to C24>xx ~ g
V = 0,
and then substituted the dimensional, linear, deep-water velocity potential given by (p = —ac exp(—my) sin mx into the right-hand-side (RHS) of the CKDFSBC and obtained for y = 0: g(f)y — c 4>xx & —m a c sinm*, where the RHS is a resonant forcing term at the third approximation. In order to avoid resonance, Stokes made the following very clever assumptions: (1) the unknown 4> on the left-hand-side (LHS) of the CKDFSBC was identical to the linear velocity potential
316
Waves and Wave Forces on Coastal and Ocean Structures
(ii) a known c on the RHS of the CKDFSBC given by the linear deepwater solution c2 = g/m. Substituting these assumptions into the CKDFSBC gave c2 = - ( l + m V ) . (6.12) m Computing the square root of Eq. (6.12) and applying the binomial expansion by Eqs. (2.23) in Chapter 2.3.4 to approximate the radical, he then computed a velocity of propagation that depended on the wave amplitude, viz.,
(l + | « V J .
(6.13)
Stokes (1847, Sec. 13, p. 212) observed that the general procedure for solving the BVP to any order of approximation will result in a phased-locked free surface rj(x) in terms of cos mx only, in a velocity potential 4> (x, y) in terms of sin mx only, and that the value of the velocity of propagation c2 will be determined by equating to zero the coefficients of sin mx in the CKDFSBC (i.e. avoids secular behavior at resonance). Because of its historical position in the theory of nonlinear surface gravity water waves, it is strongly recommended that anyone conducting research in the field of nonlinear surface gravity waves review this classic treatise.
6.3. Traditional Stokes: Lindstedt-Poincare 4th Order Perturbation Solution In the following formulation of the BVP for nonlinear surface gravity waves, the BVP of Stokes from Sec. 2 is followed including the assumptions that the scalar velocity potential (j) will be a series in sin(nxm) only and that the free-surface rj will be a phased-locked series in cos(raxm) only (Stokes, 1847, Sec. 13, p. 212), where xm is a moving coordinate. Rather than applying successive approximations, the BVP is scaled in order to apply the perturbation method that is a more structured technique than is successive approximations for addressing the nonlinearities and for suppressing resonate forcing (Nayfeh, 1981). The vertical z axis is positive up; the fluid velocity vector q is computed from the negative gradient vector of a scalar velocity potential; and only real-valued transcendental functions are evaluated so that an index symbol of i is applied in summations and it is not to be confused with the imaginary unit
317
Nonlinear Wave Theories
Dimensional Boundary Value Problem (BVP) First, the dimensional fluid velocity vector in a fixed coordinate system {x/, z] may be computed from the negative gradient vector of a scalar velocity potential (f> by $ = -V0. (6.14) Following Stokes, the field equation for the total dimensional pressure P is given by the unsteady Bernoulli equation - = — --\V
d2d>
7 T + T T = 0, dxi ozA
l*/l
-h
(6.15b)
The boundary conditions posed by Stokes were that at the horizontal, impermeable bottom there is no flow normal to the bottom. This gives the following kinematic bottom boundary condition (BBC): dd)
— = 0, z = -h. (6.15c) dz Contemporary formulations specify both a kinematic and a dynamic boundary condition at the free-surface (Mei, 1989); and the unknown free-surface displacement is eliminated by combining the two equations to obtain the CKDFSBC. An equivalent derivation of the CKDFSBC requires computing the Stokes material derivative of the pressure at the free-surface (Chapter 2.2.10 and Stokes, 1847) giving DP _dP
dcp 8P
3(pdP _
Dt
dxf dxf
dz dz
dt
'
The boundary value problem is formulated in a fixed inertial coordinate system {xf,z}. In this inertial coordinate system there is an explicit dependency on time. Stokes made a Galilean transformation (Morse and Ingard, 1968) to a moving coordinate system defined by
318
Waves and Wave Forces on Coastal and Ocean Structures
where x/ is the fixed coordinate, C(= X/T) is the wave celerity,1 X is the wavelength, T is the wave period, t is time, and xm is the moving coordinate. In the moving coordinate system, the dimensional velocity potential may be written as 4>(x,f,z,t) =
d
The unsteady Bernoulli equation (6.15a) for the dimensional pressure and the DFSBC in Eq. (6.15d) are transformed to P
d(b
1 - ?
- = -C-^ + Q-gz--\V(t>\2, (6.16a) p oxm I DP f d
7d
d) dx2
dd> dz
3 dx z
,
1 30 3 2 dx dx
= n, P = 0,
d
|V0| 2 , (6.17a)
'Stokes (1847, Sec. 3, pp. 202-3) defined "the velocity of propagation" and gave two definitions.
319
Nonlinear Wave Theories
that does not contain explicitly the unknown dimensional free-surface r\. The DFSBC recovers the location of the dimensional unknown free-surface rj according to 1 90 (6.17b) •|V0| z + 0 , z = Vf /, = 0 . ox Again, the dimensional Bernoulli constant Q in Eq. (6.17b) is a result of the Galilean transformation (Morse and Ingard, 1968) to a non-inertial moving coordinate system. The free-surface boundary conditions are applied at the free-surface z = rj. The two principal problems that are associated with this condition are that: (1) the location of this surface is unknown, and (2) solving the BVP problem on this unknown free-surface will lead to a B VP that does not separate in any of the thirteen separable coordinate systems (Morse and Feshbach, 1953, Chapter 5.1 and pp. 655-666). In order to avoid these two problems, the free-surface conditions are applied at the still-water-level (SWL) z = 0. This is accomplished by expanding the CKDFSBC and the DFSBC in a Maclauren (Chapter 2.3.3) series about the SWL z = 0 according to ,d2(p c- dx2 m=0
m\ dz"
1 ~2
[(M dx
z = 0,
m=0
m\ dz"
(6.18a)
P = 0,
M
gn
dcp dz d
+ Q,
dx
z
= 0,
P = 0. (6.18b)
Boundary conditions are now specified at the SWL at z = 0 and at the bottom z — —h. Solutions must also be spatially periodic in the non-inertial moving horizontal coordinate x. These boundary conditions completely specify the BVP. The total dimensional pressure P may now be linearly decomposed into a dimensional dynamic component p and a dimensional static component ps according to P = P + Ps = P
d
, - pgz.
(6.18c)
320
Waves and Wave Forces on Coastal and Ocean Structures
Dimensionless Boundary Value Problem (BVP) The most important first step in the Lindstedt-Poincare perturbation method (Nayfeh, 1981) is the scaling of the BVP. The selection of the scaling parameters is not unique and different choices will lead to different scaled variables (cf, Chapter 3.5). The BVP is scaled by the following non-unique parameters (dimensional variables denoted by tildes (•)): x = kx,
(z; 1) = (z;h)/h,
(TV) = (f,t)y/§&, (0; P) = U/aJ&i;
P/pga),
C = C/y/i/k, (/?; Q) = (fj/a; Q/ag),
(6.18d-f) (6.18g-i) (6.18j-m)
where a is the linear wave amplitude, k is the wave number (= litIX) and X is the wavelength. The horizontal coordinate is scaled by the wave number k or the wavelength k. This scaling is appropriate because the wavelength scales the spatial periodicity in x. However, scaling by k is in fact scaling by an eigenvalue of the BVP that is a priori unknown. The vertical coordinate z and depth h are scaled by the water depth h that is an appropriate scale for the vertical coordinate z. However, in order for Laplace's equation to remain valid, the depth and wavelength must be of the same order of magnitude. Time is scaled by gravity because waves are a free-surface phenomenon. The dimensional Bernoulli constant Q and the dimensionless free-surface r\ should scale similarly from the DFSBC Eq. (5.17b) and where the dimensional Bernoulli constant Q is scaled similarly to grj. Apparently Stokes (1847, Eq. (50), p. 200) also observed this similarity and replaced his dimensional Bernoulli constant Q with the dimensional product gl where £ is a length scaling constant with the dimensions of [Length]. Because each term in the DFSBC must contribute equally to the physics of the BVP, both the dimensional free-surface fj and the dimensional Bernoulli constant Q should be scaled by the dimensional wave amplitude a (Dean and Dalrymple, 1991, Chapter 11.2, p. 296). The dimensional velocity potential 4> and the dimensional pressure P should also be scaled similarly. Therefore, similar scaling is applied for these 2D variables. Finally, the wave celerity C scales simply as the horizontal length and time scales. The domain of the dimensionless scaled BVP is — 1 < z < arj/h,
—TV <x<
+n (spatial periodicity).
321
Nonlinear Wave Theories
The dimensionless scaled BVP is p =
-C
r30
(ka) 1
dx
d(j)
(kh)1
2
(kh) 2
dx
~~ 7d2d)
30 dz
d2
E
'ka i kh
3 dx
1=0
«! 3z
(6.19b)
d
SH 3x 3x
,
z
= o. (6.19d)
3z 3z 2
3
3d) £
(6.19a)
(6.19c)
(ifcfc)2
1 {ka)2
2 (M)
+ Q,
0,
m\ dz"
m=0
•E
~dz
= 0, ~~ 782
(ka)c {kh)
dcf>
+
(itA)'
G?HD
1 £a
C— H ;^— 3* 2 (fc/,)2
(M»)'
G^G*)'
2;z = o (6.19e)
There are two scaling parameters in Eqs. (6.19). The first is ka = e that is the dimensionlessyree-swr/ace perturbation parameter. The second is kh = q that is the dimensionless finite-water depth scaling parameter. The italic scalar q should not be confused with the bold fluid velocity vector q. For Laplace's equation to be valid, it is obvious from Eq. (6.19b) that kh must be order unity (9(1). This places a restriction on the shallow-water limit for the application of Stokes waves. Goal of Perturbation The goal of the perturbation method is to obtain a sequence of boundary value problems (BVP) that are linear in the unknown variable at a given order em and all of the nonlinear terms are composed of products of known lower order solutions. In order to generate this sequence of linear BVPs in the unknown at order em, each of the scaled dependent variables is expanded in thefree-surface
322
Waves and Wave Forces on Coastal and Ocean Structures
perturbation parameter e = ka according to
Q = J2 *m(m+\)Q,
(6.20a,b)
m=0
V = Yl c"Wi)»7. C = Y, c "Wi)C. m=0
(6.20c, d)
m=0
P = £e m (m+i)/>-
(6.20e)
ra=0
The expanded scaled dimensionless dependent variables are: (1) (2) (3) (4)
the velocity potential (m+i)>, the dynamic pressure (m+\)P, free-surface displacement (m+i)??, the Bernoulli constant (m+i) Q.
In addition, the dimensionless celerity C is also expanded in the freesurface perturbation parameter e. This expansion of the celerity is included because there will be forcing terms appearing at higher order that will lead to secular solutions due to resonant forcing (Nayfeh, 1981). The celerity expansion in Eq. (6.20d) allows these secular terms to be eliminated by the Lindstedt-Poincare method (Nayfeh, 1981). There are two common perturbation expansions that may be applied to eliminate these secular terms: (1) expand the wave celerity C, or (2) expand the wave frequency co(= 2n/T) in a perturbation series. The expansion in the wave C results in changes to the magnitude of the wave celerity at higher perturbation orders; and the expansion in the wave frequency co results in changes to magnitude of the wave frequency at higher perturbation orders. The celerity is given by X =
~T
a> =
I'
where A, is the wavelength, T is the wave period, so that expanding C in a perturbation series will hold the period constant but allow the wavelength to change. The dimensionless BVP is now formulated, scaled, and the dimensionless scaled dependent variables expanded in the free-surface perturbation parameter e = ka. The general solution at each order em may now be developed.
323
Nonlinear Wave Theories
This can be a very laborious exercise, and, therefore, it is very important to first organize the BVP in order to manage the cascade of nonlinear terms that will evolve (vide., Mei, 1989, Chapter 12.2.1). The dimensionless BVP may be summarized as follows: ^ ^ 2 + g - 2 J ( m + l ) 0 = O,
-7T<X<+7t,
— {m+i)4> = 0, oz x(
1 ~
2
-I
< Z < 0,
z = -1,
(6.21a) (6.21b)
32 d \ + TZ5. Tl 2 dx dz ) (m+\)4> = (m+\)F = (m+\)FLHS + (m+\)FRHS,
Z = 0, (6.21c)
with initial condition (m+\)V = ~\C
7,
\-(m+l)Q - (m+l)H,
Z = 0,
(6.21d)
OX
and dimensionless dynamic pressure p: (m+l)P = —1<-
9(m+l)0 „ ^ r (m+l)H + (m+l)Q-
, , _, , (6.2le)
OX
The upper-case symbols in Eqs. (6.21) that represent the nonlinear forcing terms are chosen for the following reasons. The (m+i)FiHS are the nonlinear forcing terms of known lower order solutions that are generated from the Maclauren series expansion in z derivatives of the LHS of Eq. (6.19d); while (m+\)FRHS are the nonlinear terms of known lower order solutions that are generated from the Maclauren series expansion in z derivatives in the RHS of Eq. (6.19d). The ( m +i) n are the nonlinear forcing terms of known lower order solutions that are generated from the last bracketed {•} terms in Eq. (6.19d). Grouping the nonlinear forcing terms in this manner reduces significantly the computations required as well as the potential for errors that may result from a pedestrian term by term differentiation tactic. The general solution (homogeneous solution plus a particular solution) for the dimensionless velocity potential ( m +i)0 at each order e m is (Stokes, 1847, Sec. 13, p. 212) ( m + i ) 0 = homogeneous (n = 1) + particular (n > 1) m+l
m+1
= X I (™+\)<&n = - ^2 (m+l)Bn cosh[nq(z + 1)] sinnx. n=\ n=\
(6.22a)
324
Waves and Wave Forces on Coastal and Ocean Structures
The dimensionless free-surface elevation (m+\)il at each order em is (Stokes, 1847, Sec. 13, p. 212) m+\
m+\
(m+l)V = 2_j ("1+1)^1 = X ! (™+\)An COSAZX, n=\ n=\
(6.22b)
and the nonlinear free-surface forcing terms at each order em are m+l (m+\)F = (m+\)FLHS
+ (m+l)FRHS
— 2_] (m+\)-Fn s i n « X , n=l
(6.22c)
where n = 1 is the homogeneous solution at each order m > 0. Note that the Lindstedt-Poincare expansion for the celerity given by Eq. (6.20d) will guarantee that a resonant forcing term proportional to shut will appear at higher orders m > 0 in the forcing terms (m+\)FLHS from the LHS. This is readily seen from the expansion of the dimensionless celerity C in the first term on the LHS of Eq. (6.19d) that is given by C 2 = iC 2 + e2xC2C + €2(2C2 + liCiC) + e32(xC4C + 2C3C) + €4(3C2 + 2 2 C 4 C + 2iC 5 C) + 0(e5).
(6.22d)
Every forcing term in (m+i)FLHS that contains the product 2iC( m +i)C must be multiplied by d2(f>/dx2 at each order em for m > 0 and this will guarantee a resonant forcing term proportional to sin x. The value of the dimensionless wave celerity (m+i)C at each order em for m > 0 may be determined by requiring that the coefficients of all of the sin x terms in (m+i) F in Eq. (6.21c) are identically equal to zero. This will make the coefficient (m+i)fli for the homogeneous solution undetermined for m > 0 from Eq. (6.21c). The general strategy for solving the BVP in Eqs. (6.21) will be: (1) Compute the coefficients for the particular solutions (m+\)Bn for each harmonic n > 1 from the CKDFSBC Eq. (6.21c) at each order e m for m > 0. (2) Compute the coefficient for the undetermined homogeneous solution (m+l)B\ for m > 0 from the free-surface initial condition Eq. (6.2Id)
325
Nonlinear Wave Theories
by requiring that the coefficient of the cos x term vanish at each order em for m > 0. (3) Compute symbolically certain combinations of the nonlinear terms in (m+\)F, (m+i)H and (m+i)FI in Eqs. (6.21c,d,e), and then substitute algebraically at each order m > 0 for these combinations of nonlinear terms. This strategy of substituting algebraically for combinations of nonlinear terms at each order m eliminates the pedestrian term by term differentiation tactic that may become extremely tedious at each order m > 0 as the number of terms becomes very large and extremely complex (Fenton, 1985, p. 220, or Mei, 1989, Chapter 12.2.1). (4) Non-zero corrections to the dimensionless Bernoulli constant (m+\)Q in Eq. (6.20a) may be computed only at the even perturbation orders ra+1 = 2 , 4 , . . . , etc. because ofthe phase-independent velocity squared products in (m+i)H in Eq. (5.21d); and non-zero corrections to the dimensionless wave celerity (m+\)C in Eq. (5.20d) may be computed only at odd perturbation orders ra + 1 = 1,3,5,..., etc. Linear Operator Xn The coefficient ( m + i)6i ofthe homogeneous solution (OT+i)
2
= Xn{m+i)Bn
coshnq sinnx,
smh.nq}(m+\)Bnsmnx z = 0,
(6.22e)
where the linear operator Xn for \C2 — tanhg (vide., Eq. (6.23g) below) and for z = 0 is given by nq £n = _ . {(w - 1) sinh[(n + l)q] - (n + 1) sinh[(n - l)q]}, z = 0. 2 cosh q (6.22f)
326
Waves and Wave Forces on Coastal and Ocean Structures
With a linear differential operator Xn defined symbolically by Eq. (6.22f) for i C 2 = tanh q, the theory of linear differential operators (Friedman, 1956) may now be applied to compute each of the inhomogeneous coefficients ( m+ i)B„ in Eq. (6.22a) at each nonlinear order m > 0 for each harmonic n > 1 according to {m+l)Bn
= <2+n?±t
z = 0,
n > 1.
(6.22g)
J-n
Maclauren Series Expansions in z Derivatives The advantages of retaining terms together in the forms of ( m +i)^L#s, (m+i)FRHS, (m+i)H and (m+i)FI may now be demonstrated. In order to facilitate the algebraic substitutions of the formulas derived below, the nonlinear products of derivatives in (m+i)FLHS, (m+\)FRHS, (m+i)H and ( m + i)n are grouped in curly bracketed terms {•} that are identified by numerical subscripts that are also in curly brackets {•}{[}• For example, term {9} in Eq. (6.33d) below is given by ,d\(p 3 2qAdz
dx dx
+
d\(j) d dz dz
/n
2
q
dx
+
±\ 2"
dz
(9)
£„ and (m+i)FLHS Maclauren series expansions of the linear operator Xn in Eq. (6.22f) may be computed symbolically and substituted for algebraically at higher orders m > 0 by the following two expansions: viz., for odd-order z derivatives when z = 0 and I > 0 and integer by
qinW+^tA^jBn 2(2£ + l)!cosh x
cos^2l+1\kx)sm(nx),
and for even-order z derivatives when z qW)i2i)\
dzM
(n — l)cosh[g(n + 1)] -{n + l)cosh[(« — 1)] (6.22h) 0 and I > 0 and integer by
JBnSmnX
q{nW+»)iA(2e) 2(2£)!cosh«?
(n - l)sinh[^r(n + 1)] cos(2^(&x)sin(n;c), -in + l)sinh[(n - 1)] (6.22i)
327
Nonlinear Wave Theories
provided that iC 2 =tanh<7,
(6.22j)
where i ^ V--T. For n — 1 (i.e. the homogeneous solutions at each order ra), z = 0 and £ > 0 and integer, the even-order z derivatives of X\ in Eq. (6.22i) are
< 6 - 22k >
-&r = °«
and for I — 0, Eq. (6.22f) reduces to Eq. (6.22k) for n = 1. For example, an even-order z derivative of a Maclauren series expansion of <*£„ in term {7} in ^FIHS in Eq. (6.33b) below may be computed algebraically from Eq. (6.22i) for i = I = k = 1, j = n — 2 and for \t] = if 1, 2$ = 2^2 according to ,/?2 3 2 /
2 2d 2
2
coshg
92<M
itf fd2£2
_
. ,
2 (sinh[3g] — 3 sinh[g]) cos xsin2x
2
[2sin2x + sin4x],
where the following angle sum identities are applied from Tables (2.1 and 2.2) in Chapter 2.4.1: cos 2 xsin2x = ^(2sin2x + sin4x),
sinh[3g] = 3 sinh[g] + 4sinh 3 [#].
+I)FRHS
For (m+i)FRHS, the velocity squared term may be written symbolically as 29m$(-
dn<$>j dm<S>idn<$>j
q*-
i. -1 dx
dx
-L
dz
dz
—
ijq2mBinBj i. 2 cosh[(/ - j)q(z + 1)] cos[(z + j)x] ' + cosh[0' + j)q(z + 1)] cos[(z - j)x] (6.221)
328
Waves and Wave Forces on Coastal and Ocean Structures
where i ^ v — 1. There are two Maclauren series expansions in the z derivatives of Eq. (6.221). The first is an expansion of the x derivative of the velocity-squared term in square brackets [•] in the first term on the RHS of Eq. (6.19d) that is given symbolically by 3 f 2dm<S>idn
Vx"
dx
3 m O ! 3„
+
dx
dz
dz
2
ijq mB, inBj J (/ + j) cosh[(j - j)q(z + 1)] sin[(i + j)x] 1+0-7 ) cosh[(j + j)q(z + 1)] sin[(j - j)x] (6.22m) so that an odd-order z derivative of a Maclauren series expansion of Eq. (6.22m) when z = 0 and I > 0 and integer is given symbolically by d2i+2
21+1
*?.
q2M(2l
l)\dz2Wdx
+ 2
ij(i
2
-
TJ-
q — dx
dx
+
dz
dz
21+1
2
j )q mBinBjk<;, 2(21+ l)\
(i - j)21 sinh[(/ - j)q(z + 1)] sin[(i + j)x] +(i + j)21 sinh[(i + j)q(z + 1)] sin[(i - j)x]
(6.22n)
where i ^ \/—T. An even-order z derivative of a Maclauren series expansion of Eq. (6.22m) when I > 1 and integer is given symbolically by 2t
*£ 2t
q {2l)\
d2M
f 2dm<J>idn<S>j
2l
dz dx
ij(i
-j
\ ;2
q
dx
dx
dm^idn^j dz
dz
)q mBinBjk<; 21 2(2£)!
' (i - j)2£~l cosh[(/ - j)q(z + 1)] sin[(/ + j)x] 1 +(i + j)2i~l cosh[(i + j)?(z + 1)] sin[(i - j)x]\'
l
"
j
where / ^ V—T. When i = j , both of the z derivatives of a Maclauren series expansions given by Eqs. (6.22n, o) are equal to zero. The algebraic substitution of Eqs. (6.22n, o) into (m+i) FRHS significantly reduces the algebra required to compute the nonlinear derivatives for m > 0, and it illustrates
329
Nonlinear Wave Theories
the power of retaining the nonlinear terms together in the forms shown in Eqs. (6.21)! Specifically, when i = j and I > 1 and integer then 3(m) dzldx
f 2dm<S>idn<S>i , dm*i*n*i (/ dx dx dz !«——+ — - ^ {iq)2mBinBi
3z^3x
2 0,
3^ + 1 )
{cos[2/x] + cosh[2/(z + 1)]} (6.22p)
£>0,
for all m, n, i and j where i ^ \/—T. For example, an odd-order z derivative of a Maclauren series expansion of Eq. (6.22m) in {m+\)FRHS in Eq. (6.21c) for I > 0 for any orders m, n and k > 0 where & ^ 2JT/X may be computed algebraically from Eq. (6.22n) so that an odd-order z derivative of a Maclauren series expansion of Eq. (6.22m) in term {2} in ^FRHS i Q Eq. (6.33d) below may be computed algebraically from Eq. (6.22n) for z = I = 0, i = k = m = s = l,j = n = 2 and for i£i = \Y), \(p = \<&\ and2> = 2^2 by 2\C\r) q2
d2
2 3 l 0 32> , d\
= —6q\C\B\
2B2 \A\ cosx{sinhg sin3:c + sinh3g sinx}
6qiC\Bi2B2iAi =
-, {smhg sin4x + 4(smhg + sinh q) sin2x},
where the following angle sum identity is applied: (sinh q sin 3x + sinh 3>q sin x) cos x = j suing sin 4.x + 2(sinhg + sinh q) sin2x. The second Maclauren series expansion in z derivatives of Eq. (6.221) is an expansion of the velocity-squared term in square brackets [•] in the last term on the RHS of Eq. (6.19d). The odd-order z derivative of a Maclauren series expansion of Eq. (6.221) for I > 0 and integer is given symbolically by dU+\ kC2l+\ U+l q2M(2l+\)\dz 2
ijq m>\Bi nBj k^s 2(2£ + l)!
[ 2dm^idn^j [
dx
|
dx
dm<S>idn<5>j dz
dz
u+l
(i - j) sinh[(/ - j)q(z + 1)] cos[0' + j)x] + ( / + j)2t+l s i n h [ ( / + j)q(z + 1 } ] c o s [ ( / _ j)x] (6.22q)
330
Waves and Wave Forces on Coastal and Ocean Structures
and the even-order z derivative of a Maclauren series expansion of Eq. (6.221) for £ > 0 and integer is given symbolically by
q2i(2£)\dzu 2
=
r
9* l
ijq mBinBjktf
2(2£)!
9*
dz
dz
f (j - ;)2€ cosh[(j . _ y.)qr(z + u
l + ( / + j)
1}] C08[( . + j)x]
j
cosh[(i + y)9(z + D] cos[(i - ;)x] J ' (6.22r)
where i ^ V^T. When i = j , Eqs. (6.22q,r) become functions of z only. For example, an odd-order z derivative of a Maclauren series expansion of Eq. (6.221) in (m+i)H in Eq. (6.2Id) for £ > 0 for any orders m,n and k > 0 where & 7^ 2TT/A. may be computed algebraically from Eq. (6.22q) so that an odd-order z derivative of a Maclauren series expansion of Eq. (6.221) in 4H term {9} in Eq. (6.38a) below may be computed algebraically from Eq. (6.22q) with £ = z = 0, i = m = k = s = l and j — n = 2 and for i£i = \r), 10 = 1 <5>i and 20 = 2^2 by j7?_3_ f 2 9 i 0 320 g 3 dz I 3x 3x
3l0 9 2 0 | 3z dz j
= i#i 2^2 1A1 cosxfsinhg cos3x + 3 sinh3g cosx} = \B\ 2^2 1^1 {§ sinhg + 6 sinh3 # + (5 sinh g + 6 sinh3g) cos 2x + \ sinh cos 4x}, where the following angle sum identity is applied: (sinh g cos 3x + 3 sinh 3g cosx) cosx = I sinh q + 6 sinh3g + (5 sinh q + 6 sinh3g) cos 2x + \ sinh q cos Ax.
(hv)\q\2 Perhaps the most complicated symbolic z derivatives in a Maclauren series expansion in (m+i) FRHS m Eq. (6.19d) are generated from the term (»V) \q\2 that is given by (following algebraic substitution of x and z derivatives from Eq. (6.221) into the curly bracket {•} term below)
331
Nonlinear Wave Theories
'
3 , dtQk 2dt
2dm
3
dx
dz
3m
dx
dz
I
dz J
4
ijkq
Bk m Bi n Bj cosh[(( - j - k)q(z + l)](i sm[(i + j + k)x] + j sin[(7 + j + cosh[(i +j+ k)q(z + l)](i sin[(i - j - k)x] - j sin[(i -j+ + cosh[(i - 7 + k)q(z + l)](i sin[(/ + j - k)x] + 7 sin[(i +j+ + cosh[(i + 7 - fc)$(z + l)](i sin[(/ - 7 + k)x] - j sin[(i -j -
k)x]) k)x]) k)x]) k)x])
(6.22s) where i 7^ V—T. An odd-order z derivative of a Maclauren series expansion of Eq. (6.22s) for £ > 0 and integer is given symbolically by (U+l) pSs + q2M(2£
dU+l
\q
dn$>j
dx
dx
+
d
dx
l)\dz
2dm®i
x
,dt$k
U+1
dm^i
dx
, dt<5>k d
+
dz dz
dn4>j
dz
dz
4
ijkq
+l
4(2£ + l)! / (i - j - QV+V sinh[(i - j +(( + 7 + fc)(M+1) sinh[(i + +(i - j + k)V-l+^ sinh[(; +(i + j - k)W+V sinh[(( +
tBkmBinBjpAf^co^
\sx)
- k)q(z + l)](i sin[(( + 7 + *)*] + 7 sin[(i +j- k)x]) ' 7 + fc)$(z + l)](i sin[(i - 7 - *)*] - 7 sin[(i - 7 + *)*]) 7 + k)q(z + l)](i sin[(i + 7 - &)*] + 7 sin[(i + j + k)x]) j - k)q(z + l)](i sin[(i - 7 + k)x] - j sin[(i - 7 - *)*]),
(6.22t) where i ^ •%/—1- An even-order z derivative of a Maclauren series expansion of Eq. (6.22s) for £ > 0 and integer by it PSL
,2£ 2
< ^ ( 2 £ ) ! 3z ^
V4(2£)!
x
2
3
€
^
3x
3 3x
>Bk, \Bi nBj
3€
+
dz dz
2dm<&i
1 —dx
dn<$>j
r^+ dx
dm®i
dn$>j
dz
dz
v(2t) (sx) pA21 s cos
(i - J ~ k) cosh[(; - j - k)q(z + l)](i sin[(i + j + k)x] + j sin[(i + j - k)x]) +(i + 7 + £) z * cosh[(i + j + k)q(z + 1)](( sin[(i - 7 - k)x] - j sin[(i - 7 + *)*]) +(i - j + k)u cosh[(i - j fc)?(z + l)](i sin[(i + 7 - &)x] + 7 sin[(( + 7 + *;)JC]) +(/ + 7 - k)21 cosh[(; + 7 k)q(z + l)](i sin[(( - 7 + k)x] - j sin[(i - 7 - fc)x]) (6.22u)
332
Waves and Wave Forces on Coastal and Ocean Structures
where / ^ V —1- For example, substituting algebraically for Eq. (6.22s) in term {8} in 4FRHS m Eq. (6.33d) below with i=k = £ = m = l,j=n =2 for z = 0 and i
23i0
3 3x 3x
9i0 3 \ / 3z 3 z / \
23i0
320 3x 3X
3i0 320 3Z
3Z
= 11 B\ 2B2 {(1 + 4 cosh 2 - cosh Aq) sin 2x + (2 + cosh 2
cosh(4) = 1 + 8[sinh2g + sinh4*?],
and where i ^ V--1- The majority of the products of derivatives in the nonlinear terms in (m+\)FLHs, {m+\)FRHS, (m+i)H and ( m + i)n may be replaced for m > 0 by algebraic substitutions of Eqs. (6.22f-u); and this algebraic substitution strategy eliminates the tedious pedestrian tactic of term by term differentiation as the number of terms grow with the nonlinear order m (vide., Fenton, 1985, p. 220, or Mei, 1989, Chapter 12.2.1). Linear Solution €° The nonlinear free-surface \F and \H and pressure 1 n terms in Eqs. (6.21) are i F = 0, i # = 0, i n = 0, (6.23a,b,c) so that the linear solutions from Eqs. (6.22) are {cp
= !! = —ifii coshq(z + l)sinx, Xi = 0,
\rj = ifi = \A\ cosx, (6.23d,e) (6.23f)
333
Nonlinear Wave Theories
provided that the dimensionless linear dispersion equation from Eq. (6.21c) for i F = m = 0 and n = 1 is \C2=tanhq.
(6.23 g)
Recall from strategy #4 in the General Strategy for Solving the BVP in Eqs. (6.21) that non-zero corrections to the dimensionless wave celerity ( m+ i)C in Eq. (6.20d) are computed only at odd perturbation orders m + 1 = 1,3,5,..., etc. A zero-mean free-surface displacement may be computed from
if71 (v) = — /
rn
i
r)dx - — /
{m+\)i;ndx
= 0,
(6.23h)
so that substituting Eqs. (6.23b,d,e) into Eq. (6.2Id) for m — 0 and n = 1 gives i<2 = 0. (6.23i) Recall from strategy #4 in the General Strategy for Solving the BVP in Eqs. (6.21) that non-zero corrections to the Bernoulli constant (m+\)Q in Eq.(6.20a)arecomputedonlyattheet;en/>ertarfearionon3fersm-|-l = 2 , 4 , . . . , etc. The linear homogeneous coefficient i B\ may be computed by substituting Eqs. (6.23b,d,e,i) into the initial condition Eq. (6.21 d) for i A \ = 1 according to lAi = l,
iBx = (XC cafoq)-1
=-}—
(6.23j,k)
sinhg by Eq. (6.23g) so that cosh<7(z + 1) ~ ^sinx, (6.231) sinh^ IV = i£i = cosx. (6.23m) The linear dynamic pressure i p may be computed by substituting Eqs. (6.23g,l) into Eq. (6.2le) according to 3i0 cosh[^(z + l)] cosh[<7(z + l)] IP = -lC-r— = cosx = it]. (6.23n) ox cosh q cosh q Second-Order Solution e 1 The nonlinear free-surface terms iF in Eq. (6.21c) are 1 0 = 1
2 2F
= 2FLHS + 2^Rf/5 = X l 2 , 7 " s i n " x ' n=l
z = 0,
(6.24a)
334
Waves and Wave Forces on Coastal and Ocean Structures
where substituting algebraically for ^FLHS the odd-order derivative from Eqs. (6.22e,f,h) in term {2} below with l = Q,i = j = k — n — \,\r}= £1 from Eq. (6.23e) and \
=
2q\CiC
,d2\4>) 3x
+
2
= 2q2C sinx —
\ \f] 3 r q- dz-
a2 c29jl0 2
3x
3i0' 3z
{2}
q\C
sin2x, z = 0. (6.24b) sinh 2q Substituting algebraically for IFRHS from Eq. (6.22m) in term {1} below with i = j = tn — n — 1 and i0 = ii from Eq. (6.231) yields q
-IFRHS
9i0 3x
2q\C sin2x, sinh 2q
+
£l0 dz
(i)
z = 0,
(6.24c)
2
so that 2F
= IFLHS
+ IFRHS
= —2^2Csinx +
_~. sinnx, = / ^ n si n=l / 3qiC \ sinh 2q
sin2x
= 2^1 sinx + 2^2 sin2x.
(6.24d)
The Lindstedt-Poincare expansion of the celerity guarantees a resonant forcing term proportional sinx from IFLHS in Eq. (6.24d). In order for a uniformly convergent solution (i.e. to avoid resonance by applying the Lindstedt-Poincare method) 2C
= 2 * 1 = 0.
(6.24e,f)
Recall from strategy #4 in the General Strategy for Solving the BVP in Eqs. (6.21) that non-zero corrections to the dimensionless wave celerity (m+\)C in Eq. (6.20d) may be computed only at odd perturbation orders m + 1 = 1,3,5,..., etc. The second-order potential 2
24> = Yl2n n=\
= -2J5i cosh[q(z + 1)] sinx - 2B2Cosh[2g(z + 1)] sin2x,
(6.25a)
Nonlinear Wave Theories
335
where the coefficient 2^2 for the particular solution 2 $2 may be computed from Eq. (6.2 lc) by substituting algebraically from Eqs. (6.22e-g) with m — 1 and n = 2 and equating coefficients of the sinnx terms in Eq. (6.24d), i.e. lqXC sinh 2q'
4q sinh q cosh q
£2
2^2 = —
2B2 =
3iC
2^2
-£•2
8 sinh q
, (6.25b-d)
so that 3iCcosh2^f(z + 1 ) . „ 2
7
8
sin2x.
(6.25e)
sinh q
The coefficient 2^1 for the homogeneous solution 2^1 is undetermined from Eq. (6.21c) and it must be determined from the free-surface profile by applying the initial condition given by the DFSBC in Eq. (6.21 d) for m = 1 according to 2V = -lC^--2H ox
+ 2Q,
z=0 (6.25f)
= 2A1 COSX + 2 ^ 2 COS 2 * ,
where the nonlinear free-surface term 2H in Eq. (6.25f) following the substitution of Eq. (6.24e) and the algebraic substitution of Eq. (6.221) in term {3} below with i=j=m=n=l and \(j) = i $ i from Eq. (6.231) is iC
lH =
dx
q dx4>
+
dx
+
3l0 J J {3}
2
1 2 sinh 2q
3x3zjj2)
+
(1 - 2 sinh ,?) cos2x, 2 sinh 2q
z = 0.
(6.25g)
Equating coefficients of cos nx terms in Eq. (6.25f) following the substitution of Eq. (6.25g) yields
2A\
=2Bi = 0,
2M
=
3 + 5 sinh q + 2 sinh q 2 sinh2g sinh 2q
(6.25h-j)
336
Waves and Wave Forces on Coastal and Ocean Structures
so that (3 + 5 sinh2^ + 2 sinh4^) 2?7 = 2?2 = 2 ^ 2 COS 2x =
2 sinh2g sinh 2q
cos2x.
(6.25k)
A zero-mean free-surface displacement, following the substitution of Eq. (6.25k) into Eq. (6.23h), gives the following first non-zero correction to the dimensionless Bernoulli constant Q in Eq. (6.20a): iQ =
1 2 sinh 2q
(6.251)
Recall from strategy #4 in the General Strategy for Solving the BVP in Eqs. (6.21) that non-zero corrections to the Bernoulli constant (m+i)<2 m a y be computed only at the even perturbation orders m + 1 = 2 , 4 , . . . , etc. If the nonlinear BVP is solved in a fixed coordinate system Xf with time as an independent variable, there will be no Bernoulli constant 2<2 to cancel the constant term in 2 # in Eq. (6.25g). In this case, the free-surface r\ may be decomposed into two components; viz., a fluctuating component fj(xf, t) and a constant component fj that may be evaluated to cancel the constant term in 2H in Eq. (6.25f) that is equal to 2Q in Eq. (6.251). The nonlinear pressure term 21! in Eq. (6.2 le) following the algebraic substitution of Eq. (6.221) with i = j = m = n = 1 and \4> = I ^ I from Eq. (6.231) is
-H^?
9if dx
+
9i0
coslx + cosh[2g(z + 1)] 2 sinh 2g
(6.26a)
An interesting consequence of 21! in Eq. (6.26a) is that there is a pressure component that is proportional to cos 2x but independent of depth z; and a pressure component that is proportional to cosh 2q (z + 1) but independent of the moving horizontal coordinate x. The dynamic pressure 2P may be computed by substituting Eqs. (6.25e,l and 6.26a) into Eq. (6.2 le) according to
337
Nonlinear Wave Theories
ox =
/3cosh2^(z + l ) - s i n h 2 ^ \ 1-cosh2?(z + 1 ) cos2x H ^ . . . . (6.26b) 2, \ 2 sinh sinh 2q ) 2sinh2^
The dimensionless pressure component in Eq. (6.26b) given by /
— sinh2g 2 sinh q sinh 2q
cos2x
is a constant over depth and is elaborated in a theory for the origin of microseisms by Longuet-Higgins (1950). A seismometer is installed at the Hatfield Marine Science in Newport, Oregon as a wave recorder based on this theory (ZopfetaL, 1976). The dimensional perturbation amplitude a remains equal to one-half the dimensional wave height H/2 that is computed from the linear e° solution because H = iiji(0) - iv(7t) + ehfjiO) - 2>j(jr)] = 2a + 0,
(6.27)
and the second-order 0(e) correction elevates symmetrically both the wave crest 2»?(0) and wave trough lijin) as illustrated in Fig. 6.1.
Fig. 6.1. Finite-water depth 2nd order Stokes wave for e = 0.2, q = 0.8, S = e/q3 = 0.39.
338
Waves and Wave Forces on Coastal and Ocean Structures
Stokes obtained a dimensionless parameter S in shallow-water given by (vide., Eq. (6.8a) in Sec. 2 where m = k) ka
e
{khf
q3
that is a ratio of the free-surface nonlinearities given by e = ka to the shallow-water nonlinearities given by q3, = (kh)3. Note that the coefficient multiplying the nonlinear pressure term in Eq. (6.19a) ise/q2 and that the two coefficients multiplying the two nonlinear terms on the RHS of Eq. (6.19d) before expanding in a Maclauren series about z = 0 are e/q and €2/q3 = eS, respectively. An interesting feature of the second-order approximation is that in deepwater where q = kh ^> 1, the magnitude of the forcing 2-^2 in Eq. (6.24d) for lC = 0 is, approximately for q S> 1: ~ 3g ,9 3 q , ? ^ 6q 2F2 = - — sech2q = -^== sech^^r ^ -L exp-2q -» 0. 2iC 2 v /tanh^ VI This implies that there is no correction to the velocity potential in deep-water at second-order (i.e. 2
cos
1 2* — - cos 2x. 2
It is important to remember that even though the correction to the dimensionless second-order velocity potential is zero in deep-water (i.e. 20 —>• 0 for ^ » 1), the correction to the dimensionless water surface elevation is non-zero in deep-water (i.e. 2?7 # 0 for g > 1). The difference between a Stokes Fourier series of phase-locked cosine harmonics and a linear Fourier series of dispersive cosine harmonics is illustrated in Fig. 6.2 where the wave number for the phase-locked second harmonic Stokes wave is twice the wave number of the first harmonic (i.e. ki = 2k\). In contrast, the wave number of the dispersive second harmonic wave in a linear
339
Nonlinear Wave Theories
(o=12.5 k
6
fa>,
5 ^
4
*'° /f 1/
kk
"
7lr.
'm
0
co,2 = g*k,*tanh[klh]
0.2
k
h=30ft
.'• £-•0.4 0.6 0.8 k (rad/ft)
1.2
Fig. 6.2. Comparison between Stokes phase-locked and dispersive wave numbers.
Fourier series is computed from the dimensionless dispersion equation (6.23g) and is shown in Fig. 6.2 for h = 30 ft. to be &2 » 2k\. All of the harmonics in phased-locked Stokes waves travel at the same speed. In contrast, each of the harmonics in Fourier series waves are dispersive waves that travel at speeds that depend on the wave number kn for each linear harmonic n. Summary 0{ex) 3iC 2
2V = 2 ft =
-
8 sinh4g
cosh[2q(z + 1)] sin2x,
( 3 + 5 sinh24 +2sinh 4 ^) COS2JC,
2 smb. q sinh 2q
IP =
/ 3cosh2^(z + 1 ) - s i n h i o \ „ 1 - cosh[2(z + 1)] .o—— I cos2x + 2 sinh 2q 2 sinh q sinh 2q
iQ =
l 2 sinh 2q'
C2 = iC 2 =tanh<7.
Third-Order Solution e2 The nonlinear free-surface terms 3F in Eq. (6.21c) are
3 ^ = 2,FLHs + 3FRHS = ^sFnsinnx, n=\
z = 0,
(6.28a)
Waves and Wave Forces on Coastal and Ocean Structures
340
For 2C = 0 and substituting algebraically for the odd-order derivative from Eq. (6.22h) in term {3} below for £ = 0, i = j — k = 1 and j = n = 2 and in term {5} below for £ = 0, j = n = 1 and i = k = 2 and for the even-order derivative from Eq. (6.22i) in term {4} below for i = k = I = 1 and j = n = 2 (or, equivalently for Eq. (6.22k)) yields -IFLHS
+ 2iC
-\qh.C 2
+
\2qiC3C
a3i0
dx2
dx2dz
c2
q dz
(2)
dx2
+
d
\rj d
d i
9i0' dz
+
^JxT
+
(1}
2
W
dz
(14sinh2$-3) 16sinh q x sin3x,
a220 , 2mc
(5)
sinx
q\C
( 9 - lOsinh2,?) 8 sinh4^ (6.28b)
z = 0.
Substituting algebraically from Eq. (6.22m) into term {1} below for/ = m = 1 and j = n = 2; from Eq. (6.22n) for the odd-order z derivative into term {2} below for i = j = k=m = n = s = l and from Eq. (6.22s) into term {3} below for i = j = k = I = m = n = 1 yields —3FRHS
[2C = 0] 3 q dx 1\C d
dx^_ dx
9 ^
+
7di(j)d2(t>
dz
d\(j) dify
q -—— +
J J {3}
341
Nonlinear Wave Theories
Z = 0
— ^ - 4 - ( 3 + 11 sinh2g + 4sinh4<7) sinx 4 sinh q qx
4,
(9-sinh 2 ^)sin3x,
(6.28c)
4 sinh q so that 3
3 ^ = 3FLHs + iFRHs = Y^i^nsinnx,
z= 0
n=\
[9 + 8(sinh4<7 + sinh2*?)]' • smx 2 (4 sinh q) I (27-12sinh2i) . + 2q\C ~ sm3x (4sinh 2 ^) 2 = 3^1 sin* + 3^2 sin2x + 3.F3 sin3x,
(6.28d)
where 3^2 = 0 after equating coefficients of sin2x terms in Eq. (6.28d). In order to eliminate the resonant forcing term 2,!F\ sin* in Eq. (6.28d), the second non-zero correction to the celerity expansion C in Eq. (6.20d) requires that the coefficient of sin x in Eq. (6.28d) vanish, i.e. 3^1 = 0 and 9 + 8(sinh2<7 + sinh4*?) —^ n—^ • (o.iae) \C (4sinh 2 4) 2 Recall from strategy #4 in the General Strategy for Solving the BVP for Eqs. (6.21) that non-zero corrections to the dimensionless wave celerity (m+\)C in Eq. (6.20d) are computed only at odd perturbation orders m + 1 = 1,3,5,..., etc. The third-order potential 3^ from the expansion in Eq. (6.22a) contains both an homogeneous j,B\ and two particular 3B2 and 363 coefficients given by 3C
=
3
n=\
= — 3B1 cosh[g(z + 1)] sinx — 3#2Cosh[2g(z + l)]sin2x - 3 #3 cosh[3q(z + 1)] sin 3x,
(6.29a)
342
Waves and Wave Forces on Coastal and Ocean Structures
where the coefficients 362 and 3,63 for the particular solutions 3 $2 and 3O3, respectively, may be computed by substituting algebraically from Eqs. (6.22e-g) into Eq. (6.21c) with m = 2 and n = 2 and 3 and equating coefficients of sin nx terms to Eq. (6.28d), i.e. 3 F 2 = 0,
3*3=29iC
3B2
= 0,
( 2 7 - 12sinh2) 2
(4 sinh ^)
3S3
2
^
K
: i 3
0C3 = 24,? sinh J , ( 6 . 2 9 b - d )
3-^3
iC(9-4sinh2^)
^3
64 sinh7g
(6.29e,f)
so that iC(9-4sinh2g) 3$3 =
-
64 sinh7#
cosh[3<7(z + 1)] sin 3x.
(6.29g)
The coefficient 3 Si for the homogeneous solution 3<£>i is again undetermined from Eq. (6.21c) and must be determined from the free-surface profile applying the initial condition given by the DFSBC in Eq. (6.2Id) for m = 2 according to 3?? = ~\C-
h32
lH,
z=0
ox (6.29h)
= 3A1 COSX +3A2COS2X + 3 A 3 C O S 3 X ,
where the nonlinear free-surface term 3 H in Eq. (6.29h) following the algebraic substitution of Eq. (6.221) in term {3} below with i = m = 1 and j = n = 2; of the odd-order z derivative from Eq. (6.22q) in term {6} below with £ = 0 and i — j = m — n = I; of Eq. (6.24e) in terms {2} and {4} below and of Eq. (6.28e) in term {1} is 3# =
\C
+
3i0 dx .1}
,3201 dx
I
irn£d\±] 3x3zJ, 4 )
. l H
|
+
2 (4) dx
J {2}
w
2di
dx
fl??lC3 2 20] [ g 3 x 3 z | (5)
+
dz
J J {6)
dx
d\
dz
(3}
343
Nonlinear Wave Theories
J 2< 2^r"2 3 J C 3 Z 2 J { 7 } J 3 x 9 z | { 8 } ' -1 — . -(3 + 14 sinh2^ + 2 sinh4g) cos x 16sinh q H
j-(I - 5 sinh2 - sinh4g) cos 3x. 8 sinh q
(6.29i)
Requiring that 3 A\ = 0; substituting Eq. (6.29i) into Eq. (6.29h) and equating coefficients of cos nx terms in Eq. (6.29h) yields A n 3A1 = 0,
R
iC(3 + 14sinh 2 g + 2sinh 4 g )
3B1 =
,
,*0Q-V, (6.29j,k)
16 sinh g iC(3 + 14sinh2<7 +2sinh 4 a) , . 3<J>i = — ^ — cosh[^( z + 1)] sinjc, 16 sinh ^
_ , (6.291)
3(9 + 24 sinh2,? + 24 sinh4^ + 8 sinh6o) —» 3A3 = 0 • , 3 7 (8 sinh qY
„ nn s (6.29m,n)
and 3A2 = 0, so that 3(9 + 24 sinh2« + 24 sinh4o + 8 sinh6«) 377 = 3?3 = — cos 3*. /0 . ,., „ (8 sinlr g) 2
(6.29o)
For a zero-mean free-surface displacement, following the substitution of Eq. (6.29o) into Eq. (6.23h), gives the following: 3 2 = 0.
(6.29p)
Recall from strategy #4 in the General Strategy for Solving the BVP for Eqs. (6.21) that non-zero corrections to the Bernoulli constant (m+i) Q may be computed only at the even perturbation orders m + 1 = 2 , 4 , . . . , etc. The nonlinear pressure term 3!! in Eq. (6.2 le) following the algebraic substitution of Eq. (6.221) in term {2} below with i = m = 1 and j = n = 2
344
Waves and Wave Forces on Coastal and Ocean Structures
and of Eq. (6.28e) in term {1} below is
-3n ,3i0 dx
iC-
+
29l>920
}_( 2
q
\
dx
d\(j>d2
dx
dz
dz
(2)
3C fcosh[q(z + 1)] cosx \C V coshq cosh[g(z + l)]cos3x + cosh[3g(z + l)]cosx
+3
4 sinh q sinh 2q
(-[9 + 8(sinh24 + sinh4^r)] cosh[(z + 1)] +6cosh3g)cosx + 6cosh[g(z + l)]cos3x
(6.30a)
8 sinh q sinh 2q The dynamic pressure ^p may be computed by substituting Eq. (6.29g,l,p) and Eq. (6.30a) into Eq. (6.21e) according to 3
3/7 =
- 1 c J ] 4 ^ + 3 n + 3!2 n=\
dx
3[(1 - sinh2.? + sinh4tf) cosh[^(z + 1)] - cosh[3^r(z + 1)]] cosx 4 sinh3 q sinh 2q _ 3[8 sinh2^ cosh[(z + 1)] - (9 - 4 sinh2g) cosh[3g(z + 1)]] cos3x. 32 sinh5 q sinh 2q (6.30b) The dimensionless wave number q—kh now depends on the dimensionless amplitude € = ka correct to 0(e 2 ) because from Eq. (6.22d) with Eq. (6.23g) and Eq. (6.28e) and 2 C = 0 Cz = i C
z
| l + 2 ^c) + 0(6 J ), ,2 3
C2 = tanhq ( l + 2 ( £ a ) 2
9 + 8(sinh2^r + sinh4*?) (4 sinh 2 ^) 2
+ 0(0(6.31a)
Proving that the celerity of nonlinear surface gravity waves depends on the wave amplitude a as illustrated in Eq. (6.31a) was the motivation for Stokes (1847) to solve this BVP. At this point, it may be of some historical
345
Nonlinear Wave Theories
interest to review how Stokes determined that the wave celerity (or velocity of propagation) depends on the wave amplitude for nonlinear surface gravity waves because Eq. (6.31a) is a direct consequence of the Lindstedt-Poincare method that was not available to Stokes at the middle of the 19th century. Applying a moving non-inertial coordinate axes with the vertical y axis positive down and the method of successive approximations, Stokes (Sec. 2) and (1847, Eq. (25), p. 210) reduced the dimensional CKDFSBC at third order after applying the deep-water approximations to = -{(4>y)24>yy + 2h fa hx + (4>x)2hx},
y = 0.
where the subscripts denotes partial differentiation with respect to the Eulerian field variable J. Stokes then substituted the dimensional, linear, deep-water velocity potential given by
C2 =
k
i{\+Pa2).
346
Waves and Wave Forces on Coastal and Ocean Structures
Computing the square root and applying the binomial expasion by Eqs. (2.23) in Chapter 2.3.4 to approximate the radical, Stokes then computed a velocity of propagation that depended on the wave amplitude, i.e.
C =
Stokes (1847, Sec. 13, p. 212) observed that the general procedure for solving the BVP to any order of approximation will result in a free surface rj that will contain only cosines (vide., Eq. (6.22b)), a velocity potential 0 that will contain only sines (vide., Eq. (6.22a)), and that the value of the velocity of propagation C2 may be determined by equating to zero the coefficient of sin kx in the CFSBC (i.e. avoids secular behavior at resonance). In order to compare Stokes result with Eq. (6.31a), substitute Eq. (6.18i) and the deepwater approximations for the hyperbolic functions given by tanh q ~ 1 and sinhg ss expq/2 into Eq. (6.31a) and obtain for q ^> 1
C
2_
8
2
1 + 2(kaY
8 /exp4g 16
\exp4q/_
C = ./%[1 + (ka)2] « 1+ V k V k
(kay
that is in agreement with the celerity (or velocity of propagation) obtained by Stokes in Eq. (6.13) in Sec. 6.2 where m = k. The wave height H is twice the amplitude of the linear first-order wave a correct to second-order m = 1 fromEq. (6.27). The first-order wave amplitude a must now be recomputed at third-order m = 2 from
— = r)(0) - rj(n) a = {MO) - iri(ji)} + e{2ri(0) - 2ri(7t)} + e2{3?7(0) - 3 ^ ) } , (6.31b) ~ = 2[l+e23A3] a
+
0(e\
347
Nonlinear Wave Theories
that must be solved simultaneously for the two unknowns q = kh and a/h where h is a known dimensional water depth; i.e. qa y 9 + 8(sinh2g + sinh4g)
koh = q tanh q H h
= 1% 1 + h
h J
(4sinh 2 ^) 2
(6.32a)
qa\2
3(9 + 24 sinh2^ + 24 sinh4g + 8 sinh6^r)
hJ
(8 sinh^) 2 (6.32b)
given dimensional values for f, H and h, and where k~o = (a> = 2n/T)2/g is the linear deep-water wave number. An algorithm from MATHEMATICA™ that computes kh = q and a/h (= "ah" in the algorithm) is given below with an example with given values of g = 32.17 ft/sec2, T = 12 sec, h = 50 ft, and H = 35 ft. (* STOKES THIRD ORDER q & a/h*) TagReal[x_]:=(x/:Re[x]=x;x/:Im[x]=0;)TagReal[q];TagReal[ah]; A33=(3*(9+24*(Sinh[q]'2+Sinh[q]"4)+8*Sinh[q]'6))/(8*Sinh[q]"3)"2; C3Cl=(9+8*(Sinh[q]"2+Sinh[q]"4))/(4*Sinh[q]^2)"2; g=Input["gravitational constant: g"]; T=Input[" Wave Period: T"];h=Input[" Water depth: h"]; H=Input[" Wave Height: H"];koh=(2*N[Pi]*h)/(g*T"2); eql=koh-q*Tanh[q]*(l+q"2*ahA2*C3Cl)=0; eq2=H/h-2*ah*(l+q"2*ah"2*A33)==0; u={q,ah}/.FindRoot[{eql,eq2},{q,N[Pi]/2},{ah,H/(2*h)}]; Print["g=",g,"; "," ","T=",T," ","sec;"," h=",h,";"," ","H=",H]; Print["q=kh=",u[[l]]];Print["a/h=",u[[2]]]; epsq=koh-q*Tanh[q]*(l+q/,2*ah"2*C3Cl)/.{q->u[[l]],ah->u[[2]]}; Print["eps(q=kh)=",epsq]; epsah=H/h-2*ah*(l+q*2*ahA2*A33)/.{q->u[[l]],ah->u[[2]]}; Print["eps(a/h)=",epsah]; OUTPUT g=32.17000000000000; T=12 sec; h=50ft; H=35ft q=kh=0.2464015936149326; eps(q=kh)= -3.91205642880088081 10**-8 a/h=0.1217047716740730; eps(a/h)=-3.5904491957339246 10**-7
348
Waves and Wave Forces on Coastal and Ocean Structures
Figures 6.3 to 6.5 illustrate the water surface elevations from each order m = 1 to 3 for three values of q and e (and, correspondingly, three values of the Stokes parameter S = e/q3) that span three decades of water depths from shallow-(g = 0.3) to deep-(q = 3.0) water. In Fig. 6.4, the distortion effects of the nonlinearities are evident in the trough region of the composite wave between TT/2 < 6 < 37r/2 where the higher order nonlinearities have distorted the smooth trough that is observed in the other two free surface trough profiles in Figs. 6.3 and 6.5.
Fig. 6.3. Shallow-water depth 3rd order Stokes wave e = 0.01, q = 0.3, S = e/q3 = 0.37.
Fig. 6.4. Finite-water depth 3rd order Stokes wave e = 0.2, q = 0.8, S = e/q3 = 0.39.
349
Nonlinear Wave Theories
e Fig. 6.5. Deep-water depth 3rd order Stokes wave e = 0.4, q = 3.0, S = e/q3 = 0.0148.
Summary
0(e)
30 = 3*1 + 3 * 3
iC(3 + 14sinh2(7 + 2sinh 4 ^) cosh[q(z + l)]sinx 16 sinh5 g iC(9-4sinh24) 64 sinh7g k(,h = q tanh g 1 + 2 ?A\ // 5 — =2- 1+ h )
z
qa\
3C
h )
\C
3C
_ 9 + 8 (sinh2,? + sinh4,?) (4 sinh 2 ^) 2
3(9 + 24sinh2,? + 24sinh4,? + 8sinh6,?) (8 sinh 3 ^) 2
3(9 + 24 sinh2,? + 24 sinh4,? + 8 sinh6,?)
cos3x, 3 2 = 0, (8 sinh3,?)2 3[(1 - sinh2,? + sinh4,?) cosh[(z + 1)] - cosh[3g(z + 1)]] COS X 4 sinh q sinh 2q 3[8 sinh2.? cosh[q(z + 1)] - (9 - 4 sinh2,?) cosh[3(z + 1)]] cos3x 32 sinh q sinh2
3V = 3 ft IP
cosh[3g(z + 1)] sin3x,
C = tanhq
x\2 l+2(ka)
9 + 8(sinh 2 4 + sinh4,?) 2
(4 sinh q)
2
+ 0(e 4 ).
350
Waves and Wave Forces on Coastal and Ocean Structures
Fourth-Order Solution e 3 The nonlinear free-surface terms in Eq. (6.21c) are 4 4F = AFLHS+AFRHS
= ^4^/isinnjc,
(6.33a)
z = 0,
n=\
where for 2C = 0 from Eq. (6.24e), for 3C given by Eq. (6.28e) and substituting algebraically four times for the odd-order z derivative from Eq. (6.22h) in terms {4,5,6 and 8} below with Xn defined in Eq. (6.22f); once each for the even-order z derivative from Eq. (6.22i) in term {7} below and Eq. (6.22k) in term {9} below yields -AFLHS
=
+ 3C
2q\C
+
+
h
2
])
dx
dz
2^20 dx2
920 dz
{5}
2^1^ dx2
dl
{6}
{4}
+
2^ 2 10 dx2
(7)
+
dx2
dz
(8)
+ 4
9230
4lC
,23210
dx2dz\{3]
q
[1}
dl
ITT ^T dx2 + dz
(6.33b)
;z = 0 {91
FLHS
•• {2#4Csinjc}(i)
+ { q 1C sin 2x
27 + 78 sinh 2 q + 72 sinh 4 q + 48 sinh 6 9 16 cosh g sinh 7 9
(2)
351
Nonlinear Wave Theories 9 sinh4 q + 8(sinh6 q + 48 sinh8 q) 16 cosh g sinh7 4 J
+ { q i C sin 2x
q i C sin 2x
+
J {3)
51 93 sinh2 + 70 sinh4 q - 35 sinh6 q 0 +~ 8 2 16 cosh g sinh7 q 51
+ 45 sinh2 q + 63 sinh4 ^ - 36 sinh6 4
+ i C sin 4x
16 cosh g sinh7 q (4)
+ { q i C sin Ax
- 9 + 3 sinh2 g + 24 sinh4 9 + 12 sinh6 q 16 cosh q sinh7 9
+ < g iC(sin2x — sin4x)
(5)
Y + 9(sinh2 g + sinh4 q) + 3 sinh6 g 16 cosh g sinh7g
+ \q iC(2sin2x + sin4x)
q 1 C(2 sin 2x + sin 4x)
12 sinh6 q 16 cosh q sinh7g sinh q 48 cosh q sinh7 q
(6)
(7)
+ {0}{9( (8)
24Csinx
+ q 1C sin 2x
q 1C sin 4x
^ + ^ sinh2 ^ + 160 sinh4 q + if2- sinh6 + 8 sinh8 q" 16 cosh q sinh7 - f + 39sinh2 q + 78sinh 4 4 - f sinh64 16 cosh q sinh g
(6.33c)
Substituting algebraically three times from (6.22m) in terms {1, 3 and 6} below; once for the odd-order z derivative from Eq. (6.22n) in term {2} below; once each for the even-and odd-order z derivatives from Eq. (6.22p) in terms {4 and 5} below; twice from Eq. (6.22s) in terms {7 and 8} below and once
352
Waves and Wave Forces on Coastal and Ocean Structures
for the odd-order z derivative from Eq. (6.22t) in term {9} below yields — AFRHS
9 dx
*
\dx
7
v
7
J J {1}
2
+ q
\2ir, d 9 i 0 920 9 9 l 0 920 q 1 dx dx dz dz \< 1 dzdx
+ 2 dx
29l0
1 dx
dx
+
\lV I 2q2
dz
dz
%H%)
1 2^7 d I q dzdx 2
d\(f> 930
q 2
+
930
(2)
3
9 dz2dx
dx
+
+
%0/ 2 ' v dz
_
'{5}
(£)'
J J {6}
q
+ +
2d2(t>
9
dx dx
1
920 9
9i^\2
dz dz
dx )
jd\(j) 9 d\
q1?? 4
2q dz
9
/9if\ \dz
3 i 0 920 2 9 i 0 920 q"—-—— + dx dx dz dz
3i0 3 23i0 9 q ^—— 9x 9x + 3z 9z z = 0,
9
9l0 dx
+
)
(7}
18}
9l0 9z
J J {9} (6.33d)
1%q i C sin 4x "4 ^tftfS =
16 coshg sinh 7 g
I — =- [8 sin 2x (sinh4 + sinh 6 ^) + 2 sin Ax sinh 4 ^] 116 cosh q sinh <7
{2}
Nonlinear Wave Theories
353
? \C 16 cosh? sinh7? X
sin 2x (^ + 96 sinh2? + 32 sinh4? - 52 sin sinh 6 ?Y + sin 4x (27 + 42 sinh2? - 24 sinh4?) '
7 16 cosh ? sinh'q ? 2 [9 sinh ? + 8 (sinh4? + sinh 6 ?)] (6)
+ {0}(4) + {0}(5) 3? \C
4
r
6 =—18 sin 2x(sinh q + sinh ?) 16 cosh q sinh ? — sin4x(sinh 2 ? + 2 sinh 4 ?)]
3? \C
+
J {3}
q \Csin2x
(7)
if,0
r
=— I sin 2x (4 smlr ? - 8 sinh ?) ish<7q sinh 16 cosh sinh7a? + sin 4x (3 sinh2? + 2 sinh4?)]
(8)
- | — =-[sin2x(12sinh 6 ? + 24sinh 8 ?) - 2sin4x sinh 6 ?] 1 [ 16 cosh q sinh ? J r9, "27 _ ? i C sin 2x y + 9 3 sinh2< ? + 1 3 6 s i n h 4 ^ 16 cosh? sinh7? „„ . , * ~, . , s y * [ + 88 sinh 6 ?+24 sinh8? ? i C sin 4x r . t- -. + — =-[ - 45 - 30 sinh2? + 18 sinh4? + 2 sinh 6 ?], 16 cosh? sinh? (6.33e) so that
4F
= AFLHS
+ AFRHS
= ^4-y«
sin
rax,
z = 0
n=\
= 2? 4Csinx 122
27 81 , 4 q\C z 1 16cosh? sinh7? \ 4 + — 2 sinh ? + 24 sinn ^
sinh ? — 16 sinh ? I sin2x
354
Waves and Wave Forces on Coastal and Ocean Structures
H
4lC
f135
n.:-^2„z
=- — 16coshg sinh q \ 2
n ^ u 4 .
9 sinh <7 - 96 sw&fq
40
a \ sinh q sin4x
3
V
^
= 4^1 sinx + 4^2 sin 2.x + 4J3 sin3x + 4^4 sin4;e, (6.33f) where 4^3 = 0 after equating coefficients of sin nx terms in Eq. (6.33f). In order to have a uniformly convergent solution (i.e. to avoid resonance by Lindstedt-Poincare), (6.34a,b) 4C = 4 ^ 1 = 0 . Recall from strategy #4 in the General Strategy for Solving the BVP in Eqs. (6.21) that non-zero corrections to the dimensionless wave celerity ( m+ i)C in Eq. (6.20d) may be computed only at odd perturbation orders m + l = 1,3,5,..., etc. Note also that six of nine terms in 4F1HS and that all of the nine terms in 4 FRHS may be computed algebraically from the formulas in Eqs. (6.22h-t). The tactic of algebraic substitution has replaced a total of 36 separate products of differential terms with 16 algebraic formulas. The fourth-order potential 4^ from the expansion in Eq. (6.22a) contains both an homogeneous 4B1 and three particular 4B2, 4^3 and 4S4 coefficients given by 40 = — 4B1 cosh[g(z + 1)] sinx — 42?2 cosh[2g(z + l)]sin2x — 4fi3Cosh[3g(z + l)]sin3x — 4B4Cosh[4g(z + l)]sin4x, (6.35a) where the coefficients 4^2, 4#3 and 4B4 for the particular solutions 4O2,4($3 and 44, respectively, may be computed from Eq. (6.21c) by substituting Eqs. (6.22e-g) with m = 3 and n = 2,3 and 4 and equating to Eq. (6.33f); i.e. 4F2 =
-qiC (21 , 81 _. ,z, , . , . . 4 — =1 sinh <7 + 24 sinrTg 16 cosh g sinh g \ 4 2 122 sinh6<7 - 16 sinh8^ ) ,
q Aq sinh q £2 = —:—(sinh3 - 3 s i n h ^ ) = , cosh q cosh q
(6.35b)
4^2 4B2 = ——, X2 (6.35c,d)
355
Nonlinear Wave Theories
\C ( ? + T sinh2 + 2 4 s i n h 4 ' ? _ *r sinh6< ? ~ 4^2 =
16sinn
V)
£—«
>
(8sinh 5 ^) 2
(6.35e) 4F3 = 0, 4^4 =
X4 =
o \C
/135 =— I 16 cosh q sinh q \ 2
= 0,
(6.35f,g)
„ . ,? _ . ,4 40 . , f i 9 sinlrg - 96 sinrr # -\ sinh°g 3 (6.35h)
2g 16q sinh3g (5 + 6 sinh2g) (3 sinh5g — 5 smh3g) = , (6.35i) cosh q cosh q 1 C (*¥ ~
4F4 4 4 _
4 fi 3
^ 7 '
4
4 _
9sinh
_ 9 6 sinh4
^ 5
^ + f
2
sinh6
^)
2
(16sinh 4) (5 + 6sinh 4) (6.35j,k)
so that 1C ( T + T 4$2
sinh2
^+
=
24 sinh4
? " W
sinh6
^ - ! 6 s i n h 8 9)
(8 sinh5 q)2 x cosh[2^r(z + 1)] sin2x,
(6.36a)
\C (W- - 9 sinh2 q - 96 sinh4 + f sinh6 4) 4<>4 =
(16sinh 5 g) 2 (5 + 6sinh 2 g) x cosh[4g(z + 1)] sin4x.
(6.36b)
The coefficient $B\ for the homogeneous solution 4i is again undetermined from Eq. (6.21c) and must be determined from the free-surface profile by applying the initial condition given by the DFSBC Eq. (6.2Id) for m = 3 according to tf = -xC-
\-4Q-4H,
Z= 0
OX
= 4A1 cosx + 4A2COS2X + 4A3 cos3x + 4A4COS4X,
(6.37)
where the nonlinear free-surface term 4H in Eq. (6.37) following the substitution of Eq. (6.34a) in term {1} below; the algebraic substitution twice of
356
Waves and Wave Forces on Coastal and Ocean Structures
Eq. (6.221) in terms {3 and 4} below; twice for the odd-order z derivative from Eq. (6.22q) in terms {9 and 10} below; and once for the even-order z derivative from Eq. (6.22r) in term {13} below is 4H=\4C^\ dx
+
wswsy
dx
I
J{i)
+u
q -—— dx dx + dz dz
+ +
7
9]0 d-*,<$)
—3L
92,4>
q
dzdx
3
q2
3 i 0 83^
r
16}
(3}
+ \%cd-*\
{4}
q
dzdx (7}
dzdx J (5)
+ 52,c5* 3z3x
{8}
di
q dz 21) 3
1
dx
dz
dz
']\
{9}
^t) +
2q3 dz
6g3
dx
3z33x
4H
= -{0}(i, 3 (9 + 26 sinh2g + 24 sinh4# + 16 sinh 6 g) cos 2x 64 cosh q sinh g {2}
. 64 cosh
=-[1 + 8(sinh2 + sinh4<j) + cos 4.x ] J {3} 2
- 2 sinh ^(3 + 20 sinh g + 30 sinh4 + 4 sinh6g)
+
+ ( 64 cosh q sinh q
2
h 96 sinh^ + 32 sinh4^ - 52 sinh°g ) cos 2x '21 + I
h21sinh2<7 - 12sinh 4 9j cos4x 14)
357
Nonlinear Wave Theories + sinh 2 ? (6 + 28 sinh 2 ? + 4 sinh 4 ?) 243
sinh ?
+
64 cosh ? sinh ?
102 sinh 2 ? + 100 sinh 4 ? + 4 sinh 6 ? ) cos 2x 243
+
108 sinh 2 ? + 72 sinh 4 ? I cos Ax {5}
[18 sinh 4 ? + 16(sinh6? + sinh 8 ?)]
(1 + coslx)
64cosh? sinh 7 ?
(6)
24(3 sinh 2 ? + 5 sinh 4 ? + 2 sinh 6 ?)
(1 +cos4x)
64 cosh ? sinh 7 ?
(7)
3 ["§ sinh 2 ? + 12(sinh4? + sinh 6 ?) + 4 sinh 8 ?! (cos 2x + cos 4x)
64cosh? sinh ?
{8}
+ I
—^—=- [ ( 9 + 12 sinh 2 ?) + ( 1 0 + 12 sinh 2 ?) cos 2x + cos 4x1 64cosh?a sinh a? L\ 64cosh / J 16(3 sinh 4 ? + 5 sinh 6 ? + 2 sinh 8 ?) 64 cosh ? sinh 7 ?
J 24(sinh 4 ? + 2 sinh 6 ?) 64 cosh ? sinh 7 ?
I
4
(1 +2cos2x + cos4x) {11}
6
8
64 cosh ? sinh 4 ? 16(sinh6? + 2sinh 8 ?) 64 cosh ? sinh ?
= -
{12}
{13}
4 cos 2x
64 cosh ? sinh ? 1 +
AH
(1 + 2 cos 2x + cos 4x)
(1 +cos2x)
4 sinh 8 ? 7
cos2x {10}
4(3 sinh ? + 5 sinh ? + 2 sinh ?)
+
(9(
;
cos 4x
+ —r—
(6.38b) {14}
4 sinh2#(21 + 3 5 sinh2 + 16 sinh4.? + 2 sinh6) 64 cosh q sinh7g
[-81 - 702sinh 2 ^ - 600 sinh4,? + 312sinh 6 ^ + 112 sinh8,?] -\ = cos 2x 384cosher sinh q
358
Waves and Wave Forces on Coastal and Ocean Structures
[-135 + 1116 sinh2 + 1800 sinh4^ +480 sinh6^ + 128sinh8^] 384coshgsinh 7 g
cos4x.
(6.38c) Requiring that 4 A i = 0 and equating coefficients of cos nx terms in Eq. (6.37) following the substitution of Eq. (6.38c) yields 4A1 = 4S1 = 0.
(6.39a,b)
There are no terms in Eq. (6.38c) with a coefficient cos 3x, therefore, 4A3 = 0.
(6.39c)
Equating the remaining two coefficients of cos nx terms in Eq. (6.37) yields 4A2
- 8 1 - 648 sinh2 4 - 1260 sinh4^ -88sinh 6 4 + 1168sinh8 + 384sinh 1 % 384 cosh g sinh q
(6.39d)
and 4A4 =
405 + 2511 sinh2 + 7002 sinh4.? + 10736 sinh6 +9232 sinh8,? + 4160 sinh1 % + 768 sinh12^ 9
(6.39e)
2
384 cosh q sinh g(5 + 6 sinh ^) so that 4T]
= 4^2 + 4^4 = 4A2COS2X + 4A4COS4X,
4ft =
- 8 1 - 648sinh 2 ^ - 1260 sinh4# -88sinh6 + 1168 sinh8g + 384 sinh1 °q
cos 2.x,
(6.39f)
9
384 cosh q sinh ,?
4 £4 =
405 + 2511 sinh2,? + 7002 sinh4,? + 10736 sinh6,? +9232 sinh8^ + 4160 sinh10^ + 768 sinh12^ 384cosher sinh9,? (5 + 6sinh 2 ^)
cos4x. (6.39g)
A zero-mean free-surface displacement, following the substitution of Eq. (6.39f,g) into Eq. (6.23h), gives the following: 9 - 6 sinh2,? - 28 sinh4,? - 4 sinh6,? 4<2 =
64 cosh q sinh q
(6.40)
359
Nonlinear Wave Theories
Recall from strategy #4 in the General Strategy for Solving the BVP in Eqs. (6.21) that non-zero corrections to the Bernoulli constant Q may be computed only at the even perturbation orders m + 1 = 2 , 4 , . . . . The nonlinear pressure terms 41! in Eq. (6.2le) following the substitution of Eq. (6.221) first with i = j = m = n = 2, second with i = j = m = 1 and n = 3, third with i = m = 1, j = 2 and n = 3, and fourth with i = m = I, j = n = 3 are
n ;C
dx 1
mi
dx ) 3
q
+
h dx
dz
dx
dz
(3}
cos2x[l2sinh2^r + 56sinh 4 ^ + 8sinh6^r - ( 2 7 - 12sinh2^) cosh[2q(z + 1)]] 64 sinh6 sinh 2q 18 + (27 - 12sinh 2 ^)cosh[2^(z + 1)]
— cos Ax
+
64 sinh q sinh 2q
12 sinh2<7 + 56 sinh4 + 8 sinh 6 ^ 64 sinh 6 g sinh 2q li 6
64 sinh g sinh 2q.
cosh[2(z + l)]
cosh[4(z + 1)],
(6.41)
where 3C must be computed from Eq. (6.28e). The dynamic pressure may now be computed from Eq. (6.21e) by substituting Eqs. (6.36a,b), Eq. (6.40) and Eq. (6.41) into Eq. (6.21e) according to 4
4P =
L -iCj2^P +4^+4Q dx
n=\
9 - 6 sinh 2 g — 28 sinh4 - 4 sinh 6 g + (6 sinh 2 g + 28sinh 4 g+4sinh 6 <30cosh[2<2'(z + 1)] - 9cosh[4(z + 1)] 32 sinh 6 q sinh 2q
360
Waves and Wave Forces on Coastal and Ocean Structures
36 smh4^ + 168 sinh6^ + 24 sinh8 - (81 + 324 sinh2 + 144 sinh4 q - 632 sinh 6 ? cos 2. -192 sinh8*?) cosh[2(z + 1)] + 36 sinh4 q cosh[4g(z + 1)] 192 sinh8 q sinh 2q - 270 sinh2.? - 324 sinh4 + (6 sinh2 +28 sinh4.? + 4 sinh6) cosh[2(z + 1)] + (405 cos Ax -54sinh2 - 576sinh4 + 80sinh6)cosh[4(z + 1)]
+
+
192(5 + 6 sinh2) sinh8 q sinh 2q (6.42)
J Summary 0(e 3v ): 4>
= 4 0) 2 +
404,
\C ( f + ^ sinh2 + 24sinh 4 ^ - ^
sinh6 - 16sinh8)
(8 sinh5,?)2 x cosh[2(z + 1)] sin2x, iC (if5- - 9 siiih2^ - 96sinh 4 ^ + f sinh 6 ^) i
^
(16sinh 5 ^) 2 (5 + 6sinh 2 ^)
L
x cosh[4g(z + 1)] sin4x, 4?? = AK2 + 4 & ,
4?2 =
- 8 1 - 6 4 8 sinh2,? - 1260 sinh4,? 1 -88sinh 6 g + 1168 sinh8^ + 384sinh10 [ cos 2*, 3 84 cosh g sinh9,?
4?4 =
405 + 2511 sinh2,? + 7002 sinh4^ + 10736 sinh g + 9232 sinh8g +4160 sinh10g + 768 sinh12^ 384cosher sinh9g(5 + 6 sinh2,?) 6
cos4x,
361
Nonlinear Wave Theories
4P
9 - 6 sinh2.? - 28 sinh4.? 4 sinh6g + (6 sinh2.? + 28 sinh4.? + 4 sinh6.?) cosh[2(z + 1)] 9cosh[4^(z + l)] 32 sinh6 q sinh 2q 36 sinh4^ + 168 sinh6.? + 24 sinh8.? -(81 + 324 sinh2 q + 144 sinh4 q - 632 sinh6^ + 192 sinh8.?) cosh[2(z + 1)] + 36 sinh4 q cosh[4(z 1)] 192 sinh8 q sinh 2q x cos 2x
+
42 =
- 270 sinh2.? - 324 sinh4.? + (6 sinh2.? + 28 sinh4,? + 4 sinh6.?) cosh[2.?(z + 1)] + (405 - 54 sinh2# - 576 sinh4^ + 80 sinh6^) cosh[4^(z + 1)] 192 sinh8 q sinh 2^(5 + 6 sinh2) x cos Ax
9 - 6 sinh2.? - 28 sinh4.? - 4 sinh6.? 64 cosh q sinh7g
C2 = tanh$ l+2(£a) 2
9 + 8(sinh2 + sinh4.?) (4sinh24)2
+ 0(e 4 ).
6.3.1. Traditional Stokes: Stokes Drift A mean mass flux derived by Stokes from his second definition for the velocity of wave propagation (vide., Sec. 1) is a weakly nonlinear wave property that may be computed from linear wave theory variables for an inviscid and incompressible fluid. This inviscid theory mass flux is called Stokes drift. Longuet-Higgins (1953, Sec. 2) concluded that mass transport could not be computed solely from inviscid wave theory; and he defined mass transport as a linear combination of Stokes drift computed from inviscid wave theory plus a viscous mean flow. Stoke drift, however, is still referred to as mass transport in many references (Dean and Dalrymple, 1991, Sec. 10.2). Stokes drift may be computed from either Eulerian or Lagrangian quantities; and even though the formulas are identical, the flow distributions are, of course, different. Correct to second order 0(e 1 ), Stoke drift in the Eulerian frame is a time-mean average flow that lies wholly between the wave crest and the wave trough; while Stokes drift in the Lagrangian frame is a time-mean average flow that is
362
Waves and Wave Forces on Coastal and Ocean Structures
distributed vertically between the bottom and the SWL. Stokes drift is a timemean average quantity for periodic waves where the dimensional temporal averaging operator is defined by
((•)>* = ^ f (*)dt.
(6.43)
1
Jo One way to understand Stokes drift is to consider the Eulerian horizontal water particle velocity at the dimensional free surface rj(x,t) correct to secondorder 0(e = kH/2). The dimensional horizontal component of the water particle velocity at the free surface must be estimated from a Maclauren series expansion from the still water level (SWL) at z = 0 according to '' r](x,t) + 0(€2), z = 0, (6.44a) dz where from linear wave theory (LWT) by Eq. (4.37b) and Eq. (4.15a) in Chapter 4.3 Hcocoshk(z + h) u'x,z,t) = cos(A:x — cot), (6.44b) 2 smh kh du(x,z,t) —Hkco cosh k(z + h) „ , . , sm'kx — cot), (6.44c) dx 2 sinh kh du(x,z,t) Hkco sinhk(z + h) cos(£x — oot), (6.44d) dz 2 sinhfc/z Hcosmhk'z + h) w(x, z, t) = sin(Kx — cot), (6.44e) 2 sinh kh u{x,r],t) = u{x,z,t)+
U XZ
r]{x,t) = — cos{kx — cot),
(6.44f)
provided that co1 = gktanhkh,
(6.44g)
and Eq. (6.44a) reduces for z = 0 to Hgk fH\2kco , u{x,r),t) = —2— cos(A;x -cot) + [ — \ —[1 +cos2(^x — cot)] + 0(ez) 2co \ 2/ 2 (6.45) The flow from Eq. (6.45) is not symmetric with respect to the SWL because the positive flow at the wave crest is faster than the negative flow at the wave trough, and the dimensional time-mean average of Eq. (6.45) is (u(x,r,,t))t
/Hk\2 co 2C = l — ) — = e2-<^C,
(6.46)
Nonlinear Wave Theories
363
for 6 <S 1 that is a constant, time-mean, horizontal average drift in the direction of wave propagation. Stokes Drift The dimensional time- and depth-averaged mean horizontal mass flux per unit horizontal width is defined by (Phillips, 1977) ME(L) = lp
UE(L)dz) ,
(6.47a,b)
where UE(L) is the Eulerian (Lagrangian) horizontal component of the dimensional horizontal component of the water particle velocity. Stokes Drift: Eulerian In the Eulerian inertial coordinate frame, Eq. (6.47a) may be computed for an incompressible fluid from
ME = P UdZ
{ \L
+
fo[U^ + (^)z=oZ +
= 0 + (pwz=0)t + 0(e2) = —A 4a>
yH2k Sco
dz
lcos2(kx - cot)) \
It
E C'
by Eqs. (6.44) when UE is given by Eq. (6.44b); and where the total average energy density E is defined by Eq. (4.57c) in Chapter 4.5. Stokes Drift: Lagrangian In the Lagrangian inertial coordinate frame, Eq. (6.47b) may be computed from the dimensional horizontal x component of the dimensional vector mass flux that may be computed from ML = lp\f
j
UEdr,V
UEdz \) .
(6.49)
If the dimensional horizontal and vertical scalar components of the dimensional water velocity vector q are given by Eqs. (6.44b,e), respectively, then the horizontal x scalar component of the dimensional mass flux vector from
364
Waves and Wave Forces on Coastal and Ocean Structures
wdx
2
,4sinh kh
dz
udz ,
cosh2£(z + h){sin2(kx — cot)),
+ sinh2£(z + h)(cos\kx - cot)), dz \ =P
[cosh2k(z + h)]dz\
=
pH2cosiah2kh 16 sinh 2 M (6.50)
Stokes Drift in a 2D Channel with a Planar Wavemaker The dimensionless kinematic wave maker boundary condition (KWMBC) correct to second order e for a planar 2D wavemaker in a non-reflecting wave channel (vide., Eq. (5.5d), Chapter 5.2) contains a time mean forcing term that yields a solution that is numerically equal to but opposite in the direction of the dimensional Stokes drift E/C (Hudspeth and Sulisz, 1991). When the KWMBC in Eq. (5.5d) in Chapter 5.2 is applied on the instantaneous wavemaker boundary at x = %{z/h,t) of a planar wavemaker, the dimensional time-independent nonlinear forcing term proportional to \%(z/h,t) •V<&(x,z,t) is non-zero only for a hinged wavemaker. However, this dimensional nonlinear forcing term becomes non-zero for both piston and hinged wavemakers when the KWMBC is expanded in a Maclaurin series about the wavemaker equilibrium position at x = 0 (vide., Eq. (5.5h) in Chapter 5.2). The complete dimensionless second-order wavemaker solution for a planar wavemaker from Hudspeth and Sulisz (1991) and Hudspeth (1994) is reviewed in Sec. 7 where the scaling parameters are defined for the dimensionless variables reviewed below. Notation Because of the relatively larger number of variables required for the complete second-order WMBVP and because real- and not complex-valued variables are applied here, a change in notation from the linear WMBVP in
Nonlinear Wave Theories
365
Chapter 5.2 is required. The dimensional, complex-valued orthonormal eigenfunction expansion in Eq. (5.38) in Chapter 5.2 is now given by the following real-valued dimensionless orthonormal eigenfunction expansion: i<J>(x,z, r) = —ai^i (z/h) sin(x — T — v) - COS(T + v) ^2 am(t>m (z/h) exp(-/c m x), m=2
where (f>m(Km,z/h) =
-I:
NT2Lm = /
;
coshz Kmh(l + z/h)d(z/h)
m>\
f 2kh + smh Ikh ;m = 1 = 2icmh +4kh sm2/cmrt ;m > 2 4Kmh
provided that K\ = kfor m = I, Km = //cm for m > 2, and ?
.
f kh tanh kh; m = 1 [—KmntanKmn;m > 2
Time-Independent Second-Order Planar Wavemaker Potential Solution 2 * ( x , ^) An interesting feature of the second-order WMBVP is the dimensionless timeindependent potential 2*(x,z) in Sec. 7 that is part of the complete solution the WMBVP correct to second order in the perturbation parameter € = k A (the subscript 2 denotes a second-order solution). This time-independent solution accurately estimates the magnitude of the mean return flow in a wave flume computed by the Eulerian formula in Eq. (6.48). The dimensionless, second-order, time-independent potential 2^(x,z) must satisfy the following BVP (Hudspeth and Sulisz, 1991): V 2 2 * ( x , z ) = 0, dlV(x,z) dz
0,
x > 0 , -h
(6.51a)
x > 0, z = -h,
(6.51b)
366
Waves and Wave Forces on Coastal and Ocean Structures
fr^Ocz) = -a\ cos* Y^ am exp(-/c x) f (0i(O),0 (O)), m 5 m dz
ra=2
*>0, d2xi'(x,z) dx
z = 0,
(6.51c)
a\
->«)"£))• x = 0,
-/z < z < 0,
(6.51d)
where
h (<M0),<M0)) = |0i(O)0m(O) [ 4 + I] ,
M^)-G))=^M)(I)*© =
CO
m=2
0i(») =
sinh/c/i (1 + (.)) : ,
sinKm/i (1 + (.))
(
m > 2,
PmK») =
X (z/A) = [a(z/h) + /3] [t/(z//i + 1 - rf/A) - C/(z//i + b/h)), 9
/
co h + Kmh tanKmh = 0,
m > 1,
/ci = i = V — 1
-(••). d(z/h) Because the dimensionless time-independent solution 2^(x,z) generates a steady, time-independent velocity and not a progressive oscillatory wave velocity, the radiation condition as x —>• +00 may be relaxed to admit only bounded time-independent dimensionless solutions (Wehausen and Laitone, 1960). Because the dimensionless boundary conditions in Eqs. (6.5 lc,d) are inhomogeneous and are applied on orthogonally intersecting boundaries at x = 0 and 2 = 0 , 2 * (x,z) may be linearly decomposed into two dimensionless solutions that may be evaluated separately for each of these two dimensionless boundary conditions; i.e. (-)' =
2 *(jc,z)
= 2^e{x,z)
+ $>f(x,z).
(6.52)
367
Nonlinear Wave Theories
A solution for the first dimensionless time-independent evanescent interaction potential 2^e(x,z) on the RHS of Eq. (6.52) that satisfies Eqs. (6.51a,b) is assumed to be of the form (Hudspeth and Sulisz, 1991)
2^e(x,z)
= aicosx y ^ am exp (-Kmx) m=2
a\ sinx V J . am
x
ex
P (—Km*)
M (I) * ( * ) - * * (j) *•($]•
where <$'m(z/h) = d<£>m(z/h)/d(z/h) coefficients bm and cm are given by
(6 53
->
form > 1 and where the dimensionless
a) tanhh-Km tan icmh 5 ; = 2 tanlr/j + tan2Kmh &> /cm tanh h + tan /cm/j c m = —7T 5 z = 2 tanlr/z + tan 2 ic m h
bm =
Km T~I \> u> \Km + 1) Km Um - l) 7-5 r, 2a> \Km + 1)
(6.54a)
(6.54b)
correct to second-order e 1 ; and where the dimensionless linear solution coefficients am are defined in Eqs. (6.101x,y) in Sec. 7. The second-order dimensionless time-independent wavemaker-forced potential 2^^(x,z) must satisfy a dimensionless homogeneous free-surface condition and the dimensionless time-independent terms in the nonlinear inhomogeneous wavemaker boundary condition in Eq. (6.5Id) given by
~^x—
~2hWl r
U ' X Kh))
x = 0,
—h < z < 0.
+
" " f c T - - "'
(6.55)
A dimensionless solution for 2 * ^ ( x , z ) on the RHS of Eq. (6.52) that satisfies Eq. (6.51a,b) is assumed to be given by the following dimensionless eigenfunction expansion:
2^f(x, z) = Y2 djfj 7=0
\-\
[8j0x + (1 - <5;o) exp (-tijx)],
(6.56a)
368
Waves and Wave Forces on Coastal and Ocean Structures
where 8JQ is the Kronecker delta function (vide., Eq. (2.2) in Chapter 2.2.3) and where fj ( | ) = y/2-8j0co8[njh (1 + z/h)], j > 0,
(6.56b)
provided that the eigenvalues /Xj are computed from Hjh = jit,
j >0
(6.56c)
The dimensionless coefficients dj may be computed from (8jo - /*/) dj 2hco
h
^
m=2
(6.57a) where (••)' = 3/3(z//j)(»») and where the dimensionless depth-averaging operator (•, •) „ in Eq. (6.57a) is
{•>m)z/h = J
(:•)&/»,
(6.57b)
that is readily distinguishable from the temporal averaging operator in Eq. (6.43); and where the inner products in Eq. (6.57a) integrate to dj=2(-l)J
7(2 - 8j0) CO(8JO
— iij)h
( 4 +^) _1 -^3E=I{KI-\-IX ^ 2
amKm(pm(0) +{2Km)2
;>o. (6.57c)
For j = 0, the dimensionless coefficient doh = {loo) l in Eq. (6.57c) that is exactly equal to the dimensionless magnitude of the Stokes drift in Eq. (6.48) or
369
Nonlinear Wave Theories
Eq. (6.50) if the scaling is selected to be 2 E _ pgii k C ~ 8 Ct)
PgH2/S
k/k
2
pgA
cbJgk
1 2co'
(6.57d)
where the tildes (•) denote dimensional quantities; where the wave height H — 2A; and the scaling parameters for the dimensionless second-order WMBVP are given in Chapter 5.2 and in Sec. 7. For generic planar wavemakers where the first-order evanescent eigenseries converge only as fast as am ->• (icmh)~2 ~ [(m — 1)TT]~2, the series for dj in Eq. (6.57c) will converge at least as fast as amKm on (icmh)~5 ~ [(m — 1)TT]~5 for m ^ 2. The convergence for the eigenseries \/fj(z/h) will converge at least as fast as (/ZjA)-3 = ( J T T ) - 3 for j ^ 1. Spatial Distribution of the Time-Independent Eulerian Velocity The spatial distribution of the time-independent Eulerian velocity components computed from the dimensionless second-order time-independent potential 2^{x,z) reveals the importance of the contributions from the first-order evanescent eigenseries. The horizontal time-independent velocity component is given by TI
.
tAj, (x,z)
,
= —€
3 2 * e (*,z) : dx 2a>o
3 2 * / (Jc,z) dx
e-
/ , ^m^n m—2
01 (j-J
+ Km Sinx)
exp(-/c m x) + <
^ \h) ^'m \h) ^m °0SX ~
^ ^
- €J2dJ^j ( r ) [ ^ ° - *j e x P (~NX)]>
( 6 - 58a )
Waves and Wave Forces on Coastal and Ocean Structures
370
Z_
JL-0.2
,t,h
*•
-^—Z—i
1
1
L>.
x/h • 0.5
do ts 1
xlhz
Fig. 6.6. Spatial distribution of the time-independent velocity for full-draft piston wavemaker (Hudspeth and Sulisz, 1991).
0.2
z_
t. h
*-**
v - -
x Ih h
|r • * - * >
-: h 0
I
/
»
2
3
xl h Fig. 6.7. Spatial distribution of the time-independent velocity for fall-draft hinged wavemaker (Hudspeth and Sulisz, 1991).
Nonlinear Wave Theories
371
and the vertical time-independent velocity component is given by V\ji(x,Z) = — €
-
d2Ve(x,z) ; dz
d2Vf(x,z) 6-
dz
^amK"
\2coo m=2
(j>[ [T] 4>m ( r ) ( s i n x -
-
+^ J2 d^j^'j
(i)
ex
K
COS X
m COSX)
exp(-icmx) + m sin x) K
(6.58b)
p i-^jx) •
i=i
The magnitude of the time-independent velocity vector is given by | g v ( * , z ) | = e ~ y £/£(*, z) +
V£(x,z).
and is illustrated in Fig. 6.6 for a full-draft piston wavemaker and in Fig. 6.7 for a full-draft hinged wavemaker for two relative water depths; viz., h/Xo = 0.2 (= relatively shallow-water); and JI/XQ = 0.5 (= relatively deepwater) and where the dimensionless deep-water wavelength A-o = T2 /2n. Figures 6.6 and 6.7 illustrate the importance of the contribution of the firstorder evanescent eigenseries given by Eqs. (6.101h-s) in Sec. 7 to the mean circulation pattern in a closed wave flume; especially near the irregular point (vide., Fig. 5.18 in Chapter 5.7.1) that is located at the intersection of the free-surface and the top of a full-draft planar wavemaker. The far-field time-independent velocity is equal to do that is proportional to the Stokes drift as demonstrated by Eqs. (6.57).
6.4. Method of Multiple Scales (MMS)2 The evolution of an initially two-dimensional, long-crested wave of dimensional finite wave height H and dimensional wave period T in a fluid of constant dimensional finite-depth h may be analyzed by the method of multiple scales (MMS) (Nayfeh, 1981). Applications of the MMS to surface 2
D. E. Vaughan-Makela, MS Thesis, Dept. of Mathematics, Oregon State University, 1997.
372
Waves and Wave Forces on Coastal and Ocean Structures
gravity waves are given by Chu and Mei (1970 and 1971), Davey (1972) and Whitham (1974) (vide., Mei, 1989 for an extensive bibliography). The MMS has also been applied with much success to the Lagrangian variational approach of Luke (1967). An application of MMS to surface gravity waves over wavy bottoms is given by Kirby (1986 a, b); and applications of MMS to cross-waves in both rectangular and circular wave basins are given by Miles and his colleagues (Becker and Miles, 1992 and Miles and Becker, 1988). The dimensional fluid velocity q (x,y,z, t) in a fixed coordinate system {xf, y, z\ may be computed as in Sec. 3 from the negative gradient vector of a dimensional scalar velocity potential 0(Jc, y, z, i) by (vide., Chapter 2.2.7) q (x,y,z,t)
= -V0
(x,y,z,t).
The scaling of the boundary-value-problem BVP is the same as in Sec. 3 with the exception that the dimensional amplitude a scale will be replaced by the dimensional wave height H /2 scale so that the perturbation ordering parameter is now e = kH/2. In addition, the Cartesian coordinate system of Sec. 3 will now be a fixed dimensional inertial coordinate system with the x and y axes horizontal at the still water level (SWL) with a dimensionless y scaled as y = ky where k = 2n/k is the dimensional wave number and X is the dimensional wavelength. Because the coordinate system is now a fixed inertial reference frame, the Bernoulli constant Q in Eq. (6.19e) will now be absorbed into <j>{x,y,z,t). A dimensionless fast time will be defined by r = cot = 2nt where cb = 2n/T. The dimensionless scaled BVP in a fixed inertial coordinate system with q =kh (that should not be confused with the bold velocity vector q) is given by the following: P = {p} + [ps]
f 3$ =
1
/3
{CO
(6.59a) where {p} is the dynamic pressure and [ps] is the static pressure; 32
3 2 cD 32
x,y < |oo|, 0 < z < - l ,
(6.59b)
373
Nonlinear Wave Theories
3$
= 0,
32
3T
1
J
fS
El )--
3$ +
2
3x 9 *
x lq*
i=0
e \ s £!i!_. a / ^! Sz* H /
x
+ "37/ (6.59d)
z = 0,
1 e —
2^2 N f 22 /34> \* U _\9*/)
3r
= 0,
3z 3z 3
+ 97/
3x
3
3
+
3
x,y < |oo|,
&)
co-3r 3$ 3 dy dy
ao> 3
1 e 2^2
e
dz
3$
N
(6.59c)
x,y < |oo|, z - - 1 ,
x,y < |oo|,
/3
fej.
ra^yi
+ Kdz)
J
(6.59e)
z = 0.
Wave Beats To facilitate understanding the concept of slow and fast variables, consider the linear superposition of two, long-crested dimensional progressive surface gravity waves of equal dimensional amplitude a but slightly different frequencies and wave numbers; i.e. $(.X,t) = T]l(x,t) + T]2(x,t)
= a cos (co\t — k\x) + a cos io)2t — k2x),
(6.60)
where the dimensional tilde (•) notation is dropped and where Aco co\ = co — Ak
k\=k-
—,
0)2= 00 +
Aco
Ak r k2 = k + — ,
where Aco and Ak <
374
Waves and Wave Forces on Coastal and Ocean Structures
1 0.5
-0.5
-1 01
5
10
15
20
25
30
time it) Fig. 6.8. Amplitude modulated long-crested wave (wave beats),
very small differences in frequencies Aco and wave numbers Ak as given by t;(x, t) = 2a cos \ [(co\ — 002) t — (k\ — £2) x] x cos U [(&>! + C02) t - (k\ + ki) x] J 1 / Aco -Ak x / cos (cot — kx) = 12a cos 2 V A* = A (e\t, €2x) H cos (cot — kx),
(6.61)
where e\ and ^2 <S 1 and H = 2a. The amplitude modulated, long-crested wave in Eq. (6.61) is illustrated in Fig. 6.8 as a temporal wave record at x = 0. Note that the dimensional amplitude A(e\f,€2x) in Eq. (6.61) is a function of the slow variables e\t and QX. This is the type of amplitude modulated wave that will evolve from the MMS with fast and slow variables (Chu and Mei, 1971). The motivation for choosing e = k~H/2 may now be seen from Eq. (6.61) where the amplitude is the product of a dimensional wave height constant H and a dimensional modulated amplitude A(e\t; €2x). The dimensional constant wave height H may be applied as the perturbation parameter that is proportional to e and the dimensionless modulated amplitude A (ei t; €2x) as the coefficient in the dimensionless boundary value problem for the dimensionless scalar velocity potential
375
Nonlinear Wave Theories
vertical coordinate z will only have a fast scale (cf., Mei, 1989, for an analysis of deep-water waves with slow z) and the y scale will only have slow scales. The only modulated slow variables will be the horizontal coordinates x, y and time t that are defined as: x\ = ex,
X2 = e x,
etc.,
(6.62a)
yi = ey>
yi = £2y,
etc.,
(6.62b)
etc.
(6.62c)
n = ex, x2 = e2r,
Again, note that there is no fast y for an initially long-crested wave or slow z for finite-depth fluids. The following chain rules for differentiation with respect to the independent variables are required: +e2
— = — +e OX
OX\
OX
f__&_ dx2
dx2
+ 0(e3),
(6.63a)
0X2
o2
2(f_ \3x2
3xi dx
32 \
3
dx2dx I
3 3 -, 3 -> — = e— + e2-- + 0(e\ oy oy\ oy2 02
, 32
(6.63c)
,
— = e 2 _ + 0(63), 3r 3yf 3 3 3 9 3 i T - = — + f 7 - + e 2 — + 0(e 3 ), dr dr dTi ox2 2 32 32 32 32 \ 2 / 3 3 — Zj = — Z + 2e — — + e 2 - ^ + 2 - — + 0(e\ 3T 3T 3ri3r \3TJ ox2ox J
(6.63d) , „ (6.63e) (6.63f)
where co = co/Jgk. Note that there are no fast y derivatives in Eqs. (6.63c,d) or slow z derivatives in a fluid of finite-depth. A steady, time-independent current may be included at the zeroeth harmonic (n = 0). The perturbation expansions for the dimensionless scalar velocity potential <&(x,y,z, x) in Eq. (6.22a), the dimensionless free-surface displacement f (x, y, x) in Eq. (6.22b) and the dimensionless dynamic pressure in Eq. (6.2 le) are now given by m+l
<J> (x, y, z, x) = ^ m=0
em ^ «=0
(m+i)>„ (x, y, z, x),
(6.64a)
376
Waves and Wave Forces on Coastal and Ocean Structures m+\
f (x, y, T) = ^
m
e ^
m=0
(m+DVn (x, y, x),
(6.64b)
n=Q
p (x, y, z,t) = Y^ £m(m+\)P (x, y, z, x),
(6.64c)
m=0
where (m+\)
(m+ i)fi„
(m+l)^ (*, J, T) = (m+\)an cos (n0) + (m+i)&„ sin (n@),
sin («©)], (6.64d) (6.64e)
where the fast phase 0 = x — x. The dimensionless velocity potential coefficients (m+\)An and (m+i)B„ and the dimensionless wave amplitude coefficients (m+i)a„ and (m+i)6„ are functions only of the slow variables and are given by (m+Y)An (xi,x2,...;y\,y2,.-.;x\,x2,...),
(6.65a)
(m+\)Bn (x\,X2,...;yi,y2,...;xi,x2,...),
(6.65b)
(m+l)On ( - « l , - « 2 , - - - ; 3 ' l , y 2 , . . . ; T i , T 2 , . . . ) ,
(6.65c)
(m+\)bn (xux2,...;yi,y2,...;
(6.65d)
x\,x2,...).
Substituting Eqs. (6.64) into the scaled BVP in Eqs. (6.59) and applying the slow variable derivatives in Eqs. (6.63) leads to the following sequence of linear problems for each harmonic n at each nonlinear order m: 32 dz2 -inq)
,\ )(m+i)(j)n(x,y,z,x)
= (m+\)G„,
x,y<\oo\,
0
(W+i)G„ = 0 , 9 7T(m+i)
a2
«® ^ 2
(6.66a;)
n = m + 1, x,y<\oo\,
z =-I,
(6.66b)
a\ +
J~
)(m+l)4>n(x,y,Z,x)
= (m+l)Fn,
X,y
< \oo\,
Z = 0, (6.66c)
(m+l)»7nC*:,;y,r)
=
3 u> — (m+l)
ox
z = 0,
t) + (m+\)Hn,
X,y
< |00|,
(6.66d)
377
Nonlinear Wave Theories , (m+l)P(x,y,Z,r)
,
= 0)
x,y < |oo|,
d{m+i}(p(x,y,z,T) — h (m+l)ll,
0
(6.66e)
where Eq. (6.66a') is to be interpreted as m+\Gn = 0 at any order m + l when the harmonic n is equal to the order; i.e. m + l = n. The choice of the symbols for the forcing terms on the RHS in Eqs. (6.66) is motivated by the following: (1) (m+i) Gn represents forcing in the "Governing" field equation (6.66a); (2) (m+i)F„ represents forcing in the "Free" surface boundary condition in Eq. (6.66c); (3) (m+\)Hn represents forcing in Eq. (6.66d) for the free surface that is proportional to the wave height "H"; and (4) (m+i) n represents forcing inEq. (6.66e) for the dynamic "Pressure". Similar to Eq. (6.22e) in Sec. 3, a linear free-surface operator in the fast variables z and r may be defined as
A.) = ^ | ^ + -
(.),
(6.66f)
so that Eq. (6.66c) may be compactly written for expanding in a Maclauren series as <£(m+\)>n = (m+l)Fn,
X,y
< | o o | , Z = 0.
(6.66c')
The introduction of the slow variables in Eqs. (6.62) makes significant changes to the solution of the scaled BVP in Eqs. (6.21) derived in Sec. 3 that applied the Lindstedt-Poincare perturbation method. The governing field equation (6.66a) is now an inhomogeneous ODE with slow derivatives appearing in the forcing ( m +i)G„. More importantly, the coefficients (m+i)An and (m+i)B„ for each dimensionless velocity potential (m+\)4>n(x, y, z, r) for each harmonic n at each nonlinear order m must satisfy the inhomogeneous conditions given by both Eq. (6.66a) and Eq. (6.66c). This leads to a solvability requirement that may be obtained from the Fredholm alternative (vide., Garabedian, 1964, Chapter 10.1; or Guenther and Lee, 1996, Chapter 7.7). Solvability To obtain the solvability condition for the scaled BVP in Eqs. (6.66), multiply the governing field equation (6.66a) by the homogeneous solution at each
378
Waves and Wave Forces on Coastal and Ocean Structures
nonlinear order m + 1 denoted by ( m+ i)<^ and integrate over the dimensionless interval of orthogonality — 1 < z < 0 according to d2(m+l)
£[[•
, .2 JL (") (m+\)
— (m+\)Gn(m+\)
(m+l)4>n
n = 0, 1.
Integrating twice by parts and applying the BBC in Eq. (6.66b) and the CKDFSBC in Eq. (6.66c), the following solvability condition is obtained: (m+l)Fn(m+l)
/
(m+l)Gn(.m+l)
« = 0, 1,
(6.67)
that will generate at 0(e2) the evolutionary equations for the slowly varying amplitudes (m+\)an and (m+\)bn in Eq. (6.64e). Because of the integral condition in Eq. (6.67), the particular solutions to Eq. (6.66a) are computed more efficiently by applying the integral equations that result from the variation of parameters method (vide., Chapter 2.5.3) instead of the method of undetermined coefficients (vide., Chapter 2.5.6). The homogeneous solutions to Eq. (6.66a) for n = 0 are constants; and for n = 1 are [ cosh q(z + l)\ h <* +1 >*i {X | « ! * « ( * + 1 ) }
(6.68a) (6.68b)
with a Wronskian W[u] (vide., Chapter 2.4.3.4) that is given by W[u] = q,
(6.68c)
so that the impulse response convolution kernel h(z — u) is (vide., Chapter 2.5.3) sinhg(z-M) r , , M . h(z — u) = , (6.68d) q and the inhomogeneous solutions to Eq. (6.66a) for each harmonic n at each nonlinear order m may be computed from the following convolution integral: (m+l)(pn(x,y,z,r) =
h{z-u){m+\)Gn{u)du.
(6.68e)
379
Nonlinear Wave Theories
Because the function (m+i) !Fn (z) in Eq. (6.64d) is generally not known because it may have to be computed from Eq. (6.68e), there is no generic equation for £{•) in Eq. (6.66f) as there was in Eq. (6.22f) in Sec. 3. Linear First Order Solution e° For m = 0, iG„ = 0,
iF„ = 0,
\Hn=0
forn>0,
i n = 0.
(6.69a-d)
n= 0 Substituting Eq. (6.69a) into Eq. (6.66a) and Eq. (6.69c) into Eq. (6.66d), the fast, time- and space-independent solutions are constants given by i0o = -iAo(xi,x2,...
;y\,yi,...
;ri,r2,...),
i*7o = lflo = l&o = 0.
(6.69e) (6.69f)
n= l The dimensionless linear wave theory solution is \<j>\{x,y,z, r) = -cos\iq{z
+ 1) (iAicos© + \B\ sin©),
(6.69g)
that satisfies Eq. (6.66a,b) exactly for Eq. (6.69a); and where \!F\(z) = cosh q (z + 1) for m = 0 and n = 1. Substituting Eqs. (6.69b,g) into Eq. (6.66c) gives the following linear dispersion relation for the dimensionless eigenvalue 2cosher — sinhg [1A1 cos© + \B\ sin©] = 0, a> 2 =tanh<j.
(6.69h)
The coefficients \A\ and \B\ must be determined from the DFSBC in Eq. (6.66d). Substituting Eqs. (6.69c,g,h) into Eq. (6.66d) and equating to Eq. (6.64e) with m = 0 and n = 1 yields oo\b\ (o\a\ \A\ = — r - r - , iBi = -——. (6.69i,j) smh q sinh q The dimensionless dynamic pressure i p may be computed by substituting Eqs. (6.69d,g,h,i,j) into Eq. (6.66e) with m = 0 and obtaining 3i0i(x,y,z,T) cosh^(z + l) . \p = a> = (ifli cos 0 + \b\ sin ©). (6.69k) ax coshg
380
Waves and Wave Forces on Coastal and Ocean Structures
An important feature of the MMS is demonstrated by Eqs. (6.69ij), viz., the amplitudes \a\ and \b\ of the modulated wave are functions of the slow variables (vide., Eqs. (6.62)) and may not be determined until third-order(l) 0{€2). The dimensionless linear solutions are 100 = -\AQ{x\,X2,...
;yi,y2,... ;X\,T2,.
..),
im = 0, \
i^li(x,y,r) ip(x,y,z,r) =
= i«i cos© + I&I sin©,
coshtf(z + l ) , ^ , . ^ ( i a i c o s © + iZ?isin@), coshg
where the fast phase 0 = x — r and where \(po and i^i are the homogeneous solutions (m+i)0Q and ( m +i)0p respectively, that are required for the the solvability condition in Eq. (6.67) for n = 0,1. Second-Order Solution el For m = 1,
«=0 2Go = 0,
2^2 = 0:
(6.70a,b)
2
2q co /d\a\ d\b\ . \ 2G1 = -——cosh(z + 1) - — c o s @ + - — s i n © . suing \ 3xi ox\ /
(6.70c)
and y iFn = -2qco —— z —' n=0
— (=Ci0i)
3TIOT
g dz
w 3 J 2 /9l0i\ <7 3 T 2^0 = 0,
9
3x
, /9i0ix
+ az
(6.70d)
381
Nonlinear Wave Theories
, d\a\ d\b\ . 2F1 = — 2qco Idx\ —— cos 0 + dt\ —— sin 0 3qco hiaubi cos20 - (1a2 - xb\) sin2©l, 2^2 = : sinh 2q
(6.70e) (6.70f)
and 2
1
n=0
n=0
2#o = -co 2#1 =
-
3_10n 3TI
3iA 0 dri
d\a\ 3TI
0)1*71 9 2 101 q dzdr
1_ 2q2
3 101 dx
+
3 101 dz
liaj + ibj 2 sinh 2g / '
(6.70g)
sin 0 + —— cos 0 ,
(6.70h)
3TI
2 // 2 = \{\a\ - \b\) cos 2 0 + 2\a\ \b\ sin 2©1
2 sinh2 # - 1 2 sinh 2g (6.70i)
and
n
3 10K l_ dn ~2q~2
! = ^E n=0
2[910iy
|
/3i0ix2
3x
3z
3iA 0 , cosh#(z + 1) /3ifei = -co— 1 —— cos © 3TI 3TI coshg
3iai . ' sin 0 3TI
1
(1a2 - \b\) cos20 + 2\a\\b\ sin2@"
2 sinh 2g
+ ( i a 2 + i/32)cosh2g(z + l)
(6.70J)
n= 0 Substituting Eq. (6.70a) into Eq. (6.66a), the fast, time- and space-independent homogeneous solution is again a constant given by 20o = -2M(x\,x2,... n= l
;y\,yi,...
;TI,T2,...).
(6.71)
382
Waves and Wave Forces on Coastal and Ocean Structures
Substituting Eqs. (6.68d and 6.70c) into Eq. (6.68e), the particular solution to Eq. (6.66a) is -z
2
3iai
„ . dlh . cos 0 H sin 0
3xi
9JCI
sinh[g(z — u)] cosh[q(u + 1)] du sinh[q(z + l)] fd\ai d\b\ . = a>q(z + 1) sinh —q I —— cos© + —— s m 0 \ 3xi dxi (6.72a) Note that Eq. (6.72a) becomes unbounded in deep-water as q ->• oo and this limits the applications of the potential 2
cos 0 : a)
diai
= 0, (6.72b) 3xi d\b\ d\b\ (6.72c) sin©: w-^+ C.^-^ - 0. 3ti 3XI The dimensionless group velocity Cg in Eqs. (6.72b,c) may be defined from a dimensional linear dispersion relation in Eq. (6.69h) given by 3TI
h C„
co = gk tanh kh according to (cf. Eqs. (4.60) in Chapter 4.5) Cq =
deb dk
2q
(6.72d)
1 + sinh2^r 2k
Scaling the group velocity Cg by the same scales as the wave celerity C in Sec. 3 yields Ce =
0)
1 + smhlq)
2 \
sinh2q
(6.72e)
383
Nonlinear Wave Theories
The one-dimensional wave equation for the amplitudes \a\ and \b\ in terms of the slow variables t\ and x\ may be derived from Eqs. (6.72b,c) by noting first that
a 2 flail "Wnlifcij-
r c
a2 flail *3^1i*i|'
(6.72f) (6.72g)
and second that
-2822lTl = - ^
sl
I'f'l.
(6.72h) (6.72i)
so that (Mei, 1989) 2 a2 Jiflll
a2
r2
\\a\\
(6-72J) (6.72k)
that may be generalized to any slow order i ^ */—l according to
dxf \\b\\
(6.721) (6.72m)
8
dxf [\b\J
as well as Eqs. (6.72b,c) according to (6.72n) (6.72o) n= 2 The homogeneous solution to Eq. (6.66a) from Eq. (6.64d) is 2
+ 2B2sm2Q),
(6.73a)
where 2!F2(z) = cosh2g(z + 1) in Eq. (6.64d) and the coefficients 2^2 and may readily be determined by substituting Eq. (6.70e) into Eq. (6.66c) to 2B2 obtain 3a> = „ . t 4 4 cosh[2(z + 1)] 2(p2(x,y,z,r) 8 sinh g x hiai i f e i c o s 2 0 - ( i a f - i f e f ) s i n 2 0 J .
(6.73b)
384
Waves and Wave Forces on Coastal and Ocean Structures
The modulated free-surface elevation at second-order may be computed by substituting Eqs. (6.70g-i) into Eq. (6.66d) and separating harmonics according to diAp
iflf + ifrf
3*1
2 sinh 2q
2?70 = 2 # 0 = 2«0 = — 0)
3 2 (/>i(x,y,z,T) 2»7i (*, y, T) = o> dibi dx\
(6.74a)
h 2/JI = 2»l cos 0 + 2^1 sin 0 ,
3T
= cos©
n = 0,
co q
3ifr\ 3xi /
sin 0 I \3ri
org
3ifli 3xi (6.74b) (6.74c)
2<3 2bil j
l3ri
^dx^Wihl'
3202(*,y,z,r) = co h 2 # 2 = 2^2 cos 2 0 + 3r
ir\2{x,y,x)
. 202Sin20,
3 + 5 sinh g + 2 sinh ^ 2 02
= (l^i -
\b\j
2/>2 = 2\a\ 1&1
(6.74d)
2 sinh 2 g sinh 2q
3 + 5 sinh 2 g + 2 sinh4
n = 2.
2 sinh2<j sinh 2q
(6.74e)
The dimensionless dynamic pressure 2/? may be computed by substituting Eqs. (6.70j, 6.71, 6.72a, and 6.73b) into Eq. (6.66e) with m = 1 according to h
E -z—
IP
„ OX n=0
q(z + 1)
+
+ 2n
sinh[g(z + 1)] / 3 i a i . sin© coshg \ 3xi
3 1A0
„
cosh[g(z + 1)] fd ibi
3TI
cosher
3cosh[2^(z + 1)]
\ 3n
3 ifei —cos®
cos 0
3ri
[2 iai 1&1 sin 2 0 + (\a\ - \bf\
sin 0 cos2©l
2 sinh q sinh 2q 1 2 sinh 2q
\\a\
- 1 Z? 2 )cos2© + 2iai i ^ i s i n 2 0 +{\a\ + i / 3 2 ) c o s h 2 ^ ( z + l)
(6.74f)
385
Nonlinear Wave Theories
Stokes Second-Order Potential To recover the Stokes second-order potential for a fluid for a Stokes wave of finite-depth, recall from Sec. 3 that the free surface profile for a Stokes wave is a phase-locked cosine with \a\ = 1 and ib\ = 0 so that 2A2 = 0 in Eq. (6.64d) and 2 £2 = 0 in Eq. (6.74e). It is easily shown that 2B2 in the velocity potential in Eq. (6.73a) is equal to 2^2 i n Eq. (6.25d) in Sec. 3; and that 2
Third-Order Solution e 2 For m = 2,
a2 d2 d2 1 \-2 dxL2l dy2 dx2dx dyf
EaG„ = V E
3 G 0 = <7 I T l + 7-T 1 1 A 0,
%x\
dyf
(6.75b,c)
3^3=0,
' cosh[q(z+l)][—
sinhg
(6.75a)
d22
n=0
coq2 sin 0
10n
d2
d2
dx2
+ ' dy2
d2 + 2q(z + 1) sinh[q(z + 1 ) ] — ^
iai
dxfl
2cosh[q(z 3G1 =
V ( coq cos 0 sinhg
+ l)]
/ 32
dib
3x2 d2 \
cosh^ + D l ^ + ^ j 101
32
dx2 d\a\
)
0X2
(6.75d)
386
Waves and Wave Forces on Coastal and Ocean Structures 3(Oq2 5^>2
2 sinh q X
" • o r k 9 i a i 1*1 + COS20—- ( i f l j 2 sin 2e) 3xi 3xi
j^)
(6.75e)
and 3
£3^ n=0
+
00 3
q
3TI
1
=£
3i0« dx J
9 2
"V
3
2i/7„w2 d3i
2?7n 3«C , \r\n 3=C , ^-yPn — 29n q az q oz 2co 3 2 / 3 10« 3 20n q 3T \ dx dx
3 \
dx
+ +
dx
3 \
3 i
d i0„x
n=0
+V dz
dz
3 \(pn
2 3' 32 \ , „ ?32 0„ 0 \4>n - 2qco2 2 -qu> \ —~ + 2 3ri3r * \3r2 3r23r 1 /1Vn\2d2x 2 \ q
l
dz2
)
irjnco d2 q2 23
3 F 0 = qa>-
2
lA0
3T
2
+
3 \
dzdr
3 \
+
q
co2\
3
3
2tanh2g
2/
9TI
2a>dx\
( i a 2 + i^ 2 ), (6.75f)
387
Nonlinear Wave Theories
co l«l
3
sinh2g 3TI 1 3
—2q sin 0 2
3 \
31l «al
dx\J 3 qco
812 3iai I co-
dx\J
3TI
4
3TI
2
-co / 8 ( s i n h ^ + sinh ^) + 9 y
2sinh22g
16 sinh 4 x (1a 2 +
3F1 =
•2 -1*1
\b\)\a\
3
3 \
\ s i n h 2 g 3ri +
,
23^ 4
+co
.
#&>
3 ii& »i
1A0 + co-
3xi 2
-2q cos 0
1
4
3T2
3 \
3xi/
3 iZ?i I CO
3r\
2
/8(sinh <7 + sinh < ? ) + 9 16 sinh q x (1a 2 +
1 2 sinh 2g
\b\)\b\ (6.75g)
1 3^2 =
-
2 sinh 2g
(6g cosech2g + 9g + 2 sinh2^)
3ri
2
+ &;g(6 +
^x^
2
(\a\ — i & ) c o s 2 0 + 2(ian&i)sin20
qco sinh 2g 3F3 =
(6.75h)
3 + 5 sinh q + 2 sinh q \ 2
2 sinh q sinh 2q
I
\2qco —
Iqcosinhq
4 sinh 4 g
( i « 2 — ib^)(\a\ s i n 3 0 — ib\ c o s 3 0 ) — 2iai i&i(ifli cos 3 0 + 1&1 sin 3 0 ) _ (6.75i) Note that the last two Maclauren series expansion terms in the equation 3
2 J 3 Fn are identically equal to zero from Eqs. (6.22k,o)! This may be verified n=0
388
Waves and Wave Forces on Coastal and Ocean Structures
by noting that from Eq. (6.22k) that
because o£i>i is proportional to or — tanh q = 0 by Eq. (6.69h); and because
dx J q2 sinh 2q
\ dz 2
da - ib2) cos20 + 2\ai ibi sin20' + {\a\ + ib2)cosh2q(z
+ l)
= fl(2@) + f2(2q(z + I)), then, accordingly, 32
- 2
drdz Evolutionary Equations for \a\ and \b\ The evolutionary equations for the modulated amplitudes iai and ibi may now be obtained from the solvability condition from Eq. (6.67) for n = 0 and 1. These two conditions yield three coupled nonlinear partial differential equations for the modulated amplitudes i Ao, i«i and ibi when the sin 0 and cos 0 terms are separated for n = 1. The three coupled partial differential equations are of the following operator forms (vide., Sec. 3): Xi iA 0 = -X2{ia\
+ ib\),
(6.76a)
°C3 iai + -z— (a iai - ft ibi) = iaiX4 iA 0 , •£3 lh - -r—(a ibi - fi iai) = -ibiX5 3t2
iA 0 ,
(6.76b) (6.76c)
where a and P are constants to be defined and where Xi are partial differential operators that may be nonlinear. The procedure for solving this coupled system is: (i) Neglect transverse modulations in y (i.e. y\,yi, • • •> etc. = 0). (ii) Replace time (or space) derivatives with space (or time) derivatives by Eqs. (6.72b,c).
Nonlinear Wave Theories
389
(iii) Make a Galilean transformation to a moving coordinate system that is defined by C
8 -X\. CO
£l =X\
(6.77)
(iv) Eliminate 1 Ao in order to obtain two evolutionary equations for i^i and ib\. This procedure may now be applied. (i) For no transverse modulations in y (i.e. y\,yi,---, etc. = 0), the solvability condition in Eq. (6.67) for n = 0 from Eqs. (6.75b,e) with 30Q = i AQ, a constant, gives
'
t
2
9
2 3
dxf
lA 0
3rf # 2tanh2g
<x> \ 2/
3 3ri
<7 9 2codx\
(io? + ife?). (6.78a)
For no transverse modulations in y (i.e. yi, y 2 , . . . , etc. = 0), the sin 0 component of the solvability condition in Eq. (6.67) for n = 1 from Eqs. (6.75d,g) with z4>\ = cosh[g(z + 1)] is q 32 2codx2
co d1 + qcol 23^2 3xi3ri
8 ( s i n h 4 ^ + sinh 2 ) + 9 —co
4
16sinh g
1 2sinh 2 2g
\a\
x(ia2 + \b\) d\a\ dxj
+co~!-^- H
= - \a\\
co2
3
lsinh2g 3TI
Co d\b\ ^ —sinhg 3^2 3
+ dx\
lA 0 .
(6.78b)
390
Waves and Wave Forces on Coastal and Ocean Structures
For no transverse modulations in y (i.e. yi,y2,---, etc. = 0), the cos 0 component of the solvability condition in Eq. (6.67) for n = 1 from Eqs. (6.75d,g) with z4>\ = cosh[q(z + 1)] is q
+00
co d2
d'
+ qcoz 3XI3TI loo dx\ 2 3rj 4 2 ' 8 ( s i n h ^ + sinh ^) + 9 1 16 sinh 4 g 3 \b\ —co-3t2
= lM
3
co
Ca
3 \a\
sinhg dx2 3 ,
sinh2q dt\
\o\
2 sinh 2 2q
, lA 0 .
dx\
(6.78c)
(ii) From Eq. (6.72j,k), 23
co
2
(.) 2
3T
23
= ct
2
(.)
(6.78d)
dx\
and from Eq. (6.72f,g),
32W
3>)
n
(6.78e)
3xi 9^1 and because Eqs. (6.72b,c) imply that
3(.)
n — Ln
3t2
3(.)
(6.78f)
3^2 '
then Eq. (6.78b) reduces to <7 2
C
qco Co
2
dx\
2
8(sinh\ + sinh ^)+9 -co
4
16 sinh q x ( i a 2 + iZ?2) 3 / + — (wifli 3t2 V
1
\a\ 2
2 sinh 2q
a> - ^ - j — - \t>\ sinhg
391
Nonlinear Wave Theories
3
CO
= -
\a\
d
llsinh2g 3TI + dx\
(6.78g)
iM,
and Eq. (6.78c) reduces to
I \2co
qa> Ce I —» 8 ) dx\
H
2co
'8(sinh4<7 + sinh2<7) + 9
1
16sinh 4 g
2sinh 2 2^
-co
ih
x(ia2+i^2) •— [coim 3t2 \
\b\ \ I sinh2g dx\
— — — ifli suing
(6.78h)
[ \AQ. dx\
(iii) Transforming to a moving coordinate system given by Eq. (6.77) converts Eq. (6.78a) to an ODE in £i giving 9 2
\ ( c^ 2 )_
/
1A0
* *- * s*r CP
2 tanh 2
- ^2 )/ + f2co\ - [ ^3§i( i a 2 + ife2)' (6-78i)
where the linear operators £\(») and 0C2W in Eq. (6.76a) may be defined from Eq. (6.78i) by
£X(.) = q{q
=c2(.)
C,
2 ~C g)-3(.), s
3£f
2tanh2« ?
2/
2<w J 3£i '
392
Waves and Wave Forces on Coastal and Ocean Structures
and Eq. (6.78g) now becomes
la
Cl
i „
32
qco2Cn ,
„
_
s \2co 2co J 9^2 8( sinh4 +sinh 2 g) + 9
-co
1 2 sinh2 2q
16 sinh q x
lfll
( l « i + i*l)
3 / a) , +^— « i « i - ^-r—ibi 0X2 V sinhg = i«i
wCp
1 3iA 0
sinh 2^
J 3*1 '
(6.78J)
and Eq. (6.78h) now becomes <7
'8(sinh 4 ^ + sinh2^r) + 9 —
16 sinh g x
fflCo - il »*l
sinh2g
1*1
2 sinh 2^
dfli+i*i)
6t2 V =
^
+1
suing 3lA 0
(6.78k)
3£l '
where the nonlinear operator £•$(•) and the //wear operators £.${•) and ^5 in Eqs. (6.76b,c) may be defined from Eqs. (6.78j,k) by o
'8(sinh4(7 + sinh2<7) + 9
1
16 sinh q
2 sinh2 2q
— CO
(i«l + lb\),
393
Nonlinear Wave Theories
£4 =
coCo -1 sinh2g
c£5_
a = co, fi =
l s i n h 2 ^ + 1 J 3|i'
co sinhg
(iv) In order to eliminate 1 AQ and to obtain two evolutionary equations for \a\ and \b\, Eq. (6.781) may be integrated once to obtain
Ce
d \A0
ar\+q_\ 2 ) 2coJ ,
2tanh2g
3^1
,2
2
«(*-<#
(6.79a) where /$ is a constant of integration. Substituting Eq. (6.79a) into Eqs. (6.78j,k) gives
^- - r
-co
1
- <7«2Cg
5(sinh4^ + sinh2^r) + 9 16sinh4g x
9
sinh2g
hi.Xi,Ti),
1 2sinh22
lfli
(i«2+i^2)
(
+ 3r2 V i«i
—?
Cg
Itznhlq q(q- Ch
sinhg
lfc q_
2
2co
(i«? + Xb\) (6.79b)
394
Waves and Wave Forces on Coastal and Ocean
Structures
^ 4
2
8(sinh ^+sinh ^) + 9 16sinh4g
1 2sinh22g
1*1
x(lfl?+i&?) 9
3t2 wCo
1*1
+1 L sinh 2q —
(
u o»lb\
W
7~r—\a\ svahq
q Ce 2tanh2 q(£l-q)
fy(xi,rt),
2/
2co (i*i +
i*i)
(6.79c)
that are the evolutionary equations for the modulated amplitudes \a\ and \b\.
6.5. Stream Function Solutions The Dean Stream Function wave theory (Dean 1965) may be applied to analyses of both symmetric (theoretical) and asymmetric (real) ocean wave profiles. Dalrymple (1974) modified the Dean Stream Function algorithm for nonlinear waves propagating on a shear current where the shear current velocity was modeled by either a linear or a bilinear steady current profile. Computational procedures were modified by adding two Lagrangian constraints to the iterative numerical algorithm given by Dean (1965) that resulted in convergence of the solution to a specified wave height H and to a zero-mean free surface displacement (r?> = 0. For a design wave condition specified by the wave height H, the wave period T and the water depth h, the field variables are computed by a finite Fourier cosine series. The solution chosen satisfies the governing field equation, the bottom boundary condition, and the kinematic free-surface boundary condition exactly. The unknown Fourier coefficients are then computed iteratively such that the dynamic free surface boundary condition errors are minimized in a best least-squares sense. Von Schwind and Reid (1972) develop a stream wave function wave theory that is similar to Dean (1965). The principal difference between the two
Nonlinear Wave Theories
395
theories is that Von Schwind and Reid applied a conformal transformation of the coordinates for the boundary value problem from the complex (x + iz) plane to the complex (4> + i ir) plane. The dimensionless boundary value equations that result from the conformal transformation may be solved by a Fourier cosine series and the Fourier coefficients are determined through a numerical iterative process similar to that applied by Dean (1965). Benzi et al (1979) formulate a stream function wave theory from a variational principle, but no numerical results are given. Chaplin (1980) reformulates the problem solution technique by choosing as unknowns the wavelength A. and the free surface wave profile r\ at equal discrete phase positions between the wave crest and trough. A set of orthonormal functions are constructed from the free-surface stream function by a GramSchmidt process (Hildebrand, 1965, Chapter 1.13). A generalized Fourier cosine analysis is applied to compute the stream function coefficients. The numerical iterative process is also carried out through a minimization of the dynamic free-surface boundary condition error. Cokelet (1977) obtained an exact solution by means of a series expansion in terms of a perturbation parameter. Cokelet extended his series to higher order terms and summed with Pade approximates. Chaplin (1980) and Cokelet (1977) demonstrate that the wave celerity and other integral properties of surface gravity waves do not increase monotonically with wave steepness, but reach a maximum for waves slightly lower than the maximum steepness. Hudspeth and Slotta, (1978) demonstrate this non-monotonically increasing property from a stream function algorithm (vide., Hudspeth and Slotta, 1978, Figs. 7-8). Rienecker and Fenton (1981) apply a finite stream function Fourier series similar to the Dalrymple (1974) modified Dean Stream Function. They solve the nonlinear equations for the Fourier coefficients by Newton's method in contrast to the linear-Taylor differential correction method that is applied by Hudspeth and Slotta (1978). Their stream function solution overcomes previous difficulties encountered in obtaining convergent solutions for steep symmetric waves in deep-water by incorporating a term proportional to sech(nkd) in the stream function expansion. A copy of the Fenton numerical algorithm written in FORTRAN and the definition of the terms (nkd) may be found in Fenton (1988). The stream function solution assumes that the surface gravity wave profile propagates without a change of form at a constant speed C in an irrotational
396
Waves and Wave Forces on Coastal and Ocean Structures
7777777777777777777777777777777777 Fig. 6.9. Definition sketch for one-half of a symmetric nonlinear wave propagating at constant speed C on a steady, uniform current UQ (Huang and Hudspeth, 1984).
flow of an inviscid fluid. It is then possible to make a Galilean transformation to a coordinate system that moves with the speed of and in the same direction as the long-crested wave profile. Figure 6.9 illustrates one-half of a symmetric nonlinear periodic wave propagating without change of form on a steady, uniform current Uc- In this non-inertial moving reference frame, the wave profile does not change shape and the motion is steady relative to this steady moving coordinate system. The dimensional water particle velocity components u(x,z) and w(x,z) may be defined in terms of a dimensional scalar stream function x/f(x,z) according to di//(x,z)
u(x,z) + Uc - C = —
dz~~ '
w(x,z) =
djf(x,z) dx
(6.80a,b)
where C = k/T is the constant dimensional wave celerity, and k,T are the dimensional wavelength and period, respectively. The dimensional boundary value problem for the irrotational motion of an inviscid fluid is given by the following: V2f(x,z)
= 0,
d^{x,z) dx
\x\ < 00,
= 0,
-h
|*| < 00, z = —h.
(6.81a) (6.81b)
397
Nonlinear Wave Theories
di{r(x,z) di/r(x,z) dx ) I \ dz
dx r\(x) +
2
/djr(x,z)
V
2g
+
dz
|*| < oo, z — y{x),
d\jr(x,z) dx
\x\ < oo,
2g
= Q,
(6.81c) (6.8 Id)
z = rj(x), (6.81e)
ijr(x + L,z) = yjr(x,z),
where Q is the dimensional Bernoulli constant, g is the dimensional gravitational acceleration and the dimensional two-dimensional Laplacian operator V 2 (.) = d2(,)/dx2 + d2(»)/dz2 (vide., Chapter 2.2.9). The stream function solution for symmetric waves is assumed to be given by the following dimensional finite Fourier series (Huang and Hudspeth, 1984): X(3)
f(x,z)
T
-UC
N+2
y ^ X(n) sinh n=4
'2n(n-3)(z .
*(3)
+ h)
2n(n — 3)x cos . X(3) . (6.82)
where X(3) = k is the unknown wavelength, and where the dimensions of the dimensional coefficients X(n) in the summation J ] in Eq. (6.82) are [Length]2/[Time]. Equation (6.82) satisfies exactly all of the equations of the dimensional boundary value problem in Eqs. (6.81) except for the dimensional dynamic free surface boundary condition given by Eq. (6.8Id). Evaluating the dimensional stream function ifr(x,z) on the dimensional free surface z — rj(x) yields the following dimensional transcendental equation for the free surface r](x): X(N + 3) [X(3)/T •UC] N+2
r](x) =
+
n?4
'2JT(TI -3)(t](x) X(n) • sinh X(3) [X(3)/T UC 2n{n — 3)x X COS X(3) .
+h)'
, (6.83)
398
Waves and Wave Forces on Coastal and Ocean Structures
where X(N + 3) = ^(x,rj) is the value of the dimensional stream function at the free-surface r](x). Once the TV — 1 stream function coefficients X{n), the wavelength X(3) and the stream function at the free surface X(N + 3) are computed, the free-surface elevation may be computed by solving Eq. (6.83) iteratively. By specifying the dimensional wave height H, dimensional wave period T, and dimensional water depth h, the N + 1 unknown coefficients, X(n),n = 3 , . . . , N + 3, may be computed numerically by the linear-Taylor differential correction technique (Marquardt, 1963) applied iteratively to the following discrete total error: €T(X{n)
J
+ SX(n)) = eQ +
iKJ
7= 1
+ X(2)(H -m+
l)
€r]+€H
7=2,4,6,...
ru),
(6.84a)
where rjj = TJ(XJ) and the average Bernoulli constant Q may be computed by 1
J
Q = 1Y,Qh
(684b)
7=1
where J is an odd number that is equal to the total number of integration points along the discretized free-surface profile between the wave crest and trough in Fig. 6.9; and X{\), X(2) are the two Lagrangian multipliers (Glowinski, 1984 or Hildebrand, 1965). At each ith iteration, a small correction 8X^'\n) to the previous estimate for each nth unknown coefficient X^(n)is computed and added to the previous ith coefficient. These small corrections 8X^'\n) to the previous estimate X^(n) may be computed by solving simultaneously the following system of equations: deT(X(n) + SX(n)) = 0, d[8X(m)}
n,m = 1,2,..., TV + 3,
(6.85a)
Nonlinear Wave Theories
399
that may be expressed in matrix notation as [A]{SX] = {B}.
(6.85b)
The iterations are terminated when the dimensional total error eT(X(n) + 8X(n)) = k G | 1 / 2 + |e„| + \eH\
(6.86a)
is acceptably small (10~6, say). At each z'th iteration until convergence, the new coefficients X^'+l\n) are computed according to X(i+l\n)^X^\n)
+ 8X{i\n),
n = 1,2,..., N + 3.
(6.86b)
For either near-breaking waves or waves in extremely shallow water, the iterative corrections oscillate rapidly and only a small fraction of the differential corrections obtained from solving Eq. (6.85b) should be added to the estimates from the previous iterations by Eq. (6.86b) in order to achieve a stable solution (Marquardt, 1963). The starting estimates for X(3) and X(4) are computed from linear wave theory: h h 27th — = tanh X0 XQ) *(3). 2
HXQ) X(4) = IT
27th cosech X(3) ,
(6.87a,b)
where Xo = co /g is the deep-water wavelength from linear wave theory Eq. (4.24h) in Chapter 4.3. The elements of the matrix [A] in Eq. (6.85b) are lengthy expressions that are given by Huang and Hudspeth (1984). Dean (1974) tabulates the principal wave field variables from 40 wave cases that extend over three decades of dimensionless relative depth from shallow-water wave conditions (h/X < 1/25) to deep-water wave conditions {h/X > 1/2). The dimensionless relative depth (h/Xo) varies between 0.002 < h/Xo < 2.0. For each dimensionless water depth tabulated, four values of dimensionless relative wave steepness (H/Xo) are computed that correspond to the following four ratios of wave height to the breaking wave height (H /Hi,): 0.25,0.5,0.75 and 1.0. Figure 6.10 illustrates these 40 cases on an H/T2 versus h/T2 dissection plane (Dean, 1974).
400
Waves and Wave Forces on Coastal and Ocean Structures
io 2 i
h/
K
101
i
1
io° 1—i
10° 5 2 IO 1 (15
&
2
2. 10 3 10 2 2
H/X.
5 IO-1 2 5 10° 2 2 A/T (ft/sec2)
5 IO1
Fig. 6.10. The 40 cases on dimensionless H/T2 versus h/T2 dissection plane (Dean, 1974).
6.6. Breaking Progressive Waves Both theoretical and empirical estimates for breaking waves are available. Reid and Bretschneider (1953) combine both theoretical and empirical breaking wave estimates to form a breaking index curve over three decades of dimensional water depth from deep- to shallow-water on the dimensional Hb/gT2—ht,/gT2 dissection plane illustrated in Fig. 6.11. Separate theoretical wave breaking estimates are required for deep- and shallow-water as illustrated in Fig. 6.11, where the deep-water estimates are identified as "Michell theory" and are independent of the water depth h, and where the shallow-water estimates are identified as "Solitary wave theory" and vary linearly on the logarithmic axes with the depth of breaking hi,. Empirical estimates connect these two theoretical estimates in deep- and shallow-water over one decade of the dimensional breaking depth hb/T2 between 0.3 < hb/T2 < 3.0.
401
Nonlinear Wave Theories
.01
6
8 .1
i
i i i I
Michell theory —T?
/L
Solitary wave theory —-, /y
Hh 2 2
T
,
\7§& Breaking index curve
2
(ft/sec ) 8
>4a
6
2
'244
.OIL .01
6
" £* -( f•t /•s e>c •2 ) v s -AT h .( f t / s e c 2 ) • 0 1 o
8.1
rrti
Danel (France) D WHOI Estero Bay o B E B Tank SI O • Lake Superior Berkeley Tank (beach slopes 1:30, 1:50)-
Numbers beside plotted points designate number of waves averaged .01 6 8 1 6 8 10 2 T2 ( f t / s e c )
hJT
Fig. 6.11. Wave breaking index dissection plane (Reid and Bretschneider, 1953).
Theoretical Wave Breaking Criteria Two theoretical criterion have been identified to determine wave breaking; viz., kinematic and dynamic. Dean (1974) tabulates the relative error between the Dean Stream Function Theory and linear wave theory for each criterion. Kinematic Wave Breaking Parameter Iljf The kinematic wave breaking parameter is the ratio of the maximum horizontal water particle velocity UQ at the wave crest rjc to the wave celerity C given by IU = ^
= 1
z = tic-
(6.88a)
Dynamic Wave Breaking Parameter Ho The dynamic wave breaking parameter is the ratio of the maximum vertical water particle acceleration Dwc/Dt at the wave crest r\c to the gravitational constant g given by
nD = -
Dwc/Dt Z = rjC.
(6.88b)
402
Waves and Wave Forces on Coastal and Ocean Structures
Theoretical Breaking Progressive Wave Estimates in Deep- and Shallow-Water Miche (1951) applied the kinematic wave breaking criterion Oj^ and Stokes wave kinematics to compute the following maximum wave steepness criterion that is valid in all water depths: Hh
— = 0.14tanh£/j. Xh
(6.89)
Deep-Water h/\Q > 1/2 The Miche formula for wave steepness in Eq. (6.89) in deep-water reduces to — « 0.14, (6.90a) x ob where Xob is the wavelength of a breaking wave in deep-water. Stokes (1847, p. 227, and 1880, p. 320) computes the angle that the free-surface makes on a deep-water breaking wave crest shown in Fig. 6.12 to be 120° (vide., MilneThompson, 1968, Chapter 14.50). Michell (1893) computed the maximum wave steepness in deep-water based on a wave crest angle of 120° from Stokes wave theory to be Hh 1 -*- = 0.142 = - , (6.90b) X0b
7
that is in agreement with the Miche estimate in Eq. (6.90a). In addition, Michell (1893) computes the increase in the nonlinear deep-water breaking wavelength Xoh over the linear wavelength Ao to be ^ = 1.2, (6.90c) Xo so that the nonlinear deep-water breaking wavelength Xob is approximately 20% longer that the linear wavelength AQ = gT2/2n. Substituting Eq. (6.90c)
Fig. 6.12. Crest angle a for maximum wave steepness in deep-water.
Nonlinear Wave Theories
403
and the formula for A.o into Eq. (6.90b) yields the dimensionless deep-water ratio % = 0.027. (6.90d) Multiplying Eq. (6.90d) by the gravitational constant g in English units yields the "Michell theory" breaking index curve for the deep-water breaking waves that is illustrated in Fig. 6.11. Shallow-Water (kh «: 1) Wave Breaking on Planar Beaches with Mild Slopes Wave reflection may be neglected on mildly sloping planar beaches when tan f$ < 0.1 where /? is the angle that a planar beach makes with the horizontal (i.e. tan yS = vertical rise / horizontal run = V: H). Breaking waves on mildly sloping beaches may be classified according to the Iribarren number or Battjes surf similarity parameter $•{, that is defined as tanS
tan/3
Hb =
=
, „ s (6.90e)
(spilling),
(6.90f)
(plunging),
(6.90g)
V#*Ao y/{Hb2n)l(gT2) Galvin (1968) classifies three types of breaking waves on mildly sloping planar beaches by the parameter & according to & < 0.5 0.5 < & < 3.3 %b > 3.3
(collapsing/surging).
(6.90h)
In shallow-water kh <§C 1, the wave breaking parameters Uj in Eqs. (6.88) may be expressed as the ratio of the breaking wave height Hb to the breaking water depth hi, rather than by the wave steepness ratios in Eqs. (6.89 and 6.90). This parametric dependency is given by ^
= Yb,
(6.91a)
hb
where various values of the dimensionless parameter yb may be found and where y^ ^ pg. The Miche formula in Eq. (6.89) for kh «; 1 reduces to ^ hb
= 0.88 = yb-
(6.91b)
404
Waves and Wave Forces on Coastal and Ocean Structures
McCowan (1894) applied the kinematic wave breaking parameter O^ to solitary wave theory (vide., Wehausen and Laitone, 1960, Sec. 31) and derived ^
= 0.78 =
Yh,
(6.91c)
hb
that is illustrated in Fig. 6.11 as "Solitary wave theory" for shallow-water breaking waves. Empirical Progressive Breaking Wave Estimates in Shallow-Water Aft«l Weggel (1972) analyzed laboratory data for progressive breaking waves in shallow-water on mildly sloping planar beaches and determined that breaker heights depend on the tangent of the beach slope tan p\ His empirical formula for estimating progressive breaker heights on mildly sloping beaches where tan ft < 0.1 and where H0/lo < 0.06 is hb
bifi) - a(/3)-^ 2 = YbiP, Hb), gT
(6.92a)
1 + exp(—19.5 tan p) a(fi) =43.8(1 - e x p - ( 1 9 tan £)).
(6.92c)
The Weggel breaker height formula in Eqs. (6.92) may be applied to estimate breaker heights and depths from deep-water waves on mildly sloping beaches with straight and parallel contours (vide., Chapter 4.6). An unrefracted deep-water wave height may be defined by H'0 = H0KR,
(6.93a)
where HQ is the deep-water wave height and KR is the refraction coefficient for straight and parallel contours in Eq. (4.69) in Chapter 4.6. Komar and Gaughan (1972) or Komar (1976) modify Munk's formula (1949) that relates the breaker wave height Hb and the unrefracted deep-water wave height HQ by Hh —
0.563 =
T77 = * V 1/5
(6-93b)
H0' (// 0 'Ao) The breaker height Hb may be estimated from the unrefracted deep-water wave height HQ by Eq. (6.93b) and then the breaker depth hb may be estimated from Eqs. (6.92).
405
Nonlinear Wave Theories
6.7. Second-Order Nonlinear Planar Wavemaker Theory Havelock (1929) applied Fourier integrals to develop the first theory for surface gravity waves forced by both planar and circular wavemakers in water of both infinite- and finite-depth. The Hudspeth and Sulisz (1991) and Sulisz and Hudspeth (1993) reviews of a second-order nonlinear WMBVP for planar wavemakers are summarized here. Even though the linear first-order eigenseries will converge for any geometry of a generic planar wavemaker (vide., Fig. 5.3 in Chapter 5.2), the weakly nonlinear second-order solutions obtained from Stokes perturbation expansions will not converge for all planar wavemaker geometries. The two length scales applied to scale the coordinates {x,z\ by the Lindstedt-Poincare perturbation method (Nayfeh, 1973, Chapter 3, Nayfeh, 1981, Chapter 4.3, or Nayfeh, 1985, Chapter 4.1) for a planar wavemaker in a 2D channel in Chapter 5.2 will now be replaced by the following single length scale for the planar wavemaker geometry illustrated in Fig. 5.3a: _ (x, z, h, d, b, Ab, A) = k(x, z, h, d, b, Ab, A),
_ 2n k = -z-, A
where k is a dimensional wavelength, and where all of the other scale parameters including the dimensionless perturbation parameter e = k A «s kS defined in Chapter 5.2 remain the same. The dimensionless fluid motion may be computed from the negative gradient of a dimensionless scalar, time-dependent velocity potential <£>(x,z,t) according to q(x,z,t)
— u(x,z,t)ex
+ w(x,z,t)ez
— -V4>(x,z,0,
(6.94a)
where the two-dimensional gradient vector operator (vide., Chapter 2.2.7) is V(,) = d(.)/dxex
+
d(.)/dzez.
The total dimensionless pressure field P(x,z,t) may be computed from the unsteady Bernoulli equation (4.3) in Chapter 4.1 according to P(x,z,t)
= p(x, z, t) + ps(z) =
' dt
-T
V
2
+ B(t) J
--, €
(6.94b)
406
Waves and Wave Forces on Coastal and Ocean Structures
where B(t) is the dimensionless Bernoulli constant (cf. Q(t) in Eq. (5.3b) in Chapter 5.2 or Eq. (6.18c) in Sec. 6.3) The dimensionless scalar, time-dependent velocity potential
\z
=0,
x > € I ^ J | ( Z , t), -h
erj(x, t),
x > € I j J H-h,t), z = -h,
92
3* 3z 'e 3 e1 2, ^ ~ o V4>(x,z,0»V |V
/ 5 \ 3g(z,Q \ AI dt x = el-U(z,t),
(6.95b)
9<&(JC,Z,0
2
3<S>(x,z,Q 3x
(6.95a)
e
= 0,
(6.95c)
z = er){x,t),
/ 5 \ 3<E(*,z,Q 3g(z,Q \ AI dz dz
=
'
-h
where S is the dimensional wavemaker stroke; A is the dimensional wave amplitude and where the small parameter e = k A <$; 1. In addition, a radiation condition is required as x ->• +oo in order to insure that propagating waves are only right progressing or that the fluid velocities are bounded (Wehausen andLaitone, 1960). The dimensionless displacement of the generic planar wavemaker §(z,0 shown in Fig. 5.3a in Chapter 5.2 is £(z,0 = xU/h)sin(cot + v),
(6.96a)
407
Nonlinear Wave Theories
where the dimensionless radian wave frequency a> = 2n/T, T = the dimensionless wavemaker forcing period, and the generic shape function /(z/A) that is illustrated in Fig. 5.3a in Chapter 5.2 is given by X(z/h)
= (h/A)[a(z/h)
+ /3][U(z/h-d/h
+ l)-U(z/h
+ b/h)], (6.96b)
where £/(•) is the Heaviside step function (vide., Eq. (2.1) in Chapter 2.2.2) and where for the Type I planar wavemaker in Fig. 5.3a the dimensionless coefficients a and /3 are /
Sb\
R
A
(d
Ab
A
\
(6.96c,d)
where S is the dimensional wavemaker stroke measured at an arbitrary dimensionless elevation z/h = — 1 + d/h + Ab/h + A/h above the wave flume bottom as illustrated in Fig. 5.3a. A planar piston wavemaker is represented by Sb/S — 1, and a planar piston or hinged wavemaker of full-depth draft is represented by b/h = d/h = 0. The dimensionless free-surface profile may be computed from ri(x,t) =
d®(x,z,t) ^ ^ at
1 - -e\V4>(x,z,t)\2 2
9
+ B(t), (6.97)
x > € I -j Jt(7?,0,
z = er](x,t),
and the total dimensionless pressure from p{x,z,t)
=
d
\\ dt
1
-
- -e\V4>(x,z,t)\2 2
9
Z
- - + B(t), e (6.98)
*>
-h
The dimensionless CKDFSBC free-surface boundary condition in Eq. (6.95c), the dimensionless wavemaker kinematic boundary condition Eq. (6.95d) and the dimensionless free-surface elevation t](x,t) computed from Eq. (6.97) may be expanded in a Maclaurin series (vide., Chapter 2.3.3)
408
Waves and Wave Forces on Coastal and Ocean Structures
according to
E (*»?(*,n\ 0 ) "
n=0
x
d2
3" 'dz"
= 0,
(6.99a)
x > 0, z = 0,
£ [6(5/A)g(z,Qf n\
84>(x,z,t) dx
3" dxn
(s\
n=0
S\ \ A )
d$(z,t) dt
d
\AJ
dz
= 0,
3z
(6.99b)
x = 0, -h < z < 0,
3" K*,0-J] (^(x,or «! 3z" rc=0
' at
- - e | V O ( x , z , 0 | 2 + B(t) = 0, 2 (6.99c)
x > 0, z = 0.
In addition, the dimensionless dependent variables may be expanded by the following Lindstedt-Poincare perturbation method (Eqs (6.20) in Sec. 6.3; Nayfeh, 1973, Chapter 3; Nayfeh, 1981, Chapter 4.3 or Nayfeh, 1985, Chapter 4.1): Q{x,z,t) = ]Te m ( m + i)
ri(x,t) = ^
m=0
€m
{m+X)r]{x,t),
m=0
(6.99d,e) p(x,z,t)
= 2%m(m+i)p(x,z,0, m=0
B(t) =
J2€m(m+i)B(t). m=0
(6.99f,g) In contrast to expanding the dimensionless celerity C by Eq. (6.20d) in Sec. 6.3, the dimensionless frequency co = litjT and the dimensionless time
409
Nonlinear Wave Theories
t are now expanded by
cot = I ^ V w „ jt = \n=0
(6.100a)
I
This Lindstedt-Poincare perturbation of the dimensionless frequency to leads to the following change of variables:
(6.100b) \n=0
/
and, correspondingly, a dimensionless free-surface operator Xn (•) defined by
AW =
X>%, \n=0
dr2
+
dz
(•),
(6.100c)
where the value of each dimensionless frequency con is to be determined at each perturbation order en. Substituting the Lindstedt-Poincare perturbation expansions from Eqs. (6.99 and 6.100) into the nonlinear WMBVP in Eqs. (6.95) and following the procedure in Sec. 3 of equating coefficients of equal powers of the perturbation parameter e" yields a system of linear equations in the unknown variables at each perturbation order en. Linear First-Order Solution e° The dimensionless linear WMBVP that results from the Lindstedt-Poincare perturbation method is: V 2 I O ( X , Z , T ) = 0, d I
dz
= 0,
x > 0,
-h
x > 0, z = -h,
(6.101a) (6.101b)
410
Waves and Wave Forces on Coastal and Ocean Structures
^o{l4>(x,z,T)} + ft;o9lf(Z) = 0 ,
x > 0 , z = 0,
(6.101c)
x = 0, -h < z < 0,
(6.101d)
3T
3l$(x,z,r) 3x
s\ ag(z,r) ,
-COQ
A/
3r
where the dimensionless first-order free-surface operator Xo(») is £o(>) = \co^
32
3 + - \
(6.101e)
(.).
A radiation condition is required as x - • +oo for uniqueness that will admit only right-progressing waves or bounded evanescent eigenmodes. The dimensionless first-order free-surface elevation t](x,z) and the dynamic pressure p(x, z, r) may be computed from 9 i<&(*,Z,T) IT](X,T) = co0
ip(x,z,r)
3T 3I<E(X,Z,T)
= COQ
3T
+ IJB(T),
x > 0 , z = 0,
(6.101f)
+ IB(T),
X>0,
(6.101g)
-/J
The dimensionless linear solution e° to the dimensionless WMBVP is (Hudspeth and Sulisz, 1991): \
=
—ai>i(z//*)sin(x — r — v) •cos(r + v)^am(f)m(z/h)exp(-icmx)
\ ,
(6.101h)
m=2
where the orthonormal eigenfunctions 4>m(z/h) are ±
4>m(z/h) =
cos[Kmh(l + z/h)]
,
2
nm =
[2icmh + sm2Kmh] — , m > 1, (6.101ij)
provided that i B{x) is identically equal to zero and that coQh + Kmhta.tiKmh = 0,
m > 1,
K\ = + v —1 = +i.
(6.101k)
The dimensionless coefficients am for the dimensionless generic shape function £(z, r) for a planar wavemaker are ai =
co0h
S\ fh\
n
Dl h
/
d b a fi]
H^ I ) { A ) { ' h ' h ' '
0
'
. m= l
(6.1011)
411
Nonlinear Wave Theories
%=
( S I (i)''(* A rt'"'"' m~x
<6101m)
where 4>\ (0) cosh bu [tanh h — tanh bu] —
Di(h,^,a,p)=ph
+a
01 (0) cosh Z? x [bu (tanh £>„ — tanh h) — 1 + tanh & tanh &M]
W£-.
1+A| 1 - y
)tanh<4 (6.10 In)
(j)m(0)cosKmbu[tan
An \Kmh,-,-,a,fi h h
) = fiicmh
*(H
+
(/>m(0)cos/<:m^
/cm bu (tan Km A - tan /cm bu) — 1 — tan /cm/z tan Kmbu
+a 1 - /cm/j I 1 - - M tan Kmdu m > 2,
(6.101o)
smh/z(l + (.)) >i(») =
sin/cw/z(l + («)) ,
«i
>/«(•) =
> m > 2. «m
(6.101p,q) 6H = bU(b/h); du = dU(d/h) (6.101r,s) where £/(•) is the Heaviside step functions (vide., Chapter 2.2.2) that is required in order to ensure that negative values for the dimension b and d are not applied in the arguments of the transcenlental functions in Eqs. (6.101n,o) by Eqs. (6.101r,s). Some of the inner products that are required for the second-order €l solution may be simplified by expanding the wavemaker shape functions in the following orthonormal eigenfunction expansion:
x
(D = a Ml) W1 (D ~ ^flm,Cm0m © '
(6101t)
412
Waves and Wave Forces on Coastal and Ocean Structures
^f=(i)(l)H''©+£-^©)- <—> For generic planar wavemakers, the evanescent eigenseries in Eq. (6. lOlh) converges at least as fast as am —>• (icmh)~2 ~ [(m — 1)TT]~2 for large m for a piston wavemaker of full-depth draft and for a hinged wavemaker of variable-draft if a - 1 = 1 + A&/A. This difference in the convergence of the first-order eigenseries is important when computing nonlinear solutions for the geometries of planar wavemakers that may be computed from perturbation expansions. Far away from the wavemaker (x/h > 3, say) the dimensionless first-order free-surface profile \T](x, x) is \t](x,x) — cos(x — r — v),
(6.101v)
so that the inverse of the dimensionless wavemaker gain function in Eq. (5.49o) in Chapter 5.2 is
(^M))2
., x > 3«,
0i(O)Di(/i > (&//i),(d//O,a,/5)'
,,
im
,
(6.101w)
that is valid for both hinged (Sb ^ S) and piston (S^ = S) wavemakers of variable-draft. Substituting Eq. (6.101w) into Eqs. (6.1011,m) reduces the coefficients of the linear first-order potential i $ (x, z, x) to the following far field expressions: a\ =
,
m= 1
(6.101x)
a>o
Dm(Kmh,(b/h),(d/h),a,/3)
am = a\ K h m
D\(h,(b/h),(d/h\a,P)
J
m>2.
(6.101y)
Second-Order Solution e 1 The dimensionless WMBVP correct to second-order from the LindstedtPoincare perturbation expansion is V 2 2 * ( J C , Z , T ) = 0,
d2Q(x,z,T)
dz
x>0, = 0,
-h
(6.102a)
x > 0 , z = -h,
(6.102b)
413
Nonlinear Wave Theories
d2B(r) + (i>o dx
Xo{2^(x,z,r)} -2o)Qcoi
92i$(jc,z,r) 9 -. 22 - ^z + « O T - | V i$(x,z,r)| 9r „ dx -ir](x,x)—X0{i^(x,z,x)} az
x > 0,
9 2
(6.102c)
z = 0,
5
9§(z,r) 3 !(*,Z,T)9$(Z,T) -o»l— h dx dz dz 92I
dx2 x = 0,
$(z,x) (6.102d)
- A < z < 0.
The solution to Eqs. (6.102) must also satisfy a radiation condition as x ->• + o o that will admit only right progressing waves or bounded fluid velocities. Substituting Eq. (6.101t) into Eq. (6.96a) and then Eq. (6.96a) into the RHS of Eq. (6.102d) significantly reduces the algebra required to compute the second-order solution. Because Eq. (6.102d) is an inhomogeneous Neumann boundary condition (vide., Table 2.3, Chapter 2.5.1), any constant times x is also a time-independent solution for 2<&(x, z, t ) . The free-surface elevation 2*1 (x,x) and dimensionless dynamic pressure 2P(x, z, x) may be computed from coo
2r](x,x) =
d2<&(x,z,x) dx
+ a>i
9i
9ZI0>(X,Z,T)
+a>o \r){x,x) x > 0,
2p(x,z,x)
1 -.
d2<$>(x,z,x) = COQ dx 1 -
2
- | V I
9z9r
2
+ 2B(x) (6.103a)
z = 0, di$(x,z,r)
Vov
|VIO(X,Z,T)|Z +
dx 2
S(T),
x > 0 , - h < z < 0 ,
(6.103b)
414
Waves and Wave Forces on Coastal and Ocean Structures
where for a zero-mean free-surface elevation,
-L
*{.)dx, 2B(r) = (^-)
,
(6.104)
so that 3 2 B ( T ) / 3 T = 0 in Eq. (6.102c). In addition, the first term in the RHS of Eq. (6.102c) must vanish because 3^$(x,z, T ) / 3 T 2 is a homogeneous solution of the linear operator Xo{»} on the LHS of Eq. (6.102c) and would introduce a secular term of the form T3 I <J>(x, z, T ) / 3 T in the solution for 2^>{x, z, r). Because COQ > 0, it is required that
<wi=0,
(6.105)
that implies that the propagating wave number k is a constant correct to second-order e 1 . Consequently, the linear dispersion equations (6.101k) may be applied to reduce some of the transcendental expressions at second-order e1. However, any expression resulting from substitutions of Eq. (6.101k) would not be valid for higher order approximations e" for n > 1. For well-posed boundary value problems with inhomogeneous boundary conditions on orthogonal boundaries such as those given by Eqs. (6.102c,d), it is advantageous to linearly decompose the solution into complementary homogeneous and inhomogeneous solutions. Accordingly, the solution to the second-order e 1 WMBVP in Eqs. (6.102) may be expressed as a linear combination of four scalar velocity potentials defined by 2$(X,Z,T) = 2®S\x,Z,t)
+ 2<$>e{x,Z,T) + 2®J\x,Z,t)
+ 2*C*,z),
(6.106) where 2<&S(x,z,r) is a second-order Stokes wave potential that is independent of the wavemaker motion, 2$>e(x,z,r) is an evanescent interaction potential, 2<^>^(x,z,r) is a wavemaker-forced potential, and 2^(x,z) is a time-independent potential that is required in order to satisfy Eqs. (6.102c,d). Substituting Eq. (6.106) and Eqs. (6.10 lh-s) into the inhomogeneous boundary
Nonlinear Wave Theories
415
conditions in Eqs. (6.102c,d) yields ^o{2*S(x,^r)+2*«(x,£,r)+2*/(x,^)+2*(JC,i)} a\f\ \4>\ ( - J J s i n 2 ( x
r — v)
-a\ sin(x - 2r - 2 v ) ^ ] a m exp(-Ar m x)/ 2 (0i ( T ) , 0 m ( r ) ) m=2
- a i cos(x - 2r - 2 v ) ^ a m exp(-/c w x)/3 (0i ( r ) , 0 m ( r j ) m=2
- s i n 2 ( r + v ) ^ ^ a m a „ exp[-(/c m + Kn)x]f4 [4>m ( T ) ,0« ( r ) ) m=2n=2
- a i c o s x ^ a m e x p ( - K m x ) / 5 (0i (r)>0m ( T ) ) m=2 x > 0,
z
(6.107a)
= 0,
and
lA2*s(x>l>r)+2*e(x>hr)+2*f(x'Zh>r)+2*(x4)}
|^(^©^(D) [ 1 - c o s 2 ( T + v ) ] .^a m/ c m iy 2 (0m (£) , x (£)) m=2
x = 0,
-/z < z < 0,
(6.107b)
where the dimensionless nonlinear, free-surface interaction terms /i(»), hi; •), hi; •), f4i; •), and / 5 ( . , . ) are defined by
h (01 ( I ) ) = y[20?(-l) + 0?(O)(l -fflgtanhA)] (6.107c)
/2 (01 ( ^ ) ,0m ( ^ ) ) = 2Wo^m01 (O)0m(O),
(6.107d)
416
Waves and Wave Forces on Coastal and Ocean
0)
MMD^G))=(TW **
(0)
Structures
(DgCtanh h — Km tan Kmh) —4 tanh h.Km tan Kmh + K}„ — 1
= (y)>l(O)0 m (O)[6^
+ /c2
_
1L
\AT?I ~i Kn)
(6 107e)
+
LKmKn
ft (
/ 5 (01 (^),d>m ( I ) ) = (y)4>i(0)<M0) —&>g(tanh hm + K tan /c /i) m m = ( y ) ^ l ( 0 ) ^ ™ ( 0 ) [ 4 + l],
(6.107g)
correct to second-order el. The dimensionless nonlinear wavemaker interaction terms W\ (•, •) and Wi{; •) are defined by
1
(6.107h) m=2
*Mf)-*(s)) =3(z/A) 1
••(*G) 7 (i)y,
(6.1071)
COQ
n=2
where (..)' = d/d(z/h)(..). The dimensionless inner product terms that are required to compute the dimensionless coefficients of the eigenfunctions in the second-order el
417
Nonlinear Wave Theories
potentials 2 < ^(*, z, T) and 2^^{x, z) are defined by the integral (;»)z/h = /
(;»)d(z/h)
that may be integrated to obtain
_
|
(6.107J) where H> Qj = —kj tmkjh
^
A,- (\)
= Qj ( £ ) for kj = fij
—
0,
for A; ( | ) = ^ ( ^ ) for A.; = /i, J (6.107k)
and D(Km,Kn,kj)
= [ 0 4 + KI - k2j) + 2KmKn][(K% +Kll-
k2j) - 2KmKn]. (6.1071)
For n = 1, K\ = i = V—T and
*i =
sin[j'/i(l + z/h)]
i sinh[/i(l + z//i)]
ni
"l
(6.107m)
so that
- ((*• (f) « (!))' • 4,,=-((*• (!) « (s))' • 4 , . • - >• (6.107n) Similarly, the inner product between 4>[(z/h) and
<-K) *(!))„.=-((*© *(*))'•<) = -^1(0)^,(0).
z/h
(6.107o)
418
Waves and Wave Forces on Coastal and Ocean Structures
The dimensionless second-order Stokes wave potential 2®S(x,z, r) satisfies Eqs. (6.102a,b), a radiation condition, and the following Stokes wave component of the dimensionless inhomogeneous free-surface boundary condition in Eq. (6.107a): £o{2®S(x, z/h, T)} - a\fi (Vi ( ^ ) ) sin 2(JC - r - v) = 0,
x > 0, z = 0. (6.108a)
The dimensionless solution is (Stokes, 1847) 2$>S(x, z/h, r) = - I — - I cosech4/z cosh 2h(l + z/h) sin 2(x - z - v) (6.108b) l
correct to second-order e . The dimensionless evanescent interaction potential 2$e(x,z, r) must satisfy Eqs. (6.102a,b), a radiation condition, and the following evanescent wave component of the dimensionless inhomogeneous free-surface boundary condition in Eq. (6.107a): £o{2&(.x,z/h,T)} + a\ sin(x - 2 T - 2 v ) ^ a m e x p ( - / c m x ) / 2 (fa (-) ,4>m (fa ( r ) ) ) m=2 +
x>0,
z = 0.
(6.108c)
The dimensionless solution is assumed to be given by 2<S>e(x,r, r) oicos(x - 2 r -2v)^a m exp(-«: m x) [Am0i (^)
- a i s i n ( x - 2 r - 2v)^am
exp(-Kmx) \Am
\h)^m
\l)\
m=2
sin2(r+ v ) ^ ^ a m a „ exp[-(Km+K„)x]Cmn \
(r)*« (r)J
m=2 n=2
(6.108d)
419
Nonlinear Wave Theories
where the coefficients of the orthonormal eigenfunctions are 4[3(4^ + 4 - l ) + 2 ^ ] Am
CO0[(4CO40KI
- l ) 2 + (2 K m ) 2 ]'
(6.108e)
* m [(4a,g + KJ - l)(6o)g + 4 ~ 1) ~ 8 4 ] Bm = — 2«o [( 4w 4 + 4-i) 2 + (2 Km )2] (6.108f) [(Km + Kn)
+ 2KmKn + 6(D%\
l^mn —
^0
[(icolf
+
(6.108g)
iKn-Kn)*]
DET[Am, Bm] = ( -5- ) [ ( 4 ^ + 4 - l ) 2 + (2/cm)2],
(6.108h)
correct to second-order e and where DET [•, •] = a determinant. For generic planar wavemakers where the first-order evanescent eigenseries converges only as fast as am ->• (icmh)~2 ~ [(m — l ) ^ ] - 2 for large m, the two single-summation evanescent eigenseries in Eq. (6.108d) will converge at least as fast as amBm ex. (icmh)~l, and the double-summation evanescent eigenseries in Eq. (6.108d) will converge at least as fast as anCmn ex (Knh)~l. This restriction on the rate of convergence of the evanescent eigenseries in Eq. (6.108d) restricts the practical application of the evanescent interaction potential 2<&e(x,z, r) to those planar wavemaker geometries where the convergence of the first-order evanescent eigenseries improves to am -»• (Kmh)~3 ~ [(m — l)jr]~ 3 for large m. The dimensionless wavemaker-forced potential 2(&f(x,z, r) must satisfy Eqs. (6.102a,b), a dimensionless homogeneous Eq. (6.101e), a radiation condition, and the dimensionless KWMBC in Eq. (6.102d) that is now given by — {l<S>f(x,Z,T)} + — {2
/
aiWi [4>\
/Z\
\ {-),%(Z,T))
2h sin2(r + v) ^2amKmW2 [4>m \-\ 2h 0,
-h < z < 0.
= 0,
,£(z,r)j (6.108i)
Waves and Wave Forces on Coastal and Ocean Structures
420
The dimensionless solution is assumed to be given by f
2$>
(x,z/h,r)
= [Ei cosGSix - 2(T + v)) + Fi sinGSix - 2(T + v))]0l ( ^ ) - ^2 exp(-Pjx)[Ej
sin2(r + v) + Fj COS2(T + v)]Qj ( | ) , (6.108J)
7=2
where the orthonormal eigenfunctions are
Qj
/z\
U)
=
cos Pjh(l + z/h)
A^
N] = / =
(6.108k)
j>h
'
cos2[Pjh(l + z/h)]d (z/h)
(2P;h+ sm2fi jh) -
-—
Afijh
'
;
(6.1081)
i > 1
-
'
provided that 4a>lh + ftjhtanfijh = 0,
j > 1, fi\ — ifi\.
(6.108m)
The coefficients Ej and Fj in Eq. (6.108J) may be computed from
«1 \ ^ n=2
\(tj\Ej = -
+
(£-£)(-«)* ©)'•«/©), z/h
(om'Mi)
«=2«=2
z/h- .
(6.108n)
S) 2 (^)(KfMf))^(f) \Pj\Fj
= m=2
B
-s)((*i©*"©)'-^© (6.108O)
421
Nonlinear Wave Theories
where (••)' = 3/3(z/A)(»») and where the identity (6.108p) m=2 n=2
m=2 n=2
has been substituted into Eq. (6.107i) in order to make the double summation term derived from Wii; •) symmetric. Substituting Eqs. (6.108e-g) into Eqs. (6.108n,o) and persevering through some very tedious algebra gives, eventually
[ ^ p ] 0i (0) 2 j (0) £> m <M0) m=2
WEJ = xr
2/(0) >
[ 2 ^ 4 + ( ^ + 4 + 1) ( 5 ^ - 1 ) ] [(4 - 1 - fi)2 + VKm)2]
> a m a„(a m +a„)0 m (O)0„(O)
m=2n=2
fll«U0
l^-IF,-:
0?(O)fiy-(O)
5-
[(am + a n ) 2 - fij] (6.108q)
5^-3
2 ^ + ( 4 + l-^?)(5a,4+Jf2)
+ [ ^ ] 01 (0)2/(0) J2 "mKmCpmiO) m=2
( 4 - 1 - pzy + (itcmy (6.108r)
where fi\ =ifi\. For generic planar wavemakers where the first-order evanescent eigenseries in Eq. (6.101h) converge only as fast as am —>• (icmh)~2 ~ [(m — l)7r] - 2 for large m, the single summation series in (6.108q) will converge at least as fast as am/c~2 oc (Kmh)~4 ~ [(m — 1)]~ 4 for large m. The double summation series in Eq. (6.108q) and the single summation series in Eq. (6.108r) will converge at least as fast as amKm <x (Kmh)~l ~ [{m — 1)TT] _1 for large m. The orthonormal eigenfunctions for Qj (z/h) in Eq. (6.108k) will converge at least as fast as (j8/ft)-3 ~ [(j — 1)TT]~3 for large j . The coefficients £j and Fj in Eqs. (6.108q,r) depend on the first-order evanescent eigenseries coefficients am even for the propagating free-wave potential j = 1. This implies that the amplitude of the second-order freewave a2 computed from 2<$>f(x,z, x) depends on the first-order evanescent eigenseries and it may not be neglected at second-order. The dimensionless
422
Waves and Wave Forces on Coastal and Ocean Structures
amplitude of the second-order free-wave a2 may be computed from a[ = 2cooQi(0)JEi + Flz,
(6.108s)
and the dimensionless amplitude of the second-order Stokes wave a | may be computed from cosh h (cosh 2/2 + 2) s (6.108t) 4 sinh h The dimensionless ratio of aJ2 /ai„s 2 for a full-draft piston wavemaker is shown in Fig. 6.13 and for a full-draft hinged wavemaker in Fig. 6.14. This
0.4
0.5
hlX Fig. 6.13. Dimensionless amplitude ratio a£ /af f° r a full-draft piston wavemaker : — = with evanescent interaction potential; and = without evanescent interaction potential. (Sulisz and Hudspeth, 1993)
Fig. 6.14. Dimensionless amplitude ratio a^/a^ with evanescent interaction potential; and (Sulisz and Hudspeth, 1993)
for a full-draft hinged wavemaker : — = = without evanescent interaction potential.
Nonlinear Wave Theories
423
ratio is also compared with the same dimensionless amplitude ratio when the evanescent interaction potential i&e(x,z, r) in Eq. (6.108i) is neglected. The time independent solutions for *(x,z) may be applied to compute Stokes drift in wave channels and are reviewed in Sec. 3.1.
6.8. Chaotic Cross Waves: Generalized Melnikov Method (GMM) and Liapunov Exponents The cross wave instability shown in Fig. 6.15 is parametrically excited by the progressive waves generated by a planar wavemaker at a subharmonic frequency of the wavemaker frequency (Bowline et ai, 1999 and Hudspeth et al, 2005). Parametrically excited standing cross waves that oscillate in a direction transverse to the wavemaker forcing with crests perpendicular to the wavemaker may be analyzed by the generalized Melnikov method (GMM) and by the Liapunov characteristic exponents. The GMM is a global perturbation analysis about a separatrix and about a manifold of fixed points that are connected by separatrices for higher dimensional nonlinear dynamical systems (Wiggins, 1988, Sec. 4). The Wiggins-Holmes (1987) generalization of the Melnikov method (Melnikov, 1963 and Arnold, 1978) to higher
Fig. 6.15. Mode 2 cross wave in the 2D wave channel at the O. H. Hinsdale-Wave Research Laboratory (OHH-WRL), Oregon State University (vide., Fig. 5.1 in Chapter 5.1)
424
Waves and Wave Forces on Coastal and Ocean Structures
0.25-j 7 A o.20-
015
"
kS
8 ; * • o 10
13
1
* »
9
4
15
\ 4\
16
»i/
0.10-
•°0<>i»
0.05 -
*
U.UU-)
0.70
18
,
1
1
1
0.85
1.00
1.15
1.30
(2G) 2 /G) W O T ) Fig. 6.16. Neutral Floquet stability diagram for mode 2 cross wave at the O. H. Hinsdale-Wave Research Laboratory (o mode 2 cross waves and A no cross waves) Bowline et al., (1999).
dimensions may be applied to parametrically excited cross waves with surface tension in a long rectangular wave channel in order to demonstrate that cross waves are chaotic. The Hamiltonian for these cross waves in an inviscid fluid is homomorphic to the Hamiltonian for a parametrically excited pendulum that is an example of a Floquet oscillator that may be approximated by the Mathieu equation (Berge et al, 1984). The Luke Lagrangian density function (Luke, 1967) for surface gravity waves with surface tension contains three generalized coordinates (or, equivalently three-degrees-offreedom) that are the time-dependent components of three velocity potentials that represent three standing waves. The neutral Floquet stability diagram (Jordan and Smith, 1987, Fig. 9.2, p. 255) shown in Fig. 6.16 that was measured by Bowline et al, (1999) motivated the inclusion of dissipation in the Floquet Hamiltonian model for cross waves that is reviewed below. A generalized dissipation function that is proportional to the Stokes material derivative of the free surface is added to the action integral of the Luke Lagrangian density (Luke, 1967) so that damping appears correctly in the WMBVP. The generalized momenta and the Hamiltonian are computed from a Legendre transform of the Lagrangian (Lichtenberg and Lieberman, 1992). This Hamiltonian contains both autonomous and nonautonomous components and is transformed by three canonical transformations in order to
Nonlinear Wave Theories
425
obtain a suspended system that will survive the KAM averaging theorem (Arnold, 1978, p. 179) and the subsequent application of the GMM (Wiggins, 1988, p. 387). The Herglotz algorithm (Guenther and Schwerdtfeger, 1985) for computing canonical transformations for autonomous dynamical systems is extended to include nonautonomous systems (Appendix A; Fadel, 1998 or Orum et al., 2000). This extension includes two types of the generalized Herglotz algorithms (GHA). The system of nonlinear nonautonomous evolution equations derived from Hamilton's equations of motion must be averaged in order to obtain an autonomous system that may be analyzed by the GMM. The unperturbed autonomous system possesses hyperbolic saddle points that are connected by heteroclinic orbits. The perturbed Hamiltonian system that includes surface tension satisfies the KAM nondegeneracy requirements (Wiggins, 1988 and 1990); and the Melnikov integral is calculated to demonstrate that the motion is chaotic. For the perturbed dissipative system with surface tension, the only hyperbolic fixed point that survives the averaged equations is a fixed point of weak chaos that is not connected by a homoclinic orbit; and consequently, the Melnikov integral is identically zero. The chaotic motion of the perturbed dissipative system with surface tension is demonstrated by numerical computation of positive Liapunov characteristic exponents (Benettin et al., 1976 and Rasband, 1990). A chaos diagram of the Liapunov characteristic exponent is computed in order to search for regions of the damping parameter and for the range of values of the Floquet parametric forcing parameter where chaotic motions in the dissipative system may exist. In general, cross waves may oscillate at a frequency a>c that is half the frequency of the wavemaker forcing (or, equivalently, the progressive wave u>p) a>c = cbp/2 although other instability frequencies may occur. Cross wave wavelength eigenvalues and the cross wave modes that may be possible are determined by the channel width Lc = 2 (channel width)/« where n is the cross wave mode number and is equal to the number of half-wavelengths across the wave channel. Floquet instabilities occur when the cross wave wavelength eigenvalues also satisfy the frequency dispersion equation (4.15) in Chapter 4.3 for surface gravity waves. Energy is transferred from the progressive waves that are directly forced by a planar wavemaker to the parametrically forced cross waves through the spatial mean motion of the free surface (or, equivalently, the radiation stress); and the growth of the cross wave amplitude is
426
Waves and Wave Forces on Coastal and Ocean Structures
due to the rate of working of the transverse stresses of the wavemaker forced progressive waves. The Hamiltonian for a parametrically excited pendulum (Berge et al., 1984) is homomorphic to the Hamiltonian for subharmonic resonant cross waves; but the nonautonomous components represent significantly different physical processes. The motion of a parametrically excited pendulum is an example of a Floquet oscillator that may be modeled by the linear Mathieu equation (Miles and Henderson, 1990 andNorris, 1994) according to —-j +COQ(\ + rcos2cot)0 = 0,
0 < h «; 1,
where T is a small dimensionless parametric excitation parameter, 2co is the parametric excitation frequency, and coo is the natural frequency of the pendulum in the absence of the parametric excitation T = 0. Hamilton's equations of motion for the parametrically excited pendulum are given by — = P, at
-?- = -co20(l + P(t))8, at
P(t) = Tcos2cot,
that may be obtained from the nonautonomous Hamiltonian
P2 + ole2 H(p,0,t)
= Ho(p,6) + Hr(P(t),e,t)
=
2
°
co2 +
,
-f[P(t)e2],
where Ho(p,6) is the autonomous free oscillations component and Hr(P(t),0, t) is the 100% nonautonomous Floquet parametric forcing component of the Hamiltonian H{p,6,t). In contrast, the corresponding Floquet parametric forcing component in the cross wave Hamiltonian is 100% autonomous; and the 100% nonautonomous component represents boundary forcing by the wavemaker that is not parametric forcing. Of the various theoretical criteria available for testing for the chaotic behavior of a dynamical system, the criterion for horseshoe maps and homoclinic/heteroclinic orbits is employed. Specifically, the Melnikov method gives local criterion for the transverse intersection of stable and unstable manifolds of the perturbed system and for the resulting chaotic motion near the unperturbed (undamped and unforced) homoclinic/heteroclinic orbits. However, the Melnikov method does not signal the appearance of strange attractors (that represents persistent chaotic behavior over a global domain of phase space). The Melnikov method requires that the dynamical system be described in terms of a set of first order ordinary differential equations of the phase space
427
Nonlinear Wave Theories
variables and that the unperturbed (undamped and unforced) system possess homoclinic/heteroclinic separatrices connecting hyperbolic saddle points. The Melnikov method measures the distance between the stable and unstable manifolds of the perturbed system from the unperturbed homoclinic/ heteroclinic separatrix. When there are multiple transverse intersections of the manifolds, the Melnikov integral yields simple zeros that indicate the presence of Smale horseshoes and chaotic motion. Hamilton's Principle for Cross Waves The fluid is assumed to be incompressible and inviscid and the flow to be irrotational. Dimensional variables are denoted by a tilde (S). The dimensional fluid particle velocities q(x, y, z, t) and the dimensional total pressure in the fluid P(x, y, z, t) are computed from q(x,y,z,t)
= -V4>(x,y,zJ),
(6.109a)
P(x,y,z,i) d(x,y,z,t) 1 * ~2 ~~ ((- i n Q M _ _ -\V(/)(x,y,z,t)\ - gz, (6.109b) p dt 2 where <j>{x,y,z,i) is a dimensional velocity potential, p is the dimensional fluid mass density, g is the dimensional gravitational acceleration, and V (•) is the three-dimensional gradient vector operator (Chapter 2.2.7). The Bernoulli constant is assumed to be incorporated into the velocity potential (Eq. (6.15), Sec. 6.3 and Stokes, 1847). The fluid domain is the long wave channel shown in Fig. 6.17 that extends from the wavemaker boundary S^ at x = x(z,f)
-z
=0
z = —h
Fig. 6.17. Definition sketch for a planar wavemaker and long wave channel.
428
Waves and Wave Forces on Coastal and Ocean Structures
to an arbitrary vertical cross-section at a distance x = I from the planar wavemaker. Variational Principle The dimensional Lagrangian X for a free surface wave with surface tension is (Becker and Miles, 1992):
£
=
1 s „ , -\V(p(x,y,z,t)\z al/ ^
r /
_ _ r t(x v t - I) dV + T ^V^ -dS g (6.110a)
dt
where fj(x, y, t) is the dimensional free-surface displacement, and where
dS7l = l{x,y,t)dxdy.
l2{x,y,i)
= \ + \Vrl(x,y,t)\2.
(6.110b,c)
The positive definite dimensional kinematic surface tension f is the ratio of the conventional surface tension to the fluid mass density (vide., Chapter 3.3). The fluid domain V (t) is bounded by the following six surface boundaries (vide., Fig. 6.17): (1) (2) (3) (4 and 5)
the wavemaker boundary S% atx = x(z,i), the free surface boundary Sfj at z = fj(x,y,t), the horizontal bottom boundary S^ at z = — h, the two vertical wave channel side boundaries S±^ at y = ±b, and (6) a far field (i.e. x > 3h, vide., Chapter 5.2) vertical boundary surface S^atx = I.
The Luke Lagrangian density (Luke, 1967) is not equal to the kinetic minus the potential energy densities X j^ T — V (vide., Chapter 4.5 and Scheck, 1990, p. 92).
429
Nonlinear Wave Theories
Generalized Hamilton's Principle with dissipation (Guenther and Schwerdtfeger, 1985) The first variation of Eq. (6.110a) with an additional dissipation integral is 8 jp2
F (d^J))
i, dh + ^
dh = 0,
(6.111a)
where the first integral in Eq. (6.111 a) is the action integral of the Lagrangian density with surface tension (Luke, 1967), and the generalized dissipation function per unit mass density F(Dfj(x, y, t)/Dt) is given by
F (W*'?'*) = -aM f \
Dt
)
D^,yJ)^,y,t)
y~KJs„
Dt
s(x,y,i)
"
where a is a positive definite dimensionless damping parameter, ic is the cross wave wave number, and Dfj(x, y, t)/Dt is the Stokes material derivative (Chapter 2.2.10) of the free surface fj(x, y, t) given by Dfj(x,y,t) —=
=
dij(x,y,t) r=
Dt
£~ s _ V(p(x,y,z,t)»Vv(x,y,t),
, , n n (6.111c)
at
The dimensional WMBVP may be obtained by requiring that the independent variation of 4>(x, y, z, t) and fj(x, y, t) vanish at the arbitrary temporal values t\ and t2 in Eq. (6.111a) (Luke, 1967; Whitham, 1974; or Becker and Miles, 1992). Thefirstvariation of the Lagrangian X in Eq. (6.110a) is given by
* r f~b / " V 1 ! ^ - - - ?M2 w*>y>z>t) , ~~\,-,-,-
x?
8X = 8
I
+ fS f
I -|V0(x,y,z,OI
f
w=
hgzjdxdydz
iX{x,y,t)-\)dxdy
J-bJx
=
rn rb r£ i J-hJ-bJx \
V
Waves and Wave Forces on Coastal and Ocean Structures
430
1 3 -
+
ft
+T
...... -
-\V
dy
d
f-f
l(x,y,t){-l)(Vij(x,y,t).VSfj(x,y,t))dxdy.
(6.112)
where Eq. (6.110b) has been substituted for the differential surface dSfj in Eq. (6.111b). The following identities are applied repeatedly (cf. Hildebrand, 1976, Chapter 6.9): (1) Green's first identity (vide., Eq. (2.123a) Chapter 2.6), f(x,y,z,i)V2(j)(x,y,z,i) dV /,
+ V 0 ( i , y, z, i) • Vf{x, y, z, i) =
l$(x,y,zJ)[V
(6.113a)
(2) The Reynolds transport theorem (vide., Eq. (3.11), Chapter 3.2)
df(x,y,z,i) dV / ,V(i)
dt
d
dt
L ir(x,y,z,i)dV
JV(t)
-/,
\J/{x,y,z,t){U{x,y,z,t)*n)dS, (6.113b)
(3) Vector differentiation, F . V G = V.(GF) - (V ,F)G,
(6.113c)
Nonlinear Wave Theories
431
(4) Gauss divergence theorem (vide., Eq. (2.122a), Chapter 2.6) I V.(GF)dV Jv
= f(F.n)GdS, Js
(6.113d)
where n is the outward unit normal of the fluid boundary. The volume integral in Eq. (6.112)may be transformed by Eqs. (6.113a,b)by substituting^ = 8(f). The dynamic free surface boundary condition (in Eq. (4.5h) in Chapter 4.2) may be substituted into the first surface integral in Eq. (6.112); and the second surface integral in Eq. (6.112) may be transformed by Eqs. (6.113c,d) by substituting F=
, ' , S(x,y,t)
and
G=
8fj(x,y,t).
These transformations reduce the first variation of the Lagrangian Eq. (6.112) to d f . 8X + -z 8
f-f
,2 _ dcj>(x,y,z,t) ~ di
1 i
2\V
8fj{x,y,t)dxdy
J-b J if,
(V24>(x,y,z,t))84>(x,y,~zJ)dV
- [
+
f ~ / (n{x,y,z,i) + U~s(x,y,zJ) •ns)8<j){x,y,z,i)dS~s f f fjn(x,y,z,t)8rj(x,y,zj) -
Js*
l(x,y,t)
n
(6.114a) where the subscript n denotes a normal derivative; i.e., r)„(x,y,t) = Vfj(x,y,z,t)*n 4>„(x,y,z,t) = V4>(x,y,z,t)*h
dfj(x,y,z,i) dn d(j)(x,y,Z,t) = dn
=
(6.114b) (6.114c)
432
Waves and Wave Forces on Coastal and Ocean Structures
The differential surfaces dS$; the outward unit normal«;; and the motion of the wave maker boundary Ug(x, y, z, t) • n^ are given by the following: Sf, : dSf, = t;dxdy;
^ 1 n^ = -(-fjx,-Vy,
* fj^ • « ^ = ^L;
1);
(6.115a,b,c) Sjj : dSz = yjl + x?dydz;nx
= <-1.0,Xi).
uf, n
x' x
-. _
~
-XI (6.115d,e,f)
where the vector notation (•, •, •) in Eqs. (6.115b,e) means (•, •, •) = (•)?* + (»)ey + (»)ez. The geometric quantity V2rj/t; in the last term in the square brackets [•] in the first integral on the RHS of Eq. (6.114a) represents the total curvature of the free surface (Wehausen and Laitone, 1960, Eq. (6.18)) that may be linearized by the binomial expansion (vide., Chapter 2.3.4) y2fj + 0(\Wfj\2).
(6.116)
The last integral on the RHS of Eq. (6.114a) vanishes when either the natural contact line boundary condition fjn = 0 or the edge constraint boundary condition rj = 0 is applied (Benjamin and Scott, 1979). Decomposition of the Lagrangian Integrals The first and the second terms in the volume integrand in Eq. (6.110a) may be transformed by substituting ijr = <> / into the identities in Eqs. (6.113a,b); and the third term in the volume integrand in Eq. (6.110a) may be transformed by substituting gz = V»(FG), F = z2ez and G = g/2 into the Gauss divergence theorem Eq. (6.113d). The Lagrangian Eq. (6.110a) then reduces to
L =
d r ~ . £ + dt —r Jy jJ){x,y,z,t)dV
+ gh2b[i-x(-h,t)]-l-J
_
2 J-hJ-h
"°~
jn\gildl^-)h=xdidy 3z
433
Nonlinear Wave Theories
2Jv
4>(x,y,z,t)V
4>(x,y,z,t)dV
+
I- ( 2^"(*' *'5' ^ + U~s^' *' *' ^ *"* ' ^*' ^' *' ^ ^
i
pb
+~ /
pi
rb
/ 8V2(x,yJ)dxdy
pt
/ tf(x,y,t)
+f /
- 1] dxdy (6.117)
where /^ = / ( z = 77, F), ^ = fj(x = x,y,t), and ij| = r}(x = l,y,t). The difference [L — <*£] in Eq. (6.117) makes a null contribution to the variation of the action integral of the Lagrangian so that the Lagrangian L satisfies Hamilton's principle in Eq. (6.111a). Grouping terms according to the integrals of the surface boundaries shown in Fig. 6.17 decomposes the Lagrangian Eq. (6.117) into the following integrals: i-'V "I
Lp
(6.118a)
~\~ L^fj ~t~ -/-"T^j
where
i r
put r„
1 p
pi
Lfj = - / . /
didy,
(6.118b)
dldy,
dfj(x,y,t) : __2 4>(x, y, z, t)' ' v " ^ ' ' ' + gfjz(x, y, i)
(6.118c)
dxdy, Z—r\
(6.118d) 2
Lf = i [ I \l{x,y,T)-\\dxdy^^J J-b Jy,1
2
[ f\vij(x,y,t) dxdy, J-B J is ' (6.118e)
where Eq. (6.118e) follows from the substitution of the binomial expansion of Eq. (6.110c) by Eq. (6.116) for the total curvature of the free surface. The volume integral in Eq. (6.117) is identically equal to zero by Laplace's equation
434
Waves and Wave Forces on Coastal and Ocean Structures
(Chapter 2.2.9) so that only the following surface integrals in Eqs. (6.118) survive: (1) The free surface fj(x, y, t), (2) The wavemaker surface x(z,t), (3) The far field vertical boundary surface at x = I. The velocity potential <j>(x,y,z,t) is selected to be the field variable so that the free surface fj(x, y, t) may be expressed in terms of 4>(x, y, z, t) by the free surface boundary conditions and the contact line condition at the side walls. Scaling and Nondimensional Parameters The velocity potential >(x, y, z, t) and the free surface displacement rj(x, y, i) are assumed to be a linear combination of a progressive wave component (subscript p) that is independent of y and a cross wave component (subscript c) that is independent of x as follows: 4>(x,y,z,t) = 4>p(x,z,t) + 4>c(y,zJ), rj(x,y,t) = fjp(x,t) + fjc(y,t).
(6.119a) (6.119b)
Dimensional variables (with tilde (•)) are related to nondimensional variables (without tildes (•)) by the following scales: x = l,
y = l , ~z = l ,
k
K h = Kh,
« %=ki,
t = -^=, b = Kb,
8
cpp = (ppap
fjp = apVp, s - icac,
X
=
(6.120a-d)
y/gic (6.120e-g) 8
(pc =
fk2
-, yk
n = Vl r. + r,
I* Xr = J-,
(6.120h,i) (6.120j-m) (6.120n-q) (6.120r-t)
where y ^ pg, aw is the dimensional amplitude of the wavemaker displacement, k and ap are the dimensional wave number and amplitude of the
435
Nonlinear Wave Theories
progressive wave, respectively; and ic and ac are the dimensional wave number and amplitude of the cross wave, respectively. The dimensional progressive radian wave frequency S)p is related to the dimensional progressive wave number k by the dimensional deep-water dispersion equation with dimensional surface tension according to S?p = ~g~kxx, rx = \ + {x/XAr).
(6.121a,b)
The dimensional cross wave wave number ic must also satisfy the dimensional deep-water dispersion equation with dimensional surface tension according to S?c=gk{\ + T),
(6.121c)
where the eigenvalue solution for icT from Eq. (6.121c) must also be equal to the eigenvalue computed from the homogeneous dimensionless Neumann boundary condition Eq. (6.123d) below. The dimensional progressive wave period and wavelength are given by Tp = 2n/a>p and Xp = 2n/k, respectively; and the dimensional cross wave period and wavelength are given by Tc = 2n/u)c and Xc = 2n/k, respectively. The dimensionless velocity potential (f>(x,y,z,t) and the dimensionless free surface displacement r](x, y, t) in Eqs. (6.119) are now given by
(6.122a)
r](x,y,t) = rjc(y,t) + Tt]p(x,t).
(6.122b)
Requiring that the independent variation of the dimensional 4> and r\ vanish at the arbitrary dimensional temporal values t\ and ti in Hamilton's principle from Eq. (6.111a) yields the following scaled dimensionless WMBVP for the dimensionless velocity potential
1 d2
YX(z,t)<x<$ y<\b\ -h
<eri(x,y,t),
(6.123a)
436
Waves and Wave Forces on Coastal and Ocean Structures
d(p(x,y,z,t) dz
+s
dt](x,y,t) -x\< 1 d(p(x,y,z,t) drj(x,y,t)\ Xj d* dx d(p(x,y,z,t)3t](x,y,t)
+
By
dy
Z=
er){x,y,t),
/ J
_1_ /dcp(x,y,z,t) 21 X4r \ dx d(p(x,y,z,t) , d(p(x,y,z,t) - TI — dy d
+
YX(z,t)<x<£ y<W
(6.123b)
h r](x, y, t)
dz d2r](x,y,t)\ dx2
( 1 xj
d2r)(x,y,t)
V+ dy2 J d)(x,y,z,t) —a dz
V^>(x,y,z,t)»n
= 0,
, YX(z,t) <x < £ , y <\b\, z = sr](x,y,t),
(6.123c)
YX(z,t) <x <%, y = \b\, -h YX(z,t) <x < f ,
y < \b\, z = -h, (6.123d,e)
d(j){x,y,z,t)
dx x = YX(z,t),
Vrj(x,y,t)
• « = 0,
y .4
= —Kx\
dx(z,t)
s r dt y<\b\,
,
4d(j)(x,y,z,t)
dx(z,t)
\- YK r
r
—,
dz -h
\x = YX(z,t), y <\b\, [ y / ( z , 0 <x < £ , y = \b\,
dz
(6.123f)
-h
plus an appropriate radiation condition at x = £. The free surface curvature in the first term on the RHS in Eq. (6.123c) requires a dynamical constraint
437
Nonlinear Wave Theories
that is given by the contact line conditions in Eqs. (6.123g,h) or, alternatively, an edge constraint (Benjamin and Scott, 1979). The dimensional Lagrangian L in Eqs. (6.118) may now be decomposed into the following dimensionless integrals: (6.124a) where dx(z,t) Lx = - f / rX(n
fb rer\i=
d(f>(x,y,z,t) dx
2kf i rb ^
= - \ \ \ ^ J-b Jyxn L YXr, b 1
ri
dt](x,y,t) dt
, +
(6.124b)
dzdy,
2
t]L{x,y,t)
(6.124c)
dxdy,
-lz=er]
(6.124d) 2
rl
dzdy, /x=yx
J_ /dr,(x,y,t)\
/3r?(x,y,Q
x 2
"
dxdy,
(6.124e)
J-bJyXr,
where r\x = t](x = yx,yJ),m ~ ^x = H,y,t), a n d Xn = x(z = er),t). The wavemaker integral in Eq. (6.124b) and the free-surface integral in Eq. (6.124d) may be approximated by (Miles, 1988a) -2\.
(6.125a)
0(y2)y.
(6.125b)
~ i [•]rfz+ £??(*,y,0[»] z =o+ (£); h
[ [.]dx* JyX
f
[,]dx - yx(z,t)[.]x=0
+
JO
The dimensionless wavemaker perturbation forcing parameter y is smaller than the Floquet parametric forcing parameter s because laboratory experiments demonstrate that the standing cross wave amplitude becomes larger than the wavemaker forcing amplitude as t —>• oo. The parameter ordering is, therefore, 2
0
2
2 Y
Y
or 0 < — < y < — < 1. (6.126a,b) s s
438
Waves and Wave Forces on Coastal and Ocean Structures
The higher order terms that are neglected are 0(e2), 0(ey), and 0(y2), but terms 0(y/e) are retained. Neglecting these higher order terms, the nondimensional Lagrangian integrals Eqs. (6.124a-e) are approximated by: Ly =
-
dXJZ,t) I J Uc(y,z,t) + rkr
2e
dt
b r
+0
dX(z,t) dt
(
ric(y,t)
dzdy -b=0
\ ( 4>c(y,z,t) \+Tr)P(x,t))\+TXr(pp{x,z,t)
dy x,z=0
0{sl),
+
(6.127a) d(j)p(x,z,t)
2T3/ /
h
b
+ £
11
I ~^^'"(
rd
+rr)p(x,t)J
dx
4>dy,z,t) \~|
\+rkrp(x,z,t))\. =Q *,z=
+ o(s2),
(6.127b)
4>c(y,z,t) +Tkr(j>p(x,z,t)
3ric(.y,t) dt dr}p(x,t) +r dt
1 rb r% 2 J-Jo
dzdy
( nc(y,t) \+rrip(x,t) xd_(
x
>ic(y,t)
dxdy
+rr)p(x,t)
4>dy,z,t)
dz \+rxr<j>p(x,z,t)
z=0
"I/.
n +rxr4>p(x,z,t)
X(z,t)
x_9_/
X
dt
ric(y,t)
( ric(y,t) V \+Tnp(x,t)j
^
dy
\+Trip(x,t)J
x,z=0
(6.127c)
1
+ Oie ) b
rf
2j-bJo t
£2 (drip(x,t)\2
/dr]c(y,t^
X f \ d x )
\
<Jr2
dy
2
'
dxdy
f9Vp(x,t)\2fdriest)
r dy x,z=0
+
0(ye).
(6.127d)
439
Nonlinear Wave Theories
Trial Functions The dimensionless wavemaker forcing x (z, t) is specified as z t f(z) = l-\— h *(z,0 = / ( z ) s i n - • P f(z) = l
for a full draft hinge,
for a full draft piston. (6.128a,b) A generic planar wavemaker function is given by Eq. (5.8) and Eqs. (5.9) in Chapter 5.2. The cross wave potential
(6.129a)
4>p(x,z,t) = [Qi(t)cosx
(6.129b)
+ Q2(t) sin x]exp[z/^],
where the dimensionless temporal variables q(t), Q\(t), and 2 2 ( 0 are the generalized coordinates or, equivalently, three degrees-of-freedom. The generalized coordinate q(t) and coordinate vector q(t) are not to be confused with the fluid velocity vector q(x,y,z,t) defined in Eq. (6.109a). The free surface displacement r](x, y, t) is a solution to the linearized inhomogeneous boundary value problem given by \[£p{r)p(x,t)}-fp(x,z,t)]\ V /
£»{r](x,y,t)}=
lpK
dx drjc(y,t) dy
' = 0, = 0,
i
=0'
(6.130a,b)
x =0
(6.130c)
y = ±b,
(6.130d)
where Eqs. (6.130c,d) are the contact line conditions and where the linear operators <£^(») are T
£p{r]p(x,t)}
= ^
d2TlD(x,t)
^V
~ VP(x,t),
z=0
(6.130e)
440
Waves and Wave Forces on Coastal and Ocean Structures
ft
t\
i d(f>p(x,Z,t)
fp(x,z,t)=aXr—z—
d(/)p(x,Z,t)
x\Xr—^- dt
dz d ric(y,t) vdy,t), z = o dy2 d
~,,
,
z = 0,
A
~
(6.130f)
2
fdy,z,t)
(6.130g) 0.
(6.130h)
Denoting ordinary temporal derivatives by overdots d(»)/dt = (•) and substituting Eqs. (6.130g,h) and Eq. (6.129a) into Eq. (6.130b) yields a solution for the dimensionless cross wave free surface displacement r]c(y, t) that is given by 1 ric(y,t) = -t(-q(t)a + Tiq(t))cos(y-b). (6.131a) x
i
Substituting Eq. (6.129b) and Eqs. (6.130e,f) into Eq. (6.130a) yields the following homogeneous plus particular solutions for the dimensionless progressive wave free surface displacement rjp(x, t) that satisfy the contact line condition in Eq. (6.130c): -aQi(t) T]p(x,t)
+
k2rxiQi(t)
cos*
K^x
=
+
XrTx
sm x -\
Y
ex
P
a -\
-ft
(6.131b) where x\ and x\ are defined in Eqs. (6.120s,u), respectively. All 0(e) progressive wave self-interaction terms with either (j>p(x, y, t) or r\p (x, t) are neglected (Jones, 1984). In order to evaluate the dimensionless Lagrangian components in Eqs. (6.127), the integrals will now be separated according to their order rather than to the surfaces over which they are integrated. This ordering of the Lagrangian integral components in Eqs. (6.127) is (1) LQ integrals that are order 0(1); (2) LE integrals that are order 0(E); and (3) Ly integrals that are order 0(y). Substituting the trial functions from Eq. (6.128a or b), Eqs. (6.129 and 6.131) with no dissipation a = 0, the surface integrals in the Lagrangian
441
Nonlinear Wave Theories
Eqs. (6.127) become L = L0 + Le + LY + 0(e2, ey, y2),
(6.132a)
where bT2
[f(G?(0 + Gi(0) + Qi(OG2(0]
2T-2
L0 =
bxtV \-^ [$X4rtx(Q2(t) + Q2{t)) + 4T Gi (0 QiiO]
(6.132b)
V$Tk
-y(92(0-92(0) £Z?rriT
T
•
sbT
(6.132c)
Le = - ^ 2 7 T^i - ? 2 ( 0 G 2 ( 0 ~ 2TIA.? Z-^TT9(04(OG2(0,
/i
—yb cos
a)
r 2 rL2 £r A
r>.rTi6i(0 e/3
GI(0(^2GI(0 + V^G2(0)
q(t)q(t)
Ly =
2?
) z
rzQj(t) +
2
q (t)-q (t)
\
+y£>sin
4 2 1 (A /"i4/)2 Q (0
2 r<-\
*?*
+ A2V^GI(0G2(0 +
\ L I I rQlit))
(6.132d) where f\ is an integral of Eq. (6.128a or b). The Lagrangian component LQ represents the free oscillations in the WMBVP, LB represents the Floquet parametric forcing of the cross wave by the progressive wave (Jones, 1984), and Ly represents the 100% nonautonomous perturbed (wavemaker forcing) components, respectively. The analogy to the nonautonomous Hamiltonian for a Floquet oscillator follows. Legendre Transform (Lichtenberg and Lieberman, 1992) The Legendre transform of the Lagrangian L(q(t),q(t),t) with respect to the velocities q(f) = (q(t), Q\(t), Qlit)) yields a Hamiltonian given by
442
Waves and Wave Forces on Coastal and Ocean Structures
(Lichtenberg and Lieberman, 1983) H(q(t),p(t),t)
= p(t)q(t) + Pi(t)Qi(t) + P2(t)Q2(t) -
L(q(t),q(t),t), (6.133) where p(t) = (p(t), P\{t), P2(t)) are the conjugate momenta corresponding to the generalized coordinates q(r) = (q(t), Q\{t), £?2(0)- The conjugate momenta p(t) are computed only from the free oscillations plus the Floquet parametric forcing components of the Lagrangian function (LQ+LS) according to (Goldstein, 1980) d(L0 + Ls) a w / '
PM=
'
=
1 2 3
Pl(t) = Pit) = bt-q(t) - (J^) P2(t)
= px{t) =
bZl F K
(6134a)
>>>
q(t)Q2(0,
(6.134b)
(t-Qx{t) + 7 ^ - 6 2 ( 0 ) ,
(ebxxTx\ P3(t) = Pl(t) = - — | —
, q\t)
(6.134c)
br2r2),2 / . 2T . r + —l- UG2(0 + 74-Gi(0 (6.134d)
Inverting Eqs. (6.134) for qi{t) yields 01 (0 = q(t) = —— + — — — 3 — , WA
q2(t)
/) m
= Qm =
. _ A ^ 3(0 = Q2(t)
*
(6.135a) sx2q2{t)
A.^f1(Q-(2r/g)ft(Q
^ F ^
YW^'
4
A TAJP2(Q-(2T/g)Pi(Q
=
^ r ^
+
£rg
2
(Q
2T^R-
(6 135b)
'
(6135c)
The parameter £ that denotes a length down the channel will now be selected as an ordering parameter because £ ^> 1. This approximation by order is both a way to simplify the Hamiltonian in Eq. (6.133) and to apply the long channel assumption. This approximation is applied only for determining the order of the coefficients in each variable grouping and not to scale the different combinations of variables. Substituting Eqs. (6.132, 6.134 and 6.135) into
443
Nonlinear Wave Theories
Eq. (6.133) yields the following Hamiltonian: #(q(f),p(0,0 = # o ( q ( 0 , p ( 0 ) + He(q(t),p(t))
+ Hy(q(t),p(t),t)
0(e2,ey,y2), (6.136a)
+
where the autonomous free oscillation component Ho(») is
ibr2 flb(q(O.P(0)
=
[m\(t) + Q\(t)) + Qx(t)Q2{t)\ 2 p2(0 + (p2(0 + p2(0) 1 2 — L +-b$q (t) + 2 2Z>£ ibpvH 2xPx(t)P2(t)
(6.136b)
bx2Y2X%2
'
where the autonomous Floquet parametric forcing component HE (•) is q(t)p(t)Q2(t) 2ri£A.3 ex q2(t)P2(t) 2riT§A.5
He(q(0,p(0)
(6.136c)
£T
-q2(t)Pl(t) Tir| 2 A.?Tx
I
and where the 100% nonautonomous perturbed component Hy(») from the wavemaker forcing is /
bTKriQiit) ^
H
Qi(t)(k2Pl(t)
+
71 y cos
P
p{t)q{t)
e/3
2
\
2ft ^P2(t))
m
2xQdt)(k2P2(t)
+
^Pi(t))
2
Hy(q(0,p(0) =
M ^x / , ^2nl z
)
2
\ P (t) 2 + ~r(t) 2b$ ' 2 {X2rPx{t) + Jx~P2(t))2
br Qi(t)-
-y sin | -
bx2T2l-2X?
/ J (6.136d)
444
Waves and Wave Forces on Coastal and Ocean Structures
The autonomous Ho(q(t),p(t)) + Hs(q(t),p(t)) components represent the free oscillations and the Floquet parametric forcing of the cross wave by the progressive wave (Jones, 1984). The perturbed Hy(q(t),p(t)) component that is due to wavemaker forcing is 100% nonautonomous and will not survive the averaging theorem that is required for the Wiggins and Holmes (1987) extension of the GMM. All of the Hs(q(t),p(t)) terms are of 0(e) and are inversely proportional to the long channel parameter £. If £ -*- oo is assumed, all 0(e) terms vanish. Because the He(q(t),p(t)) terms represent the Floquet parametric forcing of the cross wave by the progressive wave, allowing the He(q(t),p(t)) component to vanish would remove all possibilities of finding chaos in the system because the nonlinear oscillations of the cross wave is required for a homoclinic/heteroclinic orbit to form in the phase space (Wiggins, 1988). Although the nonautonomous Hamiltonian in Eq. (6.136d) is homomorphic to the nonautonomous Floquet Hamiltonian for a parametrically excited pendulum, a subtle but important distinction exists. The 100% nonautonomous component in the Floquet Hamiltonian is due to parametric forcing; while, in contrast, the 100% nonautonomous component in Eq. (6.136d) is due to the external wavemaker boundary forcing that generates the progressive wave that then parametrically forces the cross wave. This sequence from external forcing to parametric forcing distinguishes cross wave parametric forcing from the Floquet forcing of a parametrically excited pendulum. Damping Forces The scaled Hamilton's principle in Eq. (6.111a) may be expressed as (Guenther and Schwerdtfeger, 1985)
8 f2 Xdt= IJ^DiSqtiOdt, Jti
Jh
(6.137)
i=1
where D = {D\,D2,03) is a set of generalized components of the damping force corresponding to the set of generalized coordinates q(t) = (q(t), Q\ (t), Q2(t)). Substituting for dS^ fromEq. (6.110b) for Dfj/Dt in Eq. (6.111c) and the scaled variables from Eqs. (6.120) into Eq. (6.137) yields the following scaled action integral of the Lagrangian: S I' £dt = - I2
Ju
f
f ad
JH J-bJo
3z
(6.138)
445
Nonlinear Wave Theories
Equating the variation of the action integral of the Lagrangian in Eq. (6.137) to Eq. (6.138) gives a damping variation of 3
Jti ~[
Jti
J-bJo
d
dy dt, (6.139a)
where 8r/(x,y,t) = >
%(?) +
(6.139b)
Integrating the Sen (t) term in Eq. (6.139b) by parts with respect to time t and noting that 8qi (t) vanishes at the arbitrary temporal values of t\ and t2 reduces Eq. (6.139a) to
Dt=a [ [ J-b JO
d
dx dy,
i = 1,2,3.
(6.140) Substituting the scaled velocity potentials from Eqs. (6.129) and the scaled free surface displacements from Eqs. (6.131) yields the following damping force components: „ ba£ ( a (6.141a) Di = — \q(t) + -q{t)\, D2 D3 =
ba^tiT7
Qi(t) +
4 Iba^xixT2 2 nA£TA 2T
1 / . +7
a —~Qi(t)),
Til2/
(6.141b)
Qiit) + a 02(0
a
(6.141c)
2I(0 + - T 2 I ( 0
Substituting q{(t) from Eqs. (6.135) into Eqs. (6.141) converts the damping forces to the canonical variables q(0 and p(0 according to
(aebr\
°i = I - ) P(0 + I ^ 3
a2b%
) 4 ( 0 0 2 ( 0 + I '-—- I qit),
(6.142a)
446
Waves and Wave Forces on Coastal and Ocean Structures
2a T
a 2 U ri
Pl(t)
Pl(t)
D2 = q\t) 'aebTz2^
+
( ^
D3 =
Gi(0 +
+
Gi(0
X2TX
(asbTx P2(0 +
(6.142b)
a2b$T2
q2(t) (6.142c)
a2bS-T2 X2 TX
Qi(t)
The long channel ^parameter 2 a 2 / ? r r 2 £ denotes a dimensionless length down the channel and is selected as an ordering parameter because £ S> 1. This approximation V ^rxX is applied only for determining the order of the coefficients for each variable grouping and not to scale the different variable combinations. Hamilton's Equations of Motion Following the Legendre transformation in Eq. (6.133), the dynamics of the damped dynamical system may be determined from Hamilton's equations of motion (Scheck, 1990) according to qi(t)
dH dpi(t)'
dH Piit)
dqtit)
-A(q(0,P(0),
i = 1,2,3, (6.143a,b)
where D,-(q(0.p(0) is a s e t of generalized components of the damping force that are computed from the set of generalized coordinates q(t) and conjugate momenta p ( 0 from Eqs. (6.142). If a canonical transformation Q = Q(q, p, t) and P = P(q, p, t) is obtained with a generating function F(u(t),\J(t),t) from the generalized Herglotz algorithm (GHA-Type I, Appendix A; Fadel, 1998 or Orum et al., 2000) then the Hamiltonian K(*) for the new set of variables is * ( Q , P,t) = H [q(Q, P, t), p(Q, P, 0 , t] -
dF(u(t),\J(t),t) dt
(6.144)
where w, (t) and £/,• (t) may be any of the four combinations of old and new variables (Goldstein, 1980; Appendix A; Fadel, 1998 or Orum et al., 2000).
447
Nonlinear Wave Theories
The transformed Hamilton's equations of motion from Eqs. (6.143) are 3
^ ( ^ i L + y n . M l , J dPi(t) f^ dPi(t)'
i = l,2,3,
(6.145a)
3
Pj(r)
^
.
r
n
M
,
i = 1,2,3,
A(Q(0,P(0,0 = A[q(Q(0,P(0,0,P(Q(0,P(0,0L
(6.145b)
« = 1,2,3, (6.145c)
where D(«) is the transformed set of generalized components of the damping force. After each canonical transformation, the damping force components may be introduced into all of the six evolution equations (6.145a,b) if qj(t) is a function of both Q(t) and P(r). Three Canonical Transformations for the GMM In order to apply the GMM to the dynamical system in Eqs. (6.145), three canonical transformations are required: First to simplify the 0(1) terms in Ho in Eq. (6.136b); second to simplify the 0(e) terms in Hs in Eq. (6.136c); and third to suspend the 0(y) terms in HY in Eq. (6.136d). The GHA (Appendix A) is required for two of the following three canonical transformations that are required in order to suspend the Hamiltonian in Eqs. (6.136) so that the GMM may be applied. (i) Rotation of Axes [q(0,p(0] =>• [q(0,p(0] The Hamiltonian component HQ in Eq. (6.136b) contains the cross-product terms 0 i ( 0 0 2 ( 0 and Pi (0^2(0 that may be removed by a rotation of axes canonical transformation. The new canonical variables and the transformed Hamiltonian components are denoted by tildes (•) that do not represent dimensional variables in this canonical transformation. The rotation transformations are q(t) = q(t),
p{t) = p(t),
(6.146a,b)
(2i(0 = Q i (t) cos 6 + Q2(t) sine,
Pi(t) = Pi(t) cos 9+ P2(t) sind, (6.146c,d)
02(0 = 0 2 (Ocos0 - gi(Osin0,
Piit) = P2(t)cose -
Pi(t)sin9, (6.146e,f)
448
Waves and Wave Forces on Coastal and Ocean Structures
that are canonical because the Poisson bracket conditions are satisfied; viz. Ui{t),qj(t)\^
= 0,
lqi(t),pj(t)\p
= 0, (6.147a,b,c) where the Poisson bracket M for any two canonical variables Rj (t), Rk(t) of the transformed set for the transformations in Eqs. (6.146) is N
r
= §ij,
dRj(t)dRk(t) dPi(t)
_
lRj(t),Rk(t)]qp = f^ld Yl qi(t)
lpi(t),pj(t)\p
dRj(t)dRk(t) 3/7,(0 dqi(t)
(6.148)
where <5,7 is the Kronecker delta function (vide., Chapter 2.2.3). Substituting the angle 6 — TC/A into Eqs. (6.146) will eliminate the cross-product terms Q\{t)Qi(t) and P\{t)P2{t) in the Hamiltonian component HQ inEq. (6.136b). The transformed Hamiltonian H(q(t), p(t), t) may be computed by substituting Eqs. (6.146) into Eq. (6.136b) with 0 = n/A and is given in component form by H=>H = H0 + He + HY(t) + O (e2,ey,y2^
,
(6.149a)
where bF
q{t)
+
Y H0 =
+ +
bT2
Wj +
^ o2,^ , •Gf(0 + 2bp2rW 2 i2
Pht) '2
^ o 2 ^ ,
{Q\(t) - Q\{t)) +
eT
(6.149b)
2bp2F2% x{P2(t) -
P2(t))
br2r2X?i;2
q(t)p(t)(Q2(t)-Qi(t))
2V2TI|A.3
HF
£ T ( 2 t + ^ T X ) , 2z q (t)Pl(t)
2^2xxT^l?rxx
e r ( - 2 r + ^ r A ) _ z2 2V2tir§ 2 A9rx
q (t)Pi(t)
(6.149c)
449
Nonlinear Wave Theories
M
/ ^ r v n ( < 2 i ( 0 + <22(0) , P(t)q(t) 2/3$ r
(Gi(0 + G2(0) (A.2-v^)A(0 y cos I —
l+(X.2r+S?)P2(t)_
20W r(Qi(t)
Hy =
+
(k2 +
Q2(t))
^)P2(t)
\
' —y sin
2 - ( G i ( 0 + G2(0) 2 + 2? O
-
^
x
G) 2bri2rH2Xf
/ (6.149d)
All of the terms that are proportional to the dimensionless surface tension parameter r in Eqs. (6.149) are of order 0 ( § - 2 ) . The coordinate-dependent q(0 terms are of higher order of magnitude O (£) than the momenta-dependent p(0 terms. The Hamiltonian component He in Eq. (6.149c) contains all the Floquet parametric forcing of the cross wave by the progressive wave. The 100% nonautonomous perturbed component HY in Eq. (6.149d) due to wavemaker forcing will not survive the averaging theorem that is required for the GMM. The free oscillations component Ho in Eq. (6.149b) motivates the next canonical transformation to action-angle canonical variables. The first three energy brackets in Eq. (6.149b) will be the action set of the new canonical variables. (ii) Action/Angle Transformation [q(0,P(0] =>• [
P;2/ L
P (i) 2b§
(6.150a) l W
2 2
2bp r $'
(6.150b)
450
Waves and Wave Forces on Coastal and Ocean Structures
hit) = p2(t)
^ 5 - 2 ^ , 22(0 + 2 "^-' •
Pjit) 2bp2r2f
(6.150c)
that satisfy the Poisson brackets in Eq. (6.147c) given by lPi(t),Pj(t)lqp
= 0,
(6.151a)
i,j = 1,2,3,
and the non-zero determinant (Fadel, 1998) dp(t) dpit) dp{t) dq(t)
3<2l(0
3<22(0
dPi(t)
dPi(t)
9Pi(0
3-7(0 dP2(t)
3<2i(0 dP2(t)
302(0 dP2{t)
3?(0
dQxit)
dQ2(t)
= b3r4^q(t)Q1(t)Q2(t)
jL 0. (6.151b)
The Herglotz auxiliary functions Xi (p(0. q(0) (Fadel, 1998) are chosen to be in a ratio form given by
XtW-4? Pit)
^3(0 =
*2(0 = ^ "" Plit)
2i(0
(6.152a-c)
Plit)
that satisfy the nonzero Jacobian requirement " {P2it) + 3(p(f),X(Q) 3(q(0,p(0)
b2r4p2$2Q2(t)) Plit) Plit))2
b^^T^Hp{t) 2
2 4 2
2
2
2
x (P (t) + b T p i; Q (t)){p (t)
#0. 2 2
+ bH q {tj)
(6.153) Solving the Herglotz algorithm for qi(p,X) gives the following three new positive definite canonical action variables p(X(?),p(0): 2b$p(t) i+bH2x2(t)'
(6.154a)
Piit) = P i ( 0 =
2b$p2r2Px(t) i+b2t;2p2r4x2(ty
(6.154b)
hit) = p2(t) =
2bw2r2P2jt) i+b2$2p2r4x2(t)'
(6.154c)
Plit) = pit) = 1
\
451
Nonlinear Wave Theories
The generating function F{; •) for the canonical transformation [q(0> P(01 =$ [qO), p(OJ may be computed from the following integral: F(p(p(0,X(0),p(0) A = - V
[Xi(t) ( dp(t) dPi(t) ~ dP2(t)\ / q(t)-^-+ Qi(t)—-r^ + Q2(t)—^— dX'At)
/=i
b$pXx{t) 2
+ jp(Otan"1(^Z1(0)
1 + bH XJ(t) b^r fi P1(t)x2(t) + ^A(Otan- 1 (^r 2 X 2 (0) i + b2p2i;2r*x2(t) bHT2p2P1{t)Xi(t) , ^ ^ ^ ^ + ^P2(Otan-1(^r2X3(0) 2 2 2 4 i+fe )8 § r x?(0 2 2
(6.155)
where an arbitrary function of time is neglected. The remaining set of three new angle variables q(q(p(0>X(0), p(p(t),X(t))) are *(0= > ?7(0^r-- + ^ 9/?,(?)
^r-77 9p,-(0
,
i = l,2,3,
(6.156a)
where & ( 0 = 9(0 = UurHbSXiit)),
(6.156b)
$2(0 = C i ( 0 = Ptm-l(bl3t; T2X2{t)),
(6.156c)
$3(0 = <22(0 = i S t a n - ^ f e ^ r ^ a C O ) .
(6.156d)
Solving the Herglotz algorithm for [q(q(0>P(0)>P(q(0»P(0)] yields the following canonical transformation to action/angle variables: (0 = J ^ P sin$(0,
p{t) = y/2b$p(t)cosq(t),
(6.157a,b)
(6.157c,d)
452
Waves and Wave Forces on Coastal and Ocean Structures
Ql(t) = J2^-
sin ( ^p-Y
p
2(t) = rPy/2b$P\(t)cos
62(0 \ (6.157e,f)
where the action variables p = (pit), P\ it), Pi(t)) are positive definite. This transformation may be confirmed to be canonical by the Poisson bracket conditions by Eqs. (6.147) given by [[;(?), <7;(0Iqp = 0,
lp~i(t), Pj(t)\j, = 0. (6.158a-c) The transformed Hamiltonian H(i\(t),\t(t), t) is now given in component form by H^H
lqiit),pjit)Jqp
= &ijt
= H0 + He + Hyit) +
Oie2,sy,y2),
(6.159a)
where [pit) + Plit) + p2it)] H0 =
Awco^MA _ tew'
+
P
2§
sin
61 (0
2 x
' ^ \ o 1* 2 (Q2(o\ ^ (P2it)\ . 2 /e 2 (o N (6.159b)
/.VA cos I 2qit)
Qiit)\ ' —I
/ Plit) -cosl2(0 + — — I HR =
spit)
Ayfbpfyi
cos I 2^(0 —
Qi(t)\
-VA(0 — cos I 2qit) +
P I Qi(t)
P
Nonlinear Wave Theories
453
Gi(0
(2T+^TX)/A
(0
1 / Gi(0 •-cos[2§(0 + 1 L.M --cos ( 2?(f)
Q\(f)
{¥) +(2T -
^T^yfhV)
\ cos (25(0 + ^ f ) -1 cos ( 2 ^ ) - ^ j j (6.159c)
7 ^ ( 0 sin ( ^ ^ ) + 7 ^ ( 0 sin
(A2-V?)(2T+^TA)
y cos
^2r,
+
^
2
t
22 (0
\ P J.
A
, p(f)sin(2^(0)
/fy =
fil(0
yPi (0 cos
m (X;}-yft)y[h
(0 cos
Gl(0
4rf +(x2 + v^)V^wcos y~-\
(i) +
^sm(^)+y^sin(M>)' +p(0(sin 2 9(0 - cos2 q(t)) (6.159d)
454
Waves and Wave Forces on Coastal and Ocean Structures
The free oscillation component HQ in Eq. (6.159b) and the Floquet parametric forcing component He in Eq. (6.159c) are functions of the canonical variables [q(0>P(OL The perturbed component Hy in Eq. (6.159d) that is due to wavemaker forcing remains 100% nonautonomous and will not survive the averaging theorem that is required for the GMM. The dependency of Hs on [<2i(0, Q 2(0] and some of the nonautonomous terms in Hy may be suspended by applying the following final nonautonomous Hamilton-Jacobi canonical transformation. (iii) Hamilton-Jacobi Transformation [q(0>P(0] =>• [q(0>p(0] The main objectives of this final canonical transformation are: (1) to obtain a completely integrable unperturbed Hamiltonian system where Q\(t) and <22(0 are the cyclic coordinates, and (2) to suspend some of the nonautonomous terms in HY in order to apply the GMM. This transformation will include near resonance cases by defining a detuning parameter that transforms the first square bracketed 0(1) term in Eq. (6.159b) [p(t) + P\(t) + P2(t)] to be 0(e) (Holmes, 1986). Both <2i(0 and Q2(t) may be eliminated from the autonomous Hamiltonian component Hs in Eq. (6.159c) by the following nonautonomous transformations for q(q(0> 0 :
qi(t) = q(t) = ^+q(t)
+ Qi(t),
(6.160a)
q2(t) = Qi(t) = t + 2PQ!(t),
(6.160b)
hit) = (22(0 = 3/ + 20(q(t) + Qi(0 + (22(0),
(6.160c)
where Eqs. (6.160) satisfy the Poisson brackets equation (6.147a) Ui(t), qj(tnqp
= 0,
i,j = 1,2,3,
(6.161a)
455
Nonlinear Wave Theories
and the non-zero determinant dq(t) dq(t)
dq(t) 30l(O 302(0 30l(O 30l(O 30i(O = 4>S2 ^ 0. (6.161b) 30l(O 302(0 dq(t) 302(0 322(0 302(0 3(0 301 302(0 The Herglotz auxiliary functions X;(q(0,p(0) a r e chosen in a ratio form X(0 = P ( 0 / q ( 0 to satisfy the nonzero Jacobian condition
3(q(0,X(0)
4£2
3(q(0,p(0)
9(001(002(0
#
(6.162)
Solving the Herglotz algorithm for p(q(0, X ( 0 , 0 yields PI(0
/>(O = 4r7rW(O-0i(O),
P2(t) = P\(t) =
2? -x2(t)
('-0l(O),
(6.163b)
(2^ + 2 ^ ( 0 - 0 2 ( 0 ) .
(6.163c)
2/3 P3(0 = Pl(t) =
(6.163a)
2/3the nonautonomous canonical transformation The generating function F (•) for ma [q(0,p(0] =>• [q(0.p(01 Y be computed from the following integral: F(q(0,p(q(0,X(0,0,0 3
i=\j
"Xi{t) L
9(0^777^ + 0l(Or^777^ + 02(0 dX[{t) 3X,'(0 ax;(o
I X j ( 0 ( 2 ^ ( 0 - 01 (0) 2 + X2{t){t - 0 i (0) 2 ' "4/J2 +X3(t)(2t + 2l3q(t)-Q2it))2
rfx;(o (6.164)
where an arbitrary function of time is neglected. The remaining set of three old variables p(p(q(0, X ( 0 , 0 , q(q(0,0) m ay be computed from Piit) =
-^qj{t) j=i
dp jit) dqt(t)
8F dqiit)'
i = 1,2,3,
(6.165a)
456
Waves and Wave Forces on Coastal and Ocean Structures
where 1 1 T ((7Rn(A 2 0 $ ( O— - G i ( 0h,(t\\Y,(t\ )Xi(0 P\(t) = p(t) = — +(Zt + 2Pq(t)-Q 2(t))X3(t) 2p
hit) = A(0 = —^[(2p^(0 - Qxit))xxit) + it-
(6.165b)
fii(0)x2(0], (6.165c)
hit)
-1 = P2it) = —^ i2t + 2fiq(t) - Qlit))X3it).
(6.165d)
The Herglotz auxiliary functions X ; (q(f), pit)) may be substituted to obtain p(p(0) according to hit)
= Pit) = Pit) - Plit),
(6.166a)
hit) = A(0 = ^ ( ^ i ( 0 - p(0),
(6.166b)
P3(0 = A(0 = ^ P -
(6.166c)
2p Because the action variables Eqs. (6.166) must be nonnegative, the ordering constraint on the new variables pit) is Plit) > Pit) > Piit) > 0.
(6.166d)
The transformations from Eqs. (6.160 and 6.166) may be confirmed to be canonical from the Poisson bracket equations (6.147). The first square bracketed 0(1) term in Eq. (6.159b) may be transformed by Eqs. (6.166) to
[p(t) + A(0 + Plit)] =(l-~\
(p(t) - Piit)) + ^0-.
(6.167)
The primary Floquet resonance condition is, approximately, (Holmes, 1986) 2S)C « cop,
(6.168a)
so that 2wc-cop
= 0(e).
(6.168b)
A detuning parameter SI may now be defined as
x
l
=,a
-k)-{ -h) -
(6 168c)
-
457
Nonlinear Wave Theories
where Q = 0(1) is the detuning parameter that defines the primary parametric Floquet resonance ratio a>p : u>c = 2 : 1. Substituting q(q(0>0 from Eqs. (6.160) and the Herglotz auxiliary functions X,-(q(f),p(0) m t o Eq. (6.164) yields the following generating function in terms of (q (f), P (0 > 0 :
p(t) ( -q(t) + 27jfii(0) + JoWW
~ Gl(0)
F(q(f),p(0,0 + 2^P2(t)(2t P
+
20q(t)-Q2(t))
(6.169) The transformed Hamiltonian H(q(t), p(t), t) is given in component form for both the autonomous and the nonautonomous terms by = HQ + HQ(t) + He + He(t) + HY + HY(t) + 0(s2,sy, y2), (6.170a) where the autonomous Hamiltonian components in Eq. (6.170a) are H-
— ^H at
HQ=
-
0r
m +' i^rx\ 1
(Px(t)-P(t))
1
+ -^P + 8yS^
2fA.«n2.
Pl(t),
(6.170b)
sQ(p(t) - P2(t)) HF =
e(p(t)-P2(t))JPi(t)-p(t) + |
[
4y/2bp$sX.9rrixk
(6.170c)
x[(2£r 2 - §^T X (A.2 - fix)) cos(2 9 (0)]
Hy =
2sy/2iiP +y
2jS-l
V/,i(0-p(Osin(2j2i(0) (6.170d) ( f t ( 0 - p ( 0 ) s i n 2 ( 9 ( 0 + Gi(0)
and where the nonautonomous Hamiltonian components Ho(t),He(t) and Hy(t) in Eq. (6.170a) are given by Fadel (1998). The free oscillation component Ho in Eq. (6.170b) depends only on the canonical variables p(t). The Floquet parametric forcing component He in Eq. (6.170c) is independent of
458
Waves and Wave Forces on Coastal and Ocean Structures
the canonical variables Q\(t) and Q2(t). The autonomous perturbed component Hy in Eq. (6.170d) that is due to wavemaker forcing will now survive the averaging theorem that is required for the application of the GMM.
Combined Transformation [q0rig(t),porig(i)] => [q(0.p(0] The original canonical variables following the Legendre transformation by Eq. (6.133) must now be expressed in terms of the final transformed canonical variables [q(0>p(01 m order to compute the generalized damping forces D ( q ( 0 , p ( 0 , 0 in Eqs. (6.142). The original canonical variables are designated as [qorig(t),Porig(t)] and may be expressed as functions of the final transformed canonical variables [q(t),p(t)]by successive substitutions of the transformed variables in Eqs. (6.157,6.160 and 6.166a-c) into Eqs. (6.146); i.e.
qorigit) = J^p(t)-P2(t)
sin (q(t) + Qi(t) + ^-\,
(6.171a)
Portgif) = J2bk7P(t)-Pi(t)
cos (q(t) + fii(0 + ±-\ ,
(6.171b)
l
6u(0 = rj2FM
y/Pi(t)-P(t)
sin ( 20i(O + -g
+^Pl(t) sin 2(q(t) + Gi(0 + Qi(t)) +
3>t
J/J
(6.171c)
Piorig(t) = rv
bfc 2
y/Pi(t)-p(t)cos +V^(0cos
2<2i(/) +
p
2{q{t) + Gl(0 + 22(0) +
3r
Jjl (6.171d)
459
Nonlinear Wave Theories
- V ^ i ( 0 - P ( 0 s i n ( 2gi(f) + -
1
3?
+ 7 ^ 0 ) sin ( 2( 9 (0 + g i ( 0 + 22(0) + j . j (6.171e)
-y/Pi(t)-p(t)
bft
*V,-,(0 = r
cos ( 2Qi(0 + -
+ ^ / ^ 0 ) cos 2fo(f) + Qi(0 + g 2 ( 0 ) +
3t
1)1 (6.171f)
The combined transformations in Eqs. (6.171) may be applied to obtain the transformed set of generalized components of the damping forces D(q(t), p(t), t) and the transformed Hamilton's equations of motion. Transformed Damping Forces D,-(q(f), pif), t) The damping forces in Eqs. (6.142) rewritten in terms of the original canonical variables designated as [q O rig(0>Pon'g(0] are D, =
a *\
(asbr\ (a2b^\ Porigit) + —J-T qorig(t)Q2orig(t) + —r V Tl \1XlK' '
a \ D2 =
/ 2ar
'asbYx2\ ^
, Jong'
i
p
q0rig(t), (6.172a)
*W0
(a2b$Y2 ,
(6.172b)
A.?Tx
asferr
x? J w*) + ( ^ r r ,(0 ^ D3 =
'2a26tr2
+
* ^2
, <x2&§r2 .
(6.172c)
,e w (o + (-^—le^w r^k
Substituting the combined canonical transformation in Eqs. (6.171) into the damping forces in Eqs. (6.172) yields the following transformed damping
460
Waves and Wave Forces on Coastal and Ocean Structures
forces as functions of the final transformed canonical variables [q(t),p(t)]\ JTMa, . / t ^ — > _ V p ( 0 - f t ( 0 c o s q(f) + Qi(t) + — Tl J2Ma2 xf
£>! =
+
\
,
ea
2A#V?r 2
Iff
/ V
t 2jS
V/KO - ft(0 sin U + Gi(0 + ^ ) rin(?
(6.173a)
Vft(0 sin (| 22(«(r) + g i ( 0 + g 2 (0) + y - V ^ i ( 0 - P ( 0 s i n ( 2Qi(0 + T>pcir(2x + kj%Tk)
V ^ i ( 0 - P ( 0 c o s f 2fii(0 + -M
Tl r A ^V2f
/^ar(-2t+x4f^)
nr^?V2f D2 =
Wrl
V^Wcos 2(^(0 + 6 i ( 0 + g 2 (0) +
3?N
-(P(0 - *2(0) sm2 \q(t) + Q\{t) + —\ JP\{t)-p(t)sm\2Qx(t)+t-\
a2rV5|"
TTMOsin ^2( 9 (0 + Gl(0 + 22(0) + ^ ) (6.173b) - V A ( 0 - P ( 0 c o s >I 2Gi (0 + + J ) (2(21(0
V V2X2T! /
+ViWcos {2(q(t) + Qi(t) + g 2 (0) + ^ J (p(t)-P2(t))
D3
sin2 U (0 + 2i(0 + ^ ) ((0
+ | g 2 r ^ ? ^2 t X ) W f i ( 0 - p ( 0 . i n « 2 f i 1 ( 0 + ^T V2M
2
a rVfc(2r+^TA)
x6r2V2M
N/P2l0sin
2(^(0 + e,(/) + 2 2 (0) +
-6/
(6.173c) that are all completely autonomous.
461
Nonlinear Wave Theories
Averaged System The transformed Hamilton's equations of motion in Eqs. (6.145) may now be rewritten in terms of the original [q0rig(t),Porig(t)] and the final [q(0.P(01 canonical variables as dH
Pi(t) =
d(qorig(t))j dpt(t) '
8H
d(qorig(t))j dqiit) ''
D dqiit)-^-7^-J2 r 7=1
1,2,3,
(6.174a)
i = 1,2,3,
(6.174b)
where H(q(t),p(t),t) and D(q(t),p(t),t) are the transformed Hamiltonian and the transformed damping forces as functions of the final transformed canonical variables [q(t), p(OL respectively. Because it is difficult to analyze the full nonlinear nonautonomous system, an averaging method may be applied to obtain an autonomous system (Umeki and Kambe, 1989). Following Holmes (1986), an averaged system of first order ordinary differential equations may be obtained by averaging the autonomous Hamiltonian components in H(q(t), p(0, t) in Eqs. (6.170) over the dimensionless cross wave period In according to {H) =
{H0(p(t),Pi(t),P2(t))) \+(He(q(t),p(t),Pl(t),P2(t))) +(Hy(q(0,p(t),Pi(t),P2(t),Qi(t)))
fa
J__
(I
(Pi(0-p(0) Bx P
' Pl(t)
Q(p(t) - P2(t)) +e
+
'(p(t)-P2(t)WPi(t)-P(0
x[(2£r 2 -$X4rxx(X2r -0r))cos(2?(f))]
+y
\2eJlW
(2fi-Vs
IV^i(0-p(0sin(2Gi(0) (P2(t)-p(t))
sin2{q{t) + Qi(0) (6.175)
462
Waves and Wave Forces on Coastal and Ocean Structures
Substituting the transformed damping forces in Eqs. (6.173) and the combined transformations in Eqs. (6.171) into Eqs. (6.174) and then averaging over the dimensionless cross wave period 2n yields the following averaged system of first order ordinary differential equations: 2P1(t) + P2(t)-3p(t) cos(2q(t)) JPi(t)-p(t) sin(2£h(0) , -yc\ SyfPl{t) - p(t) 'd2(Pi(t)-p(t)) + d3(P2(t) Pit)) sin(2<7(0) —as y/P\(t)-p{t) (6.176a) —a\ + eQ + ecti q(t) =
P(0 =
4sa3y/Pi(t)-p(t)(p(t) P2(t))sm(2q(t)) -2yc2(P2(t) - p{t)) cos(2(q(t) + g i ( 0 ) ) +ad4P2(t) p{t) +2aed2JPi(t)
- p(t)(p(t) -
(6.176b)
P2(t))cos(2q(t))
— c i y/Pi(t)-p(t) cos(2 Q j (/)) -2yc2(P2(t) - p(t))cos(2(q(t) + Q\(t))) Piit) = • (4X2r - 1)(P 2 (0 - P(0) - Pi(t) +a 2X2 n -asd5^Piit)-pit)ipit) P2it))cosi2qit))
(6.176c)
Plit) = aPlit)
(6.176d)
Gi(0 =
62(0
2, , . . (P(0-P2(t)) cos(2qit)) a\ + ot^df, + ECLI r y/P\(f)-p{t) I sin(26i(Q) (P2(f) - Pit)) . n ._ +yc\—e7V ===== + a £ t f 3 7 = = = = = = sin i2qit)) ^i(0-p(0 ' " " " V A ( 0 - p ( 0 (6.176e)
_ Ja 2 + «2^7 - sQ - 2Eais/P\it) - p(t)cos(2q(t)) \+yc2smi2iqit) + fii(0)) + usd2^P\it) - pit) sin(2$(f)) J (6.176f)
.}•
463
Nonlinear Wave Theories
where the coefficients a;, c, and d\ are
A8-l\ (ifiT2 - i-kArTXQ?r - M\
dn = rl
ds
r+
d
1/-2
r(2r+A 4 gr A )
X^rk(l2r-l) A.^nr A
l2
(TXf-Xfa) — —, „ .
=
-,
d(, =
-T+^Tk
T + I ^ T A
The RHS of Eqs. (6.176) are independent of the variable Q2O); and the ordering constraint in Eq. (6.166d) on the final canonical transformed variable p(0 is P\(t) > p(t) > Pi(t) > 0.
(6.177)
The system of first order ordinary differential equations (6.176) is now suitable for the application of the GMM. Application of the GMM The GMM determines the existence of transverse homoclinic points, i.e. transverse intersections between the stable and unstable manifolds to any invariant sets of the perturbed system given the existence of a homoclinic/heteroclinic separatrix to a hyperbolic invariant manifold in the unperturbed (undamped a = 0 and unperturbed y = 0) system (Wiggins, 1988, p. 336).
464
Waves and Wave Forces on Coastal and Ocean Structures
Geometric Structure of Unperturbed Phase Space (a = 0, y = 0) The unperturbed vector field [q(0,p(0] may be obtained by equating to zero both the perturbation (wavemaker forcing) parameter y = 0 and the dissipation parameter a = 0 in the evolution equations (6.176) and obtaining •m J. o ^ (2Pl(t) + P2(t)-3p(t)\ q(t) = -a\ + eSl + ea 3 cos(2q(t)), \ y/P\(t)-p(t) J (6.178a) p(t) = 4ea3y/Pi(t)-p(t)(p(t)
- P2(t)) sin(2q(t)),
Pi(t) = P2(t) = 0,
(6.178b) (6.178c,d)
G i ( 0 = ai + e a 3 ^ 7 = = = cos(2 9 (0),
(6.178e)
Q2(0 = fl2 - £ ^ - 2ea3^Pi(jt) - p(t)cos(2q(t)).
(6.178f)
The unperturbed vector field a = y — 0 in Eqs. (6.178) has the form of a three degrees-of-freedom Hamiltonian system with (q(t),p(t),Pl(t),P2(t),Ql(t),Q2(t))eT1
x l 1 x E 2 x T2.
(6.178g)
An important consequence of the Hamilton-Jacobi canonical transformation is that the unperturbed (y = 0) Floquet Hamiltonian {H) in Eq. (6.175) is given by {H){
=0)_\(Ho(p(t),Pi(t),P2(t)))
+ {He(q(t),p(t),Pi(t),P2(t)))\
(6 179)
'
and is independent of the variables Q\(t) and Q2(t). Hyperbolic Saddle Points For every Pi(t) and P2(t) e R2, the [q(t) — p(t)] components of the unperturbed vector field a = y = 0 possess an hyperbolic saddle point that varies smoothly with both P\(t) and P2(t). The solution from Eq. (6.178b) for p(t = 0) and for p(0) = 0 is p(0) = P2(0),
2q(0) ^nn,
n = 0,1,... .
(6.180a,b)
Substituting Eq. (6.180a) into Eq. (6.178a) with q (t) = 0 and solving for q (0) gives
465
Nonlinear Wave Theories
where Pi(0) > P2(0) + (ai~J^
.
(6.180d)
The fixed point [q(0), p(0)] given by Eqs. (6.180c,a) is an hyperbolic saddle point provided that the determinant Hit)
dq(t)
dq(t) dp(t) Bq(f)
3/7(0 dp(t) 9p(t)
( a i - £ ^ ) 2 - 4 ( P i ( 0 ) - P 2 ( 0 ) ) e 2 a | < 0,
(6.180e)
t=0
is negative and provided that [q (t), p(t)] are given by Eqs. (6.178a,b), respectively. The symmetry properties of Hamiltonian systems imply that the stable and the unstable manifolds of the hyperbolic saddle point [q(0),p(0)] that are given by Eqs. (6.180c,a) have equal dimensions (Abraham and Marsden, 1978). In the full six-dimensional phase space in Eq. (6.178g), the unperturbed system a = y = 0 has a four-dimensional (M.2 x T 2 ) normally hyperbolic invariant manifold (with boundary dM) given by the union of the hyperbolic saddle points q(0) and p(0) in Eqs. (6.180c,a) according to
M
-[
eT'xR'xR2xT2:q(0),p(0)
J"
(6 181)
"
The normally hyperbolic invariant manifold M has a five-dimensional M.1 x R2 x T 2 stable manifold Ws (M) and unstable manifold W(M) that coincide along the five-dimensional heteroclinic manifold H = Ws(M)nWu(M)-M,
(6.182)
where WS(M) and WU(M) are the set of initial conditions that approach the hyperbolic saddle points on Mas t ->• ±oo under the action of the unperturbed flow (Wiggins, 1988, p. 354). The unperturbed locally stable and unstable manifolds of the normally hyperbolic invariant manifold Mmay be denoted as Wfoc(M) and W"0C(M), respectively. Because P\(t) = P2(t) = 0, no trajectories may cross the boundary of the normally hyperbolic invariant manifold dM. However, in the perturbed system, P\(t) and Piit) may not be equal to zero, and, therefore, trajectories may cross the boundary dM.
466
Waves and Wave Forces on Coastal and Ocean Structures
Dynamics on M The unperturbed vector field a = y = 0 that is restricted to the normally hyperbolic invariant manifold M may be computed from A ( 0 = 0,
P2(t) = 0,
Gi(0 = ai,
Q2(t)=a2-ai.
(6.183a,b) (6.183c,d)
The unperturbed vector field a = y = 0 restricted to Mhas the form of a two degrees-of-freedom (i.e. four-dimensional phase space T ' x l ' x T 2 ) that is a completely integrable Hamiltonian system with the Hamiltonian given by the level energy surfaces H(Pl(t),P2(t))
=
(H)(q(0),p(0),y=0)
= alPl(t) + (a2-ai)P2(t)=E,
(6.184)
where E is a constant energy set of the system; and Pi (t) and P2 (t) are the two constants (or integrals) of motion (Helleman, 1980). The constant energy E allows the phase space motion to be reduced from four dimensions R2 x T 2 to three dimensions. The constancy of Pi (t) and P2 (t) allows a further reduction to a one dimensional surface in the three-dimensional constant energy space. On the one-dimensional surface, the angular motion is parameterized by the two frequencies associated with each degree of freedom according to a\ = a\,
(6.185a)
a2 = a2 — a\.
(6.185b)
The angular components of the motion on the normally hyperbolic invariant manifold M may be determined by integration and are given by Gi(0 = a i f + G i ( 0 ) ,
(6.186a)
Q2(t) = (fl2 - ai)t + g 2 (0).
(6.186b)
Consequently, the normally hyperbolic invariant manifold Mhas the structure of a two-parameter Pi(0> Piif) family of two-dimensional tori. For a fixed Pj (t) = Pi (t) and P2(t) = P2(t), a corresponding two-dimensional torus on
467
Nonlinear Wave Theories
Pi
W (TCA.P,)) n W (T(J»1(Pa))
?(Pi,P2)
A - * \
^H*rT 0
"n/2
n
Fig. 6.18. Three-dimensional unperturbed heteroclinic manifold Ti (after Bowline et al., 1999).
M shown in Fig. 6.18 may be denoted as T(Pi(0,ft(0 = Uq(t),p(t),
Pxit), P2(t), Qi(t), 2 2 ( 0 ) € T1 x M1 x R2 x T2 : q(0), p{0),Pl(t) = Pl(t),P2(t)
= P2(t) (6.187)
where the hyperbolic saddle points are given by g(0) and p(0) in Eqs. (6.180c,a). Each two-dimensional torus Y(Pi(t), Pi{t)) C Mis invariant, i.e. any trajectory starting at a point on the surface of the torus remains on the surface. For a constant energy value E = E, say, a corresponding level energy surface may be represented in the three-dimensional constant energy space as a family of concentric tori as shown in Fig. 6.19. The P\ (t) and P2W variables measure the radii of the circular cross section and the minimum ring of the torus, respectively, shown in Fig. 6.19. The angular variables Q1 (t) and Qi (t) measure the two angles of a point on the surface of the torus as shown in Fig. 6.19. Selecting E = E and fixing the value of P\(t) also fixes the value of Pi(t). Because both frequencies of motion are independent of P\(t) or P2(t), they do not change from one concentric torus to another concentric torus. The motion on the surface of the invariant torus T (Pi (0, ^2(0) C Mis quasiperiodic or conditionally periodic
468
Waves and Wave Forces on Coastal and Ocean Structures
Fig. 6.19. Motion of a phase space point for an integrable Hamiltonian system with two degrees of freedom given by Eqs. (6.183): (a) Invariant tori in a three dimensional constant energy space E = E (after Rasband, 1990), (b) The flow on a two-dimensional torus on M for H{Pi (t), P 2 (0) = E (after Lichtenberg and Liberman, 1992).
(Lichtenberg and Lieberman, 1992, p. 22). If the frequency ratio
(72
«2 — a\
is a rational number, the motion on the surface of the two-dimensional invariant torus degenerates into a periodic trajectory of one-dimension that closes on itself and the torus is a resonant torus. In general, the frequency ratio is an irrational number and the motion on the surface of the twodimensional invariant torus may no longer be periodic; i.e. trajectories wind densely on the surface of the torus and never close on themselves; and the torus is a nonresonant torus (Arnold, 1978, Appendix 8). The two-dimensional nonresonant invariant torus T(P\{t), P2{t)) C M has a three-dimensional R 1 x T2 stable manifold Ws(T(Pi(t), P2_(t))) and unstable manifold Wu (T(Pi (t), P2(t))) that coincide along the T(Pi (/), P2(t)) threedimensional R 1 x T2 heteroclinic manifold H for fixed values of Pi (0 = Pi (t) and P2 (0 = P2 (t) as shown in Fig. 6.18. By invariance of manifolds (Wiggins, 1988, p. 387) ^ ( T ( P i ( f ) , P 2 ( 0 ) ) C Ws(M) and Wu(T(Pi(t),P2(t))) C U W (M). On a constant level energy surface given by Eq. (6.184), neither the nonresonant invariant tori Y(Pi(0, P2{t)) nor the stable and unstable manifolds are isolated. In addition, the two-dimensional nonresonant invariant torus T ( P i ( 0 , -^2(0) C Mhas a two-dimensional center manifold Wc(T(Pi(t), P2{t))) corresponding to non-exponentially expanding or contracting directions tangent to the normally hyperbolic invariant manifold dM (Wiggins, 1988, p. 383).
469
Nonlinear Wave Theories
Heteroclinic Orbits For fixed values of [Pi(t),P2(t)] = [Pi(t),P2(t)] € R 2 , the [q(t),p(t)] components of the unperturbed vector field a = y = 0 given by Eqs. (6.178) possess a one-dimensional heteroclinic separatrix that connects the hyperbolic saddle points [#(0), p(0)] in Eqs. (6.180c,a). The heteroclinic separatrices lie on each of the level energy surfaces defined by Eq. (6.184) for fixed values of Pl(t) = Pl(t) and P2(t) = P2{t) and are solutions to [H)(y = 0) - (fliPi(0 + (a2 - ai)P2(t)) = 0,
(6.189)
where the unperturbed Floquet Hamiltonian (H)(y = 0) is given by Eq. (6.175) and Eq. (6.179). Values for q(t) on the heteroclinic separatrices may be computed from Eq. (6.189) and are given by 1 -1 q(t) = -cos
a\ — eQ, K2eaiy/P\{t) -
(6.190) p(t),
In the full six-dimensional phase space in Eq. (6.178g), the heteroclinic manifold H in Eq. (6.182) may be determined by substituting a\ — e£i
cos(2q(t)) =
K2ea^Piit)
-
(6.191a) p(t);
a\ — sQ
sin(2 9 (0)
N
K2ea3y/Pi(t)
-
(6.191b) p(t)t
into the RHS of the unperturbed vector field a = y = 0 in Eqs. (6.178a-f)and by integrating to obtain the following variables on the heteroclinic manifold H: 2 Ph(t) = p2(t) + -sech \jAt\
qh(t) = 2 t a n _ 1
+ p(0),
tanh(VA0
(6.192a) + q(Q),
(6.192b)
a\ — sQ Pik(f) = Pi(0),
(6.192c)
Plhit) = P 2 (0),
(6.192d)
470
Waves and Wave Forces on Coastal and Ocean Structures
Qlh(jt) = ait - qh(t) + (21(0) +4(0) 1
_!
ai? — - t a n
a\ — £ ^
tanh(VAO "~ +Gi(0)
G2*(0 = (a2 ~ ai)t + 22(0),
(6.192e) (6.192f)
where A = [-(ai
- eQ)2 + (Pi(t) - P2(t))B] > 0,
B = 4s2al
(6.192g,h)
and where p(0) = 0,
q(0) = {In + 1 ) | ,
n = 0,l,2,....
(6.1921J)
The trajectories of the unperturbed system a = y = 0 along the fivedimensional Rl x R 2 x T 2 heteroclinic manifold H in Eq. (6.182) may be expressed as y(Pi(t),Pi(t)) = {qh{t),pn{t),Px{Q),P2(Qi),Qlh{t),Q2h{t)}.
(6.193)
The six-dimensional phase space in Eq. (6.178g) is a direct product between a region in four-dimensional space with coordinates [#(0, p(t), P\ (t), P2{t)] and the two-dimensional torus with angular coordinates Q\{t) and Q2(t). Because P\(t) and P2(t) are constants in Eq. (6.189), the motion in the sixdimensional phase space in Eq. (6.178g) is reduced to four dimensions on which four variables [q(t), p(t), Q\{t), Q2{t)] flow as shown in Fig. 6.20. Just as in the two-dimensional phase space (Guckenheimer and Holmes, 1983), small perturbations are expected to break up the geometric structure of the unperturbed system a = y = 0 and separate the manifolds. The behavior of the perturbed systems y > 0 and a > 0 near the unperturbed heteroclinic manifold H in Eq. (6.182) is required for the application of the GMM. The distance between the stable and unstable manifolds of any surviving invariant set in the perturbed system must be computed at an arbitrary point P on the unperturbed heteroclinic manifold H in Eq. (6.182). Two types of perturbed systems are evaluated. First, the perturbed system with perturbation (wavemaker forcing) parameter y > 0 and with no dissipation a = 0 is shown to be governed by a perturbed vector field that may be derived from the averaged autonomous components of the Hamiltonian given by Eqs. (6.176). Second,
471
Nonlinear Wave Theories
<«":(f''-N£f'0)
( W ,
Fig. 6.20. The unperturbed a = y = 0 reduced four-dimensional phase space in Eq. (6.178g) with Pi (f) and Pjit) equal to constants (Hudspeth, et al., 2005)
the perturbed system with forcing y > 0 and dissipation a > 0 is shown to be governed by a perturbed vector field that is dissipative. Geometric Structure of Perturbed Hamiltonian Phase Space (a = 0,y > 0) The perturbed vector field [q(0> p(01 for non-zero perturbation (wavemaker forcing) parameter y > 0 but with no dissipation a = 0 may be determined by setting a = 0 in the evolution equations (6.176) and is given by
q(t) =
^ o^ (2P{(t) + P2(t)-3p(t)\ —a\ + e\l + saj, I ^=^= I cos (2q(t)) \ y/Pl ~ P ) sin(2Qi(0) K iy -yci ~ U= - yc2 sin(2 (q(t) + Qx(t))) Sy/P\{t)p{t) (6.194a)
P(t) =
P\(t)
Usa^P^-pitXpit) P2(t))sm(2q(t))\ 1 -2yc2(P2(t) - p(t))cos(2(q(t) + fi!(0)) J ' - y W i ( 0 - P(0cos(2<2 1 (0)
=
-2yc2(P2(t)
hit) = o,
- p(t))cos(2(q(t)
(6.194b)
(6.194c)
+ Qi(t))) (6.194d)
472
Waves and Wave Forces on Coastal and Ocean Structures
Qx(t)=ax +sa3
(
^° ^ cos (2g(Q) + — — y c i V A (0 - p(t) s^Pi(t)-
sin (2Qi (Q), p(t) (6.194e)
Q2(t) = a2-eQ-
2ea3-y/Pi(t) - p(f)cos(2$(f))
+ yc 2 sin(207(O + 0i(O)),
(6.194f)
that is a System III (Wiggins, 1988, p. 336) in the six-dimensional space in Eq. (6.178g); and provided that the condition required for the Hamilton-Jacobi transformation is given by Pl(t) > Pit) > Pi(t) > 0.
(6.194g)
Because the perturbed vector field in Eqs. (6.194) is Hamiltonian (i.e. no dissipation with a = 0), the three-dimensional level energy surfaces in Eq. (6.184) are preserved. The four-dimensional normally hyperbolic invariant manifold /Win Eq. (6.181) of the unperturbed system a = y = 0, the locally stable W[oc(M) and unstable W"0C(M) manifolds and the flow on Mmay be applied to describe the geometric structure of the perturbed phase space given by the perturbed normally hyperbolic locally invariant manifold M, by the locally stable Wf0C(M) and unstable W"oc(M) manifolds, and by the persistence of the two-dimensional nonresonant invariant tori TY(P\ (t), P2(t)). Persistence of M The perturbed system with wavemaker forcing y > 0 but with no dissipation a = 0 possesses a four-dimensional normally hyperbolic locally invariant manifold My, i.e. trajectories may leave My by crossing its boundary because Pi(t) ^ ObyEq. (6.194c) (Wiggins, 1988, proposition 4.1.16). If all trajectories eventually leave MY by crossing its boundary, then there are no recurrent motions on MY. The objective of the GMM is to locate any recurrent motions in My. The perturbed normally hyperbolic locally invariant manifold My is given by My =
(q,p,Pl,P2,QuQ2)eTl xl1 xl2xT2: q = q0 (Pi, Pi, Qu Qi; Y) = qo (Pi, Pi) + 0{y), ) , p = po (Pu Pi, Qu Qi; Y) = Po (Pi) + 0(y)
(6.195)
where the hyperbolic saddle points are given by Eqs. (6.180a,c). Moreover, on My there are locally stable Wfoc(MY) and unstable W" (My) manifolds
473
Nonlinear Wave Theories
that are of equal dimensions and are close to the unperturbed locally stable W?oc(MY) and unstable W"oc(My) manifolds, respectively. Although trajectories in Wf0C(MY) and W"0C(My) approach My as t - • ±00, they need not terminate on MY because all trajectories on MY may leave My in a finite amount of time. The perturbed normally hyperbolic locally invariant manifold My intersects each of the five-dimensional level energy surfaces given by Eq. (6.175) in a three-dimensional set of which most of a two-parameter family of two-dimensional nonresonant invariant tori persist in Eqs. (6.183) by the KAM theorem (Arnold, 1978, Appendix 8). These two-dimensional tori are KAM tori that may be slightly deformed compared to those of the unperturbed system for a = y = 0. The Melnikov integral may now be computed to determine if the stable and unstable manifolds of the KAM tori intersect transversely (Wiggins, 1988). KAM theorem (Arnold, 1978, Appendix 8) The KAM theorem may be applied in order to determine if recurrent motions occur on the perturbed normally hyperbolic locally invariant manifold My; and, in particular, if any of the two parameter family of two-dimensional nonresonant invariant tori T(P\(t), P2O)) inEq. (6.187) survive the Hamiltonian perturbation with wavemaker forcing y > 0 but with no dissipation a = 0. The unperturbed Floquet Hamiltonian (H)(y — 0) given by Eq. (6.179) satisfies the following nondegeneracy (or, equivalently, nonresonance) condition (Wiggins, 1988, p. 386): d2(H)
d2(H)
dPx2(t)
dPx{t)dP2{t)
d2{H)
d2(H)
dP2(t)dPx(t)
2
dP2 (t)
(ai - eft) 2 ~
4(Pi(t)-P2(t))2<
'
(qm,p{0y,y=o)
(6.196) where the hyperbolic saddle points (qh (0), pn (0)) are given by Eqs. (6.180c,a). Consequently, most of the two-dimensional nonresonant invariant tori in Eq. (6.187) persist as KAM tori denoted by T(Pi(t), P2(t)). Accordingly, in the phase space of the perturbed system with wavemaker forcing y > 0 but with no dissipation a = 0, there are invariant tori that are densely filled with winding trajectories that are conditionally periodic with two independent frequencies a\ and a2 from Eqs. (6.185). The resulting conditionally-periodic
474
Waves and Wave Forces on Coastal and Ocean Structures
motions of the perturbed system with wavemaker forcing y > 0 but with no dissipation a = 0 at these two fixed frequencies are smooth functions of the perturbation (wavemaker forcing) parameter y (Arnold, 1978, Appendix 8). A generalization of the KAM theorem states that the KAM tori T(P\(t), P2{t)) C MY has both stable and unstable three-dimensional manifolds denoted by Ws(Ty(Pi(t), P2(t))) and by Wu(TY(Pl(t), P2(t))), respectively (Wiggins, 1988, p. 387). Consequently, WS(TY) c Ws(My) and Wu(Ty) c Wu(My) by invariance of manifolds (Wiggins, 1988, p. 387). In order to determine if chaos exists, it remains only to determine whether or not Ws(Ty(Pi (t), P 2 (0)) and Wu (Ty (Pi (t), P 2 (0)) intersect transversely. Because the perturbation with wavemaker forcing y > 0 but with no dissipation a — 0 is Hamiltonian in the six-dimensional phase space in Eq. (6.178g), trajectories are restricted to lie in five-dimensional level energy surfaces given by Eq. (6.175). Consequently, two measurements are required in order to determine whether or not W*(T K (Pi(0, P2(t))) and Wu(TY(Pi(t), P2(t))) intersect transversely (Wiggins, 1988, Lemma 4.1.18). Moreover, because P2(t) = 0 in Eq. (6.194d), only one measurement along a constant unit vector in the P\(t) direction denoted as Pi(t) is required in order to determine whether or not ^ ( 1 ^ ( ^ 1 ( 0 , ^ 2 ( 0 ) ) and Wu(TY(Py(t), P2(t))) intersect transversely. Melnikov Integral (a = 0, y > 0) The distance between W*(T K (Pi(0, P 2 (0)) and Wu(Ty(Pi(t), P 2 (0)) at any point P e 7i may be computed from (Wiggins, 1988, 4.1.85)
M(fi,(0)) = -/«„\ _ „ r
^*<*w.ft<'»A,
(6 197a)
where *!/(•••) is defined in Eq. (6.193) and where
Hy = ^j-y/Pi(t)-p(t)sm(2Qi(t))+C2(P2(t)-p(t))sm2(q(t)
+ Qi(t)),
(6.197b) where HY = yH in Eq. (6.175). The improper integral in Eq. (6.197a) may be evaluated if the limits an,fin ->• ±oo are approached along sequences of discrete time (Wiggins, 1988, Proposition 4.1.29 and Lemma 4.1.27, y
475
Nonlinear Wave Theories
pp. 410-412). The Melnikov integral M(0i(O)) in Eq. (6.197a) represents the leading order term in a Taylor series expansion (with respect to y) for the distance between Ws(TY(Pi(t), P2(t))) and Wu{TY(Pi(t), hit))) at the point IP e H along a constant unit vector in the Pi (t) direction denoted as A ( 0 - In order to evaluate Eq. (6.197a), only the perturbed vector field with wavemaker forcing y > 0 and with no dissipation a = 0 in Eqs. (6.194) and the trajectories in Eq. (6.193) along the unperturbed heteroclinic manifold H are required. Substituting Eq. (6.197b) and Eq. (6.193) into the RHS of Eq. (6.197a) yields M(Gi(0)) lim fin an ->• oo / Pn
Ac\ ~y/h{t)
-Ph(0cOS(2fiiA(0)
dt.
~2c2(P2(t) - Ph(t)) cos(2(qh(t) +
Qih(t))) (6.198a)
Substituting Eq. (6.192e) into Eq. (6.198a) and expanding by elementary trigonometric identities (vide., Table 2.1, Chapter 2.4.1) yields M(fii(0)) = cos2«2i(0)+tf(0))x lim pn |~_4ci y/Pl(t)-ph(t) (cos (2qh(t) - 2ait)) x dt otn -2c2 (h(0 ~ Pk(0) cos (2oi0 + sin2(Gi(0) + ^(0)) lim
rPn
(;;HL
Ac\
—y/P\(f) s
- ph(t) sin2 (qh(t) -
+2c2 (hit)
(6.198b)
a\ — e£2 \2ea3y/Pi(t)
sin(2qh(t)) =
dt
- ph(t)) sin (2ai0
Substituting (?) = /,(*) from Eq. (6.192b) and p(t) Eq. (6.192a), Eqs. (6.191) may be transformed to cos(2qh(t)) =
ait)
- ph(t)/
y/A + B(h(t) leaiy/P\(t)
\
-
ph(t)) ph(t)
= ph(t) from
(6.199a) '
(6.199b)
476
Waves and Wave Forces on Coastal and Ocean Structures
where A,B are given by Eqs. (6.192g,h). Substituting Eq. (6.192a) and Eqs. (6.199) into Eq. (6.198b) and retaining only the even integrands reduces the Melnikov integral in Eq. (6.198b) to simply M(Gi(0)) = cos[2(fii(0) +
(6.200a)
The integrals /, in Eq. (6.200a) are lim
f -2c\ (ai - sQ) [P" 1 h = \ -^r / cos(2axt)dt , n -* oo [ s2a3 J_an J 2 h = j -^— / cos(2ait)sech (*/At)dt I ,
h =
~
Q 2
j
sin(2ai0tanh('v/A?)^!.
(6.200b) (6.200c) (6.200d)
£ fl3
The improper integral I\ = 0 if the limits of integration are evaluated as a limit of sequence of times an = — Tn and /?„ = Tn where Tn = nn/a\ and where n = 1,2, (Wiggins, 1988, p. 447). The improper integral li is given by Gradshteyn and Ryzhik (1980, #3.982(1), p. 505). Finally, the improper integral 1$ may be integrated by parts and then computed, again, by Gradshteyn and Ryzhik (1980, #3.982(1), p. 505). Wiggins gives the conditions (1988, Theorems (4.1.19) and (4.1.20), p.393) that are required for the transverse intersection of Ws(TY(Pi(t), P2(t))) and Wu(TY(Pi(t), P2(t))) in the fivedimensional level energy surfaces. Because all of the Melnikov components in Eqs. (6.200) are bounded and do not sum to zero, the Melnikov integral M(£h(0)) = 0when Gl(0) = Gi,,(0) = (2/i+ l ) ^ - $ ( 0 ) ,
n = 0,1,2,...,
(6.201a)
and the derivatives dM
}^0))
= -2sin2(Gi„(0) + q(0))lh + h + h] ^ 0,n = 0,1,2,...
oGi(0)
(6.201b) are never zero so that the zeroes of the Melnikov integral may be computed from Eq. (6.201a). Consequently, the stable Ws(TY(Px(t), Piit))) and unstable Wu{TY{Px{t), P2(t))) manifolds of the KAM tori TY(Pi(t), P2(t))
477
Nonlinear Wave Theories
intersect transversely yielding Smale horseshoes (Wiggins, 1990) on the appropriate five-dimensional level energy surfaces in Eq. (6.175). This implies multiple transverse intersections and the existence of chaotic dynamics in the perturbed vector field with wavemaker forcing y > 0 and with no dissipation a = 0 that is governed by the perturbed vector field given in Eqs. (6.194). Geometric Structure of Perturbed Dissipative Phase Space (a > 0, y > 0) The perturbed dissipative vector field [q(t),p(t)] for non-zero wavemaker forcing y > 0 and for non-zero dissipation a > 0 may be obtained from Eqs. (6.176) by setting a = vy where v <£ 1 to obtain sa-i
q(t) =
2Pl(t) + P2(t)
=
V^ITO" p(t)
cos(2<7)
J
-ay + etl + ygiW (q(t), p(t), Pi(t), P2(t), Qi(t), v; y) (6.202a) Usa^Pi(t)-p(t)(p(t)
=
PK)
-
P2(t))sm(2q(t))
Pit)
I +yg (q(t),p(t),Pi(t),P2(t),Qi(t),v)
Pl(t) = P~2(t) =
Gi(0 = Qlit) =
J
ygPl{t)(q(t),p(t),Pi(t),P2(t),Qi(t),v), ygP2{t)(P2(t),v), {p{t) - P2(t)) : cos(2q(t)) a\ + ea-i,JPi(t)-p(t) +ygQlit\q(t),Pit), Pi(t), P2(t), Qi(t), v; y) a2-eQ+YgQi(t)
{q(tX
2eai-s/Pi(t) - p(t)cos (2q(t)) p2(tX g l ( r ) j v. y) p(t)> Pl(t)t
(6.202b) (6.202c) (6.202d) (6.202e)
(6.202f)
where the perturbed components g' (Wiggins, 1988, Sec. 4) are given by gq{t\q(t),p(t),
Pi(t), P2{t), (2i(0, v; y)
sin (2 G i(0) C2sin(2( 9 (0 + Gi(0)) -c\- e^Pi(t)-p(t) +2iP d32(p , 2 . (d2{Pl(t) ~- pit)) p(t))+d (t)2( - p(t))\ ...... +v^ydi - vs ( ====== ) sm(2q(t)) 7 y/pl(t)-p(t) ; (6.202g)
478
Waves and Wave Forces on Coastal and Ocean Structures
!p{t)(q(t),p(t),Pi(t),P2(t),Q1(t),v) f -2c2(P2(t)
- p(t))cos(2(q(t)
-p(t) + 2ved2JPlit) r\
+ Qx{t))) + vd4P2(t)
- p(t)(p(t) -
P2(t))cos(2q(t)) (6.202h)
gPl{,Hq(t),p(t),Pi(t),P2(t),Qi(t),v) 4ci -—VPM^pJt)
cos(2 Q i (0)
-2c2(P2(t) - p(t))cos(2(q(t) + (2i(0)) (4X2r - l)(P2(t) - p(t)) - Pi(t) +v 2Xhi -vedSy/Pi(t)-p(t)(p(t) - P2(t))cos(2q(t)) 'x -k^xks = v (\6tK^X) Plit),
p g
gQM
*\P2it),v)
,
2
A
^ , ( Pitt) - Pit) \ . n . .. +v£d2 smi2qit)) \s/P\it) - Pit) J g
(6.202J)
(q(t), p(t), P^t), P2it), Qxit), v; Y) sin(2Gi(0) eJPxit)-pit)
Q.,t), U2{t)
(6.202i)
,. , . D f . iqit),pit),P1it),v;y)=\
,
(6.202k)
\c2smi2iqit) + Qlit))) + v2yd7\ ^ }. [ +ved2y/Pi(t) - pit) sin (2(0) J (6.2021)
This perturbed dissipative vector field is a System I (Wiggins, 1988, p. 336) with iqit),pit), Piit), P2it), Qiit), (22(0) e T 1 x R 1 x R2 x T 2 and where Pi(t) > Pit) > /MO > 0. The four-dimensional normally hyperbolic invariant manifold M in Eq. (6.181) of the unperturbed system a = y = 0 and the locally stable W/0C(M) and unstable W"0CiM) manifolds may be evaluated in order to compute the geometric structure of the perturbed dissipative phase space given by the perturbed normally hyperbolic locally invariant
Nonlinear Wave Theories
479
manifold Mya, the locally stable Wfoc(MYce) and unstable W"0C(MYa) manifolds and the persistence of the two-dimensional nonresonant invariant tori rYa(Pi(t),P2(t)). Persistence of M The perturbed dissipative system a > 0 and y > 0 possesses a fourdimensional normally hyperbolic locally invariant manifold Mya where trajectories may leave Mya by crossing its boundary because P\ (?) ^ 0 and P2(t) £ 0 by Eqs. (6.202c,d) (Wiggins, 1988, Proposition 4.1.5, p. 354). If all trajectories eventually leave Mya by crossing its boundary, then there are no recurrent motions on Mya. The objective is to locate any recurrent motions in Mya in order to apply the GMM. The perturbed normally hyperbolic locally invariant manifold MYC( is given by Mya = (q(t),p(t),Pl(t),P2(t),Qi(t),Q2(t))eT1xm1xR2xT2: q(t) = q0 (Pl(t), PliO, Gl(?), Ql(t)\ y) = qo (Pl(t), P2(t)) + 0(y), | , p(t) = Po(Pi(t),P2(t),Qi(t),Q2(ty,y) = p0(P2(t)) + 0(y) (6.203) where the hyperbolic saddle points qo(P\(0), P2(0)) and po(P2(0)) a re given by Eqs. (6.180c,a). Moreover, Mya has locally stable Wioc(Mya) and unstable W"0C(MYa) manifolds that are close to the unperturbed locally stable Wfoc(MYa) and unstable W"oc(Mya) manifolds, respectively. If Ws(Mya) and Wu(Mya) intersect transversely, then the Smale-Birkhoff theorem (Smale, 1963) predicts the existence ofhorseshoes and their attendant chaotic dynamics in the perturbed dissipative system a > Oandy > 0 governed by the perturbed dissipative vector field in Eqs. (6.202). Averaging Method An averaging method (Simiu, 1996) may be applied to the perturbed dissipative vector field a > 0 and y > 0 in Eqs. (6.202) in order to determine if recurrent motions occur on the perturbed normally hyperbolic locally invariant manifold Mya. If there are any tori on IWya, then the Melnikov integral may be computed to determine if the stable and unstable manifolds of these tori intersect transversely. A two-dimensional hyperbolic invariant torus Tya{P\ (?), P2(t)) may be located on Mya by averaging the perturbed dissipative vector field a > 0 and y > 0 in Eqs. (6.202) restricted to Mya over the angle variables Q\(t)
480
Waves and Wave Forces on Coastal and Ocean Structures
and Qiit). The averaging method requires nonresonance conditions resulting in trajectories filling densely the surface of the torus Tyct(P\(t), /MO)- This averaging method is not appropriate for Hamiltonian systems (Wiggins, 1988, p. 359). In the case of Hamiltonian perturbation a = 0 and y > 0, twodimensional invariant tori TY(P\(t), Pi{t)) are located on the perturbed normally hyperbolic locally invariant manifold MY by the KAM theorem. The perturbed dissipative vector field a > Oandy > 0 in Eqs. (6.202) that is restricted to the perturbed hyperbolic locally invariant manifold Mya in Eq. (6.203) may be determined from Eqs. (6.202c-f) and Eq. (6.203) and is given by p
Pi(t) = y
8
HtU K) \p
q(0) = qo(Pl(0),P2(0)), Y + = po(P2(0)), Pi(0), F 2 (0), 0i(O), v)
0(y2), (6.204a)
Pi(t) = y(gP2(t)(P2(t),
v)) + 0(y2),
01 (0 = Ql{q(t)=qo(Pl(0),P2(0)),p(0)
(6.204b) = p0(P2(0));y = 0) +
= ax + 0{y),
(6.204c)
Qlit) = Q2(q(t) = qo (Pi(0), P2(0)),P(t) = a2-ax
0(y)
= Po(P2(0)); y = 0) + 0(y)
+ 0{y),
(6.204d)
where the hyperbolic saddle points qo{P\ (0), P2(0)) and po(P2(0)) are given by Eqs. (6.180c,a), gP2(t){P2(t), v) by Eq. (6.202J); and where gPi(t)(q(t)=q0,p(t)
=
= s
P0,Pl(t),P2(t),Ql(t),v)
-VPiiO - P 2 (Ocos(20i(O) - ^ 2 - ^ 1 ( 0 2Xjx\
(6-204e)
The averaged equations are v f"2it fI>2TI PlP
{P2)
= 7r^2 {2TCY JO
/ JO
8P2(P2,v)dQ1dQ2=
=
y
\6,Tr:
-vyL -j -Pu K Ti (6.205a)
2
"ft,
K ? Ti r *
(6.205b)
481
Nonlinear Wave Theories
that have a unique stable hyperbolic fixed point at (Pi (t), P2(t)) = (0,0) with two negative eigenvalues provided that the determinant d(Pi) dPi(t) d(Pl) dPi(t)
3(Pi) dP2(t) d{P2) dP2(t)
,2„2M4
vV(^grx-T)
>0
(6.206)
2A.^TI2TX
is positive (viz., k,% rx > T). This fixed point of the averaged equations (6.205) corresponds to a two-dimensional torus denoted as Y Ka (0,0) on the perturbed hyperbolic locally invariant manifold Mya in Eq. (6.203). Applying Wiggins Proposition 4.1.6 (1988, p. 358) to the full six-dimensional phase space defined by Eq. (6.178g), the perturbed dissipative vector field a > 0 and y > 0 in Eqs. (6.202) that is restricted to the hyperbolic locally invariant manifold Mya in Eq. (6.203) lies on a two-dimensional normally hyperbolic invariant torus Tya (0,0) c Mya that has a five-dimensional stable manifold Ws {Tya (0,0)) and a three-dimensional unstable manifold Wu(Tya (0,0)). By invariance of manifolds (Wiggins, 1988, p. 359), W%Tya(0,0)) C Ws(Mya) and r ( T , , a ( 0 , 0 ) ) c r ( J M / ( I ) . The normal hyperbolicity of T y «(0,0) insures that the dynamics normal to the invariant torus dominate the dynamics on the invariant torus under the action of the perturbed dissipative flow (Wiggins, 1988, p. 319). The dissipative perturbation a > 0 and y > 0 creates two new independent vectors in the tangent space of the stable manifold Ws(Tya(0,0)) that is due to the breakup of the two-dimensional center manifold Wc(T(Pi(t),P2(t))) of the unperturbed system ot = y = 0 (Wiggins, 1988, p. 369). In order for chaos to exist, the five-dimensional stable manifold s W (Tya(0,0)) and the three-dimensional unstable manifold Wu(Tya(0,0)) must intersect transversely. For any point fp e WS(M), the tangent space of WS(M) at (P denoted as Tj>Ws{M) is a five-dimensional linear vector space. Applying Wiggins Proposition 4.1.2 (1988, p. 342), a one-dimensional vector space in (T1 x x W x T 2 ) complementary to TpWs(M) is given by N$> = span
3{H)(y = 0) d{H)(Y = 0) ,0,0,0,0 dp(t) Hit)
(6.207)
482
Waves and Wave Forces on Coastal and Ocean Structures
where {H)(y — 0) is the unperturbed Floquet Hamiltonian in Eq. (6.179) and where the derivatives in Eq. (6.207) are evaluated at 0 (Wiggins, 1988, p. 364). The Wiggins lemma 4.1.8 (1988, p. 361) assures that the components of the distance between the stable manifold Ws(Tya(0,0)) and the unstable manifold Wu(Tya(0,0)) may be set equal to zero in the directions along constant unit vectors in the P\(t) and the P2O) directions that are denoted as P\{t) and P2(0> respectively. Therefore, only one measurement along the one-dimensional vector space Np defined by Eq. (6.207) will be required to determine if Ws(Tya(0,0)) and Wu(TYa(0,0)) intersect transversely. For Hamiltonian perturbation a = 0 and y > 0, the trajectories are restricted to lie in five-dimensional level energy surfaces given by Eq. (6.175). Because the one-dimensional vector space N
M«2l(0)) f Antigq-pgp =„
11
A !™
(Xfl , Pfl
+ QigPl + Q2gP2K*(m(t);y Pl
im
-Ql(ci0(PuP2),P0(P2Y,Y = 0)ftng W ('y'Y
= 0)dt = (»dt
,
T CX)
0
-Q2(qolPl,P2).Po(Pz);Y = 0 ) / ^ n ^ ( * ( ° . ) ( r ) ; y =0)dt
(6.208a)
483
Nonlinear Wave Theories
where the hyperbolic saddle points qo(P\ (0), /MO)) and po(^2(0)) are given by Eqs. (6.180c,a) and where * ( 0 ' 0 ) (O = {qh(t),ph(t), Qih(t), G2*(0}(P,(0 = 0,^(0 = 0)
(6.208b)
is a heteroclinic separatrix of the unperturbed system a = y = 0 in Eq. (6.193) where Pi (/) = Piit) = 0 that corresponds to the hyperbolic fixed point of the averaged vector field on the perturbed hyperbolic locally invariant manifold Mya in Eq. (6.203). Because Eq. (6.180d) is valid on the unperturbed normally hyperbolic invariant manifold Min Eq. (6.181), then Pi(t) = P2(t) = 0 is a point on M provided that a\ = eQ,, (6.209a) that is the lowest point on the unperturbed heteroclinic manifold shown in Fig. 6.18. Bowline et al. (1999, p.39) identified this fixed point as a point of weak chaos for cross waves without surface tension or dissipation. Substituting Eq. (6.168c) for eQ and a\ into Eq. (6.209a) and solving for /3 yields
'44
l-All-JL
-11/2
(6.209b)
where
A
= ?7R-
(6 209c)
-
Expanding /3 by the binomial expansion (Chapter 2.3.4) gives, approximately, P - \ (l - ~|) •
(6-209d)
The Melnikov integral M(gi(0)) in Eqs. (6.208) represents (to 0(y2)) the distance between Ws(TYa(0,0)) and W"(T ya (0,0)) at any point $> e H along the one-dimensional vector space Np defined by Eq. (6.207). In order to compute the integral in Eq. (6.208a), only the perturbed dissipative vector field a > 0 and y > 0 in Eqs. (6.202) and the trajectories along the unperturbed heteroclinic manifold H in Eq. (6.193) are required. Computing the Melnikov integral in Eq. (6.208a) on the Pi(t) = Pi(t) = 0 energy level and evaluating the improper integrals as limit of discrete time sequences leads to M(<2i(0)) being identically equal to zero. This implies that the Melnikov
484
Waves and Wave Forces on Coastal and Ocean Structures
method fails to provide the necessary condition for this dissipative system to be chaotic. Liapunov Characteristic Exponents In contrast to Hamiltonian (conservative or non-dissipative a — 0) systems in which the phase-space volume is conserved by the Liouville theorem (Verhulst, 1990), dissipative systems are characterized by continued contraction of the phase-space volume with time t. Dissipative systems are characterized by the attraction of all trajectories passing through a certain domain towards an invariant surface or an attractor of lower dimensionality than the original space. If the dissipative system parameter is changed, then the motion on the attractor may also change from regular (i.e. sink or limit cycle) to chaotic (i.e. strange attractor). Rates of divergence or convergence of trajectories may be computed by Liapunov characteristic exponents. One positive Liapunov characteristic exponent is a strong indicator of chaotic motions (Rasband, 1990). A strange attractor must have at least one negative, one zero, and one positive Liapunov characteristic exponents (Parker and Chua, 1989). For the six-dimensional phase space defined in Eq. (6.178g); there are six real exponents that may be ordered as AH > M2 > M3 > M4 > M5 > M6,
(6.210)
where fi\ is the largest Liapunov characteristic exponent and where one of the remaining five exponents represents the direction along the perturbed dissipative flow and must be equal to zero. For the six-dimensional phase space defined in Eq. (6.178g), there are ten distinct strange attractors with the following spectral signs of the Liapunov characteristic exponents: (+,+,+,+,0,-), (+,+,0,0,0,-), (+,0,0,0,0,-),
(+,+,+,0,0,-), (+,+,0,0,-,-), (+,0,0,0,-,-),
(+,+,+,0,-,-), (+,+,0,-,-,-), (+,0,0,-,-,-),
l
"
}
(+,0,-,-,-,-). Numerical Calculation of /ii For chaos to exist, only the largest Liapunov characteristic exponent fi\ is required in order to determine whether nearby trajectories diverge (fi\ > 0) or converge (fii < 0) on the average (Moon, 1992). The perturbed dissipative vector field a > 0 and y > 0 in Eqs. (6.202) is integrated numerically by the fourth-order Runge-Kutta method with a time step At = 0.02 sec to
485
Nonlinear Wave Theories
Trajectory
Fig. 6.21. Numerical calculation of the largest Liapunov characteristic exponent (after Benettin et al, 1976) where do — 1, x — f, y = i/r + 8f, At — 0.1 seconds.
determine a reference trajectory i/r(t) in the six-dimensional phase space in Eq. (6.178g). The largest Liapunov characteristic exponent /JL\ is calculated from the first variation of the perturbed dissipative vector field a > 0 and y > 0 in Eqs (6.202) according to
/ Sq(t) \ Sp(t) SPl(t) SP2(t) SQl(t) \SQ2it))
dq(t) dq(t) dp(t) dq{t) 3Pl(f) dq(t)
dq(t) 8p(t) dp(t) dp(t) dP\(t) dp(t)
dP2
dP2
dq(t)
dp(t)
9Gi(f) dq(t)
32l(0 dp(t) dQliO dP(t)
dQlit) dq(t)
dq(0 dPl(t) dp(t) dPlit) 3Pi(0 dPi(t) dP2(t)
dq(t) dP2(t) dp(t) dP2(t) dPl(t) dP2(t) dP2(t) dP2(t)
9P\(t) dQ\(t) dPi(t)
dQlit) dP2it)
322(0 dPi(t)
3fi 2 (0 dP2(t)
dqit) dQlit) dp{t) dQlit) dPlit) dQlit) dPlit) dQlit) dQlit) dQlit) dQlit) dQx{t)
dqit) dQlit) dp(t) dQlit) dPlit) dQlit) dP2it) dQlit) dQlit) dQlit) dQlit) dQ2(t)
( Sqit) \ Sp(t) SPiit) SPlit) SQiit) \SQiit)/
(6.212)
Figure 6.21 illustrates a reference trajectory x//(t) and a nearby trajectory VKO + W(t) with initial conditions fo(t = 0) and f0(t - 0) + 8\J/0(t = 0), respectively, that evolve with time yielding the tangent vector 8xjr {^Q, t)'m the six-dimensional phase space defined by Eq. (6.178g) with an Euclidean norm
d[0 = ll^(^o,OII,
(6.213)
where the symbol d(t) is not to be confused with the ordinary derivative symbol dt. A renormalization procedure due to Benettin et al, (1976) and applied by Umeki and Kambi (1989) for parametrically excited surface waves is adopted
Waves and Wave Forces on Coastal and Ocean Structures
e 0.8 " 0.6 " Chaotic Motion 0.4 -
u,>o
0.2 "
1 0.2
1 0.4
-i
1 O.i
0.6
Fig. 6.22. Chaos diagram for the regions in the v, e parameter space where chaotic motion may exists (Hudspeth, et al., 2005).
to avoid overflows and other computation errors that result from the exponential growth of the Euclidean norm d(t). For computational convenience, the initial norm is chosen to be unity and H-ifr is renormalized to a norm of unity every At = 0.1 seconds. The values for Euclidean norm d(t) in Eq. (6.213) are determined iteratively with a total number of time steps N = 500. The Liapunov restart time step is NAt = 50 sec. The largest Liapunov exponent is (Lichtenberg and Lieberman, 1992, p. 315) 1 N ui(if,Sxl/) = lim >hW;(0A/_>.™ MNAt At i—t N->oo
(6.214)
(=1
The following initial conditions: ^o = {0,0,0.1,0,0,0}, Wo = {0,1,0,0,0,0},
(6.215a) (6.215b)
and the following numerical values for the dimensionless parameters: y = 0.25, X,. = 0.5, | =24TT, / I = 1, 0 = 0.45, r =0.016, b = n are selected
487
Nonlinear Wave Theories
for the numerical computations. For twenty different values of the dimensionless damping parameter 0 < v < 1 and twenty different values of the dimensionless Floquet parametric forcing parameter 0 < e < 1, the largest Liapunov characteristic exponents are calculated and a chaos diagram is constructed in Fig. 6.22 for the positive values of the largest Liapunov exponents. This chaos diagram is useful for searching for those regions of the e — v parameter space where chaotic motions may exist. In the e — v parameter space in Fig. 6.22, chaotic motions may occur inside the shaded region shown and regular motions occur outside of the shaded region.
A. Nonautonomous Canonical Transformations by the Generalized Herglotz Algorithm (GHA) In a canonical transformation from the set (q, p) to the set (Q, P), the dynamical state of the system may be defined in terms of the new set of variables (Q, P). Because the canonical transformation destroys the distinction between generalized coordinates and conjugate momenta, the set of new variables (Q, P) may no longer be restricted to sets that define the configuration of the system as did the original set of generalized coordinates q (Desloge, 1982; Guenther and Schwerdtfeger, 1985; and Guenther et al., 1996). The extension of the Herglotz algorithm to nonautonomous dynamical systems (GHA) significantly reduces the effort required to suspend a nonautonomous Hamiltonian component (Orum et ah, 2000 and Hudspeth et al., 2005). Canonical Transformations: Definition and Properties A time-dependent transformation (q, p) =>• (Q, P) where Qi = Qi (q, P, 0,
Pi = Pi (q, P, 0 ,
i = l,2,.-.,N
(A. la,b)
is canonical if and only if there exists a generating function F(u,U, t) such that (Desloge, 1982, p. 794) N
dF(u, U, 0 = ^2(vt dui - Vt dUi), !=1
(A.2a)
488
Waves and Wave Forces on Coastal and Ocean Structures
where d2F du\dU\
32F duidllj
82F duNdUi
d2F du\dUN 7^0,
(A.2b)
32F
and where the combinations of 2N new and old variables are Ui,Vi
=
\
Ui,Vt =
\Pi,-9i,
Qi,Pi
i =
l,2,...,N.
Pi,-Qi,
(A.2c,d) (A.2e,f)
Equations (A.2) represent a generalized definition for a canonical transformation (Desloge, 1982). Poisson Brackets The Poisson bracket for any two members Rj, Rk are of the transformed set of 2N variables Q = Q\,..., QN and P = P\,...,P^ for the transformation inEq. (A.l) is N
r
dRjdRk
lRj,Rklw = ^ Y, |_ dqi dpi
dRjdRt dpi dqi
(A.3)
and the transformation by Eq. (A.l) is canonical if and only if the Poisson brackets satisfy IQi, Cjlqp = 0,
IQi, Pjlqp = SU,
IPi, Pjlqp = 0,
(A.4a-c)
where Sjj is the Kronecker delta (vide., Chapter 2.2.3). Generalized Herglotz Algorithm (GHA) The Herglotz algorithm for autonomous dynamical systems (Guenther et ah, 1996) may be generalized by: (1) including time t as a parameter, and (2) defining a generating function by Eq. (A.2a) with a non-zero determinant of second derivatives with respect to u, U defined in Eq. (A.2b). The GHA transforms a set of 2N variables (u, v) to a set of 2N new variables (U, V) by choosing
489
Nonlinear Wave Theories
N new variables £/, and then computing the remaining N new variables V, uniquely from the chosen [/,• so that the transformation (u, v) ->• (U, V) is canonical. The old («,-, u,-) and new (£/,-, V,) variables may be any of the four combinations in Eqs. (A.2c-f). The transformation (u, v) -> (U, V) may be shown to be canonical by the Poisson bracket conditions equations (A.4a-c). There are two types of the GHA. Type I equates N new variables to N old variables [/,• (u, v, t); and Type II equates N old variables to N new variables m — ui (U, V, t). In both types, however, the N new variables are to be chosen. GHA: Type I Equate N new variables to N old variables Ui =
Ui(u,\,t),
i = l,2,-
,N,
(A.5a)
,N
(A.5b)
such that the Poisson bracket in Eq. (A. 3) satisfies
m,UjJw
= 0,
i,j = 1,2,
and the determinant dUi dUi
dUi dvN #0.
dVj
dUN
(A.5c)
dUN dvN
Equate N Herglotz auxiliary functions X, (u, v) to the absolute value of the ratios of the Af old variables in either of the following ratio forms: Xi=Xi(u,v)
=
i =
l,2,-..,N,
(A.5d) (A.5e)
such that the Jacobian of the chosen N new variables U and the Af Herglotz auxiliary functions X are nonzero; i.e. 9(U,X) #0. 3(u,v)
(A.5f)
490
Waves and Wave Forces on Coastal and Ocean Structures
Solve Eq. (A.5d or e) for vi = u;(u,X), substitute f,(u,X) into Eq. (A.5a) and invert Eq. (A.5a) to obtain Ui=Ui(V,X,t),
i = l,2,-..,N.
(A.5g)
Compute the generating function F(u, U, t) in Eq. (A.2a) from N
du
dF(\\{\J, X, t), U, t) = J^^i
V dlJ
i -
i
( A - 5h >
i)
i= l
by expanding to obtain N
E
/dF
N
dXi + (
N
dF
dUt
\
3 V
^ 3 V
dUi
j=\
'
dx
\
= E E J^Yt i + E J w i=i y = i
'
-
VidUi
)-
^ A - 5i )
/
Equate the coefficients of like differentials in Eq. (A.5i) and compute: (1) the generating function F(u, U, t) from the integral N
pXi
N
a
V
F(U, u, o = E / E J ^rdx'i + C(U> 0> and 1=1
7=1
(A 5j)
-
'
(2) the remaining N new variables Vi (u, U, t) from
E
duidF J-^T--Z7T>
V
i = h2,...,N,
(A.5k)
y=i
where the integration constant C(\J,t) in Eq. (A.5j) is an arbitrary additive function of both time and U that may be neglected. The 2N old («,, i;,-) and new ([/,, Vi) variables may be any of the four combinations in Eqs. (A.2c-f). The transformation given by Eq. (A.5a) and Eq. (A.5k) may be shown to be canonical by the Poisson bracket equations (A.4a-c). In order to compute the transformed Hamiltonian in terms of the new variables (U,V), the inverse canonical transformation (u(\J,V,t), v(U,V,t)) must be computed and the new Hamiltonian K(Q, P, t) is given by K(Q, P, t) = tf(q(Q, P, t), p(Q, P, t), t) +
dF(u,\J,t) \' ', at
(A.51)
491
Nonlinear Wave Theories
where F(u,V,t) is the generating function in Eq. (A.5j) for the canonical transformation (U(u, v, t), V(u, v, t)). GHA: Type II Equate N old variables to N new variables ut = m(lJ,V,t),
i = 1,2,..., AT,
(A.6a)
such that the Poisson bracket Eq. (A.3) satisfies EM/.M/IUV
= 0,
i,j = 1,2, ...,N,
(A.6b)
and the determinant dVi
dvN
dui
(A.6c)
^0.
dVi
du N dVN
8UN
dVi
Equate A^ Herglotz auxiliary functions X, (U, V) to the absolute values of the ratios of the new variables in either of the following forms: Xi = Xi(U,\)
=
\Ui/Vi\ Wi/Ui\,
i =
l,2,..-,N,
(A.6d) (A.6e)
such that the Jacobian of the N old variables u and the Af auxiliary functions X is nonzero; i.e. 3(u, X) #0. (A.6f) 9(U,V) Solve Eq. (A.6d or e) for Vt = V/(U, V), substitute V,(U,X) into Eq. (A.6a) and invert Eq. (A.6a) to obtain Ui = Ui(u,X,t),
i=
l,2,...,N.
(A.6g)
Compute the generating function F(u, U, t) in Eq. (A.2a) from N
JF(u,U(u, X, 0,0 = J2iVi i=i
dUi
~ Vi
dui)
(A.6h)
492
Waves and Wave Forces on Coastal and Ocean
Structures
by expanding to obtain N
x^{
dF
E
i^rdXi + ir-du>
*—' \ aX; (=1
dF dui
V
dXi +
N
E ( E J jY
E VJ °irdu' - u«du' I •
i=l \ y = i
7=1
'
(A-6i>
Equate the coefficients of like differentials in Eq. (A.6i) and compute: (1) the generating function F(u, U, t) from the integral N
CXi
F(U,u, o = E / ;' = 1
N
dUV
E J^hdx'i+C(u'tX 7= 1
and
(A 6j)
-
'
(2) the remaining N old variables v,- (U, V, t) from vi(V,V,t)
J^ dUj = YiVj—±-—,
dF
i = l,2,...,N,
(A.6k)
7=1
where the integration constant C(u,f) in Eq. (A.6j) is an arbitrary additive function of both time and u that may be neglected. The 2N old («,, u,-) and new (£/,-, V,) variables may be any of the four combinations in Eqs. (A.2c-f). The transformation given by Eq. (A.6a) and Eq. (A.6k) may be shown to be canonical by the Poisson bracket equations (A.4a-c). The transformed Hamiltonian A^(Q, P, t) in terms of the new variables is given by K(Q, P, t) = H(q(Q, P, 0 , P(Q, P, 0, 0
-ir^'
at
where F(u,U,f) is the generating function for the canonical transformation (u(U,V,0,v(U,V,0).
6.9. Problems 6.1.
Applying the Dean stream function coefficients from the Dean Stream Function Tables (1974), compute the dimensionless water surface elevation r)' and horizontal water particle velocity u' for the following
493
Nonlinear Wave Theories
conditions and compare with the tabulated values from the Dean Stream Function Tables (1974): Dean * Tables (1974) Case 10D
s—z+ h h 0.9 0.9 0.9
0 (deg) 130 140 130
11
H
- 0.294
, u =
u H/T
- 0.625
in which
v'm = |r - ^ !>(«)sinh I (^r) ^ +ti}cosnd> u'{6) = - 22 X(n) I
1 cosh j
cos nd.
Case 10D: 3rd Order Stream Function Wave Theory Definitions: XQ = (g/2n)T2 is the deep-water wavelength calculated from linear wave theory, H is the wave height, g is the gravitational constant, T is the wave period, X(n) is the nth stream function coefficient, X is the nonlinear wavelength, *,, is the value of stream function on the free surface r/(9). Wave parameters: H/X0 = 0.170401, h/X0 = 1.999993, H/h 0.085201, A./A.0 = 1-222070, V^/igHT) = -0.015407,
6.2.
=
Dimensionless stream function coefficients: X(l)/(HTg) = -0.488993 - 0 5 , X(2)/(HTg) = -0.863825 - 11, X(3)/(HTg) = - 0 . 6 3 8 3 0 4 - 16. Verify that Eqs. (6.82 and 6.83) satisfy exactly the kinematic free surface boundary condition given by Eq. (6.81c).
Chapter 7
Deterministic Dynamics of Small Solid Bodies
7.1. Introduction Engineering analyses generally begin with an observation of the principles of physics that govern a physical phenomena. After these principles of physics have been carefully identified, a mathematical model may be developed that approximates these principles of physics. The careful selection of a mathematical model is always done after the physics have been carefully evaluated. It is crucial that a coastal or ocean engineer never juxtapose the sequence of these two elements in the engineering design and analysis. In this spirit, an observer standing on the boat landing near one of the corner jacket legs of a pile-founded steel jacket structure would observe visually that this leg has no sensible effect on the incident wave field. Under careful observation, especially in the case of long-crested waves, the observer could see or not see the following two major physical phenomena: Seen:
The observer would see flow separation around the slender jacket leg that results in vortex shedding and wake turbulence on the down-wave side of the jacket leg. In addition, some slight decrease in the water surface elevation may be seen in this down-wave, turbulent-wake zone. Not Seen: The observer would not see the incident wave modified significantly by the slender jacket leg. Additionally, if this platform were free to respond with dynamic motions such as in a small, semi-submersible buoy, no radiated waves would be observed radiating away from the slender member or buoy. 495
496
Waves and Wave Forces on Coastal and Ocean Structures
In contrast to what is seen or not seen in the case of a large body (vide., Chapter 8), the effects of flow separation and wake turbulence require that real fluid effects must be considered in the selection of a mathematical model to approximate the wave-induced pressure forces on a small body. Physically, viscous and turbulent stresses dissipate energy. As is the case with every mathematical model for dissipative stress tensors, empirical coefficients are required. This would be the case if it would be possible to model analytically the flow field around the small body. Just as Newton's law of viscosity requires an analytical model of the flow field (the gradient of the flow field dll/dy) plus an empirically determined coefficient (fi = dynamic viscosity), the mathematical model for the wave-induced pressure forces on a small body represented by the Morison equation requires an analytical model for the flow field (wave field kinematics) and empirically determined coefficients (Cd and Cm). Figure 7.1 illustrates the physics of a fixed and oscillating small member structure. The question posed in Fig. 7.1 is whether or not the inertial Cm and
S M A L L B O O V n o l n t i v s M o t i o n M o d l s n En S«mi)ubm«rflJble Buoy
PHYSICS
Mow
Not
Scan
FLUID
-f-
W»h»«
a l nnmt
Flttttt
xv_
s
ForcaJ
X £
i
I
-zx*!W
y»».
-" . ^—
!
Forced Oscfitatlorv
)
I
Cmtti'tot
I
-v-^^
W I V A • • R • r f t » + i r > n Jkvwsv
Vlt«BU> Flow
SOtUTION Ffnnd B o d y In W a v « * (Exclltrif) P o t o t )
$•(>•*attain &
•StimiimttHt
1 C
A
^
x.x.x
? ? ?
"vl 7 7 7 '
'
1
r
•
V i s c o u s
c
)
Dynamic Rasponsa Jn Portion
MtKli M o d a l
FT = Cmu, +CdLu-
(Cm - 1) X- CdL X
Fig. 7.1. Small body dynamic response.
497
Deterministic Dynamics of Small Solid Bodies
kA
2b <—»
kA>l;kA/(kh)3>l
2b V7.
Deep-Water
2b
2b 7J^
Z7
Finite-Depth
Fig. 7.2. Comparisons of small and large bodies in deep- and finite-depth water.
linearized drag Cdi coefficients for the exciting force on a fixed body are the same as the added mass Ca = Cm — 1 and viscous damping Cm coefficients for the same body oscillating in otherwise still water. In contrast to Fig. 8.1, the small body in Fig. 7.1 that is oscillating in otherwise still water does not generate radiated waves; but, rather only local vortices due to flow separation around the body. Figure 7.2 illustrates these two extreme possibilities as well as the possibility of a complex structure that may simultaneously include both sizes of members and, consequently, both effects. Because there is at present no single unified wave force field theory for computing the entire spectrum of wave force possibilities between these two extremes, only the effects based on the two extreme member sizes will be reviewed. The first effect to be reviewed is the small body or Froude-Kriloff theory where the size of the structure makes no sensible modification to the incident wave field (Chapter 7). The second effect to be reviewed is the large body or diffraction theory that also includes the possibility of wave radiation for compliant or for semi-constrained floating bodies (Chapter 8). Figure 7.3 illustrates the three methods that are applied to estimate each of these two extreme loads. Morison, et al. (1950) introduced a semi-intuitive equation to compute wave forces on fixed immersed objects having characteristic dimensions that are small compared to the wavelength of the incident wave. This equation has since been extended to the three-dimensional analyses of arbitrarily-oriented small members on offshore structures and to the relative-motion of small members. Sarpkaya and Isaacson (Chapter 3, 1981) give a detailed analysis and history of the Morison equation. Mei (1989) gives a concise but
Waves and Wave Forces on Coastal and Ocean Structures
498
METHODS FOR WAVE LOAD PREDICTIONS
ANALYTICAL
FROUDE-XRILOFF HYPOTHESIS
DIFFRACTION THBORI
DIMENSIONAL AKALIBES
NAVIER-SrOXES EQUATION
|
BERNOULLI EQUATION
| B . I . E . M . | | FBOUDE-KBILOFF | | M F F R A C H o i r ~ | MORISON H ) .
STRIP THEORT
VSl
SPECIAL CBOHETRZ
m
U R E A S THEORY KIHHATIOS
NONLINEAR THEORY KINEMATICS
URSELL HULTHOLE | CREEK'S FUNCTION
NUMERICAL STREAM FUNCTION
VARIATIONAL METHOD STOKSS PERTURBATIONS
CONFORMAL TRANSFORMATIONS
Fig. 7.3. Three methods for computing loads on coastal and ocean structures.
cogent argument applying classical viscous flow theory to demonstrate why the Keulegan-Carpenter parameter (or period parameter) K = UmT/D is important in separated flow. The Morison equation for the horizontal force per unit length on a fixed small body is approximated by dFx(z,t) = dFm(z,t)
+ dFd(z,t),
(7.1)
where dFx{z,t) = the wave-induced, hydrodynamic, horizontal pressure force per unit length acting normal to the axis of the small body; dFm(z, t) = the inertia component; and dFd(z, t) = drag component.
7.2. Small Body Hypothesis (Morison Equation) In order to appreciate the small body hypothesis in the Morison equation, it is helpful to compare the hydrodynamic pressure force on a totally porous pile that is composed completely of the same fluid. Newton's second law of motion may be applied to determine the time rate of change of momentum on this fluid pile in a well-known integral form that is derived in elementary texts on the fundamental mechanics of fluid for unsteady flows (e.g., Schlichting, 1979). A finite control volume may be applied to demonstrate that Newton's second law for the time rate of change of momentum M(t) in an accelerating
499
Deterministic Dynamics of Small Solid Bodies
fluid is dF(t) =
dM(t) dt '
(7.2a)
where dF(t) = sum of external surface and body forces; viz., dF(t) = dFN(t) + dFs(t) + dFB(t),
(7.2b)
where dF^it) = normal surface force; dFs(t) = tangential surface shear stress force; and dFs (f) = body force. The total force in Eq. (7.2b) is equal to the time rate of change of momentum within the control volume plus the transport of momentum or flux of momentum across the control volume boundaries (i.e., the inertia of the Eulerian fluid field). The finite control volume shown in Fig. 7.4 with an outward pointing surface normal vector dS = ndS contains a mass of fluid in the two finite control volumes Vol. 1 + Vol. 2 at time = t. The same mass at time = t + At is in Vol. 2 + Vol. 3. The fluid momentum is given by M(t)
•E
pu(x,t)dV(x),
v r >
dFx \
dz
ItTfl
$
= &
Fig. 7.4. Finite control volume for a "water pile".
(7.3)
500
Waves and Wave Forces on Coastal and Ocean Structures
and for a finite control volume, the time rate of change momentum from Fig. 7.4 is dM(t) dt
=
JM2 (0 + M3 (t) [ {+^
JMj (t) + M2 (t)}
At
At
hm At^o
(0
Adding and subtracting the quantity {M\(t)}(t+At)/At terms in Eq. (7.4) gives {Afl(0 + M2(t)}{t+At)
(7.4)
and rearranging
- {Mi(0 +M 2 (01(f) At
dM(t)
=
dt
hm At^o
+
{M3(t)}{t+At)
~ {A*i(0}(f+A0
At
-if
(7.5) The first term [•]£ on the RHS in Eq. (7.5) yields the local time rate of change of momentum in the fixed control volume given by lim
{Mi(t) +M2(t)}{t+At)
- {Mi(0 + M 2 ( 0 } ( 0
dt
At
At^O
dM(t)
Substituting for Af (?) from Eq. (7.3) yields dM(t)
*7r = 7,llh^-<)dV&-
(76a)
The second term on RHS in Eq. (7.5) [ • ] / yields the transport of momentum (or flux of momentum) out of the fixed control volume given by lim Af-+0
{A/ 3 (0}(t+A0 At
{Afi(0}(r+A0 At
= (p pu(x,t)[u(x,t)
•n]dS(x).
-1/ (7.6b)
Substituting Eqs. (7.6a and 7.6b) into Eq. (7.2) gives dF(t)
dM(t) dt = —
pu(x,t)dV(x)+
j> pu(x,t)[ii(x,t)»n]dS(x)
(7.7)
501
Deterministic Dynamics of Small Solid Bodies
that relates the time rate of change of momentum of a fluid particle in the Lagrangian coordinate axes to the fluid inertia in the Eulerian coordinate axes. Now, if a slender vertical circular pile were placed within the fluid control volume Vol. 1 + Vol. 2 in Fig. 7.4 and if the velocity distribution could be determined both within the control volume and across the surface of the finite control volume, then the total effect of the sum of all the external surface forces (normal and tangential) and body forces could be computed from the inertia of the fluid by Eq. (7.7). Because a closed-form analytical solution is not known for the exact velocity distribution around a slender cylinder in a real viscous fluid because of flow separation and wake turbulence in unsteady fluid flows, it is necessary to apply approximate methods to estimate the integrals in Eq. (7.7). If the slender cylinder radius is small compared to the incident wavelength (b/k <£. 1), one approximate method for estimating these integrals would be to expand the integrands in a Taylor series about the vertical centerline of the slender cylinder. Retaining only the linear terms in this series expansion would be equivalent to estimating the wave particle kinematics at the centerline (i.e., xo, yo, Zi) of the slender cylinder as though the cylinder were not present. For the horizontal force component in the positive x direction, Eq. (7.7) reduces to
<1FX(XQ + Ax, jo + Ay,
n,t)
9 + —idFx(xo, ox
«s dFx(x0,yo,Zi,t)
+ — [dFx(x0,yo,Zi,t)]Ay dy ^ ~dt / / /
y0,Zi,t)]Ax
+ 0(Ax)z
Pu(xo>yo,Zi,t)dV(x)
+ (f> pu(x0,yo,Zi,t)u(x0,yo,Zi,t) Jcs du(x0,yo,Zi,t) P / / / dV& dV(x) Jt
•hdS(x)+
0(Ax)
•Iffcv >
+ pu(xo,y0,Zi,t)
(t u(x0,yo,Zi, t) • ndS(x) + 0(Ax) Ics
502
Waves and Wave Forces on Coastal and Ocean Structures
Ifu(xo,yo,Zi,t)
• « ~ \u(xo,yo,Zi,t)\cos9,
/// dV(x) = 7Tb2dzr, JJJcv
and if
(b dS(x) = n2bdzi Jcs
for the z'th circular disk of height dzi in Fig. 7.4; then, dFx(x0,y0,Zi,t)
«s (prtb2)
' ' " at
dzi +
p(7T2b)u(x0,yo,Zi,t)\u(x0,yo,Zi,t)\dzi.
(7.8) In order to obtain an equality in Eq. (7.8), two empirically determined coefficients Cm and Cd/4jt reduce Eq. (7.8) to dFx(xo,yo,Zi,t) dzi 2
du(xQ,yo,Zi,t)
= Cm{pTtb )
1 h -Q(p^)«(xo,jo,Z(,OI"Uo,) ; o,Z(,OI(7.9a)
Examining carefully the sum of the external forces in Vol. 1 in Fig. 7.4 illustrates that the momentum transport theorem is equal to both the skin (friction) and form (wake) drag forces. The total resultant force in Eq. (7.2b) includes: (1) a normal total pressure force p(b,6,zi)dS and (2) a tangential shear force x{b,6,Zi)dS. The total force per unit length in the direction of wave travel in the x coordinate direction shown in Fig. 7.4 is given by (Dean andHarleman, 1966) dFx(xo,yo,Zj,t) dzi = j -/
P(b,0,zi,t)bcos6d6-
= {-/
lPo(zi) + 8p(b,0,Zi,t)]bcos6d6-
U
2it
/
r(b,6,Zi,t)b
sin0d9\
r(b,e,Zi,t)bsm9de\
p2n
Sp(b,9,Zi,t)cos6d6-
I
I
T(b,9,zi,t)sin9d9\.
(7.9b)
503
Deterministic Dynamics of Small Solid Bodies
Equating Eq. (7.9a) to Eq. (7.9b) demonstrates that the Morison equation (7.2) is equal to both the dynamic pressure force 8p(b,9,Zi)dS and tangential shear force r(b, 6, Zi)dS but not to the hydrostatic (body force) pressure force po(Zi)dS. Designs of slender body structures by the Morison equation must account for the collapse of the slender body due to the hydrostatic pressure force separately from the hydrodynamic pressure forces in Eq. (7.9b) computed by the Morison equation. Comparing Eq. (7.9a) with Eq. (7.9b), it may be observed that the hydrodynamic pressure forces on a slender circular cylinder computed by the Morison equation may be obtained from Eq. (7.7) by replacing the first volume integral on RHS with an inertia force dFm(z,i,t); and by replacing the second surface integral on RHS with a drag force dFd(zi,t). Assuming that the fluid is incompressible, Dp/Dt = 0 and that the small body hypothesis is valid, Eq. (7.7) may be written as dFx{z,t) = dFm(zi,t)+dFd(Zi,t),
(7.10a)
where for a right circular cylinder of diameter D = 2b; AV < r ( 7tL>2\ du(xo,yo,Zj,t) j t\ dFm(x0,yo,Zi,t) = Cm I p—— I — dzt,
rn,nU\ (7.10b)
dFd(xo,yo,Zi,t)
(7.10c)
= -Cd(pD)u(xo,yo,Zi,t)\u(xo,yo,Zi,t)\dzi.
Identification from load data and selection of the force coefficients Cm and Cd are deferred until Sec. 7.6. The Morison equation (7.2) for a fixed slender body may be extended heuristically to a dynamically responding slender body by the modified Wave Force Equation (WFE, relative motion generalization of the Morison equation, American Petroleum Institute API RP2A, 1987) that is given by AT? i *\ ,du(zi,t) dFm(zi,t) = pVi———
fdu(zi,t) + (Ca)pVi I ——
dFd{zi,t) = -CdpAi[u(zi,t)
- Xi(t)]\u(Zi,t)
» \ Xiit) I , - Xi(t)\,
(7.11a) (7.11b)
where Xt(t) = horizontal Lagrangian motion of the ith element of a slender body and where the over dots denote ordinary temperal derivatives of Lagrangian solid body motions (vide., Eqs. (3.48) in Chapter 3.5). Applications of a linearized form of the WFE in Eq. (7.11) are reviewed in Sec. 7.8
504
Waves and Wave Forces on Coastal and Ocean Structures
for single-degree-of-freedom (SDOF) articulated tower and for an uncoupled three degrees-of-freedom (MDOF) semi-immersed moored pontoon made of rigidly-connected double horizontal circular cylinders. The Morison equation (7.10) may be integrated for a total depth-integrated force on a vertical circular cylinder by applying linear wave theory kinematics (vide., Sec. 4.4). The horizontal water particle velocity and local acceleration at the vertical centerline of a circular cylinder at x = 0 are given by u(z,t) =
H gTcoshk(z 2 A.
+ h)
cosh kh
2nt cos
Hgn cosh k(z + h) (2nt cosh kh sin
du(z,t) dt
(7.12a) (7.12b)
V T
Substituting Eqs. (7.12) into Eq. (7.10a); assuming that the force coefficients Cm and Cd are constants; and integrating over the still-water-depth —h
ynD HKmsm[
4
r
Cd
\ T •}
/2nt
+ ^-yDH2Kdcosl Km =
—
2nt cos ——
VT
tanhkh
(7.13a) (7.13b)
2kh d = a 1+ sinh 2kh
(7.13c)
K
An example of the variations of Fx(t), Fm{t) and Fd(t) over one wave cycle for the following conditions: H = 34 ft, D = 4 ft, is illustrated in Fig. 7.5.
T = 12 s,
A = 8 5 ft Cd = 0J
Deterministic Dynamics of Small Solid Bodies
505
_3 000 o t T nTTlTrTTI111TtrtTrTTr1llTITtTTTTTI -180 -120 -60 0 60 120 180 S = 2m/T Fig. 7.5. Morison equation total and component depth-integrated forces on a vertical circular cylinder.
7.3. Drag dFd and Inertia dFm Forces An application that requires some knowledge of wave kinematics and dynamics is the response of small bodies to wave-induced velocities (kinematics) and pressures (dynamics). The adjective small restricts this analysis to body dimensions that are a small fraction of both the wavelength (b/X.
506
Waves and Wave Forces on Coastal and Ocean Structures
U(t)
WAKE
kWWWM
Fig. 7.6. Three types of elliptical cylinders.
ideal fluids may be applied to estimate these forces in unsteady flows. These potential flow estimates are also modified by real fluid effects. Drag Forces dFd: Two general types of drag forces are encountered; viz. (1) skin (friction) drag and (2) form (wake) drag. Figure 7.6 illustrates this general case for an elliptical cylinder of various ratios of the semi-minor to the semi-major axes. For an elliptical cylinder with a zero semi-minor axis (i.e., horizontal flat plate), only skin (friction) drag occurs. For equal semiminor and semi-major axes (i.e., a horizontal circular cylinder), both skin and form drag occurs. For an elliptical cylinder with a zero semi-major axis (i.e., a vertical flat plate), only form (wake) drag occurs. Form drag is associated with the stagnation pressure caused by a body blocking local velocities. Consider the verticalflatplate oriented perpendicular to the horizontal flow in Fig. 7.7. The pressure rise from point 1 to point 2 in Fig. 7.7 may be estimated from the unsteady form of the Bernoulli equation (Eq. (3.46) in Chapter 3.4) for horizontal flow by 30 at
p p
U2 2
•
Q(t)
d4> "dt
uz + - + — + gz P
J2 P 2 that for steady flow (d(»)/dt = 0) at the same vertical elevation (zi = n) reduces to P2 ~ Pi
ul
Ap
P where Ap = the local rise in pressure at the body due to the flow blockage effect and U\ = steady free-stream velocity. The force may be estimated
Deterministic Dynamics of Small Solid Bodies
507
•^ i
U(t)
3T
Fig. 7.7. Form (wake) drag past a vertical flat plate.
by integrating the pressure over the frontal area only of the body normal to the flow
r uf Fx « / pdA a pA-£that may be made into an equality by defining Cd to be the required constant of proportionality; i.e., Fd = \cd{pA)U\,
(7.14)
where Cd = dimensionless drag coefficient that is a function of the body shape and turbulence state that may be determined by the Reynolds parameter R from Chapter 3.6; p = fluid mass density; U\ = horizontal steady velocity; A = frontal area perpendicular to U\. Although the drag coefficient Cd varies with the Reynolds parameter R, fully turbulent values are approximately 1.0 for a cylinder and 2.0 for a sphere. The Reynolds parameter dependence for a circular cylinder in steady flow (cable, pile, etc.) is shown in Figs. 7.8 and 7.9. Equation (7.14) must be modified for oscillatory flows by Fd(z,t) = -Cd(pA)u(zi,t)\u(z,t)\
(7.15)
to insure that Fd(z, t) acts in the same direction as u(z, t) for oscillatory flows.
508
Waves and Wave Forces on Coastal and Ocean Structures
III III
1 1 1 1 III
D[mmJ •> 0.05, • 0.1 • RJ • 1.0 H83SWC6 " 3 0 if • 7J Wieselsberger » *ZJ> e 60.0 • 300.0 ---Thcotydueto lamb
,
_--
s \
L-
Fig. 7.8. Drag coefficient for circular cylinders in steady flow (Schlichting, 1979).
I
\ Co
s %
* f!S^ \
®
•
Uehster
® ' • "
v
"^
"V
MAX NJJ a.
*" *=»,
\ \ »<:
«r»' * e V ?
4 6
-e
V2 ' V
?
*£V* *SV?
4S
V2
*eV
e_B3 *~ v
Fig. 7.9. Steady-state drag coefficient, Cds, for a smooth cylinder in the transition and transcritical regimes (Schlichting, 1979).
7.3.1. Inertia Forces Inertia forces are the body reaction to a fluid accelerating in the vicinity of the body. The volume of fluid accelerated is proportional to the volume of the body. The constant of proportionality is Cm where Cm = an inertia mass
509
Deterministic Dynamics of Small Solid Bodies
coefficient that integrates the effects of all local fluid accelerations in terms of (1) the displaced fluid volume V and (2) the approach acceleration 3 U/dt; i.e., Fm(z,t) = CmPVd-^^-. at
(7.16)
The inertia coefficient Cm may be computed from potential theory for flow around bodies having simple geometric shapes (e.g., cylinders, spheres, bodies of revolution) by solving for a scalar velocity potential O from Laplace's equation constrained by the boundary conditions; solving for the pressure p from <E> by the Bernoulli equation; integrating p around the body surface to obtain Fm; then dividing by pVdU/dt to obtain Cm. For example, Cm = 1.5 for a sphere and Cm = 2.0 for a circular cylinder. These values are exact and are not determined experimentally. However, they need to be adjusted empirically to account for real fluid effects that are not included in the potential flow solution.
7.4. Comparison Between a Fixed Cylinder in Accelerating Flow and an Accelerating Cylinder in Still Fluid To illustrate the concepts of an added mass coefficient Ca and an inertia coefficient Cm — 1 + Ca, consider the following two examples: (i) an infinitely long circular cylinder accelerating in an unbounded still fluid (i.e., no free surface or other free or fixed boundaries); and (ii) the acceleration of an unbounded fluid (i.e., no free surface or other free or fixed boundaries) past an infinitely long fixed circular cylinder. In both of these examples, the infinitely long circular cylinder may be either vertical (x = r cos 9; y = r sin 6) or horizontal (x = r cos 9; z = r sin 9) in a 3D Cartesian coordinate system with the z axis positive up. The velocity vector is defined by the negative gradient of a scalar velocity potential according to
-
,
,
^
q = {ur,ue} = -V
f9*-
19$
-
—er + -—e# [ dr r d9
(7.17)
510
Waves and Wave Forces on Coastal and Ocean Structures
7.4.1. Accelerating Cylinder in Still Fluid For the horizontal circular cylinder in Fig. 7.10 in an unbounded otherwise still fluid where x = r cos 0 and z = r sin 0 a scalar velocity potential
0
(7.18)
b2 *(r,0,f) = -U(t)— cos<9,
(7.19)
rdr\dr is given by (Lamb, 1932, pp. 62-65)
where the cylinder is accelerating in the negative x direction. The total pressure is given by the unsteady form of the Bernoulli Eq. (3.46) in Chapter 3.4 p p
[a
l
[ dt
2
2
na<&\2"
[(£) + [r
gz +
86 )
Q(t)
(7.20)
The x component of the pressure force per unit length may be determined from -In
dFx = - f
pbcos9d9;
r = b.
(7.21)
Jo
Substituting Eq. (7.19) into Eq. (7.20) gives the total pressure on the cylinder r = b as P_ P
b
dU U2 n cosO -\ dt 2
gz +
Q(t)
(7.22)
Substituting Eq. (7.22) into Eq. (7.21) and integrating gives ,dU ,dU dFx = Capnb — = onb — dt dt
(7.23)
that gives an added mass coefficient Ca — 1.0. 7.4.2. Fixed Cylinder in an Accelerating
Flow
For a fixed cylinder in an accelerating unbounded fluid domain, the solution to Eq. (7.18) must be modified to account for the accelerating fluid in the
511
Deterministic Dynamics of Small Solid Bodies
Fig. 7.10. Cylinder accelerating in still fluid.
positive x direction according to Q(r,0,t) = -U(t)r{l
+
cos 9.
(7.24)
The total pressure on the cylinder r = b from Eq. (7.20) becomes 7 . 9 .
dU
£
b—cosO
P Pm
Pv
P
P
+ U2 sin2 6>| - gz + Q(t)
(7.25a)
(7.25b) + — + Q(t) p that gives both an inertia and drag component of pressure. Substituting Eq. (7.25a) into Eq. (7.21) and integrating gives ,dU
dFx = CmpTtbL'—dt
= IpTtb2
dU — dt
(7.26)
that gives an inertia coefficient Cm = 2.0. The inertia coefficient may be expressed as t-'/n — 1 "T C a .
(7.27)
From Eqs. (7.23 and 7.26), there is no net horizontal drag component of the force per unit length on a cylinder in an ideal fluid (d'Alembert's paradox, Milne-Thompson, 1968). This is a consequence of neglecting real fluid effects that result in flow separation and the formation of a wake. From Eqs. (7.25), the dynamic pressure in brackets {•} may be expressed as a linear sum according to PD - Pm + Pv
(7.28)
512
Waves and Wave Forces on Coastal and Ocean Structures
Fig. 7.11. Comparison between ideal and real fluid flow around a circular cylinder (R 2.0 x 105).
The dynamic pressure distribution around a circular cylinder in steady flow (i.e., dU/dt = 0) from Eqs. (7.21 and 7.20) with r = oo is Cn
Poo 2
pU /2
1 -4sin20,
(7.29)
where Cp = pressure coefficient (cf., the inverse of the Euler parameter in Table 3.2 in Chapter 3.6). Figure 7.11 compares Eq. (7.29) for a real and ideal fluid flow around a circular cylinder for a Reynolds parameter R = 2.0 x 105. To help interpret the difference between accelerating a cylinder in a still fluid from accelerating the fluid past a fixed cylinder, it is necessary to consider the momentum equation. The buoyant force in a static fluid may be computed from dz
+ y =0
(7.30a)
for a fluid volume V by
y(V) = FB =
-[d^-\v.
(7.30b)
The x-component of the horizontal momentum in an incompressible fluid from Euler's equation (Eq. (3.41) in Chapter 3.5) is 9
Du Dt
dx
(7.31)
513
Deterministic Dynamics of Small Solid Bodies
or for a circular cylindrical volume of fluid of unit height (d V = nb2) •) du
T
dp
dFm = P(7tb2)— = -{itb2)-f. 6t ox
(7.32)
Comparing Eq. (7.32) with Eq. (7.23) demonstrates that the inertia coefficient includes the effect of the pressure gradient in an accelerating fluid. Establishing that oscillating a cylinder in a still fluid is kinematically similar to oscillating a fluid past a fixed cylinder is most easily demonstrated by a noninertial moving coordinate axes attached to the center of a horizontal circular cylinder (Garrison, 1980 and 1990). The total derivative of the fluidvelocity vector that appears in the Navier-Stokes equations may be written in terms of moving coordinates attached to the oscillating cylinder, that oscillates horizontally with instantaneous velocity U as Dq Dq — = Dt Dt
dU ex, dt
v(7.33)
'
where Dq/Dt denotes the total derivative operator defined in moving coordinates where q denotes the fluid-velocity vector measured in moving coordinates. The vector gradient operator V(«) in Chapter 2.2.7 is the same in either the moving- or fixed-coordinate system; so that the complete boundary value problem written in moving coordinates is given by (Garrison, 1980 and 1990) -7-e x + - 7 = dt Dt p where v = kinematic fluid viscosity where V»qr = 0,
q =—Uex, x -* oo,
+ W2~q,
q = 0 on r = b.
(7.34)
(7.35a-c)
If the equivalent boundary value problem is formulated for oscillation of the fluid with instantaneous velocity £/ex about a.fixedcylinder, the result is Dq _
=
-WP _
+
7^ „V>,,
(736)
where V » ^ = 0,
q = -Uex,
x -» oo,
q = 0 on r = b.
(7.37a-c)
514
Waves and Wave Forces on Coastal and Ocean Structures
At an infinite distance x —>• +00 from the cylinder, the viscous and convective acceleration terms inEq.(7.36) vanish leaving only the following x component from Eq. (7.36): dU 1 3Poo ~r = ^> (7-38) v dt p dx ' where Poo denotes the pressure in the oscillatory uniform external flow. This pressure field Poo = 0 in Eqs. (7.34 and 7.35), but dPoo/dx is needed to accelerate the flow in Eqs. (7.36 and 7.37) so that P = Poo + P*, (7.39) where P* denotes the pressure due to the disturbance caused by the cylinder; and Poo = pressure in the absence of the cylinder. Substituting Eqs. (7.38 and 7.39) into Eq. (7.36) gives £/_ Dq 1 -. ^ — e x + - ^ = — V P * + vV 2 ?, (7.40) dt Dt p that demonstrates that the boundary value problem for the oscillating cylinder in Eq. (7.34) is identical to the boundary value problem for oscillating flow about a cylinder given in Eq. (7.36) with Eq. (7.36) replaced by Eq. (7.40) (Garrison, 1980 and 1990). Because the boundary value problems are identical, the solutions must also be identical; and, therefore, P in Eq. (7.34) is the same as P* in Eq. (7.36). As noted above, P* is the pressure due to the effect of the cylinder in both cases. The only difference between oscillating a cylinder in still fluid and oscillating the fluid past a fixed cylinder is that oscillatory flow past a fixed cylinder includes the additional pressure term Poo that is required to accelerate the flow with magnitude dU /dt. The integrated pressure Poo on the surface of the cylinder generates an inertia force per unit length equal to p(nb2)dU/dt. The integral of P* inEq. (7.40)andof P inEq. (7.34) generates a drag and an inertia coefficient in Morison equation. Because the integral of Poo is exact, the inertia coefficient may be written as Cm = 1 + Ca for oscillatory flow past a fixed cylinder where the 1.0 component accounts for the horizontal pressure gradient force due to dPoo/dx that is required to accelerate the fluid in the absence of a body; and the added mass coefficient Ca accounts for the oscillation of a cylinder in still fluid when dPoo/dx = 0 (Garrison, 1980 and 1990). The preceding review by Garrison (1980 and 1990) demonstrates that data obtained by oscillating a cylinder in still water is applicable to the wave force computations; or more precisely, it demonstrates that oscillating a cylinder in
515
Deterministic Dynamics of Small Solid Bodies
still water is kinematically identical to oscillating water past a fixed cylinder. The significance of the kinematic similarity between these two experiments is that data measured from oscillating cylinders in a still fluid are equally valid for wave force computations as are data measured from a fixed cylinder in an oscillating stream. This is important in that the oscillation of a cylinder in still water as a testing method has some advantages over fixed-period apparatuses such as oscillating U-rubes or variable-period wave tanks where the Reynolds parameter and the displacement ratio for a fixed cylinder size are coupled. Specifically, oscillating a cylinder in still water usually has the versatility that allows independent control over both the amplitude and frequency of oscillation. This allows independent variation of both the Reynolds parameter R and Keulegan-Carpenter parameter K (or amplitude ratio or period parameter G). In addition, when the cylinder is oscillated in still water, the horizontal pressure gradient term dPoo/dx that contributes 1.0 to the inertia coefficient is eliminated and, accordingly, so is the requirement to subtract from the measured force the force due to the pressure gradient that accelerates the free stream flow (Garrison, 1980 and 1990).
7.5. Maximum Static-Equivalent Force/Moment (Fixed-Free Beam) For constant values of Cd and Cm in Eqs. (7.10) (i.e., no dependence on R or K through depth z), a particularly useful dimensionless form for the inertia dFm (xo, yo, Zi, t) and drag dF^(xo, yo, Zi, t) force components may be obtained for vertical circular cylinders. The total integrated force/moment loads in the static-equivalent fixed-free cantilevered beam analogy shown in Fig. 7.12 where the distributed hydrodynamic wave-induced pressure loading is replaced by a total depth-integrated force Fx (r) and a moment arm I where the dimensionless time x = 2nt/f. In a 2-D non-inertial moving coordinate system {xm,z}, dimensionless variables may be scaled by dimensional variables (denoted by a tilde ~) according to: 2nxm ~ ~ § = —~— = k(xfJ - Ct), X r)(6) = —z-, H
u(%,st) =
~
2nt x = —^, T ~
(H/T)
9
s z+h s = - = —=—, h h Dx
(H/T2)
516
Waves and Wave Forces on Coastal and Ocean Structures
^a£ ^^.y
M(x) "777777777777777777777777777777,
Fig. 7.12. Static-equivalent fixed-free cantilevered beam.
and the dimensionless total depth-integrated inertia force/moment for x/ = 0 are given by
Fm(t) =
Fm(cbi)
Cm (pnfr/A)
(Hh/f*)
-L
Jo
(7.41a)
DT
r\+Hr,{z)fh
Cm (pwD 2 /4) (Hh2/f^
D M ( T > 5 )
ds,
Jo
Mm{&t)
Mm(r) =
\+Hr,(x)/h
sDu(r,s)
Dx
ds, (7.41b)
and the dimensionless total depth-integrated drag force/moment for x/ = 0 are given by
Fd(r) =
Fd{S>t)
Cd~pb~h/2 (H/ty Md(x) =
Md{(bt) 2
CdpDh /2
(Hit)'
-L
\+Hrt(x)/h
u{x,s)\u(x, s)\ds,
(7.41c)
su(x,s)\u(x,s)\ds,
(7.41d)
l+Hr,(r)/h
=1
517
Deterministic Dynamics of Small Solid Bodies
where xf = 0. A dimensionless total depth-integrated force/moment may be defined by a linear combination of Eqs. (7.41a-d) as Fx{t)
Cm7tD
Fx(a>t)
=
_
Cd (pbh/2)
_\2
My {cot)
My{x)
Fm{x) + Fd{x),
Cm TTD
2
Mm{x) + Md{x),
(7.41f)
Cd 2H
2
cd (pbh /i)
(7.41e)
Cd 2H
(H/T)
(H/T)
where nD/2H = TT2/2K and K = Keulegan-Carpenter parameter defined by Eq. (7.72). Dean (1974) tabulated Eqs. (7.41) applying his nonlinear stream function wave theory (vide., Chapter 6.5) (Dean, 1965). The solutions in dimensional moving coordinate axes {um, z\ for a dimensional streamfunction ^(^,s) and water surface elevation ??(§) are given by Eqs. (6.82 and 6.83) in Chapter 6.5. The dimensional horizontal water particle velocity u{i;,s) and acceleration Du{%,s)/Dt are given by {vide., Dean (1965); Rienecker and Fenton (1981) or Huang and Hudspeth (1984)) N+2
2ir{n — 3)s
-J^— ) V {n - 3 ) i ( n ) c o s h
.
cos((n-3)£),
X{3) (7.42a)
Du(M —
=
Dt
W{£,S)
X^'
—
'
T
du{%,s)
—
\X{3)J
h W{1;,S)
9«(g,£)
3§ (7.42b)
and the dimensional vertical water particle velocity w{%,s) and acceleration Dw{%,s)/Dt are given by N+2
w($,s)
= ( - = ^ - ) J2^n .X{3)/ n=4
3
27z{n — 3)s
)*(ra)sinh
.
X{3)
sin {{n - 3 ) £ ) ,
. (7.42c)
DwGJ) z
Dt
~(b~, = I w{t,s)
X{3)\( —
2JV —
\dw{$)
I "
T
\X{3)J
3§
dw{%,5) h w{£,s)
ds (7.42d)
518
Waves and Wave Forces on Coastal and Ocean Structures
The dimensionless integrals for Fm{x), Fd(t), Mm{x) and Md{x) are tabulated by Dean (1974) and may be applied to quickly evaluate the total dimensionless integrated loads on the vertical circular members of a spaceframe structure for dimensionless combinations of H, f and h if the inertia and drag coefficients Cm and Cd are assumed to be constants that are independent of the R and K parameters over depth. The load components may be plotted on scaled transparent paper and the space-frame structure may be loaded graphically by a wave progressing at an angle a to the structure as shown in Fig. 7.13 (Dean, 1972). The dimensionless forces and moments computed by the nonlinear Dean stream function theory may be compared with the dimensionless forces and
PLATFORM SUPPORX PILING,
0 DISTANCE (FT.)
40
80
120
Fig. 7.13. Graphical force loading on a pile supported offshore platform from the Dean Stream Function Tables (Dean, 1972 and 1974).
519
Deterministic Dynamics of Small Solid Bodies
moments computed with LWT kinematics with constant values for Cm and Cd- For a dimensional horizontal water particle velocity u(xf,z,i) given by Eq. (4.37b) in Chapter 4.4 and the local water particle acceleration computed by du(xf,z,t)/dt, the total depth-integrated inertia force Fm(r) may be computed for x/ = 0 from
Fm(r) =
kh
Fm(o)t)
L
~2~
Cmfn^H tanh kh
1
cosh.(khs) cosh kh
sinr
X — sm T. 2X0
sin T = 2
(7.43a)
where y = pg; Xo = gf2/2n = deepwater wavelength; and the wavelength X must be computed from the LWT dispersion equation according to cb2h
k$h =
(7.43b)
khtanhkh.
The total depth-integrated drag force Fd (r) may be computed for x/ = 0 from
Fd(t)
=
Fd(cbt)
gh
cdyDH2 2
4^2
cosh (khs) ^^ds JO cosh kh
gh[koh + (kh sech kh)2 COST COST
2(2Ckh)2
cost cost (7.43c)
where the wave celerity C = cio/k = X/f. The total depth-integrated inertia moment Mm{x) may be computed for Xf = 0 from
Mm(x) =
Mm(ut)
_
kh
2 cosh kh _/ ' Jo sech kh + koh — 1
Cmyn^-Hh
2kh
SUIT.
s cosh(khs)ds sinr (7.43d)
520
Waves and Wave Forces on Coastal and Ocean Structures
The total depth-integrated drag moment Md(x) may be computed for Xf = 0 from 2
Md(x) =
Md(cbt) 2
CdyDH h 2
ACkh
L
2C
scosh (khs) Y^—as cosh kh
[(1 + (khf)sech2kh
COST COST
+ k0h - 1] COST|
COST|.
(7.43e)
7.6. Parametric Dependency of Force Coefficients Cm and Ca In general, the Morison equation (7.1) is applicable to wave loadings where: (i) the wavelength is large with respect to a characteristic body dimension; i.e., b/k <3C 1; and (ii) the characteristic body dimension is a small fraction of the water depth;
i.e.,b/h « 1. The small body condition (b/k « 1 ) assures that the incident wave kinematics and dynamics are not modified (diffracted or scattered) by interacting with the small body. 7.6.1. Relative Importance ofdFm (zi, t) and dFj (zi, t) The Morison equation (7.10a) quantifies both ideal and real fluid forces through the inertia dFm{zi,t) and drag dFd(zi,t) terms, respectively. In general, both of these force components contribute to the total force on a small body. The relative contribution of each force component may be estimated from Eqs. (7.10a-c) for the x component by dFx(zi,t) = dFd(zi,t) = dFd(zi,t) 1 +
fdFm(zj,t)
\dFd(zi,t)J] L,m ~C~d
Ttb
ut(xo,yo,Zi,t) u(x0,yo,Zi,t)\u(xo,yo,Zi,t)\/
Deterministic Dynamics of Small Solid Bodies
521
where the subscript notation (»)f denotes partial temporal differentiation of the Eulerian velocity field according to ut(xo,yo,Zi,t)
-
du(xo,yo,Zj,t) dt
Linear wave theory scales give H u <x um(zi) oc —cof(zi), 2
du H , — <x -umco a - — a r / f o ) > at I
f(zi) = cosech kh coshk(zi +h),
a> = —
and consequently dFx (zi, t) reduces to 21
dFx{zt,t)
=dFd(zt,t) 1 - | ^ ) Cd) K
(7.44a) 7t2b
1 f(zt)\CdJ\
= dFd(zt,t) 1
H
(7.44b)
where the Keulegan-Carpenter parameter is defined by um(zi)T (7.44c) D Figure 7.14 (Garrison, 1978b) demonstrates the regions of validity for Eqs. (7.44) by identifying the regions where viscous and diffraction effects K =
4.0 "' I 1 '
3.0 •\ tl
HI
p
UU \
§2.0 Hi
C
\ \ \ \
\ diffraction \^_/^effects
.x V
/ \ viscous I f effects
1.0//2X/X/>// / / / / / / / I
0
I
•• E&fFRACTION THEORY • • • _^ i - - 1- —— i
0.1
0.2 D/X
0.3
0.4
Fig. 7.14. Parametric regions of relative importance of drag and inertia force components on a vertical pile (Garrison, 1978b).
Waves and Wave Forces on Coastal and Ocean Structures
522
(U
H
'5 '10 X/D
TTOO
' 1000*
Fig. 7.15. Hydrodynamic loading regimes for small bodies.
dominate the physics of the wave-induced hydrodynamic loadings. Figure 7.15 demonstrates the region of validity of the Morison equation in deep-water where the limiting steepness H/X — 1/7 is imposed. The full Morison equation (i.e., including both inertial and drag force components) is valid in deep-water when 1.0 < H/D < 10 and X/D > 20. 7.6.2. Computing the Force Coefficients Cm and Cj In order to more clearly understand the parametric dependency of the empirical force coefficients Cm and Q , it is important to note that there have been four basic types of laboratory tests plus field tests in the ocean; and that there have been three basic types of analyses applied to obtain these coefficients. These empirical evaluation methods are summarized in Table 7.1. Figure 7.16 from Garrison (1982) compares portions of the results for the drag Q and inertia Cm coefficients from the U-tube tests of Sarpkaya (1976); from the oscillating cylinder tests of Yamamoto and Nath (1976) and from Garrison, et al. (1977c). The Reynolds parameter range shown extends beyond the transition region and it appears that the post-critical region for these oscillatory flows lies well beyond R = 106 for a smooth cylinder. Sarpkaya (1977) analysed cylinders ranging in diameter between 2.0 < D < 6.5 inches. Yamamoto and Nath (1976) oscillated horizontally a cylinder with D = 12 inches in a wave channel filled to a depth of 13.0 feet. Garrison, et al. (1977c)
523
Deterministic Dynamics of Small Solid Bodies Table 7.1. Summary of 4 types of tests and 3 types of empirical evaluation methods. Test Type 1. LABORATORY i) Progressive Wave ii) Standing Wave Node iii) Oscillating U-tube iv) Oscillating Cyl Oscillating Cyl Oscillating Cyl 2. OCEAN i) Wave Project I Wave Project II ii) Ocean Test Structure (OTS)
Kinematics
Analysis
Theoretical
Phase
Morison etal. (1950)
Measured
Fourier
Measured
Fourier
Measured
Least Sqs
Theoretical Theoretical
Fourier Fourier
Keulegan and Carpenter (1958) Sarpkaya (1975, 1976, 1977) Yamamoto and Nath (1976) Garrison et al. (1977c) Rodenbush and Gutierrez (1983)
Theoretical
Least Sqs
Theoretical
Phase
Theoretical and Measured
Least Sqs and Phase
References
Dean and Aagaard (1969) Kim and Hubbard (1975) OCEAN IV Proceedings (ASCE, 1979)
Sarpkaya(1976) Yamamoto & Nath (1976) Garrison et al (1977c)
-' Roshko (1960)j steady flow 0.4
0.6
HD
2.0
4.0
6.0
ta
20.
4a
R x 10~5
Fig. 7.16. Drag and inertia coefficients in oscillatory flow past smooth circular cylinders (Garrison, 1982).
524
Waves and Wave Forces on Coastal and Ocean Structures
oscillated horizontally cylinders of D = 2.0 and 3.0 inches in a channel 2.0 feet wide, 16.0 feet long and 42 inches deep. Although exact correlation does not exist between results of the three independent laboratory experiments shown in Fig. 7.16, it is noteworthy that the trends of the data are all very similar and the agreement is actually as good as exists for equivalent steady flow experiments. However, a certain amount of disagreement should be expected in the transition region because the drag coefficient is fairly sensitive to free stream turbulence (Garrison, 1982). Some variations no doubt existed in the rate of dissipation of the turbulence level and vibration of the test cylinder. At the high Reynolds parameters the disagreement appears to decrease and the results appear to converge with the asymptotic values of Cd increasing with decreasing values of the displacement ratio 2A/D (Garrison, 1982) where A — wave amplitude = H/2. It is significant that relatively good agreement exists between these different experiments. This may have been unexpected compared to the large degree of scatter that characterizes the drag and inertia coefficient data obtained from ocean test platforms and from small-scale wave tank test (Heideman, et al., 1979). Because these results cover the transition region where the coefficients are sensitive to a number of different effects, it appears that a correct answer does not exist from the viewpoint of practical applications. However, the postcritical Reynolds parameter region is generally of more interest in design and, although data in this range for smooth cylinders is still somewhat lacking, it appears that the data in Fig. 7.16 do tend to converge at high Reynolds parameters. 7.6.3. Methods of Analyses Three methods have been applied to obtain the force coefficients Cm and Q from measured force data by minimizing the error between the measured force and the predicted force. These three methods are: (i) Phase method (ii) Least-squares method for both continuous and discrete time series (iii) Fourier analyses method Each of these three methods are reviewed briefly. If the cylinder is horizontal instead of vertical (such as in an oscillating U-tube or at the node of a
Deterministic Dynamics of Small Solid Bodies
525
standing wave), the horizontal force component dFx(zi,6) is the total force per horizontal unit length at vertical elevation z,. (i) Phase Method: Evaluating Cm and Cd from measured force traces depends on the intended application of the values. For maximum wave forces, it appears reasonable to reduce the data in a manner similar to the method applied by Morison, et al. (1950) and by Kim and Hibbard (1975). The drag coefficient Cd was computed from the phases of the measured force trace where the horizontal velocity was a local maximum and the total force was mostly drag. The inertia coefficient Cm was computed from the phases of the measured force trace where the horizontal acceleration was a local maximum and the total force was mostly inertia. In order to estimate the phase of the wave where the force predicted by the Morison equation is a maximum, the dimensional wave kinematics may be approximated for simple harmonic motion from linear wave theory by u{z,9) = um(z)cos9,
ut(z,0) = —coum(z) sinO
(7.45a,b)
Aco cosh k{z + h) 7-r-r. (7.45c) sinh kh where A = wave amplitude; and 6 = cot. The predicted dimensional waveinduced pressure force per unit length may be expressed by the Morison equation for simple harmonic data as um(z) =
dFx(z,0) = dFm(z,e) + dFd(z,e) -dFm (Z) sin 6 + dFd (z) cos 6 | cos 0 |
(7.46a)
where dFm(z) = Cm (p^-
J coum{z),
dFd{z) = -CdPDu2m (z).
(7.46b,c)
The phase at the time of the maximum force occurs near 0 = 0° and may be determined from (Morison, et al., 1950) ddFx(z,0) ^ r ^ = 0,
6 = 6max
(7.47a,b)
Waves and Wave Forces on Coastal and Ocean Structures
526
that requires that cos9max[dFm(z) + 2dfd(z) sin# max ] = 0.
(7.48)
The two possible solutions to Eq. (7.48) may be computed from cos #max = 0;
and
[dFm (z) + 2dFd(z) sin 0max] = 0
giving the following two possibilities: it
(i) #max = (2n + 1)—,
n > 0 and integer
r-\ • a dFm(z) (u) sm# max = 2dFd(z) that imply that either:
f ~ cos6»max = J 1 - sin i
(i) the maximum force is inertia-dominated and the Morison formulation is not valid
2
y/2
dFx (z, (In + 1 ) | ) = dFm{z)
(7.49a)
or (ii) the maximum force is drag-dominated and given by, approximately
dFx(z,0m!a)
=dFd(z)
1 +
/dFm(z)Y 4\dFd(z)J
(7.49b)
527
Deterministic Dynamics of Small Solid Bodies
The dimensionless force ratio dFd(z) = W dFm(z)
(7.49c)
appears repeatedly in the analyses of wave force data and is called the O'Brien parameter W in recognition of his pioneering contributions to coastal and ocean engineering. Least-Squares Method An alternative to the phase method that is readily adaptable to strictly periodic fluid or structure motion with period T = 2n/to is to apply a least-squares fit of the measured force trace to a computed Morison equation over a complete cycle of the forcing. For periodic motion past a fixed vertical circular cylinder of diameter D, a mean square error between the measured force dFj(z, 9) and the computed force dFx(z, 6) may be defined by e 2 = ([dFT(z,e) -dFx(z,8)]2)
(7.50a)
where the temporal averaging operator ((•)) is defined for simple harmonic data by ((•)> = ^~ f \*)d0
(7.50b)
and where dFj (z, 9) = measured force per unit length; 9 = cot; and the computed force per unit length dFx{z, 9) is computed from the Morison equation (7.46) by dFx(z,6) = dFm(z,6) + dFd(z,9)
(7.51)
where dFm(z, 9) = an inertia force component and dFd(z, 9) = a drag force component. Minimizing Eq. (7.50a) with respect to Cm and Cd in a best least-squares sense requires that de2 dCm
n
0,
de2 — dC- = 0. d
(7.52a,b)
528
Waves and Wave Forces on Coastal and Ocean Structures
For simple harmonic motions given by Eqs. (7.46), Eqs. (7.52) give the following formulas for Cm and Q : 2
« dFT{z,9)
. sm9d9, pDuliz)
(7.53a)
2
Cd =
* dFT(z,9)
V w Jo
pDul(z)
cos 9\ cos 9\d9,
(7.53b)
where the Keulegan-Carpenter parameter K is defined by um{z)T
(7.53c) D The coefficients Cm and Cd are functions of the vertical location z. This vertical dependence is reflected in the parameters K and R through the amplitude of the velocity um{z). K
Fourier Analysis Method Keulegan and Carpenter (1958) analyzed the standing wave-horizontal cylinder system shown in Fig. 7.17 and applied a Fourier analysis of the measured force data to obtain the following expansion for the dimensionless predicted, wave-induced force per unit length: dFx(z,9) pDum(z)
= YJ[A2n(z)cos(2n9)
+ A2n+i(z)cos[(2n + \)9]
n=0
+B2n (z) sin(2n6) + B2n+i (z) sin[(2n + 1)9]],
(7.54a)
Fig. 7.17. Definition sketch for Keulegan and Carpenter experiments (1958).
529
Deterministic Dynamics of Small Solid Bodies
where the coefficients Aj (z) and Bj (z) are again functions of the vertical location z. These coefficients may be computed in a best least squares sense by substituting Eq. (7.54a) into Eq. (7.50a) and minimizing according to 3e 2 dA2n
0,
3^ 2 dA2n+\
3e 2 n = 0, 3B2n
n
= 0,
de2 n — = 0 d#2n+l
lncA^
, (7.54b-e)
that give -2;r
A2n{z) = I" A2n+i(z)
* . 1 f ^
d F
^ \
cos(2n9)d9,
zjr 1 f*2n dFT(z 9) = - / „ ' cos((2n + l)0)0, jr J0 pDulm(z)
"2» dFT(z,9) . " ' ' K?"' sin(2n9)d9, pDuUz) 7T Jo -n-''~s
B2n(z) = - [
1 C2n dFriz 9) Bin+liz) = - / „ ' { sin((2n + \)9)d9.
(7.55a) (7.55b) (7.55c) (7.55d)
The vertical dependence on z is again reflected in the parameters K and if through the amplitude of the velocity um{z). If the dimensional measured force per unit length dFj(z, 9) is assumed to be given by Eqs. (7.46), then Eqs. (7.55) reduce to Azn(z) = 0,
A2n+i(z) =
4(-P"+1 Cd_ ,— -, r o -3 n [8n + 1 2 n 2 - 2 n - 3 ] '
fl2»(z) = 0,
B2n+1(z) = -Cm"Jn0,
(7.56a,b)
(7.56c,d)
where 5,7 = Rronecker delta function (vide., Chapter 2.2.3). Substituting Eq. (7.56d) with n = 0 into Eq. (7.55d) gives Cm =
2K f2n T /
dFT(z,9) ' V , \in 9d9,
(7.57a)
and substituting Eq. (7.56b) with n = 0 into Eq. (7.55b) gives 3 /,2?r d = -A\ 4;0
C
dFT(z,9) *\, cos 9 d9. pDuUz)
(7.57b)
530
Waves and Wave Forces on Coastal and Ocean Structures
1.6.4. Linearized Drag Force Both the least-squares and Fourier methods have been applied to provide a better fit between the Morison equation and a measured force trace. However, it is of equal importance to be able to compute the added mass and damping coefficients associated with the motion of compliant bodies. In the case of damping, it appears to be less important to accurately represent the exact temporal variation ofthe force. Rather, it appears to be more important to accurately represent the energy or work done during a periodic cycle of the motion. This concept is called the Lorentz (1926) method of equivalent work (Ippen 1966, Chapter 10.4). If the work done by the measured force is equated to that done by the velocity-squared term in the Morison equation dFd(z, 9), a definition for Cdi is obtained that is appropriate to damping of compliant bodies. The result is identical to the Fourier method; i.e., the Fourier average drag coefficient Cdi is defined in such a way as to preserve the energy dissipated over a complete periodic cycle of the motion (Garrison, 1982). The linearized drag force is equivalent to assuming that the quadratic drag force in Eq. (7.46a) may be replaced by a linearized form of the drag force by applying the LWT kinematics in Eq. (7.45a-c) according to dFdt(z,9) = dfdi{z) cos6,
(7.58a)
where dfdl(z) = Cdi^-um(z)
(7.58b)
and Cdi = a linear drag coefficient that has the dimensions of velocity um [Length/Time]. The time average rate of work (WL) done by a linear drag force dFdi(z, 9) may be equated to the time average rate of work ( WNL) done by the nonlinear drag force dFd(z,9). The time average rate of work done by the linearized drag force dFdi(z, 9) over one wave cycle is "27T
Pi = (n't) = j - /
dFdl(z,e)u{z,9)de
CdWD 2 f2* 2 cos^ 9d9 -«m(z) / An JO CdlpD
2
—-.—u m {z).
(7.59a)
531
Deterministic Dynamics of Small Solid Bodies
The time average rate of work done by the nonlinear drag force dFd (z,9) over one wave cycle is 1 f2jT PNL = {WNL) = — / 2TT JO
dFd(z,6)u{z,6)dd
CdpD 1 3 f ^ „2z cos 6>|cos#|d6» An -um(z) I Jo 2CdpD 3 = —: um(z). in
(7.59b)
Equating the average linear power P^ to the average nonlinear power P ^ L yields g Cdi = -z—CdUm(z). in
(7.60)
The linearized drag force given by Eqs. (7.58a,b) may be equated to the nondimensional A\{n = 0) coefficient term in Eq. (7.56b) for n = 0; i.e., dFdi(z,0) = PDu2m(z)AiU) (4C u (z)\ = pD I d m
cose \um(z)cos6
(7.61)
by Eq. (7.56b) for n = 0. Equating Eq. (7.61) to the linearized drag force in Eqs. (7.58a,b) also gives Eq. (7.60) proving that the Lorentz method of equivalent work (Ippen, 1966) is equal to the first term in a Fourier series expansion. The Fourier definition for Cd appears to be the most appropriate to apply to compliant bodies while the least-squares definition appears to be the most appropriate for representing the temporal variation of the wave force (Garrison, 1982). Garrison (1982) has compared the values of Q computed by both of these methods for three types of the four laboratory tests listed in Table 7.1. Garrison (1980 and 1982) demonstrates that the Fourier method gives slightly higher values than the least-squares method; however, it does not appear to be significant to distinguish between Cd coefficients computed by these two methods for practical design purposes.
532
Waves and Wave Forces on Coastal and Ocean Structures
7.6.5. Laboratory U-Tube Data Sarpkaya (1975) applied Fourier analysis to data from the oscillating U-tube shown in Fig. 7.18 to obtain drag Cd and inertia Cm coefficients on horizontal cylinders in oscillating flow without a free surface. He concluded that these two force coefficients were reasonably well-correlated with the period or Keulegan-Carpenter parameter K = UmT/D; but that plots "versus Reynolds number (Re = UmD/v) have shown that there is absolutely no correlation between the said parameters and the Reynolds number" (Sarpkaya, 1975, p. 35). Miller (1975) replotted both the original standing wave data from Keulegan and Carpenter (1958) and the oscillating U-tube data from Sarpkaya (1975) and demonstrated that both of these data demonstrate a clear Reynolds parameter dependence. Sarpkaya (1977) then re-examined both the original Keulegan and Carpenter data (195 8) as well as his own data (1975) and determined that a frequency parameter f5 = D2/vT and roughness Reynolds parameter R€ = ew m /v where e = characteristic length of the cylinder roughness could explain the earlier incorrect interpretation regarding the total lack of dependence on Reynolds parameter. The frequency parameter f) = D2/vT represents a ratio of the Reynolds parameter R to the Keulegan-Carpenter parameter K. This ratio may be computed from Eq. (3.58) in Chapter 3.6 by dimensional analysis. If the Reynolds parameter if = pb^fgbl^ is divided by the dimensionless parameter T^/g/b in Eq. (3.58) where the time was scaled by gravity g, then the frequency parameter is given by fi = R/K = b2/vT.
_v_ -VJ
Y^
5Z. / / / / / / / / / / / A
\ 777777777777777777
Jy
Fig. 7.18. Definition sketch for oscillating U-tube.
Deterministic
Dynamics of Small Solid Bodies
533
In U-tube laboratory tests, oscillatory flow past test cylinders have been obtained by sinusoidal current flow represented by u{6) = Um cos# where Um = Hn/T where H = 2 A that is twice the amplitude of the U-tube oscillation shown in Fig. 7.18. The period or Keulegan-Carpenter parameter K is a measure of the wake effects in oscillatory flows. The Keulegan-Carpenter parameter differs from the relative amplitude parameter by TT, a constant for harmonic laboratory fluid flows or sinusoidally forced body motions in still water. In the supercritical Reynolds parameter region R > 106, the drag Q and inertia Cm coefficients are sensitive to wake effects that depend on the Keulegan-Carpenter parameter K (or relative displacement or period parameter H/D) and to the relative roughness e/D as shown in Figs. 7.19 and 7.20 (Garrison, 1980 and 1982). For smooth cylinders in steady flows, the drag coefficient Q = 0.6 in Fig. 7.19; and the effect of relative roughness e/D is to increase the drag coefficient with increasing relative roughness. The additional increase in the drag coefficient Cn in the supercritical Reynolds parameter region is due to the wake effects that correlate with the amplitude parameter H/D. As the wake is washed back and forth over the cylinder in oscillatory flows, values for the drag coefficient are elevated over the values for roughened cylinders in steady flow with the more elevated values for the drag 2.4
' ' ""I ' ' ' ' ""I Typical in-service roughness
2.2
-<
• -
2.0
HID = 6.4
1.8 1.6 -
U-Tube i Sarpkaya" (1976)
1.4 1.2 1.0
r o u g h n e s s ^ l ^ y , rough, 2A/D = oo
0.8
6ffect
0.6
TTT77T7-T
' steady, smooth-*
0.4 0.2
e/D
0.0
10 s
10 4
Fig. 7.19. Effects of period parameter H/D (Garrison, 1980 and 1982).
io- 3
= 2A/D
lO"2
on drag coefficient Q
IO"1
forR
> 10 6
534
Waves and Wave Forces on Coastal and Ocean Structures
CO 2.0 1.8
1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 1C
Fig. 7.20. Effects of period parameter H/D = 2A/D on inertia coefficient Cm for/? > 106 (Garrison, 1980 and 1982).
coefficient occurring for relative small values of the amplitude parameter H/D. However, as the amplitude parameter H/D increases, the oscillatory flow or cylinder motion appear more like steady flow conditions and the drag coefficient values reduce toward the steady flow values (Garrison, 1980 and 1982). The effects of the relative roughness parameter e/D and the relative amplitude parameter H/D on the inertia coefficient Cm are compared in Fig. 7.20 to a smooth cylinder where Cm = 2.0 (cf., Eq. (7.26) in Sec. 7.4) in the supercritical Reynolds parameter region R > 106. In contrast to the wake effects on the drag coefficient Cd illustrated in Fig. 7.19, wake effects decrease the values of the inertia coefficient Cm from the smooth cylinder values. Again, wake effects are most pronounced at relative small values of the amplitude parameter H/D; and the wake effects are reduced as these values increase. The highest value of fi measured in the Sarpkaya data (1977) for the force coefficients Cd and Cm appears to be approximately fi = 8370. A representative experiment measuring wave forces on a vertical circular cylinder in the O.H. Hinsdale-Wave Research Laboratory (OHH-WRL) at Oregon State University, USA (vide., Fig. 5.1 in Chapter 5.1) would be for a cylinder diameter D = 1.0 ft., a wave height H = 3.0 ft. and a wave period T = 3.5 sec giving a /J = 30 000. A representative design specification for a tubular member on a jacket-type offshore platform may require a diameter D = 13 ft ^ 4^- ft, a wave height H = 48 ft and a wave period T = 13 s ^ An s.ForUm & Hn/T, P = 1.3 x 106. Because none of the /? values from the oscillating U-tube experiments are of these magnitudes, the graphs for the drag and inertia coefficients from Sarpkaya (1977) are not included in this review.
535
Deterministic Dynamics of Small Solid Bodies
7.6.6. Ocean Wave Data Dean and Aagaard (1970) analyzed measured wave pressure forces on two prototype offshore oil platforms in the Gulf of Mexico from Wave Projects I and II. Force coefficients Q and Cm were computed from the data by a least-squares regression algorithm applying theoretical nonlinear water particle kinematics computed by the theoretical stream function theory (vide., Chapter 6.5). They found that the drag coefficient varied with Reynolds parameter and that the inertia coefficient was essentially a constant. The Dean and Aagaard regression analysis illustrates the parametric dependence of the coefficients and explains some of the scatter observed by other investigators. Dean and Aagaard (1970) discretized continuous force records by tn — n A t and computed Cd and Cm from a least-squares fit to the discretized measured forces per unit area (psf) dFj (z,ri)by minimizing the following mean square error for discrete time series: 1 N e = -J2idFT(z,n) 2
- dFx(z,n)f
(7.62)
n=\
where dFj{z,n) = discretized measured force per unit area (psf) and dFx (z, n) = discretized predicted force per unit area (psf) given by / TCD\ Du(z,n) /p\ dFx(z,n) = Cm I p—J ^ + Cd ( | ) u{z,n)\u{z,n)\
(7.63)
where the Stokes material derivative operator D(»)/Dt is defined in Chapter 2.2.10. Note that both the measured dFj (z,n) and predicted dFx (z,n) pressure forces have the dimensions of pressure = [Mass]/[Length]/[Time]2. Minimizing the mean square error Eq. (7.63) by Eqs. (7.52a, b) yields
CmAN + —CdBN
=
TtD
Cm—BN TtD
+ ——CdbN (nD)z
-FiiN,
(7.64a)
pnD
=
-F2,N, pnD
(7.64b)
536
Waves and Wave Forces on Coastal and Ocean Structures
where N
\\ 2
/ r> I
AN=
2S\—~j)t—J
N
'
BN =
Z-,—^—u(z,n)\u(z,n)\,
N
bN = y^M4(z,w), n=l AT
Du(
Fl,N = ^2dFT(z,n)——j—, Dt
,
N
F2,N =
n—\
^dFT(z,n)u(z,n)\u(z,n)\. n=\
Solving Eqs. (7.64a, b) by Cramer's rule gives the following formulas for the force coefficients: r m
— FI>NDN ~ F2,NBN ~ p7t(D/4)DET(Cm,Cd)'
d
_ F2,NA„ - F\yNBN ~ (p/2)DET(Cm,CdY
(
' °'
where DET(Cm, Cd) = ANDN
2
- BZN.
(7.64e)
Dean, etal.{\ 91 A) reanalyzed the Wave Projects I and II data from Dean and Aagaard (1970) and their results for the inertia coefficient Cm are illustrated in Fig. 7.21 and for the drag coefficient Cd in Fig. 7.22. Because the direction of wave advance varied azimuthally around the vertical piles on the platforms in Wave Projects I and II, an inline direction of wave travel over the crest portion only of each wave profile analyzed was computed in a best-least-squares sense and an in-line wave force computed with the in-line angle for each wave crest time sequence analyzed. A resultant wave force was also computed over each wave crest time sequence analyzed. A comparison between the inline drag force coefficients and the resultant drag force coefficients is illustrated in Fig. 7.23.
537
Deterministic Dynamics of Small Solid Bodies
2J0 T V
.5 ?
08
?»Ul + •
*
AO
• a
IJO
X
• X
D
J,
V
0.6
-mO.4
0.2
0.1IXI0 4
4
8 IXIO5
6
2 Reynolds
4 6 No.R
D/A Symbol (Ft.)
4
8 1X10*
UF Symbol
Resultant
2*
o
•
©
t
'!
&
•V
+
I *
•
0
#
•
•
X
TV
It
w a
4*
371*
* f
Resultant
8 IXI0 T
Wave Project
Inline
Inline
6
Water Depth ~ 3 3 Ft. Water Deplh~ICO Ft.
Fig. 7.21. Comparison of inertia force coefficients Cm from Wave Projects I and II (Dean, et al., 1974 and Dean and Aagaard, 1970). 2.0
*
i.o
*%< x*
0.8 0.6
i r? »
va DO
"<*0.4
0.2
6
8 1X10°
2 Reynolds
DA Symbol (Ft.) 2* 3* 4* 3.7.t
Inline
Resultant
o
• *
A V
o
* f
4 6 No./}
8 IXIO"
UF Symbol Inline
Resultant
o
t
T
+ o
. •
•
X
it
4
6
8 1X10'
Wave Project i *
it
Wfaler Depth ~ 33 Ft. Water Depth ~ 100 Ft.
Fig. 7.22. Comparison of drag force coefficients C^ from Wave Projects I and II (Dean, et al., 1974 and Dean and Aagaard, 1970).
538
Waves and Wave Forces on Coastal and Ocean Structures
;
—r
i /^Resultant
J
~^J — —
- Inlin
IXIO4
2
4
6
8 IXIO*
2 Reynolds
4 6 No. A
8 IXIO*
•
•
2
4
6
8 IXH3T
Fig. 7.23. Comparison of inline and resultant drag force coefficients Q from Wave Projects I and II (Dean, et ah, 1974 and Dean and Aagaard, 1970).
7.7. The Dean Eccentricity Parameter and Data Condition Much effort has been directed toward resolving the parametric dependency of the two empirical force coefficients Cm and Cd that are applied to estimate the wave-induced pressure loads on small members by the Morison equation (cf, Sarpkaya and Isaacson, 1981 or Chakrabarti, 1987). The two most commonly applied parameters are the Reynolds parameter R = um D/v and the Keulegan-Carpenter parameter K = um T/D where um = amplitude of the horizontal water particle velocity. However, only Dean (1976) appears to have recognized the importance of the condition of the data when identifying these two empirical force coefficients in any parameter identification algorithm. Although the error ellipse concept was originally proposed by Dean (1976) to demonstrate geometrically the condition of data for identifying Cm and Q ; his original development did not demonstrate that the alignment of the axes of the error ellipse depends explicitly on either J? or K. Because the data are relatively better-conditioned for identifying the empirical force coefficient on the axis that is parallel to the semi-minor axis of the error ellipse, it is essential to be able to demonstrate that the alignment of the axes of the error ellipse depends explicitly on either R or K or, preferably, both of these parameters. It is possible to connect the parametric dependency of Cm and Cd with the Keulegan-Carpenter parameter K by the condition of the data through the Dean eccentricity parameter E as illustrated in Fig. 7.24.
Deterministic Dynamics of Small Solid Bodies
539
? CONNECTION
Fig. 7.24. Connection between parametric dependency of Cm and C^ and the data condition (i.e., orientation of error ellipse axes).
It is possible to demonstrate that, for data with kinematics that are simple harmonic, the Dean eccentricity parameter E is proportional to the Keulegan-Carpenter parameter K. Thus, the Dean eccentricity parameter E provides an explicit measure of the parametric dependency of the alignment of the semi-minor axis of the error ellipse on the parameter K. Specifically, when E < 1.0, then K < 11.40 and the semi-minor axis of the error ellipse is parallel to the Cm axis and the data are relatively better conditioned for computing Cm. Conversely, when E > 1.0, then K > 11.40 and the semi-minor axis is parallel to the Cd axis and the data are relatively better conditioned for computing Cd- When E = 1.0, then K = 11.40 and the error ellipse is a circle with zero eccentricity. It is interesting to note that a value of K = 11.40 is approximately the value of K where the peak in Cd and the trough in Cm occur in the replotted Keulegan-Carpenter data (vide., Fig. 7.26, Sarpkaya and Isaacson, 1981 and Chakrabarti, 1987) and, also, where the amplitude of the transverse lift force in the Hayashi and Takenouchi (1979) data is completely stable (vide., Fig. 7.27). Dean postulated an algorithm that is capable of including the effects of errors in the amplitude/phase of the wave kinematics on the estimates of Cm and Cd (Hudspeth, 1974). The Dean error ellipse methodology (1976) may be compared geometrically with his earlier amplitude/phase error methodology (Hudspeth, etal., 1988). In addition to demonstrating geometrically the importance of the condition of the data, the amplitude/phase error methodology also
540
Waves and Wave Forces on Coastal and Ocean Structures
(Cml
IT
Fig. 7.25. Dean error ellipses for determining the condition of wave force data to compute force coefficients (Dean, 1976).
demonstrates the importance of errors in the measured amplitude/phase of the wave kinematics. In contrast to the Dean error ellipses illustrated in Fig. 7.25, the amplitude/phase error methodology demonstrates geometrically the parametric dependency of Cm and Cd on the parameter K (or 1J) by the magnitude of the slopes of the contours of the dimensionless O'Brien force ratios W = \dFd\/\dFm | that pass through a zero error in phase at cox — 0. The advantage of the amplitude/phase error methodology is that the infinite suite of separate graphs shown in Fig. 7.25 that are required by the Dean error ellipse methodology for each constant value of the O'Brien force ratio W = \dFd\/\dFm\ may be replaced by a single graph with contours of constant values of W. The effect of W on the alignment of the Cm and Cd error ellipse axes in the Dean error ellipse methodology may be demonstrated by the slopes of the contours of W that pass through a zero error in phase. Comparisons with synthetically phaseshifted laboratory data for E ^ 1.0 (or equivalently, K ^ 11.4) are excellent for phase shifts in the range of \u>x\ < n/S (Hudspeth, et al., 1988). A definition for the term data condition may be appropriate here. Data condition is defined as the ability of a least-squares algorithm to locate a global minimum on an error surface for given wave kinematic/force data (vide., Marquardt, 1963). It is, of course, related to the numerical condition number of a least-squares error matrix (vide., Atkinson, 1989). Matrix condition numbers are computed by four standard measures for least-squares error analysis of the Morison equation. The matrix condition number is identically equal to unity whenK = 13.16 andE = 1.15.
541
Deterministic Dynamics of Small Solid Bodies
Because the Morison equation represents the inertia of the fluid (Sec. 7.1 and Sec. 7.2), it does not contain an explicit constitutive relationship for the viscous stress tensor. Therefore, it is not possible to demonstrate a similar explicit dependency of the alignment of the axes on the Reynolds parameter/?. 7.7.1. Dean Error Ellipse and Eccentricity Parameter E (Geometric) The mean square error e2 between the true force per unit length (denoted by upper case unprimed letters) dFj{z,9) and the computed force per unit length (denoted by lower case primed letters) df'x (z,0) may be estimated from Eq. (7.50a) in Sec. 7.6.3 where the temporal averaging operator (•) is defined for simple harmonic data by Eq. (7.50b) in Sec. 7.6.3 and where 9 = 2nt/T. The true force per unit length is assumed to be represented exactly by the two-term Morison equation (7.1) and is given by dFT(z,9) = dFm(z,e)+dFd(z,e)
= KmUt{z,d) +
KdU(z,9)\U(z,9)\ (7.65)
and the computed force per unit length is given by dF^z,9)^dF^(z,9)+dFd(z,9)
= k'mut(z,9) +
k'du(z,9)\u(z,9)\, (7.66)
where (»)f = 3 (•)/dt= partial temporal derivative of an Eulerian velocity field and where the true and computed generalized inertia and drag coefficients, are, respectively Km = Cm ( — - — I,
4
r
C = C(^),
Kd = Cd\ - r - I,
"
" 2
4-ci(f).
where Cm and Cd are the true coefficients and C'm and C'd are the computed coefficients. The computed inertia and drag coefficients are denoted by superscript primes (•)' in order to distinguish them from the true coefficients that are unprimed.
542
Waves and Wave Forces on Coastal and Ocean Structures
Substituting the computed force in Eq. (7.66) into Eq. (7.50a) and expanding yields e2 = [ ( / > D / 2 ) 2 ( M ( Z , 0 ) 4 ) ] X2 + [(pnD2/4)2(ut(z,9)2)]
Y2
2[(pD/2)(P7tD2/4){u(z,e)\u(z,e)\ut(z,e))'jXY
+
-2[(pD/2)(FT(z,9)u(z,9)\u(z,6)\)]X - 2 [(p7zD2/4)(FT(z,8)ut(z,9))~]
Y
+ (FT(z,9)2},
(7.67)
whereX = C^and7 = C^. Equation (7.67) is a conic section equation for an ellipse (Dean and Daliymple, 1991 or Thomas, 1965) with an origin that has been translated and rotated; i.e., (aX)2 + 2HXY + (PY)2 - 2GX - 2JY + C = 0.
(7.68)
The coordinates of the translated and rotated origin Xo = (C^)mjn and yo = (C'm)mm may be computed from the data according to (ut(z,9)2){dFT(z,9)u(z,9)\u(z,9)\)
\
-(u(z,9)\u(z,9)\ut(z,9))(dFT(z,9)ut(z,9))J
v
•^0 — (C^)min —
(p D/2) DET[X0, Y0] (7.69a) 4
lu(z,9) )(dFT(z,8)ut(z,8)} -(u(z,9)\u(z,9)\ut(z,9))(dFT(z,9Mz,9)\u(z,9)\)/ ^0 — (^m)min —
(p7zD2/4)DET[Xo,Y0] (7.69b)
DET[XQ,Y0]
= (u\z,9))(ut{z,9)2)
- (u(z,e)\u(z,9)\ut(z,9))2,
(7.69c)
and the angle of rotation may be computed from only the kinematics by co. 20 = ( " f a < » 2 > - ( ^ > 2 ( - f o ^ ) . JrD(ii(z,9)|n(z,0)|ii,(z,»)>
(7.69d)
543
Deterministic Dynamics of Small Solid Bodies
If the true and computed kinematics are simple harmonic oscillations that are given by U(z,9) = Um(z)cos(9),
Ut(z,9) = -Um(z)cosm(9),
(7.70a,b)
u(z,6) = um(z)cos(6),
ut(z,d) =-um(z)eosm(6),
(7.70c,d)
then the time averaged quantities (•) that are required in Eq. (7.67) are given by / /
m4\
(u(z,9y)
3
"m(z)
/
= —-—,
,
m
(coum(z))2
2\
(ut(z,9n
=
,
{u(z,e)\u(z,9)\ut(z,e))
(7.70e,f)
=0
(7.70g)
and Eqs. (7.67 and 7.69a-d) reduce to e2 = l(PD/2)2(3um(z)4/8)]X2
l(P7t2D/4)2(2um(z)4)/K2jY2
+
-2[(pD/2){dFT(z,0Mz,9)\u(z,e)\)]X - 2[(P7TD2/4)(dFT(z,9)ut(z,9))]Y ,r,. 0 = (Q) m in =
+ (dFT(z,9)2),
l6(dFT(z,9)u(z,9)\u(z,9)\)
v X
,
4
2K2(dFT(z,9)ut(z,9))
Y
0 = (Cm)min
(7.71a)
=
5
2
47^
'
(7.71b) (7-71c)
n pu^(z) where the Keulegan-Carpenter parameter K is defined for simple harmonic kinematics by K = ^ f l .
(7.72)
Equations (7.71a-c) demonstrate the parametric dependency of the translation of the coordinates of the origin (XQ, YO) on K for simple harmonic kinematics. Equations (7.71a-c) are similar to those given by Dean and Dalrymple [1991, Eq. (8.47), p. 226] but that are not dimensionally homogeneous and that do not demonstrate an explicit parametric dependency on K. In order to demonstrate explicitly the parametric dependency of the eccentricity of the error ellipse on the parameter K for simple harmonic kinematics,
544
Waves and Wave Forces on Coastal and Ocean Structures
note that in Eq. (7.68) that: a2 = (pD/2)2(u(z,0)4) R2 =
p2 = (P7zD2/4)2(ut(z,9)2)
> 0,
> 0,
(S/3){dFT(z,9)u(z,9)\u(z,9)\)2/u4m(z) + 2{dFT{z,9)ut{z,9))2/{coum{z))2
+ e2 -
(dF2(z,9))
and that H = 0 in Eq. (7.68). This implies (Thomas, 1965) that the translated axes of the error ellipse are not rotated and are parallel to the Cartesian axes X = Cd and Y — Cm as illustrated in Fig. 7.25. Dean and Dalrymple (1991) demonstrate that it is more illustrative for the case of simple harmonic data to write the conic section equation by completing the square of Eq. (7.68) in the following manner:
(X-Xp)2
-WaT
|
+
(F-FQ)2
-(KW- = h
(7 73a)
"
The eccentricity of the error ellipse may be defined by the Dean eccentricity parameter E given by E2 = (a/p)2 = 3(K/2n2)2.
(7.73b)
The eccentricity determines geometrically the condition of the data for identifying C'd and C'm and may be evaluated from the ratio a/p. For K < lit21V3 = 11.40 the eccentricity parameter is E2 < 1.0; and R/a identifies the semi-minor axes parallel to the Y(=C'm) axis and R/fi identifies the semi-major axis parallel to the X{=C'd) axis. For K > 2n2/*j3 = 11.40 the eccentricity parameter is E2 > 1.0; and R/a identifies the semi-major axis parallel to the Y(=C'm) axis and R/fi identifies the semi-minor axis parallel to the X(=C'd) axis. The data are relatively better conditioned for identifying the force coefficient on the axis that is parallel to the semi-minor axis as illustrated in Fig. 7.26. The eccentricity e2 of the error ellipse determines geometrically the condition of the data for identifying C'd and C'm and may be computed from (Thomas, 1965) +2
a2 = l0_E(±)\
>
E<10
> «/0 < 1-0, E > 1.0, a/p > 1.0.
(7.74a) (7.74b)
545
Deterministic Dynamics of Small Solid Bodies
R-15,000
100 K
K = ll.4
Kzll.4
E
E = I.O
E>I.O
C,
1.0 0.5
-R= 25,000 - / "-R= 20,000 R-15,000 E
.4
_ i —
10
20
, 1 ) 1 1 1
100
K
Fig. 7.26. Replotted Keulegan-Carpenter data (Hudspeth, 1991).
The parametric dependency of the eccentricity of the error ellipse and the alignment of the axes are now shown to depend explicitly on the Keulegan-Carpenter parameter K by the Dean eccentricity parameter E defined in Eq. (7.73b). There appear to be at least two sets of data where the significance of the Dean eccentricity parameter is obvious. The first data set is the well-known replotted Keulegan-Carpenter force data for a horizontal circular cylinder (vide., Sarpkaya and Isaacson, 1981 and Chakrabarti, 1987). The second data set are the transverse lift forces on circular cylinders reported by Hayashi and Takenouchi (1979). Figure 7.26 (cf., Sarpkaya and Isaacson, 1981 and Chakrabarti, 1987) demonstrates that the peak in the Q graph and the trough in the Cm graph of the replotted Keulegan-Carpenter data occur approximately at a Dean eccentricity parameter of unity or K = 11.40. A Dean eccentricity parameter of unity identifies the value of A" where the eccentricity of the error ellipse is zero and the semi-major and semi-minor axes of the error ellipse are equal (i.e., the error ellipse is a circle). Data for values of A" < 11.40 are relatively well-conditioned for identifying Cm (i.e., the semi-minor axis
546
Waves and Wave Forces on Coastal and Ocean Structures
is parallel to the Cm axis for E < 1.0); while data with values of K > 11.40 are relatively well-conditioned for identifying Cd (i.e., the semi-minor axis is parallel to the Cd axis forE > 1). It is obvious in Fig. 7.26a that the values of Cm all collapse onto a single line for E < 1.0 (or K < 11.40). However, this is not the case for E > 1.0 (or K > 11.40). Conversely, in Fig. 7.26b, there is less correlation for Q for E < 1.0 (or K < 11.40); but the data are relatively better correlated for E > 1.0 (or K > 11.40). The Dean eccentricity parameter E identifies the peak in Cd and the trough in Cm with data that are equally well-conditioned for identifying both of these two coefficients. It also delineates the two regions where the Reynolds parameter R must also be considered in addition to K. Figure 7.27 (Hayashi and Takenouchi, 1979) illustrates the transverse lift forces on a vertical circular cylinder under periodic wave excitation for various values of the Keulegan-Carpenter parameter K. The irregular behavior of the transverse lift force is evident for all values of the Keulegan-Carpenter parameter except for K = 11.8. At K = 11.8, the amplitude of the transverse lift force is completely stable. This could be explained by the error ellipse
Run No. (1)
W*H^|jh^^
RMSKC = 5.2 ,\^www-y\vWM\iviw
(3)
"iMv^^:.,;^..'.*^,,.;, .,,.^ f ,.,. J l ,,
(4)
RMS KC = 10.1
RMSKC=12
(5)
(6)
RTvlS KC = 6 2
RMS KC= 18.2
j*i§i^^
RMSKC=11.8
20 sec Fig. 7.27. Transverse lift forces on vertical circular cylinders. Note the completely stable behavior when K(= KC) = 11.8 (Hayashi and Takenouchi, 1979).
547
Deterministic Dynamics of Small Solid Bodies
being a circle at approximately K = 11.8. Other possibilities exist and the dependence on R requires further analyses. In order to determine the suitability of the data to resolve the force coefficients Cm and Cd, changes in the minimum values of these coefficients for a given mean square error must be minimized from the following Taylor series expansion (vide., Chapter 2.3.3):
e2 = (e2)ffiin + ^r(SCd) + ^(SCm) otd
+
^hL
oCm
+
JCJ-2L
+ J"
dCdaLm
(SCd)(8Cm)
+m
( 5)
-
"
The slope of the error surface at the minimum is zero; therefore, Eq. (7.75) reduces to 9 2 e 2 ,o„ w*„ , d2€2(8Cd)2 €2 = (e2)min + ——— x (8Cd)(8Cmm) + \ - 2 x ' ' dCddCm ' dC 2
, +
d2€2(8Cm)2 8Cm 2
For Cm = constant, it may be shown that this minimization of the changes in the minimum values gives (Dean, 1976)
8cd = ~y and for Cd = constant (Dean, 1976)
2 JJ -
(^
J(ut(z,6)2)
P*D
The ratio of the axes of the error ellipse may be computed from E by
8Cd
TTD]) (ut(z,0)2)'
K
"
;
For simple harmonic oscillations, Eq. (7.76) reduces to „
V3.fi:
Cm V3 \dFd\
where \dFd\ = Cdpu2n(z)/2 harmonic motions.
Cm V3
and \dFm\ — CmpjtDcoum(z)/4
for simple
548
Waves and Wave Forces on Coastal and Ocean Structures
Table 7.2. Suitability criteria for determining Cd and Cm (Dean, 1976). E
<0.26 0.26-4.0 >4.0
K
<3 3-46 >46
Relatively well-conditioned to determine Cm Cd and Cm
cd
Equation (7.77a) demonstrates that data are relatively well-conditioned to determine Cm for small values of K and are relatively well-conditioned to determine Q for large values of A". The error surface for simple harmonic oscillations in dimensionless form is e2 - (c2)mm
\fdV
Q/S)(8Cd/Cdy
2
+
(P^-ml^m)
2W1
(7.78)
Dean (1976) gives criteria for estimating the suitability of the data for determining Cd and Cm that are summarized in Table 7.2. 7.7.2. Amplitude/Phase Method
(Geometric)
The two-term Morison equation coefficients Q and Cm for a cylinder in waves depend on the correct measurement of amplitudes and phase shifts between the ambient wave kinematics and the measured force. Hudspeth, et at. (1988) apply a regression analysis to develop an algorithm that illustrates how these force coefficients will change if incorrect amplitudes or phase shifts are introduced in the analysis due to: 1) errors in the data acquisition; 2) numerical data reduction techniques; or 3) natural variations due to vortex shedding. The algorithm assumes that the two-term Morison equation models the measured forces exactly and that linear wave theory models the wave kinematics exactly. A least-squares analysis of the time-averaged, mean square error between measured and predicted forces is applied. Dimensionless variations in the force coefficients are shown to depend on two dimensionless parameters: 1) a dimensionless force amplitude ratio W (the O'Brien parameter
549
Deterministic Dynamics of Small Solid Bodies
Eq. (7.49c)) that is proportional to the Dean eccentricity parameter E and to the Keulegan-Carpenter parameter K); and 2) a dimensionless velocity amplitude ratio V(z) that is a function of the vertical elevation in the water column z. Their algorithm combines both the effects of data condition and of the wave amplitude/phase that complements the earlier development by Dean (1976). Good agreement is obtained with laboratory data of wave forces on a vertical, sand-roughened cylinder wherein the force measurements were purposefully phase shifted with respect to the wave phase in small increments up to ±33.8 degrees (±3n/l6 radians). Variability observed in the values of Cd and Cm may be due to several causes: viz., the accuracy of the two-term Morison equation; incorrect estimates or measurements of the wave kinematics; the influence from unknown roughness; measurement errors; poor condition of the data; wake encounter effects; or the inadvertent introduction of erroneous amplitudes or phase shifts into the data acquisition or the numerical analysis. There are several possible causes for a phase shift error. For example, there may be a spatial separation between the wave profile recorder, the current meter (if used), and the pile where the force is measured (vide., Fig. 7.28). The electronic or numerical filtering of data signals may introduce both a phase shift and an amplitude distortion. The sequential sampling of multiple data channels by analog-to-digital recorders introduces a small phase shift. If these potential amplitude and phase shift errors and the conditioning of the data for parameter estimation are not considered appropriately, variations in the values of the force coefficients will result. •
Wave staff
Instrumented pile
D Currentmeter
ggg •*- Force transducer
Fig. 7.28. Typical wave and wave force measurement configuration.
550
Waves and Wave Forces on Coastal and Ocean
Structures
Figure 7.28 illustrates a typical experimental configuration. The wave staff and current meter (that may not be superimposed as shown) are located at the origin, while the pile is located at some distance from the origin. A phase shift in the measurements will result from this spacing and must be taken into account. The true force is assumed to be represented exactly by the two-term Morison equation given by (Eq. 7.65). An erroneous phase shift cox between the computed force and the computed kinematics is denoted by 0 = to it + T) in Eqs. (7.70c,d). Minimizing the mean square error between Eqs. (7.65 and 7.66) by Eqs. (7.50) with respect to the computed coefficients (denoted by superscript primes) according to 0,
dk'm
dk'
(7.79a,b)
0
yields the following two equations: -Kd{U\U\ut)
- Km{Utut)+k'd{u\u\ut)
+ k'm(u2t) = 0,
-Kd(Uu\U\ \u\) - Km (Utu\u\) + k'd{u4) + k'm{u, u\u\) = 0,
(7.80a) (7.80b)
where the true and computed kinematics are defined in Eqs. (7.70). Equations (7.80) may be rearranged to give a dimensionless inertia coefficient ratio em and a dimensionless drag coefficient ratio e^ defined by the following: €/n —
c
Ca 2 (U\U\ut) + (Utut) Cm7tD (M2)
c
(Uu\U\ \u\) +
(7.81a)
(Cm/Ca)(7TD/2)(Utu\u\
(7.81b) (u4) Cd It is not a trivial task to evaluate some of the integrals in Eqs. (7.81a, b) that require absolute values of elementary transcendental functions (Hudspeth, et al., 1988). Both negative and positive phase shifts must be considered. The dimensionless inertia coefficient ratio em is given by *d
em = V(z) cos(
sin(<wr) 5TT
\(OX\ < 7T
(7.82a)
551
Deterministic Dynamics of Small Solid Bodies
and the dimensionless drag coefficient ratio ed by 1
T
ed = — V2(z)[2n + 3sm(\2(or\)-2\2a)x\ 3n + ^ - sin(|2a>T|)], 5W
- (|2WT| - 7r)cos(|2«r|)
\2cox\ < it,
(7.82b)
where the dimensionless velocity amplitude ratio V(z), the dimensionless O'Brien force amplitude ratio W, and dimensionless Keulegan-Carpenter parameter K, are defined by , Um(z) V(z) = — — , T//
W=-
Cd K j '
K
Um(z)T = —Fi—•
(7.83a-c)
The O'Brien force ratio W is the ratio between the true drag force and the true inertia force (vide., Eq. (7.49c) in Sec. 7.6.3). The magnitude of W may be evaluated to determine the condition of the data to estimate the force coefficients because it is directly proportional to the Dean (1976) eccentricity parameter E; i.e.,
where the Dean eccentricity parameter E is defined in Eq. (7.73b). The dimensionless force coefficient ratios defined by Eqs. (7.82a-d) incorporate not only the effects of amplitude/phase shift errors but also the condition of the data for estimating the force coefficients through the parametric dependency on the Dean eccentricity parameter E. The parametric dependency on the two dimensionless parameters V (z) and W (or E and K) are evaluated separately. Figure 7.29 illustrates the parametric dependency of the dimensionless drag coefficient ratio €d on the dimensionless force amplitude ratio W (or E or K) for a constant dimensionless velocity amplitude ratio V(z) = 1.0 (i.e., the computed velocity amplitude equals the true ambient velocity amplitude). For relatively large values of W(> 4.0), €d is not sensitive to the magnitude of the phase shift near the origin \cor\ = 0 . Relatively large values of W (or E or K) imply that the data are drag-dominated and are relatively well-conditioned for
552
Waves and Wave Forces on Coastal and Ocean Structures
COT ( r a d i a n s ) Fig. 7.29. Parametric dependency of dimensionless drag coefficient ratio on O'Brien force ratio W for V(z) = 1.0.
determining the drag coefficient C'd. Note that if Cd ^ 0.9 and Cm ~ 2.0, then*" ~ 22W. For relatively small values of W(< 0.1), u is very sensitive to the magnitude of the phase shift near the origin. This implies that for small values of W{< 0.1) (or K < 2.2), the data are relatively ill-conditioned for determining the drag coefficient C'd. The slope Sd of the dimensionless drag coefficient ratio ed near the origin provides additional insight into the condition of the data for estimating C'd and will be examined in detail later. Figure 7.30 illustrates the parametric dependency of the dimensionless inertia coefficient ratio em on the dimensionless force amplitude ratio W (orE or K) for a constant velocity amplitude ratio V(z) = 1.0. For relatively small values of W(< 0.1), em is not very sensitive to the magnitude of the phase shift near the origin \a>x | ~ 0. Relatively small values of W (or E or K) imply that the data are inertia-dominated and are relatively well-conditioned for determining the inertia coefficient C'm. For relatively large values of W(> 4.0), em is very sensitive to the magnitude of the phase shift near the origin. This implies that for large values of W(> 4.0 or K > 88), the data are relatively ill-conditioned for determining the inertia coefficient C'm. The slope Sm of the dimensionless inertia coefficient ratio €m near the origin provides additional insight into the condition of the data for estimating C'm and will be examined in detail later.
553
Deterministic Dynamics of Small Solid Bodies
COT (radians) Fig. 7.30. Parametric dependency of dimensionless inertia coefficient ratio on O'Brien force W ratio for V(z) = 1.0.
For small phase shifts (\cor\ ~ 0), values for Eqs. (7.82a, b) are given by the following approximations:
V{l-|'£)(«r>
(7.84a) (7.84b)
Q
Differentiating Eqs. (7.84a, b) with respect to COT gives the following formulas for the slopes: dem *Jm —
d(cor)
nWV\ wv\ 3TT J '
_= Sd
da Jo_ d(coz)
/64\ ,9^
/V 2 N F
) .
(7-85a,b)
Figure 7.31 illustrates that the slope Sm is negative near the origin, independent of cot and proportional to the product V(z)W. This confirms the earlier observation that for inertia-dominated data {W < 0.1) changes in em are relatively small and nearly independent of the phase shift cox near the origin. For the data that are ill-conditioned for determining the inertia coefficient (W > 4.0), the slope Sm and the changes in em are relatively larger near the origin. Figure 7.32 illustrates that the slope Sd is positive near the origin, independent of cor, and proportional to the ratio V2(z)/W. This confirms the earlier
554
Waves and Wave Forces on Coastal and Ocean Structures
TT "4"
-rr
n
"T
n
TT 8
4
COT (radians) Fig. 7.31. Parametric dependence of slope 5 m on the ellipse eccentricity E near the origin.
IT
Tr
4
" 8
n
jr
JT
8
4
OJT (radians) Fig. 7.32. Parametric dependence of slope Sd on the ellipse eccentricity E near the origin.
observation that for drag-dominated data (W > 4.0) changes in Q are relatively small and nearly independent of the phase shift near the origin. For data that are ill-conditioned for determining the drag coefficient (W < 0.1), the slope Sd and the changes in Q are relatively larger near the origin. Figure 7.33 compares em and Q with synthetically phase-shifted laboratory data on a vertical circular cylinder measured in the O.H. Hinsdale-Wave
555
Deterministic Dynamics of Small Solid Bodies
_| [ _ TT 4
1
1 _TT 8
,
L —, O
COT (radians)
, TT 8
,
1 TT 4
Fig. 7.33. Comparison of em ande^ with synthetically phase-shifted laboratory data (Hudspeth,
era/., 1988).
Research Laboratory at Oregon State University (Hudspeth, et ah, 1988). These synthetically phase-shifted data are for both a relatively small value of K = 7.9 (E < 1.0) and a relatively large value of* = 23.3 (E > 1.0).
7.7.3. Matrix Condition Numbers
(Numerical)
The Dean error ellipse methodology (Sec 7.7.1) and the amplitude/phase error methodology (Sec 7.7.2) provide geometric interpretations of the condition of the wave kinematic data to identify the drag and inertia coefficient C'd and C'm. Because both of these methods are derived from a least square error, standard techniques from error analyses are available to determine matrix condition numbers (Atkinson, 1989). These matrix condition numbers provide numerical measures of the sensitivity of the computed empirical force coefficients to small perturbations in the wave kinematic/force data. These numerical measures of the condition of the data may again be related to the two geometric methodologies by the Dean eccentricity parameter E. Minimizing the mean squared error defined in Eq. (7.50a) with respect to C'd and C'm gives the following matrix equation: AX = B
(7.86)
556
Waves and Wave Forces on Coastal and Ocean Structures
"^ •
~**W
RUN 131 ' K= 30.8
RUN 129' K= 28.76
VTVAJ RUN 133 = K = 17.18
RUN 145= K= 13.02
Fig. 7.34. In-line and transverse lift forces on a vertical circular cylinder forced to oscillate in still water (Maull and Milliner, 1979).
where the scaled matrices in Eq. (7.86) are given by A=
3K
V
0
1
\in °1 0
X =
1
c
(7.87a)
(7.87b) (7.87c)
LWJ
(Fu Ut t) B = ¥\}f(Fu\u\) \.].
(7.87d)
Matrix A is Hermitian and unitary. It becomes a unit matrix with matrix condition numbers identically equal to unity when A" = 4TT 2 /3 = 13.16 and E = 2/y/3 = l.5. The transverse lift force on a vertical circular cylinder forced to oscillate in still water measured by Maull and Milliner (1979) are shown in Fig. 7.34. The transverse lift forces are stable and repeat exactly only when K = 13.02 for RUN 145. The four standard measures of the condition number of the error matrix A defined in (Eq. 7.87a) are summarized in Table 7.3. The matrix condition number when K = 11.40 or E— 1.0 are also tabulated in column 3 of Table 7.3.
557
Deterministic Dynamics of Small Solid Bodies
Table 7.3. Summary of matrix condition numbers. Matrix condition number (Atkinson, 1989)
K= 13.16 £=1.15 (2)
# = 11.40 E= 1.0 (3)
1.0
1.15
1.0
K < 13.16 E < 1.15 (4)
K > 13.16 E> 1.0 (5)
(1) Cond (A)i
4JTZ _
2
7
3K
£V3
2
IK ~ £\/3"
4;r
1.15
4 ^ _ £A/3 3tf
3A"
1.15
4*f
Cond (A) c Cond (A) 2 Cond (A)*
1.0
=
3# ~ £ v ^
4JT
3tf 4jr
£V3
2
£V3
2
The four standard matrix condition numbers in column 1 of Table 7.3 are defined as follows (Atkinson, 1989): Cond(A)i = Cond(A)oo Max | A. |
Cond(A)2 =
kecr[A x A] Min \k\
lAI.IA" - i i
1/2
" Max W '
Cond(A*) =
kea[A] Min \\\
Xeo[A x A] where | • | = a matrix norm; A - 1 = matrix inverse; A. = eigenvalue of the matrix;CT[»]= spectral radius of the matrix ([•]); and A* = complex conjugate transpose (Atkinson, 1989). Summary The condition of wave kinematic/force data to identify the empirical force coefficients Cd and Cm in the Morison wave force equation for small bodies may be evaluated by three methods; two geometric (Sees. 7.7.1. and 7.7.2) and one numerical (Sec. 7.7.3). The two geometric methods are the Dean error ellipse methodology and the amplitude/phase error methodology. The Dean error ellipse demonstrates geometrically the condition of the data by the alignment of the Cd and Cm axes of the error ellipse. A separate error ellipse is required for each value of the O'Brien force ratio W as shown in Fig. 7.35. The amplitude/phase methodology demonstrates geometrically the condition
558
Waves and Wave Forces on Coastal and Ocean Structures
e
2>el / " ~ \
E >I.O IT > 1.0
E< 1.0 w< 1.0
Cm 63 >6|
'CmW
6£0
'^d'min
I
* 'cd'min Fig. 7.35. Suite of Dean error ellipses required for each value of W.
of the data by the magnitude of the slopes of contours of the force coefficient ratios W passing through a zero phase error cox = 0 as shown in Figs. 7.31 and 7.32. Each of the separate graphs in Fig. 7.35 for each separate value of W required by the Dean error ellipse methodology may be replaced by a single graph with contours of W as shown in Figs. 7.29-7.32. Both of these two error methodologies may be related to the Keulegan-Carpenter parameter K by the Dean eccentricity parameter E. The Dean eccentricity parameter E = */3K/(2TT2) provides a geometric measure of the condition of wave force data on circular members for estimating the force coefficients C'd and C'm. Two sets of data appear to illustrate the physical significance of the Dean eccentricity parameter. The variability in Cm for E > 1.0 (or K > 11.40) and in Cd for E < 1.0 (or K < 11.40) in the replotted Keulegan-Carpenter data in Fig. 7.26 may be explained by dividing the data into two parts determined by a Dean eccentricity parameter equal to unity (E = 1). The unusual stable behavior of the amplitude of the transverse lift force in the Hayashi-Takenouchi (1979) wave force data in Fig. 7.27 may possibly be explained by a Dean eccentricity parameter of unity. The axes of the Dean error ellipse are shown to be parallel to the Cm and Cd axes for simple harmonic kinematics (i.e., (u\u\ut) = 0). The Dean eccentricity parameter E may be incorporated into an error analysis that also includes errors in the amplitudes/phases of the kinematics. Comparisons shown in Fig. 7.33 with synthetically phase-shifted laboratory data are quite good for phase-shifts \cox\ < T T / 8 .
Deterministic Dynamics of Small Solid Bodies
559
Four measures from standard matrix error analyses may be computed to determine the matrix condition numbers for the least square error between the measured and predicted wave forces. Each of the four error matrix condition numbers are identically equal to unity when K = 13.16 or when E = 1.15. The only stable transverse lift force in the Maull and Milliner (1979) data shown in Fig. 7.34 occurs when K = 13.02. The matrix condition numbers tabulated in Table 7.3 are equal to 1.15 when K = 11.40 and when E -\.Q. The Dean eccentricity parameter E may be applied to compare each of the three methods to evaluate the condition of the wave kinematic/force data to identify the force coefficients Cj and Cm. It also connects each of the methods to the Keulegan-Carpenter parameter K. To illustrate the connection of the eccentricity E to design conditions, a typical design specification for a member on a jacket-type offshore platform may require a pile diameter D = 13 ft «s 4TV ft (4 m) for a design wave height H = 48 ft (15 m) and a wave period T = 13 s « 4n s. For Umax « HTT/T, thenA" = 12andi? = 1.5xl0 7 and/J = 1.3 x 106. There are very few data sets available for Cm and Cd in the narrow interval between 11.40 < K < 13.16 with correspondingly high values of the Reynolds R or frequency fi parameters.
7.8. Modified Wave Force Equation (WFE, Relative Motion Morison Equation) The linearized modified Wave Force .Equation (WFE, relative motion Morison equation, American Petroleum Institute, 1987) is applied to two types of structures that consist of small members. The first type is a vertical circular cylinder that is an articulated tower and that oscillates about a base that is connected to the ocean bottom. This is a single-degree-of-freedom oscillator (SDOF) that responds to a wave induced, relative motion moment load M$(t, ao, ©5, 6)5). The second type consists of two semi-immersed horizontal circular cylinders that are tautly moored to the ocean bottom, connected on each end by rigid horizontal members, and floats horizontally as a pontoon system. This is a three (multiple)-degree-of-freedom oscillator (MDOF) that responds to two wave induced, relative motion force loads F2(x{,Zi,t,ao,X2,X2) and Fi(xi,Zi,t,ao,Xi,Xi) and a wave induced, relative motion moment load Ms(xi,Zi,t, ao, ©5, ©5). These two examples illustrate both translational and rotational dynamical systems of small members. The goal of both of these
560
Waves and Wave Forces on Coastal and Ocean Structures
two examples is to solve the following dimensionless dynamic equations of motion in the y'th mode of oscillation/rotation:
Xj(r) + 2!ja>JhXj(T) + &2jhXj{x) = SstjS>2Jh J ^ J (T -ft),
j = 1,2,3, (7.88a-c)
&j(r) + 2Cj5>jh®j(T) + a>2jh®j(r) = AstjS>2h {**| (r -
ft),
j = 4,5,6, (7.88d-f)
that may be readily solved by the well-known method of undetermined coefficients (vide., Chapter 2.5.6) and where the overdots (•) denote ordinary temporal derivatives of a Lagrangian solid body; i.e., d (i) = —,
di'
.. _ d2 (i) (#) = - j .
~ dJi
(7.89a,b)
The linearized WFE separates the wave induced relative motion loads into two components: viz., 1) an added mass and a fluid viscous damping component that are proportional to the Lagrangian solid body acceleration Xj(z,aj)/@j(T,<Xj) and velocity Xj(r,aj)/&j(r,aj) in the ;'th mode, respectively; and 2) a wave induced exciting force/moment load F?(t, ceo, Pj)/M?(t, ao> Pj) m the jth coordinate direction on a fixed body. Because this linear decomposition of the WFE loads results in a wave induced hydrodynamic force/moment load on a fixed body, fixed bodies are not analyzed separately because fixed bodies experience the same exciting force/moment loads (correct to the linear approximation) as bodies that respond dynamically! The damping ratio fy in the jth mode of oscillation/rotation is due entirely to the fluid hydrodynamic force/moment from the linearized WFE. Once the dimensionless dynamic equations of motion have been derived, the fundamental problem solution technique is an application of the method of undetermined coefficients (vide., Chapter 2.5.6) given by the following sequence: (i) Compute the amplitude F?/M? and phase fij of the jth mode exciting force/moment load on a fixed body by the linearized WFE
561
Deterministic Dynamics of Small Solid Bodies
according to:
Ff(t,a0,Pj) Mf(t,a0,Pj)
•ao±Pj).
(7.90a) (7.90b)
(ii) Select the form of the solution for the dynamic response in the jth mode of oscillation/rotation by the method of undetermined coefficients in Chapter 2.5.6 from the form of the family of the exciting force/moment load F?(t, ao, f5j)/MJ(t, ao, fij) according to [Xj(f,aj)\ _ Uj 1 (sin)
{0,(^ 7 .)J-i^Hcosl
_
(ft;f
(7.91a)
ao±
^'
(7.91b)
where ao = phase of the free surface wave profile r](x,t,oto) and the choice of {sin/cos} depends on the family of the exciting force/moment loads in Eqs. (7.90). (iii) Compute the dimensionless natural frequency for the jth mode with hydrodynamic effects (viz., added mass) a>jnfromthe stiffness Kj /KJ and mass/mass moment of inertia m/Ij by lKj/(m -2 °>jh =
w
-2
+ Ma)
(7.92a) )
'
[ Kj/Vj + Ia) J
(7.92b) where Ma/Ia = the dimensional added mass/mass moment of inertia from the linearized WFE. Compute the dimensionless damping ratio £j from the dimensional damping coefficient Cj from the linearized WFE and critical damping Cjcr by
0=
-jcr
y/Kj{m
t-' jcr — -^
I V^^J
+ M_a)\ +
la) V
(7.93a,b)
(iv) Compute the static displacements 8stj/Astj from the amplitudes of the exciting force/moment load ^ ^ / M ? and stiffnesses KJ/KJ (vide., Chapter 2.5.6) by (7.94a) •
=
'
[ Kj
)
(7.94b)
562
Waves and Wave Forces on Coastal and Ocean Structures
and dynamic magnification factor D (&>;%) (vide., Chapter 2.5.6) from DOSJfc) = L(l - 5>~]l? + Qtjl&jhfr1'2.
(7.95)
(v) Compute the dimensionless amplitudes |§/|/|£27-| of the y'th mode of oscillation/rotation from the static displacement 8st/Ast. (vide., Chapter 2.5.6) and dynamic magnification factor D \(o~^) by l£/l 1
TW~-2N f &stt \
(7.96a) (7.96b)
(vi) Compute the phase angel aj in Eqs. (7.91a,b) from the family of the exciting force/moment solutions by the method of undetermined coefficients (vide., Chapter 2.5.6) by first defining tj = arctan ( /
__, J .
(7.97a)
For the family of exciting force/moment given by Eq. (7.90a) with the response solution given by Eq. (7.91a) and with the same choice of ± sign in each equation, the phase angle aj is given by a;3 = arctan
/ tan 8 ± tan e \ . VlTtanetanB/
v(7.97b,c)
'
For the family of exciting force/moment given by Eq. (7.90b) with the response solution given by Eq. (7.91b) and with the same choice of ± sign in each equation, the phase angle aj is given by a,1 = arctan
/ tan B :p tan e \ }. Vl±tanetanB/
v(7.97d,e)
"
The following two examples in Sec. 7.8.1 and 7.8.2 demonstrate this problem solution technique outlined above in Steps (i)-(vi) by the method of undetermined coefficients (vide., Chapter 2.5.6) for both SDOF and MDOF dynamical systems.
563
Deterministic Dynamics of Small Solid Bodies
Fig. 7.36. Definition sketch for articulated SDOF tower.
7.8.1. Articulated Circular Cylindrical Tower: SDOF System The dynamic equation for the SDOF motion in the pitch mode response rotating about a bottom-fixed base located at s = z + h of the articulated tower shown in Fig. 7.36 is /5©s(0 = [-FBhB
+ Wbhb + Wghg + Wphp]&s(jt) +
= -K*5e5(t)+
M5(t,ao,®5,®s) (7.98a)
M5(t,a,&5,G5),
where the over dots represent ordinary temporal derivatives of a Lagrangian tower according to ..
d2& dtz
.
d@ dt
(7.98b,c)
where I5 = total mass moment of inertia about the bottom-fixed base; Fg = buoyantforce(=pg7rD 2 /i/4); Wb = total weight of ballast; Wg = total weight of the tower material; Wp = total lumped-mass weight of the platform superstructure; M5(t,ocQ,@5,®5) = wave-induced hydrodynamic load moment computed by the linearized WFE where ao = an arbitrary phase angle for the free surface profile r)(x,t,ao); K^5 = rotational stiffness due to excess buoyancy that is given by K;5
= Wgh
El
m-imymm-m (7.99a)
564
Waves and Wave Forces on Coastal and Ocean Structures
and where 0 5 (O = |fi 5 | cosioot - a0 - a5) - RQ{Q5 exp ~i(cot - ao)}.
(7.99b)
The contribution from each individual moment in the term in square brackets [•] in Eqs. (7.99a) to the total restoring moment K^5®s(t) about the point of rotation with coordinates {XR,SR} = {0,0} may be computed from the geometry shown in Fig. 7.36 according to J^rRj
xFj
= - £ ( X - XR)j(±Fj)ey,
j
j = b, B,g,p,
(7.99c)
J
where each moment arm (X — XR)J may be approximated for small rotations ©5(f) from the vertical equilibrium geometry shown in Fig. 7.36 by (X - XR)j = hj tan &s(t) « hj&s(t).
(7.99d)
The total mass moment of inertia I5 for rotations about the base of the articulated tower located at the ocean bottom at s = z + h = 0 is a linear combination of the individual mass moments of inertia about the bottom-fixed base from the ballast component l\,\ from the structural tower component Ig; and from the lumped mass platform component Ip: h = h + Ig + IP,
(7.100a)
where the mass moment of inertia about the center of gravity IJCG for each y'th structural component may be computed from non-inertial axes x'j located at the center of gravity of each 7'th structural component from IJCG
= /
rfdnij
Jm i
PjrfdVj;
rf = xf + yf,
dVj = r'jdcpdr'jds,
- / / / . (7.100b-d) where the homogeneous mass density pj for the y'th structural component may be computed from WPj = -rjr, (7.100e) where Wj — the total weight of the jth structural component and where the vertical integration over ds is over the total length Lj of each yth structural
565
Deterministic Dynamics of Small Solid Bodies
component at equilibrium ©5(0). The mass moment of inertia Ij about the point of rotation at s = X = Y = 0 for the y'th structural component in Eq. (7.100a) may be computed by the parallel axis theorem according to
/,- = IJCG + h2j^ = fffWj PF?dv + ^ 5 y ^ / gn{D>l/4Lj)J0 W D2
, u2^j
28 4
+
J
/ Jo
_u2Wj
= hi
^
+ h1:^J 8
/ r'fd^drjds J Jo J 1 +
1 /D 8 \h
j = b,g,p.
(7.100f)
Because the platform super structure is idealized as a lumped mass with no size (i.e., D = 0), Eq. (7.100f) reduces to h2pWp
h=
(7.100g)
8
Substituting Eqs. (7.100f, g) into Eq. (7.100a) and neglecting the D/hj terms gives for the total mass moment of inertia / 5 = [Wbh\ + Wgh] + Wph2p]/g Wgh? 8
h,
+
(7.101)
where the heights hj from the fixed base bottom at 5 = X = Y = 0 are illustrated in Fig. 7.36. The wave-induced, hydrodynamic load moment M${t, ao, ©5, ©5) may be approximated by the linearized modified wave force equation (WFE, relative motion generalization of the Morison equation) according to M it „ & & \
I f0hdMm(s,t,ao,&5) +JQ dMdi(s,t,a0,®5) [ +Jo dMFK(s,t,a0)
I J (7.102)
where the relative motion differential inertial component dMm (s, t, ao, ©5) is dMm(s,t,a0,&5)
= Ca(p7TD2/4)s[ut(s,t,ao) - s@5(t,a0)]ds,
(7.103a)
566
Waves and Wave Forces on Coastal and Ocean Structures
where the temporal partial derivative of an Eulerian field variable is denoted by a subscript according to 3(.) (•)t
=
dt '
and where the relative motion differential linearized drag component dMdi(s,t,a0, ©5) is dMdi(s,t,a0,&5)
= CaiipD/2)s[u(s,t,ao)
- s@5(t,ao)]ds,
(7.103b)
and the Froude-Kriloff component dMp# (s,t,ao) is dMpK(s,t,ao) = (picD /4) sut(s,t,ao)ds
(7.103c)
where the Froude-Kirloff component is the inertia of the fluid in the absence of the body (vide., Chapter 3.3.1). Assuming that linear wave theory is valid and that the free surface displacement is given by H
\H
rj(x, t, ao) = ~r COS(KX — cot + ao) = Re | — exp i(kx
— cot + ceo)
(7.104a) where the arbitrary phase angle «o m Eq. (7.104a) is introduced for data analyses by Finite Fourier Transform (FFT) algorithms (vide., Chapter 9.2), then the horizontal components of the wave-induced velocity u(s,t, ao) and acceleration ut (s, t, ao) may be expressed by Hco cosh ks u(s, t, ao) =
2 sinh kh COS(KX — cot + ao) Hco cosh ks expi(kx — cot + ao) = Re 2 sinh kh — RG{UM(S) exp i(kx
— cot +
ao)}
(7.104b)
and by Hco2 cosh ks . ut(s,t, ao) = — sin(o: — cot + ao) 2 sinh kh Hco2 cosh ks = Re • exp i (kx — cot + ao) 2 sinh kh Re{—icoUM(s) exp i(kx — cot + ao)},
(7.104c)
567
Deterministic Dynamics of Small Solid Bodies
where the depth dependent velocity amplitude Um (s) from linear wave theory is given by Hcocoshks ,„,„..,, UM(s) = -^r^-rvT(7.104d) 2 sinh kh If the slope of the instantaneous water surface profile r\x (t, ao) and the temporal derivatives of the slope of the instantaneous water surface profile evaluated at x = 0 are given by r]x(t,ao) = sm(cot — ao) = Re ji—— exp — i(cot — ao)} (7.105a) Vxt(t,cio) =
Vxtt (t, ao) =
cos(cot — ao) = Re I —— exp —i(a>t — ao)} (7.105b) kHco2 f Ww2 , 1 — sm(cot - a 0 ) = Re j -i —-— exp -i(cot - a 0 ) | (7.105c)
then the wave-induced water particle kinematics in Eqs. (7.104a, b) may be defined in terms of the temporal derivatives of the slope of the free surface profile at x = 0 according to Hu> cosh ks 1
{
{
————- exp-i (cot -ao)\ = 2 sinh kh J Hco2 cosh ks -i—
2
hV(s)r)xt(t,ao), (7.106a)
1
. exp -i(cot - a 0 ) } = smh A:rt J
h^(s)r)xtt(t,a.o),
(7.106b) where the dimensionless depth dependency is cosh ks kh sinh Krt The relative motion differential component of the inertial dMm (s,t, ao, ®s) and of the Froude-Kriloff dMpK (s,t, ao) moments may be combined with the second temporal derivative of the slope of the water surface profile x]xtt (t, ao) according to dMm(s,t,a0,@5)+dMFK(s,t,ao) = (P7tD2h/4)(s/h) [Cmh^(s){r]xtt(t,ao)}
- Cas{®5(t,a0)}]ds,
(7.107a)
568
Waves and Wave Forces on Coastal and Ocean Structures
that may be integrated over the SWL depth h in Eq. (7.102) if both Cm and Ca are constants to obtain Mm(t,a0,®5)
+MFK(t,a0)
/•Ffl/i 2 \ f"
fh
= / {dMm(s,t,a0, Jo
„,,?/,
3C
2 l
,, ,
1
1
©5)
+dMFK(s,t,a0)}
cosechMX
= (-IT j [ ^- ( ~ ^ r + - £ - ) x{rixtt(t,ao)}-Ca{G5(t,a0)}
,
(7.107b)
where Cm = 1 + Cfl,
FB = ynD2h/4.
(7.107c,d)
The linearized drag coefficient Cdi{s) is a function of the depth 5 and is related to the quadratic drag coefficient Cd in Eq. (7.60) in Sec. 7.6.4 by f27T 2 Cdi(s) — / cos (a)t — ao)\ cos(a>t — ao)\d(o)t) n Jo SCdUM(s) (4Ha>\ = \ = Cdkh — - *(*) (7.108) CdUM(s)
that has the dimensions [Length/Time] of the velocity Um(s) in Eq.(7.104d). The differential component of the linearized relative motion drag moment dMdi(s, t, ao, ©5) may also be combined with the first temporal derivative of the dimensionless wave slope r)xt(t,ao) according to dMdi (s,t,UQ,®5) (pD\ = Cd l^-J
{4Hkhco\
r ?
•
1
( — 3 — ) (sh) [* 2 (j){ifc,(f,a 0 )} - ^(5){© 5 a,Qfo)}J ds (7.109a)
and may be integrated over the SWL depth h if Cd is a constant to obtain fh Mji(t,ao,&s) ^ ^
= I Jo
dMdi(s,t,ao,®5)
(H/D)(ko/k)(FBh/3) a>{Tckh)2
(2kh (kh + sinh 2kh) + 1 - cosh 2kh \ . ,2,, V sinh kh / -S[kh(kh 2(k/k0))+2]{®5(t,a0)}
{lxt(t,a0)}
(7.109b)
569
Deterministic Dynamics of Small Solid Bodies
The dimensional dynamic equation of motion given by Eqs. (7.98a-c) may be scaled by the following dimensionless variables (with tildes ~ except for a dimensionless time r): T
= cot, x = kx, ®5(r) — (2h/H)®5(t,a0),
fj(r,ao) =
rj(t,ao)/(H/2).
Dimensionless temporal and spatial derivatives are now given by m(r,a0)
= nx(t,a0)/(kH/2),
fjxzir, «o) = r)xt(t,a0)/(kHco/2), 2
ilxxr(r,c(o) = r,xtt(t,ot0)/(kHco /2),
05(T) =
®5(t,a0)/(H/2h),
05(r) =
®5(t,a0)/(Hco/2h),
05(r) =
®5(t,a0)/(Hco2/2h).
Equations (7.98) may now be scaled by dividing by a buoyant moment = (FBh/3)(koH/2) to obtain the following dimensionless equation of motion: L/5 + C a J© 5 (r,ao) + Bv®5(t,a0)
+ K5®5(i,a0)
= M 5 £ (r,a 0 ,&) =
Km{fjxtt(r,ao)} +
Cddfjxtir,ao)}, (7.110a)
where M$ (r, ao, fis) = exciting moment on a fixed tower. The dimensionless mass moment of inertia is given by
<%)m+my<m)\ <•»*> FB.
the dimensionless added mass by (7.110c) the dimensionless viscous damping moment coefficient by Bv = Cd
(tor)
kh(kh-2(k/ko))+2 ,
(7.110d)
(Ttkh)3
and the dimensionless rotational stiffness moment coefficient by ^5
=
Mi^KF \)
w
g
B
~(FB\(hB\
(Wb\ fhb
l\wj\h J \wg)
m -m
U
(7.110e)
570
Waves and Wave Forces on Coastal and Ocean Structures
The dimensionless inertia moment coefficient is Km =
(kh)(k0h)
[koh — 1 + sech kh],
(7.110f)
and the dimensionless linear drag moment coefficient is '2kh(kh + sinhlkh) + 1 - cosh2A:/z Cdi = Cd ( —
(nkh)2 sinh2 kh
(7.110g)
Figure 7.37 illustrates the parametric dependencies on the dimensionless deepwater wave frequency koh = co2h/g of the dimensionless viscous damping moment coefficient B1' B" =
Bv Cd(H/DY
(7.11 Oh)
of the dimensionless inertia moment coefficient K'„ K
m =
(7.110i)
-fT>
and of the dimensionless linear drag moment coefficient Cidi
(7.110J)
r" -dl —C (H/DY d 100
0.01 0.01
0.1
1
10
kjl
Fig. 7.37. Dimensionless hydrodynamic coefficients for an articulated SDOF tower.
571
Deterministic Dynamics of Small Solid Bodies
The solution kh to the following dimensional linear dispersion equation is also shown: co2h
(7.110k)
= koh = khtanfakh.
The dimensionless dynamic equation of motion in Eq. (7.110a) may now be expressed in the following dimensionless form (cf, Eqs. (2.99) in Chapter 2.5.3) for a SDOF damped harmonic oscillator: 0 5 ( r , a o ) + 2£ 5 <»5A©5(T, «O) + a>ih@s(T,a0) = Mf(T,a 0 ,ft),
(7.111)
where the dimensionless natural frequencies <wo, /o and period To without hydrodynamic added mass are given by
where the dimensionless critical damping Bcr and damping ratio £5 are given by
Bcr-2y/K5(I5
+ Ca),
^5 =
Bv ^ , Brr
(7.112d,e)
and where the dimensionless frequency ratios a>$h, fs„ and period fsn with hydrodynamic added mass are given by (cf., Eqs. (2.99) in Chapter 2.5.3) u\h
K5
a>5h
(h + Ca)
(JO''
T Tsh
— fsh>
(7.112f-i)
where Eq. (7.112g) for fen is the inverse of the traditional definition for the dimensionless frequency ratio (vide., Chapter 2.5.3 or Clough and Penzien, 1975 p. 53, Eqs. (4-8)). The dimensionless exciting load moment on a fixed tower is given by M5£(r,a0,#0 =
h + Ca
= M5 cos (T — ao + P5),
(7.113a)
572
Waves and Wave Forces on Coastal and Ocean
Structures
where the dimensionless exciting moment amplitude M^ and phase ^5 in Eq. (7.113a) are
- F J (Km)2 + (Q/)2 Mf = 1 = , h + Ca
05 =- arctan I -z^\Cdi
(7.113b,c)
The solution to Eq. (7.110a) by the method of undetermined coefficients (vide., Chapter 2.5.6) is assumed to be a near-surface slope follower r\x given by the real part ofEq. (7.99b) as 05(r,ao) = ^5 cos(r — ao — as),
(7.114a)
where Q5 = D [(S)5hr2]
A« 5 ,
Ast5 = ^ - ,
Si5h
D [(^5/z)-2] = [ (1 - &jfy + as5/u5hy
(
(7.114b,c)
1/2
(7.114d)
( €5 = arctan I \ 1 - a>.5h-2 , (7.114e,f) where the signs in Eq. (7.114e) are changed from the signs in Eq. (2.101c) in Chapter 2.5.3 because the sign of the forcing phase angle ^5 in Eq. (7.113a) is changed from the sign of the forcing phase angle fi in Eq. (2.99a) in Chapter 2.5.3. The dimensionless damped natural frequencies a>5h,fsd and period 7 ^ are .,2 5d_
/ T \ 2
^
\ + Ju /
*h = ^T = (^-)
tan65 — tanfis \ , l+tane5tanft/
/ i \ 2
=^d-?52)-
=fsd = (^-)
(7.H4g-k)
\ •« 5 d /
The dynamic magnification factor DL(ft>5/,)~2J and phase angle €5 are shown in Fig. 7.38. The dimensional (without tilde ~) natural frequencies are easily shown to be col = a?w\,
co\h = to2 &\h,
a>\d = co2 d)25d.
(7.115a-c)
The dimensional (without tilde ~) inertia, damping and stiffness are /s =
'FBh^ 3g
k;Bv = (^\
k0hBv;K5 = H^kohKs
(7.116a-c)
573
Deterministic Dynamics of Small Solid Bodies
(a)
>J3 P
(b) 18° 160 140 120
10
^ ° f» 80
|c. =«h-* _/^^r===== [771 1MJ I—i
MJ
"li/^^^^^^^^ ^ **\i3\ -|oT|
60 40 20 0
Hi.o [ 0.5
1.5
2.5
w"« = r,,, IT Fig. 7.38. Parametric dependencies on the dimensionless damping ratio £5 of the a) dimensionless dynamic magnification factor D[ftW ~\ and b) relative phase angle 65.
7.8.2. Two Semi-Immersed Horizontal Circular Cylinders: MDOF System A single vertical circular cylinder is analyzed in Sec. 7.8.1 as a SDOF articulated tower that is constrained to rotate about a bottom anchored base in the x-z plane in response to a wave induced moment load. Here two circular cylinders are rigidly connected at their centers as shown in Fig. 7.39 and floated horizontally as a tautly-moored semi-immersed pontoon system. For normally incident surface gravity waves and a collinear current Uc, this system will oscillate as an uncoupled 3 degrees-of-freedom (MDOF) oscillator in the sway X2O), heave X3O) and pitch &s(t) modes of oscillation (vide., Chapter 8.1 for the definitions of these modes of oscillation). The pontoon system is constrained by two taut mooring lines with equal tension magnitude
574
Waves and Wave Forces on Coastal and Ocean Structures
Fig. 7.39. Two semi immersed, tautly moored horizontal circular cylinders of length B.
Fig. 7.40. Locations of the five separate coordinate axes on the MDOF pontoon system.
P in each mooring that are attached to the rigid connecting members "a-c-b" at point "c" at each end as shown in Fig. 7.39. The two horizontal circular cylinders, each of equal length B, are connected at their centers "a" and "b" at each end with a rigid connector "a-c-b" of length x as shown. The two pontoons float horizontally under the action of a steady current Uc only with two rigid connections "a-c-b" at each end of the pontoons that are located at the still water level (SWL). This MDOF application illustrates the static equilibrium conditions required between the Archimedes buoyant forces and external mooring constraints (vide., Chapter 8.2 where this static equilibrium constraint is generalized in the dynamic boundary conditions on a 6 DOF large Lagrangian solid body). The five separate sets of coordinate axes shown in Fig. 7.40 at the static horizontal equilibrium position of the semi immersed, double cylinder pontoon system under the action of a steady current only are required in order to analyze the dynamics of this relatively simple 3 degrees-of-freedom
Deterministic Dynamics of Small Solid Bodies
575
(MDOF) oscillating system. The first axes set is an inertial coordinate system x — {x,y,z} that is located at time t = 0 at the attachment point "c" of the two taut mooring lines with the vertical z axis positive up and the x-y axes in the horizontal plane at the SWL at z = 0. The second axes set is a body-fixed, non-inertial, moving coordinate systemic' = {x',y',z'} that is also located at time t — 0 at the attachment point "c" of the two taut mooring lines with the vertical z' axis positive up and the x'-y' axes in the horizontal plane at the SWL at z' = 0. The third axes set is also a body-fixed, non-inertial, moving coordinate system*^ = WR,y'R,z'R} for the point of angular rotation and is also located at the attachment point "c" of the two taut mooring lines with the vertical z'R axis positive up and the x'R-y'R axes in the horizontal plane at the SWL at z'R = 0 at time t = 0. Both of these latter two body-fixed, non-inertial, moving coordinate systems x' = {x',y',z'} and x'~ = [x'R,yR,zR} translate in linear momentum, but neither the point of translation x' = {x' = 0, y' = 0, z' = 0} nor the point of rotation xR = {XR = 0, y« = 0, ZR = 0} will also translate under angular rotations. The contributions to translational momenta from angular momenta is generalized in Chapter 8 to allow for additional translational displacements due to angular rotations when the origin of the translational axes x' = {x' = 0,y' = 0, z' = 0} and the origin of the rotational axes XR = {XR = 0, yR = 0, ZR = 0} do not coincide at t = 0. However, for this relatively simple two-dimensional planar 3 degrees-of-freedom oscillator, no additional translations due to angular rotations are required because the point of translation located at x' — [x' = 0, y' — 0, z' = 0} is identical to the point of rotation located at xR = {x'R, y'R, z'R); and, consequently, the vector r' between the point of translation x' and the point of rotation x'R is identically equal to zero; i.e., r' = x' — xR = 0. Finally, the fourth and fifth axes sets are also non-inertial, body-fixed, moving coordinate systems that are located at the center "a" of the large cylinder^ = [x'L, y'L,z'L] and at the center "b" of the small cylinder x's = {x's, y's, z's], respectively, with the vertical z/ axes positive up and the Xj'-y/ axes (j = L or S) in the horizontal plane at the SWL at z/ = 0 at time / = 0 for both coordinate systems. These latter two axes sets are required in order to compute the polar mass moments of inertia for each of the two cylinders by the parallel axis theorem.
576
Waves and Wave Forces on Coastal and Ocean Structures
The stiffnesses of each of the two mooring lines Kj in the vertical (j = z = 3) and in the horizontal (j = x = 2) modes are
K2 K3
sin2 8^ (2AE -P) cos2 8t
+P
(7.117a) (7.117b)
where 8i = the angle between each taut mooring line and the vertical z axis (vide., Fig. 7.41); P = the magnitude of the tension that is equal in each mooring line; I = total length of each taut mooring line; E = linear modulus of elasticity of the mooring lines; and A = cross sectional area of each mooring line. The large cylinder "a" weighs WL with a diameter Di; and the small cylinder "b" weighs Ws with a diameter D$.
Static position with only a steady current Uc The static horizontal equilibrium position of the semi-immersed, tautlymoored double cylinder pontoon system under the action of a normally incident steady current only is shown in Fig. 7.41. The origin of the inertial Cartesian coordinate system x — [x — y = z = 0} is located at the taut mooring line attachment point "c" with the x-y axes horizontal along the SWL at z — 0 at the static equilibrium position shown at time t = 0. The two mooring lines are attached to the rigid connector "a-c-b" at point "c" that may be computed from the requirement that the rigid connector "a-c-b" be horizontal at the SWL at
Fig. 7.41. Static free-body-diagram for only a normally incident steady current Uc.
577
Deterministic Dynamics of Small Solid Bodies
static equilibrium by summing the static moments about "c" according to
= (X - x0)Ws - x0WL - (x - XQ)FBS + XQFBL
(%)M?K
(7.118a)
where x and xo are defined in Fig. 7.39. Solving for a dimensionless distance xo/x f° r t n e attachment point "c" gives *o
<«w*>-(£)>-(?)
F
£lL
—.
(7.118b)
X WLFBL + Ws - FBs A numerical value for the circular cylinder "mass densities p" may be defined by 4Wi (7.119a) for j = L, S, Pi TtgD)B for the Archimedes buoyant forces FBj by jtD2- B FBJ=P8—^~
for j = L,S,
(7.119b)
and for the drag forces FQ for a normally incident steady current by FCj=CdjjDjBU?
for j = L,S,
(7.119c)
where C^. = drag coefficient for a normally incident steady current and may have different values for each cylinder because of their parametric dependencies (vide., Sec. 7.6). Substituting Eqs. (7.119a-c) into Eqs. (7.118a, b) gives xo
«*-'>(&)'-%(%)('+% (a?
X
(25 L - 1) + (2SS - 1) ( g )
(7.120)
2
where Sj = Pjg/pg = Yj/y = numerical value for the "specific gravity" of the j'th cylinder. For the special cases when In
\gxJ
CdL \DL)
« (25 s
(2SS - 1) * (2SL - 1),
1)
m
(7.121a) (7.121b)
578
Waves and Wave Forces on Coastal and Ocean Structures
then a very simple approximation for Eq. (7.120) is given by (Ds/DL)2
XQ
-r. (7.122) X 1 + (DS/DL)2 The equal tensions P in each of the two taut mooring lines may be computed from the following static equilibrium equations: Yl F* = FCL + Fcs ~ IP sin h = 0, (7.123a) Y^ Fz = - WL + FBL - Ws + FBs - IP cos h = 0,
(7.123b)
that may be solved for P and 8 values given by
(l-2SL)
7iD2LB
(£)'<'"
+
Pg
+
2CdL ( U, Uc\(l gDL)\
•
2SS)
CdsDsV CdLDL)_
*L (Ui U i | C"SDS\ n \gDL) y ^ cdL DL)
(7.123c)
2C
hi = arctan
(7.123d)
(l-25 L ) + ( g ) (1-2S 5 ) For the special cases when 5s
C
ds
(7.123e,f)
C
dL
then Eqs. (7.123c, d) reduce to P P8—§
TTDIB
-i2
(1 - 2SL)
2 QL
+L
it
8i — arctan
I
1+
U
C
\gDLJ
El DS
1+ V DL
(7.123g)
(7.123h)
a-25 L )(i+(gy
579
Deterministic Dynamics of Small Solid Bodies
Dynamic response an The wave-induced water particle velocity q(x,z,t,ao) d dynamic pressure po(x,z,t,ao) may be computed from a scalar velocity potential $>(x,z,t,cto)by q(x,z,t,ao)
= -VQ>(x,z,t,ao),
po{x,z,t,ao)
= p
d$>(x,z,t,ao) dt
(7.124a,b) where ao = an arbitrary phase angle for the free surface profile r](x,t,ao). A solution to the linear wave BVP in Chapter 4.3 is given by ®(x,z,t,ao)
= [ -^- ) V(kz)sm(a>t - a 0 - kx),
(7.125a)
where
V(kz)
exp(fcoz), coshpUi(l + f)l ooshkh
'
k0h>7T
(7.125b)
k0h
( 7 - 125c >
provided that co2h g
\koh, = koh = \ ikhtmhkh,
koh > TZ
(7.125d) (7.125e)
koh < n
The free surface displacement r)(x,t,ao) for linear waves that are collinear with the steady current normally incident to the pontoon system as shown in Fig. 7.39 may be determined from t](x,t,ao) =
1 d$>(x,0,t,a0) , g dt
z= 0
(H\ = 1 — 1 cos(
(7.126a)
580
Waves and Wave Forces on Coastal and Ocean Structures
the horizontal [vertical] water particle velocity u(x, z, t, ao) [w(x, z, t, ao)] by 9<J>(x, z, t, ao) u(x,z, t,ao) dx 9$(x, z, t, ao) w(x,z,t,ao) dz (Hgk\\ * (kz) cos(a>t — ao — kx) ^kz(kz) sm(cot — ao — kx) \2co (%)*(kz)ri(x,t,ao) {0)^kz{kz)m{x,t,ao)
"I-
(7.126b) (7.126c)
|'
UM(kz) cos(cot - a 0 - kx) 1 -Wntikz) sin(cot - a 0 - kx) J '
(7.126d) (7.126e)
where the amplitudes Uniikz) [W^fe)] of the horizontal [vertical] velocities are ,'Hco\/kh\ /Ho) v UM(kz) = — — *(**)> W (*z) = M ~' 2 y \kohJ ~" '»^ \ 2 J \kohJ (7.126f,g) and the local horizontal [vertical] water particle acceleration « ; (x,z,/,ao) [w f (x,z,?,ao)] are ut(x,z,t,ao)\ wt(x,z,t,a0)\
_ ~
(Hsk\\ y(kz)sin(cot-ao-kx) V 2 / {*a(^)cos(&)r - ao (%)*(kz)rit(x,t,ao) (^)^kz(kz)mt(x,t,a0)
{
UM(kz)co sin(a>? — ao — kx) WM(kz)(ocos{a>t — ao — kx)
1 kx)\ (7.126h) (7.126i) (7.126J) (7.126k)
where the dimensionless derivative ^kz (kz) in Eq. (7.16c, g and i) is given by (7.1261) exp(koz), koh > jr *kz(kz) = sinh[kh(l + (z/h))] (7.126m) koh < 7i cosh kh and where the subscripts t and kz denote partial derivatives of Eulerian fields; viz., 3W 9(«) (•)t (•)kz = 9? ' 9*z'
581
Deterministic Dynamics of Small Solid Bodies
Fig. 7.42. Sway mode Xj 0) free-body-diagram for coUinear, normally incident steady current and waves.
Sway mode Xi(i) For coUinear and normally incident steady current and waves shown in Fig. 7.42 the horizontal sway mode response X2U) may be computed from WL
1+
DsY SL
d2X2(t) dt2
DL)
= - 2 ^ 2 * 2 ( 0 + ^2 lxitZi,t,ao,
2
,
I,
(7.127)
where F2(xj,Zj,t,ao,d2X2(t)/dt2,dX2(t)/dt) = wave-induced hydrodynamic sway force load computed by the linearized WFE. The horizontal sway stiffness coefficient K2 is computed from the two taut mooring line stiffnesses Kx — K2 in Eqs. (7.117a, b). For normally incident waves, the wave-induced hydrodynamic sway force load may be computed from the linearized WFE (relative motion Morison equation with a linearized drag force) applied separately to each cylinder according to F2(xj,Zj,t,ao,d2X2(t)/dt2,dX2(t)/dt) = F2L (XL,ZL, t, a0, d2X2(t)/dt2, 2
2
+ F2S (xs, zs, t, a, d X2(t)/dt , L
=E
(C2mj
dX2{t)/dt) dX2(t)/dt)
DP ( K~T-) \utj(xj,zj,t,a0)
+ C2dijP ( ^ ) [uj(xj,zj,t,ao) (nD2.B\
+ P I —f— I utj
(XJ,zj,t,ao)
-
d
^P]
^ f ]
(7.128)
582
Waves and Wave Forces on Coastal and Ocean Structures
where C2mj {C2dij) = inertia coefficient (linearized drag coefficient that has the dimensions of velocity; vide., Eq. (7.60) in Sec. 7.6) for the y'th cylinder. The force coefficients are assumed to be different for each mode of oscillation and cylinder size because of their parametric dependencies (vide., Sec. 7.6) and interference or sheltering effects. Substituting Eqs. (7.126d,f and j) into Eq. (7.128) with zj = -Dj/4; approximating that V(-kDj/4) ^ 1 in Eqs. (7.125b or c); and separating terms that are proportional to only the pontoon motion from terms that are proportional to only the fluid motion gives
M
d2X2(t)
c
dX2(t)
> - ^ r - + > — + »«**« = F2L(t,x,xo,a0) = 3r2E(x,xo)cos(cot = ^(X^xo)
F2Es(t,x,x0,a0)
+
-ao + fi2)
[cos#2 cos(<*>/ — ao) — sin /% sm(cot — ao)] >
(7.129)
where K2 for each mooring line is defined in Eq. (7.117a) and where F2E(t,x,xo,ao) = the wave-induced, exciting sway force load on the fixed pontoon system that is a linear combination of the sway mode force load on the large cylinder F2E (t, x, xo, ceo) and the sway mode force load on the small cylinder F2E (t,x,xo,ao)- Scaling Eq. (7.129) by the following dimensional variables (tildes ~ denote dimensionless variables except for dimensionless time r and damping ratio £2):
(Ot == x,
dX2(t) dt
(Hco\ dX2(r) ~\ 2 ) dt '
&--= -^-,
c2l
(7.130a,b)
X2(t) = 1 — 1 X2(z), d2X2(t) dt2
C2cr = 2y/2K2M2;
{—)-*?-'
(7 13 C d)
- °'
(7.130e,f)
Deterministic
/ ^ / i v2
~2
?St2 ^Jf 2
=
583
Dynamics of Small Solid Bodies
(f)
2^2
#2
M2
M 2
a>2d = [ — ) = J l - & a>2fc
(7.130g,h)
J-2
(7.130i)
(^2//)'
where the natural frequency with hydrodynamic added mass a>2h is defined in Eqs. (7.112f-i) and the damped natural frequency a>2d is defined in Eqs. (7.114g-k). For ZJ = -Dj/4 and approximating ^(-kDt/4) « 1 in E Eqs. (7.125b or c), the dimensionless magnitude !F2 and phase angle fij of the normally incident wave induced hydrodynamic sway exciting force load are ^ (x.*o) =
y 2 £ (x.*o)
p 0*0 (¥)(&)
C 2 m L cosfcto + C 2 m s ( c f ' ) COS&(/ - x0) +
(*?) (A) (*W \CuL Sinkx0 ~ C2ds(£t) ™Hx - *0))_
(7.130J)
C2„,L sin kx0 - C2ms ( jfc ) sin *(x - * 0 )
( $ ) (A) ( W (CMt C°S^° + C2* (^f) C°S*(X - *o)). C2mL coskx0 + C2ms y^-j
c
/S2 — arctan
sin
cosk(x
-x0)
CM
+ G&) ( 4 : ) ( $ 0 ( ^ **° - S ( I £ ) sin*(x - *<»)_ C 2 m L sintoo- C2ms (75^) sin*(x - *o) " ( $ ) ( ^ l ) ( $ ) (C2"L
cos
^ 0 + C 2 , s ( £ * ) cos k(X - x0))
(7.130k) A dimensionless total cylinder and added mass M2 may be computed from M2 =
M2 pit DIB
2SL
+ (Clm, - 1) + (C2ims
SL
\DL)
1) El
(7.1301)
DL
where Sj — Pjg/pg = Yj/Y = numerical value forthe "specific gravity'' of the y'th cylinder. A dimensionless hydrodynamic viscous damping coefficient C2
584
Waves and Wave Forces on Coastal and Ocean Structures
may be computed from
C2 =
C2
p (H/2T) (4BDL/3)
(kh/k0h)
= CMi(l + g ^ ) .
(7.130m)
The dimensionless horizontal sway mode equation of motion is now given by (cf, Eq. (2.99a) in Chapter 2.5.3)
^+2^2*^ + ^
X2(r) - 8st26>2h COS(T - a 0 + &).
(7.131)
Applying the method of undetermined coefficients in Chapter 2.5.6, the dimensionless sway mode displacement X 2 (r) may be assumed to be given by X 2 (r)
= \h\ COS(T - ao +
(7.132a)
a 2 ),
where the dimensionless sway amplitude |f2| and dynamic magnification factor T>[(a)2h)~2] are 1/2 2
\h\ = D[(& 2 / ! )- ]^ 2 ,
2
T)[(w2hr ]
=
1 - (2A)~
-I
\(02h/
(7.132b,c) and the sway mode phase angle a 2 is tan ^2 — tan e2 _ 1 + tan €2 tan y$2 (2£2/<w2/,)
a 2 = arctan tane 2 =
- 22]i • [1 - {&2h)-
(7.132d) (7.132e)
The dynamic magnification factor D [(w2/!) 2 J and phase angle e are shown in Fig. 7.38 in Sec. 7.8.1.
Deterministic Dynamics of Small Solid Bodies
585
Fig. 7.43. Heave mode XT, (t) free-body-diagram for collinear, normally incident steady current and waves.
Heave mode X3 (t) For normally incident and collinear steady current and waves shown in Fig. 7.43, the vertical heave mode response X3(t) may be computed from WL
g
1+
mm
d2X3(t) dt2
V ^ - L p f
TMAt
^*3(0
= -K3X3O) + F 3 lxj,-Di/4,t,a,
dX3(t)\
,—
I,
(7.133)
where F-i(xj,—Dj/A,t,ao,d2X3(t)/dt2,dXi,{t)/dt) = wave-induced hydrodynamic heave force load computed by the linearized WFE. The vertical heave stiffness coefficient K3 is a linear combination of the water plane stiffAi nesses K3 from each cylinder and of the taut mooring line stiffness in each of the two taut mooring lines K3 from Eq. (7.117b) that may be expressed as AT3 = K^L + K*s + 2K3
= pgBDL(l
+
^\+2K3
(7.134a)
and a dimensionless vertical heave stiffness coefficient £3
K3 =
*3
2K3
i+
fpgBDL(i + %y 2/T 3
(7.134b)
When the total dimensional stiffness from both of the two taut mooring lines 2K3 is large compared to the water plane area or, equivalently, when &i «» 0
586
Waves and Wave Forces on Coastal and Ocean Structures
(that is frequently true for small pontoon bodies), then KT, « 2K$. For normally incident waves, the wave-induced hydrodynamic heave force load may be computed from the linearized WFE (relative motion Morison equation with a linearized drag force) applied separately to each cylinder according to F2(xj,-Dj/4,t,a0,d2X3(t)/dt2,dX^t)/dt) =
\F2L(xL,-DL/4,t,a0,d2X3(t)/dt2,dX3(t)/dt) 1 1 +F.35 (xs, -Ds/4, t, a0, d2Xj (t)/dt2, dX3 (t)/dt) J L
(Csmj - DP (xD2B/s)
-E
[wtj(xj,Zj,t,a0)-d2X3(t)/dt2]
+ C3dijP {DjB/2) [wj(xj,zj,t,a0) + p (nD2B/8)
-
dX3(t)/dt]
wtj (XJ, ZJ, t,a0) (7.135)
where Csmj {C^dij) = inertia coefficient (linearized drag coefficient) that has the dimensions of velocity; vide., Eq. (7.60) in Sec. 7.6) for the jth cylinder. The force coefficients are assumed to be different for each mode of oscillation and cylinder size because of their parametric dependencies (vide., Sec. 7.6) and interference or sheltering effects. Substituting Eqs. (7.126e, g and k) into Eq. (7.135) with ZJ = -Dj/4; approximating that Wkz(-kDj/4) « 1 in Eq. (7.1261) or «s tanhkh in Eq. (7.126m); and separating terms that are proportional to only the pontoon motions from terms that are proportional to only the fluid motion gives d2X3(t) =
dXi(t)
Ff(t,X,xo,ao)
= PfL (t, X, XQ, a 0 ) + Fis (f> *' x ° ' a ° ) = £3 £ (x,*o) cos(cot - ao - £3) = ^E(X> xo)[sin /% sm(cot — ao) + cos $3 cos(a>t — ao)],
(7.136)
where F3£ (?, x, *o, «o) = the wave-induced, exciting heave force load on the fixed pontoon system that is a linear combination of the heave mode force load on the large cylinder FfL(t,x,xo,<xo) and the heave mode force load on the small cylinder Ffs(t, X,XO,UQ). Scaling Eq. (7.136) by the following
587
Deterministic Dynamics of Small Solid Bodies
dimensional variables (tildes ~ denote dimensionless variables except for dimensionless time x and damping ratio £3): cot = x,
d2X3(t) dt2
Hco\ dX3(x) dx
dX3(t) dt
C
3 fc C , C3cr = 3cr
~2
(7.137a,b)
X3(t) = ( y ) X3(r),
d2X3(x) -, (7.137c,d) dx2 (7.137e,f)
2^/K3M3,
/ ^3/1
K3
K3
1
M3
M3W
<w /
(Hco2\
,
<^3rf\
co3d = [ —^- ) = ^ / l - f32, <W3ft/
(7.137g,h) Stf3 =
£ 5^ •^3
°,S*3
(H/2)
(7.137i)
(K3H)
where the natural frequency with hydrodynamic added mass u>2h is defined in Eqs. (7.112f-i) and the damped natural frequency coja is defined in Eqs. (7.114g-k). For zi = - A / 4 and assuming that tykz(-kDi/4) % 1 in Eq. (7.1261) or « tanhfc/r in Eq. (7.126m), the dimensionless magnitude !F^ and phase angle /% of the normally incident wave induced hydrodynamic exciting heave load are y3£(x,*o)
rf(x,*o)
Hco2] C3mLsmkx0-Cims(%l) C
- {&) (A) \k%) { ^L
cos
smk(x-x0)
^0 + C3ds (£*) cosk(x - , 0 )) .
C 3 m L coskx0 - Cims ( - ^ j cosfe(x - *o)
-
+
( $ ) {£) ( $ ) ( C ^ Sinfa0 + C3ds (§*) sin<:(x - *„)) (7.137J) - C 3 m ^ sintao - C3m^ (-p^-j sinfe(x - *o) C Cosfa C3rf cos
/J3 = arctan
' (fa?) ( # ) (ffi) ( ^
°+
5 ( jfc) *(* ~ *o))
C 3 m i coskx0 - C 3 m s ( - ^ - j cosA(x - *o)
• ( $ ) (BT) ( $ 0 ( C ^ s i ^ 0 - C3,s (£*) sin*(x - *„)) (7.137k)
588
Waves and Wave Forces on Coastal and Ocean Structures
A dimensionless total cylinder and added mass M 3 may be computed from M3 =
M3
2SL
(PTTD2LB/S)
SL
+ (C3mL
\DL)
1) + (C3,ms
1)
:
m
,
(7.1371)
where Sj — Pjg/pg = Yj/y = numerical value for the "specific gravity" of the jth cylinder. A dimensionless hydrodynamic viscous damping coefficient C3 may be computed from C3 =
C3
p (H/2T) (&BDL/3) (kh/k0h)
= CM,
1 +
(7.137m) CM,
DL
The dimensionless vertical heave mode equation of motion is now given by (cf, (2.99a) in Chapter 2.5.3)
H?
+2 h
^ Tx
+
3h X3(r) = 8st3 h cos(T
'
^
~ a°+ ^3)'
(7 138)
'
Applying the method of undetermined coefficients in Chapter 2.5.6, the dimensionless heave mode displacement X 3 (T) is assumed to be given by *3 (T) = \h I COS(T - a0 +
(7.139a)
a?),
where the dimensionless heave amplitude |f3| and dynamic magnification factor D [(<53/;)~2] are 1/2
2
2
If31 = D [(^h)- ] ht3, D [(u3h)- ] =
1 - (mhY
\mhj
J
(7.139b,c) and the heave mode phase angle a 3 is tan yS3 — tan e3 a 3 = arctan _ 1 + tan e tan /3 _ 3 3
tane 3 =
?
^.
[i-{mh)-2]
(7.139d,e)
589
Deterministic Dynamics of Small Solid Bodies
The dynamic magnification factor D [(w^) in Fig. 7.38 in Sec. 7.8.1.
2
] and phase angle 63 are shown
Roll mode &s(t) about the mooring line attachment point "c" Because the point of translation "c" for sway and heave motions is the same as the point of rotation "c" for angular roll modes, the translational dynamics are simplified significantly (vide., Chapter 8 for applications where this simplification is not valid). For normally incident and collinear steady current and waves, a relatively simple approximation for the roll mode response ®s(t) about point "c" may be computed from the moments induced by the horizontal, relative motion sway mode load F2(xj,—Di/4,t,ao, d2X2 (t)/dt2, dX2(t)/dt) and by the vertical, relative motion heave mode load 2 2 F^{XJ, 0, t, oto,d Xi(t)/dt , dX^(t)/dt). In this approximation, however, the moment load may be reduced to only a moment due to the exciting sway mode load on a fixed pontoon system F^ (t, x, xo, ao) because the horizontal, relative motion of each cylinder Sj(t) may be shown to be of second order (i.e., st(t) a @5(0> where Sj(t) is defined in Fig. 7.45), and therefore, may be neglected for small amplitude motions. However, this will not be the case for the moment from the vertical, relative motion heave mode force load. For normally incident and collinear steady current and waves, the roll mode response @5(0 about "c" may be computed from dzG5(t) ' dt2
= -K5@5(t) + M5
__^2,_±2),
(7 ,40,
where M5(xj,Zj,t,ao,d2&5(t)/dt2,d@5(t)/dt) = wave-induced, relative motion hydrodynamic moment load on the jth pontoon that may be computed by the linearized WFE. This moment load is linear combination of the waveinduced, exciting sway force load on thefixedpontoon system F^ (t, x, xo, ao) from Eq. (7.129) and from the vertical, relative motion heave mode force Ff(t,x,xo,ao) fromEq. (7.136). The total polar mass moment of inertia 15 = a linear sum of the polar mass moments of inertia for each cylinder Im (j = L or S) that may be computed by the parallel axis theorem about the horizontal y-axis through the mooring
590
Waves and Wave Forces on Coastal and Ocean Structures
point "c" in the x-z plane according to
h = iic+A ( Y)
+7
& + {x~Xof ( y ) '
(7 141a)
-
where IJ^ = polar mass moment about the center of gravity "C.G." in the x'j-z'j plane of the ;'th cylinder. If each of the 7'th cylinders is an homogeneous circular cylinder of diameter Dj and wall thickness dj, then IJ^ may be computed from
"&>m-<®
(djD]B'
(7.141b)
A solid homogeneous circular cylinder may be computed from Eq. (7.141b) by substituting dj = Dj/2. Because the pontoon system is rolling about the point "c" that is the point of attachment of the two taut mooring lines, the roll stiffness coefficient K5 is not a function of the taut mooring line stiffnesses but is solely a function of the moments from the vertical heave mode water plane stiffnesses K^1 of each cylinder and may be expressed as K5S5(t) = x0K*LZL(t)
+ (X-
x0)K*sZs(t),
(7.142a) A •
where the vertical heave mode water plane stiffnesses K^J of each cylinder j = L or S are given in Eq. (7.134a). Because xo ^ x/2, the vertical displacements Zj (?) for each cylinder are not equal. The unequal vertical displacements Zj(t) for each cylinder j = L or S may be converted to angular roll rotations ®s(t) from the geometry shown in Fig. 7.44 (cf., Sec. 7.8.1) according to sin 0 5 (O =
7
^
L
Xn
= - ^ Y — Xn
« 6 5 ( 0 + O (&l(t)) . V
(7.142b,c)
/
Substituting Eqs. (7.142b, c) into Eq. (7.142a) yields K5@5(t) = [X%K}L
+ (x - xo)2 K*s] 0 5 (O.
(7.143)
For normally incident and collinear steady current and waves, the waveinduced, relative motionroll mode moment load M5 (xy-, ZJ , f, an, d2@s (t)/dt2,
591
Deterministic Dynamics of Small Solid Bodies
Fig. 7.44. Heave-roll modes geometry.
d@5(t)/dt) may be derived from a linear combination of the moments about "c" that are induced by the horizontal, relative motion sway mode force load F2(xj,-Dj/4,t,ao,d2X2(t)/dt2,dX2(t)/dt) that is applied at the location Zj = —Dj/4 and by the vertical, relative motion heave mode force load F3 (XJ, 0, t, <XQ, d2X-$(t)/dt2, dX3(t)/dt) that is applied at z = 0 according to M5(xj,zj,t,a0,d2®5(t)/dt2,d@5(t)/dt) =
M52 (xj,-Dj/4,t,a0,d2&5(t)/dt2,d@5(t))} + M53(xj,0,t,ao,d2@5(t)/dt2,de5(t)) j (-xo,^,t,ao,d2X2(t)/dt2,dX2(t)/dt^
~^F2L ~^F2s
[x
-xQ,^,t,aQ,d2X2{t)/dt2,dX2{t)/dt)
, ( xoF3L(-x0,O,t,ao,d2X3(t)/dt2,dX3(t)/dt) + {-(X-xo)F3s(x-xo,0,t,ao,d2X3(t)/dt2>dX3(t)/dt)l
1 K
'
'
The relative motion sway-roll mode moment MS2(XJ,—DJ/4, t, ao, d2&5(t)/dt2, d@5(t)/dt) about the point of rotation "c" from the horizontal sway loads in Fig. 7.42 is a negative moment (counterclockwise); and the relative motion heave-roll mode moment load Ms3(xi,0,t,ao, d2@${t)/dt2,d®s{t)/dt) about the point of rotation "c" on both the small and large cylinders from the vertical heave loads in Fig. 7.44 are positive moments. In order to demonstrate that the heave-roll mode moments on both cylinders in Fig. 7.44 are positive, the moments due to the relative velocities
592
Waves and Wave Forces on Coastal and Ocean Structures
and accelerations may be computed from
I
dZL(t)
(d2ZL(t)\)
dt \ dt2 ) I = wL(t)(wtL(t)) \
-xo
0 0
0 tdZiXt) \ dt
(d2ZL{t)\ dt2 )
\
{ wL(t)(wtL(t)) rfZifr) /rf2zL(p^
dt2
dt
= +XQ
nh.
wL(t)(wtL(t))
rs x
dZs(t) dt
w
ws(t)(wtJt))
ex (X-xo) 0
J
0
ez 0
0
' dZs{t) dt \
(d2Zs(t)\ dt2 )
ws{t){wtJt)) 2 • dZs(t) (dI*Z2 s(t)\\
+ (x-xo)>
\ dt2 ) 5V ws(t)(wts(t)) J dt
The relative motion sway-roll mode moment load d2@5(t)/dt2,d@5(t)/dt) is
Ms2(xi,—Di/4,t,ao,
M52{xj,-Dj/A,t,a(i,d2®s{t)/dt2,d®s{t)/dt\
A
(7.145)
The relative motion sway-roll mode moment Af52(JC,-, — D,/4,?,ao, d2&5(t)/dt2, d@5(t)/dt) in Eq. (7.145) due to the horizontal, relative motion sway mode force load on each y'th cylinder F2(xj,0,t,ao,d2Sj(t)/dt2, dsj{t)/dt) in Eq. (7.128) may be approximated by a moment load that is due only to the sway mode exciting force load on a fixed pontoon system F2~(t,x,xo,(Xo) from Eq. (7.129) and Eqs. (7.130J, k) by an analysis of the geometry illustrated in Fig. 7.45. Similar to the heave-roll mode geometry illustrated in Fig. 7.44, the horizontal displacements Sj(t) for each cylinder
593
Deterministic Dynamics of Small Solid Bodies
z,{t),
S+Jz
*i (0 . ©5 (<) \*M5
(t) *,»
Fig. 7.45. Sway-heave-roll modes geometry.
j = L or 5 in Fig. 7.45 may be converted to the angular roll rotation ©5(0 from the geometry shown in Fig. 7.45. For the large cylinder DL, this approximation is sL(t) w
1
=cos©5(0~l
XQ
sL(t) XQ
©2(0 x~ +
2! ©2(0 2!
(7.146a)
and for the small cylinder D$ 1 -
ssit) X
= COS © 5 ( 0 ^ 1
-XQ
ssit) X -*o
©jKo +
2! ®i(t) 2!
•••
(7.146b)
so that
«(.> = (f)e3w. ^ ^ = (?) rfei<0 d? d2sL(t) dt2 ss(.t) =
=(?)
d2Q2(t) 5 Jr 2
2
(7.147c) X-x
® 5 (0,
<*2*s(0 J?
(x0^ 2/
X - x o \ c / 2 0 2»(5
(7.147a,b)
0
\ d©|(0
(7.147d,e) (7.147f)
Waves and Wave Forces on Coastal and Ocean Structures
594
Because of the small amplitude motion assumption, higher order terms of 0{(d\(t)) are neglected. Consequently, the relative motion sway mode moment load in Eq. (7.145) reduces to only the sway-roll mode exciting moment load on a fixed pontoon system Mf2(t, X>*o, «o) given by
Mf2(t,a0)
= M%2L(t,a0) + M%2s(t,ao) = Mf 2 (x,*o) sin(cot -a0-
fcl)
— Mf2(x>*o) [cos ^52 sin (art - ceo) - sin £52 cos(otf - «o)], (7.148)
where the dimensionless moment magnitude M^2(X> x o) is
Mf2(x,*o):
Mf2(x,*o)
*(*)(¥)(&) C2mL sinkx0 - C2m$ ( ^ - ) sink(x - x0)
+
_ (3)
( $ ) (C"L
(A)
COskx
° + C2ds(%t)2
C2mLcoskxo + C2ms(-^) • ( 3 ^ ) (A)
cosHx -xo)
cosk(x-xo)
( $ ) (C2
sm«X - *o))
(7.149a) and phase angle /S52 are
C2mL sin/jx0 - C2ms(j^ ,852 = arctan
sin*(x - x0)
• (£) (A) ( $ ) {C"L COS^0 + C™S (% f OP*** - *<>>) C2mLcoskx0 + C2msyifi) +
cosk(x-x0)
{&) (A) ( $ ) ( C ^ sin^0 - C2ds (ft f Sink(X - xo)) (7.149b)
595
Deterministic Dynamics of Small Solid Bodies
The relative motion heave-roll mode moment load in Eq. (7.144) is M5i(xi,0,t,a0,d2e5(t)/dt2,d@5(t)/dt)
' V^, -1) ("?**) (u,,,-*™) - L X ~xo)
+cus{Zl){pWM<*?)DsB){«>s
d
T) (7.150)
The relative motion heave-roll mode moment load in Eq. (7.150) that is due to the vertical, relative motion heave mode force load on each jth cylinder in Eq. (7.135) may not be simplified to a moment load that is due solely to the heave mode exciting force load on a fixed pontoon system F^(t,x,xo, ao). In addition, some care must be exercised with the relative motion moment load due to the vertical heave mode force load on the small cylinder because this moment load is negative and would result in a negative added mass and viscous damping coefficients if the temporal derivatives of the vertical displacement Zs(t) in Eq. (7.150) are not converted to the angular roll rotation ©5(?) correctly from the geometry shown in Fig. 7.44. For positive roll rotations ©5(0 (clockwise rotations) in Fig. 7.44, the vertical displacement of the large cylinder Zi(t) is positive (up); but the vertical displacement of the small cylinder Zs (t) is negative (down). However, both moments have already been demonstrated to be positive moments following Eq. (7.144). The temporal derivatives of the vertical displacements Zj(t) for each jth cylinder in Eq. (7.150) may be converted to angular roll rotations ©5(f) from the geometry shown in Fig. 7.44 by Eqs. (7.142b, c). Substituting Eqs. (7.126e-g) into Eq. (7.150) with ZJ = 0; approximating that tytz(0) = 1 in Eq. (7.1261) or = tanh kh in Eq. (7.126m); and then separating terms that are proportional to only the pontoon motions from terms that are proportional to only the fluid
596
Waves and Wave Forces on Coastal and Ocean
Structures
motions gives for the heave-roll mode exciting moment load M^3(t, x,x0,a0)
= M^L(t,x,x0,ao)
-
Ms3s(t,x,x0,ao)
= Mf 3 (x,*o) sin(cot -ao- £53) = Mf3(x,xo)[cos^53 sin(cot - ao) - sin £53 cos(tot - ao)], (7.151) where the dimensionless moment magnitude M^3 (x, XQ) and phase angle ^53 of the normally incident wave induced hydrodynamic exciting heave-roll moment load are Mf30(,xo) =
M^3(x,-*o) 1
( I 4 £ ) ff^'o)
+(&) (A) K •**» + C 3 4t) ( ^ ) sin*(* -xo)) +
- ( $ ) (&) K cos **° - C ^(^) (^r)cos*(* ^ 0) )
(7.152a)
C 3 m L c o s ^ 0 - C 3 m s ( ^ f ) ( 2 ^ f i ) c o s t ( x -XQ) C sinfa C
%) {A) ( ^L
/S53 = arctan
0 + Us (jfc) ( ^ ) ™*<X ~ XQ))
CimL sinkx0 - C3ms ( ^ - j ( ^ ^ T ) " & ) (A)
K
COS
sin
* ( * ~ *<>)
^0 - C3rf5 ( £ * ) ( * i f ) eos.(x - ,0)) J I
(7.152b) The wave-induced roll mode exciting moment on a fixed pontoon system Mf(t, x,XQ,a(j) may now be written as a linear combination of Eqs. (7.148 and 7.151) as M5 =
(t,x,xo,a0) M
? 5^( ' X, XQ, ao) + M5£3 (r, x, ^0, «o) 52 (t,X,x<
= (Mf 2 (x,*o)cos/352+ Mf 3 (x,x 0 )cos^53)sin(tt.? - ao) - (Mf 2 (x,x 0 )sin y 6 52 + Mf 3 (x,x 0 ) siny653) cos(a.f - a 0 ) = Mf (x,*o) sin(&;? - ao - ft) = Mf (x,xo)[cosft sin(ft>? — ao) — sinft cos(atf - ao)],
(7.153a)
597
Deterministic Dynamics of Small Solid Bodies
where
Mf(x,*0) =
1/2
[Mf 2 (x,xo)cos^52 + Mf3(x,xo)cosj653]2 + [Mf 2 (/,x 0 ) sin ^52 + Mf 3 (x,x 0 ) s i n ^53] 2
(7.153b) l$5 — arctan
[Mf2(x,x0)sinj652 + Mf3(x,A:o)sin;653] [Mf 2 (x,*o) c o s f e + Mf 3 (x,x 0 ) cos £53] (7.153c)
Separating terms in the relative motion roll mode moment in Eq. (7.144) that are proportional to only the pontoon roll mode rotations ©5 (t) from terms that are proportional to only the fluid kinematics transforms Eq. (7.140) to d2G5(t) dt2
d@5(t) dt
F
(7.154)
Because the pontoon system may be sited in deep-water conditions, scaling the roll mode rotations ®s(t) by the water depth h as was done for the articulated tower in Sec. 7.8.1 is not appropriate and, instead, will now be scaled by the following variables (tildes ~ denote dimensionless variables except for dimensionless time x and damping ratio £5): x = cot, © 5 ( T ) =
£L
d@5(x) dx
DL\d@s(t) Hco) dt
d2@5(x) dx2
'
C5 & = --
(7.155a,b)
©5(0,
H
DL \ d2@5(t) dt2 Hco1
(7.155c,d) (7.155e,f)
C5cr = 2Jic5l5,
*~-5cr ~2
(U>5h\2
K5
K5
h
hco'
A,(J =
Mf
«>5d
<5,
(7.155g,h)
\C05h/
(7.155i)
*5
Scaling Eq. (7.154) by the same scaling as Mf3 in Eq. (7.152a) (viz., p7tD2^BHco2xo/l6) gives for the dimensionless pontoon cylinder polar mass and added mass (from the heave-roll moment load Eq. (7.150)) moments of
598
Waves and Wave Forces on Coastal and Ocean Structures
inertia I5
h=
h (pnD\BxQ/16)
(pnDlLo/16) + 2(C3 ms
+ 2(C3mL
-»m
" 1} \lk (X -
XQ)2
(7.155J)
DLXQ
from Eqs. (7.141a and 7.150); for the dimensionless viscous damping coefficient C$ from the heave-roll moment load Eq. (7.150) C5 =
C5 (p7t2DlBx0/ST) 32 3n2
H
XQ
~D~L
~D~L
El
C^dr + C-ids I "=—
XQ
xo (7.155k)
for the dimensionless rotational stiffness £5 K5
(PTTDIBQ)2X0/16)
(7.1551)
and for the amplitude of the dimensionless roll mode exciting moment load M5£(X,*o) Mf(X.Jco)
Mf(x,x0) (p7TD2LBHco2xo/l6) [ M f 2 ( X , x o ) ( ^ ) ( g ) c o s ^ 2 + M 5 £ 3(X^o)cos^ 3 ] 2 + [ f i * ( X ,* 0 ) ( $ ) ( g ) s i n f e +M 5 £ 3 (x,xo)sin^ 3 ]^ (7.155m)
599
Deterministic Dynamics of Small Solid Bodies
The dimensionless roll mode equation of motion Eq. (7.154) is (cf, Eq. (2.99a) in Chapter 2.5.3) ^ 2 + 2^S)5h'd^
+ &
5h
0 5 ( T )=
*«A
sin T
( - "o - ft). (7.156)
Applying the method of undetermined coefficients in Chapter 2.5.6, the dimensionless roll mode rotation ©s(r) is assumed to be given by ©5(r) = |£25|sin(r - ao -as),
(7.157a)
where the dimensionless roll amplitude | Q51 and dynamic magnification factor D[(<S5/i)"2] are • 1/2 2
l«51 = D [ ( & 5 A ) " ] A , / 5 ^ ,
2
D[(5 5 *)" ] =
2 2
[1 - (&5hr ]
+, _
.
<»5h/
(7.157b,c) and roll mode phase angle 0:5 is tan Ps — tan €5 1 + tan 65 tan $5 _
,„ , ._ , . =-. (7.157d,e) [1 - ( ^ ) - 2 ] 2 The dynamic magnification factor D[(a)5n) ] and phase angle €5 are shown in Fig. 7.38 in Sec. 7.8.1. tanas =
(2ft/<W5ft)
tanes =
7.9. Transverse Forces on Bluff Solid Bodies The bluff solid body shown in Fig. 7.46 is subjected to three forces; viz., an in-line hydrodynamic force per unit length dfx(t) that may be due to either a steady flow Uc or an oscillatory flow u(t) or both; a transverse force per unit length dfj(t) that is due to viscous forces and eddy shedding and a resisting force per unit length dfR (t) that is the dynamic response of the body due to the alternate shedding of vortices. The fluid viscosity [i causes a velocity defect near the bluff body boundary that results in the flow in the boundary layer coming to rest and forming the separation points and a wake as illustrated in Fig. 7.46. In the wake region between the lines formed at the two separation points, vortices are formed and shed from the bluff solid body. In steady flow,
600
Waves and Wave Forces on Coastal and Ocean Structures
dfAt)
separation point
Uc,u(t)
JKy
'
point
Fig. 7.46. Definition sketch for the transverse force dfj, in-line hydrodynamic force dfx and resisting force dfn loading a bluff solid body.
these vortices are shed alternately from both sides of the body and create an oscillatory force transverse to the direction of the flow shown in Fig. 7.46 that is called a transverse force dfy (t).In unsteady flow, the same vortex formation due to flow separation in the boundary layer occurs; but now this phenomenon is much more complicated because the location of the wake now oscillates from the up-wave side of the body to the down-wave side over a wave period T as the wave-induced hydrodynamic pressure oscillates from the up-wave side of the body to the down-wave side of the body (vide., Chapter 4.5). A semiconstrained bluff body also induces additional dynamic pressures on the fluid as it responds dynamically to the vortex-shedding forces. The force that is due to the dynamic response of a bluff body to the vortex-induced forces is called a resisting force dfR(t). Sarpkaya (1979 and 1989) reviews in detail the computational methods available for analyzing vortices and vortex-induced motions. Keulegan and Carpenter (1958) identify the location of the oscillating separation points illustrated in Fig. 7.46 as a cause of additional difficulty in estimating wave-induced pressure forces on small bluff bodies. They introduce a physical interpretation for their periodparameter (or the Keulegan-Carpenter parameter K = Um T/D in Eq. (7.44c) in Sec. 7.6.1) that is based on the length of travel of a fluid particle past a small bluff body during one wave period T. Specifically, the length of travel of a fluid particle past a small bluff body during exactly one-half of a wave cycle T/2 provides great insight into the stability of eddy shedding in oscillatory flow of viscous fluids. A ratio between the wave period T and the duration of the shedding of a single eddy Ts may
601
Deterministic Dynamics of Small Solid Bodies
be computed from the following product between the Strouhal parameter St from Chapter 3.6 and the Keulegan-Carpenter or period parameter K:
When the ratio T/Ts = 2 in Eq. (7.158), only one single eddy will be formed and separated during exactly one-half of a wave period T. For values of the Strouhal parameter in the range 0.18 < St < 0.2 (Dean and Harleman, 1966) and for T/Ts = 2, then the Keulegan-Carpenter or period parameter K « 11. This value is a critical value that is associated with an eccentricity of the Dean error ellipse of unity (vide., Sec. 7.7.1) and implies that the wave force data are equally well-conditioned for identifying both of the force coefficients Cm and Cd- In the replotted Keulegan-Carpenter data shown in Fig. 7.26, the minimum value for the inertia force coefficient Cm occurs at a value of the Keulegan-Carpenter parameter K « 12.5 and only one single eddy was observed to form and be shed during the experiments (Keulegan and Carpenter, 1958, p. 439, Fig. 29b). The critical values of the Keulegan-Carpenter parameter between 11.40 < K < 13.16 may be connected to the condition of the data when the inertiald/ m and the drag dfd force components per unit length contribute equally to the mean square error between the measured force data and the forces computed by the Morison equation on small piles (vide., Sec. 7.7). Additional experimental data that support the existence of stable transverse forces for critical values of the Keulegan-Carpenter parameter between 11.40 < K < 13.16 are given by Hayashi and Takenouchi (1979) in Fig. 7.27 in Sec. 7.7.1 and by Maull and Milliner (1979) in Fig. 7.34 in Sec. 7.7.3. The physical interpretation that the Keulegan-Carpenter parameter K is a measure of the length of travel of a fluid particle past a small bluff body during one-half of a wave period T is apparently related to the wake effect that may explain the elevation of values of the drag coefficient Cd in the supercritical region where the Reynolds parameter R > 106 in Fig. 7.19 in Sec. 7.6.5 (Garrison, 1980 and 1982). Garrison (1980) defines a relative amplitude parameter by 2D
D'
(7.159)
602
Waves and Wave Forces on Coastal and Ocean Structures
where 2 A = H and that is proportional to the Keulegan-Carpenter parameter K for simple harmonic flows when UmT (Hn\ T H K=-^—^[—r\= —n=2nG. D \ TJD D
(7.160) '
v
Garrison (1980) demonstrates that the supercritical region where R > 106 may be separated into two effects: viz. 1) an effect due to pile roughness e/D and 2) an effect due to vortex wakes that correlate with the relative amplitude parameter G in Eq. (7.159) (vide., Sec. 7.6.5). In addition to the Keulegan-Carpenter experiments, transverse forces have been measured in two very different types of experiments. Hayashi and Takenouchi (1979 and 1985) measure transverse forces on vertical circular cylinders in progressive waves. Maull and Milliner (1979) measure transverse forces on a vertical circular cylinder forced to oscillate in still water. Hayashi and Takenouchi (1979) measure the transverse forces on fixed vertical circular cylinders in progressive laboratory wave as shown in Fig. 7.27 in Sec. 7.7.1. All of these transverse forces are all irregular except when the Keulegan-Carpenter parameter K= 11.8 (vide., Sec. 7.7 for critical values of K) where the transverse force is stable and periodic. In addition, their data records indicate that there are exactly two vortices shed over each wave period T. Hayashi and Takenouchi (1985) also reproduce stable two-vortex transverse forces for two values of the Keulegan-Carpenter parameter K — 10.5 and K = 11.2 in experiments with a 4 cm diameter cylinder at higher values of the Reynolds parameter R than the values that are illustrated in Fig. 7.27. Maull and Milliner (1979) measure the transverse forces on a vertical circular cylinder forced to oscillate in still water as shown in Fig. 7.34 in Sec. 7.7.3. The oscillograph traces for each run shown in Fig. 7.34 represent 15 continuous cycles so that each upper trace measured at each Keulegan-Carpenter parameter K represent 5 repetitions superimposed from each of the 3 cycle recordings. The one-vortex stable transverse force measured for K — 13.02 (vide., Sec. 7.7 for critical values of K) demonstrates remarkable stability. The conclusion from these experiments for transverse forces is that when the values for the Keulegan-Carpenter parameter K are in the critical range 11.40 < K < 13.16, then the transverse forces are stable and periodic. As
603
Deterministic Dynamics of Small Solid Bodies
a consequence of this periodicity, Fourier series approximations with deterministic force coefficients for the transverse forces on small bodies may be appropriate. For values of the Keulegan-Carpenter parameter K that are not in the critical range, the transverse forces are irregular and non-periodic. Consequently, only Fourier series transverse force models are reviewed. A transverse force per unit length dfy is given by dMO =
l
-CLpAu2{t),
(7.161)
where CL = dimensionless lift coefficient, p = fluid mass density, A = cross-sectional area of a body perpendicular to the in-line flow, and u(t) = a characteristic in-line flow velocity that is perpendicular to the direction of the transverse force (e.g., a maximum amplitude of a time-dependent in-line flow, an instantaneous value of a time-dependent in-line flow, a steady flow amplitude, etc.). An alternative to Eq. (7.161) for a transverse force per unit length assumes a velocity w(t) that is in-line with the direction of the transverse force and perpendicular to the direction of the in-line flow velocity u(t) in Eq. (7.161); viz., dMO = -CLpAw2(t).
(7.162)
The transverse force per unit length on a bluff body may be approximated by dMt)
= dMt)
- dMO,
(7-163)
where dfj{t) = a time-dependent transverse force per unit length, dft(t) = a time-dependent lift force per unit length and dfR(t) = a time-dependent resisting force per unit length. Chakrabarti (Chapter 6.3, 1987) proposes an alternative to the linear combination of a transverse and resisting forces in Eq. (7.163) that is a relative motion transverse force model given by dMO = \cLpA(.w{t)
- y(0)\w(t)
- y(0\,
(7.164)
where y(t) = transverse velocity of the oscillating bluff body in Fig. 7.46 and the overdot denotes ordinary temporal derivative of a Lagrangian solid body;
604
Waves and Wave Forces on Coastal and Ocean Structures
viz., y(t) = dy(t)/dt. Because the fluid velocity w(t) that is transverse to the wave direction must be known in order to apply Eq. (7.163), the model in Eq. (7.164) is difficult to apply with wave forces. Relative velocities could also be applied in Eq. (7.163). In unsteady flows, vortex shedding from bluff bodies becomes very complicated and the transverse force is irregular and random. Because of this random behavior, the transverse force per unit length may be expanded in a Fourier series (vide., Sec. 7.6.3 and Chapter 9.2) with frequency dependent lift coefficient Cm according to (Chakrabarti, 1987) / r ( 0 = ^ ^ " m a x J2 n
C t n C0S M
( nt
+ €n),
(7.165)
where umax = the amplitude of the periodic wave velocity that is perpendicular to the transverse force. A maximum Cz,max and a root-mean-square Cirms lift coefficients may be computed from the maximum fTmax and root-mean-square fTrms transverse forces, respectively, from Chmax =
•^jTmax • 7 ,
pAu*
„ CLTMS =
^JTr . 9
pAu{
•
(7.166a-b)
A model for a one-half cycle transverse force per unit length proposed by Chakrabarti (1987) is given by 1 T frit) = -C L pAw m a x cos — (1 -cosa>0 + * sin,2z o;/, 2'
(7.167)
where ^ = a phase angle that depends on the history of the transverse flow. Transverse forces per unit length on marine risers may be modeled with an amplification ratio according to dfrit) = l-CLPAu2mw, (~\
cos(cot - * ) .
(7.168)
Finally, a relatively simple transverse force per unit length that is periodic and that has been applied to tension leg platform tethers is given by dfT(t) = ^CLpAulmcos((ot).
(7.169)
605
Deterministic Dynamics of Small Solid Bodies
7.10. Stability of Marine Pipelines The algorithm derived by Huang and Hudspeth (1982) to determine the lateral static stability of bottom-laid pipelines under finite-amplitude deterministic waves is reviewed. The dynamic behavior of deep ocean bottom-laid pipelines is analyzed by Zimmerman, et al. (1986). An unsteady, wave-induced oscillatory flow u(x,y,z,t) combined with a steady, uniform current Uc exerts hydrodynamic pressure forces on a pipeline laid on a rigid, impermeable bottom that may be expressed by lateral and vertical force components as shown in Fig. 7.47. The dynamic response of a pipeline or cable protector to these hydrodynamic forces may be divided into the two categories defined by Eagleson and Dean (1959) for sediment motion under waves; viz., incipient motion and established motion. Only the case of incipient motion will be considered for pipeline stabilization. The incipient motion criterion for the pipeline requires that the x-component of the hydrodynamic force Fpx in Fig. 7.47 be less than the sum of the two resisting forces represented by the tensile strength of the pipeline FT and the Coulomb friction force Fp. This stable equilibrium condition may be expressed as FP • 2 r
x \FT\l\FF\WF»ex < 1.
W+
(7.170)
Neglecting the ratio \Fr \/\Fp | in Eq. (7.170) results in a conservative lateral stability criterion given by FPX
=
FFX
dFm+dFd
< h
(?
i?i)
FFx
where the x -component of the lateral inertia dFm and drag dFd forces per unit length on the pipeline may be computed from the Morison equation Eqs. (7.10a-c) evaluated at the bottom z = —h{y) according to (TTD2\
dFm = Cmp
——
ut cos #00,
dFd = '-CdpD^u1 + U'£+ 2uUc cos0(y) [u cos6»(y) + UC],
(7.172a) (7.172b)
Waves and Wave Forces on Coastal and Ocean Structures
Fig. 7.47. Definition sketch for bottom-laid pipeline stability (Huang and Hudspeth, 1982).
where the horizontal {x, y} fluid velocity vector q evaluated at z = —h(y) is q = (ucosO + Uc)ex + u sin6ey;
|^| 2 = u2 + t/£
+2uUccosO, (7.172c,d)
where Q = drag force coefficient; Cm = inertia force coefficient; D = pipeline diameter; and u and ut — instantaneous wave-induced horizontal water particle velocity and local acceleration, respectively, evaluated at z = —h(y); p = mass density of water; and 6(y) = local incident wave angle in Fig. 7.47.
607
Deterministic Dynamics of Small Solid Bodies
The friction force per unit length of pipe Fp may be computed from Coulomb's law of friction with a coefficient n by F F = ^(|Fiv|-|i^l)cos£2x,
(7.173a)
where fi = beach slope; and where 71D2
(7.173b)
FN = PgiYs ~ 1) FL = -CLpDyJu2
+ UC + 2uUc cos6>(y) [u cos6>(y) + Uc] ez, (7.173c)
where the specific gravity of pipe ys = ps/p',ps = mass density of pipe; g — gravitation constant; and CL = lift force coefficient. The application to design conditions of the lateral stability criterion Eq. (7.171), where both the inertia dFm and drag dFd forces per unit length are included, may be reduced to Cd l+li[-^-)cosp\ Cd
yju2 + UC + 2uUc cosG(y) [u cos 9{y) + Uc]
TtD
+ Cm ( — ) ut cos6{y) < ixg(y.• s - l )
—
cos/3.
(7.174)
Dimensionless horizontal water particle velocity and acceleration (denoted by tildes ~) u and ut evaluated at the bottom z = —h{y) and a dimensionless steady, uniform horizontal current Uc are defined by H(y)/T'
ut ut = H(y)/T2' '"
Uc UC = ^ H(y)/T
(7.175a-c)
that reduces Eq. (7.174) to the following dimensionless form: H(y)
D < V(Ys ~ l)?r — cos/S, x [ ^ + cos 0(y)\+Cm^^ut
cos 6(y) (7.176)
where the dimensionless force coefficient ratio G is G = 1 +ii\
— )cos/3,
(7.177)
608
Waves and Wave Forces on Coastal and Ocean Structures
H = friction coefficient and XQ = gT2/2n = the linear theory wavelength in deep-water given by Eq. (4.24h) in Chapter 4.3. For a bottom-laid pipeline with a diameter that is small compared to a wavelength (D/X <^ 1), the dimensionless lateral water particle bottom velocity and acceleration u\j and Ubt computed at the center of the pipeline (i.e., z = —h(y)[l — D/2h(y)] « —h (y) in Fig. 7.47 may be computed to represent the water particle kinematics at the bottom at z — —h(y) in the Morison equations (7.10a-c). For the special case where Uc — 0, Eq. (7.176) may be reduced further to PB < Ps,
(7.178)
where the dimensionless wave force acting on the pipeline PB is PB = ul + T^Gubt
(7.179a)
and the dimensionless pipeline stability parameter Ps is t Ps = -P11 — _ -T, 2 {(#(v)Ao) QGcos0Oo}'
(7.179b)
where G is a dimensionless force ratio defined by the following: Cm G =
D (7.179c)
Cd[l+n(CL/Cd)co*p]H(y) that is a measure of the relative contributions between the inertia force component dFm and the drag force dFd component. The friction resistance between a pipeline and a rigid, impermeable bottom depends on the following factors (Anand and Agarwal, 1980): 1) the surface coating of the pipeline; 2) the benthic material (viz., the geotechnical properties, the angle of the internal friction, the cohesion of the benthic material, and the degree of saturation); 3) the contact area between the pipeline and the bottom; and 4) the direction of the pipeline motion. Potynody (1961) estimates values for the friction coefficient (x between various soils and pipe materials by conducting direct shear tests where the contact surface between the pipe materials and soils was horizontal. Valent (1979)
609
Deterministic Dynamics of Small Solid Bodies
Table 7.4. Summary of friction coefficient from Potynody (1961) and Valent (1979). Pipeline Material Smooth steel
Soil Material
M
Sand Dense cohesionless silt Clay
0.64 0.68 0.50
Coralline sand Oolitic sand Foram sand-silt Rough steel
Sand Dense cohesionless silt Clay Coralline sand Oolitic sand Foram sand-silt
Smooth concrete
Sand Dense cohesionless silt Clay Coralline sand Oolitic sand Foram sand-silt
Grained concrete
Rough concrete
0.17-0.18 0.31 0.37 0.80 0.75 0.80 0.55 0.50-0.51 0.66 0.80 0.87 1.00 0.56 0.52-0.54 0.67
Reference Potynody (1961)
Valent (1979)
Potynody (1961)
Valent (1979)
Potynody (1961)
Valent (1979)
Sand Dense cohesionless silt Clay
0.88 0.96 1.00
Potynody (1961)
Sand Dense cohesionless silt Clay
0.90 1.00 1.00
Potynody (1961)
Coralline sand Oolitic sand
0.59 0.57
Valent (1979)
also computes experimental values for the friction coefficient /x for pipelines made of steel and of concrete by conducting direct shear tests using three different beach sand materials and two different building materials. Both the Potynody (1961) and Valent (1979) tests results are summarized in Table 7.4.
610
Waves and Wave Forces on Coastal and Ocean Structures
Coulomb friction coefficients for concrete-coated pipelines on both sandy and silty soils are reported by Lyons (1973) and by Anand and Agarwal (1980) using both model and prototype pulling tests. There are no laboratory or field measurements for the friction coefficient for pipeline materials on an impermeable rock bottom. For design purposes (if no accurate site survey data are available), a value of /u. = 0.33 for the friction coefficient is recommended for stone masonry and wet undisturbed ground (Eshbach, 1974). The maximum dimensionless transverse hydrodynamic force exerted on a bottom-laid pipe may be divided into three distinct regions on a dimensionless relative displacement H/D relative water depth h/Xo dissection plane where the linear wave theory deep-water wavelength ko = gT2/2TZ by Eq. (4.24h) in Chapter 4.3. These three regions are determined from values of the ratio of the amplitude of the drag force \dFd\ to the amplitude of the inertia force \dFm | computed by linear wave theory from \dFd\ = (Cd\ W = ^\dF \ = \C (^L)) m m
H_ = (Cd\
K_
— = \C ( 7^) "J. TTD m) TT<
(7180>
where the Keulegan-Carpenter parameter K — Hn/D for linear wave theory kinematics (Eq. (7.44c) in Sec. 7.6.1). For design purposes, two values of W may be shown on this dissection plane for W — 20 (drag force component dominates) and for W — 0.05 (inertia force component dominates). Region A in the upper-left corner of Fig. 7.48 applies to design conditions where the drag force dominates over the inertia force. In general, steep waves in shallow water create the conditions encountered in this region. In Region C in the lower right corner of Fig. 7.48, the situation is reversed. In Region B, both the drag force and inertia force components should be considered for design. The wave height H(y) in Eq. (7.176) may be expressed in terms of the deep-water swell height Ho, a refraction coefficient Kr(y), and a shoaling coefficient Ks(y) by the following
H(y) = Kr(y)Ks(y)H0,
Kr(y) = Jjiy),
K^l(y) = J 2 n ( y ) ^
Deterministic Dynamics of Small Solid Bodies
611
1000
Fig. 7.48. Parametric regions of importance for ratio of drag/inertia force components (Huang and Hudspeth, 1982).
where Eq. (4.60b) in Chapter 4.5 gives n(y) = - [1 +
2kh(y)cosech2kh(y)]
and bo and b{y) = the horizontal distances between wave orthogonals in deepwater and at a specific pipeline location y, respectively, shown in Fig. 7.47 and Fig. (4.12) in Chapter 4.6. Substituting for u, ut and H(y) from linear wave theory and invoking the small argument approximation z ~ —h(y), the general stability criterion given by Eq. (7.176) may now be expressed by: Kr(y)Ks(y)
Kr(y)Ks(y) ( § )
CdG
cosh kh(y)
cos 2 cot
y 1 + (Uc7t~l sinh[fc/i(;y)]cos_1 cot)2 + 2Ucn~1 sinhlMOOlcos -1 cotcosO(y) x[Ucsinh[kh(y)]cos •C,
l
cot + n cos6(y)]
-Hi) cosh lkh(y) smcot cos0(y) J™?'
(7.181)
Waves and Wave Forces on Coastal and Ocean Structures
612
Fig. 7.49. Contours of dimensionless wave force on a pipeline P# computed by Dean stream function wave theory (1965) for G = 5 (Huang and Hudspeth, 1982).
The maximum value of the LHS of Eq. (7.181) for the special case in that Uc = 0 is given by CmnD sinh[kh(y)] V
[Kr(y)Ks(y)cosech[kh(y)]y < fi(ys - 1)
Ho)
2Kr(y)Ks(y)HoCdGj \H0
(QG^cosjS
cos^(v) (7.182)
and defines the lateral stability criterion for linear wave theory kinematics with UC = 0. Because of the infinite number of wave-current combinations that are possible, Huang and Hudspeth (1982) applied Eq. (7.178) to represent the lateral stability for a bottom-laid pipeline under a shoaling, finite-amplitude wave without a current Uc- The local values for the wave height H(y) and wave direction 0{y) at a specific location may be determined iteratively following the procedures proposed by Dean (1974) or directly from site survey data. The maximum values of the dimensionless wave force PB given by Eq. (7.179a) are computed by the Dean stream function kinematics and contoured on a wave height parameter H/T2 versus water depth parameter h/T2 dissection plane. Figure 7.49 illustrates the parametric dependency of contours of the dimensionless wave force PB for four wave height H to breaking wave height
613
Deterministic Dynamics of Small Solid Bodies
HB ratios H/HB (viz., A = 25% breaking, B = 50% breaking, C = 75% breaking and D = 100% breaking in Fig. 6.10 in Chapter 6.5) when the dimensionless force ratio G equal to 5.0. The dimensionless relative errors for PB computed by the stream function wave theory compared to the linear wave theory are summarized by Huang and Hudspeth (1982) for the 40 Dean stream function wave cases. The values of the dimensionless force ratio G may be calculated for the following three design conditions: 1) /x = 0.33 (hard-rock), 2)11 = 0.55 (sand), and 3) //, = 1.0 (silty-soil), with zero-slope (cos ft = 1.0). The results are presented in Figs. 7.50-7.52 as contours of the dimensionless force ratio G on a dimensionless relative displacement H/D versus dimensionless relative water depth h/Xo dissection plane. These results provide an approximate range for values of the dimensionless force ratio G that may be applied as design aids for estimating the lateral stability of a bottom-laid pipeline. In conditions where bottom-laid pipelines are subjected to shoaling finiteamplitude waves, it is necessary to estimate the maximum water depth where lateral stability protection is required. Between the shoreline and this maximum depth offshore, it is necessary to either bury a pipeline or to use other means of stabilization in the case of a rigid, impermeable beach. This maximum depth may be determined from the dimensionless design curves of PB
,o 2
H D 10°
K Fig. 7.50. Dimensionless force ratio G computed from linear wave theory for /J, — 0.33 (rock) and p = 0° (Huang and Hudspeth, 1982).
Waves and Wave Forces on Coastal and Ocean Structures
614
Fig. 7.51. Dimensionless force ratio G computed from linear wave theory for ^ = 0.55 (sand) and yS = 0° (Huang and Hudspeth, 1982). 10 Os"
1
10
H D 10°
' S /
..-1 10
2 5
/ ^ -2 10
/,
-1 10
Fig. 7.52. Dimensionless force ratio G computed from linear wave theory for fi = 1.0 (siltysand) and p = 0° (Huang and Hudspeth, 1982).
from Huang and Hudspeth (1982) by the following procedure: (i) Estimate the design wave height H(v) wave period T and the direction of wave propagation 9(y) along the proposed pipeline route, (ii) Compute Kr(y) and Ks(y) along the proposed pipeline route, (iii) Determine the local wave height H (v) and wave direction 9 (v) along the proposed route. (Steps (ii) and (iii) may be omitted if site survey data are available that give both local wave height and direction.)
615
Deterministic Dynamics of Small Solid Bodies
(iv) Beginning seaward of the breaker zone determined in Steps (ii) and (iii), compute the wave height parameter H/T2, depth parameter h/T2, and the dimensionless force ration G either by applying Eq. (7.179c) or Figs. 7.50-7.52. (v) Estimate the value of the dimensionless force PB from the design curves given by Huang and Hudspeth (1982) (e.g., Fig. 7.49) by entering a vertical line through the design depth parameter h/T2 and a horizontal line through the design wave height parameter H/T2. The value of the dimensionless force PB may then be estimated at the intersection of these two lines using the values from the adjacent contours of constant PB • Numerical interpolation may be used for values of G that differ from the three values selected by Huang and Hudspeth (1982). (vi) Compute the dimensionless pipeline stability parameter Ps from Eq. (7.179b). (vii) If PB > Ps then the pipeline at this depth must be stabilized, (viii) Repeat Steps (iv) to (vii) moving seaward along the proposed route until PB < Ps- The maximum depth for pipeline stabilization is at this location.
7.11. Problems 7.1.
Given the following properties for the articulated tower analyzed in Sec. 7.8.1: h = 135 m; D = 4.5 m; Lp = Lg = 150 m; Mg = 880,000 kg; Wp/Wg = 0.2; Wb/Wg = 0.5; hb/h = 0.45; Lg/h = 1.111; Lt,/h = 1.0; Lp/h = 1.111 and the following wave conditions: H(m) T(sec)
7.5/10/15/20/30 9.8
15/20/30 13.9
10/30 17.4
a) Compute the natural undamped period To and frequency /o in air in the pitch 05 mode about the base at z = — h. b) Select an appropriate hydrodynamic load model for the waveinduced hydrodynamic pressure loading and justify your selection.
616
Waves and Wave Forces on Coastal and Ocean Structures
c) Compute the undamped Tn and damped Td natural periods in pitch 05 about the base including hydrodynamic effects. d) Compute the following dynamic response constants: Ast = rads; £2s = rads; 0:5 = rads e) Compute the damping ratio £5. 7.2. Assume that the measured force/unit length on a small vertical circular cylinder may be written as dFr(z,t) = —dfm(z) sin cot + dfd(z)coscot\ cos cot \ Apply Fourier analyses to show that if the Morison equation is given by /pitD2\
dFx(z,t) = -Cm I —-— I um(z)co sin cot D 2 + CdP-rum(z) coscot\ coscot\ then the Morison equation force coefficients Cm and Cd may be computed from r
4dfm(z) pit D2coum{z)'
_
2dfd{z) pDu^iz)
7.3.
Applying linear wave theory kinematics from Chapter 4.4, compute the total force on a vertical circular pile for Q = 1.0; Cm = 1.5; D = 3.0 ft; H = 30 ft; T = 10 sec;h = 550 ft(F max = 13.2kips). 7.4. Assume that the inertia and drag coefficients in the Morison equation areC m = 1.40 and Cd = 1.05. Apply the Morison equation with linear wave theory kinematics from Chapter 4.4 and determine horizontal and vertical forces on the pontoon shown in Fig. 7.39 in Sec. 7.8.2 for h = 50ft;// = 10ft; andT = 12 sec. Plot these forces as a function of time for one wave period T and estimate the maximum values from these plots. 7.5. A vertical circular cylinder of D = 4 ft is located in h = 100 ft of water. a) what is the Froude parameter Fr and Reynolds parameter R for the maximum velocity of a wave with a height H = 42 ft and a period T = 13 sec by linear wave theory kinematics?
617
Deterministic Dynamics of Small Solid Bodies
b) compute the maximum force the pile by the Morison equation with linear wave theory kinematics from Chapter 4.4. 7.6.
Two horizontal cylinders, each 40 ft long, are connected at each end with a rigid connector as shown in Fig. 7.39 in Sec. 7.8.2. Assume that the weight ofthe connector is included in the cylinder weights. The large cylinder diameter is DL = 4 ft and weighs 5,220 lbs; the small cylinder diameter is Ds — 2 ft and weighs 1,350 lbs; / = 11 ft. The water depth is h = 80 ft; wave height H = 4 ft; wave period T — 4.1 seconds; uniform current is Uc = 1.0 knot. Compute the following: a) with the current only acting, determine the length I of the mooring line at each end and the rigid connector distance xo so that the rigid connector line a-c-b will be horizontal at the still water surface. Each of the two mooring lines has a cross-sectional area of .44 in2. The linear modulus of elasticity E of the mooring is 20 x 106 psi. Assume that the ultimate breaking strength is very large. The stiffnesses in x (horizontal) and z (vertical) per mooring line are given by Eqs. (7.117) in Sec. 7.8.2. b) compute the dimensionless dynamic amplification factor D [a)j£] for sway j = 2 and heave j = 3 about the steady state position with current. In addition, compute the dynamic magnification factor D[<w -h ] for roll j = 5 about the points, a, b, and c. c) tabulate each of the quantities in the following table: MODE SWAY (j = 2) HEAVE O" = 3) ROLL (J = 5)
7.7.
^(Mf)
Pj
«>jh
*stj
O l^-KI^I)
a
j
Compare the general quadric error surface given by the equation (7.68) in Sec. 7.7.1 (aX)2 + 2HXY + OSy)2 - 2GX - 2JY + C = 0. with the least-squares equation (7.67) for Cd and Cm given by A
(CdP\2
. 4,^ (r
x (u\u\ut) - 2 (^\
P*D\2
2
(CdP\
(CmPitD\
(FTu\u\) - 2 (CmP*D\ {FTUt)
+
{F2}
618
Waves and Wave Forces on Coastal and Ocean Structures
where Cd = X and Cm = Y. Compute the equations (7.69a-c) for the origin (Xo, Fo) of the translated axes for Q and Cm if the equation for the translated origin is derived from the simultaneous solution to HXQ + p2Y0 -J
= 0,
a2X0 + HY0 -G
=0
Compute Eq. (7.69d) for the rotation of the Cd and Cm axes. 7.8. Verify that the following phase method Eq. (7.49b) is correct: dFx(z,9max)
= dfd(z) 1 +
1 (dfrniXyV 4\dfd(z)J
(7.49b)
7.9. Verify that the following least squares Eqs. (7.53) are correct.
2
Cd =
*
\M) Jo
°2* dFT(z,e) sin 9 dO, o pDum{z)
(7.53a)
dFT(z,9) cos 9\cos0\d9 pDum(z)
(7.53b)
7.10. Verify that the following Fourier Method Eqs. (7.55) are correct K2
—If
Mn(z) =
* dFT(z,6) cos(2n9)d0, pDum(z)
(l+8„o)7t]Jo • Sn2o)n J Jo ' * dFT(z,9) A2n+\(z) cos((2« + \)9)d9, n Jo pDum(z) 271
B2 n(z)
= - f0 •x Jo
B2n+l(z)
7.11.
=-f
In
0
dFT(z,9) sin(2n9)d9, pDum(z)
dFT(z,9) pF>um(z)
sin((2n + \)9)d9.
(7.55a) (7.55b) (7.55c) (7.55d)
7T Jo
Verify that the following linearized drag coefficient Cdi in Eq. (7.60) is correct: Cdi =
—CdUm(z) 571
(7.60)
Chapter 8
Deterministic Dynamics of Large Solid Bodies
8.1. Dynamic Response of Large Bodies: An Overview The importance of focusing on the principles of physics before selecting mathematical models to analyze the dynamic response of coastal and ocean structures is highlighted in Chapter 7.1 and will not be repeated here. In contrast to the dynamics of small solid bodies, the effect of large bodies on an incident wave field is not negligible. Carefully observing what is seen or not seen when long-crested surface gravity waves interact with large solid bodies identifies the following two major physical phenomena: Seen:
Waves may be observed radiating away from large, semiimmersed solid bodies and the incident wave may be modified significantly. Not Seen: When the solid body is sufficiently large compared to the incident wave height and wavelength, no wakes due to flow separation may be observed. Figure 8.1 illustrates the linear superposition of an exciting force on a. fixed body plus the restoring force (wavemaker theory) on a large solid body. In contrast to Fig. 7.1 in Chapter 7.1, the motion of the large body in otherwise still water generates radiated waves (wavemaker theory). Also in contrast to Fig. 7.1 in Chapter 7.1, the added mass Hjk and radiation damping kj^ coefficients from the radiated force are not equal to the exciting force coefficients C,f and C,f. 619
Waves and Wave Forces on Coastal and Ocean
PHYSICS
Structures
LARGE BODY DYNAMICS SCATTERED & RADIATED WAVES v X T L O W SEPARATION IDEAL & INVISCID POTENTIAL THEORY
SEEN NOT SEEN FLUID SOLUTION EXCITING FORCE on a FIXED BODY
+
RADIATED FORCE (WAVEMAKER THEORY)
•^s\r il
I
II DYNAMIC RESPONSE LINEAR SUPERPOSITION
MATH MODEL
F(t) = FE(t) + F'(t) = CEuu,(t) + £EU u(t) - aik u,(t) - Xik u(t)
Fig. 8.1. Large body dynamic response.
When no flow separation occurs, real fluid effects due to fluid viscosity may be neglected and a potential theory for an ideal fluid may be applied to compute the wave loads. A boundary value problem (B VP) for the motion of an ideal fluid may be solved to account for this type of wave-structure interaction. Because both the radiation conditions and the fundamental singular solutions to Laplace's equation (viz., Green's functions) are different for 2D and for 3D geometry, 2D solid bodies must be distinguished from 3D solid bodies. The solid body boundary between an ideal fluid and a solid body requires special care. Because this boundary isfree, both kinematic and dynamic boundary conditions are required. Because the fluid motion is an Eulerian field with time and space as independent variables and the solid body is a Lagrangian body with time and the body as independent variables, these kinematic and dynamic boundary conditions will be different from the free surface boundary that separates two Eulerian fields; viz., air and water (cf, Chapter 4.2). The boundary between the fluid and solid body separates an Eulerian field from a Lagrangian body and these boundary conditions are somewhat more complicated.
Deterministic Dynamics of Large Solid Bodies
621
Fig. 8.2. (a) Definition sketch for semi-immersed MDOF Lagrangian solid body, (b) Locations of the origins of the inertial and non-inertial coordinate systems.
The solid body motion of the semi-immersed body shown in Figs. 8.2 may be described in Lagrangian coordinates by three translational displacements [X\(t)= surge; X2(t)— sway; X^(t) = heave] and three rotational displacements [04(f) = roll; ©5(0 = pitch; and ©6(0 =yaw]. The direction of the unit normal to the solid body is ambiguous and must be defined as either into or out of the solid body surface. The unit normal vector in Fig. 8.3 points out of the body and into the fluid and is defined by n = n\e\ + «22 + «3
(8.1)
with direction cosines defined by n\ = cos(0i,*'),
«2 = cos (02,/),
«3 = cos(03,z'),
(8.2a,b,c)
where 9j = angle between the unit normal vector n and the unit basis vector ej. The direction cosines nj defined by Eqs. (8.2) are dimensionless. Pseudodirection cosines for the three rotational modes about an arbitrary point of rotation R in Figs. 8.2 in the body-fixed, non-inertia coordinates that are defined by the coordinate vector XR = {x'R,y'R,zR}
(8.3a)
and by an arbitrary point of translation C in Figs. 8.2 that is defined in bodyfixed, non-inertial coordinates by the coordinate vector xc = {x'c,y'c,z'c}.
(8.3b)
622
Waves and Wave Forces on Coastal and Ocean Structures
Fig. 8.3. Unit normal points out of the Lagrangian solid body and into the fluid.
The pseudo-direction cosines are defined by the following: rc x n = nAe\ + n5ez + n6ei,
(8.4a)
r' — v ' - v '
(8.4b)
where from Fig. 8.2(a)
and » ' = Sjkm rj nk em,
j,k,m
(8.4c)
- 1,2,3,
where Sjkm = the Levi-Civita permutation symbol (vide., Chapter 2.2.4); and r
'd =x'c~
X
n
4 = {rc2"3 - r'c^m),
(8.5a,b)
r
C2 =y'c-
y'R>
" 5 = {rC3wl -
C! W 3},
(8.5c,d)
{r'Cln2-r'C2ni}.
(8.5e,f)
r
'R>
C3=Zc-4.
n'6 =
r
The pseudo-direction cosines n'- defined by Eqs. (8.5b,d,f) have the dimensions of [Length].
Deterministic Dynamics of Large Solid Bodies
623
The fixed inertial x axis is defined by the longitudinal axis (or generator) of the body and not by the direction of wave propagation. These motions are defined by the body coordinate axis and not by the wave-field coordinate axis. The three translational modes of oscillation for a Multiple-Degree-OfFreedom oscillator (MDOF) are given by X(t)=X(t)+R(t),
R(t) = 0(t)xr'c,
(8.6a,b)
where r'c in Figs. 8.2 is defined by Eq. (8.4b). The displacement R(t) is time dependent and defined by Eq. (8.6b), whileic^ is a time independent coordinate vector in the body-fixed, non-inertial coordinates defined by Eq. (8.3a). For strictly periodic solid body motions with period T = 2n/co, the harmonic time dependence may be given by exp — (icot). The complex-valued, scalar components of the Lagrangian displacement, velocity and acceleration of the solid body are given by the real parts of Xj(t) = Re{i%j exp - icot} = |£/ | sin(cot - ctj), ®j(t) = Re{i£2j exp - icot} = \Qj\ sin(cot - ctj),
j = 1,2,3, j = 4,5,6,
Xj (t) = Re{co%j exp - icot} = co\i-j | cos(cot -ctj),
j = 1,2,3,
®j(t) = Re{coQj exp — icot} = co\£2j\cos(cot — ctj),
(8.7c)
(8.7e)
j = 1,2,3,
(8.7f)
®j(t) = Re{-ico2Qj exp - icot} = -co2\Qj\ sin(cot - ctj), = Re{-i<w©;(0},
(8.7b)
j = 4,5,6, (8.7d)
Xj(t) = Re{-ico2t-j exp - icot} = -co2\i-j\sin(cot - ctj), = Re{-icoXj(t)},
(8.7a)
j = 4,5,6,
(8.7g) (8.7h)
where the over dots represent ordinary Lagrangian temporal derivatives according to , ,
x
dXj(t)
.
d®,(t)
where the complex-valued amplitudes of the translational l-j (j = 1,2,3) and the rotational modes £2j (j = 4,5,6) are defined by S; = IfyI expiaj,
j = 1,2,3,
Slj = \Slj\ expictj,
j = 4,5,6. (8.7k,l)
624
Waves and Wave Forces on Coastal and Ocean Structures
The free boundary between the Eulerian fluid and the Lagrangian solid body may be described by a Stokes material surface S(x,y,z,t) (vide., Chapter 2.2.10) that requires two boundary conditions; viz., 1) an Eulerian .Kinematic 5ody .Boundary Condition (KBBC) for an ideal fluid; and 2) a Lagrangian Dynamic .Boundary Condition (DBC). It is important to focus on the differences in the boundary conditions for 1) a free boundary between a Lagrangian solid body and an Eulerian fluid field; and 2) a free boundary between two Eulerian fluid fields such as water and air (vide., Eqs. (4.5e-h) in Chapter 4.2). The fundamental fluid unknowns for the flow of an incompressible fluid are the velocity (a vector q) and pressure (a scalar p). Two boundary conditions are required (just as in the case of a deformable free surface fluid boundary between two Eulerian fluid fields) because the solid boundary isfree to respond to dynamic pressure loads; i.e., the motion is not prescribed (vide., John, 1949 and 1950). Free boundaries between either two Eulerian fluid fields or between an Eulerian fluid field and a Lagrangian solid body require both a .Kinematic (KBC) and a Dynamic (DBC) .Boundary Condition. The principal distinction between the Lagrangian solid body boundary S(x, y, z, t) and the Eulerian free surface fluid field boundary F(x, y, z, t) is that the Lagrangian solid body free boundary S(x, y, z, t) is a non-deforming solid boundary between a Lagrangian solid body and an Eulerian fluid field. In contrast, the Eulerian fluid field free surface boundary F(x, y, z, t) is a deformable boundary between two Eulerian fluid fields. The Problem Solution Technique (PST) for linearly decomposing the boundary value problem for a scalar velocity potential <$> (x, y, z, t) for the Eulerian fluid field motion into three separate boundary value problems is easily demonstrated and motivated by a close examination of both the .Kinematic (KBC) and Dynamic (DBC) .Boundary Conditions on the Lagrangian solid body free boundary S(x, y, z, t). Free boundaries between an Eulerian fluid field and either another deformable Eulerian fluid field or a non-deforming Lagrangian solid body require both a .Kinematic (KBC) and a Dynamic (DBC) boundary condition because the boundaries are free and because there are two fundamental fluid unknowns for an incompressible fluid; viz., velocity and pressure. The exact forms of these two boundary conditions depend on 1) whether or not the fluid is a real or an ideal fluid; and 2) whether or not the^ree boundary is a nondeforming boundary between a Lagrangian solid body and an Eulerian fluid field or a deformable boundary between two Eulerian fluid fields.
Deterministic
625
Dynamics of Large Solid Bodies
Because the Lagrangian solid body is assumed to be a dimensionally large body with respect to the dimensions of both the wavelength X and wave height H, no fluid flow separation will occur around the Lagrangian solid body; and, consequently, a potential theory solution for an ideal, inviscid and incompressible fluid may be computed. The fundamental fluid unknowns of velocity q(x, y, z, t) and dynamic pressure p(x, y, z, t) may be computed from a scalar velocity potential <£>(x, y, z, t) according to the following: q(x,y,z,t)
= {u(x,y,z,t),v(x,y,z,t),w(x,y,z,t)}
=
-VQ(x,y,z,t), (8.8a)
p(x,y,z,t)
= P
dQ(x,y,z,t) at
•
(8.8b)
From physical observations, it is easy to motivate a linear decomposition of the scalar velocity potential <£(x, y,z,t) into an incident wave potential ^wi*, y, z, t) and a body disturbance potential <&b(x, y, z, t) that is due to the presence of the solid body. This linear decomposition is valid only for small oscillations about an equilibrium position of the solid body QS(X, y, z, 0). The .Kinematic Body boundary Condition (KBBC) for an ideal (incompressible, inviscid) fluid is given by the following single scalar equation that is evaluated at the equilibrium position oS(x, y, z, 0): -V4> • n=X»ii
+
&»n',
- ( V ^ u , + V0>fe) • n=X»n
+
Q»n',
— (w + (S>b)+X*n + ®»n'= 0; on 0S(x,y,z,0), (8.9a) an where the six solid body Lagrangian velocities are defined by Eqs. (8.7c, d). The scalar kinematic boundary condition transforms the solid body Lagrangian velocities X(t) and 0(?) into Eulerian fields through vector dot products with the spatially dependent unit normals « and n' to the Lagrangian solid body 0S(x,y,z,0). Because the free boundary is a boundary between an non-deforming Lagrangian solid body and an Eulerian fluid field, the Dynamic boundary Condition (DBC) is not expressed in terms of the local Eulerian fluid field pressure as in the case of the deformable free surface boundary in the DFSBC in Eq. (4.5f) in Chapter 4.2. Instead, the Dynamic 5ody boundary Condition
626
Waves and Wave Forces on Coastal and Ocean Structures
(DBBC) is expressed in terms of the integrated fluid pressure force/moment through the following vector equations of Lagrangian solid body motions evaluated at the equilibrium position QS{X, y, z, 0):
m — _^_
'
iX(t) —— — —e
. iO(0 ,
J
J
• +
•
—— —
£lifr;(0 j
(8.9b) (8.9c) (8.9d) ~(8~9e)~ (8.9f) (8-9g)
where e = kA = dimensionless perturbation scale for the smallness of the time-dependent motions about equilibrium QS{X, y,z,0); and where both the time-dependent (subscript = 1 ) and time-independent (subscript = 0) forces/moments are derived in detail in Sec. 8.2. The notation in Eqs. (8.9b-g) and in all subsequent equations for linear and angular momenta is an implied stacked notation with the three translational momenta stacked vertically above the three angular momenta and separated by a horizontal dashed line [ ];i.e.,
lX(t)
|
10(0 J
iXi(0 1*2(0 1*3(0 104(0 105(0 106(0
The forces/moments on the RHS of Eqs. (8.9b-g) include: 1) stiffnesses due to buoyancy, body weight and water plane surface areas; 2) external mooring constraints; 3) the exciting forces/moments on a fixed, still body; and 4) the restoring forces/moments on a body oscillating in otherwise still-fluid (i.e., the wave maker problem in Chapter 5.). The Lagrangian vector dynamic boundary conditions in Eqs. (8.9b-g) transform the Eulerian dynamic pressure field in Eq. (8.8b) and hydrostatic pressure ps = —yz into a time-independent and a time-dependent Lagrangian solid body force/moment by integration over the immersed surface area of the Lagrangian solid body at the equilibrium position 0S(x,y,z,0).
Deterministic
Dynamics of Large Solid Bodies
627
The scalar Eulerian .Kinematic fiody boundary Condition (KBBC) Eq. (8.9a) and the Lagrangian vector Dynamic .Body .Boundary Conditions (DBBC) given by Eqs. (8.9b-g) represent seven scalar equations in seven unknowns; viz.,
(8.10)
where <£>s (x, y, z, 0 = scattered wave potential and <£>r (x, y, z, 0 = radiated wave potential. Substituting Eq. (8.10) into Eq. (8.9a) gives fd
on
0S(x,y,z,0),
(8.11) that gives a diffraction boundary value problem for <&D(x,y,z,t) = $>w(x,y,z,t)
+
$>s(x,y,z,t)
= Re{ A((j)w (x, y, z) + (/>s(x, y, z)) exp - icot}
(8.12a)
with the following homogeneous .Kinematic .Body .Boundary Condition (KBBC) at the equilibrium position oS(x, y, z, 0): ^ = 0 , an
on
0S(x,y,z,0)
(8.12b)
628
Waves and Wave Forces on Coastal and Ocean Structures
or, equivalently, the following inhomogeneous kinematic 5ody boundary Condition (KBBC) for the scattered potential <&s(x, y, z, t): a^
=
an
_ 3 ^
o n
z>0)_
o 5 (
( 8 1 2 C )
an
Because of the identically equal sign [=] in Eq. (8.11), the second term in brackets {•} on the left-hand-side of Eq. (8.11) may also be linearly decomposed into six separate .Kinematic 5ody .Boundary Conditions (KBBC) by the following linear superposition for <&r (x, y, z, t): 6
<$>r(x,y,z,t) = ^ l > y ( x , j , z , 0 7=1
3
(8.13) v
6
'
= ] T Xj (t)4>j (x, y, Z) + J2 ®j (*)4>j (*, y, z) 7=1
7=4
and substituted into the second term in brackets {•} in Eq. (8.11) to obtain
*l(0 {^7 +nx\+ ±2{t) {^T + m\ + *m \lt+n3} +
@4(t)\^1+nf4\+@5(t)\^p-+n'5 [an
J
+ ©6(0{^+<}=0,
[ dn
on
0S(x,y,z,0).
(8.14a-f)
By factoring explicitly the unknown amplitudes of the six Lagrangian solid body velocities Xj(t) and ®j(t) for each of the unknown radiated (or wavemaker) potentials y (x, y, z, t), six inhomogeneous kinematic .Body .Boundary Conditions (KBBC) are obtained that depend only on the equilibrium position of the solid body oS(x,y,z, 0) and not on the amplitude of the motion of the solid body (i.e., d(j)\/an = — n\, dfo/dn = — nj, etc.)! Of course, these unknown amplitudes must still be determined; and they are readily computed from the Dynamic .Boundary Conditions (DBC) in Eqs. (8.9b-g). In the linear diffraction theory, the interaction of a fixed Lagrangian solid body with the incident wave field generates a scattered wave. If a semi-immersed body is freely floating or semi-constrained, the response also generates radiated waves. Because of linear superposition, the total fluid
629
Deterministic Dynamics of Large Solid Bodies
motion resulting from this wave-structure interaction is assumed to be a linear combination of the incident wave <&w(x, y,z,t) plus a scattered wave <£>s(x,y,z,t) plus six individual radiated waves each resulting from one of the six modes of solid body motions from <&r(x,y,z,t). Because the incident wave and the resulting Lagrangian solid body motions are assumed to be simple harmonic in time (viz., oc exp— icot), each of these three uppercase scalar velocity potentials may now be defined by the real part of a linear sum of complex-valued, time dependent amplitudes and lower-case spatially dependent velocity potentials by Eqs. (8.12a and 8.13) according to <&(*, y, z, t)= <&D(x, y, z, t) + ®r(x, y, z, t) I" A[cj)w(x,y,z) +
fa(x,y,z)]
3
~]
6
+co^2^j
~ L
•/=!
XP
lC
°
'
J
7=4
(8.15) where each of the lower-case scalar velocity potentials fa (x,y,z) are functions only of the spatial coordinates (x,y,z); and where the amplitudes of each velocity potential A, %j and Qj are complex-valued amplitudes of the incident wave and the translational and rotational Lagrangian solid body displacements, respectively. The boundary value problems for each of the seven unknown lower-case scalar potentials fa (x,y,z) plus the known incident wave potential are identical with the exception of: 1) the inhomogeneous kinematic 5ody boundary Condition (KBBC) on the Lagrangian solid body and 2) a radiation condition to be imposed at large horizontal distances from the equilibrium position of the Lagrangian solid body oS(x, y, z, 0). The governing field equation for each of the lower-case scalar velocity potential fa (x, y, z) is Laplace's equation (2.12a) in Chapter 2.2.9 for the conservation of mass of an ideal (incompressible and inviscid) fluid in irrotational flow; i.e., V2fa=0,
|x|
\y\
-h
(8.16)
where the scalar Laplacian operator in Eq. (2.12a) in Chapter 2.2.9 is defined by V 2 (.) = 3 2 (.)/3x 2 + 9 2 (*)/3y 2 + 3 2 (.)/9z 2 . Each scalar potential fa(x,y,z) must satisfy a .Kinematic boundary Condition (KBC) on the horizontal impermeable bottom given by a Stokes material
630
Waves and Wave Forces on Coastal and Ocean Structures
surface B = z + h = 0 according to DB D(z + h) = - V 0 ; • n = V(pj • e3 = 0 Dt Dt that reduces to
(8.17a)
^p- = 0, | J C | < O O , | y | < o o , z = -A (8.17b) oz for a rigid, impermeable, horizontal seabed. Each potential >j must satisfy the linearized Combined kinematic and Dynamic Free Surface boundary Condition (CKDFSBC) given by -P-- ko>j = 0, \x\
+ ®(t)»n'
= 0,
on
0S(x,y,z,0),
that yields seven inhomogeneous kinematic body boundary conditions (KBBC). Again, this scalar kinematic 5ody boundary Condition (KBBC) converts the Lagrangian motions of the solid body X(t) and 0 ( 0 to Eulerian fields through the vector dot product with the spatially dependent unit normals it and n' to the equilibrium position QS{X, y, z, 0) of the Lagrangian solid body. For the diffraction of waves by a fixed solid body, the inhomogeneous kinematic .Body boundary Condition (KBBC) for >s may be computed from V ( 0 U , + 0 J ) « » = O,
on
0S(x,y,z,0)
(8.19a)
or d
30u
,
on
0S(x,y,z,0)
(8.19b)
on on For the radiation of waves by a solid body moving in otherwise still fluid (i.e., the wavemaker boundary value problem from Chapter 5), the inhomogeneous linear kinematic 5ody boundary Condition (KBBC) becomes „ .
—^
3
= HXJ^ni 7=1
6 +
H®}(t)n'j> j=4
on
oS(x,y,z,0),
(8.20)
631
Deterministic Dynamics of Large Solid Bodies
that will yield an inhomogeneous .Kinematic Body boundary Condition (KBBC) for each of the six radiated wave potentials 4>j(x,y,z). Substituting Eqs. (8.7c,d) into Eq. (8.13) provides a separate radiated wave potential for each jth mode of oscillation according to
3 =
CO
6
] T %j
Qj(/)j (x, y, z)exp — icot.
j=4
(8.21a) Substituting Eqs. (8.7c,d) into Eq. (8.20) and equating to the normal derivative of Eq. (8.21a) for each of the potentials cj>j (x, y, z) to the negative of the unit normal results in
Wj
\-njlj
= 1,2,3 1
1 7 = {-»$;;• = 4,5,6 J
^ :
°
n
o*(W,0).
(8.21b) (g21c)
By factoring explicitly, a priori, the unknown, complex-valued amplitudes of the velocities of the Lagrangian solid body defined by Eq. (8.7c,d), the inhomogeneous .Kinematic Body Boundary Condition (KBBC) for the radiated wave potentials requires knowing only the equilibrium geometry QS(X, y, z, 0) of the Lagrangian solid body defined by the direction cosines rij and pseudo-direction cosines n'-. In principle, the unknown amplitudes have been assumed symbolically to be known and the computation of their numerical values is deferred until the Lagrangian Dynamic Boundary Conditions (DBC) Eqs. (8.9b-g) are solved for these Lagrangian solid body motions. For the translational motions denoted by j = 1,2,3, the dimensions of
632
Waves and Wave Forces on Coastal and Ocean Structures
required to account for local pressure effects near the solid body (vide., the wave maker evanescent eigenmodes in Chapter 5.2). The radiation condition eliminates both incoming wave energy and unbounded non-wave, spatially transient solutions. This kinematic Radiation boundary Condition (KRBC) depends on the choice of the time dependence exp ±icot. Two-dimensional surface waves behave like exp ± [sgn(x)]?' Kx; while three-dimensional waves behave like [exp±iKr]/^J~Kr where the free-surface coordinates r 2 = x2 + y 2 > 0. Accordingly, the .Kinematic Radiation boundary Condition (KRBC) at large horizontal distances from the Lagrangian solid body must be (Stoker, 1957) lim (Kr)^-^2 M=>oo
\ ~ ± [8n2 + Snl sgn(x)] IK) 0, = 0, I on J
(8.22)
where n = the number of horizontal surface coordinates; i.e., n = 1 for 2D waves; and n — 2 for 3D waves. The choice of the ± sign is determined totally by the choice of the sign of the argument of exp ±icot. The premultiplication in Eq. (8.22) by */Kr in the case of 3D waves (n = 2) is required in order to insure that only the coefficients in the solution are applied to satisfy the radiation condition and not the radially decaying behavior of the solution given by r _ 1 / / 2 .
Summary The complicated irrotational motion of an ideal fluid that results from the interaction of a semi-immersed Lagrangian solid body with surface gravity waves may be linearly decomposed into three less complicated boundary value problems (BVP) by invoking the following linear superposition. First, the Lagrangian solid body is completely removed from the fluid and a boundary value problem for the incident wave potential (pw(x,y,z) is obtained. Engineering design specifications are given by the wave height H = 2A, the wave period T, and the water depth h and allow this potential to be completely quantified (vide., Chapter 4.2). Second, a fixed Lagrangian solid body is placed in the incident wave field and a scattered wave potential
Deterministic Dynamics of Large Solid Bodies
633
Condition (KBBC) on the Lagrangian solid body given by the normal gradient of the incident wave potential having a known complex-valued amplitude. Third, the Lagrangian solid body is oscillated in otherwise still fluid (i.e., the wavemaker boundary value problem in Chapter 5). The complex-valued amplitudes of oscillations are assumed to be known symbolically; and six radiated wave potentials
=
[u(x,y,z,t),v(x,y,z,t),w(x,y,z,t)}
= -VQ(x,y,z,t)
(8.23a)
634
Waves and Wave Forces on Coastal and Ocean Structures
and the dynamic fluid pressure from the linearized Bernoulli equation d
The three boundary Falue Problems (BVP) for each of the potentials 4>„,(x, y, z, t), <$>s (x, y, z, t), and $>r (x, y, z, t) may be summarized as follows: <&(x,y,z,t) = <S>w(x,y,z,t) +
3
6
7=1
7=4
exp — icot (8.24a)
Laplace's equation 3
2
2
coJ^^V2(t)j{x,y,z)
A[W 4>w{x,y,z) + y 4>s{x,y,z)'\ +
7=1 6
+ a ) ^ ^ 7 ' V 2 0 ; - ( x , ) ' , z ) = 0;
\x\ < oo,
\y\ < oo,
— h < z < 0.
7=4
(8.24b) Combined kinematic and dynamic free-surface boundary condition (CKDFSBC) 3
A[Lf((f>w) + Lf((/)s)] + coJ2^Lf((l)j) 7=1 6
+ o^QjLfi^j)
= 0,
|JC| < oo,
\y\ < oo,
z = 0,
(8.24c)
7=4
where the linear free-surface operator £ / ( • ) is defined by L / ( . ) = [3/3z - *<>](•).
(8-24d)
Bottom boundary condition (BBC) d&w dj>s_ + CO dz dz
3
907 I>^dz + £ V dz 7= 1
x\ < oo,
~~
\y\ < oo,
7=4
z = —/?.
= 0,
(8.24e)
635
Deterministic Dynamics of Large Solid Bodies
Eulerian kinematic body boundary condition (KBBC) d*w dn
9<*>* dn
+
dn
+ X(t)»n + @(t)»ri = 0 on 0S.
(8.24f)
Kinematic radiation boundary condition (KRBC) 3
6
(8.24g) where the deep-water wave number &o = (*)2/g and the linear radiation condition operator Loo(») is defined by Loo(.) = , lim (Kmr)(n-l)'2
\ — ± [8n2 + 8nl sgn(x)]K„ (•), (8.24h)
where n = 1 for 2D waves; n = 2 for 3D waves; and £"i = — ik because the time dependence is chosen to be exp — icot; and the Lagrangian dynamic body boundary condition (DBBC)
m — _/*_
•
(8.24i) (8-24J) (8.24k)
lX(t) — — — = e
. 16(0
£iM„(f)
+
Eo^«
(8.241) (8.24m) (8.24n)
where both the time-dependent 0(e=kA) and time-independent 0(€°) forces/moments are derived in detail in Sec. 8.2. These forces/moments include the following: 1) stiffnesses due to buoyancy, weight and water plane surface area (vide., Eqs. (8.53c-h) in Sec. 8.2); 2) external constraints (vide., Eqs. (8.55a-d) in Sec. 8.2); 3) the exciting forces/moments on a still body (vide., Eq. (8.43) in Sec. 8.2); and 4) the restoring forces/moments on a body oscillating in otherwise still-fluid (i.e., the wave maker boundary value problem and vide., Eqs. (8.44) in Sec. 8.2). The Lagrangian vector dynamic body boundary conditions in Eqs. (8.24i-n) transform the Eulerian dynamic and static pressure fields into time-dependent and time-independent, respectively, Lagrangian solid body forces/moments by integration of the total
636
Waves and Wave Forces on Coastal and Ocean Structures
fluid pressure field over the immersed surface area of the equilibrium position QS(X, y, z, 0) of the Lagrangian solid body. Linear superposition requires the identically equal [=] notation in the kinematic 5ody boundary Condition (KBBC) Eq. (8.24f). Both the scattered 4>s{x,y,z) and radiated
8.2. Linearized MDOF Large Solid Body Dynamics A mathematically consistent theory for evaluating the large solid body Lagrangian motions of the semi-immersed large solid body shown in Fig. 8.4
Fig. 8.4. Free body diagram for large solid body.
Deterministic Dynamics of Large Solid Bodies
637
was developed by John (1949, 1950). Extensions of this theory are given by Peters and Stoker (1957), Stoker (1957), Wehausen (1960, 1971), Newman (1961, 1977) and Mei (1989). The extensions by Newman (1961) include three separate perturbation parameters that represent the orders of: 1) the wave height/wavelength ratio (H/k); 2) the beam/wavelength ratio; and 3) the magnitude of the oscillations of the body. This less ambitious summary follows John (1949, 1950) with the exception of the orientation of the vertical z axis instead of the y axis; the negative gradient of the scalar velocity potential to recover the fluid velocity q = — V4>; and the addition of zero-order external constraints (Mei, 1989). There are two well-known spring stiffness constants for the hydrostatic stiffness due to the change in water plane area for semi-immersed large solid bodies. One of these two spring constants is the vertical heave mode X$(t) linear spring given by the well-known formula K& = pgA0, where AQ = the water plane area at equilibrium (e.g., AQ = nD2/4 for a vertical circular cylinder of diameter D). The second spring constant is the pitch mode ©5 (t) rotational spring constant for rotations about the origin of the body-fixed coordinates x' given by the well-known formula
where the second moment of area may be computed from
0A=
Jj(x'-x'R)2dx'dy', Ao
where 1^ , = rrD4/64 for a vertical circular cylinder and Ao = water plane r r
i i
area at equilibrium (i.e., t = 0). When these two well-known spring constants are derived from the fluid pressure given by Eq. (8.47) below, it is then possible to compare the rotational spring constant due to excess buoyancy at equilibrium K$5 with the spring constant due to changes in water plane area at equilibrium A55 for semi-immersed solid bodies (cf, the articulated tower in Chapter 7.8.1 where K^5 is neglected compared to K^5).
638
Waves and Wave Forces on Coastal and Ocean Structures
8.2.1. Kinematic Body Boundary Conditions (KBBC) The instantaneous wetted-surface of the solid body shown in Fig. 8.4 may be described by a Stokes material surface S(x, y, z, t) given by S(x,y,z,t)
= z-f(x,y,t)
= 0.
(8.25)
The solid body kinematic .Sody boundary Condition (KBBC) may be computed from the Stokes material derivative of the Stokes material surface (vide., Eq. (2.13a) in Chapter 2.2.10) according to DS of d
onS(x,y,z,t).
(8.26a)
The unit normal surface vector may also be determined from Eq. (8.25) by n= -^— =
±|VSI
dx
dy
2
(8.26b)
2
^[m +m +i'_
where the ambiguous sign of the modulus of the gradient vector ± | V 51 must be determined from whether the unit normal points out of the surface and into the fluid or into the surface and out of the fluid (Hildebrand, 1976, Chapter 6.12). The perturbation procedure involves expanding the scalar velocity potential <£>(x,y,z,t); the translational displacements X(t) of the center of mass x'c (x'c,y'c,z'c) of the solid body; the rotational displacements 0 ( 0 about an arbitrary point of rotation xR(xR,yR,zR); the external loads F(t) and M{t); and the Stokes material surface S{x,y,z, t) in a power series ordered by a perturbation parameter given by the wave amplitude/wavelength ratio e — kA = 2nA/X > 0 (cf., Chapter 6). Simple coordinate transformations relate the fixed inertial coordinate system x to the moving body-fixed, non-inertial coordinate system x' where x = x' at t = 0. Note that x may be a function of time; but x' is never a function of time. Solidbody rotations in the moving body-fixed, non-inertial axes x' that coincide with the fixed inertial axes x at t = 0 may be about any arbitrary point of rotation xR(x'R, y'R,z'R) shown in Figs. 8.2 at point R. Substituting the linearized Bernoulli equation for the total fluid pressure (i.e., static plus dynamic) into the body surface integrals evaluated over the equilibrium wetted-surface area of the immersed body o^(x, y, z, 0) yields: 1) zero-order
639
Deterministic Dynamics of Large Solid Bodies
e° equilibrium equations for the Archimedes buoyant force/moment F /M and 2) a first-order e 1 linearized dynamic pressure force/moment components Fp /Mp. The Archimedes buoyant force/moment F /M may be combined with the body weight force/moment F /M to give the hydrostatic pressure -> k
-> k
force/moment stiffness components F /M . Equality between two different forms of the kinematic boundary condition for an ideal fluid on the solid body boundary may also be demonstrated (Mei, 1989). The perturbation expansions are given by (cf. Chapter 6.2) $>(x,y,z,t)
+ e22®(x,y,z,t)
= €i®(x,y,z,t)
+•••
= JV„
(8.27a)
n=\
X(t) = €lX(t) + €22X(t) +-.. = J2 €\X(t)
(8.27b)
n=\
0 ( 0 = e i 0 ( O +e22®(t)
+--- = ^€nn@(t)
(8.27c)
n=\
z = f(x,y,t)
= of(x,y)
+€if(x,y,t)
+•••=
^2ennf(x,y,t)
n=\
(8.27d) P(x,y,z,t)
= op(z) +e\p{x,y,z,t) = ^j£nnP(x,y,z,t)
+€22p{x,y,z,t)
-\ (8.27e)
n=0
that imply that there is nofluidmotion for e = Oandop(z) = hydrostatic pressure. The Stokes material surface S(x, y, z, t) and the pressure P(x, y, z, t) all contain zero-order equilibrium components that are independent of time; viz., of(x, y) andop(z). The complex-valued scalar velocity potentials <&(x, y, z, t) represent only wave-induced fluid motions and, therefore, do not contain zero-order e°, steady-state, time-independent components. The coordinates in Fig. 8.2a to an arbitrary point of translation xc(xc,yc,zc) iQ the wave inertial coordinate system fixed at the still water level equilibrium position are equal to the moving, body-fixed non-inertial coordinates xc{x'c,y'c,z'c) at t = 0 (John, 1949). An arbitrary point of rotation expressed in the moving body-fixed non-inertial coordinates x'R(x'R,y'R,z'R) in Fig. 8.2a need not be
640
Waves and Wave Forces on Coastal and Ocean Structures
Fig. 8.5. Translation and rotation of a large solid body.
located at the body center of mass x'c(x'c, y'c,z'c) (cf., the articulated tower Chapter 7.8 and waves in a 2D channel Chapter 5.2). Again, x = x' at t = 0; and x may be time-dependent but 3c' is always time-independent. An expression for \f(x,y,t) in terms of the static equilibrium position of(x,y) is required in order to proceed further with the linearized problem (John, 1949). In order to obtain this required expression for i / (x, y, t) in terms of of(x, y), the moving body-fixed, non-inertial coordinate system x must be related to the fixed, wave inertial coordinate systems; and the Stokes material surface S(x, y, z, t) must be expanded in a Taylor series about the equilibrium position oS(x, y, z, 0) of the solid body in order to be able to equate terms of 0(e). It is important to remember that the goal of the following derivation is to replace the time-dependent location of the Lagrangian body at i / ( x , y, t) with the time-independent location at equilibrium at of(x, y). The coordinate transformations between the fixed, wave inertial coordinate system fixed at the still water level at the body equilibrium position x and the moving, body-fixed non-inertial coordinate system x' may be obtained by a linear combination of pure translation without rotation x(t) — x' + e\X(t) + 0(e 2 ) (vide., Fig. 8.5) plus pure rotation without translationR(t) = ei®(t) xr' + 0(e2) about an arbitrary point R in the moving body-fixed, non-inertial coordinates x' that is aligned with the fixed inertial axes x at t = 0 according to
x(t) =x' + €[iX(t) + 10(0 x (x'-x'R)] = 3c' + 6 i x ( 0 + O(6 2 ),
+ 0(e 2 ) (8.28a)
641
Deterministic Dynamics of Large Solid Bodies
that may be rearranged to X{t)-X'
= 6lX(0 + O(6 2 ) + 0(e2)
= e [ i £ ( 0 + !0(f) x (x'-xR)]
(8.28b)
with the following scalar components: x'-x(t)=
- e [ i X i ( 0 + i® 5 (0(z' - z'R) - i®6(t)(y'-
y'R)],
(8.29a)
y' - y(t) = - e [ i X 2 ( 0 + i©6(0(*' - 4 ) ~ i©4(0(z' - z'R)],
(8.29b)
z' - z(0 = - e [ i X 3 ( 0 + i © 4 ( 0 ( / - 4 ) - i®s(0(*' - 4 ) ]
(8-29c)
because 3c = x correct to 0(e). At equilibrium or, equivalently, at time t = 0, z(0) = z' and the Stokes material surface becomes oS(x,y,z,0)
=
z-of(x,y)=
(8.30)
z / - o / ( x / , / ) = 0. The Stokes material surface may be expanded in a Taylor series about the equilibrium body position o5(x, y, z, 0) by first expanding f(x,y, t) by Eq. (8.27d) to obtain S(x,y,z,t)
= z-
f(x,y,t)
« [z' - of(x',y')]
- eif(x,y,t)
+ 0(e2) + •••
(8.31a)
and then expanding S(x, y, z, t) by S(x,y,z, t)*0S
+ - V ( z ' - z(t)) + 1rr(x'dz ox' 6 nS , + ^r(y'y(t)) + • • •
x(t))
dy' *(z'-0f(x',y')) ^~f
dx'
+
(z'-z(t))
(x - x(t))
—
(y - y(t))
+ ••• .
dy' (8.31b)
642
Waves and Wave Forces on Coastal and Ocean Structures
Substituting Eqs. (8.29a, b) into Eq. (8.31b) and equating 0(e) terms gives the following first-order approximation for i f(x,y, t) in terms of of(x, y) (John, 1949, p. 24, Eq. (2.2.4)): e1 :
lf(x,y,t)
= iX3(t) +
l®4(t)(y'-y'R)-l®5(t)(x'-x'R)
- - ^ - { i ^ i W + i0 5 (O(z' - z'R) - 1 © 6 ( 0 ( / - y'R)) ri
-f
- -jriiXjit)
+1©6(0(*' - 4 ) - i©4(0(z' - z'R)}
= 1X3(0 - ofxiXiiO - ofyiX2(t)-
(8-32)
Substituting the temporal derivative of Eq. (8.32) into the .Kinematic .Body .Boundary Condition (KBBC) given by Eq. (8.26a) yields, correct to 0(e) dif(x,y,t) dt
=
diQdpf dx dx'
|
d^dpf dy dy'
3iO dz
= jX 3 (0 + i ® 4 ( 0 ( / - y'R) - i® 5 (0(*' - x'R) - -2L{xXi(t)
+ i©5(0(z' - zR) - i © 6 ( 0 ( / - y'R)}
~ - ^ 7 { l X 2 ( 0 + i© 6 (0(*' - 4 ) - 1©4(0(Z' - ZR)} ay (8.33) or
3$
A
— +X»n an
A
+ @»ii' = 0,
on
0 5(x,v,z,0),
(8.34)
that readily demonstrates the well-known equality between the .Kinematic .Body .Boundary Condition (KBBC) for 1) a boundary between a solid body and an ideal fluid given by a Stokes material derivative (vide., Chapter 2.2.10) and 2) the continuity of the normal component of the velocity of a solid Lagrangian body and the normal component of the velocity an ideal fluid (Dean and
643
Deterministic Dynamics of Large Solid Bodies
Dalrymple, 1991 orMei, 1989) i.e., 3
A
A
DS
\-X»n + Q»n
= 0=
,
on
0S(x,y,z,0),
(8.35a)
where the unit normal on oS(x, y, 0) is computed from Eq. (8.26b) by n=
H>/*,-o/y,l) 2
.
(8.35b)
2
- A Ao/, + o/v + l]
The scalar kinematic Body Boundary Condition (KBBC) in Eqs. (8.35) transforms the large Lagrangian solid body motions X{t) and 0 (t) to Eulerian fields by the vector dot product with the spatially-dependent unit normals and the body motions are now functions of both time and space in the KBBC.
8.2.2. Dynamic Body Boundary Condition (DBBC) Because the semi-immersed body in Fig 8.6 is free to respond to the wave induced pressure loadings, a dynamic boundary condition must also be prescribed in addition to the kinematic Body Boundary Condition (KBBC) given by Eq. (8.34). The three translational displacements \X{t) may be referenced to any arbitrary differential point of mass dm in the body shown in Fig. 8.6. The Lagrangian linear and angular momentum principles for a solid body may be
Fig. 8.6. Locations of 2 axes: 1) Inertial {x, y, z} and 2) Non-inertial {x', y', z'
644
Waves and Wave Forces on Coastal and Ocean
Structures
written as n
(8.36a)
x(t)dm = '(Ox where r{t) = (x-xR)
+ eiR(t) = r'(t) + erR(t) +
0(e2)
(8.36b)
ic(0 =x' + ehX(t) + iR(t)] = ehX(t) + i © ( 0 x r'] + 0(e 2 ); (8.36c) x(t) = eiX(0 = ehX(t) + i 0 (O x r'] + 0(e 2 )
(8.36d)
The notation in Eq. (8.36a) and in all subsequent equations for linear and angular momenta is an implied stacked notation with the three translational momenta stacked vertically above the three angular momenta and separated by a horizontal dashed line [ ]; i.e.,
n
EF2„(o n
£iW)l
X>„(0 (8.37a) n
I>2„(0 X>3Bc) . n
Careful consideration must be given to r(t) when transforming from linear momentum to angular momentum. In Fig. 8.6, r(t) may be seen to be the (8.37b) r(t)=? + €iR(t). vector sum of static r' and dynamic R(t) components according to
645
Deterministic Dynamics of Large Solid Bodies
Large Solid Body Lagrangian Inertia (Crandall et al., 1968, Ch. 3) The Lagrangian inertia of an homogeneous large solid body correct to O (e) becomes
d dt / / /
x(t)dm .'(Ox \X(t)dm
- / / /
r x (?) x r'dm + 0(e2)
"l[l{y.¥ m\X(t) Sijkm(r'c)j
\Xk{t)ei/
( + €
Sijkm(r'c)k
lO(0 Iff \r'.r'\dm-
i@j+3(t)ei
i&(t) • ffjV
r'dm )
(
m\X{t)ei
= e \msijk(r'c)j
(
\Xk{t)ei
msijk(r'c)k i0 7 + 3 (Oe; __ \,i,j,k / r v.i ©,-+3 ( 0 5 , - / ^ 1 ©j+3 (0?«
= 1,2,3,
(8.38)
where e = &A = the perturbation parameter; Sijk = the Levi-Civita permutation symbol (vide., Eq. (2.5) in Chapter 2.2.4); and repeated subscripts imply tensor summation. The total mass m and center of mass x'c of the homogeneous
646
Waves and Wave Forces on Coastal and Ocean Structures
large Lagrangian solid b o d y are defined b y III
dm = m,
x'c = m~ III
vm
x dm,
(8.39a,b)
vm
w h e r e Vm is the total v o l u m e of the h o m o g e n e o u s large solid body. T h e first(translational inertia) a n d second- (rotational inertia) m o m e n t s o f total m a s s v o l u m e for a large h o m o g e n e o u s solid b o d y are defined b y / / / r'jdm = \ \ \ (x' -x'R)jdm vm
= m(x'c -x'R)j
=m(r'c)j,
j = 1,2,3
vm (8.39c) (rc)i =x'c
-x'R,
(rc)i = y'c - y'R,
(rch = z'c - z R (8.39d-f)
'rt = Iff ^'^ = fff(X' ~X'R)i{X' ~X*)idm vm
vm 2
[(x' - x'R) + (y' - y'R)2 + (z' - z'R)2] dm
= jjf vm r r
l l
V2
r r
33
where I™, = mass m o m e n t o f inertia (Crandall et ah, 1968, p . 172) a n d where qr,
= jjf
r'/jdm = Jjj(x'
vm
x'R)jdm
vm
= I™x, -m(xR)i(x'c)j i,j
- x'R)i(x' -
= 1,2,3
-m(x'R)j(x'c)i
+m(xR)i(x'R)j (8.39h)
w h e r e I"], — products of mass m o m e n t s inertia (Crandall etal., 1968, p . 172); and where Eq. (8.39h) is a convoluted form o f the parallel axis theorem for a large solid h o m o g e n e o u s body. For a non-inertial axis x" fixed to the large solid h o m o g e n e o u s b o d y at the center of mass x'c, the parallel axis theorem
647
Deterministic Dynamics of Large Solid Bodies
for rotations about the body-fixed non-inertial coordinates xR is given by ==
x jll
v Q == (x
v
x D)
\^c
x"x'-dm = J I fix' - x'c)iix' -
vm
^ R)
x'c)jdm
vm =
Qx'. ~ m(xc">i(x'ch>
iJ = 1 ' 2 ' 3 -
External Forces/Moments F/M The total sum of the external forces/moments F/M may be evaluated by a variety of methods; but the method that most easily illustrates the spring constants K^ and K^5 is of primary interest. The most common method is to compute these spring constants from the total fluid pressure from Eq. (8.27e); i.e., hydrostatic opiz) plus dynamic eipix, y, z, t). The total sum of the external forces/moments include the following vectors: r n •
=
-P
F it)
+ •
'
J^Mnit)
' Pw '
1
( ->C • +
•
-» w M it)
P
M it)
1
F it) MCit)
where the superscript P = total Pressure forces/moments; the superscript W = total fTeight forces/moments and the superscript C = total mooring and anchoring Constraints. These forces/moments may be expanded and combined to obtain f - F
f ~R
1
F it)
F \t)
n •
=
• +
•
E
M it)
R
M it)
^
\PKit)} +| K
[M it)
f ->C
1
F it) • +
•
MCit)
n
where the superscript E = Exciting dynamic pressure force/moment on a fixed body; the superscript R = the hydrodynamic Restoring force/moment on a large Lagrangian solid body oscillating in still water (i.e., the wavemaker BVP).
648
Waves and Wave Forces on Coastal and Ocean Structures
Total Fluid Pressure Forces/Moments K(t)/M(t) The total fluid pressure loads may be computed from /
FP(t)
fjp(x,y,z,t) P
M {t)
I , \
\
1
hdS = -e
^ r(f)x /
oS
(
I
\p(x,y,z,t)
QS
ndS
V
rx
)
\
1
hdS
//OPCZ) y(f'
Fp(t)
+
€lX(t)+€1R(t))x/
F\t) • +
p
M (t)
(8.40)
•
b
M {t)
where o S is the wetted surface of the large solid body at equilibrium or at t = 0; and where the minus sign is required because the unit normal n is directed out of the body and into the fluid. This method includes the body force/moment ^b
- b
loads F (t)/M (t) through the hydrostatic pressure component op(z). The total fluid pressure is computed from the linearized Bernoulli equation and includes both hydrostatic e° and dynamic pressure components correct to the first-order 0(e); i.e., P(x,y,z,t)
= op(z) + €ip(x,y,z,t) = -pgz
+ ep—-
at
= -pg[of(x,y)
+ 0(e )
+ 0(e2)
+ €if(x,y,t)]
(8.41a) + €P-^-
+ 0(e2)
(8.41b)
at
by Eqs. (8.27d, e). The zero-order hydrostatic component op(z) represents a body force/moment contribution to the total sum of the external forces/moments. It is this body force component that will eventually contribute to the hydrostatic spring constants in Eq. (8.50a) below; viz., K&
and
K& + KL
649
Deterministic Dynamics of Large Solid Bodies
Dynamic Pressure Force/Moment The dynamic pressure force/moment loads may be computed by substituting Eq. (8.24a) for I<J>(JC, y, z, t) from Sec. 8.1 into Eqs. (8.41a, b) for \p(x, y, z, t) and then by substituting into the first curly bracket term {•} on the RHS of Eq. (8.40) to obtain (Fp{t)\
(
1
\ ndS
= -Re \MP{t))
y rx J
0S
(
1
\ wrfS
= -Re
V rx / /
1
\
ipoj / / A[4>w +4>s]
= Re
ndS
exp — icot
oS 3
6
/ , ?m0m + 2 J £lm4>ir, m=\ m=4
+ Re
ndS
i pay
/ /
3
exp — ittrf
6
/ , Hm4>m + 2 J ^fflfc .m=l
( J\t)
m=4
\ (8.42a-f)
+ 6
\iM (t)J -> E
—E
where i F {t)/\M
{t) = the .Exciting force/moment on afixedbody given by /
(i^v) = eRe £
\,M (f)y
1 \
'W/'
ipoj I I A[
ndS exp — ('a;/
(8.43a) (8.43b) (8.43c) = eRe
d S e x p — icot
i/ooj / / A[0„, +4>s] " / / •
oS E
\Mj (t)
V«/
(8.43d) (8.43e) (8.43f)
650
Waves and Wave Forces on Coastal and Ocean
Structures
The \F (t)/\M (t) = the Restoring force/moment on a large solid Lagragian body oscillating in otherwise still fluid (i.e., the wavemaker BVP in Chapter 5). Substituting Eqs. (8.2 lb,c) from Sec. 8.1 for each of the j unit nor-* R
-* R
mals and Eqs. (8.7c,d) for the velocities into Eqs. (8.42) for \F (t)/\M and expanding each
(8.44a)
0 7 = R e { 0 7 } + nm{0 7 }, -* R
(t)
-* R
the Restoring force/moment \F (t)/\M
{t) may be expressed as
xFfit)
\Mf{t)
3
J2ixm(t)fj (Re{^ }
+
m
6
iIm{4>m})dS
OS
3«
+ X ) i©m(0 [J(*B{4>m} + i \m{4>m})dS m=4
= eRe
icop
>J = 3
J2iXm(t)jj (Re{4> } + ilm{
m
oS
J2 i®«<') /7(Re{0 m } + i lm{4>m})dS m=4
dn
oS
(8.44b-g)
651
Deterministic Dynamics of Large Solid Bodies
where the two integral terms in the elements of each of the two 3 x 6 matrices in the RHS of Eq. (8.44b) represent typical matrix elements. Substituting the Lagrangian ordinary temporal derivatives of Eqs. (8.7f,h) given by iXj(t) = -ia>iXj(t),
j = 1,2,3,
l©j(t) =-ia>i®j(t),
j =4,5,6
into Eq. (8.44b-g) gives
iFfit) J =
\Mf{t)
3
E
j j Re{
,
m=l
E m=4
co j j \m{cj>m}d-^-dS
\Xm{t) +
j j Re{
LloS
l©m(0 +
j
\Xm{t)
oS
co Jj
Iml^^dS l©m(0
I o5
-epRe 3
E
//
Re{c/>m}—J-dS \Xm{t) + an
co Jj' \m{4>m}d-^-dS
\Xm(t)
m=l
E
jJ' Re{4>m)d-p-dS on
l©m(0 +
co J j'lm{<M^S l©m(0
m=4
(8.44h-m)
652
Waves and Wave Forces on Coastal and Ocean Structures
that may be reduced to
/ , i^mj 1 Xm (t) + ^mj \ Xm (t)] m=l
+ 6
m (') + ^mj 1 ®m (01
lF/(0
• m=4
= -Re
lMf{t) / , [Mm; 1 ^m (f) + A.my-1 X m (f)] m= l
+ 6
XI [iJ-mj 1 ®m (0 + \mj i ©m (/)] -m=4
(8.44n-s)
where the added mass coefficients iJLmj and the radiation damping coefficients kmj in Eq. (8.44n-s) may be computed from
fimj = P if o5
RelM-^-dS;
Xmj = pco jj
Im {
(8.45a,b)
oS
where, again, the square bracket terms [•] in the elements in each of the two 3 x 6 matrices in the RHS of Eqs. (8.44n-s) represent typical matrix elements. The restoring force/moment matrix in Eqs. (8.44) may be linearly decomposed into two matrices; viz., a matrix for the accelerations Xm(t)/@m(t) and the added mass fimj and a matrix for the velocities Xm (t)/®m (t) and the
653
Deterministic Dynamics of Large Solid Bodies
radiation damping Xm; coefficients according to
\Ff{t)
\Mf{t) •
xFfif)
•
1
lF|(0
2
•
•
• - eRe
-eRe
3
e
,
J=
-
•
•
4
lMf(t)
\M)(t)
5
•
•
6
3
6
^ Mm; 1 *m (t) + Y^ Hmj m=\ m=4
\®m(t)
= -eRe 3
6
XI V-mj\Xm(t) + ^2 Vmj\®m(t) m=\
3
m=4
6
/ , -^mj 1 ^m (t)+Y^kmjl®m(t) m=\
m=4
1 2 3
(8.45c) (8.45d) (8.45e)
-eRe 4
22 Xmj 1 Xm 0) + 22 ^mJ l ®m ('} m=\
m=A
5 6
(8.45f) (8.45g) (8.45h)
654
Waves and Wave Forces on Coastal and Ocean Structures
The added mass matrix elements in Eqs. (8.45) and the radiation damping matrix elements in Eqs. (8.45) may be transposed to the LHS of the dynamic equation of motion and coupled into the mass matrix and the damping matrix. Hydrodynamic Stiffness Fb(t)/Mb(t) Substituting the hydrostatic pressure component op(z) Eq. (8.41a) into the third integral in Eq. (8.40) yields
P\t) M\t)
oS
—pgz from
1
= P8
ff-
=
ndS. v
(8.46)
{F + eiA"(0+ *!*(*)}x
Substituting Eq. (8.31a) for z into Eq. (8.46) gives F (0 = pg
z
ndS -epg
M\t)
l f(x, y,t)\
r x -*Pg
r x
ffof(,x',y')l JJ
\hdS
\ndS. x
ix(0x
The gradient Theorem in Eq. (2.122b) in Chapter 2.6 converts the surface integral over dS to a volume integral over d V according to
llFidS = lll*FdVoS
Vo
where Vz = g3 (Hildebrand, 1976, Chapter 6.12); where Vo = submerged volume of the solid body at equilibrium; and where the horizontal projection dxdy of the wetted surface may be computed from dx dy = n-}d S =
dS
y/ofi + of? + l'
Deterministic
655
Dynamics of Large Solid Bodies
so that Eq. (8.46) becomes
F\t) M\t)
= M [[[I
\dV-epg [fifl
~€p8fff\J J
l
| hdxdy
)dv'
Vvx(z;x(0)/
where QA = the water plane area at equilibrium. Collecting order e terms for bodies that are symmetric at the still-water-level (SWL) gives F\t)
pgVoh
M\t)
yPgV0[(y' - y'R)ex + (*' - x'R)e2] (
[OA1X3(O + / r | l 0 4 ( O - / r | l 0 5 W ] ^
+ £pg
I$lX3(t) + I * 104(O - / I , 105(O / l x 3 ( 0 + /;l,i©4(0 - / l , 105(0
+
\
+ epg
6 (o-/Vi0 4 (ol2i + kVi0 L l 3 J r
r
kVi0 5 (o-/Vi0 6 (o L
r
3
r
2
+ epg _-VoiX2(Oei + VoiXi(0«2_
^
656
Waves and Wave Forces on Coastal and Ocean Structures
and combining zero-order e° and first-order e 1 terms gives: xF\t)
OF"
+ €
\M\t)
oM"
pgVoh pgVo [(y' - y'R)h + (*' -
x'R)e2\
- [oA iX 3 (0 + 7^184(0 - /^105(O]«3 + epg - fa 1*3(0 + fa . + 7?) 1©4(0 - 7* 105(0 - # 1 0 6 ( 0 " V iX (/)l e 0 2 X L r2
V r2r2
r
r r
3/
l 2
r
l
J
+ [ # i X 3 ( 0 + / 4 , i ©4(0 - fa, + iY) 105(0 + 7Vi06(O - v 0 iXi(ol h L rl
r
2rl
V rlrl
r
3/
r
2
J
•
(8.47) where # = / Y (*' - 4 ) , . r f * ' ^ ; J J0A
# r , = / / " (*' ' J
4),-(JC'
-
x'R)jdx[dx'2
J J0A
(8.48a,b) IVr, = \ \ \ (*' - 4)« dV = V0(x'B - x'R)i,
(8.48c,d)
and where x'B is the vector coordinate to the center of buoyancy B in Fig. 8.4. -w ^w Body Weight Force/Moment F /M The body weight force/moment due to gravity may be computed from ™ge3
- w M (0
p'c+eiX(0]
oF
-mge 3
»* oM
w
- w \M (0
mg(yc ~ y'R^i ~ mS(x'c - x'R)e2 +6
- m g [ i © 6 ( 0 ( 4 - X'R) ~ i®t(0(Zc - Z'R) " i^2(0]2i +mgh®5(t)(z'c - z'R) - i®6(t)(y'c - y'R) + iXi(0]? 2
(8.49)
657
Deterministic Dynamics of Large Solid Bodies
where (vide., Fig. 8.4) r'c = (xc ~ X'R) «i + (y'c ~ y'lt) *2 + (z'c - 4 ) & Stiffness due to Buoyancy and Weight The stiffness due to buoyancy and weight may be written as FK{t)
i
vb 1
=1
b
\ J\t) |
• +€
•
I oM
K
M (t)
• +
+e
•
.M W (0
\M\t) \FK(t) '
oFK •+ €
\MK{t)
oMK where o ^ K
(-mg + pgV0)h
1
[-mg (y'c - y'R) + pgV0 (y'c - y'R)] ex + [-mg (x'c - x'R) + pgV0 [x'c - x'R)] e2
(8.50a)
and where Pg [ A O I X 3 ( 0 + 7.|i© 4 (0 -
i^(0
> 8 ifrXrt)
+ [p g ( / £ , + 7rV) - mg (z'c - z'R)\ 0 4 ( 0
-pg I*Ai05(O lMK(t)
• '2
Irli@5(t)]e3
- [pg /rV - mg ( 4 - jc^)] 106(0
ei
p g / . | i X 3 ( 0 + PgI^r[ 104(0
[ p « ( / r | r ; +/ r V) " m g (Z'c- Z'«)] 105(0
-[pg/rV-mg(^-^)]i©6(0 (8.50b) External Constraining Forces/Moments F (f)/M (t) The total external constraining forces/moments represent both static and dynamic external constraints such as mooring line guys, tension leg tethers, anchors, etc. These external constraining forces/moments may also be
658
Waves and Wave Forces on Coastal and Ocean Structures
expanded in a perturbation series according to F (t) = 0F +€rF
(0 + f V w
+ -
= £ f " / ( 0 ,
(8.51a)
n=0
MC(t) = 0MC + eiM C (0 + e22MC(t) + • • • = J ] e"„M C (0.
(8.51b)
n=0 -*c
•* c
Zero-order, time independent constraints o-F /o-^ are required, for example, in a tension leg platform that has excess buoyancy (vide., the articulated tower in Chapter 7.8.1 and the two semi-immersed horizontal pontoons in Chapter 7.8.2 in Chapter 7.8). Collecting terms only to 0(e) gives oFC
FC(t) MC(t)
C
+ [?; x 0FC]
0M
iFC(t) (8.51c)
+€ C
C
C
iM (f) + [r'e x lJF (f)] + [ix(0 x 0F ] lFC(t)
oF oM°
•+•
(8.5 Id)
liMfw + iM^OJ
where C
= 0MC + [r'e x 0 F C ]
(8.51e)
MCx(t) = e\MC{t) + [r'e x e , F C « ]
(8.51f)
M°2{t) = [ei*(0 x 0 J F C ] + [ei*(0 x <>FC]
(8.51g)
0M
and r'e is(are) the position vector(s) from the point of rotation R in Fig. 8.4 to the point(s) where the external constraint(s) is(are) applied. If the point of rotation J?
Deterministic Dynamics of Large Solid Bodies
659
in Fig. 8.4 is the only single point of external constraint r'e in Fig. 8.4, then MC2{t) = [€lX(t) x 0 F C ] + [ei*(f) x 0 F C ] •
(8.5 lh)
Summary The Lagrangian dynamic equations of motion for a homogeneous large solid body are given by ~m~ _M_
•
iX(0 —
~K~ +€ _K
.iO(0. oF'
f l
— \+ 0M J f
r
_
\X(t)' — l&(t) r
->C i
-> F
l
,^(0
1^(0 — 0M -C
+ €|
iFC(t)
• — e
£
C
[iM (0. ~\
+e•
•
R
iM (t) (8.52a-c)
J
.lMfa ) + \M £(0.
(8.52d-f)
where the hydrodynamic stiffnesses in Eq. (8.50b) have been transposed to the LHS in Eqs. (8.52) and where the 36 element inertia matrix (Crandall et al., 1968) is given by
L.AU
where the first three rows of the mass matrix are
[m] =
m 0 0
0 0 0 m 0 —w(zc —z'fl) 0 m m (y'c — y'R)
m (z'c - z'R) -m (y'c - y'R)~ 0 m (x^ — x^) —m (x'c — x'R) 0 (8.53a)
Waves and Wave Forces on Coastal and Ocean Structures
660
and the last three rows of the mass moment of inertia are
0
-m(z'c-z'R) m (y'c - y'R)
m(z'c-z'R) m
- (y'c-y'i()
_
(4
-Ijj •2
-m{x'c-x'R)
0
m
\i"!, +ir r V V2
-Im r'r' r r
r r
V Vl
0
4)
I
3 3/
~'jj
3 2
W l
+H , (8.53b)
The corresponding 36 element stiffness matrix is given by "K" _K _
where the first three rows of the stiffness matrix are 0 0 0 0 0 0 0 0 0 0 [K] = 0 0 pgtf -pgrf ^33
0 0 0
(8.53c)
and the last three rows are M =
0
0^ 3
C
pgI$-0F% r 2
0
.+0^(4-4)-
-oF${y'e-y'R)\ r
04
r/4 r,r 2
-pgrf+oFC
~l
^ 5 5 + *55
-0Ff(x'e-x'R)
_-of 3 C (4-4).
tf
-off
0
-PS? +mg {x'c - x'R)
o^ & -4)
z
0^2 ( «
Z
~ 'R)
_+oFC (x'e-x'R)
r -p8V
2
+mg b e - y'R) c MFc,{y'e-y' R)
"- 0 f 2 (4-4) .-off ( 4 - 4 ) (8.53d)
and where the water plane stiffness elements in Eqs. (8.53c, d) are given by K^ = pg0A, K55 =P80r"
K& = pgl^,
KlA = Kv55=pgll-mg{z'c-z'R),
(8.53e,f) (8.53g,h)
661
Deterministic Dynamics of Large Solid Bodies -»E
-* E
where I"}, is defined in Eqs. (8.39g,h); \F (t)/\M (t) are defined in Eqs. (8.43); \FR {t) / XMR {t) are defined in Eqs. (8.44); and 1$, 1^, and /,? are defined in Eqs. (8.48). If the center of mass x'c and the center of rotation xR coincide {i.e., r'c = xc — xR = 0), both the mass and stiffness matrices are simplified significantly. Static Equilibrium Applying Archimedes buoyant principle, the weight of the large Lagrangian solid body plus the displaced weight of the fluid may be equated to the residual of the time-independent external constraints at static equilibrium (e°) to obtain
P8 z
ff
r li l
" 1 " 1
oF mge-i =
ndS-
/x.
(8.54) C C 0M + [P'e x 0F ]
that gives the following: Forces h • PgVo -mg = 0F3 ,
(8.55a)
Moments
h:pgll
- (y'c ~ y'R) mg = 0Mf+ [(y'c - y'R) 0 F 3 C - (z'c - z'R) 0 F 2 C ] , (8.55b)
h: -Pgtf
+ {x'c - x'R) mg = 0M2C + [(z'c - z'R) 0 / f - (x'c - x'R) 0 F 3 C ] , (8.55c)
h: 0 + 0 = 0M3C + [(x'c - x'R) o / f - (y'c - y'R) 0Ff] .
(8.55d)
-* c -c If there are no external constraints [oM = 0 and o^ = 0] and if x'c — xR = 0 the stiffness matrix is simplified further.
662
Waves and Wave Forces on Coastal and Ocean Structures
Two-Dimensional Applications Many applications do not require solving the Lagrangian dynamic equations of motion for afloatingbody in three dimensions (i.e., six degrees offreedom).An example is an asymmetric, semi-immersed body of revolution (Isaacson, 1982 or Hudspeth et al., 1994). In the two-dimensional case, translational motions are restricted to surge X\(t) and heave X^it) only, while the rotational motion is restricted to pitch &s(t) only. Because the equations of motion in Eqs. (8.52) for the three-dimensional case are written in matrix form, the conversion to the two-dimensional case is very simple. Because the sway X2 (t), roll ©4(f) and yaw @6(0 modes are neglected in the 2D case, the rows and columns associated with these motions in the matrices Eqs. (8.29) may be omitted. The reduced mass and stiffness matrices for 2D dynamics are m (z'c - z'R) -m (x'c - x'R)
0 m
m [m'] = 0 m (z'c - z'R)
(8.56a)
0 -pgIA
K
yi
[k'] = A
-PgI
K?5 + K} -oFf(x'e-x'R) l-oF3c(z'e-z'R)j]
(8.56b)
and may be simplified further if the point of translation xc coincides with the point of rotation*^.
8.3. Froude-Kriloff Approximations for Potential Theory The Froude-Kriloff hypothesis for estimating the wave-induced hydrodynamic pressure loads on small bodies formed the basis for the Morison equation force/moment loadings in Chapter 7. The linear decomposition of a 3D scalar velocity potential into an incident <$>w(x, y, z, t), a scattered ®s(x, y, z, t), and a radiated <$>r(x, y,z, t) wave potential for estimating the wave-induced loadings on large bodies also yields a Froude-Kriloff loading. In this case, the hydrodynamic pressure field induced by only the incident wave potential Q>w(x, y,z, t) yields a Froude-Kriloff load. This approximation forms the first estimate for the wave-induced loadings on large bodies; and may be applied to calibrate the magnitude of the diffraction effects on large fixed bodies. The scattered wave
663
Deterministic Dynamics of Large Solid Bodies
potential Q>s(x,y,z,t) results in a first-order correction to the hydrodynamic pressure field induced by the incident wave. If the diffracted wave load on a large fixed body is normalized by the Froude-Kriloff load, the magnitude of the scattered wave load is easily quantified. Because of this, all diffracted wave loads on largefixedbodies may be normalized by the Froude-Kriloff load due to only the incident wave pressure field. Examples in both Cartesian and circular cylindrical coordinates are reviewed.
8.3.1. Froude-Kriloff Load in 2-D Cartesian Coordinates The relatively large three-dimensional body shown in Fig. 8.7 is submerged with the depth of the center of gravity of the body located at z = —d. The effects of flow separation around the sharp corners are neglected. A normally incident, 2-D surface gravity wave in the absence of the body may be described by a scalar velocity potential given by &w(x,z,t) = Re{Acbw(x,z)exp-i(cot = Re \i
+ v)}
cosh kh(\ +z/h) exp i (kx — cot — v) > cosh kh coshfc/?(l + z/h) sin(kx -cot — v), (8.57) cosh kh
where the phase angle v in Eq. (8.57) is introduced for data analyses by Finite Fourier Transform (FFT) algorithms (vide., Chapter 9.2). The wave-induced fluid
nH
I
J f
JVV\.V*\.V*VJ
x SWLX7
1
~t'"7
IWWWW\
Fig. 8.7. Definition sketch for 2-D body in Cartesian coordinates.
664
Waves and Wave Forces on Coastal and Ocean Structures
velocity is computed from the negative gradient of a scalar potential according to q(x,z,t) =
-V®w(x,z,t)
The total linearized fluid pressure (viz., the hydrostatic ps(z) plus hydrodynamic p(x, z, t) in Eq. (4.47c) in Chapter 4.5) is computed from the linearized Bernoulli equation by P(x,z,t) = ps{z) + p(x,z,t) = -pgz + p at
f f coshkh(l+z/h) = — PS\z — Re A — expi(kx — cot — v) [ |_ cosh kh (8.58a) = -pg\z-A [
, ' cosh kh
cos(kx-cot-v)
.
(8.58b)
J
The net pressure loading on the body may be estimated from F(t) = - IJPhdS,
(8.59a)
where the unit normal n is directed out of the body surface and into the fluid and oS is the wetted surface area of the body in Fig. 8.7 at equilibrium at t = 0. The horizontal components of the surface area vector n\ dS in Eq. (8.59a) in the horizontal ±x -directions are nidS = ±dzdyei-
(8.59b)
Substituting Eq. (8.58b) and Eq.(8.59b) into Eq. (8.59a) for the horizontal components of the net hydrodynamic pressure load and integrating over the total vertical surfaces of the body in the horizontal ±x -directions gives the following dimensionless horizontal load component: rr^
Fl(T)
Spg(WBD)(A/h) sin kB sinh kD [1 = ko h sinh kd [1 — — (kh/koh) (kh/koh) coth coth kd] kd] sin(r sin(r + v), kD kB (8.60)
665
Deterministic Dynamics of Large Solid Bodies
where the deep-water wave number ko — co2/g and x = cot. The dimensionless horizontal load in Eq. (8.60) may be verified from the horizontal component of the Euler vector equation (3.41) in Chapter 3.3 that is given by du(x,z,t) dt
dFi(x,z,t),
(8.61a)
where u(x,z,t) =
d
=
COS(KJC — co t — v)
co
(8.61b)
cosh kh
and where du coshkh(l+ z/h) — = Agk sm(kx — cot — v). (8.61c) dt cosakh In order to obtain the net horizontal component, Eqs. (8.61) must be integrated over the fluid volume displaced by the large body at equilibrium Vo in Fig. 8.7 according to Fl(0 = fffdFM = / / / p3-^±dV, (8.62) Vo
Vo
where d V = dxdydz. The dimensionless load from Eq. (8.62) may be verified to be equal to Eq. (8.60). For body dimensions that are small compared to the wavelength A. (i.e., kB, kd and kD « ; 1), Eq. (8.60) reduces to FI(T) « kd(h/d)[k0d - l]sin(r + v).
(8.63)
For a semi-immersed body where d = D, Eq. (8.60) reduces to Fi(r) =
kh sinfcZ? / cosh2fcD - 1 2kD kB \
kh \ smh2fcD sin(r + v), koh )
(8.64a)
and for body dimensions that are small compared to the wavelength X, Eq. (8.64a) may be further reduced to Fi(r) « -kh(k/ko) sin(r + v).
(8.64b)
666
Waves and Wave Forces on Coastal and Ocean Structures
The vertical components of the surface area vector n^dS in Eqs. (8.59a, b) in the vertical indirections are h$dS = ±dxdyes.
(8.65a)
Substituting Eqs. (8.58b and 8.65a) into Eq. (8.59a) for the vertical components of the net hydrodynamic pressure load and integrating over the total horizontal surfaces of the body in the vertical ±z-directions gives the following dimensionless vertical load component: F 3 (T) =
F3(T)
Sy(WBD)(A/h) h sm kB sinh&D = h *o« suiakd [(k/ko) — cotakd] cos(r + v), A kB kD (8.65b)
where h/A = dimensionless Archimedes buoyant load (cf. Eq. (3.47b) in Chapter 3.5). Again, Eq. (8.65b) may be verified from the vertical component of the Euler vector equation (3.41) in Chapter 3.3 as dw(x,z, t) = dF3(x,z,t)-Y, (8.66) ' dt where y = pg and where d
COS(KX — cot — v).
dt cosh kh Integrating Eq. (8.66) over the fluid volume displaced b y the large body at equilibrium according to t) y dV (8.67) F3(,) = ///^3fe Z ,.) = / / / dw(x,z, dt vb v0 that verifies that the dimensionless load from Eq. (8.67) is equal to Eq. (8.65b).
667
Deterministic Dynamics of Large Solid Bodies
For body dimensions that are small compared to a wavelength k (i.e., kB, kd and kD < 1), Eq. (8.65b) reduces to h F 3 (r) « — + kh(kd - k0/k) COS(T + v). A
(8.68)
For a semi-immersed body where d — D, then Eq. (8.65b) reduces to h kh sinkB / , „, , fco^ . , ^, ^ , F 3 (r) = - + cosh2fcD - 1 - — sinh2£D COS(T + v) A 2kD kB \ kh ' (8.69a) and for body dimensions that are small compared to the wavelength k, Eq. (8.69a) reduces to F^ ( t ) «*
h A
(8.69b)
Ko/Z COS(T + V).
8.3.2. Froude-Kriloff Load in Circular Cylindrical Coordinates The relatively large, three dimensional vertical right circular cylinder shown in Fig. 8.8 is submerged with the location of the center of gravity C. G. of the body located at z = —dcG and the point of rotation located at R. Fluid flow separation effects around the bottom are neglected. In order to integrate the incident wave field potential around the circumference of the cylinder, the Cartesian coordinates for the horizontal plane (x, v) must be transformed into the circular
.OgTl
SWL
y
Fig. 8.8. Definition sketch for 3D body in circular cylindrical coordinates.
668
Waves and Wave Forces on Coastal and Ocean Structures
cylindrical coordinates (r, 9) by x = r cos 9 and y = r sin# and the outward pointing unit normal in circular cylindrical coordinates by n = cos 9e\ + sin 9Si. If £ = exp/(# + 7r/2) = exp(/^/2) expi9 = iexpi9 and f = kr, the generating function for Bessel functions Jm(t;) is given by (vide., Eq. (2.57a) in Chapter 2.4.3.1 or Hildebrand, 1976, p. 177)
exp[?(f - l/£)/2] = £ ^ ( 0 that gives the following Jacobi-Anger circular cylindrical coordinate expansion from Eqs. (2.57) in Chapter 2.4.3.1: gi8
e -(6
+
exp ikx = exp ikr cos 9 = exp ikr = ^ « ' m expO'm 9)Jm(kr)
'jfcr = exp
= ^ ( 2 - 5 m0 )J m
T
(*"-£):
cosm9Jm(kr),
(8.70)
m=0
where /«,(•) = Bessel function of the first kind of order m in Eqs. (2.46) in Chapter 2.4.3 and <5mo = the Kronecker delta function in Eq. (2.2) in Chapter 2.2.3. The incident wave potential defined by Eq. (8.57) in Cartesian coordinates may now be expressed in circular cylindrical coordinates by ®w(x,y,z,t)
= ®w(r,9,z,t)
= Re{A<)>w(r,9,z)exp-i(cot
.Ag coshfc/i(l + z/h) expi(kx = Re \i u> cosh kh = Re i
Ag coshM(l + z/h) co
cosh kh
Ag cosh.kh(\ + z/h) co
cosh kh
+ v)}
— cot — v)
exp i (kr cos 9 — cot — v) 00
m+l exp(im9)Jm(kr) Re ^2 i
exp - iicot + v) \
—oo
J
(8.71a) Ag coshfc/i(l + z/h) co
cosh kh (
00
x Re | £ ] ( 2 - Sm0)im+l lm=0
cosm9Jm(kr)
exp - i(cot + v)\.
(8.71b)
669
Deterministic Dynamics of Large Solid Bodies
The total linearized pressure is now given by Eq. (4.47c) in Chapter 4.5 as P(r,8,z,t)
= ps(z) +
p(r,6,z,t)
d
coshkh(l+z/h) cosh kh
m x R e y~] i sxp(im6)Jm(kr)
exp — i(wt + v) (8.72a)
z-A = ~Pg
xRe
coshkh(\+z/h) cosh kh
y ^ ( 2 - Smo)im cosm9Jm(kr) exp — i(cot + v) jn=0
(8.72b) The horizontal component of the surface area vector in Eq. (8.59b) in the xdirection is (8.72c)
«i dS = b cos 9d0 dz ei.
Substituting Eqs. (8.72b, c) into Eq. (8.59a) for the horizontal component of the net linearized pressure load yields the following dimensionless horizontal load component: ^l(T)=
*l(T)
pgnb2d(A/h) 2koh Ji(kb) (kd)(kb)
kh 1 — cosh kd + ( koh
) sinh kd sin(r + v), (8.73a)
where the deep-water wave number ko = co2/g and x = cot. For bodies that are small compared to a wavelength k,kb <§; 1, J\ (kb) ~ (kb)/2 (vide., Eq. (2.58a) in Chapter 2.4.3.2), Eq. (8.73a) reduces to Fi(r)
koh
Id
1 — cosh kd +
kh\
. sinh £ J sin(r + v). si
(8.73b)
koh J
For comparison with the MacCamy-Fuchs (1954) dimensional diffraction load on a full-draft vertical circular cylinder (d = h), the dimensional and dimensionless Froude-Kriloff load on a full-draft vertical circular cylinder where the
670
Waves and Wave Forces on Coastal and Ocean Structures
draft d = h from Eq. (8.73a) is Fi(r) = (
) J^kb)
kbkh
sin(T
+v>
( 8 - 73c >
knh J\(kb) Fl(T) =
Th ~lcir
(T + v)
(8-73d)
"
Similarly, the vertical component of the horizontal surface area vector in Eq. (8.59a) is n3dS = ±rdrdG e3-
(8.74a)
Substituting Eq. (8.72b) and Eq. (8.47a) into Eq. (8.59a) for the vertical component of the net linearized pressure load yields the following dimensionless total vertical load component on the bottom of the cylinder at z = —d: F3(z) =
p gn
_ h_ ~ A
+
b2d(A/h) 2(k0h)Ji(kb)
/
kh\
(kb){kd)
cosh k d — sinh k d cos(r + v). (8.74b)
For bodies that are small compared to a wavelength A. so that tt « Mkb) % (kb)/2 and Eq. (8.74b) reduces to h kh ( fk0h\ \ FI(T) « - H coshkd - - ^ - smhkd COS(T + v). A kd \ \kh J J
1, then
(8.74c)
If the draft d is also small compared to a wavelength A so that kd <£ 1, then Eq. (8.74c) further reduces to h F
3
I' kh
\
( T ) « - + ( — - * o A j c o s ( T + v).
(8.74d)
The moment about the point of rotation R in Fig. 8.8 may be computed from MR(t)
= - / // p(r,9,z,t)l p(r,9,z,>
x ndS,
(8.75a)
where d S for the vertical boundary of a vertical circular cylinder is given by dS = bd6dz.
(8.75b)
Deterministic Dynamics of Large Solid Bodies
671
If the point of rotation R in Fig. 8.8 is located on the vertical axis of symmetry z at r = 0, then the pseudo-direction cosines i x n may be computed from e\ I x n = b cos cos 9
e% b sin 0 sin 0
ej, z + dR 0
= - [ ( z + dfl) sin#]2i + [(z + d«) cos 6>]e2-
(8.75c)
Substituting Eq. (8.72b) and Eqs. (8.75b, c) into Eq. (8.75a) yields only one moment component in the y, ei coordinate direction that is due only to the dynamic pressure component p(r, 0, z, t) in Eq. (8.72b) and that is given by
Afs(f) =
pgAlnbh2 J\{kb) (kh)2 kdR tanh kh — 1 + k (d — dR) cosh kd[tanh kh — tanh kd] sin(o)t + v). (8.75d) + cosh kd(l — tanh kd tanh kh)
A dimensionless moment Ms(r) may be computed from M5(r) M5(r) pgnb d2 (A/h) 2
2khJi(kb) kbikd)2
kdR tanh kh — 1 + k(d — dR) cosh kd[tanh kh — tanh k d] sin(r + v). + cosh kd{\ — tanh &d tanh kh) (8.75e)
If the point of rotation R is located at the draft depth, then dR = d and Eq. (8.75e) reduces to M5(T)
=
2khJi{kb) kb(kd)2 x [kdtauhkh
— 1 + coshfcd(l — tanhfcrftanh kh)] sin(r + v). (8.75f)
672
Waves and Wave Forces on Coastal and Ocean Structures
For comparison with the MacCamy-Fuchs (1954) diffraction moment about the bottom dR = h on a full-draft vertical circular cylinder, the dimensional and dimensionless Froude-Kriloff moments on a full-draft vertical circular cylinder from Eqs. (8.75d, f) are MsW =
pgA2nbh2 J\(kb) ,,,,2 [k0h - 1 + sech kh] sin(cot + v),
(8.75g)
\KTI)
M5(r) =
2Jx(kb) , , , [k0h - 1 + sech kh] sin(r + v). kbkh
(8.75h)
Finally, if the cylinder radius b is small compared to the wavelength k then tt« 1 and J\ «« (kb)/2, and Eqs. (8.75g, h) reduce to pgAnb2 h M5(t) «s — [k0h - 1 + sech kh] sm(cot + v), (kh) [koh — 1 + sech kh] M5(r) « sin(r + v), 7r^-2 (kh)
(8.75i) (8.75J)
where Eq. (8.75j) is independent of the small cylinder radius b.
8.4. Diffraction by a Full-Draft Vertical Circular Cylinder There is a very simple closed-form solution for the diffraction of monochromatic simple harmonic linear waves incident on a vertical circular cylinder of constant diameter that extends over the entire fluid depth d = h in Fig. 8.8 in Sec. 8.3. The reason that the solution is so simple is because the water particle velocity in the radial direction r computed from the incident wave may be completely cancelled by a scattered propagating wave from the homogeneous diffracted wave kinematic body boundary condition (KBBC) on the cylinder by Eq. (8.12b) and, consequently, no evanescent eigenmodes are required to satisfy exactly the kinematic body boundary condition on the vertical cylinder. The theory is credited to MacCamy and Fuchs (1954) and there are also additional solutions for circular cylinders by potential theory that are given by Morse and Ingard (1968), by Mei (1989) and by Ingard (1990). The diffraction boundary Falue Problem (BVP) may be obtained from Eqs. (8.24a-n) in Sec. 8.1. A dimensional scalar diffraction velocity potential Q>D(r,6,z,t) for the full-draft circular cylinder shown in Fig. 8.8 with dR = d = h and a wave
673
Deterministic Dynamics of Large Solid Bodies
period T = 2n/co may be defined by a linear combination of the real parts of a linear incident wave potential
= Re[<j>D(r,9,z) exp - i(cot + v)} = Re{[0 r a (r,0,z) + & ( r , 0 , z ) ] e x p - i ( < w f + v)},
(8.76)
where the phase angle v in Eq. (8.76) is introduced for data analyses by Finite Fourier rransform (FFT) algorithms (vide., Chapter 9.2). The fluid motion may be obtained from the negative gradient of a scalar diffraction velocity potential ®D(r, 9, z, t) according to q(r,6,z,t)
= ur(r,0,z,t)er
+ ud(r,6,z,t)eg
= -VQD(r,0,z,t)
+ w(r,0,z,
t)e3
=Rs{-V4>D(r,0,z)exp-i(cot
+ v)}
= R e { - [ v 0 t u ( r , 0 , z ) + V 0 J ( r , 0 , z ) ] e x p - i ( a > f + v)J, (8.77a) where the three-dimensional gradient operator in circular cylindrical coordinates given by Eq. (2.10b) in Chapter 2.2 with ez = e3 is f 3 .,
vw
=| ^
/ 1 \ 3 _
+
(7)w*
3^1
+
ta«|w-
<8 77b)
'
The total linear pressure field P(r, 6, z, t) may be computed from the unsteady Bernoulli equation in circular cylindrical coordinates with the incident wave pressure given by Eqs. (8.72a, b) in Sec. 8.3 according to P(r,e,z,t)
= p(r,9,z,t) = P
+ Ps(z)
d
Pgz
at
= Re{-ipco
+ v)} - pgz
+ >s(r,9,z)]exp — i(a>t + v)} -
pgz. (8.77c)
674
Waves and Wave Forces on Coastal and Ocean Structures
The free surface elevation r](r,9,t) may be computed from Eq. (5.71f) in Chapter 5.3 according to i
d
rj(r,9,t) =
, g
z= 0
dt CO
= Re{-i—4>D(r,8,z)exp-i(cot + v)}, z = 0 g CO = Rel-i-[(/)w(r,0,z) +
z = 0. (8.77d)
The incident wave potential (pw(r,9,z) may be expressed by the FroudeKriloff velocity potential from Eqs. (8.71a,b) in Sec. 8.3 with the Jacobi-Anger expansionfromEqs.(2.57) in Chapter 2.4.3.1 in circular cylindrical coordinates as .Ag cosh k(z + h) (pw(x,z) = i ———expikx, co cosh k h
(8.77e)
A c a ^ . Ag COSh k(z + h) 4>w(r,6,z) = i —— expikr cosd co cosh k h oo
Agcoshk(z -^—^/ + h) ^m + 1,expO'm0)./ m+1 m (fcr) co cosh -oo oo
Ag cosh k(z + h) <^ +1 — — > (2-yim+lcosfflfl4(^), co cosh K « m=0 •*—'
(8.71a) ,.,.., (8.71b)
where 5mo = Kronecker delta function from Eq. (2.2) in Chapter 2.2.3. The BVP for the scattered velocity potential (f>s (r, 0, z) is (vide., Chapter 5.3) V2
r > b, 0 < 9 <2JT,
- h < z < 0,
d
(8.78a) (8.78b) (8.78c)
675
Deterministic Dynamics of Large Solid Bodies
lim
i Kn
\
d
d(j)w(r,0,z) dr
dr
r = b, 0 < 9 <2n, h < z < 0.
(8.78d)
(8.78e)
The solution to the BVP in Eqs. (8.78) may be computed from an orthonormal eigenfunction expansion from the following amplitude modulated circular wavemaker solution in Eqs. (5.86a and 5.79c) in Chapter 5.3: 00
00
(8.79a) where the orthonormal eigenfunctions are (vide., Eqs. (5.78) in Chapter 5.3) cosh Kn(z + h) n = 1,2,3,..., (8.79b) *„(Kn,z) = Nn where the propagating eigenvalue A^ = k for n — 1 and where the evanescent eigenvalues Kn = iKn for n > 2 so that the evanescent eigenmodes are defined from Eq. (8.79b) as Hi l
\
COS Kn(Z
+ h)
(8.79c) rT , n>2, N„ and where the dimensionless normalizing constants Nn are computed from *«(Kn,z) =
tf-/
"0
»*>*.* (i+ £)<((£)
2 k h + sinh 2k h 2Knh + sm2Knh AKnh
n=1 ,
(8.79d)
n > 2 and integer (8.79e)
provided that the eigenvalues k and Kn for n > 2 and integer are computed from koh =
\khtanhkh, n= \ —Knh tanKnh for n > 2,
(8.79f) (8.79g)
676
Waves and Wave Forces on Coastal and Ocean Structures
where ko = co2/g = deep-water wave number. The complex-valued Bessel functions of the third kind or, equivalently, Hankel functions of the first kind H„ (Knr) may be defined as (vide., Eq. (2.49a) in Chapter 2.4.3 or Hildebrand, 1976, p. 146, Eq. 92) H£\Knr)
=
Jm(Knr)+iYm(Knr).
(8.79h)
The orthonormal eigenfunction solution in Eq. (8.79a) satisfies exactly Eqs. (8.78a-d) from the BVP. The unknown coefficients Cmn may be computed in a best-least-squares sense from the inhomogeneous kinematic boundary condition in Eq. (8.78e). Substituting Eq. (8.71b) for >w(r,6,z) and Eq. (8.79a) for (ps(r,9,z) into Eq. (8.78e), equating terms of cosm0 and integrating over the dimensionless interval of orthogonality — 1 < z/h < 0 yields *~-mn —
Ag Ni co cosh kh
on\
dJm(kb)/d(kb) dH^\kb)/d{kb)'
(8.80a)
where the derivatives ofthe Bessel functions d(»)/d (kb) are evaluated at the cylinder boundary r = b and are dimensionless derivatives with respect to the argument kb of the Bessel functions. The diffraction velocity potential in Eq. (8.76) is now given by <J>o(r,6,z, t) = ~Re{(pD(r,6,z) exp - i(cot + v)} = Re{[4>w(r, 0, z) +
Ag coshfc(z + h) £](2-S m o)/ m + 1 cos(m0) co cosh kh m=0 Jm(kr) -
dJm(kb)/d(kb) dHJn\kb)/d(kb)
(8.80c)
l
Hi \kr)
Because the vertical circular cylinder of constant diameter lb extends over the entire fluid depth (or, equivalently, the total dimensionless interval of orthogonality — 1 < z/h < 0), only the propagating eigenmode W\(k,z) for n = 1 is required in order to satisfy the zero normal flow kinematic boundary condition in Eq. (8.78e).
671
Deterministic Dynamics ofLarge Solid Bodies The dynamic pressure p(r,9,z,t) in Eq. (8.80c) by Eq. (8.77c) is
p(r,0,z,t)
computed from the diffraction potential
= Re{-( pco^oir ,9, z) exp - i (cot + v)} PSA
coshfc(z + h) s-^ TT^TTT. Z ^ cosh kh
2
~ 8>"0)l COS (mG)
m=0
= Re
Jm(kr)
dJm(kb)/d(kb) dH^\kb)/d(kb)
H$\kr)
exp — / (cot + v) (8.80d)
and the free surface profile r\(r,B, t) in Eq. (8.77d) for the total diffracted wave is r)(r,G,t) =
1 3<& p (r,9,z,Q , R Bt
= Re{-i-(f>D(r,8,0)
z= 0 exp - i(cot + v)},
z-0
^ ( 2 - 5 m 0 ) « m cos (mO) = A Re
m=0
Jm(kr)
dJm(kb)/d(kb) l
//(1)
H^(kr)
dHi \kb)/d(kb)
m
exp — i(cot + v) (8.80e)
where, again, the derivatives of the Bessel functions d(»)/d(kb) are evaluated at the cylinder boundary r = b and are dimensionless derivatives with respect to the argument kb of the Bessel functions. The series in Eq. (8.80e) may be interpreted as a normalized wave profile and applied to compute the wave runup around the circumference of the cylinder (MacCamy and Fuchs, 1954 and Mei, 1989). Isaacson (1978b) extended the relatively simple approximation to Eq. (8.80e) given by MacCamy and Fuchs (1954) for the wave runup on a small cylinder to
V(b,6)
= Aj\
+ (kb)2 2 c o s ^ + l n ( | ) +
( ^ i
(8.80f)
where yE = 0.5772157 = Euler's constant from Eq. (2.41e) in Chapter 2.4.2.
678
Waves and Wave Forces on Coastal and Ocean Structures
Horizontal Pressure Force The horizontal pressure force in the +x direction of the incident wave propagation may be computed by substituting Eq. (8.80d) into Eq. (8.59a) in Sec. 8.3 and obtaining
Fl(t) = -hJ
bdO I pgAhb cosh kh
= Re
p(b,e,^,t\cosOd(^\ Wm(kb)
f
x ;
=Re
exp - i(wt + y) ^
dH^\kb)/d(kb)
cos mO cos Ode /
Jo pgAnhbkoh 2Sml
w
(2 - Smo)i"
m= 0
cosh(kh(l+z/h)d
W(kb)
i exp — i {cut + v)
dH$\kb)/d{kb)
(8.81a)
where the dimensionless Wronskian W(kb) is
Wm(kb) = Jmikb)
dH^ikb) dikr)
-Hml\kb)
dJm(kb)
2i
d{kb)
Ttkb
(8.81b)
that is independent of the 0m mode m; and where the dimensionless derivative of the Hankel function is dH%] (kb)/d(kb)
1 dH^\kb)/d{kb)
exp - ie„ Rmikb)
dHml\kb)/d{kb)
Rmikb) = ^[dJm{kb)/d{kb)f em = arctan
+ [dYm(kb)/d(kb)]2,
~dYm(kb)/d{kb)' dJm{kb)/d{kb)
(8.81c)
(8.81d) (8.81e)
r(2)/ where ffj (kb) = Hankel function of the second kind in Eq. (2.49b) in Chapter 2.4.3. The dependence of the diffracted forces and moments on the coefficient \/{kbR\(kb)) is shown in Fig. 8.9. Substituting Eqs. (8.81b, e) into Eq. (8.81a) reduces the horizontal force F\{t) in the +x direction of the
679
Deterministic Dynamics of Large Solid Bodies
incident wave propagation for m = 1 to
F i W =(
ApgAhbkph \
U(wWcos(
+v 61)
~
= C(W?) cos (cot + v — ei).
(8.82a) (8.82b)
In order to evaluate the contributions of the scattered potential to the total horizontal force, the amplitude of the diffraction force coefficient C(kb) in Eq. (8.82b) may be scaled by the amplitude of the Froude-Kriloff force in Eq. (8.73c) in Sec. 8.3 to obtain C{kb) =
C(kb) (2pg7Tb2Ak0h/kbkh)
Jx {kb)
nkbRi (kb)Ji (kb)
(8.82c)
that is illustrated in Fig. 8.9. The diffraction force amplitude C(kb) in Eq. (8.82b) may then be computed by simply multiplying the dimensionless diffraction force amplitude from Eq. (8.82c) that is illustrated in Fig. 8.9 by the Froude-Kriloff force amplitude in Eq. (8.73c) in Sec. 8.3. There are two significant points that are illustrated in Fig. 8.9. First, the value of the amplitude of the diffraction force Cikb) in Eq. (8.82b) is approximately = 2 for kb -> 0 that is equivalent to the inertia coefficient Cm in Eq. (7.16) in Chapter 7.3 for a circular cylinder. Second, the values of the amplitude of the dimensionless diffraction force amplitude C(kb) inEq. (8.82c) are approximately equal to unity between values of 1.8 < kb < 2.5 in Fig. 8.9; and this implies that the diffraction effects are negligible in this restricted range of values for kb.
Fig. 8.9. Parametric dependence of the MacCamy-Fuchs diffraction force coefficients and phase shift.
680
Waves and Wave Forces on Coastal and Ocean Structures
In contrast to the dimensionless Froude-Kriloff horizontal force in Eq. (8.73c) in Sec. 8.3, the diffraction force in Eq. (8.82a) is phase-shifted from the incident wave phase by the phase angle ei that is defined in Eq. (8.81e). Consequently, the diffraction force per unit length may be compared with the linearized wave force per unit length on small bodies where the linearized drag force per unit length is given by Eq. (7.60) in Chapter 7.6.4. The horizontal water particle velocity of the incident wave evaluated at the vertical centerline of the vertical circular cylinder at x = r = 0 may be computed from Eq. (8.71b) in Sec. 8.3.2 by u(z,t) = Re {
exp — i(cot + v) r=0 dr Agkcoshk(z + h) = cos(oot + v), co cosh kh
(8.83a)
and the horizontal water particle local acceleration of the incident wave evaluated at the centerline of the vertical circular cylinder at x = r = 0 may be computed from Eq. (8.83a) by ut(z,t) =
du(z,t) dt
+ h) = -AgkCOShkiz (8.83b) Sm(cot + v). cosh kh An inertial and a linearized drag diffraction force per unit length may be expressed in a form following Eq. (7.23) in Chapter 7.4 as dFx(
f't) = pnb2 (CMut(z, t) + CDLu(z, t)), (8.83c) dz where CM = an inertia coefficient for a diffraction force per unit length and CDL = a linear drag force coefficient that is due to radiation damping by the scattered wave and not to viscous fluid damping as in Chapter 7. The complex-valued amplitude of the diffraction force per unit length df\(b,z) may be computed by modifying Eq. (8.81a) to dfdb,z) dz
= ipoobf Jo An
-pgA2nb
(pD(b,6,z)cosdd9 , coshk(z + h) — Jm(kb) coshra
ApgAb cosh&(z + . kbR\{kb) coshM
exp - iei
dJm(kb)/d{kb) (1) " HX>(kb) dHln\kb)/d{kb) (8.84a)
681
Deterministic Dynamics of Large Solid Bodies
and the total diffraction force per unit length dF\ (b, z, t) is given by dFx(b,z,t) dz
= Re {I
{
\df\{b,z) exp —i (cot + v + ei) dz 4pgAb coshk(z + h) .,
1
—^-2 — exp -i (cot + v + ei) F kbR\(kb) cosh kh J ApgAb cosh£(z + h) _ . / r kbRi(kb) cosh kh {cos ei cos(cot + v) — sin ei sinfwf(8.84b) + v). Equating Eq. (8.84b) to Eq. (8.83c) and then equating the coefficients of the sin(») and cos(«) terms gives the following force coefficients (Mei, 1989):
4
sin.,
=4<,YIm,
2
n(kb) Ri(kb) =
4co
cos 6i
(kb)2R2(kb)
n
4codMkb)/dm
=
2
n (kb) Ri(kb)
TZ
(kb)2R2(kb)
Again, the analytically-computed force coefficients CM and CDL in Eqs. (8.84c, d) must be distinguished from the empirically-computed inertia Cm and linear drag Cdi coefficients in Chapter 7. In addition, the viscous drag coefficients Cdi are positive-definite, but the radiation damping coefficients CDL are both positive and negative.
Pressure Moment about the Bottom The hydrodynamic pressure moment about the vertical axis of symmetry at r = 0 and z — —h may be computed by substituting Eq. (8.80d) and Eqs. (8.75b, c) into Eq. (8.75a) from Sec. 8.3 with the vertical point of rotation evaluated at J« = h and obtaining
M5(t) = - h 2 f
n
bdO j
p(b,9,^,t\(l
ApgAh2b Ms(t) = ,,,,,.2 2 n /i TN [koh ~ * + kb(kh) Rx(kb)
secn
+ ^1
c o s
^\e2dU (o)t + v — e\)
= C(kb) [k0h - 1 + sech kh] cos (cot + v - e i ) .
(8.85)
682
Waves and Wave Forces on Coastal and Ocean Structures
In order to evaluate the contributions from the scattered potential to the total moment, the amplitude of the diffraction moment coefficient C(kb)[koh — 1 + sechkh] in Eq. (8.85) may be scaled by the amplitude of the Froude-Kriloff moment in Eq. (8.75g) in Sec. 8.3 to obtain, again, Eq. (8.82c); i.e., C(kb) =
C(kb)[k0h- I+ sech kh] (pgA27tbh2Ji(kb)/(kh)2) [k0h - 1 + sechkh]
nkbRi(kb)Ji(kb)' (8.82c)
8.5. Reciprocity Relationships Haskind's Theorem for Wave Forces on Fixed Semi-Immersed Lagrangian Bodies The total 3D velocity potential <£>(x,y,z,t) for the fluid motion with a 3D, semi-immersed Lagrangian solid body may be represented by a diffraction wave potential <&D(X, y, z, t) that is a linear combination of an incident wave potential ®w(x, y, z, t) plus a scattered wave potential <&s(x, y, z, t)(<$>D = 4>w + $?) pl u s a six degrees of freedom radiated wave potential <S>r (x, y, z, t) from Eq. (8.24a) in Sec. 8.1 according to &(x,y,z,t)
= <S>D(x,y,z,t) + <S>r(x,y,z,t) Aw[
= Re
3
+ico ^
6
%j
7=1
&j
7=4
(8.24a) where a> = 2n/T. Each of the spatial velocity potentials
\x\
\y\
-h < z < 0, (8.24b)
|x|
\y\ < oo,
z = —h,
|JC|
\y\ < oo,
z = 0,
(8.24e) (8.24c)
Deterministic Dynamics of Large Solid Bodies
683
where the deep-water wave number ko = co2/g. The scattered potential
= 0,
on the body equilibrium surface o 5
(8.86)
or d
d
(8.87) on the body equilibrium surface QS. dn dn For the radiated O r BVP, the six components of the Lagrangian solid body velocities of the forced oscillations (wavemaker) may be generalized as in Eqs. (8.7c, d) to CO
l©;(0
'b Qi
1,2,3 4,5,6
exp — icot
(8.88a-c) (8.88d-f)
The radiated wave velocity potential cbj may be determined from the kinematic boundary condition on the oscillating Lagrangian solid body boundary oS in six degrees-of-freedom from d
= Ke\co 'Sj
exp — icot:
IPJ J
nj(x,y,z) n'j(x,y,z)
1,2,3 J =
_4,5,6J (8.89a-f)
Each of the six radiated wave velocity potentials cpj (JC, v, z) in Eq. (8.24a) may be computed from the direction itj{x, y, z) or the pseudo-direction n'- (x, y, z) cosines for a Lagrangian solid body oscillating with six degrees-of-freedom (MDOF) according to d
-nj(x,y,z)\ -n'j(x,y,z)\
J =
1,2,3 4,5,6
(8.90a-f)
n\ = cos(9\,x),
surge,
(8.90g)
«2 = cos(^2, y),
sway,
(8.90h)
«3 = cos(#3, z),
heave,
(8.90i)
- z cos(6»2, y),
roll,
(8.90J)
n'5 = zcos(9\,x)
— x cos(^3,z),
pitch,
(8.90k)
n'6 = xcos(02,y)
— ycos(9i,x),
yaw,
(8.901)
n'4 = y
COS(9T„Z)
684
Waves and Wave Forces on Coastal and Ocean Structures
where Oj = angle between the unit normal n and the basis vector gy (vide., Fig. 8.3 in Sec. 8.1). The exciting forces/moments on a fixed solid body may be computed from a surface integral over the immersed solid body at its equilibrium position oS and the hydrodynamic pressure p{x, y, z, t); i.e.,
F/(0 Mf(t)
dS,
I =
1,2,3 4,5,6
(8.91a-c) (8.91d-f)
where the hydrodynamic pressure may be computedfromthe linearized Bernoulli equation by p(x,y,z,t)
=p
d<&l(x,y,z,t) dt
= Re{—picoA^e(x,y,z)
exp — icot} ,
I = w,s, 1,2,... ,6, (8.92)
where Ag = amplitude of the Ith radiated velocity potential <$>i(x, y, z, t). An integral for the hydrodynamic pressure force/moment on a fixed body QS that is independent of the scattered velocity potential cps may be computed from the following volume integral that relates the scattered wave potential
/// VsS/2(t)i ~
^^1dv=°-
(8.93)
Integrating once by parts and substituting Eq. (8.24b) gives: d
/ / 35
d(ps
dS = 0,
(8.94)
where the surface integral is to be evaluated over the following total surface dB that bounds the total fluid volume V shown in Fig. 8.10: dB = Sf + oS + Soo + S-n
(8.95a)
Deterministic Dynamics of Large Solid Bodies
685
Fig. 8.10. Surfaces Sj bounding fluid domain V.
and the integral over the total surface 3 B in Fig. 8.10 may be linearly decomposed into the following sum of integrals:
jj (.)dS = jj (.)z=0 dS + jj (*)dS + jj (.)r^oo dS dB
Sf
Soc.
QS
+ jj(.)z=_hdS
= 0,
(8.95b)
S-h
where the horizontal coordinates r 2 = x2 + y2 and the integrand (•) in each integral is
(•) = (Ps-
0/^—•
(8.95c)
on dn The integral across the horizontal, impermeable bottom boundary S-h vanishes by the bottom boundary condition (BBC) in Eq. (8.24e); so that, accordingly dQj d
//I
dZ
dz
\z=_h
Similarly, the surface integral across the free surface Sf also vanishes because the combined kinematic and dynamic free surface boundary condition (CKDFSBC) given by Eq. (8.24c) requires that d(ps ko
//I
686
Waves and Wave Forces on Coastal and Ocean Structures
so that Eq. (8.94) now reduces to
jfwiS = -//(.),
,dS.
Finally, the integrand on the vertical boundary in the far-field integral at infinity Soo vanishes due to the kinematic radiation boundary condition (KRBC) from Eq. (8.24g) in Sec. 8.1 for n = 2 and for a temporal dependency of exp — /cot; i.e., lim sfkr \ — r-*oo
ik | ^ = 0
\ dr
so that
//{
d
d
dS = 0
Consequently, Eq. (8.94) now reduces to
//{
84>j 4>s-r, on
d
(h-stds=Jf^ds-
(8.96)
The exciting hydrodynamic pressure forces/moments from the (fixed body) diffraction BVP for $D(JC, y, z, t) may be computed from the following surface integral over the fixed body surface boundary QS:
[Mf(f)J-
Re
ipcoAw exp— icot
/ /
{4>w +
[1,2,3,1 [4,5,6 J
\dS
(8.97a-f) The direction cosines m and the pseudo-direction cosines n't are equivalent to the normal derivatives of the radiated wave potential dcpj/dn from Eqs. (8.21 b, c). Substituting fromEqs. (8.21b, c)forthedirectionorpseudo-directioncosinesyields ME(t)\
=i (aAw
P
I
Ww+M-^-dSexp-icot
OS
ipcoAu
exp — icot,
I ••
f 1,2,31 |4,5,6
687
Deterministic Dynamics of Large Solid Bodies
Substituting Eq. (8.96) and the kinematic body boundary condition for the diffraction BVP given by Eq. (8.87) into the second integral gives Flit) Mf(t)
\ =ipcoAw
= iAwTl
II
exp — (icot),
fa
I =
dfao dS exp- - (icot) dn [1,2,31 I4,5,6j )
(8.98a-c) (8.98d-f)
where the generalized exciting force/moment amplitude per unit incident wave amplitude Tl may be computed from the integral
= p(
d(j>t d
n
°ff
/ / / Kv2<^ _4>jV24>*,}dv = j j U J ^ -
dB
where the integral over the entire boundary surfaces dB shown in Fig. 8.10 that bounds the fluid domain V is again given by Eq. (8.95b). Again, the integrals over S-h and Sf vanish by the BBC (Eq. (8.24e)) and the CKDFSBC (Eq. (8.24c)) giving:
//{^-^}'—//{^-^L«The surface integral over QS may now be replaced by an asymptotic surface integral according to •FtHt) \Mf(t)l~
=
—ipcoAy lHW w
"
= —ipcoAwTl
/ /
dfa
exp — (icot),
dcj)w fa-rdn
£ =
1,2,3 4,5,6
dS
exp — (icot)
(8.100a-c) (8.100d-f)
688
Waves and Wave Forces on Coastal and Ocean Structures
that is a remarkable result! The magnitude of the exciting force/moment on a fixed body may be computed without solving for the scattered potential 4>s or integrating over the body surface QS. The radiated wave potentials
8.6. Green's Functions and Fredholm Integral Equations There are three fundamental methods commonly applied to evaluate numerically the wave-induced pressure forces on fixed and floating large solid structures: 1) the Ursell source method (Ursell, 1950), 2) the Schwinger variational method (Black and Mei, 1969 and Black et ah, 1971), and 3) the Fredholm integral equation method (John, 1949 and 1950). Only the Fredholm integral equation method is reviewed in detail because the Ursell source method and the Schwinger variational method are generally restricted to special structural and fluid geometries that are separable in one of the 13 separable coordinate systems (vide., Morse and Feshbach, Chapter 5, 1953). This review of Green's functions for surface gravity waves follows John (1949 and 1950) and may also be found in Mei (1989). Alternative derivations may be found in Stoker (1957, Chapter 9), Wehausen and Laitone (1960) and Newman (1977). An excellent application of singular integral equations that is relatively easy to understand and that is applicable to surface gravity waves is given by Lamb (1932, Sec. 240-244). A very robust analyses of Green's functions with many physical applications is given by Courant and Hilbert (1966, Vol I and 1962, Vol II) and by Stakgold (1979). John (1950) derives four forms for Green's functions for surface gravity waves in both two-dimensions (2D) and in three-dimensions (3D). These eight forms of Green's functions for surface gravity waves may be applied as resolvent kernels in four forms of Fredholm integral equations (Guenther and Lee, 1996, Chapter 7-3). Green's functions and Fredholm integral equations may be considered to be analogous for boundary value problems to the impulse response function and the Duhamel convolution integral for initial value problems (vide., Chapter 2.5.3). Morse and Feshbach (1953, Chapter 7) derive the fundamental singular solutions for the Laplace equation in 2D and in 3D in terms of the position vector Rj(x,xo) shown in Fig. 8.11 where j = 2 for 2D and 7 = 3 for 3D
689
Deterministic Dynamics of Large Solid Bodies
P
o(*o)^-pP(x)
Fig. 8.11. Observer PQ(X, y, z) and source P(XQ, yo, ZQ) point vector definitions in 3D.
position vectors from the source point Po(*o) to the observer point P(x) and where y — yo = 0 for j = 2. The four fundamental Green's functions for 2D and for 3D are reviewed first; and then the four forms for the Fredholm integrals are reviewed.
Green's Functions in 2D (y — jo = 0) The 2D position vector ^2(^,^0) in Fig. 8.11 is given by (8.101a)
/?2(£,£o) = Rx2(x) - J?o2(*o) x = xe\ + zey, R2(x,xti)
xo = xoe\ + zo?3
= (x - xo)e! + (z -
zo)h
Jf 2 (x,*o)| = [(* - xo)2 + (z - zo) 2 ]
(8.101b,c) (8.101d)
1/2
(8.101e)
Morse and Feshbach (1953) derive the fundamental singular solution for the 2D Laplace's equation from the flux of fluid across the arc 0:2 of the circumference of a small disk of radius e in the vertical x-z plane according to 9G2(£,£o) 8r
dSe =
dG2(x,x0) dr
rd9 = o?2,
(8.102a)
where r = radial coordinate in the non-inertial circular cylindrical coordinate axis located at the center of the small disk of radius e. Integrating around the arc a2
690
Waves and Wave Forces on Coastal and Ocean Structures
of the circumference of the disk of radius e yields a2G2(x,xo) = ct2 In |J?2(x,3co)| G2{X,XQ)
(8.102b)
= In |-R2(*,*o)l
(8.102c)
where a2 = the arc on the circumference of the small disk of radius € in 2D. John (1950, p. 100) derives the following 2D integral form of the Green's function for deep-water h -> oo: G2(x,x0) = ]n(\R2(x,xo)\) +
ko + k
-k
cos (k{x — XQ) ) — exp
exp (k(z + zo))
&_
(8.103a)
V where ko = co2/g = deep-water wave number; and where the integral j> (•) denotes a Cauchy principal value integral (Morse and Feshbach, 1953, Chapter 9.2, and Hildebrand, 1976, Chapter 10.6). The corresponding 2D integral equation for finite water depth is (John, 1950, Eq. (A25), p. 99) ln(j?2(*,ico)h +ln^« 2 (Jc,3c 0 )h G2(x,xo) =
f°°
-2lnh•
cosh[k(z+h)]cosh[ic(z0+h)]cos\k\x-xo\\ ko cosh kh — k sinh kh
+
+ exp(kh)
dk k
(8.103b) where the position vector singularity ^2(^,^0) that is reflected below the horizontal bottom at z — — h is defined by R2(x,xo) = (x - xo)ei + (z + 2h + zo)h
(8.103c)
\R2(x,x0)\ = [(JC - * 0 ) 2 + (z + 2h + zo)2]
(8.103d)
and where the linear combination of the two logarithmic terms in Eq. (8.103b) satisfy exactly the kinematic bottom boundary condition (KBC) for 2D surface gravity waves that requires no flow across a horizontal impermeable bottom boundary given by Eq. (8.17b) in Sec. 8.1.
691
Deterministic Dynamics of Large Solid Bodies
Finally, John (1950, Eq. (A2), p. 100) expands the 2D Green's function in the following orthonormal eigenseries ty„(Kn,z/h) (cf, Eqs. (5.30) in Chapter 5.2): G2(x,h)
ij^(KnhrlVn(Kn,z/h)Vn(KH,zo/h)expi(Kn\x-xo\),
= 1
„=i (8.103e)
where *n(Kn,Z/h)
COShK 1+Z/h)
^
;
„ = 1,2,...
2kh+«&2kht
r° Nl = /
=
2
cosh Knh{\ + z/h)d(z/h)
where K\ =k;Kn
=
l
(8.103 g)
->
(8103h)
n =
2lc h
" +™2lc"h,
(8.103f)
n 2
= iKn for n > 2; and provided that
= koh = kh tarih kh =—Knh tan Knh;
n>2.
(8.103 i, j)
g
Green's Functions in 3D] (y—yo ^ 0) The 3D position vector ^3(3c,ico) in Fig. 8.11 is given by Rl(5i,xo) = % 3 ( x ) -i?o 3 tfo) x = xei + yh + zey, R3(x,x0)
= (x-
(8.104a)
*o = ^o?i + yoh + zoh
(8.104b, c)
xojei + (y - yo)h + (z - zo)h
(8.104d)
l*3(*,*o)l = [(* - *o) 2 + (v - yo)2 + (z-
zo) 2 ] 1 / 2
(8.104e)
Morse and Feshbach (1953) derive the fundamental singular solution for the 3D Laplace's equation from the flux of fluid across the sector 0:3 of the surface of a small sphere of radius e in 3D according to 3G3(3c,3c0)
3G3(Jc,Jc0) \r=e dSe =
dr
dr
2 • , . , ,Q \r=e r smfdfdO =013,
, c ,n* x (8.105a)
692
Waves and Wave Forces on Coastal and Ocean Structures
where r = radial coordinate in the non-inertial spherical coordinate system located at the center of the small sphere of radius e. Integrating around the section a3 of the surface of the small sphere of radius e yields (8.105b) R3(X,X0)
G3(x,*o)
1
(8.105c)
/-* -* ' l^3(*,*0)l
where a3 = section of the surface of the small sphere of radius e. John (1950, Eq. (A24), p. 98) derives the following 3D integral form of the Green's function for deep-water h ->• oo:
G3(x,x0) =
_
/»0O
+J
1 \R3(x,xo)\
[(tS)
exp
fa
+ Zo)
) J° [* (\*x,y(Xx,y,xo*j) |]] dk~ (8.106a)
where /o(«) = Bessel function of the first kind of order 0 (vide., Eqs. (2.46) in Chapter 2.4.3) and where the horizontal position vector Rxy (•, •) in the argument of the Bessel function is (John, 1950, p. 93) Rxy(Xxy,xoXj) = (x- xo)et + (y- yo)e2,
\RxAx^'x^)\
= [(* - xo)2 + (y- yo)]
(8.106b)
•
(8.106c)
The corresponding integral for finite depth is (John, 1950, Eq. (A9), p. 93) G3(x,x0) \R3(x,x0)\
+— \R3(x,x0)\
/•oo (^o + k) exp -kh cosh[ic(z + h)]cosh[fc(z0 + h)]Jo (k\Rxj>(xXty,xo )|) +2
(8.106d) where the reflected position vector singularity R3(x,x§) below the horizontal bottom at z — — h is R3(i,x0) = (x- x0)h + (y- yo)h + (z + 2h + zo)h, \h(x,x0)\
= [(x - xo)2 + (y - y0)2 + {z + 2h + z 0 ) 2 ] 1 / 2
(8.106e) (8.106f)
693
Deterministic Dynamics of Large Solid Bodies
and where the linear combination of the modulus of the two inverse position vector terms in Eq. (8.106d) satisfy exactly the kinematic bottom boundary condition (KBC) in 3D that requires no flow across a horizontal impermeable bottom boundary given by Eq. (8.17b) in Sec. 8.1. Finally, John (1950, Eq. (A16), p. 95) expands the 3D Green's function in the following orthonormal eigenseries tyn(Kn,z/h) (vide., Eqs. (5.30) in Chapter 5.2): 00
G3(x,x0) = i |
£(tf B Ar 1 * B (tfi.,z/A)*»(tf»,zo/fc)
x #0(1) (Kn [|jkj, (xXj, *o,,,)|]) ,
(8.106g)
where K\ = k; Kn = iic„ for n > 2; *„(•) are given by Eqs. (8.103f-j); H^\») = Hankel function of the first kind of order zero (vide., Eq. (2.64a) in Chapter 2.4.3.2); and provided that = k0h = khtaahkh = —Knhtan.Knh ,
n > 2,
(8.106h,i)
8 and where (vide., Eq. (2.54) in Chapter 2.4.3) r;r//0(1) (iKn Rxy (xX)y, x 0xj ,)|j = 2 K0 [Kn | ^ j , (xXJh ioX(V) J ,
n > 2, (8.106J)
where Ko(») = the modified Bessel function of the second kind of order zero (or Kelvin function, vide., Eq. (2.54) in Chapter 2.4.3). Loads on vertical, axisymmetric 3D bodies may be computed from 2D vertical plane integrals with an axisymmetric Green's function (Fenton, 1978 and Isaacson, 1982). Hudspeth et al. (1994) discuss in detail the convergence criteria for axisymmetric Green's function series. Fredholm Integral Equations Lamb (1932, Sees. 57-5 8) provides an excellent introduction to singular Fredholm integral equations that is reviewed below. More detailed derivations are given by Morse and Feshbach (1953, Chapter 8) and by Courant and Hilbert (1962, Vol II, Chapter IV). A treatment that is directed specifically toward engineers and physicists is given by Hildebrand (1965, Chapter 3). The Fredholm integral
694
Waves and Wave Forces on Coastal and Ocean Structures
equations reviewed here are boundary surface integrals over all or only part of the boundary surfaces dB that bound the entire fluid domain in either 2D or 3D as illustrated for a 3D domain in Fig. 8.11. The portions of the fluid boundary that are required to compute a scalar velocity potential
xeD
V2 G (x,x0) = amS(x - x0)S(y - y0)S(z - z 0 ),
(8.107a) x, x0 e D.
(8.107b)
where m = 2 or 3. Substituting Eqs. (8.107) into Green's second identity (vide., Eq. (2.123c) in Chapter 2.6) yields
V
dB
A, ff(JG rd(t>\^ = G + Jj{+-to- to)dS'
am
m=
J2for2£> l3for3D,
dB
(8.107c) where am =fractionof an arc on a circular disk about the singular point at x = xo for m = 2 in 2D or the fraction of the surface of a sphere about the singular point at x = xo for m = 3 in 3D. Equation (8.107c) is a Fredholm integral equation of the first kind or homogeneous Fredholm integral equation of the second kind (Guenther and Lee, 1996, Chapter 7-3; Hildebrand, 1965, Chapter 3.1; or Morse and Feshbach, 1953, Chapter 8.3). When the fundamental Green's function from Eq. (8.102c) in 2D or from Eq. (8.105c) in 3D or the deep-water integral Green's function Eq. (8.103a) in 2D or from Eq. (8.106a) in 3D is substituted into Eq. (8.107c), the integral must be computed over the entire boundary surfaces dB because the fundamental Green's function does not satisfy any boundary conditions except for the deepwater integrals where the integrals vanish as z ->• — oo. Black (1975a, b) applies Eq. (8.107c) to compute wave forces on large solid bodies.
695
Deterministic Dynamics of Large Solid Bodies
When the combination of reflected fundamental Green's functions plus an integral over all wave numbers k in Eq. (8.103b) in 2D and Eq. (8.106d) in 3D are substituted into Eq. (8.107c), the surface integral over the fluid boundary dB reduces to an integral over only the immersed surface of a solid body oS that is illustrated in Fig. 8.10. Finally, when the Green's function that is constructed from the orthonormal eigenfunction expansion in Eq. (8.103 e) in 2D or Eq. (8.106g) in 3D are substituted into Eq. (8.107c) as degenerate kernels (Guenther and Lee, 1996, Chapter 7-5), the surface integral reduces to an integral over only the immersed surface of a solid body oS that is illustrated in Fig. 8.10. The integral in Eq. (107c) is not a very efficient integral for computing wave forces on solid bodies. There are two other Fredholm integral equations that require computingfirstan auxiliary function from a Fredholm integral equation of the second kind and then computing the scalar velocity potential from a Fredholm integral equation of the first kind. The derivation that is reviewed here follows Lamb (1932, Sees. 57 and 58). The surface boundary integral over the entire fluid boundary 3 B in Eq. (8.107c) may be reduced to only a surface integral over the immersed equilibrium surface of the solid body QS by requiring that the Green's function satisfy certain boundary conditions on all of the other surface boundaries illustrated in Fig. 8.10. The surface boundary integral over dB may be linearly decomposed into separate surface boundary integrals over each of the four boundary surfaces identified in Fig. 8.10; viz.,
dB
dB
= [[(•)dS + dS+ [[(•)dS+ S-h
Soo
ff(»)dS+ Sf
ff(»)dS.
(8.107d)
0^
The boundary value problems (BVP) for both scattered <j>s and radiated <pr wave potentials are prescribed in Eqs. (8.24) in Sec. 8.1. The boundary conditions for both wave potentials are identical on three of the four surface boundaries in Eq. (8.107d) except for the inhomogeneous Neumann boundary conditions (Hildebrand, 1976, Chapter 9.2) that are prescribed on the boundary surface of the immersed solid body oS. If the Green's function G (X,XQ) is constructed so that it satisfies the same boundary conditions on the three boundary surfaces
696
Waves and Wave Forces on Coastal and Ocean Structures
that both
am
(8.107e)
oS
that may be transformed into a Fredholm integral equation of the second kind when the inhomogeneous Neumann boundary condition Eq. (8.24f) in Sec. 8.1 is substituted for dfa/dn. Lamb (1932, Sees. 57 and 58) derives formulas for two auxiliary functions from Eq. (8.107e) that are more computationally efficient. It is convenient to represent the inhomogeneous Neumann boundary conditions in Eq. (8.24f) on the immersed solid body boundary surface oS by the following generic notation:
dn
=
Uj=\
- ^ f , diffraction, ; = s, rij oxn'j, radiation, ;' = 1,2,3,4,5,6.
(8.107f) (8.107g)
The BVP may be separated into two domains; viz., an exterior problem that is defined inside the immersed solid body QS and exterior to the fluid domain V; and an interior problem that is defined inside of the fluid domain V and exterior to the immersed solid body QS. The potential for the exterior problem is defined asty(x) and the potential for the interior problem as
am
(8.108a)
oS
- / / (
an
dn
where the normal derivatives to the immersed solid body surface o S are defined by
M_-»W. dn
dn
(8,08c)
697
Deterministic Dynamics of Large Solid Bodies
Adding Eq. (8.108a) and Eq. (8.108b) and substituting Eq. (8.108c) yields (Lamb, 1932, Sees. 57 and 58) otm
G (^ V9"
+ ^)dS on J
+ [[ (4, - ir^dS. on JJ0s
(8.108d)
Lamb (1932) identifies two possibilities for computing the potential <j> from (8.108d) that depend on whether the boundary conditions on the surface boundary are inhomogeneous Dirichlet or Neumann conditions (vide., Table 2.3 in Chapter 2.5.1). Each of these two possibilities correspond physically to whether the tangential velocities or the normal velocities are continuous on the body surface. Continuous tangential fluid velocities: Source Density a (Lamb, 1932) If the tangential fluid velocities are continuous on oS, then \jr =
where a = the first auxiliary function. The source density method requires computing numerical solutions to two Fredholm integral equations that are singular integral equations. First, the source density auxiliary function a must be computed from the inhomogeneous Neumann boundary condition on oS given by Eq. (8.24f) / dG umOj -
= Uj,
(8.109b)
that is a Fredholm integral equation of the second kind and where § denotes a Cauchy principal value integral (Hildebrand, 1976, Chapter 10.15). Once the auxiliary source density function a is computedfromEq. (8.109b), then the scalar velocity potential may be computed from Eq. (8.108d) with
j = s, 1,2,3,4,5,6.
(8.109c)
698
Waves and Wave Forces on Coastal and Ocean Structures
Continuous Normal Fluid Velocities: Dipole Source Density /A (Lamb, 1932) If the normal fluid velocities are continuous on QS, then an
an
an
an
by Eq. (8.108c) and the first surface integral in Eq. (8.108d) vanishes. Consequently, the tangential fluid velocities are discontinuous; and the resulting fluid motion is caused by a surface distribution of dipole source densities /A that may be defined by fji, = ^-f,
(8.110b)
where fi = the second auxiliary function. The dipole source density method also requires computing numerical solutions to two Fredholm integral equations that are singular integral equations. First, the dipole source density auxiliary function fx must be computed from the inhomogeneous Neumann boundary condition on 0S given by Eq. (8.24f) / a2G ocmH-j -
(8.110c)
that is a Fredholm integral equation of the second kind and where
Deterministic Dynamics of Large Solid Bodies
699
8.6.1. Orthonormal Eigenfunction Expansion of Green's Function for 2D Wavemaker The 2D wavemaker theory reviewed in Chapter 5.2 focuses on the connection between the wavemaker boundary value problem (BVP) and the radiated wave potential given by Eqs. (8.21) in Sec.8.2 for dynamically responding large Lagrangian solid bodies. Consequently, it is appropriate at this point to construct a Green's function for a planar wavemaker in a 2D channel with the same orthonormal eigenfunctions given by Eq. (5.30) that are derived in Chapter 5.2 from the wavemaker BVP. In order to do this, it is first necessary to construct a BVP for a Green's function in a 2D wave channel. However, at this point it may be instructive to identify an analogy between the Green's function G(X,XQ) as a resolvent kernel in a Fredholm integral equation for a BVP and the convolution kernel h (x, £) (or impulse response function) in the Duhamel convolution integral for an initial value problem (vide., Chapter 2.5.3). The convolution kernel is constructed from the homogeneous solutions to the governing momentum equation for a damped harmonic oscillator in Eq. (2.92a) in Chapter 2.5.3; while the Green's function may be constructed from the homogeneous boundary conditions in a BVP. If the structural properties of a Lagrangian oscillator remain unchanged, the solution to a different loading requires only convolving the convolution kernel with the new loading because the convolution kernel is unchanged. Analogously, if the boundary geometry remains unchanged for a BVP, a solution for a velocity potential for new inhomogeneous boundary conditions may be computed simply by convolving the Green's function with the new boundary conditions in a Fredholm integral equation. The BVP given by Eqs. (8.24b-h) in Sec. 8.1 for either the scattered (f>s or the radiated (pr velocity potential consists of a homogeneous field equation given by Laplace's equation (8.24b) with an inhomogeneous boundary condition given by Eq. (8.24f) on the boundary of the Lagrangian solid body at the equilibrium position oS. The BVP for the Green's function may be constructed from Green's second identity in Eq. (2.123c) in Chapter 2.6. The result is that the BVP for the scalar velocity potential with an homogeneous second-order partial differential equation (Laplace's equation) with inhomogeneous boundary conditions is traded for a BVP for the Green's function with an inhomogeneous second-order partial differential equation (Poison's equation) with homogeneous boundary conditions.
700
Waves and Wave Forces on Coastal and Ocean Structures
It may not be easier to construct a solution for the Green's function BVP than for velocity potential BVP, but once the Green's function is constructed for a particular fluid domain, a solution for the velocity potential with different inhomogeneous boundary conditions on the body boundary may be computed simply by convolving the Green's function with the new boundary conditions in a Fredholm integral equation. Specifically, the focus of the Green's function BVP is on the boundary condition on the wavemaker surface boundary at QS and may compared by V2G = a2S(x - x0)S(z - zo), 3dG G dn - o ,
where U is given by Eq. (5.7e) in Chapter 5.2. The homogeneous Laplace's equation with an inhomogeneous Neumann boundary condition U on the wavemaker surface boundary for the wavemaker BVP is traded for the inhomogeneous Poison's equation with an homogeneous Neumann boundary condition on the wavemaker surface boundary (Courant and Hilbert, 1962, Vol II, Chapter IV or Stakgold, 1979, Chapter 3). The BVP for the Green' function for a wavemaker begins with the governing field equations that is Laplace's equation (8.107a) for the velocity potential and Poison's equation (8.107b) for the Green's function. Substituting Eqs. (8.107) into Green's second identity in Eq. (2.123c) in Chapter 2.6 and expanding the boundary dB over the entire 2D or 3D boundary of the fluid domain leads to Eq. (8.107d) that is given by
a2
= f
on J
JdB
(«)rf€(jc0) + I (*)^(*o) + f
+ f Wdlixo),
(m)d£(x9) (8.107d)
where the dimensions of a2 = [Length] and where the line integrals are integrated over the JCO coordinates. The procedure for constructing the boundary conditions for the Green's function BVP discussed in Sec. 8.6 may now be demonstrated from the four boundary surfaces by substituting the boundary conditions for the velocity potential into the four boundary surfaces on the RHS of Eq. (8.107d).
701
Deterministic Dynamics of Large Solid Bodies
The boundary conditions that are required for the velocity potential are given in the wavemaker BVP in Eqs. (5.7a-e) in Chapter 5.2. The first surface boundary on the RHS of Eq. (8.107d) is integrated across a horizontal, impermeable bottom boundary S-h that is given by /
Ud-^--Gdf)dl{x*)=
f °°4>(^-)
dx0
(8.111a)
by the bottom boundary condition in Eq. (5.7b) in Chapter 5.2. In order to make this integral vanish, the bottom boundary condition for the wavemaker Green's function BVP must be dG — = 0 , 0 < x < + o o , z = -h. (8.111b) dz The second surface boundary integral on the RHS of Eq. (8.107d) is integrated across the free surface boundary Sf that is given by f ( dG d4>\ J U—-G-£\dl(xo)
,
f+0° =]
(dG \
dx0
(8.111c)
by the combined kinematic and dynamic free surface boundary condition in Eq. (5.7c) in Chapter 5.2 where the deep-water wave number is ko = co2/g. In order to make this integral vanish, the combined kinematic and dynamic free surface boundary condition for the wavemaker Green's function BVP must be k0G = 0,
0<x<+oo,z
= 0.
(8.111d)
dz The third surface boundary integral on the RHS of Eq. (8.107d) is integrated across the vertical surface boundary S^ as x ->• +oo that is given by / U^- - G^-) JSoo\ dx dxJx^+00
dl(x0) = f 4>(^-~ J_h \dx
iKnC)
dzo /^+oo (8.1 lie)
702
Waves and Wave Forces on Coastal and Ocean Structures
by the radiation boundary condition in Eq. (5.7d) in Chapter 5.2. The minus sign in Eq. (8.1 lie) corresponds to the assumed negative sign for the temporal dependency of the wavemaker potential
iKnG\=0, J
x -+ +oo, -h < z < 0.
(8.111f)
The fourth and last integral in Eq. (8.107d) is integrated across the wavemaker surface boundary at equilibrium oS that is given by / ( V ^ - G ^ W i c o ) = / U^--Gu] J0S\ dx 8xJ J_h\ dx
Jx
dzo
(8-lllg)
=0
by the kinematic wavemaker boundary condition in Eq. (5.7e) in Chapter 5.2. In order to convert this integral into a Fredholm integral equation for the wavemaker potential, the kinematic wavemaker boundary condition for the wavemaker Green's function must be 9G
„ — = 0, dx and Eq. (8.107d) reduces to
(8.111h)
x = 0, -h < z < 0
a2(p --= - /
(8.1 Hi)
GUdzo-
J-h The BVP for the wavemaker Green's function is now complete and may be compared with the wavemaker BVP in Eqs. (5.7a-e) in Chapter 5.2; i.e., = 0;
V2G = -^8(x-x0)8 h
(-- * h)' \h
d
k0G = 0;
\-h
| 0 < x < +oo, 0 < x < +oo, z= 0
(8.112a,b) (8.112c,d) (8.112e,f)
Deterministic
703
Dynamics of Large Solid Bodies
lim I — -iKn^>\=0; x-»+oo I dz
lim
I
*->-+oo
d(p _ dx
dG dz
iKnG
dG ' dx
[-h < z
x —> + 0 0 ,
= 0;
-h < z < 0, (8.112g,h) (8.112ij)
The alternative to the Poison's equation (8.112b) for the wavemaker Green's function BVP is to solve an homogeneous Laplace's equation for the Green's function and specify a jump and a continuity condition (Stakgold, 1979). Because the boundary conditions for G in Eqs. (8.112d, f) are homogeneous and are identical to the boundary condition for (/> in Eqs. (8.112c, e), the solution for G may be expanded in the orthonormal eigenseries tyn(Kn,z/h) from Eq. (5.30a) in Chapter 5.2 and the jump and continuity conditions may be determined by integrating Eq. (8.112b). The wavemaker Green's function G(x, z : xo, zo) may be expanded in the following orthonormal eigenseries: G(x,z : x0, zo) = YlXr>(x
:x
°> zo)^n(Kn,z/h),
(8.113)
n
where the orthonormal eigenseries tyn(Kn,z/h) from Eq. (5.30a) in Chapter 5.2 is given by Vn{Kn,z/h)
=
cosh Knh{\
+z/h)
n=
l,2,...,
(5.30a)
where the nondimensional normalizing constant Nn is Nj = I cosh2 Knh{\ + z/h)d(z/h) =
2kh+smh2kh 4kh ' 2K„h+sm.2icnh 4K „h '
n = 1,
(5.30b,c)
provided that ^1 = k for n = 1, Kn = IK„ for n > 2, and that the eigenvalues are computed from koh =
khtanhkh, -Knh tan Knh,
where k0h = co2h/g (John, 1949 and 1950).
n=1 n > 2,
(5.30d) (5.30e)
704
Waves and Wave Forces on Coastal and Ocean Structures
Substituting Eq. (8.113) into Poison's equation (8.112b) yields V2G(x, z : XQ, zo) = -rS(x j
~2
~ x°^8 {j, ~ ~h) '
+ K;Xn(x : xo, zo) | Vn(Kn,z/h)
*
n
'
= ^(x-x
0
)*g-^).
(8.114a)
Multiply both sides of Eq. (8.114a) by a member of the orthonormal eigenseries ^m{Km,z/h) and integrate over the dimensionless interval of orthogonality - 1 < zjh < 0 to obtain d Xn(x:x0,zo) dxz
+ K2
x
.^
o^S(x _ h
zo) =
Xo)yn(KnfZo/h)
(8 114b)
Define the solutions for x < xo and for x > xo by the following symbols: X„ forO < x < xo,
X„ forxo < A: < oo.
(8.114c,d)
Integrating Eq. (8.114b) with respect to x, yields the following two ODE BVP's in the x coordinate: a An yx . * 0 . « w
,
K ! V
< , .
xo, zo) = 0,
(8.114e)
*o, zo) = 0,
(8.114f)
dx2
8>X>(x:2X0,zo)+K2K(x dX<(x:x0,zo)=^ dx lim *->-+oo
x = Q^
to-""
Z;CI
(8.114g) xo, zo) = 0.
(8.114h)
The required jump and continuity conditions, respectively, may be obtained by integrating Eq. (1.114b) once to obtain dX>(x : xo, zo) —-— 6x
dX<(x:x0, — ax
zo)
«21I( . „ ... = —Wn(Kn,zo/h), n
X>(x:x0,zo)-X<(x:xo,zo)
= 0,
x = x0,
/on>r\ (8.114i)
x = x0. (8.114J)
The solution to Eqs. (8.114e, g) is X<(x : XQ, zo) = an cosKnx,
(8.115a)
705
Deterministic Dynamics of Large Solid Bodies
and the solution to Eqs. (8.114f, h) is X>(x : xo, zo) = An expiKnx.
(8.115b)
The two coefficients an and An may be determined by substituting Eqs. (8.115a, b) into the jump condition in Eq. (8.114i) and the continuity condition in Eq. (8.114j), respectively, according to an co$K„xo — An expiKnxo -anKn
= 0,
s i n £ „ * 0 - iAnKn expiK„x0
(continuity),
(8.115c)
= — V„(Kn,zo/h), h
(jump). (8.115d)
The determinant of Eqs. (8.115c, d) is the Wronskian of Eqs. (8.115a, b) and is equal to —/. The solutions for the coefficients an and An from Eqs. (8.115c, d) are .a2^n{Kn,zo/h) an = i — Knh
exp i Knxo,
An = i
.a2^n{Kn,zo/h) — cos Knxo Knh (8.115e,f)
and the 2D planar wavemaker Green's function is G(x, z : xo, zo) .^n(^-n,<.0/n)^nK^n,<-/"J .Vn(K n,Z0/h)Vn(Kn,Z/h) Knh i nh ^.Vn(Kn,zo/h)Vn(KKn,z/h) a2 > i —— Knh
E
.„
„
exp i KnXQ cos Knx,
x < xo,
exp i Knx cos K„xo,
x > xo, (8.115g,h)
provided that ^ i = k;Kn = iicn and their values are computed from Eqs. (5.30d, e). The wavemaker potential may now be computed from Eq. (8.111 i) by ,o
c,z) = -— I G(x,z:0,zoW(zo)dzoG(x,z : 0,
(8.116a)
The critical element now for computing Eq. (8.116a) is to determine which of the two Green's functions in Eqs. (8.115g, h) is to be substituted into the integral. Because the Green's functions are symmetric with respect to z and zo, the decision
706
Waves and Wave Forces on Coastal and Ocean Structures
rests with whether x is greater than or less than XQ. Because xo = 0, then x > x$ and Eq. (8.115h) is the correct choice and Eq. (8.116a) reduces to
expiKnxf
Vn(Kn,z0/h)U(z0)d
(^) (8.116b)
The dimensional integral in Eq. (8.116b) is identical to the dimensionless integral Eq. (5.40) in Chapter 5.2 that computes the wavemaker potential coefficients Cn. Again, in an analogy to the convolution kernel in the Duhamel convolution integral for initial value problems in Chapter 2.5.3, the degenerate Green's functions in Eqs. (8.115g, h) may be constructed from a BVP with homogeneous boundary conditions. Once the Green's function is constructed for a given fluid domain, the coefficients for either the scattered or radiated velocity potentials may be easily computed from an integral over the equilibrium surface of the solid body QS for any new values for U.
8.7. Wave Loads Computed by the FEM The geometries of many ocean and coastal structures are not separable in one of the thirteen separable coordinate systems (Morse and Feshbach, 1953, Chapter 5). Consequently, the hydrodynamic wave pressure loads on these structures must be computed by numerical solutions of integral equations. The two most commonly applied numerical methods are the boundary integral Element Method (BIEM) and the finite Element Method (FEM). An example of the BIEM is given by Hudspeth et al. (1994) who compute numerically wave loads on vertically axisymmetric bodies from a Fredholm integral equation with a resolvent kernel that is an orthonormal eigenfunction expansion of an axisymmetric Green's function. Only the FEM is reviewed here. The fundamentals of the FEM are given by Zienkiewicz (1967), inter alios. Analyses by the FEM for wave interference between two horizontal floating circular cylinders in oblique seas is given by Leonard et al. (1983); for wave interference between multiple vertical surface piercing cylinders of arbitrary cross-section shapes with radiation boundary dampers is given by Huang et al. (1985a); and for 3D wave interference between multiple permeable and nonpermeable semi-immersed structures of arbitrary cross-sections with radiation
707
Deterministic Dynamics of Large Solid Bodies
boundary dampers is given by Huang et al. (1985b). These FEM analyses are reviewed below. Because the matrix equation for the dynamic response of a single solid body in six degrees-of-freedom given by Eqs. (8.52) must now be replaced by matrices for NB bodies, it is notationally convenient to assign all six degrees-of-freedom motions and the forces and moments to single variables instead of the two variables previously assigned to the translational and rotational modes and to the forces and moments in Sec. 8.2. The single subscripted variable assigned to the motion modes of the rath body is defined by R e f e e x p - ^ } = (Zj'"W' 1
'
j =
123
j
=4,5,6,
\njn{t),
'''
(8.117a)
where co = 2n/T, where T = wave period and where the complex-valued amplitude £,-„ is defined by £yi, = l£/i,|expi« / l I ,
7 = 1,2,3,4,5,6.
(8.117b)
The Exciting (E) and Restoring (R) forces and moments on the nth body are defined by
where the complex-valued amplitude / • ffnR)
= \ffnR)\
«P»"ft».
is defined by 7 = 1,2,3,4,5,6.
(8.117d)
The three boundary Falue Problems (BVP) for the incident wave potential ^w (x, y, z, t), for the scattered wave potential <E\ (x, y, z, t) and for the radiated wave potential <$>r (x, y, z, t) that are summarized in Sec. 8.1 must now be modified for multiple NB body interactions. The total dimensional scalar velocity potentials for the diffraction wave potential & (x, y, z, t) and for the radiated wave potential <J>r(x, y, z, t) in Eq. (8.15) in Sec. 8.1 are now given by
+
A\
L +££fi£$=i^B(*,:v,z)
J
j (8.118)
708
Waves and Wave Forces on Coastal and Ocean Structures
^ wi\JC\?A
ds,FS n*
Fig. 8.12. Definition sketch of Ng large solid bodies in 3D seas.
where NB = number of large bodies and where each of the lower-case scalar velocity potentials 4>i(x,y, z) for I = w,sn or jn are functions only of the spatial coordinates {x, y, z). For an incident linear wave propagating at an arbitrary oblique angle of incidence a in Fig. 8.12, the spatial incident wave potential 4>w (x, y, z) in Eq. (8.118) in given by cosh k (z + h) 4>w (x, y, z) = — ——— exp / (kx cos a + ky sin a), KQ cosh kh
(8.119a)
where the deep-water wave number &o = &>2 /g and the wave number k = 2n/k is the solution to the linear dispersion equation koh = kh tanh kh
(8.119b)
and where X = wavelength. Each of the dimensional scalar velocity potentials (pi{x,y,z) in Eq. (8.118) must be a solution to Laplace's equation (2.12a) in Chapter 2.2.9 that is given by V 4>i(x,y,z) = A
d24>e(x,y,z) dz2
I = w,sn, jn (8.120a)
and to the following boundary conditions for I = w, sn, jn:
dz — dz
= 0,
kofa = 0,
=
(8.120b)
z = 0.
(8.120c)
z
709
Deterministic Dynamics of Large Solid Bodies
The scattered and radiated scalar velocity potentials must satisfy a radiation condition given by lim Vkr — r^oo [ dr
[ik\
— )(j)i 2r J
= 0,
I =
sn,jn
(8.120d)
where the magnitude of the horizontal position vector is r 2 = x2+y2 inFig. 8.12. The body boundary conditions for the scattered and radiated scalar velocity potentials are ^
+
^ L
dnn
=
0on305„,
d(pjn = rijn ondoSn, dnn \L
"nm)
d
l,2,...,NB
(8.120e)
n,m — 1 , 2 , . . . ,NB
(8.120f)
n=
n,m — 1,2,
...,NB
(8.120g)
where Snm = Kronecker delta function (vide., Eq. (2.2) in Chapter (2.2.3)); and where itjn = the jth direction and pseudo-direction cosines for the nth body that are defined by Eqs. (8.1 and 8.4) in Sec. 8.1. The hydrodynamic pressure field may be computed from
p(x,y,z,t)
=Re
poo L
NB
6
n=\
j=\
exp — icot
(8.121)
where p = fluid mass density. The hydrodynamic exciting load on the fixed nth solid body in the j t h degree of freedom may be computed from R*{/£exp-iarf}
1 = Re poo A
/ / {fiw + ^fak)
\j =
l,...,6,
\n =
l,...,NB,
njndS exp — icot
(8.122a)
710
Waves and Wave Forces on Coastal and Ocean Structures
where 3o5„ = the wetted surface area of the nth solid body and where ff„ = the complex-valued hydrodynamic exciting load coefficient on the nth body in the y'th degree-of-freedom. The hydrodynamic restoring load on the nth solid body in the yth degree-of-freedom may be computed from NB
2
Re{/£exp-ia)f}=Re
6
pco ^2 ^
. .
Hjn / / (pmknjndSn exp - it doS„
f NB 6
(8.122b) where Ffnmk = the complex-value hydrodynamic restoring load coefficient on the nth body in the jth degree-of-freedom due to the motion of the &th body in the mth degree-of-freedom. The restoring load coefficient in Eq. (8.122b) may be separated into real and imaginary components according to F
fnmk
= P™2 / /
+ io)Xjnmk,
(8.122c)
doSn
where \Xjnmk = the added mass coefficient and'kjnmk= the radiation damping coefficient for the nth body in the y'th degree-of-freedom. Applying Green's second identity in Eq. (2.123c) in Chapter 2.6 to the radiation potential yields the following symmetric identities: r
_
R
_
pR
inmk ~ tmkjn^
_
\n,k=\,...,NB,
l^jnmk — H-mkjn, *-jnmk — A-mkjn,
i
[J,m=l,...,6. (8.122d,e) The dynamic responses of each of the NB bodies may be computed from the solutions to the following simultaneous equations of motions: -a)2(mjnmk + fijnmk)-ia)kjnmk
I
6
El
m=l
+ Efcf 1 0 - &kn) [-(02Hjnmk
= ffn,
\
i =
\j =
\'-'
6
'
l,...,NB,
+Ffnmk + Ffnmkj^jn
- lCoXj„mk + Ffnmk + Ffnmk\ $mk
(8-123)
711
Deterministic Dynamics of Large Solid Bodies
where Ffnmk = the hydrostatic stiffness coefficients in the stiffness force/moment loads Eqs. (8.50) in Sec. 8.2 and F
FEM Algorithm Solutions to the 3D BVP's for the scattered and the radiated scalar velocity wave potentials may be computed from the standard FEM. Because boundary dampers are applied in the FEM algorithm, the kinematic radiation boundary condition (KRBC) in Eq. (8.120d) may be applied at only moderate distances from the structures; and all of the closely spaced NB structures are enclosed within this reduced radiation boundary. Radiation boundary dampers, Bettess and Zienkiewicz (1977) The .Kinematic .Radiation boundary Condition (KRBC) in Eq. (8.120d) is applied at a finite distance rp according to
"^\^~(ik~2~r~)(pl}
= 0, l=zsnJn
-
(8 124)
-
Kinematic fictitious bottom boundary condition In order to avoid discretizing the fluid domain below the depth where wave dynamics and kinematics are negligible in deep-water applications, the kinematic bottom boundary condition in Eq. (8.120b) may be modified for the scattered and the radiated wave potentials to the following: d dz
ktanhk(z + h)
tn
= 0,
i = s,j,
(8.120b')
that may be applied at afictitiousbottom z = —hf where hf < h and that reduces to Eq. (8.120b) when z = -h. The FEM solutions may be computed from variational functional that are integrals over the entire fluid domain V, over the free surface boundary dSps, over the radiation boundary surface dSrD, over the bottom surface boundary dSh and over the NB body surface boundaries doS„. The FEM functional for
712
Waves and Wave Forces on Coastal and Ocean Structures
the scattered velocity potential BVP is given by
L
V
dSFs
4>2sdS - - / / ktemhk(z + h)
n(&) =
dS/,
dSr[)
+ :Ln l
- JJ
-<j)sdS
dnn
(8.125a)
and for the radiated velocity potential BVP by
Uff\(°th(^hmV-lff^ SSFS
n (*;„)
-\ fj («'* - ^ j ) 4>)ndS - l- fjkUahkiz + h)4>%dS dSrD
- I
3 S/,
These functionals are discretized into finite elements; and the unknown potentials are expanded in each finite element in a set of nodal unknowns tye and prescribed shape functions N" where q = \,...,NE where Ng = the total number of nodal points in each finite element. A typical element may be denoted by the superscript e. Expanding the functionals in Eqs. (8.125) in the nodal parameters ifr* gives the following combinations of discrete
713
Deterministic Dynamics of Large Solid Bodies
functionals:
n(&) =ni(0,) - n2(&) - n3(&) - n4(&) + n5(&) N
NV
NFS
q=\
q=\
rD
n
= J2 nf (&) - J2 K&) - E n 3^) q=\
N
hf
NB
NNB
q=\
n=\
q=\
(8.126a)
u{(j)jn) =nl{(t>jn) - n2(4>jn) - n 3 ( ^ „ ) - n4(>7„) - n5(
NFS
"rD
q=\
q=\
J= l
Nh
f
NB
n
NNB
-E 4<w-EEn5(u n—\
7=1
(8.126b)
o=l
where Afy = the total number of finite elements in the fluid domain V; Nps = total number of surface area elements on the free surface boundary; NrD = the total number of damper area elements on the radiation boundary at ro', Nhf = total number of surface elements on the bottom boundary at z = — h or on the fictitious bottom boundary in deep-water at z = — ft/ and NBE = total number of surface area elements on each of the NB structural boundary surfaces. Each of the five functionals for the scattered velocity potential 4>s{x,y,z) in Eq. (8.126a) is minimized with respect to the nodal parameters tye over the boundaries of each element according to
anf(&) ab
-III
dx
dx
dy
dy
dz
dz
VadV (8.127a)
9n|(&) ab
-SI
dS%
k0NeaNebradS
(8.127b)
714
Waves and Wave Forces on Coastal and Ocean Structures
(8.127c) dSS
rD
+ h)NeaNebfeadS
= ffktanhk(z
(8.127d)
3 st
II —K dS
(8.127e)
a
J Je dnB 3oS„
and for the five almost identical functionals for the radiated velocity potential
HI
ab
-IS!
dx
dx
dy
dy
dz
dz
VadV (8.128a)
= ff' k0NeaNebradS
(8.128b)
as*FS
= ab
.
are
ff(ik-£)NMr°dS
*n L -If
k tanh k(z + h)NeaNebfeadS
(8.128c)
(8.128d)
dSi
dne5(
Hi
ab
= / / * *bNebnjndS.
(8.128e)
doS<
The functionals in Eqs. (8.127a-d) for the scattered potential are identical to the functional in Eqs. (8.128a-d) for the radiation potentials; and only the inhomogeneous kinematic boundary conditions in Eqs. (8.127e and 8.128e) are distinct. The minimization and assemblage of the functional derivatives in Eqs. (8.127 and 8.128) follow the standard porcedures that are outlined
Deterministic Dynamics of Large Solid Bodies
715
in standard FEM texts and references (e.g., Zienkiewicz, 1967). The system matrices for the scattered and the radiation potentials, respectively, are given by [A]{<M = {Us},
[A]{
(8.129a,b)
The system matrix [A] in Eqs. (8.129) is symmetric and banded and is identical for both BVP's. Consequently, this system matrix need be assembled only once for both BVP's. However, this system matrix may be very large for 3D BVP's; and a very efficient Gauss elimination solution method with block form that uses secondary computer storage (Wilson et al, 1974) may be required for some computer systems. Semi-immersed discus buoy Huang, et al. (1985b) computed the dimensionless surge amplitude 2|£i|///; the dimensionless heave amplitude 2|§ 3 |/# and the dimensionless pitch amplitude |£5 \/{H/2b) for a semi-immersed discus buoy by the FEM algorithm given above and compared these three dimensionless motion amplitudes with data from Hoffman et al. (1973) and with the three dimensionless motion amplitudes computed from the BIEM by a 3D source Green's function (Garrison, 1975) and by an orthonormal eigenfunction expansion of an axisymmetric Green's function (Hudspeth et al, 1994). In the FEM algorithm, the fluid domain is discretized into 3D rectangular prism volume elements and 2D
1 .1875b
150.
±
TZ .225b
375b •
175b
h,
fictitious bottom
Fig. 8.13. Dimensions and FEM mesh for a semi-immersed discus buoy where b (h/b = oo;hf/b = 1.0; h\jb = 0.5) (Huang et al., 1985b).
2.0 ft
716
Waves and Wave Forces on Coastal and Ocean Structures
2.0
1.5
1.0
0.5
0. 0. 0.5 1.0
1.5 2.0 2.5
3.0 3.5
Fig. 8.14. Dimensionless dynamic response amplitudes ofa semi-immersed discus buoy (h/b = oo;hf/b = \.0;}n/b = 0.5)(— : 2 | £ 3 | / # & \!-5\/(H/2b) BIEM 3D source Green's function, Garrison, 1975; • & • : 2|f 3 |/ff, & \$5\/(H/2b) experimental data, Hoffman etal, 1973; + , • & • : 2II] \/H,2\h\/H, & \^\/(H/2b) BIEM axisymmetric Green's function, Hudspeth etal, 1994; A, o&D: 2 | ^ \/H,2\^\/H, & \^\/{H/2b) FEM, Huang et al, 1985b).
quadrilateral area elements. Two isoparametric finite elements with quadratic shape functions are applied, viz., a 20-noded prism and an 8-noded quadrilateral. The FEM mesh and buoy geometry are illustrated in Fig. 8.13 where the radius of the discus buoy b = 2.0 ft, the weight mg = 198.7 lb, and the polar mass moment of inertia / = 7.20434 slug-ft2. Cylindrical radiation dampers were applied at a distance rD = 1.3b and a fictitious bottom athf = b where b = radius of the discus buoy. Comparisons with data and the two BIEM algorithms are illustrated in Fig. 8.14.
8.8. Problems 8.1
8.2
Derive a Green's function for each of the two circular wavemakers analyzed in Chapter 5.3 by orthonormal eigenfunction expansion method applied in Sec. 8.6.1. Compute the maximum horizontal force and overturning moment about the bottom on a full depth vertical circular storage tank of diameter
111
Deterministic Dynamics of Large Solid Bodies
8.3
D = 300 ft in a water depth of h = 300 ft from a progressive wave with H = 45 ft and period 7 = 15 sec. The dimensionless roll mode equation of motion for a semi-immersed Lagrangian solid body is ' d2 _ d ~-j~2 +2^ + 2i; + <*4 [ 0 5 ( r ) = A5cojh sin(r - a0 - fa) 5co 5a>5h^r; 5h-rL dx dx where r — cot. Applying the method of undetermined coefficients from Chapter 2.5.6, the dimensionless roll mode rotation @5(r) is assumed to be given by @5(r) = | £251 sin(r - aQ - as). Prove that the dimensionless amplitude Q5 and dynamic magnification factor D [((Ss/,)-2] are \Q5\=B^(co5hr2'JA5co25h;
2
2 2
D[(^r ] = |[i- ( ^r ]
+
(gy
-1/2
and that the sway mode phase angle a$ is tan 0:5 = 8.4
tan /S5 + tan e$ 1 - tane 5 tan£5 J '
tan 65 =
(2£5/<W5fc) [l - (co5h)~2]'
Prove that the added mass coefficient fiji and radiation damping coefficient kji are given by
ff
d
ff
30;
\Ljt = P
8.5
Re{(f)j}—-dS; kjt = pco / / lm{<j)j}—dS. JJ0s dn JJ0s dn In the articulated tower analysis in Chapter 7.8.1, the pitch mode freesurface spring constant K$5 was neglected compared to the pitch mode excess buoyancy spring K^5. Compare the ratio K^5/K^5 of the pitch mode free-surface spring constant to the pitch mode excess buoyancy spring for a large homogenous circular cylinder fixed to the bottom that are given by
*55 = Pgrfy, 5 < 4 = ^ 5 = S^ ~ mg(z'B - l'R)
Chapter 9
Real Ocean Waves
9.1. Introduction Kinsman (1965, p. 386) is careful to point out that a valid specification of real ocean waves must integrate the following three concepts: 1) Fourier and spectral analyses of random processes, 2) probability theories applied to stochastic processes, and 3) hydrodynamics. Techniques from the first two concepts that do not depend on the physics of the hydrodynamic processes are available for analyzing real ocean waves. However, only those techniques from concepts that may be related rigorously to the physics of the hydrodynamics of real ocean waves are reviewed. The theoretical techniques reviewed are applicable to stationary ergodic processes and are limited strictly to short term statics. Isaacson and MacKenzie (1981) give an excellent review of long term statistical and probabilistic techniques applied to real ocean waves. The significance of the stationary ergodic hypothesis is that the ensemble average E[x(t\)] at the same time t\ shown in Fig. 9.1 of an infinite number of finite length time series x\(t\),X2(t\),x-s(t\), ,*oo(^l) is equivalent to the temporal average over all times shown in Fig. 9.2 of an infinitely long single time series X\(t\),X\(t2),Xl{tl),....,X\(tooY,i-S;
E[x(tn)]=
lim — I TR^OO
1R
719
JO
R
xi(t)dt.
(9.1)
720
Waves and Wave Forces on Coastal and Ocean Structures
Xjrft)
-A/l/yAA^JjV/ x3(0
x2(t)
-X1
xtf)
'
!/ *
/
/
Fig. 9.1. Two ensemble averages of finite length records *,- (t) at times t\ and ti (Bendat and Piersol, 1986).
Xrft)
Fig. 9.2. Temporal average of a single time series x\ it) of infinite length.
9.2. Fourier Analyses Definitions of Fourier coefficients Fourier coefficients are defined separately for deterministic and for nondeterministic (or, equivalently, random) analyses.
721
Real Ocean Waves
Deterministic: Fourier coefficients are those coefficients that provide the best least squares fit to the data. Non-Deterministic: Fourier coefficients are those coefficients that explain the contribution by each frequency to the total variance of a random process. When both time and frequency are continuous, the Fourier transform pairs are given by the following integrals:
r+0°
i
r)(t) = —f==. 'In j — /oo J—c +
i
F(a))exp±(i(ot)d(D,
(9.2a)
t](t)exp^(icot)dt,
(9.2b)
r °°
F(a>) = - = V2TT J—OO
where u> = 2nf = radian frequency; where F(co) = Fourier transform of the time series r){t); and where the plus + sign in Eq. (9.2a) must be paired with the negative — sign in Eq. (9.2b). Both choices for the ± signs in the arguments of the exponential functions exp(«) in Eqs. (9.2) may be found in the literature; and the placement of the normalizing constant 2n is also arbitrary (Lighthill, 1964). If the frequency / is given in Hertz, the normalizing constant 2n does not appear and Eqs. (9.2) reduce to
+00
/
F(f)exV±(i2nft)df,
(9.3a)
r,(t)expT(i2nft)dt.
(9.3b)
-00
/+0O -00
Data records that are continuous in time are termed time series; and data records that are digitized to discrete values of time are termed time sequences. Modern Fourier analyses employ discrete/inite Fourier transform (FFT) algorithms that are designed to take advantage of high speed digital computers. Discretization of Eqs. (9.3) requires a finite record length TR and discrete
722
Waves and Wave Forces on Coastal and Ocean Structures
values of both time and frequency; i.e.; t ->• t„ = nAt,
f ->• fm=
mAf,
co -» com = 2nfm =
InmAf, (9.4 a-c)
where At(=T[t/N) and Af(=l/NAt) are constant fixed temporal and frequency intervals, respectively, of N total discrete values of time and frequency. A discrete Fourier transform pair may be approximated from Eqs. (9.2) by N-l
rj(n) = A / ^
F(ra)exp ± {ilizrimAf At),
n = 0,1,2,... ,N - 1,
m=0
(9.5a) N-l
F(m) = At y ^ rj(n)exp ^(UnnmAtAf),
m = 0,1,2,.. .,N — 1,
n=0
(9.5b) where the total number of discrete time and frequency values are equal to ./V; and where the FFT coefficients F(m) are complex-valued quantities. The mean value of the discrete time sequence r\ (n) is given by the real-valued FFT coefficient ^(0); and the FFT coefficient F(N/2) is also real-valued at the Nyquist or folding frequency fy = (N/2)Af in Eq. (9.5a). The discrete positivedefinite frequencies fm > 0 in Eqs. (9.2 and 9.3) are represented by the indices 1 < m < N/2 — 1; while the discrete negative-definite frequencies fm < 0 in Eqs. (9.2 and 9.3) have the indices N/2 + 1 < m < N - 1. The complex-valued FFT coefficients F(m) at these discrete negative-definite frequencies are the complex conjugate values (denoted by superscript asterisks *) of the FFT coefficients at the corresponding positive-definite frequencies; i.e., F(N -m)
= F*(m),
N 1 < m < — - 1.
(9.5c)
Both of the summations of the series of discrete values in Eqs. (9.5) begin with a zero index n = m = 0 that are compatible with modern digital computer algorithms. However, older FORTRAN versions of the discrete algorithm
723
Real Ocean Waves
Eqs. (9.5) did not permit zero indices and Eqs. (9.5) were given by only positive-definite indices m, n > 0 according to N
rj(n) = A / ^ F ( m ) e x p ± ( / 2 ; r ( M - l)(m - 1 ) A / A 0 , n =
l,2,...,N (9.6a)
N
F(m) = At^t](n)exp
=F(/2jr(n - \)(m - \)AtAf),
m = 1,2,...
,N.
(9.6b) The mean value of r)(n) in Eqs. (9.6) is given now by the real-valued FFT coefficient F(l); and the real-valued coefficient F(N/2 + 1) is now at the Nyquist or folding frequency /N = (N/2 + 1)A/. The discrete positivedefinite frequencies fm > 0 in Eq. (9.5a) now have the indices 2 <m< N/2 in Eq. (9.6a); while the discrete negative-definite frequencies fm < OinEq. (9.5a) now have the indices N/2 + 2 < m < N in Eq. (9.6a). The complex-valued FFT coefficients F(m) at these discrete negative-definite frequencies are the complex conjugate values of the FFT coefficients at the positive-definite frequencies; i.e., F(N + 2-m)
= F*(m),
2 < m < —.
(9.6c)
A normalizing constant CN appears in all modern FFT algorithms that is related to the constant discrete intervals At and A / . Just as there is no standard convention regarding the normalizing constant 2n in the transform pair in Eqs. (9.2) for continuous time series (Lighthill, 1964), there is also no standard convention for where to place the normalizing constant CN in the FFT transform pairs in Eqs. (9.5) or Eqs. (9.6) for discrete time sequences; and it must be determined uniquely for each FFT algorithm before applying them. To illustrate how this normalizing constant CN is related to A? and A / , consider the geometric sequence
sN = i + z + z2 + N-l
«=0
3 z
+ ... +
N 1 z
-
724
Waves and Wave Forces on Coastal and Ocean
Structures
that may be reduced to the following compact result: -;\-zN
,
z#l
, N If z is a complex-valued variable given by
z= l
SN
—
\-z
(9.7b) (9.7c)
z = exp ± {i2n(m — m)/N),
(9.8)
where m and m are integer constants; then Eq. (9.7a) becomes N-l
SN = ^
exp[±i2jr(#n -
m)/Nf
(9.9a)
1 — exp ± \i2n{m — m)\ 1 — exp ± [ilTz{m — m)/N]
(9.9b)
n=0
and Eq. (9.7b) is SN
=
N, 0,
m — m = 0, 0 < m — m < N — 1.
(9.9c)
The normalizing constant CN may be determined by substituting Eq. (9.5a) into Eq. (9.5b) with a careful change of dummy indices from m to m and obtaining N-l
N-l
F(m) = AtJ2
A / y ^ F(m) exp ± (ilnnmAf
At)
n=0
x exp == | {ilnnmAf N-l
At) (N-l
= A?A/ ] T F(m) | ] T exp ± (ilnim m=0 N-l
m)nAfAt)
ln=0 N-l
= AtAf J^ F^) m=0
J]{exp
±
(i2n(™ ~ m)AfAt)f.
(9.9d)
n=0
If AtAf = N 1; then the term in curly brackets {•}" = N (5^m by Eqs. (9.9b, c and 9.5b) reduces to F(m) = ( A f A / ) F ( m ) { A ^ m )
(9.9e)
725
Real Ocean Waves
that is possible only if At • Af • N = I, so that Af A / = i
(9.10)
= CN
and the normalizing constant CN is inversely proportional to N. For example, the FFT algorithm in the symbolic software MATHEMATICA™ employs the finite Fourier transform pair Eqs. (9.6), pre-multiplies the complex-valued FFT coefficients F(m) by a normalizing constant CN = s/N and selects the minus (—) sign for the exponential function in Eq. (9.6a) and the positive sign (+) for the exponential function in Eqs. (9.6b). With this convention, Eqs. (9.6) become N
r](n)= ] [ ^ f l ( m ) e x p - ( i 2 j r ( n - l)(m - \)/N);
n=
l,2,...,N, (9.11a)
N
B(m) = */NF(m) = ^
rj(n) exp+(/27r(n - l)(ra -
m = l,2,...,N.
l)/N); (9.11b)
Because Eq. (9.11a) applies the minus (—) sign for the argument in the exponential function exp(«), the complex-valued FFT coefficients B(m) are expressed with a negative (—) sign for the phase by B(m) = \B(m)\ exp - ia(m),
(9.11c)
where a(m) = phase angle at the discrete frequency mAf. Figure 9.3 illustrates the frequency domain representation of the amplitudes (moduli) \B(m)\ in Eq. (9.11c). To illustrate a simple numerical test that may be applied to determine where a particular FFT algorithm places the normalizing constant CN, synthesize the following time sequence that consists of a non-zero mean F(l); first F(2) and fourth F(5) harmonic components with positive and negative phase
Waves and Wave Forces on Coastal and Ocean Structures
726
\B(m)\ ,
I-J I i I U,J_I
I .
t i > i s 1 e • \i|J«J.!!« "•,M,-mHH"'' "
fff
Fig. 9.3. Frequency domain representation of the amplitudes of the FFT coefficients \B{m)\. 11 0.5 , 0
1 i
"-•
"sT -l .
-2 -
1
1
1
1
i
r J! 1 • L I • 1I
i
r
i
i
T
r
J1 1 L I »-H I
•
.
j
r
•
'
J1 • 4_ ! 11 L » L I I -i i I I
i
*
i
i
4
7
10
13
1
n Fig. 9.4. Discrete nondimensional time sequence for N = 16.
angles a(2) and a(5), respectively; and a total number of discrete sequences N = 16 = 2 4 (NOTE: «/w«ys select N = 22M where 2M must always be an even integer so that if the FFT algorithm applies the square root J convention for the normalizing constant CN then the normalizing constant CM will always be a rational number). Consider the discrete time sequence
3
(Inn
TT
+
1
(litAn
C S
n
2 ° U^~4
(9.12)
that is illustrated in Fig. 9.4. A program for synthesizing the normalized discrete time sequence in Eq. (9.12) by the FFT algorithm in MATHEMATICA™ is listed below. The amplitudes \B(m)\ of the complex-valued FFT coefficients are illustrated in Fig. 9.5; and the phases aim) of the complex-valued FFT coefficients are illustrated in Fig. 9.6. Note the change in sign of the phases a(m) in Fig. 9.6 as a result of the sign convention defined for the
727
Real Ocean Waves 5 4 ~^~ 3 i i
-r-i
1 0 7
10
13
16
m Fig. 9.5. Amplitudes of the FFT coefficients for the discrete time sequence r)(ri) in Fig. 9.4.
1
4
Fig. 9.6. FFT phase angles for the discrete time sequence shown in Fig. 9.4.
FFT coefficients in Eq. (9.11c). Table 9.1 lists both the expected amplitudes \F(m) | of the Fourier coefficients without regard for the normalizing constant CN and the amplitudes \B(m)\ from the program MATHEMATICA™. By dividing the values of the amplitudes \B(m)\ from MATHEMATICA™ by the expected amplitudes \F{m)\ of the Fourier coefficients in Table 9.1, it is easy to obtain the normalizing constant CV = V16 = 4. Note that in Table 9.1 that the negative-definite frequencies in the FFT algorithm are stored in the discrete frequencies identified by the indices N/2 + 2 < m < Af. Accordingly, the complex-valued FFT coefficients B(m) for the negative-definite frequencies with indices N/2 + 2 < m < JV are the complex conjugate values B*(m)
728
Waves and Wave Forces on Coastal and Ocean Structures
Table 9.1. Determination of normalizing constant Qy and phase angles a (m) from MATHEMATICA™ FFT algorithm for N = 16 = 4 2 {terms in (•) are the amplitudes and phases of the complex-conjugate FFT values}.
m
\F(m)\
1 2(16) 3(15) 4(14) 5(13) 6(12) 7(11) 8(10) 9
1.0 0.75 (0.75) 0(0) 0(0) 0.25 (0.25) 0(0) 0(0) 0(0) 0
\B(m)\ =
CN\F(m)\
4.0 3.0(3.0) 0(0) 0(0) 1.0(1.0) 0(0) 0(0) 0(0) 0
a(m) (Radians)
CN = JN 4.0 4.0 (4.0)
-(-) -(-) 4.0 (4.0)
-(-) -(-) -(-) -
n -JT/4(+TC/4)
0(0) 0(0) +jr/4(-7r/4) 0(0) 0(0) 0(0) 0(0)
of the positive-definite frequencies with indices 2 < m < N/2; i.e., B(N + 2-m)
= B*(m),
N 2 < m < —.
(9.13a)
The mean of the time sequence is stored in the first real-valued FFT coefficient B(\); and the Nyquist or folding frequency is fy = (N/2)Af and is stored in the real-valued FFT coefficient B (N/2 +1) for 1 < m < N when the positivedefinite index notation of Eqs. (9.6) are applied. Both the mean B(l) and the complex-valued FFT amplitude at the Nyquist or folding frequency B{N/2 + 1) are real-valued because, for the Nyquist frequency at m — N/2 + 1, the complex-valued FFT coefficients B*(N/2 + 1) = B(N/2 + 1) by Eq. (9.13a). For FFT algorithms that employ the positive semi-definite notation for 0 < m < N — 1 given by Eqs. (9.5), the complex-valued FFT coefficients F(m) for the negative frequencies with indices N/2 + 1 < m < N are the complex conjugate values F*(m) of the positive-definite frequencies with indices 1 < m < N/2 — 1 given by F(N -m)
= F*(m),
N 2 <m < —- 1
(9.13b)
Real Ocean Waves
729
and the Nyquist frequency is /N = (N/2)Af. Again, both the mean F(0) and the complex-valued amplitude at the Nyquist frequency F(N/2) are both realvalued, because for the Nyquist frequency m = N/2, F*(N/2) = F(N-N/2) by Eq. (9.13b). The algorithm from the software package MATHEMATICA™ is not optimized and is dated. It is intended only to illustrate one of many possible algorithms that may be applied to determine the normalizing constant CN and the sign of the phase angles a(m) for complex-valued FFT coefficients. Some of the commands illustrated in the program may need to be changed for later versions of MATHEMATICA™. (* PROGRAM DEBUGFFT.MTH FOR DEBUGGING FFT MATHEMATICA *) (* RULES*) TagReal[x_] := (x/: Re[x] = x; x/: Im[x] = 0;) Unprotect[Arg]; Arg[0.0] :=0.; Arg[0] :=0;Protect[Arg]; els :=Run["cls"] (* TIME SEQUENCE w/MEAN AND 2 HARMONICS *) fn :=N[m + al Cos[2 Pi n/npts - Pi/pl] + a2 Cos[8 Pi n/npts - Pi /p2]] f=Table[fn/.{al->1.5,a2->0.5,npts->16,pl->-4.,p2->+4.,m->-l.},{n,0,15}] f=Chop[fJ (* SEQUENCE OF INTEGERS FOR N = 16 *) t=Table[t,{t,0,15}];tf=Table[{t[[i]],f[[i]]},{i,l,16}] tsplot=ListPlot[tf,Frame->True,FrameLabel-> {"n dt","f(n dt)"}, FrameTicks-> {t,Automatic,t,Automatic}, GridLines->Automatic, Prolog->AbsolutePointSize[10]] Display ["ts.plt",tsplot]; (* COMPUTE FFT COEFFICIENTS B(m) *) bnX2=Fourier[fJ;bn=Chop[bnX2];absbn=Abs[bn]; tbn=Table[{t[[i]],Abs[bn[[i]]]},{i,l,16}] bnplot=ListPlot[tbn,Frame->True,FrameLabel-> {"m df","F(m)"}, FrameTicks->{t,Automatic,t,Automatic},GridLines-> Automatic, Prolog- > AbsolutePointSize[ 10]] Display ["bn.plt",bnplot]; phase = N[Table[Arg[bn[[i]]],{i,l,16}]];tph=Table[{t[[i]],Phase[[i]]},{i,l,16}] tphplot=ListPlot[tph,Frame->True,FrameLabel-> {"m df","phase(m)"}, FrameTicks-> {t,Automatic,t,Automatic}, GridLines->Automatic, PlotRange->All,Prolog->AbsolutePointSize[10]] Display ["tph.plt",tphplot]; finv=InverseFourier[bn];finv=Chop[finv]
730
Waves and Wave Forces on Coastal and Ocean Structures
SameQ[finv==f] (* OUTPUT FILE TO BE IMPORTED TO SPREADSHEET "QPRO" *) SetDirectory["C:\lfn\"]; z=Table[ {t[[i]],f[[i]],absbn[[i]],phase[[i]]}, {i, 1,16}] ColumnForm[z];»debugfft.prn
9.3. Ocean Wave Spectra Spectral representations of stationary ergodic random seas are required for engineering analyses of random ocean waves. Several two-parameter theoretical spectral models have been applied to engineering applications along with five- and six-parameter spectral models. The parameters applied in the original derivations of these spectral models are varied. Among the most commonly applied parameters are wind speed Uw, fetch length F, significant wave height Hs and period Ts. For those two-parameter spectral densities, it is convenient for engineering design to replace the original parameters with the zeroeth spectral moment mo (= to the variance a2 of the time series or the area under the spectrum) and the frequency of the spectral peak /o or COQ. The mean (= p,\), the mean square value (= \xi), the variance (= a% = mo) and the standard deviation (= <JX) of a continuous time series may be computed from a stationary ergodic random process by either an ensemble average of an infinite number of finite length records (vide., Fig. 9.1) or from a temporal average of a single record of infinite length TR (vide., Fig. 9.2) or from probability moments /x„(«) by the following (Bendat and Piersol, 1986): +oo
/
=
xnp{x)dx
-00
lim — [ *xn(t)dt, TR-^oo TR JQ
n = l,2,3,...,
(9.14a)
+oo
/ 1
=
f
TR
(X -
pLX)2p(x)dx
-00
->
9
lim — / (x(t) - n\Ydt TR^oo TR JO
= 112- p\,
(9.14b)
731
Real Ocean Waves
where £[•] = an ensemble averaging operator defined in Eq. (9.1); p(») = the probability density function (pdf) for the random variable (•); and 7> = a temporal record that is infinitely long in Eq. (9.1). Data are frequently normalized by subtracting the mean \x\ and then dividing by the standard deviation ax of x(t) in order to obtain the following zero-mean, unit variance data record: =
*(O-MIW
(914c)
Ox
The Wiener-Khinchine Fourier transform pair is similar to the Fourier series transform pair from Eqs. (9.2) but relates the covariance function Cxy(x) to the two-sided spectral density function Gxy(co) (Bendat and Piersol, 1986) according to 1 r+°° Cxy(r) = —— / Gxy(oo)exp ±(ioor)dco, V2TT J—oo
i Gxy(co) = —=
(9.15a)
r+°° I
Cxy(r)exTp^(ia)T)dr,
(9.15b)
V2TT J-OO
where Gxy (co) = a complex-valued, two-sided cross-spectral density function that may be expressed as Gxy(co) = Cxy(co) ± iQxy(a>),
(9.15c)
where Cxy (oo) = coincident spectral density function and Qxy (co) = quadrature spectral density function. Alternatively, the complex-valued, two-sided cross-spectral density function in Eq. (9.15c) may be expressed as an amplitude and a phase by Gxy(co) - \Gxy(oo)\exp ±iaxy(co),
(9.15d)
where the amplitude \Gxy(oo)\ and phase axy(co) are computed from Gxy(oo)\ = JC^Jco) + QlJco),
axy(co) = arctan
Qxy (CO) Cxy(00) _
(9.15e,f)
732
Waves and Wave Forces on Coastal and Ocean Structures
A real-valued coherence function may be computed by 2
YIy(w) =
\Gxy{co)\2 \
< 1,
(9.15g)
where G f f (&>) = areal-valued, two-sided spectral density function for the time series £(/)• A real-valued one-sided spectral density function Sm{co) for the time series r) (t) may be computed from the two-sided spectral density function Gxy{(o) by Sm{a>)=2Gm(a>)U{a))
(9.15h)
where U(a>) = the Heaviside step function in Eq. (2.1) in Sec. 2.2.2. Onesided spectral density values Snri (a>m) may be computed for the discrete radian frequency com from two-sided, complex-valued discrete FFT amplitudes | Bm \ in Eq. (9.11c) by Sm(a)m) =
2\Bm\2 ' "' IjcdfC^i
\Bm\2 = —^Ndt, TTCN
(9.15i)
where Eq. (9.10) has been substituted for ^f and where CN =the FFT normalizing constant defined in Eq. (9.10) for the FFT coefficients computed by MATHEMATICA™. Similarly, values may be computed from Eq. (9.15i)for the random wave simulations in Sec. 9.6 by 2\Bm\2 S^icom) = ^ - ^
(9.15J)
where CN = the FFT normalizing constant defined in Eq. (9.10) for the FFT coefficients computed by MATHEMATICA™. One-sided spectral density functions Sm (•) may be expressed as functions of the independent variables (•) of radian frequencies a>(=27tf), or of cyclesper-second (cps) frequencies / , or of wave periods T, or of scalar wave numbers k{=2iz/X) or of vector wave numbers k. In order to determine the relationship between spectral densities expressed with different independent variables, equate the differential area under each spectral density in a small incremental interval of the independent variable according to -S^mdT
= Snr}{(o)da> = Sr,„(f)df = Snn{k)dk = Snri{k)dk.
(9.16a)
733
Real Ocean Waves
For co = 2nf, the transformation Jacobians required for Eq. (9.16a) may be determined from dco = litdf
= -^rdT
= -27tf2dT;
= -—dT
z
T
(9.16b)
2n
so that, accordingly, for the independent variables T, f and co 7
Sm(T) =
T f2Sm(f)
CO
= — Sm(a>), lie
(9.16c)
S^f) = 2nSr,n(eo). (9.16d) Transformations between frequency / and wave number k spectra require the linear dispersion equations (4.15) in Chapter 4.3 given by co2 = (2TT/) 2 = gktanhkh;
(9.16e)
and the corresponding Jacobian transformation from Eq. (9.16e) is given by Eq. (4.60c) in Chapter 4.5
where CG =the wave group velocity. The Jacobians from Eq. (9.16f) for the deep- and shallow-water approximations, respectively, for Eq. (9.16e) are dco
co
TT
=^r>
UK deep-water
ZICQ
dco
^
co
/—-
,„ , ^ , s
=T = Vgk,
(9.16g,h)
UK shallow-water
K
where the deep-water wave number ko = co2/g. Correlation-covariance definitions There do not appear to be consistent definitions with respect to the time series Cxy{x) (Kinsman, 1965). The function Cxy(r) is defined as a crosscorrelation function if the times series x(t) and y{t) are scaled by their means [i\ and standard deviations at in accordance with Eq. (9.14c) such that each time series has zero-mean /xi = 0 and unit variance of = 1. If the time series x{t) and y(t) of record length TR are not scaled by Eq. (9.14c), then the function Cxy (r) is defined as a cross-covariance function.
734
Waves and Wave Forces on Coastal and Ocean Structures
The two-sided spectral density function Gxy(a>) is always defined as a crossspectral density function regardless of whether or not the time series x(t) and y(t) are scaled in accordance with Eq. (9.14c). If the two time series x(t) and y(t) are identical, then Eq. (9.15a) is defined as an auto-covariance (or -correlation) and Eq. (9.15b) the auto-spectral density function. Historically, the covariance (or correlation) function was computed in order to efface the randomness from a time series of a random process in order to expose the invariant statistical anatomy of the process (Wiener 1964, p. 6). The cross-covariance and cross-correlation functions are computed from time series x(t) and y{t) by 1
fTR/2
Cxy(x) = lim — / TR^OO
1R
x(t)y(t + r)dt,
|r| < oo,
(9.17a)
J-TR/2
TR-+oo 1R J-TR/2
|r|
(9.17b)
where Eqs. (9.17a, b) are symmetric according to
Cxy(-r) J = j Cxy{x) 1 <W-T)J iQ^OOj
(9i7c) (9.17d)
and where Eqs. (9.17) are always real-valued functions and symmetric about r = 0. The auto-covariance Cxx{x) and auto-correlation C^x^y{x) functions are computed from 1
Cxx(r)=
f>TR/2
lim — / r f i ^ o o 1R
c
?x&(T) =
Tlmi
—
x(t)x(t + r)dt,
(9.18a)
J-TR/2 /
—2
dt,
(9.18b)
735
Real Ocean Waves
where Eqs. (9.18a, b) are always real-valued functions, symmetric about x = 0 and equal to the variance 0% of x(t) for r = 0; i.e., Cxx(0) 1 = }a x 2 ] lCfc&(0)J 11J
(9!8c) (9.18d)
Unless x(t) is a strictly periodic time series, Eq. (9.18b) is also proportional to the mean ii\ (x) for r —> ±00; i.e., j y/Cxx{±oo) I _ j ^ j ^ J
(9.18e)
IV^(±°°)J~1 ° J
(9.18f)
The first analyses of random data computed the cross- (or auto-) covariance (or correlation) function from the time series by Eqs. (9.17a, b or 9.18a, b) and then applied the Wiener-Khinchine Fourier transform Eq. (9.15b) to obtain the two-sided spectral density function. Modern analyses of digitized discrete time sequences (tn = ndt, Eq. (9.4a)) employ the discrete finite Fourier transform (FFT) (fm — mdf, Eq. (9.4b)) to compute the complex-valued discrete FFT coefficients Fx(m) and Fy{m) of the discrete time sequences x(n) and y(n); and then apply these discrete coefficients to compute either the discrete cross-covariance Cxy{n) (or cross-correlation Cxy{n)) function or the two-sided Gxy(m) (or one-sided Sxy(m)) cross-spectral density functions. A comparison of these two methods for obtaining discrete spectral estimates from an FFT algorithm is illustrated in Fig. 9.7 where the notation FFT implies the forward transform from Eq. (9.11b); and the notation F F T - 1 implies the inverse transform from Eq. (9.11a). To illustrate how either a two-sided Gxy(m) or a one-sided Sxy(m) crossspectrum may be computed by either of the two paths in Fig. 9.7 by a discrete FFT form of the Wiener-Khinchine Fourier transform pair in Eqs. (9.15a, b), a random time series rj{t) of six cosine waves is synthesized from the amplitudes and phase angles listed in Table 9.2 and is illustrated in Fig. 9.8a. The amplitudes Am, phase angles am and discrete frequencies com = lizmdf for each of the six cosine wave are summarized in Table 9.2. Continuous time and frequency are discretized by Eqs. (9.4a, b) for application of an FFT algorithm where dt = 0.2 sec, AT = 64 and df = 1/Ndt = 0.078125 Hz. The auto-correlation function C^(x) for the normalized random time series f(/) computed from x (t) = rj (t) by Eq. (9.14c) may be computed from the discrete
736
Waves and Wave Forces on Coastal and Ocean Structures
{x(n),y(n)}
C„ {n) "
Fx (m),Fy (m)
Gxy(m),Sxl,(m) Fig. 9.7. Comparison of methods for computing spectra by FFT and by covariance functions. Table 9.2. Parameters for the random time sequence dx = 0.2 sec, N = 0.078125 Hz, MI (?) = 0 ft frequency m 5 1 9 11 13 15
the six cosine waves for in Figs. 9.8a and b (dt = 64, df = 1/Ndt = and tr 2 = 64ft2).
a>m (rads/ sec)
A m (ft)
am (rads)
2.4544 3.4361 4.4179 5.3996 6.3814 7.3631
2.0 4.0 8.0 6.0 2.0 2.0
1.0192 2.0579 5.2495 5.3168 2.6336 0.6556
Fourier coefficients F f (m) by the FFT coefficients Bm for f (n) following the horizontal path in the middle of Fig. 9.7 according to C ff (T„) = |F f (m)| 2 = - ^ - ,
(9.19a)
where Civ=the FFT normalizing constant defined in Eq. (9.10) for the complex-valued FFT coefficients Bm computed by MATHEMATICA™. The two-sided discrete amplitude spectrum \G^(m)\ in Eq. (9.15d) may be computed from discrete FFT coefficients Bm by \Gu(m)\ = \F((m)\2 = ¥^-,
(9.19b)
737
Real Ocean Waves
0
2
4
6
8
10 12 14
t (sec)
JJJ.._LLL J.1L JJJ.._LLL J-LL JJJ...LLL J-LL
_LLL -LLL
Fig. 9.8a. Random time sequence r){tn) of six cosine waves in Table 9.2 where tn = n(0.2) sec. .LLL AAL JJJ. .LLL .LLL AIL -UJ. .LLL .LLL AIL .UJ. L L L
JJJ..-LLL AIL _UJ. .LLL AAL JJJ. L L L JJJ. .LLL J-LL JJJ. .LLL J-LL -UJ. L L L JJJ..JJJ. JJ.L JJJ. .LLL J 1 L -UJ, L L L JJJ.. .LLL J 1 L JJJ. .LLL J-LL -UJ- L L L JJJ, .LLL J 1 L JJJ. .LLL i l l -UJ_ L L L JJJ.. .LLL J L L J J J . .LLL J L L -UJ- L L L JJJ,. .LLL AIL JJJ. .LLL J 1 L _UJ_ L L L JJJ.. .LLL ALL JJJ. .LLL J 1 L -UJ. L L L JJJ. ALL JJJ .LLL J 1 L
—J, -J—
0
-UJ*
8 16 24 32 40 48 56 64 m
Fig. 9.8b. Two-sided discrete amplitude spectrum computed from FFT coefficients for six cosine waves in Table 9.2 for fm = m/Ndt Hz.
where CM =the FFT normalizing constant defined in Eq. (9.10) for the FFT coefficients computed by MATHEMATICA™. The two-sided discrete amplitude spectrum |G w (m)| for the time series rj(t) computed from FFT coefficients Bm by Eq. (9.19b) is illustrated in Fig. 9.8b. The symmetry about the Nyquist or folding frequency m — N/2 + 1 = 33 in Fig. 9.8b of the two-sided discrete amplitude spectrum \Gm(m)\ is a consequence of the negative-definite frequencies being represented by the discrete frequency interval N/2 + 2 < m < N — 1. The mean and standard deviation of the random time series r](t) synthesized from the parameters in Table 9.2 are ii\{r)) = Oft and or] = 8.0ft, respectively. The random time series r](t) is normalized by this mean ii\(,rf) and standard deviation <7n in accordance with Eq. (9.14c) and the resulting normalized random time series C, (t) is illustrated in Fig. 9.9a. The two-sided discrete amplitude spectrum \G^(m)\ may be computed from discrete FFT coefficients
Waves and Wave Forces on Coastal and Ocean Structures
738
2
4
6
8
t (sec)
10 12 14
Fig. 9.9a. Normalized random time sequence f (tn) of the six cosine waves in Table 9.2 where
0.25 _
H-4 +
0.2 ' 4 4 +
10.15 to. 0.1 0.05 0
43± -1-4 + "lOT
i-l-ti-i—n-tTO TTT44-HI4T Ed: 4-14- -t-H- + 4-4-4+4- 44 + L I J - l L U - xrt xnr + 4-h -4 + 31
1-4 +
uremic 431 B J ' itti±tJit
trnnTrriTT
331 -U + 111
0 8 16 24 32 40 48 56 64 m
Fig. 9.9b. Two-sided discrete amplitude spectrum computed from FFT coefficients for the normalized time sequence r (t) in Table 9.2 for fm = m/Ndt Hz.
by Eq. (9.19b) and is illustrated in Fig. 9.9b. This illustrates the computational procedure for computing the two-sided discrete amplitude spectrum \Grr(m)\ following the FFT path on the right side of Fig. 9.7. The two paths illustrated in Fig. 9.7 for computing the two-sided discrete amplitude spectrum \Grr(m)\ for the normalized random time sequence f (tn) in Fig. 9.9a will both be followed in order to demonstrate their differences. First, the auto-correlation function Crr(t) is computed by Eq. (9.19a) from the FFT coefficients Bm for the normalized random time sequence £(?„) and is illustrated in Fig. 9.10a. Algorithms for constructing covariance or correlation functions from FFT coefficients are given by Brigham (1974, p.206, Fig. 13-6a) or by Bendat and Piersol (1986, Chapter 11.6.2, pp. 406-407). Note in Fig. 9.10a the symmetry about T = 0 of the auto-correlation function Crr (T) in accordance with Eq. (9.17d); and the limiting values of the autocorrelation function Crr (T) ~ 0 = the mean /AI(£) as r ->• ±7>2dt = 6.4 sec in accordance with Eq. (9.18f). Next, the two-sided discrete amplitude
739
Real Ocean Waves
- 6 - 4 - 2 0 2 4 6 T (sec) Fig. 9.10a. Auto-correlation function for normalized time sequence f (?„) in Table 9.2.
0.25
-H4-
0.2 FmF!-" 444;0.15:m
fc
-t-H--n-i--n-t--i-n111:311:311 -t-i-t--t-t-t--i-t-t-
tctltcaitzlit
H 4 -H4- 4-H- 4J-H4 ^4-
•l-H- 4 H - 4 +
0.1 --H-I- : P 3 : : c q : 0.05
: B I bzfci l t d : rp: ttt 444*^4-
- 1 ^ * 1 . -|—i*r« -jpr-i
4 : S nrx
t.+. t. . . . . J - H -
mcdumtar '
-T-T—I"
-144 -144-
-+4*444
innttnnir :•! 41-1-4
Q ^ . . i | r 5 y h » | T r T iIBIWi 0 8 16 24 32 40 48 56 64
m
Fig. 9.10b. Two-sided discrete amplitude spectra for the auto-correlation function Cjf (r) computed from Table 9.2 by the FFT coefficients from f (r){»»} and by the Wiener-Khinchine Fourier transform pair {xx}.
function \G^(m)\ is then computed from the FFT coefficients of the autocorrelation function C^(t) by Eq. (9.19b) following the path on the lower left side in Fig. 9.7. Second, the two-sided discrete amplitude spectrum | G^ {m) \ is computed from the FFT coefficients of the normalized time sequence £(/) by Eq. (9.19b) following the path on the lower right in Fig. 9.7. Both of these two-sided discrete amplitude spectra \G^(m)\ are compared in Fig. 9.10b. The non-deterministic definition for Fourier coefficients in Sec. 9.2 is illustrated in Fig. 9.10b. Both of the two-sided discrete amplitude spectra in Fig. 9.10b have unit variance; but contributions to the variance by the amplitudes from each frequency are very different. The discrete spectrum computed from the FFT coefficients for f (f) by the path on the right side in Fig. 9.7 are limited to only the six discrete frequencies in Table 9.2. In contrast, the discrete
740
Waves and Wave Forces on Coastal and Ocean Structures
spectrum computed from the auto-correlation function for £ (?) distributes contributions to the total variance at more than the six discrete frequencies listed in Table 9.2. However, the total variance computed from the discrete amplitudes of both spectra are identically equal to unity but the contributions from the discrete frequencies are different. For this reason, two-sided discrete amplitude spectra are rarely computed from covariance or correlation functions and are computed directly from the FFT coefficients by the path on the right in Fig. 9.7 for stochastic processes. 9.3.1. Generic Four-Parameter
Wave Density Spectrum
A generic four-parameter one-sided wave density spectrum that may be related to many theoretical wave spectra may be expressed as a product of two frequency functions that are dimensional (denoted by tildes ~) and that are given by F\(5>) =
mW
F2(a>) = exp
p
m'
p,q>0,
(9.20a,b)
CD
where the dimensional parameter A is a constant; m is a characteristic radian wave frequency that may also include other multiplicative constants; and p and q are positive-definite integer constants. A generic dimensional fourparameter, one-sided spectrum derived from the product of Eqs. (9.20) is 5(o>) =
m\qm {5>)P
exp —
0 < a> < oo.
(9.21)
CO
If Eq. (9.21) represents a one-sided wave density spectrum, the dimensions of Sw(
[Time] =
lit
m
(9.22a,b)
741
Real Ocean Waves
then a dimensionless generic four-parameter one-sided wave density spectrum S(Q) may be expressed by the product of the two dimensionless functions F\ (Q) and F2(fi) that are defined by (9.23a,b)
mourp '
m Fi(fl) = — ,
F 2 (Q) = exp - «"«,
(9.23c,d)
and that may be combined according to the following product
m0 = —exp-ft_,
0 < f t = — < oo. m
. (9.23e)
Dimensionless values of F\ (Si)/A, F2 (fi) and S(fi) are illustrated in Fig. 9.11 and demonstrate that F\{Q)/A controls the spectral behavior at frequencies higher than the dimensionless spectral peak frequency £2 > £2o = a)/mo = 1 and that F2(Q) controls the spectral behavior at frequencies lower than the dimensionless spectral peak frequency Q < £2o = «/<^o = 1- Note that for dimensionless frequencies less than unity that F2(£2) approaches zero faster than F{(Q)/A approaches infinity. The dimensionless parameter A may be related to dimensionless spectral moments that correspond to characteristic radian frequencies (Vanmarcke, 1983, Chapter 4.1). The n th dimensionless spectral moment m„ is
0
1
2
__
3
4
Q =co/i3 Fig. 9.11. Dimensionless generic spectral functions.
742
Waves and Wave Forces on Coastal and Ocean Structures
defined by
mourn
Jo
\m J mo/m
'
/<00
nn-pexp-Q-idQ,
= A
n = 0,1,2,...,
(9.24)
Jo that may be integrated dimensionlessly to obtain A mn = -T[q \
(p-n-\\ q
,
p-n-l>0,
(9.25a)
J
where T(«) = Gamma function defined by Eq. (2.6a) in Chapter 2.2.5 and the dimensionless zeroth moment mo = 1. For each dimensionless spectral moment n in Eq. (9.25a), there corresponds a dimensionless characteristic radian wave frequency defined by (Vanmarcke, 1983, Chapter 4.1, Eq. (4.1.4))
On = (^)l/" m I
= (-J^-Y" = (m,)1'-, »>0. \mozrr"
(9.25b)
Spectral peak frequency a>o The dimensional spectral peak frequency u>o may be computed from Eq. (9.23e) by — ^ ail
= 0,
n = «0 = - = 1 coo
(9.26)
so that the dimensional characteristic radian wave frequency parameter m may be defined by COQ as m = coo (-)
(9.27)
where (p/q)l^q is the multiplicative constant noted following Eqs. (9.20). The dimensionless generic parameter A in Eq. (9.23e) may be replaced by
143
Real Ocean Waves
Eq. (9.25a) with n = 0 and with the dimensionless variance of the time series mo = 1.0 (or, equivalently, the area under the dimensionless spectrum 5(fi)); i.e., A=
r « / > - i)A?)
(9.28)
and the dimensional characteristic radian wave frequency m from Eq. (9.27) so that Eq. (9.23e) becomes jP/o
S(Q0)
=
q(.(j>/9)-Dr ((p 0 <
SIQ
\)/q)
Q0
p
exp — | — Q0
q
u> = — < oo. coo
(9.29a)
A dimensional form of Eq. (9.29a) for arbitrary values of the exponents p and q may be obtained by multiplying Eq. (9.29a) by mo/m with m defined by Eq. (9.27) to obtain S(co,mo,coo,p,q)
= 1
(iq + \-p)lq)
m0 p(d-p)/«) co0r((p - \)/q)
x exp
P_ im q\co
\!o (9.29b)
A number of theoretical wave spectral density functions have exponent values of p — 5 and q = 4. The parameters of several of these theoretical wave spectral density functions may be converted to the parameters of mo and coo to obtain a generic dimensional two-parameter spectral density function given by mo (coo S(co, mo, coo, P = 5, q = 4) = 5 COQ \ co
exp
5 /coo ~4\~co~
(9.29c)
and that are tabulated in Table 9.5. A dimensionless plot of the generic spectral density function Eq. (9.29c) is shown in Fig. 9.12. It is not an easy task to determine the spectral peak frequency coo from measured wave data because of the variability in real spectral estimates obtained from dimensional FFT algorithms. An estimate of the spectral peak frequency may be computed from FFT coefficients in a best least-squares sense by
Waves and Wave Forces on Coastal and Ocean Structures
744
1.9
i
1
[.../.
mju>a n K-
i -4
n 0
0.5
i 1
1
\ 1.5
2
2.5
3
G5/W„
Fig. 9.12. Dimensionless two-parameter generic spectral density function for p = 5 and q=4.
applying the linear Taylor differential correction method (Marquardt, 1963). Because the dimensional FFT coefficients are indexed to discrete frequencies m, the method may be defined with discrete dimensional radian wave frequencies a)m = mAco = m2nAf for the dimensional FFT algorithm from Eq. (9.5b). A dimensional mean-square error e~ between a dimensional measured spectral density estimate SM(W) computed from the dimensional two-sided discrete FFT coefficients at discrete frequencies m by Eq. (9.15i) and the dimensional theoretical generic spectral density S(m) computed by Eq. (9.29c) may be defined by
«1 =
MC
1 Mc
2
J2
-Ms
[SM("0
- 5(m)] ,
(9.30)
m—Ms
where Ms = the starting index for the first significant dimensional FFT coefficient and Mc = the index of the cut-off frequency above which the dimensional FFT coefficients are negligible in the measured spectral density SM(»»)- By restricting the linear Taylor differential correction method to only those few frequencies in the vicinity of the estimated value of COQ, the algorithm is very efficient. The generic dimensional theoretical spectral density S(m) from Eq. (9.29c) may be expanded in a Taylor series about the dimensional spectral peak frequency u>o by S(m) = S(m) +
dS(m) dcoo
-<5
(9.31)
Substituting Eq. (9.31) into Eq. (9.30) and minimizing with respect to Sa)0 according to dScbo
= 0
745
Real Ocean Waves
gives j:
(SM(m) - S(m))
m=Ms SCOQ
9<wo
(9.32)
=
An initial estimate for <SQ may made from the dimensional spectral estimates and the (j + 1) estimate for a)J0+ may be computed from the jth correction computed from Eq. (9.32) by
&l+l = a>J0 + 8&1.
(9.33)
The iterations are terminated when the corrections 8a)JQ are stable and acceptably small (10~ 6 ,say). Note that the theoretical spectrum S(m) from Eq. (9.29c) must be recomputed after each iteration because of the newly computed value of <5Q . Table 9.3 summarizes an application of this algorithm to Hurricane Carla spectra (Hudspeth, 1975). There are commercially available software packages that will compute spectral peak frequencies from spectral estimates.
Average or mean radian wave frequency 5) Because it is easier to compute the spectral moments mn than it is to compute a>o from data, the dimensional characteristic radian wave frequency m in the dimensionless generic four-parameter spectrum in Eq. (9.23e) may be replaced Table 9.3. Summary of least-squares fit to Hurricane Carla data for MQ — Ms = 305 and N = 4096 (Hudspeth, 1975).
Record No. 06885/1 06886/1 06886/2 06887/1
Initial col. [rad/sec]
Final <J>Q
Final <S<WQ
[rad/sec]
[rad/sec]
Final j
0.52 0.50 0.50 0.50
0.4990 0.5187 0.4847 0.5199
-6xl0"6 - 1 xlO-6 -97xl0~6
5 4
4xl0"6
9 5
el s [ft 2 /(rad/sec]
m0 [ft2]
5.29 5.83 12.23 6.52
22.31 28.76 24.96 28.63
746
Waves and Wave Forces on Coastal and Ocean Structures
by the dimensional average or mean radian wave frequency cb or period T that may be computed easily from the spectral moments in Eqs. (9.25) for n = 1 by (Vanmarcke, 1983) m\ {mQTu)m\ „ \ i J co = — = —— — = mmi = w—)—-Y = mA2\(p,q) m0 (m0)(m0 = 1) H ^ )
(9.34a)
from Eq. (9.25a) and where Aijip, q) =
)
q
.{.
(9.34b)
A dimensional characteristic radian wave frequency m and dimensionless frequency ratio Q may be replaced by OJ
m = coAn(p,q),
fit
= —,
(9.35a,b)
OJ
where A12 is the multiplicative constant noted following Eq. (9.20) and Eq. (9.23e) becomes S(Q)=
qA\2{p,q) ' " " ' "
exp-
(An(p,q)\q "-?'1' ,
CO
0
(9.36)
The dimensional average or mean radian frequency may be interpreted as the distance of the centroid of a one-sided spectral density function with unit variance from the frequency origin in analogy to the average of a non-negative random variable computed from a probability density function (Vanmarcke, 1983). Zero-crossing coz or root-mean-square a>s frequency A dimensional zero-crossing eoz or root-mean-square u>s radian wave frequency may be defined as mi / (mom2)m2 _ .— a>z=Q)s = J— = J - T — — = mjmi V mo V (mo)(mo = 1)
.- __ . (9.37a)
1A1
Real Ocean Waves
and the dimensionless second-moment m,2 from Eq. (9.25a) with the constant A defined by Eq. (9.28) for mo = 1 as m2 =
) \[
= A31 (p,q),
Vz = ^-
(9.37b,c)
and Eq. (9.23e) becomes •S("z) = —7
rv
TZv
'(V)
«r
exp-'
v «?
0 < ^ z = - ^ < oo.
(9.38)
Vanmarcke (1983) defines a generic characteristic wave frequency Qk as «jfc =
mA
—
i/k •
(9-39)
" ( m0/
9.3.2. Wave and Spectral Moments mn
Parameters
Computed from
Spectral
All of the variables in this section Sec. 9.3.2 are dimensional variables; and, consequently, the tilde (•) notation applied in Sec. 9.3.1 is not applied here to denote dimensional variables. Dimensional spectral density moments are computed from a dimensional one-sided spectrum Sm{a)) by /•OO
I con Snr)(oo)da). (9.40) Jo Many of the wave and spectral parameters that are computed below are evaluated for the dimensional generic two-parameter spectrum Sm{ma,a>o,co, p = 5,q = 4) in Eq. (9.29c); consequently, the first four dimensionless spectral moments computed by a dimensionless Eq. (9.40) are summarized in Table 9.4 where the Gamma function F(») is defined by Eq. (2.6a) in Chapter 2.2.5. mn=
Spectral shapes Dimensional spectral moments mn computed from Eq. (9.40) are functions of the shape of the dimensional spectral density Snn{co). These dimensional
748
Waves and Wave Forces on Coastal and Ocean Structures
Table 9.4. Summary of dimensionless spectral moments mn for a dimensionless generic two-parameter spectrum S^imo,ci)o,a>,p = 5,q = 4 ) where mn = mn /<3)g.
n
0
1
2
mn
1
ar-®
4
f1 r~
[5 I
! ) -
r(0) = oo
spectral density moments have been applied to compute a number of important and useful wave and spectral characteristics for design; e.g., the probability distributions of extreme value statistics; the maximum wave height; characteristic spectral wave frequencies; the maximum structural displacement due to wave loading, inter alia. A few of these wave and spectral parameters are derived below and are also summarized along with their formulas for a dimensionless generic two-parameter dimensional spectrum Sm{mQ,a>o,5),p = 5, q = 4) in Eq. (9.29c) for the first five dimensionless spectral moments n = 0 - 4 in Table 9.4. Spectral bandwidth parameters e,q and v A dimensionless spectral bandwidth parameter e that may be applied to compute some extreme value statistics for a Gaussian process (vide., Sec. 9.4) is the following (Cartwright and Longuet-Higgins, 1956): e2 = 1
2
-,
(9.41)
where the dimensional spectral density moments mn are computed from Eq. (9.40). For a narrow-banded spectrum, e -> 0 and the maximum values for a stochastic process that is represented by a narrow-banded spectrum are Rayleigh distributed (vide., Sec. 9.4.2). For a broad-banded spectrum, e -» 1 and the maxima values for a stochastic process that is represented by a broadbanded spectrum are Gaussian distributed (vide., Sec. 9.4.1). For the generic two-parameter dimensional spectral density 5^^(mo,a>0!<w> P = 5,g = 4) in Eq. (9.29c), Eq. (9.41) is
rHm_ = r(i)r(0)
749
Real Ocean Waves
where r(0) = oo in Table 9.4. Consequently, all of the generic two-parameter dimensional spectra Svr){m,Q,coQ,CL>,p — 5, q = 4) that are tabulated in Table 9.5 may be interpreted as being dimensional broad-banded spectra; i.e., e —> 1. However, even though the dimensionless spectral bandwidth parameters e, q and v discussed here are useful for estimating certain statistical quantities, the effects of the variability in the spectra computed from measured realizations limit the interpretation that these measured spectra are broad-banded for engineering applications. The effects of this variability in the realizations from these spectra are evaluated by the Hilbert transform and the envelope function for engineering applications to damage estimates for rubble mound breakwaters in Sec. 9.5. The fourth spectral moment m.4 is undefined by Eq. (9.40) and in Table 9.4 for the generic two-parameter dimensional spectra S(mo,coo,co,p = 5,q = 4) in Eq. (9.29c). In addition, m.4 is also non-converging when computed numerically from data. For this reason, the spectral bandwidth parameter e may not be computed reliably in engineering applications from data; but it does continue to serve a useful purpose for evaluating the theoretical distributions of the maximum values in time series analyses (vide., Cartwright and Longuet-Higgins, 1956, Sec. 4, and Rice, 1954). Extremal statistics may also be computed from the dispersion of the wave frequency spectrum about the central spectral wave frequency from the following alternative dimensionless spectral bandwidth parameter q (Vanmarcke, 1972): mi2 1 q = 1 —, (9.43a) mom2 where the bold q in Eq. (9.43a) should not be confused with the italic q in the exponent in Eq. (9.21). Because the spectral bandwidth parameter q in Eq. (9.43a) does not require the fourth spectral moment ra4,q2 may be computed reliably from data. Vanmarcke (1983, Chapter 4) determines that extreme values are Poisson distributed. For the generic two-parameter spectrum S(mo,(Oo,(o,p = 5,q = 4),q 2 from Eq. (9.43a) with the dimensional spectral moments ra„ from Table 9.4 is given by ,
2
27
q = 1
r 2 (3/4) - ^ = 0.152787. Jn
(9.43b)
Table 9.5. Conversion of two-parameter theoretical spectra to generic spectral parameters mQ, a>o a n ( l £/(•) = Heaviside step function. SPECTRUM S(u>)
m0 0.3932
S-M-B: {Hs,cos) (Bretschneider 1958, 1966) 1.618f(^)5exp [-1.03(f)4]
j-)a> (¥)
"exp
e x
5^(!f)5exp[-!e)4]
0.9528&).
24/|^(t)6exP[-3(^)2]
64
'•\uww) ]
P-N-J: Pierson, etal. {a,Uw} (1955) (vide., Pierson and Moskowitz, 1964 ) ^
tf02
3CJT
Neumann {C, Uw] (1943) (Kinsman, 1965)
S(mo, COQ, a>)
a>0
0.877 ( ^ )
5^(^)5exp[-I(f)4]
0.0625 if?
Q.11U5)
5^(!f)5exP[-I(^)4]
s
0.0625/f?
0.71ft>7
5^e)5exp[-fe)4]
8
f
ft)o
.1227(^)1
<*£
-0.74 \UuCo) ]
P
ISSC{Hs,
4
0-HOvf (f) exp [-0.4427(f) ] ITTC(//s,(uo}(1966) 0.0795f
5
4
(f) exp[-0.318(f) ]
Scott {Hs,w0} (1965) (Darbyshire, 1959) 0.214#/exp
(W-OJQ)2
r]
[ 0.065(o)-a) 0 +0.26) •
AU = {[/(ft) - &>0 + 0.26) -U(co-
AU
co0 - 1.65)}
WALLOPS IP,p,coo) (Huang, etal., 1981)
4
(^r-p[-f(^) ]
2
Mitsuyasu {A, F,g, and w} (1971)
*&)*""--[*&)-"'((,)'
,?<"-'> ft)o
p>\ mo 5A
(?)
)f
(a)-a)p) [ 0.065(a)-wo-
} , / 2 l AU
£
+ 0.26) J J A<7 = {[/(&) - ft)0 + 0.26) - U(co - co0 - 1.65)}
^ fe2r[V]
SPECTRUM S(f)
3.424mo exp
J ( ^ 5 ) r[l( P -i)]
(^"exp^fe)4] p>l
/o
S(m<), /o. / )
7?0>)«P-
as a. 55
(ff
751
Real Ocean Waves
Longuet-Higgins (1975) defines a spectral narrowness parameter v that may be related to the Vanmarcke parameter q according to v2 =
mom _ m\
1 =
/momA q 2 \ m\ ]
{gM)
Goda dimensionless spectral peakedness parameter Qp Goda (1970) identifies a parameter Qp that correlates with wave groups in Sec. 9.5 and is defined for dimensional one-sided spectral densities Sm(f) by 2 r00 QP = —,\ fSl(f)df. vv
(9.45)
The Goda spectral peakedness parameter Qp appears to be less sensitive to the cutoff frequency that is required in order to compute spectral moments mn from data and also less sensitive to wave nonlinearities than the spectral narrowness parameter v in Eq. (9.44). The generic two-parameter spectrum Sm (mo, coo, co,p = 5,q = 4) in Eq. (9.29c) may be converted from radial frequencies co = 2n f to Hertzian frequencies / by equating differential spectral areas from Eq. (9.16d) according to S(f)df S(f)
= S(co)dco = S((o)2 7tdf, = 2TZS(CO).
For the generic four-parameter spectrum S^ (mo, coo, <*), p, q) in Eq. (9.29b), Qp may be computed by substituting Eqs. (9.29b and 9.16d) into Eq. (9.45) and obtaining r[2(p-l)
QP = 2 + 2 « q-
L
*
2p / / 0 \ T V
r*m
/J
-"0
>(2(p-l)\
r2
(¥)
where r[; •] = incomplete Gamma function (vide., Eqs. (2.8), Chapter 2.2.5).
752
Waves and Wave Forces on Coastal and Ocean Structures
For the dimensional generic two-parameter spectrum S^ (mo, U>Q, CO, 5,4) in Eq. (9.29c), Qp is QP = h
(9.47)
that is approximately equivalent to Qp for a broad-banded Gaussian white noise spectrum (Goda, 1985). 9.3.3. Multi-Parameter
Theoretical Spectra
All of the variables in this section Sec. 9.3.3 are dimensional variables; and, consequently, the tilde (•) notation applied in Sec. 9.3.1 is not applied here to denote dimensional variables. Multi-parameter theoretical spectra include both variance-preserving variable shape spectra and multiple peak (bi-modal) spectra. Goda-JONSWAP variance-preserving spectrum The dimensional variance-preserving Goda-JONSWAP one-sided wave spectrum is (Goda, 1985 or Chakrabarti, 1987) Snri(f) = a *
Hi /o
exp -1.25
exp[-(/Vo) 2 /(2r, 2 / 0 2 )] ;
(9.48a) where „_ _
r a = 0.07
0.0624 0.230 + 0 . 0 3 3 6 / - 0.185(1.9 + y ) - 1 ' if/0;
rb = 0.09
if / >/<,,
(9.48b) (9.48c,d)
and where 1 < y < 10 = spectral peakedness parameter; r; = spectral shape parameter (i — a or b in Eqs. (9.48c, d)); Hs = 4*Jm0 = dimensional significant wave height in Table 9.6; and /o = spectral peak frequency that may be computed by Eq. (9.26). When the spectral peakedness parameter y = 1 is substituted into Eq. (9.48b), Eq. (9.48a) is a relatively broad-banded spectrum similar to the generic two-parameter spectrum Sm{mo,a)Q,OL>,p = 5,q = 4) in Eq. (9.29c)
753
Real Ocean Waves
derived in Sec. 9.3.1 if Eq. (9.29c) is converted from radian frequency co to Hertzian frequency / by Eq. (9.16d) and Hs = 4^/mo from Table 9.6. When the spectral peakedness parameter y = 10, Eq. (9.48a) is a relatively narrowbanded spectrum. Examples of these two extreme values for y are applied to rubble mound breakwaters in Sec. 9.5 in order to evaluate the effects of spectral shapes on wave groups and the corresponding damage to rubble mound breakwaters. Goda (1998) modified the variance preserving coefficient a in Eqs. (9.48); but these modifications do not appear to be a good as the variance preserving coefficient a* in Eqs. (9.48). OCHI-HUBBLE six-parameter (bi-modal) spectrum Ochi-Hubble (1976) derive a theoretical six-parameter (bi-modal) wave spectrum consisting of both low and high frequency peaks that results in a double peaked spectrum. Each of these two frequency components require three parameters: viz., a significant wave height HSj, a modal peak frequency /o 7 , and a shape (or peakedness) parameter Xj where the subscript j = 1 for the low frequency components and j = 2 for the high frequency components. Each of the two three-parameter spectral peaks may be combined into the following single double peaked, one-sided spectral density function:
'nv
(/) = i E j=i
(^-f r(kj)
4A.J + 1
H
f
2TT/O,
4A./ + 1 x exp
(9.49a)
f
Ai = 2.72, X2 = 1.82exp(0.0277/,) = 1.82exp(0.108v^o). (9.49b,c) The dimensional significant wave heights Hs. in each of the two frequency regions j may be replaced by the total variance of the wave spectrum according to (9.50a)
H's• = " i +• H^ = 16m 0
Hi
HU] -«i Rfrf]
= 16mo-
(9.50b,c)
754
Waves and Wave Forces on Coastal and Ocean Structures
f/foi Fig. 9.13. Hinged wavemaker laboratory simulation of Ochi-Hubble 6 parameter wave spectrum (| = raw unsmoothed spectral density estimates; = smoothed estimate; and = theoretical spectral density function).
In order to compare laboratory or field spectra with Eqs. (9.49), it is convenient to scale Eq. (9.49a) by mo/2^/o, and obtain Snnif/foj)
_
mo/27T/ 0l 4((4A1 + l)/4)*l f, , (H^\2\ i +
r(M)
j
U J J \foJ
, 4/ 0l «4A2+l)/4)*2 J /o 2 r(x 2 ) 11
+
(X-\ (4A, + 1)
+
Oh.}2]1 \nn) I
exp
(J-\ -(4X2+1) \fo2)
-*[-m(*r] (9.51)
An example of a laboratory simulation of random waves by Eq. (9.51) in a 2D wave channel (vide., Fig. 5.1 in Chapter 5.1) by a hinged wavemaker at the O. H. Hinsdale-Wave Research Laboratory at Oregon State University in the USA is shown in Fig. 9.13. The parameters applied in the laboratory simulation in Fig. 9.13 were A.i — 2.72,^2 = 1.8,/o 2 //o, - 3.5,/ 0 l = 1/7Hz, fo2 = 1/2Hz, HS2/HS1 = 1.0 and mo = 0.011m2. 9.3.4. Spectral Directional Spreading
Functions
The ocean wave spectra reviewed in Sec. 9.3 and tabulated in Table 9.5 are unidirectional spectra. The Wave Project I and II hurricane wave force records
755
Real Ocean Waves
reviewed in Chapter 7.6.6 identify in-line and resultant force coefficients Cm and Cd • This distinction is due to the directionality ofthe random waves passing the instrumented offshore platform. The spectral models that incorporate this directionality of random waves are reviewed briefly (vide., Borgman, 1969a for an extensive review) A fundamental method for incorporating wave directionality is to assume a separation of variables model according to ~Sm(eo, 9) = Sm{o))D(e, n,-), D(0,n.-) = % ^ ,
(9.52a) (9.52b)
where Sm (a>) = the unidirectional one-side spectra models reviewed above and D(0,Tli) = a directional spreading function that may depend on several empirical parameters n , and/or radian wave frequencies w and that must satisfy the variance-preserving constraint required by 71
/. n
D(9,Tli)d9 = l
(9.52c)
so that the variance of the time series of random waves may be computed from m0 =
tf=
/
SnJ]{(o)D{e,Ui)dedo.
(9.52d)
Dirac delta distribution for uni-directional wave spectra For the special situation where the random waves propagate uniformly in the same direction, the spreading function may be represented by the Dirac delta distribution in Chapter 2.2.3 that is given by S(9-9 ) D(e,Yli)= \ 0 °\
(9.53)
27T
where 0$ = the direction of propagation of the unidirectional wave spectrum.
756
Waves and Wave Forces on Coastal and Ocean Structures
Cosine raised to even integer powers Mitsuyasu (1971) proposed a fetch-limited directional wave spreading function given by C(s) cos2s
o-e0
-n < 9 < n
D(9, n,o
n
C(s)cos2's(0-9o) 2(2s-l)
C(s) =
it
r
2(j
+
2
j)
r(2s + i ) '
C(s) =
<
_
(9.54a) (9.54b)
1 l\s + l)
(9.54c,d)
where 9Q = principal direction of propagation of the unidirectional wave spectrum, the parameters n,- = J and 5 = empirically determined integer constants that control the amount of angular spreading about the principal direction #o; and the Gamma function T(«) is defined by Eq. (2.6a) in Chapter 2.2.5. The parametric dependency on the empirical parameter s of the width of the spreading function is illustrated in Fig. 9.14. When the integer parameters s = s = 1, then T2(2) = 1, T(3/2) = *Jn72, r(3) = 2 and Eqs. (9.54) reduce to (Borgman 1969a and 1972b) C(s = 1) cos" 0(0,n,-) s > f = 1 =
'9-90'
-it
Cis = 1) cosz(9 - do),
1 C(s = 1) = - , it
2 C(s = l) = -= n
< 9 < n,
n < 9< - -z
2C(s = 1).
TC
(9.55a) (9.55b) (9.55c,d)
-180-120 -60 0 60 120 180 (9-e„) (deg) Fig. 9.14. Parametric dependency of cos2i(fl — 9Q) spreading function on the shape parameters.
757
Real Ocean Waves
Mitsuyasu, et al. (1975), Goda and Suzuki (1975), inter alios observe a dependency of the directional spreading parameter s in Eq. (9.54a) on the wave frequencies co and scale the wave frequencies co and spreading parameter s by CO
coU
OOQ =
11.5
co0U
so
-5/2' CO,1
(9.56a-c)
o where coo — spectral peak frequency defined in Eq. (9.26); U = wind speed in units consistent with the units of the gravitational constant g and so — directional spreading function parameter at the spectral peak frequency cooThe following parametric dependency of the shape parameter s on frequency co is recommended: CO
s
,
coo
so
co < coo,
(9.56d)
CO >
(9.56e)
2.5 I
,
COQ.
CO
Fourier series Longuet-Higgins, et al. (1961) develop a Fourier series expansion for a directional spreading function given by N
1 D(0, 111 = Wn) = — + J^ Wn[an cosnO + bn sin/i0],
(9.57)
n=l
that satisfies Eq. (9.52c) and where W„ = Bayesian weighting functions. This method is very computationally intensive because of the large number of coefficients that must be estimated from the data. Von Mises circular normal (Mardia, 1972). Von Mise proposes D(9,Tli =a) =
exp[a cos(# — 6Q)]
(9.58a)
2TZIQ{O)
\e-Oo\<
=
71
2'
where /<)(•) = modified Bessel function of the first kind of order zero defined by Eq. (2.52) in Chapter 2.4.3 and where the shape parameter a may be related
Waves and Wave Forces on Coastal and Ocean Structures
758
1.4i — j _ . —f_ -\— a = 10-i-j--H1.2 — i - « = 8 / NT~ i -t //" a=6 ^ 1 _ _ _ u . —fM" CD 0.8 — h a = 4 ~ ^ .J_l_ ~3T^ Q 0 . 6 — h - — r- y a -- ^v\ - ; - r 0.4 H___;_. afO
0.2 0 s^^s
__L
- 1
-fi!
i
— J — i —
H^Sf^
-90 -60 -30 0 30 60 90 (6-e0) (deg) Fig. 9.15. Parametric dependency of circular normal spreading function on the shape parameter a.
to the shape parameter 5 in Eqs. (9.54) by 2arccosfl - — } = 4arccos[(0.5)1/2*].
(9.58b)
Figure 9.15 illustrates the parametric dependency on the shape parameter a of the circular normal directional wave spreading function.
SWOP wave spreading function A wave directional spreading function derived from stereo photographs of ocean waves is the SWOP model (Stereo Wave Observation Project, Cox and Munk, 1954a, b) given by 1 + I 0.50 + 0.82 exp D(6, &) = - • it
(O =
+ 0.32 exp
KI)
-*©' 0)'
cos 20
it
cos46>
(9.59a) (9.59b)
U5'
where Us = wind speed measured at 5 m (16.4 ft) above the sea surface.
759
Real Ocean Waves
Wrapped normal Borgman (1969a) proposes the following two equivalent equations for the wrapped normal directional wave spreading function: I
oo
— + £ exp 2TC
(non)2
cos[n(# — 6>o)L
(9.60a)
n=l
0(0,0-,) = 00
£ n=—oo
exp •
(6> — 6>o — 2;r/z)2 2^| CT^V 27T
TT |0-0ol<-,
(9.60b)
where a„ = the standard deviation of the water surface r](t). 9.3.5. Confidence Intervals for FFT Estimates Bendat and Piersol (1986) and Otnes and Enochson (1972) derive formulas for placing confidence intervals on FFT estimates. Borgman (1972a) reviews applications for computing confidence intervals for ocean wave spectra. One method for computing confidence intervals assumes that the real and imaginary components of each complex-valued raw (unsmoothed) FFT coefficient Bm are independent Gaussian estimates; and, consequently, each raw FFT estimate Bm comes from a sampling distribution that is a chi-squared / variable (Bendat and Piersol, 1986, Chapter 8.5.4, Eq. (8.151)) with two degrees of freedom (i.e., one degree for each real and imaginary component of the complex-valued raw (unsmoothed) FFT coefficients Bm); viz., Br,
B„
2 '
(9.61)
where Bm = the complex-valued FFT coefficient computed by the expectation operator E[»] in Eq. (9.1) in Sec. 9.1 for a stationary (ergodic) Gaussian process. A smoothed FFT estimate Bm may be computed by applying an averaging method; e.g., averaging at the same frequency fm several raw (unsmoothed) FFT estimates Bm computed from several shorter sequential subrecords obtained from a single long continuous record (viz., segment averaging); averaging several raw (unsmoothed) FFT estimates Bm over several adjacent frequencies from a single long continuous record (viz., box-car
760
Waves and Wave Forces on Coastal and Ocean Structures
averaging); inter alia. A smoothed two-sided spectral amplitude estimate \Gm(m)\df may be computed by averaging according to \Gm(m)\df
1 Ns = — J^ \Bm\2nr, s
(9.62)
1
where Ns < N = total number of raw (unsmoothed) FFT estimates Bm that are averaged over several A^ shorter subrecords at the same frequency m df (segment averaging) or the total number of raw (unsmoothed) FFT estimates Bm that are averaged over several adjacent frequencies from the same single long continuous record (box car averaging). The (1 — a) confidence interval for the two-sided spectral amplitude \Gm{m)\df with Nd = 2NS degrees of freedom may be computed from (Bendat and Piersol, 1986, Chapter 8.5.4, Eq. (8.159)), 2
< \Gw(m)\df
<
^Nd;a/2
j
•
( 9 - 63 )
X-Nd;l-a/2
Tables for the values of xjy .. maybe found in Bendat and Piersol (1986), Zelen and Severo (1968), inter alios. For example, a = 10% for 90% confidence intervals and Ns — 8 for segment averaging over 8 subrecords or box car averaging over 8 adjacent frequencies (i.e., 4 frequencies before m df and 4 frequencies after m df), so that Nd = 2NS = 2(8) = 16; and Xi6;o.os = 2630 a n d Xi6;0.95 = 7 - 9 6 (Bendat and Piersol, 1986, p. 524, Table A.3) An example of the application of confidence intervals at the 90% confidence level is illustrated for the single wave record CARLA85 from Hurricane CARLA in Sec. 9.7 below.
9.4. Probability Functions for Random Waves Two probability functions that describe linear random waves are the cumulative distribution function (cdf) P(x) and the probability density function (pdf) p(x) for the Gaussian (Normal) and the Rayleigh distributions. The cdf is the probability that the random variable of time x(t) is less than or equal to some value £, say; i.e., (Papoulis, 1984) />(£) = Prob[x(t) < | ] ,
(9.64)
761
Real Ocean Waves
where P ( - o o ) = 0.0
and
P(+oo) = 1.0.
The probability that x(t) lies between the two values x\ and x2 is Prob[x\ < x(t) < x2] = P(x2) - P(xi) = /
p(x)dx,
(9.65)
where P(x\) < P(x2)
if
xi < X2
and where — OO < XI < X2 < + 0 0 .
The pdf p(x) is related to the cdf P(x) by the derivative (Papoulis, 1984) pM = * £ £ ) . ax The total area under the pdf curve in Fig. 9.16b is unity because
(9.66)
+00
p(x)dx = P(+oo)-P(-oo) /
= 1.0
(9.67)
-OO
and the probability that x(t) = % = a specific value of x(t) is zero from Eq. (9.65) because P(x(t) = §) = J p(x)dx = 0. Specific values of a random variable may not predicted; only the probability of having values in an interval by Eq. (9.65) is possible. It also follows from Eq. (9.67) and Eq. (9.65) that Prob\$ < x(0] = /
p(x)dx = 1 - P(f) = 1 - /
Jt;
p(x)dx.
J-OO
The mean value ix\ (x) (or expected value or average value) of the random variable x(t) is defined by the integral in Eq. (9.14a) for n = 1 according to +00
/
xp(x)dx -oo
(9.68)
762
Waves and Wave Forces on Coastal and Ocean Structures
and the «th probability moment ixn{x) of the random variable x(t) may be computed from the integral in Eq. (9.14a) by oo
/
xnp(x)dx.
(9.69a)
-00
Four moments computed by Eq. (9.69a) that appear repeatedly in probabilistic analyses are the following: mean IJL\{X): oo
xp{x)dx, /
(9.69b)
-00
variance a^ about the mean faix): oo
/
(x-lxl(x)fp(x)dx,
(9.69c)
-00
skewness fiiix): oo
/
x3p(x)dx,
(9.69d)
x4p(x)dx.
(9.69e)
-oo
kurtosis /U,4(x): kurtosis IIA{X): oo
/
-00
The mode or most probable value xmode of the random variable xit) may be computed from the maximum value of the pdf by the derivative — dx
= 0 , X = Xmode-
(9.70)
The median value xmed of the random variable x(t) may be computed from the cdf by the value at P{x) = 0.5,
x = xmed.
(9.71)
763
Real Ocean Waves 0.8
-I i
1 i
[—[; i
1 i
1 i-
i-
1 Problx, £ x £ x2 J
0.6
/ 4-U-U-^^Ll V m
0.4
n& fpw*
0.2
i J --4. 1 1
+~\~-Jt
ll 4
—1—,—: - ^
-6
-4
,i
-2
1 J 1 -1
t
1 i 1 1 U
0 ^ 2
Fig. 9.16a. Probability properties of a non-zero mean Gaussian pdf (ix\ (x) ^ 0).
1
•
ed f
— i
0.8 *—T
p .li-
0.6
r , — J.
0.2
r —
-
T i
2
i
—
-
—i/
4 •-I
\ / (._/
1
t fr 1\
/
i
0.4
•t
/ 4
S=£_!
/-
f
1
T
_ U ^
^ , =*_*,= *_„,.-0.5
1
0
1 2 x Fig. 9.16b. The mean ji\, the mode xmode ar>d the median xme,i for a symmetric Gaussian distribution with n\ = xmmie == xmej = 0.5 and standard deviation ax = 0.5.
TheprobabilitypropertiesforP(£)inEq.(9.64)andforPro6[.xi < *(0 < * 2 ] in Eq. (9.65) are illustrated in Fig. 9.16a for a symmetric Gaussian distribution function with a non-zero mean. The mean fxi (x) in Eq. (9.68), the mode xmoa-e in Eq. (9.70) and the median xmed in Eq. (9.71) are illustrated in Fig. 9.16b for a symmetric Gaussian (Normal) distribution function with n\(x) = xmoa-e = xmed = 0-5 and a standard deviation ax = 0.5. The nth probability moment jln (x) about the mean \i \ (x) is called a central moment that may be computed from
+00
/
(x-^(x))"p(x)dx. -00
(9.72)
764
Waves and Wave Forces on Coastal and Ocean Structures
The variance a2 about the mean ju-i(x) may be computed from /Z2OO in Eq. (9.72) by
A2OO = ol = Ey{x - Mi(x))2J = E[x2] - 2fi\(x)E[x] + ix\{x) = E[x2] -
rfix).
(9.73)
The root-mean-square (rms) of the random variable x(t) is the positive square root of the mean square or standard deviation ox. These probability properties for a Gaussian (Normal) pdf with a non-zero mean are illustrated in Figs. 9.16a and b.
Moment generating and characteristic functions (Papoulis, 1984, Chapters 5-4 and 5-5) Moments fin(x) of a random variable x(t) from Eqs. (9.69) may also be computed from a moment generating function *XC?) defined by oo
tyx(s) = E[exp(sx)]=
I
/J—00 -00
p(x p(x)exp(sx)dx.
(9.74)
If s = ico, then the characteristic function irx(co) of a random variable x(t) is the Fourier transform of the pdf p{x) and may be computed from Eq. (9.74) by yjrx{(o) = £[exp icox] = I
p(x)exp icoxdx.
(9.75)
• / • ./ — 0 0
Moments p,n(x) of the random variable x(t) may be computed by differentiating the moment generating function ^ ( s ) n times with respect to 5 according to Vnix) =
dnl
-*As) d~sn
"I
/-00
= / _ls=0
J-oo
xnp(x)exp(sx)\s
= 0dx
= E[xn].
(9.76)
765
Real Ocean Waves
This may be verified by expanding exp sx in a Taylor series about x = 0 (vide., Eq. (2.18), Chapter 2.3.3); (sx)2 exp sx « 1 + sx -\—— 2!
1
(sx)3 3!
1
1
(sx)n n\
1
,
that may be substituted into Eq. (9.74) to obtain (Papoulis, 1984, Chapter 5-5), f°°
(
As) = pix) l+sx
*
L {
(sx)2
+
(sx)2
+
+
(sx)m
+
^r ^r --- ^r
r°° r°° s2 r°° 2 = / p(x)dx + s J xp(x)dx + — / x p(x)dx J—OO
_3
J—00
/*oo
\ ,
+
---)dx
*•' J—OO sm
/-oo
+ — / x3p(x)dx + . . . + — / xmp(x)dx H -00 ' " • •/—00 3! J-oo ml y_oo , 2 ^ 2 , 3/^3 , , m Mra , /*0+*/*l + 5 z—- +^ J —- + ---+sm—+ ••• 2! 3! m!
< 9 - 77 >
= £ HX n=0
and Eq. (9.76) follows. The pdf p(x) of the random variable x may be computed from the Fourier transform of the characteristic function irx(co) (Papoulis, 1984, Chapter 5-5), according to i r' 0c 0 p(x) = —i fx(co) exp ~ (icox)dco, 2n J_oo
(9.78)
where the normalizing constant 2n has been placed with Eq. (9.78) (cf, Eqs. (9.2) in Sec. 9.2).
9.4.1. Gaussian (Normal) Probability
Distribution
Linear random waves in deep or finite depth water may be approximated by a linear boundary value problem (BVP) from Chapter 4.2 with solutions that represent a Gaussian (Normal) process. A Gaussian (Normal) probability
766
Waves and Wave Forces on Coastal and Ocean Structures
density function (pdf) for a random variable x (t) with mean fi \ (x) and standard deviation ax is given by 1
p(x) =
''2.11ax
-(x - ixijx))2-1 2a}
exp
(9.79a)
with a cumulative distribution function (cdf) defined by P(x)=
/
p{x')dx' =
J—oo
/
exp
fn(x'))2
(x'-
2
y/2lCCTx J—oo
dx'. (9.79b)
The mean /JL\ (X) of X ( 0 may be computed from Eq. (9.68) or from Eq. (9.69b) with n = 1; and the nth central moment [in (x) of x(t) may be computed from Eq. (9.69a). The variance a} of x(t) is given by the second central moment (n = 2) from Eq. (9.69c) as oo
/
(x' - in (x'))2p(x')dx'.
(9.79c)
-oo
Because the Gaussian (Normal) pdf is symmetric, the odd central moments are zero; i.e., n = 1, 3, 5, . . . , etc. The even central moments may be computed from oo
/
(x' - ^{x'))2*1
p{x')dx'
-00
= {In - \)a2n;
n = 1,2,3,...
(9.79d)
The central moment value for the kurtosis or fourth moment for n = 2 from Eq. (9.79d) is jjL4(x) = 3CT*. It is convenient to normalize Gaussian (Normal) random variables in order to obtain a dimensionless normalized random variable t;{t) for a Gaussian (Normal) random process that has zero mean (£ii(£) = 0) and unit variance (a2 = 1) and that is defined by (cf, Eq. (9.14c) in Sec. 9.3) x(t) -
?(*) =
Or
ii\(x)
(9.80a)
161
Real Ocean Waves
so that the pdf from Eq. (9.79a) for the dimensionless normalized random variable f (?) reduces to
p(0 = ^ L e x p ( - ^ .
(9.80b)
Because Eq. (9.80b) is symmetric with respect to the mean /2i(f) = 0, the mean of the dimensionless normalized random variable f (?) is equal to zero. Then th central moment £«(£) about the mean /zi(£) of £(?) may be computed from Eq. (9.72) or, alternatively, for the dimensionless normalized random variable £(?) by oo
UniX)
/
lnp(X)dl
(9.81a)
-00
The mean/ti(^)fromEq. (9.81a)is equal to zero; and the variance/12(C) = °f is equal to unity. The odd moments from Eq. (9.81a) are equal to zero; and the even moments may be computed from Eq. (9.79d). Consequently, the kurtosis £4(£) for a normalized Gaussian (Normal) random variable £(?) is equal to three; and the excess of kurtosis /X4(f) for a normalized random variable is defined as oo
/
<X?p{l)dl - 3,
(9.81b)
-00
where f is a normalized (possibly Gaussian) random variable defined in Eq. (9.80a). For a zero-mean, unit-variance Gaussian (Normal) random variable f (?), the cdf is ;?
_ -
(f
1
=/-„' *H
,+Brf
(^)
(9.82)
where the Error Function Erf (•) is defined in Eq. (2.9a) in Chapter 2.2.6. The pdf and cdf for a zero-mean, unit-variance Gaussian (Normal) random variable f (?) are illustrated in Fig. 9.17. Very useful polynomial and rational approximations to the pdf Eq. (9.80b) and cdf Eq. (9.82) for a zero-mean, unit-variance Gaussian (Normal) random variable f (?) are given by Zelen and Severo (1968). A polynomial approximation for the pdf Eq. (9.80b) of £(?)
Waves and Wave Forces on Coastal and Ocean Structures
768
- 4 - 3 - 2 - 1 0 1 2 3 4
C Fig. 9.17. Zero-mean, unit-variance Gaussian cdf P(?) and pdf p(t;).
for - o o < f < oo is (Zelen and Severe, 1968, p. 933, Eq. (26.2.21)) p(f) = (b0 + b2S2 + b^A + hi;6 + btf* + Z^oC10)"1 + e(f),
(9.83a)
where b0 = 2.5052367,
Z>6 =0.1306469,
fe2 = 1.2831204, fc8 = -0.020249, b4 = 0.2264718,
bio = 0.0039132,
|e(f )l < 2.3 x 10" 4 . A polynomial approximation for the cdf Eq. (9.82) of £ (?) for 0 < £ < oo is (Zelen and Severe, 1968, p. 932, Eq. (26.2.19))
P(jr >0) = l--(l+d^+d2^2
+
+ d^3+d4;A
6\-16 rf5?5+4r) +e(C)
(9.83b)
and for —oo < f < 0 P ( - o o < f <0) = - ( l + J i i a + ^ 2 l f | 2 + * K
3
6\-16
+ ^4ia4 + ^5K| 5 +^6lf| 6 )
(9.83c)
769
Real Ocean Waves
where d\ = 0.0498673470,
d4 = 0.0000380036,
d2 = 0.0211410061,
d5 = 0.0000488906,
d3 = 0.0032776263,
d6 = 0.0000053830,
|e(C)| < 1.5 x 10 - 7 . Probability plots of Gaussian (Normal) random variables may be easily constructed by the inverse rational approximation to the random variable £ in Eq. (9.83b) from (Zelen and Severo, 1968, p. 933, Eq. 26.2.23) Co + C\t + C2tl
£/>< = t -
1 +d\t + d2t2+dit3 + «(P), In
(ir1p)
1
'- 1 " 1 ?
0.5 < p < 1.0,
0 < p < 0.5,
(9.83d) (9.83e)
where co = 2.515517,
d\ = 1.432788,
c i = 0.802853,
d2 = 0.189269,
c2 = 0.010328,
d3 = 0.001308,
\e(p)\ < 4.5 x 10" 4 . An example of a comparison between linear and nonlinear simulated and measured ocean waves plotted on Gaussian (Normal) probability paper is shown in Fig. 9.18 (Hudspeth and Chen, 1979). The differences between the nonlinear random waves and Gaussian (Normal) random waves is seen to be significant only outside ± two standard deviations ±2o$ (i.e., the upper and lower tails of the distribution). The symmetric Gram Charlier pdf does not accurately model the asymmetry of nonlinear random wave crests and troughs (Hudspeth and Chen, 1979). For a Gaussian (Normal) random variable f(f) that is Normally distributed with zero-mean and unit-variance 7V[0, 1], the amplitudes A are Rayleigh
Waves and Wave Forces on Coastal and Ocean Structures
770
I
I
I I
i
i
i
I
I
I
I
i™ i
I
I T
4
iy—
/x -
2&F
3
-
2 1 0
-
j/1*^
GRAM-CHARLIER X MEASURED REALIZATION
<*T^ -1 •2
4T
k N o • MEASURED a • SCOTT
-
-3
-
-4
-/
A j B *
O.OI
1 .
O.I
A NEUMANN
o • BRETSCHNEIDER PIERSON-MOSKOWITZ 1 1
I I
I
10
1
1 1 1 1 1
30 50
70
1
I I
90
I I
99
II
99.9 99.99
P(0 Fig. 9.18. Comparisons of linear (L) and non-linear (N) DSA simulations (Hudspeth and Chen, 1979).
distributed and the phase angles a are uniformly distributed £/[0,2JT], where U should not be confused with the italic U (•) that is the Heaviside step function defined in Eq. (2.1) in Chapter 2.2.2. A Gaussian (Normal) random variable f ( 0 may be defined by t;(t) = a cos(a>t) + fi sin(wO = A cos((ot - a),
(9.84a,b)
where the amplitudes [a, 0] = N[0,1]; the amplitude A is Rayleigh distributed (vide., Sec. 9.4.2) and the phase angle a is uniformly distributed U[0,2n]. The relations between the amplitudes [a, ft] and the amplitude and phase angle [A, a] are or = A cos a, a \ = 0, E\ ~
/? = Asina
(9.84c,d)
„2
E { a- 2 \ = 1,
P2
(9.84e,f)
111
Real Ocean Waves
where £{•} is the expectation operator defined in Eq. (9.1). The auto-correlation function C^(t) for Eq. (9.84a) may be computed by C ff (T) = E{?(f)?(f + T)} = E{(acoscot + /3 sin cot) (a cos co(t + r) + fi sm.oo{t + T))} = - coscor[E{a2} + E{fi2}] + - cosco(2t + r)[E{a2} - E{p2}] + £{a}£{/S}sin(w(2?-|-T) =
(9.84g)
COS<WT.
Alternatively, the auto-correlation function Q - J ( T ) for Eq. (9.84b) may be computed by C f f ( r ) = E{$(t)£(t + T)} = E{A cos(cot - a) A cos(co(t + r) 1 , 1 Z = -cosoor E{A } + -cosoo(2t + 1
-a)}
, r)E{Azcos2a]
9
+ - s'mco(2t + T ) £ { A Z sin2a} = -cos«r[£{a2} 1 + - sin
+ E{p2}] + -cosco(2t + t)[E{a2} - E{fi2}] r)E{a}E{p} (9.84h)
COS COT.
The value of the auto-covariance function C^ (0) = 1 = unit variance. To verify that the amplitudes A are Rayleigh distributed and that the phase angles a are U[0,2n] when [a, ft] are Normally distributed N[0, 1], the equality of differential pdf areas requires that p(a)p(P)dadfi
=
p(A)p(a)dAda.
(9.85a)
The Jacobian of the transformation [a, /?] —> [A, a] is 9(«, J8) 3(A, a)
cos a sin a
—A sin a A cos a
= A
(9.85b)
772
Waves and Wave Forces on Coastal and Ocean Structures
so that 9(g,j8)
dad ft = J
(9.85c)
dAda
d(A,a)
and Eq. (9.85a) becomes (Papoulis, 1984, Chapter 6-2,) p(a)p(fi)dadfi exp
(-T)
da exp (
=
p(A)p(a)dAda,
) dp
-M-THI£I
V27r v2;r {Normal pdf N[0,1]} {Normal pdf N[0,1]} = {Rayleigh pdf}
{£/[0,2jr]Uniformpdf}. (9.85d)
9.4.2. Rayleigh Probability
Distribution
The Rayleigh probability density function (pdf) for the wave amplitudes A(= H/2) is given by
p(A) = - j exp a'
U(A), •
-
(
a > 0,
(9.86a)
-
2 \a
where a = Rayleigh parameter (Hoffman and Karst, 1975 ) and £/(•) = Heaviside step function defined in Eq. (2.1) in Chapter 2.2.2. The Rayleigh cumulative distribution function (cdf) for wave amplitudes is
P(A) = 1 - e x p
U(A),
a > 0.
(9.86b)
The Rayleigh pdf Eq. (9.86a) and the Rayleigh cdf Eq. (9.86b) are illustrated in Fig. 9.19.
773
Real Ocean Waves
Fig. 9.19. The Rayleigh pdf p(A/a) and the Rayleigh cdf P(A/a).
The following Rayleigh distributions with four different Rayleigh parameters a in Eq. (9.86a) may be found: 2"
TZA
p(A) = —=- exp 2H\ p(H) =
2H
H exp 4ra0
U(A),
yck > 0,
(9.87a)
4 V/W H
exp
p(A) = — exp mo p(H) =
--f-y
U(H),
tlrm,
A /mo H
U(A),
U(H),
Hrms > 0,
m0 > 0,
mo > 0,
(9.87b)
(9.87c)
(9.87d)
where IXA= average of the amplitude A; Hrms = root-mean-square wave height; mo = cr% — variance of the time series for the water surface elevation rj(t) and [/(•) = Heaviside step function in Chapter 2.2.2. When the Rayleigh parameter a in Eq. (9.86a) is defined to be equal to the standard deviation of the time series an = ^/mo, then with the following change of variables da d$ = § s/mo' /mo and equating differential areas of the pdf's da p{$) = p{a)—,
774
Waves and Wave Forces on Coastal and Ocean Structures
yields a dimensionless form of the pdf in Eq. (9.86a) given by
UG)
/?(§) = § exp
(9.88)
where U (£) = Heaviside step function. The four forms of the pdf in Eqs. (9.87) may be derived from Eq. (9.86a) by solving for the generic Rayleigh parameter a as outlined below. Average wave amplitude a = fiA In order to relate the Rayleigh parameter a to the average wave amplitude /ZA, the percent or fraction \/n of wave amplitudes A(= H/2) greater than wave amplitude A\/n{= H\/n/2) may be computed from Eq. (9.86a) by
1
f°°
f 1 fM/n
- = / pia)da = exp - " J Aim 2\ and the natural logarithm of Eq. (9.89a) gives
(9.89a)
a
M,n
= j2Wn), n > 1. (9.89b) a The average wave amplitude A \ /„ of those wave amplitudes greater than A \ /„ may be determined from /•OO
A\/n
\
/-00
p(a)da = I
J A\jn
ap(a)da,
(9.90)
JA\/„
that may be integrated by parts to obtain A\jn
aV21n(ra) JAU„
1 (a\2 da . 22 v\a «y
by Eqs. (9.89). Finally, integration of the last integral in the equation above gives ^
= V2hr(^) + « 7 | [ 1 - Erf(Vm^0)]
(9.91a)
= V21n(«) + nJ^Erfc(Jh^)),
(9.91b)
775
Real Ocean Waves
where Erf(«) = the error function and Erfc(») = complementary error function (vide., Eqs. (2.9) in Chapter 2.2.6). The Rayleigh parameter a may now be related to the average wave amplitude defined as HA, by setting n — 1 in Eqs. (9.91) and obtaining M/(n=l) _ M_ _ AM_ _ [i*_ V2 a a a (9.92)
a = VA\I —
so that substitution of Eq. (9.92) into Eq. (9.86a) gives Eq. (9.87a). The average wave amplitude A\/(n = \) = A \ = \x& may also be computed from the table of integrals given by Gradshteyn and Ryzhik (1980 p. 337, Eq. (3.461.2)), according to _
/•OO
/"OO
rOQ
A\/n I p(a)da J A\jn
= A \ I p(a)da = I ap(a)da, JO JO [°° /a\2 1 /flx2 f°° \iA = I v( —y I exp ( — ) da = a I Jo « 2 \a/ J0 VA
= a
f°° p„ 2 t exp -brdt Jo
t exp
1 2
1 dt,
a(2/3-l)!! / F = ——^-—^—J—, 2(ib)P y b'
where £ = 1; b = 1/2; and the odd-factorial = (2)8 - 1)!! = 1 • 3 • 5 • • • (2/3 — 1). Substituting these values for b and /J yields, again, Eq. (9.92)
MA
=aJ^-
Alternatively, Eqs. (9.91) may be derived directly from special Gamma functions defined in Chapter 2.2.5 according to poo
^l/« / p(a)da J A\/n
poo
= I J A\/n
ap(a)da A
l/n
2a2 where the generalized incomplete Gamma function T{f5,zo,z\) Eq. (2.7) in Chapter 2.2.5 as
r(j8,z0,zi) = r(/3,zo)-ro3,zi)
is defined in
776
Waves and Wave Forces on Coastal and Ocean Structures
and where the incomplete Gamma function T(fi, z) is defined in Eqs. (2.8) in Chapter 2.2.5 as oo
/
Y(fi,z)=
f Jo
t?-1
exp(-t)dt,
t^expi-Odt.
Substituting for r(fi,zo,zi) results in
with /3 = 3/2; zo = 0 and z\ =
A2/n/2a2
It
r
r(/3,o) = r
r(lM =r X 2a2
\ftexp(—t)dt
-L
V7exp(—t)dt
Ai/J**2
^ ^ =
+
^[i-Erf(V5w)]
VEfrj + ^ E r f c ( V ^ } n
2
giving, again, Eqs. (9.91). Root-mean-square wave height a = Hrms In order to relate the Rayleigh parameter a to the root-mean-square wave height Hrms, make the following change of variables in Eq. (9.86a): v= a ,
dy = 2ada = 2^/yda
and equate the differential areas of the pdf's p(a)da = p{y)dy giving da p{y) = p(a) — dy 1 2^2 exp
(9.93) y '2a2
U(y).
111
Real Ocean Waves
The average wave amplitude squared A2 may now be related to the Rayleigh parameter a2 from A2 = A2ms =y=
f
yp(y)dy
Jo
hQX*[-iAdy
=1 z 2a..2
(9.94)
and the Rayleigh parameter a may be evaluated by Arms
a=
^f
Hrms
=
^M
,
=
^
=
/n
°-
n-
,..
(9 95a d)
- "
Substituting Eqs. (9.92 and 9.93) into Eq. (9.86b) gives £lll
=2 / ^
+ «Erfc(yinOO),
(9.96a)
^
= V^W
+ ^Erfc(V/ir^)),
(9.96b)
— ^ = V81n(n) + nV2;rErfc(Vln(«)).
(9.96c)
The significant wave height Hs or the average of the highest 1/3 wave heights #1/3 may be computed from Eq. (9.91b) with n — 3 according to #1/3
Hs
- ~ = -^= 2.00215, 2a 2a till, 2a = - ^ 2.00215
(9.97a) (9.97b)
V
J
Substituting 2a from Eq. (9.97b) into Eq. (9.91b) and defining A 1/n = #i/„/2 gives
778
Waves and Wave Forces on Coastal and Ocean
Structures
Substituting a from Eq. (9.95c) into Eq. (9.97b) gives Us (= #1/3) = 4.0043 Vm^.
(9.98b)
Mode a = Amode To compute the mode (or most probable) value of A(= H/2), the maximum value of Eq. (9.86a) occurs when dp (a) = 0,
da
a = Amode
(9.99)
giving Amode
= OC,
Hmode
=2«
(9.100)
and "•mode
I **
Hmode
flA
V TC '
Hrms
1
Hmod^=Hmod1=2^
V^o
an
(9.101a,b)
2' *node={2sml5)-y
(9.10lc,d)
Hs (= H\/i)
Mean = median The mean or median of A occurs when P(a) = 0.5 in Eq. (9.86b). The natural logarithm of Eq. (9.86b) for P{Amode) = 0.5 with n = 2 is A-med
a
= y/2 ln(2)
(9.102)
giving A ^ IAa
=
^ ^ ^ ^ ^timed _ /41n(2) V JT
=
^
ln(2) = A(2).
( 9 1Q3
^
These values for the Rayleigh parameter a in Eqs. (9.86) that are equal to various probable wave heights are summarized in Table 9.6. The spectral energy-based significant wave height Hmo from Sec. 9.3 may be computed directly from the variance ofa time series a^ = mo according to Hmo = 4^/mo.
Real Ocean Waves
Table 9.6. Dimensionless wave heights for four values of the Rayleigh parameter a. Height H
0)
# s ( = #1/3) #1/10 #1/100
(3)
(4)
(5)
l
(2)
= 1/2)
H\ (= mean)
H
H tirms
"mode Hmed(P
H
H MA
2.0 3.19497 4.06198 5.32423
2.0
(2.00215) _I
V2 Vln(2) •Jit
V81n(2)
V2 In (2) 2.00215
-J2n
2 1.41573 1.79992 2.35924
4.0043 5.09094 6.67293
2.00215 1.0 1.27137 1.66644
Consequently, all of the values listed in column 5 in Table 9.6 may be approximated by dividing by 4 all of the values in column 4 in Table 9.6.
Generalized Rayleigh distribution Ochi (1978) generalized the Rayleigh probability density function (pdf) to the following three-dimensional generalized Rayleigh distribution with four dimensionless parameters: p(A) = -—-XCmA(Cm-1}exV-kc(Ac r(m)
+ lxc)l0(CX.2fix)U(A),
(9.104)
where C, m, X and n, are constant dimensionless parameters, A = an independent random variable for the maxima and 7o(») = modified Bessel function of the first kind of order zero (vide., Eq. (2.52) in Chapter 2.4.3). All of the families of pdf's that may be recovered from Eq. (9.104) are illustrated in the three dimensional parametric plot in Fig. 9.20. In order to recover the Rayleigh pdf Eq. (9.86a) from Eq. (9.104), set C = 2,
m — 1, A.
1
H= 0
and because T(l) = 1 and 70(0) = 1, then Eq. (9.86a) follows.
780
Waves and Wave Forces on Coastal and Ocean Structures
/IVIAXWELL/" HYDROGRAPH
Lai/ T~7
i
lEXPONENTIALl / GAMMA
-w IWEIBULLI
Fig. 9.20. Families of pdf s derived from the generalized Rayleigh pdf (Ochi, 1978).
In order to obtain the generalized two dimensional Gamma pdf from Eq. (9.104), set the dimensionless parameter /x = 0 and obtain
p(A) =
C
_kCmA(Cm-l) T(m)
exp
_(XA) C £/(A).
(9.105)
The remaining three dimensionless parameters in Eq. (9.105) must be estimated from data; and, in general, these estimated parameters should be the most efficient estimators that satisfy the Rao-Cramer condition (Kendall and Stewart, 1961). Ochi (1978) gives an algorithm for estimating the dimensionless parameters from data based on the Stacy-Mihram method (Stacy and Mihram, 1965). The solutions obtained from his algorithm may not always be stable or reliable when the parameter m is large and when A is less than zero. The generalized two dimensional Gamma pdf in Eq. (9.105) is very useful for evaluating maxima from very broad-banded spectra or from weakly nonlinear waves with high frequency nonlinear harmonics.
9.4.3. Distribution of the Maxima Cartwright and Longuet-Higgins (1956) apply Rice's theory of noise (1954) to evaluate the distribution of the maxima of a Gaussian random time series representing surface gravity waves and the motions of a ship. In their theory,
781
Real Ocean Waves
moments mn are computed from one-sided spectral densities by Eq. (9.40) in Sec. 9.3 and not from probability density functions by Eqs. (9.69) in Sec. 9.4. A Gaussian random time series with zero-mean (e.g., surface gravity waves) and its first two temporal derivatives are given by £lit) =x{t)
= Y^Rm m
cos((Dmt -
(9.106a)
am),
& ( 0 = x(t) = - ] T comRm sm(comt m
am),
(9.106b)
& ( 0 = x(t) = -^co2mRm
am),
(9.106c)
cos(a>mt -
where the overdots x(t) denote ordinary temporal derivatives and where the amplitude Rm at the radian frequency com may computed from a one-sided spectral density function Sw by R2 -f
(9.107)
= S(com)dco
and the random phase angles am are uniformly distributed U[0,2TI]. The statistical averages or expectations of Eqs. (9.106) are defined by the following moments: 3,v =
Emj] mo 0 —ni2
0 m2 0
— nt2 0 iri4,
(9.108)
where £ [ • • ] = the expectation operator from Eq. (9.1) in Sec. 9.1. The joint probability density function (pdf) for the zero-mean Gaussian random variables in Eqs. (9.106) is
P(£i,&,&) (2TT)- 3 / 2
eA/mo/W2m4 exp
1
~2?
ni2
m0
+
2T/T
(9.109)
782
Waves and Wave Forces on Coastal and Ocean Structures
where the spectral bandwidth parameter e from Eq. (9.41) in Sec. 9.3 is m%
ez = 1 -
(9.41)
mom.4 The maxima of Eq. (9.106a) occur when &(O = 0,
&(0<0.
(9.110a,b)
Normalizing Eqs. (9.106a, c) by the spectral moments mo and rm, according to x] = -4==,
« = -7=
(9.111 a,b)
and with the Jacobian p(rj,0,u)
= p($u 0 , ^ ) 7
3(fl,0,§ 3 )' _ 3(r/,0,«) _
= p(£i,0,£ 3 )Vmom4
(9.111c)
reduces Eq. (9.109) for the maxima of?? to (2^)"3/2 /?(»7, 0, u) = exp / __
(TJ2 + 2 V l -e 2 ?7j< + u2) 2^2
(9.112)
The mean frequency of occurrence of the total maxima of r\ is computed by oo
Ni
/
rv
I
\u\p(rj,0,u)dr]du.
(9.113)
-00 J—00
The mean frequency of occurrence that a maximum of rj lies in the small interval £1 < x < £1 + d£i is P(ji)dr) = V"U
/" I" \p(r),Q,u)dr]du
(9.114)
J—00
and the pdf for the maxima of r? is P(r]) =
(9.115)
783
Real Ocean Waves
The total mean frequency N\ may be easily evaluated from Eq. (9.113) with the following change of variables: OO
/-U
/ /
up(n,0,u)di]du
--00 0 0 «/—00
"00
/>0
(rj1 + 2Vl -e2?7M + w2)
uexp
= -^*wLL
dr\du.
2e2
(9.116) Evaluating Eq. (9.116) for A/i is simple if the square of the argument of the exponential function in the integrand is completed according to rj2 + 2^1 - €2r]u + u2 = (rj + y/\ - e2u\
+ eV
(9.117)
so that (?72 + 2Vl -€2t]u + u2) exp
(t] + exp
2?
(-T)
exp
Vl-e2u)2' 2ei
Integrating Eq. (9.116) first with respect to drj with the following change of variables: 2 n+ VT^l u , drj — edq
1
gives
f
(f?Wl-62K)2
exp
drj
2e2
=e
exp(-Yjdq
= eV2n, (9.118)
and integration now with respect to du yields Ni
u exp (
2n V/n4/m2 lit
2n 2n
J du
f° I exp(—x)dx J+oo
I Jo
exp(—x)dx (9.119)
784
Waves and Wave Forces on Coastal and Ocean Structures
Similarly, the mean frequency of occurrence P(r]) from Eq. (9.114) is P(tl) =
*Jm4/m2 f
(u2 + 2Vl -e2r)u + t]2) du, (9.120) 2e2
exp
that may, again, be easily evaluated by completing the square of the argument of the exponential function in the integrand in Eq. (9.120) according to u2 + 2\/\ -e2r]u + r)2 = (u + Vi - €2r))2 + e V
(9.121)
so that exp
(w2 + 2Vl -e2r)u + r}2) 2e2
exp ( - ^ - ) exp ( - ^ ) ,
(9.122)
where, with the following change of variables: q — u + V1 — e2r],
dq = du,
Qo = r\y 1 — e 2 ,
Eq. (9.120) becomes
Go
- /
1^
(9.123)
With the following change of variables for each of the two integrals I\ and h in Eq. (9.123):
h: h •
?V2
= y> dq — e^/ldy,
q r - j = f>
2
785
Real Ocean Waves
integration gives Pin) 0
/-Go/evT
-W-{-h)-*x=*\L+L /•oo
x e x p ( - / ) d y + el I
e{2n)V2
exp
2
(4)j f ,y|yr^[ 1+Erf ( 2
exp(-t)dt
ejl
/ *7
+e^ exp - '
(2jr)3/2
: exp
("&*) + ly^J^w
1+Erf
(-y)
eV2 (9.124)
The pdf for the maxima of?? from Eq. (9.115) with JVI from Eq. (9.119) is P07) Pin) e
Jin
eXP
("^)
+
2^
T T
^
e X P
(~^
1+Erf (9.125)
The parametric dependency of the pdf for the maxima of r\ on the spectral bandwidth parameter e is shown in Fig. 9.21. Note that for strictly narrowbanded spectra when e = 0 and Erf (oo) = 1, then Eq. (9.125) reduces to the
Waves and Wave Forces on Coastal and Ocean Structures
786
0.8 0.7 0.6 ^0.5 ^0.4 ^0.3 0.2 0.1 0.
Fig. 9.21. Parametric dependency of the pdf for the maxima of r/ on the spectral bandwidth e.
Rayleigh pdf Eq. (9.86a) with a = 1; while for strictly broad-banded spectra when e — 1 and Erf(0) = 0 then Eq. (9.125) reduces to the Gaussian pdf Eq. (9.80b). The interpretation of these two limiting values of e are that for strictly narrow-banded spectra (e = 0) the distribution of the dimensionless maxima rj is identical to the distribution of the dimensionless amplitudes §; and that for strictly broad-banded spectra (e —I) there are as many negative maxima as there are positive maxima and the pdf is a Gaussian (Normal) distribution. The pdf Eq. (9.125) for p(t]) may now be applied to determine the distribution of various average maximum values r)\/n in the same way that average wave amplitudes A\/n (or average wave heights H\/n) are determined from Eq. (9.90) for the Rayleigh distribution. The percent or fraction 1 /n of maxima rji/n (= ^i/^/mo) may be computed from Eq. (9.125) according to 1 f°° - = / P(v)dr] n
Jm,n
787
Real Ocean Waves
For strictly narrow-banded spectra e = 0, Erf (oo) = 1 and Eq. (9.126) reduces to n = exp giving, again, Eq. (9.89a) with a = y/ino = 1.0andAi/„ = r\\jn. The average rji/„ of those maxima greater than r)\/n may be determined from
min (°° pitfdr, = ^
= r
vpWdn.
(9.127)
The values of the lower integration limit r)\/n in Eqs. (9.126 and 9.127) may be computed numerically from Eq. (9.126) as a function of the spectral band width parameter e. Table 9.7 tabulates selected values of the lower integration limit r]\/n for n = 2,3,5 and 10 for ten values of e that range from 0.0 to 1.0 by 0.1. Note that for e = 0, the limits rji/n in Table 9.7 are identical to the amplitudes A\/n for the Rayleigh pdf Eq. (9.89b) when the Rayleigh parameter a = ^/mo. A program from MATHEMATICA™ that computes the values of the lower integration limits rj\/n is also listed. Cartwright and Longuet-Higgins (1956) integrated Eq. (9.127) numerically for a\,a\/2,a\fi,ai/5, and ai/io normalized by ^/mo as a function of the spectral bandwidth parameter e (vide., Sarpkaya and Isaacson, 1981 p. 499, Fig. 7.10). Table 9.7. Numerical values for the lower integration limits fll/ne
m/2
ni/i
m/s
11/10
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
1.17741 1.17313 1.15994 1.13666 1.10097 1.0494 0.97757 0.879346 0.743798 0.543538 0.0
1.48230 1.47891 1.46847 1.45014 1.42228 1.38203 1.32495 1.24444 1.12897 0.95067 0.430727
1.79412 1.79128 1.78271 1.76764 1.74486 1.71209 1.66546 1.59847 1.49938 1.34005 0.841621
2.54547 2.14362 2.13643 2.12388 2.10495 2.07786 2.03939 1.98366 1.89923 1.75804 1.28155
788
Waves and Wave Forces on Coastal and Ocean Structures
The algorithm from the software package MATHEMATICA™ is not optimized and is dated. It is intended only to illustrate one of many possible algorithms that may be applied to determine the lower limits of integration. Some of the commands illustrated in the program may need to be changed for later versions of MATHEMATICA™. (* MATHEMATICA PROGRAM TO COMPUTE LOWER LIMITS FOR MAXIMA INTEGRALS *) (* SET THE PATH TO THE DIRECTORY FOR YOUR OUTPUT FILES *) SetDirectory["path"]; today=Date[]; TagReal[x_]:=(x/: Re[x]=x;x/: lm[x]=0;) TagReal[np];TagReal[e];TagReal[x]; 11 =e*Exp[-(x/(e* Sqrt[2]))*2]/Sqrt[2*Pi] ;I2=Sqrt[ 1 -e"2] *x/2*Exp[-x~2/2]; I3=I2*Erf[Sqrt[l-e"2]/e*x/Sqrt[2]]; Int 1 ind=Integrate [11 ,x] ;Int 1 inf=Int 1 ind/.x- > Infinity ;Int 1 np=Int 1 ind/.x- > np; Int 1 =Simplify [Int 1 inf-Int 1 np]; Int2np=Integrate[I2,{x,np,Infinity}]; Int2=Int2np[[2]]; Int3ind=Integrate[I3,x];Int3inf=Int3ind/.x->Infinity;Int3np=Int3ind/.x->np; Int3=Simplify [Int3inf-Int3np] ;probnp=Simplify [Int 1 +Int2+Int3 ]; nprobnp=N[Simplify[Intl+Int2+Int3]];epnp=Simplify[Table[nprobnp, {e,. 1,1,. 1}]]; xp2:=Table[npp=np/.FindRoot[epnp[[i]]==l/2,{np,.01}],{i,l,10}]; OutputForm[u=TableForm[xp2,TableSpacing- > {0} ] ] > > xp2 xp3:=Table[npp=np/.FindRoot[epnp[[i]]==l/3,{np,.01}],{i,l,10}]; OutputForm[u=TableForm[xp3,TableSpacing-> {0}]]> >xp3 xp5 :=Table[npp=np/.FindRoot[epnp[[i]]==l/5, {np,.01}], {i, 1,10} ]; OutputForm[u=TableForm[xp5,TableSpacing-> {0} ]] > >xp5 xpl0:=Table[npp=np/.FindRoot[epnp[[i]]==l/10,{np,.01}],{i,l,10}]; OutputForm[u=TableForm[xp 10,TableSpacing- >{0}]]>>xpl0 It is possible to partially integrate Eq. (9.127) by parts to obtain € / 2 m/n = n i + v T^e
vr^
+ •
( m/n W/»
exp
ln(l-€2)
l-Erf(^
+
VV2
m/n
V2
1+Erf
Cexp-(^)Erf (9.128)
789
Real Ocean Waves
1 exp(-z2) ;
:W
0.5 0
;/
II M/
-0.5
Erffe)'
III/
-1
-5-4-3-2-1012345
z
Fig. 9.22. Integrands in the integral in Eq. (9.128).
The integrand of the only integral term left in Eq. (9.128) is a product of a symmetric function {exp — z2} and an anti-symmetric function {Erf (z)} that are illustrated in Fig. 9.22 and that may be easily integrated numerically . As a check that Eq. (9.128) gives the same result as Eqs. (9.91) for the Rayleigh distribution, substitute e = 0 in Eq. (9.128) so that Erf [1/0] = Erf[oo]=l giving ^ l l - E r f hl/"~h + ^ e x p - ( ^ l m/n = n
oo
+
^
(l+Erf[oo])
- • -7= ) Vrf [oo]dr}
c J
= n \ rn/n exp •
^
V2
min
m/n
V2
V
+.
%{
1-Erf
m/n
LV2
(9.129)
Substituting Eq. (9.89b) with a = y/ino given by M/n /-. , . frll/n\ = - p = = m/n = V21n(n), n = exp —, a Vfflo V V2 / reduces Eq. (9.129) to Eq. (9.91b). Two functions of fj\/n are easily determined from Eq. (9.128); viz., fj\ and ijrms (Cartwright and Longuet-Higgins, 1956, p.219, Eq. (4.8) or Sarpkaya A, U/n
Waves and Wave Forces on Coastal and Ocean Structures
790 3 2.5
2
1
0.5 0 0
0.2
0.4
0.6
0.8
1
€ Fig. 9.23. Parametric dependency of the average dimensionless maxima on n.
and Isaacson, 1981). For n = 1, the lower integration limit r)\/n -+ - o o and
The rms value ofthe maxima fjrms may be computed with r) from Eq. (9.111 a) by oo i? 2 p(i7)rfi?,
/ -oo
jj rBW = y/2 - e 2 .
(9.131)
The dimensionless average maxima fji/n for values of n = 1, 2, 3, 5, 10 and rms computed from Eq. (9.128) are illustrated in Fig. 9.23. Cartwright and Longuet-Higgins (1956) identified two significant features from Fig. 9.23. First, the dimensionless average maxima r\\jn are relatively insensitive to the spectral bandwidth parameter e for values 0 < e < 0.5. The significance of this lack of dependency of f\\/n on e in this interval is that the assumption that the maxima are Rayleigh distributed may be extended to spectra that may not be strictly narrow-banded; i.e., 0 < e < 0.5. Second, the slopes of all of the dimensionless maxima (except for fjrms) are equal to - o o and independent of n as c —> 1.0.
9.5. Wave Groups Wave groups provide a tool for analyzing random waves that requires less data than the ensemble average of all of the data x\(ti),X2(ti),X3(ti),... ,x„(t\)
Real Ocean Waves
791
Fig. 9.24. Families of realizations Jjj^(f) with the same wave group A,(r) and one-sided spectrum Sv(f) (Fassardi, 1993).
(vide., Fig. 9.1 in Sec. 9.1) but more data than the single one-sided spectral density function 5^(/) as illustrated in Fig. 9.24. Many realizations rnf(t) from an ensemble may have the same wave group characteristics that may be analyzed by wave group parameters that are measures of this groupiness. Each wave group shown in the center column in Fig. 9.24 consists of a family of realizations r]if(t) and is much smaller than the ensemble of all of the finite length realizations shown in the left column but, at the same time, is much more robust than the single one-sided wave density spectrum 5^(/) shown in the right column. Several wave group parameters have been proposed that may be applied to evaluate the effects of wave groups for engineering design. Medina and Hudspeth (1990) have shown that many of the wave group parameters may be correlated and interrelated. A wave group parameter that has shown some promise for evaluating damage to armor breakwaters is the envelope exceedance coefficient a(t) combined with the run length or spectral shape y applied by Medina, et al, (1994).
792
Waves and Wave Forces on Coastal and Ocean Structures
Rye (1982) reviewed many different wave groups parameters and methodologies and concluded that wave groups measured from field data compared quite well with those numerically simulated with linear algorithms. The validity of the linear model was also verified by Goda (1983), Elgar et al. (1984, 1985) and Battjes and Vledder (1984) in relatively deep-water. In his classic treatise on random noise, Rice (1954) developed an extensive theory of noise that may also be applied to linear surface gravity waves (vide., Bracewell, 1986; Bendat and Piersol, 1986; Cartwright and Longuet-Higgins, 1956; and Dugundji, 1958). The envelope A{t) of the sea surface elevation rj(t) appears to be an appropriate tool for analyzing wave groups. Medina and Hudspeth (1987) and Hudspeth and Medina (1988) applied the envelope A(t) of the sea surface elevation t](t) to analyze wave groups in random seas. A stationary, ergodic and Gaussian random sea surface elevation at a fixed location having a one-sided variance spectral density function 5^(/) may be approximated by M
7/(/) = J^
A
m cos(2nfmt + 0m),
(9.132)
m=\
where M = total number of wave components in the realization and A m , fm, and 9m = the amplitude, the discrete frequency and random phase angle, respectively, of the mth wave component. The random phase angle 6m is uniformly distributed in the interval f/[0, In] by Eq. (9.85d) in Sec. 9.4.1. The amplitude Am of each mth wave component and the one-sided variance spectral density Sj,(f) are related by (Tuah and Hudspeth, 1982). A2m = CmSr,(fm)Afa = -2\n(Um)Sr,(fm)Afa,
m = 1, 2, 3, . . . , M, (9.133a,b)
where Cm = a chi-squared random variable with two-degrees of freedom; Um = a random variable that is uniformly distributed in f/[0,1]; and Afa = discrete, constant frequency interval in the one-sided variance spectrum Sn(f). The Hilbert Transform H[f (§)] of the function £(£) is formally defined by Bendat and Piersol (1986) and by Bracewell (1986) as »[?(£)] = - /
^ r -
(9-134)
793
Real Ocean Waves
For sinusoidal time series, Eq. (9.134) may be interpreted as a TT/2 phase shift of £(§). For the random wave time series given by Eq. (9.132), the Hilbert transform fj(t) is M
rj(t) = ^
(9.135a)
Am sin(27tfmt + 9m)
m=l M
= ] P Am cos {lnfmt
+ 0M - | ) .
(9.135b)
m=l
An analytic function z(t) is a complex-valued function defined by (9.136a)
z{t) = r]{t) + iri(t), where i = ^T or in polar form as z(t) = A(t)exp[i(@(t) +
(9.136b)
where the envelope function A(t) is defined by (9.136c)
A(t) = ±y/riHt) + f)Ht) and the instantaneous phase function is computed from [n(t)' & (t) +
.
(9.136d)
In Eq. (9.136c) the positive values (+) of A(t) are the loci of the positive maxima of the time series r](t) and the negative values (—) of A(t) are the loci of the negative maxima of the time series rj(t). The analytic function z(t) is illustrated in Fig. 9.25 where r](t) and the Hilbert transform fj(t) are the vertical and horizontal displacements, respectively, of a point in the wave free surface. The following instantaneous time-dependent functions of the Hilbert transform fj(t) may now be defined: (i) the wave height function H(t) as H(t) = 2A(t),
(9.137a)
(ii) the frequency £2 (?) as Q (t) =
1 d , {arctan
lit dt '
nit)
(9.137b)
Waves and Wave Forces on Coastal and Ocean Structures
794
vio VM
Fig. 9.25. Analytic function of r)(t) and the Hilbert transform Hilbert transform fj (t) (Bracewell, 1986).
50
60
70
80
90 100
t (sec) Fig. 9.26. Wave surface r](t), Hilbert transform rj(t) of 77(f) and the envelope function A(t) from a realization from Hurricane CARLA85 measured in the Gulf of Mexico.
and (iii) orbital velocity V(t) as V(t) = 2jtA(t)Q(t)
=
nH(t)Q(t).
(9.137c)
The instantaneous functions defined by Eqs. (9.137) are constants for strictly periodic waves. The envelope A(t) for a random wave realization measured during Hurricane Carla in the Gulf of Mexico from a relatively broad-banded wave spectrum is illustrated in Fig. 9.26 along with the wave realization rj(t) and the Hilbert transform fj(t) of rj(t).
795
Real Ocean Waves
The wave celerity C and group velocity CG have been formally defined only for monochromatic wave by Eqs. (4.60) in Chapter 4.5. However, useful approximations for a mean celerity C and a mean group velocity CG may also be obtained for random waves. The spatial and temporal dependency of the sea surface elevation may be approximated by M V(x,t)
= J2AmCos[2Tt(Amx
- fmt) + em],
(9.138)
m=\
where Am = the inverse of the wavelength Xm corresponding to the mth wave component computed from the linear dispersion relationship fl = ^-Am
tanh(27r A m /i),
(9.139)
lit
where h = water depth and g = acceleration due to gravity. The Hilbert Transform fj(x, t) of rj(x, t) is defined as M
fj{x,t) = J^
A
m sin[2;r(Amx - fmt) + 0m].
(9.140)
m=\
The envelope A(x,t) and wave height H(x,t) functions are still defined by Eqs. (9.136c and 9.137a), respectively, butnow with two independent variables x and t. Families of phase-shifted realizations rnf(x, t) shown in Fig. 9.24 that have the same envelope function Ai(x,t) and the same flux of energy may be expressed by M
jtyCM) = J^ Am cos[27r(Amx - fmt) + (<9m - ^ ) ]
(9.141a)
m=\
= r](x,t) cos ^r + fj(x,t) sin i/f,
(9.141b)
where i^ = constant phase shift given to each m wave component. The one-sided variance spectral density 5r,(A) in the inverse of the wavelength space domain is related to the one-sided variance spectral density 5^ ( / ) in the time domain by the frequency dispersion relationship Eq. (9.139) and by the equality of differential spectral energies in each domain according to 'Jf)(.A m )dA m — £>r)\Jm)djm-
(9.142)
796
Waves and Wave Forces on Coastal and Ocean Structures
Envelope spectral density functions with unit variance in the frequency r^ (/) and in the inverse wavelength Tn{K) domains, respectively, are defined by the following spectral correlation integrals (vide., Hudspeth and Medina, 1988): Tv{f)
= \ !
Sr,(y +
r„(A)
= \ \ SJZ + Y) Sn{z)dz. mQ Jo V A/
rriQ Jo
f)Sn{y)dy,
(9.143a) (9.143b)
The Tn{f) and r^(A) defined by Eqs. (9.143) are not related in the same way as the variance spectra of the sea surface elevation are by Eq. (9.16a) in Sec. 9.3. The spectra of H(t) and H2(t) are approximately SH(f)
« (8 - 27r)m 0 r,(/),
SHi(f)
« 64mgr„(/).
(9.143c,d)
Estimates for a mean celerity C and a mean group velocity CQ m ay be computed from k
(9.144a,b)
AG
where r /max
r ^-max
/ / =
fS,(f)df
/
** /min
y
/*/max
/
Sn(f)df
/ /*AX
fTn{f)df
/ Jo
i
/•A/
/ Jo
Sr,fr)dk
** ^min
/•A/
fc =
(9.145a,b)
/*^max
** /min
/ Jo
kSn(k)dk
J ^-min
'
u
Yr,(f)df
kTn(k)dk (9.146a,b)
pAX
/ Jo
Tn{k)dk
/min;
(9.147a)
AA = A m a x — A m j n ,
(9.147b)
A / = /max
where /max(min) = maximum (minimum) cut-off frequencies; Amax(min) = maximum (minimum) cut-off wavelengths for the inverse wavelengths related to /max(min) by Eq. (9.139); / = mean frequency; A =mean inverse
797
Real Ocean Waves
Fig. 9.27. Spectral and envelope spectral representations in the frequency and inverse wave number space domains.
wavelength; / c = m e a n wave group frequency; and AG=mean wave group inverse wavelength. Figure 9.27 illustrates the functions defined by Eqs. (9.143-9.145) for a Goda-JONSWAP variance-preserving spectrum Eq. (9.48) in Sec. 9.3.3 (Goda, 1985). The Hilbert transform method for constructing envelopes is an exact low pass filter (Bendat and Piersol, 1986) and may be compared with other low pass filters that operate on the square of the wave surface profile such as the Smoothed instantaneous Wave Energy //istory (SIWEH) introduced by Funke and Mansard (1979) or the Local Variance Time Series (LVTS) introduced by Thompson and Seelig (1984). Hudspeth and Medina (1988) have compared the distortions in wave height envelopes that have been computed from the Hilbert transform methods in both time and frequency domains with the SIWEH. The results of their comparisons for two Goda-JONSWAP spectra that are broad-(y = 1) and narrow-(y = 10) banded are illustrated in Fig. 9.28. This comparison suggests that only the exact low pass filter computed by the Hilbert transform should be applied to evaluate wave groups by their envelopes. The definition for the energy flux per unit horizontal surface area (or, equivalently wave power (P), in Eq. (4.59) in Chapter 4.5) for monochromatic waves is H2
(9.148)
Waves and Wave Forces on Coastal and Ocean Structures
Fig. 9.28. Comparisons of envelope functions for a broad-banded (y = 1) and for a narrowbanded (y = 10) Goda-JONSWAP spectrum (Hudspeth and Medina, 1988).
where p = fluid mass density, g = gravitational constant, Co = group velocity defined by Eqs. (4.60) in Chapter 4.5, and H = monochromatic wave height. The monochromatic wave definition in Eq. (9.148) may be extended to random waves in order to estimate the instantaneous energy flux per unit surface area (P(*,f))/as H2(x,t) (9.149) = pg-CG. 8 Medina et al, (1994) have applied the extension in Eq. (9.149) to random waves in order to evaluate the effects of wave groups for engineering design. (?(x,t))t
Real Ocean Waves
799
lit)
c)
^2J=1 ,ilh(3te3 fo(2M 2Lh(3E>3
Fig. 9.29. Representation of (a) wave record; (b) sequence of wave heights; and (c) run lengths (Medina and Hudspeth, 1990).
Efforts to correlate wave groups with parameters that are related only to spectral shapes have met with varying degrees of satisfaction. Medina and Hudspeth (1990) review some of the wave group parameters and found that they could be interrelated. One of the parameters is related to run lengths defined by the number of consecutive wave heights that exceed a specified minimum threshold level h as illustrated in Fig. 9.29 where i L h (w) is the mth length of a run of wave heights greater than the minimum threshold h and 2Lh(m) is mth length of a total run of wave heights that exceed the minimum threshold h. In Fig. 9.29, the i L h (m) is defined by the interval between the first upcrossing of the threshold h denoted by the symbol ® and the next succeeding downcrossing of the threshold h denoted by the symbol*. The length of a total run of wave heights 2Lh (m) is defined by the interval between two consecutive upcrossings of the threshold h denoted by the symbols ®. Some of the parameters that are related to spectral shapes and correlation coefficients are interrelated by Medina and Hudspeth (1990) as shown in Fig. 9.30. One of the possibilities for the inconsistent correlation of wave groups with wave group parameters that depend solely on spectral shape was identified by Mase and Iwagaki (1986) as being due to the variability of realizations from spectral shapes as illustrated in Fig. 9.31. They identify two independent parameters in Fig. 9.31 that effect wave groups; viz., 1) the mean run length that is determined by the spectral shape y and 2) an envelope exceedance
800
Waves and Wave Forces on Coastal and Ocean Structures
} r HH (l)
0 4 . SWELL (GODA.1983) + ++ KIMURA (GODA.I983)
Fig. 9.30. Comparison of 5 wave group parameters for 2 cut-off frequencies (Medina and Hudspeth, 1990).
High SM
Wave Height Variability
S.0)
^k / H
~f H
Low ' '
Short
Long
Fig. 9.31. Wave height variability and run length for different spectral shapes y.
coefficient a (t) that is a measure of the wave envelope variability. The possibility that two very different spectral shapes (y = 1 and 10, say) or, equivalently, different run lengths may have the same wave height variability a(t); or the possibility that the same spectral shape {y = 1 or 10, say) may produce two very different values of wave height variability {ait) = 0.5 and 2.0, say)
Real Ocean Waves
801
LENGTH OF RUNS
WAVE HEIGHT VARIABILITY
Fig. 9.32. Wave group definitions and parameters (Fassardi, 1993).
are illustrated in Fig. 9.31. The wave group definitions and parameters are illustrated in Fig. 9.32. Medina et al, (1994) tested the envelope exceedance hypothesis on the damage to breakwater armor units from wave groups. In their experiments at the O. H. Hinsdale-Wave Research Laboratory (OHH-WRL) a single 3.7m wide breakwater section in a 2D wave channel was divided equally into two 1.85 m wide sections with the two different armor rock weights and sizes listed in Table 9.8. The wave group and spectral parameters tested are listed in Table 9.9. In these experiments, an average envelope exceedance coefficient a was defined as '
1
N
n=\
H(xo,nAt) H*
-1,
(9.150a,b,c)
802
Waves and Wave Forces on Coastal and Ocean Structures
Table 9.8. Armor rock characteristics for the O.H. Hinsdale-Wave Research Laboratory (OHH-WRL) experiments. Weight W (N)
Diameter Dn (m)
128.5 99.1
0.168 0.154
Table 9.9. Wave group parameters for the O.H. Hinsdale-Wave Research Laboratory (OHH-WRL) experiments. Envelope
Y
a
El E2 E3 E4
10 1 10 1
1.8 1.6 0.5 0.5
where E(a) = expected value of a; £/(•) =Heaviside step function defined by Eq. (2.1) in Chapter 2.2.2; N = total number of discrete wave height exceedance values; H(XQ, nAt) =wave height function measured at XQ in the 2D wave channel; and H* = characteristic design wave height, H\Q, say. Figures (9.33a, c) confirm that the two odd-numbered envelopes El and E3 from the relatively narrow-banded spectrum y = 10 produced different damage results for both armor rock weights and sizes with the envelope having the relatively higher wave height variability coefficient (El and a = 1.8) consistently producing the higher damage. Figures (9.33b, d) illustrate the same results for the relatively broader-banded spectrum y = 1. The damage data presented in the Shore Protection Manual (SPM) (1984) was fitted with the following empirical formula: r
,• n 0.2
#10
HND
1.6
(9.151)
where Dlk = damage from the kth. wave run to the j'th rock size; and HND — nodamage design wave height. These limited experiments on breakwater damage due to wave groups suggest that both the run length (or spectral shape y) and the
803
Real Ocean Waves
0.8 0.9
5
11
1.2
1.3
1.4
1,5
1.6
1.7
1.8
1.9
2
Hw/H^
Fig. 9.33. Comparisons of armor damage to breakwaters by wave groups (Medina, et al, 1994).
wave height variability a should be considered when evaluating the parametric dependency of wave groups on breakwater amor damage.
9.5.1. Resolving Incident and Reflected Random Wave Time Series In order to compute wave groups, wave heights or envelope exceedance coefficients, the incident wave time series must be known. Because of wave reflections, a method to resolve the incident and reflected random wave time series from wave gauges is required. A Finite Fourier Transform (FFT) algorithm computes the complex-valued wave amplitudes from which the incident and reflected random wave time series may be resolved. Goda and Suzuki (1976) and Goda (1985) modified an algorithm developed by Thornton and Calhoun (1972) to resolve the incident and reflected wave spectra from wave gauges separated spatially by a distance A/ (vide., Fig. 9.34). Kimura (1985)
Waves and Wave Forces on Coastal and Ocean Structures
804
Wave Gauges
ox, ox 2 /////////////////////y Fig. 9.34. Definition sketch for two wave gauges.
extended the Goda and Suzuki amplitude only algorithm to include wave phases in order that both incident and reflected time series could be resolved. Goda and Suzuki (1976) and Goda (1985) analyzed simultaneously by an FFT algorithm two wave records that were recorded by two closely spaced wave gauges that were aligned in the direction of wave propagation shown in Fig. 9.34. If the incident mr\j and reflected m £/ wave time series at the y'th wave gauge location at frequency 2jzfm are given by my]j
= am cos[27T(fmt - AmXj) ± sm\,
mtj
= bm COS[27T(fmt + AmXj)
±
Pm],
(9.152a) (9.152b)
then the composite wave profile at the y'th wave gauge may be expressed as mVj +m$j = mAj cos(2nfmt) + mBj sm(2nfmt),
(9.153)
where ; A\
= am cos * m + bm cos m,
^B\ = am sin * m - bm sin 3>m,
(9.154a) (9.154b)
mA2
= am cos(27rA m Al + * m ) + bm cos(2^A m A£ + $ m ) ,
(9.154c)
mB2
= am sin(27r Am A£ + * m ) - bm sin(2^A m A£ + <*>„),
(9.154d)
At = %2 — X\,
(9.154e)
805
Real Ocean Waves
where the phase angles of the incident * m and reflected <&m waves, respectively, are defined as * m = 2n Amxi =F £m
(9.155a)
$ m = 2nhmx\
(9.155b)
± pm
and xi and X2 are the spatial wave gauge positions shown in Fig. 9.34. The coefficients mAj and m Bj may be related to the complex-valued FFT coefficients fl/(m)inEq.(9.11c)by Bj(m) =
IffJ tj i
(9.156)
where? = V ^ . G o d a and Suzuki (1976) and Goda( 1985) solved Eqs. (9.154) sequentially for estimates of the amplitudes am and bm only. Kimura (1985) extended the Goda and Suzuki algorithm to include a different reflection coefficient for each incident wave component in the complex-valued FFT spectrum. Kimura defined incident and reflected wave profiles by Eqs. (9.152) and the spatial phases by Eqs. (9.155). The solutions for the amplitudes and phases from Eqs. (9.154) are (mM — mM cos27rAmA£ — mB\ sm.2nhmAl)2 + (m#2 + mM sin27rAmA£ — mB\ cos2^A m A£) 2 *2m —
2| sin27rA m A£| -m^-2 + mAi cos2;rA m A£ + mB\ sin2^A m A£ mB2 + mM sin27rA m A£ tB\ COS2TT Am/S.£
* m = arctan
(9.157a)
(9.157b)
bm =
(mM — mM cos2:rA m A£ + mB\ sin27rA m A£) 2 + (m^2 — mA\ sin27rAmA£ — mB\ cos2^A m Al) 2
2|sin27rA m A£|
(9.157c)
1A2 — mM cos2:7rAmA£ + mB\ sm2nAmAl. _mB2 — mM sin27rAmA£ tB\ cos27rAmA£
= 27rAmxi
±Bm.
(9.157d)
806
Waves and Wave Forces on Coastal and Ocean Structures
For random waves, the time series are the superposition of many wave components given by Eqs. (9.152). Because Eqs. (9.154) apply to each mth Fourier component in a random wave time series, linear superposition may be applied to resolve the time series from a spectrum of complex-valued FFT components (Goda, 1985). Goda and Suzuki (1976) observed that the spectral estimates become unbounded at frequencies where the solutions for the linear wavelength Am from the linear dispersion equation (9.139) are 2nAm Al « mix for n = 1, 2, . . . , because the term | sin 27r Am At | in the denominator of the equations for the amplitudes am and bm in Eqs. (9.157a, c) becomes small and errors from noise are amplified. Consequently, the wave gauge spacing A€ determines the upper and lower frequency limits of a band pass filter from which the wave components may be separated into incident and reflected time series. They suggest that the wave amplitudes am and bm may be resolved effectively for Fourier component frequencies in the interval 0.17T < 2nAmA£ < 0.9n. Goda and Suzuki (1976) recommend the following effective band pass limits for experimental conditions: 0.03A.max < At < 0.45A.min,
(9.158a)
where A.max and Am;n denote the wavelengths computed from the linear dispersion Eq. (9.139) that correspond to the lower / m j n and upper / m a x frequency limits, respectively, of the band pass filter. Although Kobayashi, et al. (1990, p.723) identify and discuss several reasons for the low coherence at both high and low frequencies in their algorithm for resolving time series from random waves, the reason for the low coherences at both low and high frequencies in their data is that there are no data at these frequencies because their algorithm is also a band pass filter. Goda and Suzuki (1976) also recommend that the wave gauges be located at least one wavelength Xmax away from both the coastal structure and the wave generator in laboratory experiments. In order to improve the resolution of incident and reflected waves, Kimura (1985) recommends a slightly more conservative band pass frequency interval given by 0.15A.max < Al < 0.35A.min.
(9.158b)
In the rubble mound breakwater experiment of Medina, et al. (1994) conducted at the O. H. Hinsdale Wave Research Laboratory, the algorithm in Eqs. (9.154 and 9.157) was applied to resolve the incident and reflected
Real Ocean Waves
807
time series. Three sonic wave gauges were aligned in the direction of wave propagation and were centered in the middle of the wave channel 10 m from the toe of the breakwater and separated spatially by At — 1.22 m as illustrated in Fig. 9.35. A short temporal sample of the time series from each of these three wave gauges is shown in Fig. 9.36. The time series for the incident and the reflected waves for the run E1P1L7 were resolved from sonic wave gauges 1 and 2 as shown in Fig. 9.35 by Eq. (9.153) and are shown in Fig. 9.37.
A0 = 1.22m At At Sonic Wave Gauges
n
k >j< >| I I I O O O
— • Rubble Mound Breakwater
h=3.05m
1°2°3° 10m
Fig. 9.35. Sonic wave gauge locations for rubble mound experiments at the O. H. HinsdaleWave Research Laboratory.
Time Series - Run E1P1L7
35
37 Time (s) - G A U G E 2 - - - GAUGE3
Fig. 9.36. Time series from the three wave-gauges shown in Fig. 9.35.
Waves and Wave Forces on Coastal and Ocean Structures
808 T i m e
S e r i e s
- R u n
E l I * ! IJ'7
f^fefe/w <so T i m e Cs> T i m e
S e r i e s
- R u n
E 1 P 1 L 7
T i m e s <s> I n c i d e n t
T i m e
S e r i e s - R u n
E 1 P 1 L 7
60 T i m e <s> R e f l e c t e d
—r
T i m e
S e r i e s - R u n
E 1 P 1 L 7
0.4 - •o-»
"*J\p\f\J\f\jK^"-J^r^.rvr\f^
•• ••^^"u-lr^^v^^A^AA/V/\l^\
=S - 0 . 2 -
so T i m e C*0
Fig. 9.37. Time series resolved from sonic wave gauges 1 and 2 in Fig. 9.35.
9.6. Random Wave Simulations Digital simulations of random time sequences are required for both numerical analyses and the generation of laboratory waves by digital-to-analog converters (DAC) (Hudspeth and Borgman, 1979). Algorithms for digital
Real Ocean Waves
809
simulations are available for both time and frequency domain simulations (Borgman, 1969b). Only the frequency domain simulations are reviewed here (Hudspeth and Borgman, 1979). Rice (1954) derives two mathematical algorithms for representing electronic noise currents as stationary Gaussian processes. Both of these algorithms have proven to be so robust that their applications to many otherfieldsmay be found (vide., Cartwright and LonguetHiggins, 1956, Sec. 4; inter alios). The two noise algorithms of Rice (1954) are applied to simulate Gaussian white noise spectra that may then be filtered in the frequency domain by a theoretical wave amplitude spectrum from Table 9.5 in Sec. 9.3.2 or from a measured target wave amplitude spectrum in order to obtain a Gaussian random sea time sequence by Fourier inversion of an FFT amplitude spectrum. Consequently, this algorithm may be shown to be equivalent to filtering digitally Gaussian white noise (Tuah and Hudspeth, 1982). Accordingly, applications of the Rice algorithms to coastal and ocean engineering include estimating wave-induced hydrodynamic loads on small member structures (Borgman, 1969b or Grecco and Hudspeth, 1983); generating random waves digitally in wave flumes (Hudspeth and Borgman, 1979); and modeling both stationary (Bily and Bukoveczky, 1976) and non-stationary (Cacko and Bily, 1979a, b) coastal and ocean processes. The FFT algorithm defined in Sec. 9.2 may be applied to both conditional and unconditional simulations in the frequency domain. A random time sequence may be simulated unconditionally by applying either the Nondeterministic Spectral Amplitude model (NSA) or the Deterministic Spectral Amplitude model (DSA) (Tuah and Hudspeth, 1982). The unconditional simulation of a random time sequence with zero mean is most efficiently computed in the frequency domain by an FFT algorithm (Borgman, 1982 and Hudspeth, et al., 1999). The descriptions for NSA and DSA simulations that follow apply the FFT notation in Eqs. (9.5) from Sec. 9.2 with the frequency index in the interval 0< m < N — 1; and the one-sided spectral density function Sm(m) defined by Eq. (9.15i) for radian frequencies com and by Eq. (9.16d) for Hertzian frequencies fm in Sec. 9.3 at discrete frequencies m df. However, the normalizing constant Cjy that appears in all FFT algorithms and reviewed in Sec. 9.2 following Eq. (9.10) is not included in the algorithm derived below; and this normalizing constant Cyy must be determined uniquely as in Table 9.1 in Sec. 9.2 for each FFT algorithm.
810
Waves and Wave Forces on Coastal and Ocean Structures
Nondeterministic Spectral Amplitude simulation (NSA) The NSA simulation algorithm requires the following steps: (i) For each discrete frequency fm = mdf in the interval 1 < m < N/2 — 1, generate two independent Gaussian normal variables [am,bm] with zero mean and unit variance N[0, 1] by (9.159a)
N[0, 1].
These two Gaussian variables may be generated from uniform random numbers [U\(m) and U2(m)] in the interval U[0,1]; and the amplitudes am and bm in Eq. (9.159a) may be computed by (Zelen and Severo, 1968) am
cos(2nU2{m))
(9.159b)
sin(27r U2(m))
(9.159c)
= [-ln(t/i(m))]2 bm
The random deviates for Gaussian white noise am and bm may also be computed using an acceptance-rejection method (Zelen and Severo, 1968); or from intrinsic functions in most software. (ii) In the positive frequency interval 1 < m < N/2 — 1, initialize each complex-valued FFT coefficient Bu (m) for the white noise amplitude spectrum for the unconditional simulation by Bu(m) = {flm
-ibm)
Svv(m)df
exp —i(kmx),
1 < m < N/2 — 1. (9.160a)
In the negative frequency interval N/2 + 1 < m < N — 1, the complex-valued FFT coefficients are the complex conjugates B*(m) of Eq. (9.160a) and are computed by BU(N - m) = B*(m),
N/2 + 1 < m < N - 1
(9.160b)
and for the mean of the time sequence at m = 0 and for the Nyquist or folding frequency at m = N/2 Bu(m) = 0,
m = 0 and
m = N/2.
(9.160c)
811
Real Ocean Waves
Again, the normalizing constant Cjy that appears in all FFT algorithms as discussed in Sec. 9.2 following Eq. (9.10) is not included in this algorithm; and this normalizing constant Cjy must be determined uniquely as in Table 9.1 in Sec. 9.2 for each FFT algorithm. (iii) Synthesize by inverse FFT the sequence Bu(m) to obtain a discrete unconditional time sequence r]u(n) by N-l
,~
Bu(m) expj ( Lizmn -
*lu(n) = ^
*
m=0
n = 0,l,2,...,N-l.
(9.161)
N
Deterministic Spectral Amplitude simulation (DSA) The DSA simulation algorithm requires the following steps: (i) For each discrete frequency m df in the interval 1 < m < N/2 — 1, generate a single independent uniform random number U(m) that is uniformly distributed in the interval £/[0,1]. (ii) Obtain a sequence of random phase angles 0m for each frequency fm = mdf that are uniformly distributed in the interval U[0,2n] by multiplying each uniform random number U(m)by2n. (iii) In the positive frequency interval 1 < m < N/2 — 1, initialize each complex-valued FFT coefficients Bu{m) for the unconditional simulation according to Bu{m) = ^
Sr,V
2
f
exp -i(kmx
+ 6m),
1 < m < N/2 - 1. (9.162a)
In the negative frequency interval N/2 + \ <m < N — 1, the complex-valued FFT coefficients are the complex conjugates B*(m) of Eq. (9.162a) and are computed from BU(N - m) = B*(m),
N/2 + 1 < m < N - 1
(9.162b)
and for the mean of the time sequence at m = 0 and for the Nyquist or folding frequency at m = N/2 Bu(m) = 0,
m=0
and
m = N/2.
(9.162c)
812
Waves and Wave Forces on Coastal and Ocean Structures
White noise spectrum W(m)
W(m) 1.0
—
:°nii
X
m
X
Target spectrum Sm(m)
• •ttllllllllllLlllllllllltt>...
l...tlllll
TlM
iMl
L_ m
FFT spectrum \B„{m)\2
\Bu(m)['
JllllllUllllIlk,
**~~\ m
m
DSA METHOD
NSA METHOD
Fig. 9.38. Comparison of DSA and NSA random wave simulations in the frequency domain (Tuah and Hudspeth, 1982).
Again, the normalizing constant Cjy that appears in all FFT algorithms as discussed in Sec. 9.2 following Eq. (9.10) is not included in this algorithm; and this normalizing constant CJV must be determined uniquely as in Table 9.1 in Sec. 9.2 for each FFT algorithm. (iv) Synthesize by inverse FFT the sequence Bu{m) to obtain a discrete unconditional time sequence rju(n) by .
7V-1
»?«(«) = Yl m=0
B
u(m)expi I ^
2nmn N
,
n = 0,l,2,...,N-l.
(9.162d)
813
Real Ocean Waves
In the NSA simulation, the target wave amplitude spectral density function fluctuates about the target spectrum much like measured spectral estimates from real ocean waves. The randomness in the target energy spectrum in an NSA simulation is produced by the nondeterministic generation of both the amplitudes and the phases in Eq. (9.160a). In contrast, only the phase angles 6m are stochastically generated in the DSA simulation while the amplitudes are deterministic. Therefore, the target wave amplitude spectral density function in a DSA simulation is identical to the target spectrum. Both methods produce realizations having approximately equivalent Gaussian properties (Tuah and Hudspeth, 1982). In most FFT algorithms, the unconditional time sequence is contained in only the real part of the complex-valued FFT coefficients. For FFT algorithms applied in computers with limited CPU memory, the time sequence may not be as long as desired because of this limitation. In these cases, the length of the simulated time sequence may be doubled by using a stacked FFT algorithm (Hudspeth and Borgman, 1979). Figure 9.38 demonstrates that both the NSA and the DSA random wave simulations are equivalent to filtering Gaussian white noise in the frequency domain (Rice, 1954, p. 180 and Borgman, 1969b).
9.6.1. Conditional Wave
Simulations
A conditional simulation may be defined as a numerical method for embedding a deterministic sequence of waves or wave groups into a random wave simulation having a specified target variance spectrum. The conditional simulation may be performed either in the time domain or in the frequency domain. Both time and frequency domain conditional simulations of water waves are given by Hudspeth et al. (1999); but only the frequency domain conditional simulation is reviewed here. Hudspeth et al. (1999) compare numerical conditional simulations of water surface elevations computed in the frequency domain with measured laboratory waves generated by a planar hinged wavemaker in a 2D wave channel at the O. H. Hinsdale-Wave Research Laboratory at Oregon State University. Their numerical FFT algorithm is derived from an algorithm given by Borgman et al. (1993) for conditionally simulating many wave field properties. The numerical stability of conditional simulations depends on the condition of a covariance matrix that is a function of: 1) the ratio of the length
814
Waves and Wave Forces on Coastal and Ocean Structures
"Hco
•J \ I \ -%
4—+>ndt
U
(a)
TJC
o| /
\
Xs \\ \ \ \ \ I—\
\...A1/ 1/ 4-/
A / \ » ndt
\/
U
(b) Fig. 9.39. Schematic representation of: a) a deterministic embedded design wave; and b) a deterministic embedded design wave group in a random unconditional wave simulation (Hudspeth etal., 1999).
of the embedded wave sequence to the length of the unconditional sequence; 2) the compatibility of the variance of the embedded wave sequence with the variance of the target spectrum; and 3) the size of the discrete simulation time step dt. Numerical conditional simulation is an efficient method for generating a random time sequence with an embedded deterministic wave or wave group. An embedded deterministic wave of period Tp or wave group of length Tg (vide., Fig. 9.39) may begin at any arbitrary discrete time t\ = n\dt and may end at any arbitrary discrete time later tv = (n\ + v)dt, where the embedded time interval (v + l)dt = Tp(g) for a periodic design wave (or wave group). The two steps required to produce a conditional simulation are: 1) to generate an unconditional NSA or DSA random wave simulation from a target variance spectrum by the procedures outlined in Sec. 9.6 (Rice, 1954; Borgman, 1969b and 1972b; Tuah and Hudspeth, 1982, inter alios) and 2) to embed a deterministic time sequence into the unconditional simulation generated in step 1) above.
815
Real Ocean Waves
A conditional water surface time sequence r)c(ri) at discrete times tn = ndt may be expressed by
Vc(n) = J_/Bc(m)expil——\, m=0
^
n = 0, 1, 2,... ,N - 1, (9.163a) '
where the conditional FFT coefficient Bc{m) at discrete wave frequencies fm — m df at a location x is Bc{m) = (Ci(m) — /C2(m))exp— i(kmx),
m = 0,1,2,... ,N — 1. (9.163b)
The normalizing constant C^ that appears in all FFT algorithms as discussed in Sec. 9.2 following Eq. (9.10) is not included in Eq. (9.163b); and this normalizing constant CN must be determined uniquely as in Table 9.1 in Sec. 9.2 for each FFT algorithm. The discrete wave numbers are solutions to the linear frequency dispersion equation komh = kmhtanhkmh,
(9.163c)
where the deep-water wave number kom = (2nfm)2/gA conditional simulation with an embedded time sequence r\e (n) at discrete times tn = ndt in the interval n\ < n < n\ + v requires an unconditional simulation rju(n) and two covariance matrices Cn and CnAn unconditional simulation of the water surface r)u (n) may be synthesized from
riu(n)=J_jBu(m)exVi[-—-\, m=0
^
n = 0, 1, 2 ... ,N - 1, (9.164a) '
where Bu(m) = (Um - iVm) exp -i(kmx),
l<m
(9.164b)
Again, the normalizing constant CN that appears in all FFT algorithms as discussed in Sec. 9.2 following Eq. (9.10c) is not included in Eq. (9.164b); and this normalizing constant CN must be determined uniquely as in Table 9.1 in Sec. 9.2 for each FFT algorithm.
816
Waves and Wave Forces on Coastal and Ocean Structures
The coefficients Um and Vm in Eq. (9.164b) for an NSA simulation by Eq. (9.160a) are jSw(m)df Um — «m\;
„
_ >
\Sm(m)df
"m — Pm\
(9.165a,b)
where am and fim are independent Gaussian normal variables with zero mean and unit variance N[0,1] that may be computed by Eqs. (9.159) and where Sm{m) = a one-sided spectral density function at discrete frequency fm = mdf. The coefficients Um and Vm in Eq. (9.164b) for a DSA simulation by Eq. (9.162a) are Um =
Sm(m)df
cosOm,
Vm =
Sm{m)df
_. sm8m,
(9.165c,d)
where 9m = random phase angle that is uniformly distributed in the interval U[0, In]. The variances of Um and Vm are defined as
EK]=E[^] =
^
.
(9.165e,f)
One of the two required covariance matrices is a [(v + 1) x (v + 1)] partitionedToeplitzauto-covariance matrix Cn = [Cov(t>„, vu)] (Press etal., 1986). The column vector vu contains the unconditionally simulated time sequence t]u(n) in the discrete time interval n\ < n < n\ + v where the embedded time sequence t]e(n) is to be embedded; i.e., r)u(n\)
(9.166)
v„ = riu(n\ + v)
The elements in the covariance matrix Cn may be expressed in the frequency domain by an auto-covariance function defined by Eq. (9.18a) in Sec. 9.3 that may be computed from either a target spectral density function from either Table 9.5 or from the multiple parameter target spectra in Sec. 9.3, for example, or from a measured wave spectrum by FFT coefficients where each scalar element is the auto-covariance value for the discrete time lags
817
Real Ocean Waves
L — 0, 1, 2, . . . v. Each scalar element of the covariance matrix C n may be expressed in FFT notation as
(2nmi m=0 V
/
,
t = 0, 1, 2 , . . . , v .
(9.167)
V
Because ??M(n) is related structurally to Bu(m) by Eqs. (9.5) or by Eqs. (9.11), the covariance between r)u(n) and the real and imaginary components of the complex-valued amplitude FFT coefficients Bu(m) denoted as Um and Vm, respectively, may be computed directly by 2n mn n (JT <M (Sm{m)df\ Cov{Um,t]u(n)} = I ->-!— I cos N
ICmX
(9.168a)
and by Cov{V m , r)u(n)} = [ -u—
I sin
2nmn N
(9.168b)
ftmX
The second required covariance matrix Cu between the scalar elements of the column vector v„ and the FFT coefficients Um and Vm is given by C12 = [Cov{I/m,vn],Cov{Vm,vu}] (2nmn
- y
kmxj ,
N
(2-nm{n\ + 1)
C12 = ( « f )
cos \
N
\ — kmx I ,
2nm(n\ + v) - kmx I, N
(9.169a)
. (2itmn\ ^ sin —— kmx . / 27tm(n\ + 1) kmx\ sin ~~N
sin I
—
kmx 1 (9.169b)
and the transpose Cj2 is given by
1
c — *-12 —
J
(m)Af\
(
2nmn\ cos | —
nn
sin
N 2nmn\ N
\\ (2jtm(n\ + v) kmx ) , . . . , cos | ^
N kmX \ . (2Ttm{n\ + v) kmx ) , . . . , sin | N
^r~ ~ )
kmx kmx (9.170)
818
Waves and Wave Forces on Coastal and Ocean Structures
The amplitudes C\(m) and C2(m) of the conditional simulation FFT coefficients Bc(m) in Eq. (9.163b) may be computed from
>}=<£<-<*-„> + {£)
C\(rn) Cjim
(9.171a) (9.171b)
A conditional time sequence r]c(n) may be simulated by either the NSA or DSA models by the following steps: (i) Simulate an unconditional time sequence r]u(n) by either the NSA or DSA method that is outlined in Sec. 9.6 and determine the discrete time interval n\ < n < n\ + v where the embedded time sequence vector ve is to be embedded. The embedded time sequence r]e(n) is stored in the embedded time sequence column vector ve according to
ve =
(9.172a)
The unconditional time sequence rju(n) in the same time interval as Eq. (9.172a) is stored in the column vector v„ given by Eq. (9.166). (ii) Solve the following vector equation for the column vector X: C „ X = Ve
(9.172b)
Because the covariance matrix Cn is a symmetric Toeplitz matrix, Eq. (9.172b) may be solved efficiently by a bordering method (Press et ah, 1986). (iii) Compute the C\ (m) and Ci (ni) amplitudes of the FFT coefficient Bc (m) in Eq. (9.163b) in the discrete frequency interval 0 < m < N/2 for the conditional simulation from 'Ci(m)' Urn = C\2X + C2(m)
(9.172c) (9.172d)
In the negative-definite discrete frequency interval N/2 +1 <m
819
Real Ocean Waves
to the amplitudes in the positive-definite discrete frequency interval 1 < m < N/2 — 1 according to Ci(N - m) = C\{m)
and
C2(N - m) = C2(m),
1 < m < N/2 - 1. (9.172e,f)
For m = 0 (= the mean) and m = N/2 (= Nyquist or folding frequency) d ( m ) = C2(m) = 0.
(9.172g)
(iv) For each discrete frequency fm=m df in the interval 0 < m < N/2, initialize the complex-valued FFT coefficient Bc(m) for a conditional simulation at x according to Bc(m) = [Ci(m) - iC2(m)] exp -i(kmx),
0 < m < N/2.
(9.172h)
In the discrete frequency interval N/2 <m
= B*(m),
N/2 < m < N - 1,
(9.172i)
where B*(m) =the complex conjugate of Bc(m). (v) Synthesize the sequence Bc{m) by inverse FFT to obtain the conditional time sequence Bc(m)expil——\, m=0
^
n = 0,1,2,..., AT-1.
(9.172J)
'
Again, the normalizing constant CM that appears in all FFT algorithms as discussed in Sec. 9.2following Eqs. (9.10) is not included in Eq. (9.172i); and this normalizing constant C/v must be determined uniquely as in Table 9.1 in Sec. 9.2 for each FFT algorithm. An example of this conditional simulation algorithm for an embedded wave group is illustrated in Fig. 9.40. A proof that the conditional simulation r)c{n) given by Eq. (9.172J) contains the desired embedded sequence of a deterministic wave or wave group r\e (n) in a specified interval n\ < n < « i + v i s given by Hudspeth et al, (1999). A proof that only the values of the unconditional
820
Waves and Wave Forces on Coastal and Ocean Structures
100
110
130
time (sec)
time (sec)
(c)
1.(0 (m) 80
90
100
110
time (sec) Fig. 9.40. Embedded wave group (H = 4.0 m; Tg = 12.8 sec) in a Goda-JONSWAP (DSA) unconditional simulation (Hs = 4.0 m; / 0 = 0.27 Hz; y = 1.0; N = 2048; dt = 0.1 sees): a) unconditional simulation rju(t); b) embedded periodic wave group rje(t) with embedded sequence length v + 1 = 128; and c) conditional simulation r)c(t) (Hudspeth et al., 1999).
821
Real Ocean Waves
-3 A
80
1
1
100 120 time (sec) conditional simulation unconditional simulation
Fig. 9.41. Modification of the conditional simulation shown in Fig. 9.40 only near the ends of the embedded sequence at t\ and tv following the embedding of a deterministic wave group (Hudspeth et al., 1999).
time sequence r)u{ri) near both ends ofthe embedded interval at t\ — n\dt and at tv — (v + \)dt are modified may be found in Hudspeth et al. (1999) and is illustrated in Fig. 9.41 for the conditional simulation shown in Fig. 9.40. Numerical Instabilities The column vector X in Eq. (9.172b) becomes unstable due to an illconditioned covariance matrix Cn for the following three conditions that have been determined numerically (Hudspeth et al., 1999): (1) Length of embedded sequence compared to the length of unconditional sequence The covariance matrix is well-conditioned for inverting when the ratio Rv/w ofthe embedded sequence length v + 1 (= Tp(g)) to the total record length N (= TR) is less than 2~4 for base 2 FFT algorithms; i.e., J\
1R
(2) Compatibility ofthe variance of embedded sequence with the variance of target spectrum The numerical stability improves when the variance ofthe embedded sequence is approximately equal to the variance ofthe target or measured spectrum. (3) Size of discrete simulation time step dt and FFT interpolation If the discrete time step dt is too small, the covariance matrix Cn becomes ill-conditioned. This may occur, for instance, when simulating digitally the
Waves and Wave Forces on Coastal and Ocean Structures
822
motion of a planar wavemaker (Hudspeth and Borgman, 1979). A small simulation time step produces numerical instabilities in the Cn matrix that may be avoided by applying an FFT interpolation algorithm. Stable numerical conditional simulations may be computed for a relatively large time step dt and a relatively small number of time values N. Applying an FFT interpolation algorithm, the simulation may be transformed to drive a planar wavemaker by reducing the size of the time step by the ratio dt/ j and subsequently increasing the number of time values by jN so that the frequency distribution of the energy in the target variance spectrum will remain unchanged. This may be accomplished if df =
T777
Ndt
=
TTT
=
UN) (?)
>r
Njdtj
= a constant.
(9.174)
For base 2 FFT algorithms, both Nj and j must be equal to 2 raised to an integer power. Figure 9.42 illustrates a comparison made at the O. H. Hinsdale-Wave Research Laboratory at Oregon State University between a measured and a conditionally simulated embedment of a single deterministic wave group of significant wave height Hs = \3 ft, wave group period Tg = 15.36 sec and a discrete embedded length of v + 1 = 64. This deterministic single wave group
70
75
80
85
95
time (sec) Fig. 9.42. Comparison between rjc (solid —) and measured ? ; ( - • - • - ) conditional simulation with an embedded wave group in the 2D wave channel at the O. H. Hinsdale-Wave Research Laboratory at Oregon State University (Hudspeth et al., 1999).
Real Ocean Waves
823
sequence was embedded in the unconditional time sequence at t\ = 72 sec. The wave channel simulation was measured at a distance* = 136.2 ft. fromaplanar hinged wavemaker in a channel depth h = 7.0 ft. The deterministic single wave group was embedded into a DSA conditional simulation from a GodaJONSWAP variance preserving spectrum in Eqs. (9.48) in Sec. 9.3.3 with parameters in Eqs. (9.48) of y = 1.0, m0 = 0.11 ft 2 (#, = 1.3 ft), f0 = 0.3 Hz. The conditional simulation time sequence of length N = 2048 and discrete time step dt = 0.24 sec was transformed by an FFT interpolation algorithm with the parameter j = 22 = 4 in Eq. (9.174) to a new simulation time sequence of length Nj = 8192 and discrete time step dtj = 0.06.
9.7. Data Analyses: An Example from Hurricane CARLA Methods for the analyses of time series are sophisticated and robust (vide., Blackman and Tukey, 1959; Bendat and Piersol, 1980 and 1986; Box and Jenkins, 1976; Otnes and Enochson, 1972; inter alios). An example from a wave record from Hurricane CARLA measured in the Gulf of Mexico is given only to illustrate some of the methods that are available for analyses of random data and is not intended to be a substitute for the more detailed references noted above. The following analyses are applied to a water surface elevation record r](t) measured during Hurricane CARLA that are digitized at dt = 0.2 sec intervals and contains N = 4096 data values in a record that is TR = 819.2 sec in length. The example analyses include the following: (i) Compute the 4096 two-sided, complex-valued FFT coefficients Bm for the time series rj(t) for Hurricane CARLA85 digitized at dt = 0.2 sec; compute the mean n\ and variance //,2 by Eqs. (9.14a, b) in Sec. 9.3 for the N = 4096 values of the time sequence; and plot both the time sequence r](n) and the unsmoothed raw two-sided amplitudes of the complex-valued FFT coefficients | Bm | to obtain a two-sided amplitude spectrum of the water surface. (ii) Divide the total discrete time sequence of r}(ri) with NR = 4096 discrete values into eight equal length subrecords of Ns — 512 values that are each Ts = 102.4 sec in length. Compute the complex-valued FFT coefficients Bm for each of these Ns = 512 value subrecords; compute the mean/xi and variance fi2 from Eqs. (9.14a, b) in Sec. 9.3 for each of
824
Waves and Wave Forces on Coastal and Ocean Structures
the eight subrecords; compute a smooth one-sided amplitude spectrum 21 Bm | by averaging the eight FFT coefficients from each subrecord at each discrete frequency fm = mdf; compare the smoothed measured spectral density function 2\Bm\/df computed from the averaged FFT coefficients with the generic two-parameter spectral density function from Eq. (9.29c) in Sec. 9.3.1; and place 90% confidence intervals from Eq. (9.63) in Sec. 9.3.5 on the ten most energetic spectral estimates around the spectral peak frequency /o = coo/ln. (iii) Normalize the discrete time sequence r)(n) in accordance with Eq. (9.80a) in order to obtain a zero mean, unit variance realization f (n) and compare the pdf for this normalized time sequence with a Gaussian pdf in Eq. (9.80b) for a zero mean, unit variance process. (iv) Compute the envelope function A(t) by Eqs. (9.136) in Sec. 9.5 for wave groups by the Hilbert transform fj (t) for periodic time sequences from Eqs. (9.135). Plot the wave profile rj(t), the Hilbert transform f){t); and the envelope function A(t) for the first subrecord #1 of subrecord length NS = 512. Results (i) The time series t](t) for the hurricane waves and the unsmoothed raw twosided amplitudes of the TV = 4096 complex-valued FFT coefficients \Bm\ for the hurricane time series are shown in Figs. 9.43 and 9.44, respectively. Note in Fig. 9.43 near t = 720 sec that there is a wave of approximately H = 40 ft in wave height that occurs in a water depth of approximately 100 feet. The mean from the time sequence is JJL\ = 0.01 ft and the variance mo = 22.31ft2. 25 20
&10
S5 -5 -10 -15 -20 0
120 240 360 480 600 720 840
t [sec] Fig. 9.43. Time series of hurricane waves from Hurricane CARLA85 measured in the Gulf of Mexico (MI =0.01 ft and n2 = 22.31 ft2).
Real Ocean Waves
825
Hi
m
-I—r
~n~
:i:
•
m in 0.001
it 1024
- i — i -
i—r
•bn*E SE3i = P=i: -I—I-
t±=tzt: -•-I—[I I I 2048
3072
4096
m Fig. 9.44. Amplitudes of FFT coefficients for Hurricane CARLA85 waves in Fig. 9.43 ( d / = 1/819.2Hz and M2 =22.31 ft2).
(ii) In order to obtain smoothed spectral amplitude estimates | Bm | from the raw FFT amplitude estimates \Bm | shown in Fig. 9.44 the raw FFT amplitude estimates are smoothed by the segment averaging method of eight distinct and disjointed subrecords of 512 data values each from the original wave time sequence of 4096 data values (Bendat and Piersol, 1986, Chapter 8.5.4). Figure 9.45 illustrates the eight time series segments (subrecords) from the Hurricane CARLA85 wave record of 4096 data values. Note in Fig. 9.45 in subrecord #8 near t = 710 sec the nearly 40 foot high wave that was identified near t = 720 sec in Fig. 9.43. The reason for this small difference in the time of approximately 10 sec for the time of occurrence of this large wave between Figs. 9.43 and 9.45 and that the wave time series in each figure do not appear to be continuous from one subrecord to the next subrecord is due to the 2.4 sec truncation of each subrecord to exactly Ts = 100 sec for the purpose of plotting the subrecords (i.e., Ts = 102.4 sec vs Ts = 100 sec in Fig. 9.45). The FFT analyses, however, were computed on the full N = 512 data values and T$ = 102.4 sec for each subrecord. The one-sided FFT amplitude estimates 2\Bm\ computed from the raw complex-valued FFT coefficients for each of the 8 time sequence subrecords shown in Fig. 9.45 are illustrated in Fig. 9.46 as one-sided energy amplitudes computed by Eq. (9.15j) for comparisons later with wave energy spectral densities. Only the 40 most energetic, low-frequency amplitudes estimates are illustrated in Fig. 9.46 in order to avoid plotting the large number of zero amplitude estimates shown in Fig. 9.44 and, consequently, to improve the identification of the energy contributed by each frequency to the total variance (cf., the
826
Waves and Wave Forces on Coastal and Ocean Structures
40
140
60
160
180
200
380
400
580
600
t (sec)
t (sec)
300
300
320
340
360
t (sec)
500
t (sec)
600 620 640 660 680 700
t (sec)
500
520
540
560
t (sec)
740
760
t (sec)
Fig. 9.45. Time series of hurricane waves from eight subrecords from Hurricane CARLA85.
827
Real Ocean Waves
10
^-0.1 0.01
1 1 13 44
m
E
i i U'
mmm
"-i—+•—t — i -
II
0.001
BB HhMWM
I I
1
~
isg=E: xaxlxt:
.iiaiL
'as
:ii3:
MM
30
a
:sig=e
_n i
m
0.01
rti3=t i i I*I iiiai -t-t-
#3
'ill IS
PI- T"PM
0.001 20
10
i
' ' "!Jq
30
0.01
0.001
as
1~Ms**stl
:nx I I I
10
:t±=ti
^•o.i 0.01 0.001
iap=t rfis_„ f a i n
•1-t-l-l-
•t-t-i-t-
•-f-t-l-h
III3EE I I 3 I C
II33IC
10
30
20 m
m "!1im w H^ ™-IS UPP PW .fe^dd-tefe^^^
40
i3E 133=E : M 3 = c
!
10
•1—t-l-
Sf
20
iiiii
mm M ill! 111 i^^= EI3:
10
20
= E33=E
0.001
m¥E
10
~
10
r i E
L_L_| '
m
30
40
0.01 0.001
"t-t-I 1 "
5%5x^4
II3ZC
'xilxc
EEIQElE -4-44-11
20
I I I'l
30
-4-44-1-
E « g = i = E3E3:=1=1= E E 3 3 E I E = EEESEF - 4 - 4 - 1 -1— - 4 - 4 4 - 1 - - • • 4 4 y ^ 1 1*! = £33=1H EE53EE : E 3 3 = F -4-44-1- -4-44-1- -4-44-11 1 I I 1 1 1 1 1 I I 1
10
ttsy
I I I I
5 S 3 3=1= = S3=I=F= =t±=t;J= :£±3=l= -4-4—1—1— - 4 - 4 4 - 1 - - 4 - 4 4 - 4 M i t e = g | 3 = E E3==f=!=E= 1 T
^0.1 I
as
!!s
40
#7
S33SE x a d i t
:iIS
0.01
J-4-4-
#6|
!H irti
I • > *~f
40
£333 + 44-
kti!±b^jS ,444. =Baa=e j s a s s e
ixnnx
30
33 = E5E1E + 4-1-
WWWBMlBWBaB'
' I
I=t=t= x%-i^
30
20
10
^ 0 . 1 \i
Sill
iill
-4—I—1-
l
1JJ=U igg=E
iiiim
IUXDZIX I I I I
40
irxazt-
• i i l l
n 0.01
i
i*B_o.i !
I 3 3 3 3 E E 5 3 E #5 + 44 :J,j,J_ULj,4_l.
t-44!+
—
10
a .r. i
1 0
a ~
0.001
E333E i l l #2 4-44-I-
0.001 0
—
0.01
_
wmWrnM
'3HPH
40
WP !Hfl IB! ; sag =e
10
i^.0.1
1- -1~t-t»l:£33E
20 m
10
10 = #1
20 m
40
:*:»> 4 4 4 1-
^
:
;1tr
30
40
Fig. 9.46. One-sided raw FFT amplitudes for the hurricane waves from the eight subrecords shown in Fig. 9.45 from Hurricane CARLA85 (df = 1/102.4 Hz.).
non-deterministic definition of Fourier coefficients in Sec. 9.2). Because the total number of discrete values has been reduced from NR = 4096 values for the full time sequence shown in Fig. 9.44 to Ns = 512 values for each of the eight subrecords shown in Fig. 9.45, the total number of discrete frequencies
828
Waves and Wave Forces on Coastal and Ocean Structures
Table 9.10. Summary of peak frequency discrete values mPN for the Hurricane CARLA85 time sequences (dt — 0.2 s). N mPN f0 = mPNdf = mPN/Ndt too = 27r/b(rad/s)
(Hz)
A^=4096
Ns = 512
65 0.079 0.499
8 0.078 0.490
in the spectra of each of the eight subrecord has also been reduced by eight by Eq. (9.10). For the Hurricane CARLA85 wave amplitude spectrum, the radian peak frequency is coo = 0.499rads/s (vide., Table 9.3 Column 3, Record No. 06885/1 in Sec. 9.3.1). The discrete value mPN for the radian peak frequency o>0 = mPN2ndf for a discrete sequence of length N may be estimated by m
^ p _ lizdf
=
mmt lit
=
0-499
ftads/B)
s =
a o
In
where N dt — l/df from Eq. (9.10). The values for mPN for both the full NR = 4096 record and each of the eight subrecords of Ns = 512 discrete values are summarized in Table 9.10. Because the number of data values NR = 4096 in the total Hurricane CARLA85 time sequence was divided by eight for each subrecord, the value of the peak frequency mPN for the full NR = 4096 length record was also reduced by approximately eight for each subrecord. However, the numerical value for the peak radian frequency coo remained constant for each of the eight subrecords and approximately equal to COQ = 0.499 rads/s by Eq. (9.175). The reason that the spectral peak frequency COQ for each of the eight subrecords is not exactly coo = 0.499 rads/s is because the integer m is not exactly m = 8 but, rather, is equal to m — 65/8. The mean [i\ and variance JJL2 are random variables for each of the eight subrecords and the ensemble averages of these two random variables are tabulated for comparison with the temporal average of the total CARLA85 time sequence in Table 9.11. This comparison of ensemble averages with a single temporal average is an illustration of the stationary ergodic hypothesis from Eq. (9.1) in Sec. 9.1 The last two columns in Table 9.11 " S / 8 " and "4096" demonstrate that mean ji\ and variance (X2 computed from the ensemble average of the eight
829
Real Ocean Waves
Table 9.11. Comparison of means \i \ and variances n 2 for the C ARL A85 time sequences in Fig. 9.45. Record # Mi(ft) /x2(ft)2
#1
#2
1R
m
115
116
in
;/8
E/8
0.05 -0.22 -0.04 0.18 -0.24 0.12 0.17 -0.12 -0.01 25.41 21.49 16.13 22.26 24.64 17.32 24.72 26.64 22.34
10
20
30
4096 -0.01 22.31
40
m Fig. 9.47. Comparison between the smoothed segment averaged one-sided spectral amplitudes from the eight Hurricane CARLA85 subrecords with the generic two-parameter spectrum for spectral parameters mo = 2 2 - 3 1 ft2; <»0 = 0.499 rads/s; 90% confidence limits a = 0.05 (*.) and a = 0.95 Or).
discrete time sequences in Fig. 9.45 in the column " E / 8 " are statistically equivalent to the mean /xi and variance fj.2 computed by the temporal average over the entire NR = 4096 values of the full Hurricane CARLA85 time sequence in column the "4096". The eight raw two-sided spectral amplitude estimates | Bm | for each discrete frequency m df are averaged over the eight subrecords or segments; and the two-sided smoothed spectral amplitude estimates | B,„ \ are converted to one-sided spectral densities by Eq. (9.15i) in Sec. 9.3. Figure 9.47 compares the eight smoothed segment averaged one-sided spectral amplitude estimates 2\Bm\ computed from the two-sided complex-valued FFT amplitude estimates Bm from the eight Hurricane CARLA85 subrecords or segments with the two parameter generic spectral amplitudes computed from the generic
830
Waves and Wave Forces on Coastal and Ocean Structures
0
10
20
30
40
m Fig. 9.48. Comparison between the smoothed segment averaged one-sided spectral densities from the eight Hurricane CARLA85 subrecords with the generic two-parameter one-sided spectrum for spectral parameters mo = 22.31 ft , COQ = 0.499 rads/s; 90% confidence limits a = 0.05(A) and a = 0.95(T).
spectral density function in Eq. (9.29c) with the spectral parameters of a mean /U-i =0.0 ft; a variance/^2 = mo = 22.3 lft 2 ; and a peak radian wave frequency &>o = 0.499 rads/s. Figure 9.48 compares the eight smoothed segmentedaveraged one-sided spectral density computed from the two-sided FFT amplitudes by Sm(m) = 2\Bm\2/df from the eight CARLA85 subrecords or segments with the two parameter generic one-sided spectral density from Eq. (9.29c) in Sec. 9.3.1 with the spectral parameters of a mean ii\ — 0.0 ft; a variance mo = 22.31 ft2; and a peak radian wave frequency a>o = 0.499 rads/s. The conversion from spectral densities to spectral amplitudes is given by Eq. (9.15i) in Sec. 9.3. Because N dt = 512(0.2) sec = 102.4 sec, the spectral density ordinates in Fig. 9.48 are approximately 100 times larger than the spectral amplitude ordinates in Fig. 9.47 by Eq. (9.15i). The Hurricane CARLA85 amplitude and density spectra are both plotted as "dots • "; and the Generic 2-parameter values as "solid lines —" in Figs. 9.47 and 9.48. The 90% confidence intervals computed from Eq. (9.63) in Sec. 9.3.5 are plotted in both Figs. 9.47 and 9.48 for the ten most energetic frequencies around the spectral peak frequency m = 8. The a = 0.05 lower limits are plotted as vertical triangles • ; and the a = 0.95 upper limits are plotted as inverted triangles T. (iii) The 4096 time sequence values rj(n) are normalized in accordance with Eq. (9.14c) in Sec. 9.3 or Eq. (9.80a) in Sec. 9.4.1 in order to obtain a dimensionless zero-mean /xi = 0 and unit variance /X2 = 1 time sequence £(«)
831
Real Ocean Waves
(=(
that may be compared with the zero-mean, unit variance Gaussian pdf from Eq. (9.80b). The normalized time sequence values £(«) are shown in equal 10% probability bins of unequal widths in Fig. 9.49 (Bendat and Piersol, p. 93, 1986) and are compared with a Gaussian pdf for a zero-mean, unit variance process. Because more than 20% of the time sequence values from Hurricane CARLA85 are in the ±2 bins compared to only 5% for a Gaussian pdf, the Hurricane CARLA85 waves are definitely not Gaussian! There are many more large positive and negative time sequence values of the water surface in the hurricane waves than in linear Gaussian surface gravity waves, (vide., Fig. 9.18 in Sec. 9.4.1) (iv) The envelope A{t) computed by Eq. (9.136c) for a random wave realization from the relatively broad-banded spectrum from subrecord #1 from Hurricane CARLA85 shown in Fig. 9.45 is illustrated in Fig. 9.50 along with the wave realization t){t) and the Hilbert transform /)(?) of the realization computed by Eqs. (9.135) (vide., Sec. 9.5 for wave analyses by the Hilbert transform and the wave envelope function).
9.8. Random Wave Forces on Small Circular Members Wave forces due to random waves on the pile members of steel-jacketed, space-frame offshore platforms may be modeled by the Morison equation
832
Waves and Wave Forces on Coastal and Ocean Structures
-20 -I
0
i
1
20
i
1
40
1
1
60
i
1
80
i
1
1
100
t(sees) Fig. 9.50. Water surface 77(f), Hilbert transform ij(t) and envelope function A{t) for a random wave realization from subrecord #1 from Hurricane CARLA85.
(vide., Chapter 7.2). A brief literature review of the early publications on this topic may be found in Grecco and Hudspeth (1983) and in Tuah and Hudspeth (1985b). A probability density function (pdf) computed from the Fourier cosine transform of a moment generating function and a covariance function for the wave force per unit length on a small vertical pile are derived for the Morison equation (Tuah and Hudspeth, 1985b). The dynamic response of an idealized steel-jacketed, space-frame offshore platform from Wave Project II (WPII) is evaluated from a one-dimensional wave spectral density function and the linearized Morison equation (Grecco and Hudspeth, 1983). 9.8.1. Probability Density Function p(Y) (pdf) and Covariance Function CfTFT{r) for Nondeterministic Wave Force per Unit Length for a Small Vertical Circular Pile A probability density function p(Y) (pdf) computed from the Fourier cosine transform of a moment generating function ^YC?) and a covariance Function CfTpT (T) are derived for a wave force per unit length Y(t) computed from the Morison equation (Chapter 7.2) for small diameter vertical piles and linear, Gaussian waves (Tuah and Hudspeth, 1985b ). The pdf and covariance functions may be compared with alternative functions derived by Borgman (1965) and by Hino (1969) and with measured wave forces from WPII on a single vertical pile on a prototype steel-jacketed, space-frame offshore platform.
833
Real Ocean Waves
Wave force per unit length probability density function p(Y) (pdf) The wave force per unit length probability density function p(Y) (pdf) is derived from the Fourier cosine transform of a moment generating function * Y O ) (vide., Sec. 9.4 or Davenport and Root, 1963). The dimensionless wave force per unit length Y(t) from a linear, Gaussian sea on the small vertical shown in Fig. 9.51 may be computed from the Morison equation in Chapter (7.2) by Y(0 =
dFT{t) Cmp(7tD2/4)^
Kuu\u\ + Ka
du dt'
(9.176)
where Cm = dimensionless inertia coefficient; p = fluid mass density; D — vertical pile diameter; oa — standard deviation of the horizontal water particle acceleration; u = horizontal water particle velocity; du/dt = horizontal water particle acceleration; and the coefficients Ku and Ka are defined as Ku = Ka =
Cd_
pnD OaJ 1
(9.177a) (9.177b)
On'
n!m/>n>)>i>>>iiu>p)fyii I I
>>)>>>>} mil >)ii> > in t I
Fig. 9.51. Definition sketch for wave forces on a small vertical circularpile (Tuah and Hudspeth, 1985b).
834
Waves and Wave Forces on Coastal and Ocean Structures
A characteristicfunction I/T (co)fromEq. (9.75) in Sec. 9.4 for the random variable Y(t) may be computed from TM<W)
du Kadi
= £[exp(r'wY(0)] = E Kuu\u\ +
=
P dud \dtJ' L Looexp r VuUM + Kap{u^)\'to) V' ^ )
VaF
(9.178) where £[•] = expectation operator in Eq. (9.1) in Sec. 9.1 and the joint probability density function p (•, •) for u and du/dt is given by
p(u,du/dt)
=
1 ITTOUO,
du/dt
1 "2
exp
(9.179)
On
where cru,aa = standard deviation from Eq. (9.14b) in Sec. 9.3 of the horizontal water particle velocity u and horizontal water particle acceleration du/dt, respectively. Scaling the horizontal water particle velocity and acceleration by their standard deviations according to du
V<7 =
Ou
a =
dt
and substituting Eq. (9.179) into Eq. (9.178) yields
2n
f Jo
/•OC
JO
= -\f 1 2TT
cos(Kucr^coq) exp (-^)
dq
•s/q
cos(Kaaaa>y/a) exp (=^L)
da
y/tt
fi(a),q)exp[-j-\dq
A[l)x£2(I
x / (?)*"jf
f2(co,a) exp I — ) da (9.180a)
835
Real Ocean Waves
where the Laplace transforms £;(•) in Eq. (9.180a) are defined by (Oberhettinger and Badd, 1973)
cos 2 I arctan CX ( - ) = V27T
CO
.?J 1/4
1+
CO
1
£2 ( - ) = V 2n- exp
(9.180b)
(9.180c)
-(£ a <7 a
where £ =
(9.180d)
2*W
Substituting Eqs. (9.177b and 9.180b-d) into Eq. (9.180a) yields the following moment generating function ^YC?) for Y(f):
cos 2 I arctan *Y(*)
=
-,1/4
exp
(9.181)
1+ r The probability density function p(Y) (pdf) for the symmetric dimensionless random wave force per unit length Y may be computed from the Fourier cosine transform of the moment generating function ^yfa) according to (Papoulis, 1984) P(Y) n Jo
VY(s)
cos(Ys)ds.
(9.182)
Substituting Eq. (9.181) into Eq. (9.182) and making a change of variables s
- = V rr
836
Waves and Wave Forces on Coastal and Ocean Structures
gives
COS^Y-V/T") cos(arctan
P(Y)
2n JQ
=
( x ex
Pi
_
T I JT
x exp
^(l+T)1^
£ f°° Y~ / /(^' Y ' T ) 2TT
y/x)
¥ —T
^
T
(9.183)
l 2
where £.(•) inEq. (9.183) is the Laplace transform of / ( § , Y , r) that may be computed from
£
= / exp
'y) v I
(^
,x
exP|
(?+
+ exp
,Y) 4
)D_1/2(g+Y)
(I - Y) 2
D_i/2(£-Y) (9.184)
where D_i/2(«) = the parabolic cylindrical function (Miller, 1965) given by
D_i /2 (X) =
ArX 2 fn\K
"/_i / 4 ( £ ) - /i/4 ( f ) ] , [/_, /4 ( ? ) + /i/4 ( ¥ ) ] ,
X>0
(9.185a)
X < 0 , (9.185b)
where the modified Bessel function of the first kind Is(») of order 5 (vide., Eq. (2.52) in Chapter 2.4.3) may be expanded in the following polynomial
837
Real Ocean Waves
series (Luke, 1975):
^/4J2bnOOSn
I1/4(0 =
l?l<8
arccos —
(9.186a)
w=0 oo
1/4
/-i/4(0 = T J2Cn
cos n
|?|<8,
arccos —
(9.186b)
n=0
16
^-i/4(o = - ^ E * cos n V(2^0 n=0 1-1/4(0
=
exp(-f)
7 -)]
oo
,
f > 8, (9.186c)
10 COS ft
arccos
rc=0
•-o: + h (0, /4
S > 8. (9.186d)
The series in Eqs. (9.186) converge very rapidly; and the coefficients bn,cn, dn and en are tabulated by Luke (1975) and by Tuah and Hudspeth (1985b, Appendix). The pdf p(Y) in Eq. (9.183) may now be expressed as a parabolic series according to ex 2 PW = \H~ V an P(-Y )
expp±P^]D_1/2(§+Y) + exp[^^]D_i /2 (?-Y)
(9.187)
The symmetric parabolic pdf in Eq. (9.187) has a higher and narrower peak than a Gaussian (Normal) pdf in Sec. 9.4 as illustrated in Fig. 9.52 for a normalized dimensionless wave force per unit length f = Y/^/ay for £ = 0.53. Statistical moments for the dimensionless wave force per unit length Y may be computed from the derivatives with respect to s of the moment generating function ^YC?) according to (Davenport and Root, 1963) E[YnexpsY]
= ^VY(s),
s = 0,
(9.188a)
where the odd moments n = 1, 3, 5, 7 , . . . are zero because ^YC?) is symmetric; and the even moments n = 2, 4, 6, 8,... may be computed from 2m i
E[Yzm] =
d2m
ds 2m
*Y(S),
5 = 0 and
m = 1, 2, 3, . . . .
(9.188b)
838
Waves and Wave Forces on Coastal and Ocean Structures 0.6
1
1
1
Q5
1
r
_ -
\
/ .
If—
Q2 -
\\ \\ \\
A / 1//
CI °-4
1
// \ \
0.1 -
/ -3
«-=-! -2
/ // s/
.
\\ \ \ ^ \
1 -I
0
1
r^="*»^—L
I
2
3
4
C=Y/\/oY Fig. 9.52. Comparison between a Gaussian (Normal) pdf (—) and parabolic pdf in Eq. (9.187) (- - -) for £ = 0.53 (Tuah and Hudspeth, 1985b).
The mean (n = 0) and skewness (n = 3) are equal to zero. The standard deviation ay and excess of kurtosis A4Y for the dimensionless wave force per unit length Y are given by
aY = JlK^
+ l,
X4Y = l05Kua* + ISK^al
(9.189a) (9.189b)
Even though the moment generating function in Eq. (9.181) is different from the moment generating function derived by Borgman (1972b), the pdf in Eq. (9.187) computed from Eq. (9.181) is identical to the pdf given by Borgman (1972b). Both Hino (1969) and Borgman (1972b) derive formulas for the standard deviation of the dimensionless wave force per unit length Y(t) as a function of the variance of the horizontal water particle velocity ou. The formulas given by Hino (1969), by Borgman (1972b) and by Eqs. (9.189) are different; but numerical values computed from each of these formulae are approximately equivalent (Tuah and Hudspeth, 1985b).
Autocovariance Function CFTFT(T) f° r the Wave Force per Unit Length FT(t) In the following derivation, the dimensionless horzontal water particle velocity u(t) and acceleration du(t)/dt are assumed to be real-valued functions. If the real part of a complex-valued function is taken, then functions with temporal dependencies of t + r below must be the complex conjugate values of the
839
Real Ocean Waves
complex-valued functions (Tuah and Hudspeth, 1985b). The auto-covariance function CpTFT ( T ) f° r m e wave force per unit length may be computed from the expectation operator £[•] in Eq. (9.1) of the dimensionless wave force per unit length Y(t) according to
Kuu{t)\u(t)\ + Ka E[Y(t)Y(t + x)] = E x
du{t) dt
Kuu{t + x)\u(t + x)\ + Ka
du(t + x) dt
= K„E[u(t)u(t + r)\u(t)\\u(t + r)|] +
KuKaE u(t)\u(t)\
+
KuKaE
+ K2aE
du{t + x) dt
du(t) u(t + T)\u(t + r)\ dt
du(t) du(t + r) dt dt
(9.190)
where the functions with temporal dependencies t + x (e.g.; u(t + r), etc) are complex conjugate values if complex-valued kinematics u(t) and du(t)/dt are substituted. The following change in notation is introduced for compactness of notation:
u(t) = ut, du(t) dt
= at,
u(t + T) = uz,
(9.191a,b)
du(t + x) — =aT. dt
(9.191c,d)
The following auto- and cross-covariance functions C(„)(r) from Eqs. (9.17 and 9.18) between ut,uT,at and ar are defined:
Cuu{r),
Cua(x),
Cau{x),
Caa(x).
(9.191e-h)
840
Waves and Wave Forces on Coastal and Ocean Structures
The following joint pdf's are required in order to evaluate Eq. (9.190): (pu(u2 + u2) exp p(u,,ux)
=
=
=
(9.192b)
- C2a
(oua, + oau2x 2(auaa -
2Caua,uT) C2) (9.192c)
2nJouaa - C2U exp
p(at,aT) =
(9.192a)
(aauf + aua2 - 2CuautaT) 2(auaa - C2) 2n^auaa
exp p(at,uT)
2K - Clu) Txyjol - C2UU
exp p(ut,az)
2Cuuutux)
2
(oa(at +ax) 2nja2
-2CaaataT) (9.192d)
- C2a
With the following change of variables: C, R = -J£,
u,
= ut^au(\-R2),
ux = ux^au(l-R2),
(9.193a-c)
O",,
each of the expectation operators in E [•] in Eq. (9.190) may now be evaluated. First, K2E[utux\ut\\uT\]
= K2a2(l
- R>Z2YN 2 ; E{utux\ut\\ux\\
2-K
J—oo J—oo
x exp — -{ut + uz —
11
it2
2Rutux)dutdux
smh(Rutux)Gxpl—-^-\du
dux, (9.194a)
841
Real Ocean Waves
that may be integrated by Laplace transforms (Oberhettinger and Badd, 1973) to obtain KuE[utuT\ut\\ur\] 8
= K2uau \lR2 + \
2
s
/3 3 7
= K2uou<S>(R),
(9.194b)
where the hypergeometric function 2^1 (•>•>•>•) (Oberhettinger, 1965)
ma
Y be computed from
,„(,,»,,,„-'!> fr*;•>?»+»>£:. ,„<,.(9,940 r(a)r(o)•'—•'
r(n + c)
n\
n=0
where T(«) = Gamma function in Eqs. (2.6) in Chapter 2.2.5. Second, combining the following two expectations in Eqs. (9.190) and substituting Eqs. (9.192b, c) gives KuKa [E[ut\ut\ar] + E[atur\uT\]] />00
OO
/ /
= KuKa
-00 OO
/
r
^V ° "° a
{/.
x {1
00
_
/>00
/
(atuT\uT\)p(at,uT)datduT
-00 •/—00 OO
KuKa / r
(ut\ut\aT)p(ut,aT)dutdaT
J—00
"TI"TI Ma /
C
-00
a ; cosh ( O^Oa -
OnU
a"z
exp — 2((r a - Cl„) u a
^ 52 - l exp Ct ua
ouat dat \ duT 2{puoa - Cla)
= 0
(9.195)
because Cua = — Cau and because the integrand in curly brackets {•} is an odd function of at (Tuah and Hudspeth, 1985b). Finally, the fourth expectation operator in Eq. (9.190) is the definition for autocovariance function for the horizontal water particle acceleration Caa', i.e., KaE[atar] — KaCaa.
(9.196)
842
Waves and Wave Forces on Coastal and Ocean Structures
The expectation of the dimensionless wave force per unit length Y(t) in Eq. (9.190) may now be expressed only in terms of the auto-covariance functions for the horizontal water particle kinematics Cuu and Caa by £[Y(0Y(/ + T)] = < a > (R = ~ ^ )
+ K2aCaa(r).
(9.197)
Substituting Eqs. (9.177) into Eq. (9.197) yields the following auto-covariance function for the dimensional wave force per unit length Fjit): CFTFT(T)
=
I — — I au
^2
1 + 1
I Caa(r) (9.198)
where <&(R) is tabulated in Table 9.12. Tuah and Hudspeth (1985b) compare the /(/?) function in the auto-covariance function in Eq. (9.198b) with similar functions derived by Hino (1969) and by Borgman (1972b). Table 9.12 lists each of these auto-covariance functions where R is defined in Eq. (9.193a) Even though the three auto-covariance functions in Table 9.12 are symbolically different, numerical values computed from each function are essentially equivalent (Tuah and Hudspeth, 1985b)! The two-sided spectral density function GFTFT(W) in Eq. (9.15b) may be computed from Eq. (9.198) by the Wiener-Khinchine Fourier transform pair in Eqs (9.15) according to " 0c 0
1 GFTFT{U>)
—
f
/ \l1lt
Cf r F r (t)exp-|-(/(i>r)rfr.
(9.199)
J-oo
Table 9.12. Comparison of auto-covariance functions where R is defined in Eq. (9.193a). Reference/Equation No.
/ (R)
Hino (1969)
- IY2 + 4R2) arcsin(fl) + 6Ry/l - R2~\ R(2n-l) Y* IT ^ {In- 1)!
Borgman (1972b)
-
n—1
*(tf)(Eq. (9.193a))
2fl 2 + 1 - ~ ( l
-
tf2)5/Vl
(
2 2' 2'
,
843
Real Ocean Waves
POO (%) Fig. 9.53. Comparison between cdf s for dimensionless measured (DDD); theoretical (—) and (—) Gaussian wave force per unit length (Tuah and Hudspeth, 1985b).
Tuah and Hudspeth (1985b) computed numerically the cumulative distribution function (cdf) P(Y) for the dimensionless wave force per unit length Y(t) by substituting the pdf in Eq. (9.187) into Eq. (9.66) and integrating Eq. (9.66). The results were compared with a measured cdf computed from a force transducer on an offshore platform during Hurricane CARLA by Wave Project II. Figure 9.53 illustrates that the comparison between theory and data is very good between ±2 standard deviations ay. The differences between both the theoretical (—) and the measured (DDD) cdf's with the Gaussian cdf (—) in Fig. 9.53 is a consequence of the strong nonlinearities in the wave forces per unit length on small members (cf., Fig 9.18 in Sec. 9.4.1). 9.8.2. Stochastic Response of Space-Frame Offshore Structure Grecco and Hudspeth (1983) evaluate the dynamic response of a prototype space-frame offshore structure in the Gulf of Mexico from Wave Project II (WPII) to both measured and simulated stochastic wave forces. They include references to early research on the topic that may also be found in Clough and Penzien (1975). The nonlinear drag force in the modified wave force equation (WFE in Chapter 7.8) or relative motion Morison equation is linearized in a time-average, mean-square sense and applied to a two-dimensional, idealized, lumped-mass structural model in a frequency domain, spectral analysis. Linear ocean waves from a one-dimensional wave spectrum from Sec. 9.3
844
Waves and Wave Forces on Coastal and Ocean Structures ELEVATION +1
»
FEET (METERS) 290 (88.40)
260 (79.25) 250 (76.20) -222 (67.67) 190 (57.91) -165 (50.29) 150 (45.72)
%—s/VW—
SOIL RESPONSE SPACED AT £ 10 FT.(3.05M) „„. , INTERVALS AA 28'—•Arv—# • —v\A/ -^ 32-WWV—,
—•WV— '33 ^-^/WV—"
361-VWV—^ '^MM—* 40'-AW/— ^ j^,»„ y -^A/V\Ar-4 41
<«—vAAAA—'
-*~JC Fig. 9.54. Global x-z coordinates and structural nodes for an idealized WPII Gulf of Mexico space-frame offshore platform (Grecco and Hudspeth, 1983).
are simulated as a zero-mean, stationary, ergodic stochastic process (vide., sec. 9.1). Dynamic equations of motion for the coupled wave-structure system (vide., Chapter 7.8) are solved by a normal mode superposition method. The coupled generalized damping matrix is diagonalized through an optimization procedure assuming Rayleigh damping (Clough and Penzien, 1975). The spectral analysis of Grecco and Hudspeth (1983) is reviewed below. The global x-z coordinate axes and lumped-mass node locations for the WPII Gulf of Mexico space-frame structure are shown in Fig. 9.54. The degrees of freedom are restricted to only X\{t) surge and ®s(t) pitch oscillations (cf, Chapter 8.1). The K independent equations of motion for the displacement of each of the numbered, lumped-mass nodes Xk(t) in Fig. 9.54 are [Mj]{Xk) + [Cjk]{Xk] + [Kjk]{Xk} = {FkE},
(9.200)
845
Real Ocean Waves
where the over dots (•) denote ordinary temporal derivatives of a Lagrangian variable; [Mj] = a diagonal mass matrix, [Cjic] = & square damping matrix and [Kjk] = & square stiffness matrix. The column matrices {X^}, {X^} and {Xic}=the Lagrangian accelerations, velocities and displacements, respectively, of the kth numbered node in Fig. 9.54; and the column matrix {F^} = a wave-induced exciting force on the kth numbered node in Fig. 9.54.
Linear soil springs The soil-platform interaction illustrated in Fig. 9.54 may be modeled by a system of linear Winkler springs (Penzien, et ah, 1964). The soil response S 5 (z) at elevation z may be estimated from a linear spring model by Fs(z) =
(9.201a)
Ks(z)Ss(z),
where the Winkler subgrade modulus Ks (z) is given by e+z sinh D/2 .D/2 D/21 - 2(D/2)2z + lz2 + z3 + 3fl 2 J(D/2)2 + (l + z)2 sinh
8TT
A:S(Z) = —£,(z)
z-l 3 \_J(D/2) + (l-z)2
z J(D/2)2+z2
2
+ -.
z3 - 2(D/2)2z sJ(D/2)2+z2
(D/2)2z + lz2 + z3 ((D/2)2 + (£ + z)2)V2
(D/2)2z + z3 ((D/2) 2 +z 2 ) 3 / 2
(9.201b) where I = the embedded length of pile; D/2 = the radius of the embedded pile; and Es(z) = Young's modulus for the soil that may be estimated from laboratory tests that compute the soil shear modulus Gs (z) by Es(z) = 2(1+
v)Gs(z),
(9.201c)
where v = Poison's ratio for the soil. The linear soil spring in Eqs. (9.201) are elements in the global stiffness matrix [Kj^] in Eq. (9.200).
846
Waves and Wave Forces on Coastal and Ocean Structures
Modified wave force equation (WFE, API RP2A, 1987) The relative motion form of the Morison equation (WFE) in Chapter 7.8 for each node number k is given by
FkE«) =
(Cm - V)pVk I —— + Cd^Ak(uk(t)
Xk(t) I + pVk
- Xk(t))\(uk(t)
-
,
(9.202)
Xk(t)\
where the horizontal wave particle velocity uk{t) and acceleration duk(t)/dt are applied at the undeflected &th numbered node on the submerged portion of a pile of the platform between 0 < z < — h; Cm and Cd = inertia and drag coefficients, respectively (cf., Chapter 7.6); Ak and Vk = kth numbered node cross sectional area and node volume, respectively; and p = fluid mass density. Equivalent linearization of the relative motion hydrodynamic drag force In contrast to the deterministic linearization of the quadratic hydrodynamic drag force in the Morison equation in Chapter 7.6.4 by either temporally averaging over a deterministic wave period T or by the Lorentz's method of equivalent work, equivalent linearization of the relative motion hydrodynamic drag force for stochastic wave forces requires stochastic averaging methods. The nonlinear hydrodynamic drag force in Eq. (9.202) may be linearized by the method of Krylov and Bogoliubov (Foster, 1970 ). This optimization method requires first defining a relative motion velocity rk(t) for the kth numbered node by rk{t) = Xk{t)-uk{t).
(9.203)
An optimized damping matrix [C J is a linear combination of a square structural damping matrix [C] and a diagonal equivalently linearized hydrodynamic drag matrix [C] defined by [C] = [C] + [C],
(9.204a)
where a typical element Ck in Eq. (9.204a) is given by Ck = CdLk = CdP-Ak\rk\.
(9.204b)
847
Real Ocean Waves
The linearized drag force elements in Eq. (9.204b) are added only to the diagonal elements of the square damping matix [C]. Linearization yields the following error matrix: 2 ,2
EC] + Cd^Ak
{**} =
(9.205)
{h\h\} - ICktih)
Minimizing Eq. (9.205) with respect to an optimized diagonal damping coefficient Ckk requires that
H\ dCkk
J\,„{Ckk
= 2
\\
* ,., . ^ p
- Ckk)h + Cd-Akh\hI 2
W \h
= 0,
(9.206)
where (•) = temporally averaging operator over an infinitely long record (i.e., the ergodic hypothesis in Sec. 9.1); and Eq. (9.206) may be solved for the optimized damping coefficient Ckk according to Ckk = Ckk + Ckk p (rtlhl) = Ckk + Cd'-Ak ^ - ^ . 2 r,
(9.207a)
The optimized damping coefficient for the numbered nodal elements in the hydrodynamic loading regime between 0 < z < — h are given by Cjk = (1 - 8jk)Cjk + SjkCjk,
(9.207b)
where Sjk = Kronecker delta function (vide., Eq. (2.2) in Chapter 2.2.3). The probability density function (pdf) for the relative velocity r~k for a linear, zero-mean Gaussian process is (cf, Eq. (9.79a) in Sec. 9.4.1)
848
Waves and Wave Forces on Coastal and Ocean Structures
where afk = standard deviation of the relative velocity rk. For an ergodic stationary process (cf., Sec. 9.1) 2
1 f00 ., . 1 = — / r | r t | e x p t '2TT
2
{r k\h\) = E[r k\rk\]
Z
UOr. J-oo
=
rk
drk
a.Wlc (9.209a)
^l—cr,Wk' n
\JLTZOrkrk J-oo
z
dh = ohh \°rkrk /
(9.209b) and Eq. (9.207a) reduces to Ckk = Ckk + Ckk P /8 = Ckk + Cd — AkJ —Orkrk. 1 V Tt
(9.210)
Numerical values for Eq. (9.210) may be computed iteratively with a convergence criterion established for the standard deviation for the relative velocity Orkrk (Grecco and Hudspeth, 1983). When the convergence criterion is satisfied, then the diagonal damping elements with an optimized hydrodynamic damping coefficient at the z'th iteration are computed from l
Ckt
= Ckk +
Cd-Akyj-al
— w
(9.211)
where a\ • = the standard deviation of the relative velocity for the ith iteration. 'k'k
Equivalent linearization of the hydrodynamic drag force reduces the wavesoil-structure system in Fig. 9.54 to the following set of linear, coupled ordinary differential equations: [M + (Cm - l)PV]{X(t)} = [CmpV]
+ [C]{X(t)} + [K]{X(t)}
du(t) dt
+ [C]{K(0}.
(9.212)
The system given by Eq. (9.212) may be uncoupled by the normal mode superposition method (Clough and Penzien, 1975).
849
Real Ocean Waves
Normal mode superposition solution The modal component displacement of the kth node in the rcth normal mode Xkn(t) may be represented by the product of the mode shape vector (£>kn and the modal amplitude response Yn(t) according to {X} = miY}.
(9.213)
Substituting Eq. (9.213) into Eq. (9.212) yields [M + (Cm - l)PV] {
+ [C]{u}.
(9.214)
The dynamic equations of motion for system are now represented by the set of coupled equations (9.214) that may be uncoupled by orthogonality of the normal modes to obtain [M*]{Y(t)} + [C0]{Y(t)} + [K*]{Y(t)} = {FE*(t)l
(9.215)
where the generalized mass matrix [M*] is [M*] = [4>]T[M +
(Cm-l)pV]m;
the coupled optimized damping matrix [Co] is [C0] = [4>f[C][
mT[K]m,
and the generalized hydrodynamic exciting force [FE (t)} is {FE*(t)} = mT
([CmPV] l^p-
+ [C]{«(01
The damping matrix [Co] is now coupled between normal modes as a consequence of the optimization procedure and must now be uncoupled by another iterative procedure (Grecco and Hudspeth, 1983).
850
Waves and Wave Forces on Coastal and Ocean Structures
Uncoupled damping matrix [C*] The original noninteraction structural damping matrix [C] was an uncoupled orthogonal damping matrix from the assumption of Rayleigh structural damping (Clough and Penzien, 1975, Chapter 13-3). The now coupled interaction modal damping matrix [Co] may again be uncoupled by minimizing the following mean square error: {€2} = ([C0]{Y] - [C*]{Y})\
(9.216a)
where [C*] = uncoupled damping matrix. Minimizing Eq. (9.216a) according to
\dC*
2
)=\
( E
C
(9.216b)
that may be solved for C*m to obtain C*mm = CQmm + £ ( 1 - 8nJ-j^, n=\
(9.216c)
\ ml
where N = number of normal modes. The optimization process for the uncoupled damping matrix [C*] is iterative and a convergence criterion is required for all normal modes such that the diagonal damping coefficient is given by (Grecco and Hudspeth, 1983) C * + 1 ) = C L + f > - 8nm)Clm ^
= C%\
(9.216d)
where the variance o\ • of the velocity of the generalized normal coordinate Ym is computed from poo c
tmYnW
= I
f2Symyn(fW
= o]Jn.
(9.2l6e)
Modal analysis Following diagonalization of the optimal damping matrix [C*], the n th normal mode of oscillation of the platform may be computed from the following modal
851
Real Ocean Waves
equation for a damped harmonic oscillator Chapter 2.5.3: Yn(t)+2^co0J„(t)
FE*(t) + co0iYn(t) = ^ - ^ , M*
n = 1,2,3,..., N, (9.217a)
where the modal damping ratio £„ for the nth normal mode may be computed from !» = ^r~ = T-nkrri < 9 - 217b > 2M>0„ 2^K*M* and the undamped natural frequency for the n th normal mode may be computed from col = ^ - .
(9.217c)
n
The Af linearly independent equations in Eq. (9.217a) may be solved in the time domain for the N normal modes by the Duhamel convolution integral (vide., Chapter 2.5.3 or Clough and Penzien, 1975) given by Y„(t)= f h(t,s)F^(s)ds, Jo
n = 1,2,3,..., N,
(9.218a)
where the convolution kernel (or unit impulse response function) h(t,s) in Eq. (2.100d) in Chapter 2.5.3 is hit, s) =
—
sin [codn(t - s)\
(9.218b)
Mn^dn
and the damped natural frequency a>dn in Eq. (2.99h) in Chapter 2.5.3 for the n th normal mode with hydrodynamic effects is a>dn=a>nhy/l-ti,
(9.218c)
where conh = natural frequency with hydrodynamic effects (vide., Eqs. (7.112 f-i) in Chapter 7.8) and where f„ = damping ratio in Eq. (9.217b) for the n th normal mode. The N normal mode solutions to Eq. (9.218a) may be computed more efficiently in the frequency domain by spectral analyses (vide., Sec. 9.3).
852
Waves and Wave Forces on Coastal and Ocean Structures
Frequency domain spectral analysis The cross-covariance function CymYn(r) from Eq. (9.17a) in Sec. 9.3 for the mth and n th normal modes of the generalized normal coordinates Ym(t) and Yn(t) is defined as TR/2
CrmYn(T)=
lim - j - f
Ym(t)Yn(t + r)dt.
(9.219)
-TR/2
The Wiener-Khinchine Fourier transform pair in Eqs. (9.15a, b) in Sec. 9.3 for a one-sided spectral density function SymY„ (f) from Eqs. (9.15b, h) in Sec. 9.3, following the substitution of Eqs. (9.218), yields oo
exp /
ilnfx
-00 OO
/
/-OO
/
hm(@i)hn(®2)CFE*FE*(r
- 02 -
Gi)d®id@2dT.
- 0 0 J —OO
(9.220) The Fourier transform in Eq. (9.3b) in Sec. 9.2 of the convolution kernel (or unit impulse response function) ht(®j) in Eq. (2.100d) in Chapter 2.5.3 is given by oo
/
hd®j) exp - (27tf®j)d®j
(9.221a)
-00
or in terms of a generalized matrix D (///,) by D(//7i) =
F z -, T(9.221b) M* [ ( 2 ^ ) 2 ( 1 - (f/ftf + 2f,-(f/ft))] The Fourier transform of the wave force cross-covariance function CFE* FE* (V) 1
is the cross-spectral density function SFE*FE*(/) forces given by r K'
sF£.FE.{f)=
m
l
n
for the generalized wave
X ; E * 5 >kn*
j=r k=k'
CijCmkSaMduJL(f) dt
+ CjCikS
u
j
+ CilCkSauj ( / ) - „ vUk \ . (9.222) auk {f) + CjCkSUjUk(f) 3t
dt
853
Real Ocean Waves
Substituting Eqs. (9.221b and 9.222) into Eq. (9.220) yields the following one-sided cross-spectral density function Sym ym (/) for the generalized normal mode response: SYmYn(f) = Vm(f/fm)T)*n(f/fn)SFE*FE*(f),
(9.223)
where the superscript asterisk * = complex conjugate Fourier transform of the convolution kernel (or unit impulse response function). Finally, the spectral density function for the j t h numbered structural nodal displacement may be computed from Eq. (9.223) and Eq. (9.213) by N
SXjXj(f)
N
= E E *J>n*JnK(f/fm)Vn(f/fn)SFE.FE.(f). m=\n=\
(9.224)
The cross-spectral density functions for the wave kinematics in Eq. (9.222) may be computed from linear wave theory by
"^
( / )
Ssuk(f)
f2 coshMzj + h)] cosh[k(zk + h)] ~ cosh2^ x exp[-i(xj - x^S^if), = ifSUjUk(f),
(9.225a) (9.225b)
"j at
S3ju_uk(f) = -ifSUjUk(f),
(9.225c)
dt
S*ul^{f) = fSUjUk(f),
(9.225d)
provided that (2TT/) 2
= gk tanh kh.
(9.225e)
Optimization of damping and drag coefficients The variance uf • (and standard deviation <Jfkfk) of the relative velocity rk(t) required in Eqs. (9.208-9.211) in order to linearize the nonlinear relative
854
Waves and Wave Forces on Coastal and Ocean Structures
motion drag force term in the modified wave force equation (WFE) or relative motion Morison equation is computed from /•OO
C
°hh = hh (0) = / Shfk (f)df, Jo
(9.226a)
where Srkrk(f) = SUkUk{f) + Skkkk{f)
- {SUk±k{f) + SkkUk(f)).
(9.226b)
At each ith iteration, the new spectral density function for the structural displacements Sj ^ ( / ) and for the cross-spectral density functions between the structural velocities and wave horizontal velocities Su %Af) and S% (f) are computed from Eqs. (9.226); and the new linearized drag force may then be computed for the next iteration. Next, in order to uncouple matrix elements in the generalized damping force between the normal modes, the generalized diagonal damping matrix is computed from the cross-covariance function for the generalized velocity of the normal coordinates CY Y (0) by Eq. (9.216e). The relative amount of coupling between normal modes is a function of the number of iteration cycles that are required to meet a specified convergence criterion. Comparison with data for a space-frame offshore platform from WPII Grecco and Hudspeth (1983) compare this linearized analysis with data from pressure transducers mounted on one of the vertical pile supports on the WPII steel-jacketed, space-frame offshore platform in the Gulf of Mexico. Both the spectral forces and the statistics of maximum quantities from Sec. 9.4 are computed. In general, the comparisons are very good even though the measured forces and maxima quantities are from strongly nonlinear hurricane generated waves (vide., Fig. 9.53). Table 9.13 lists the first four normal mode frequencies and periods without (i.e., in air) and with (i.e., hydrodynamic damping with added mass) hydrodynamic effects for the WPII space-frame offshore structure (Grecco and Hudspeth, 1983). Figures 9.55 and 9.56 illustrate the first four normal modes of vibration.
9.9. Frequency Domain Input-Output Transfer Functions Linear solutions to compute the deterministic dynamic response of coastal and ocean structures may be computed either in the time domain by the
855
Real Ocean Waves
Table 9.13. Normal frequencies and periods for the WPII space-frame offshore platform in the Gulf of Mexico (Grecco and Hudspeth, 1983). Normal mode
No. i 1 2 3 4
In air
Hydrodynamic damping with added mass
w,(rads/s) /KHz) Ti(s) a>ih (rads/s) fihQb) 1.213 8.723 8.821 14.410
0.193 1.390 1.404 2.294
V\\\\\\\
5.181 0.719 0.712 0.436
0.958 8.644 8.766 13.010
0.153 1.376 1.395 2.070
^(s) 6.556 0.727 0.717 0.481
VXSWXVN
Fig. 9.55. Normal modes 1 and 2 for the WPII space-frame offshore platform in the Gulf of Mexico (Grecco and Hudspeth, 1983).
Duhamel convolution integral from Eq. (2.97a) in Sec. 2.5.3 or in the frequency domain by the method of undetermined coefficients from Eq. (2.119) in Sec. 2.5.6. Linear systems are systems whose solutions may be added because no powers or products of the solutions are required. Solutions to non-deterministic dynamic response of coastal and ocean structures may also be computed either in the time domain by cross-covariance (-correlation) functions from Eqs. (9.17) in Sec. 9.3 or in the frequency domain by spectral density (amplitude) functions as in the space-frame offshore structure
856
Waves and Wave Forces on Coastal and Ocean Structures
\\\\\ Fig. 9.56. Normal modes 3 and 4 for the WPII space-frame offshore platform in the Gulf of Mexico (Grecco and Hudspeth, 1983).
in Sec. 9.8.2. The objective of this review is to demonstrate that convolution and correlation computations in the time domain are equivalent to multiplication computations in the frequency domain. In fact, the time domain convolution integral is required in order to demonstrate that crosscovariance (-correlation) computations are equivalent to multiplication in the frequency domain. Because the FFT algorithm introduced in Sec. 9.2 is very CPU time efficient, frequency domain computations are usually selected to compute kinematic and dynamic variables that are required in order to compute the dynamic response of coastal and ocean structures. Two formulas are reviewed in Sec. 9.2 for representing Fourier transform pairs; and the radian wave frequency co option given by Eqs. (9.2a, b) is applied in the review here. A brief comment on the similarities between convolution and crosscovariance (-correlation) operations may be appropriate here (Brigham, 1974, Chapter 4, Fig. 4-11, p. 65). The graphical comparison of these two operations given by Brigham (1974) are summarized in Table 9.14. The primary difference between the two operations that are compared in Table 9.14 is the folding {faltung) of x (—r) or, equivalently, making the mirror image of x(r) in the convolution process. The consequence of the folding or mirror imaging of the convolution kernel h(x, £) in the Duhamel convolution
857
Real Ocean Waves
Table 9.14. Comparison between convolution and correlation operations (Brigham, 1974). Correlation
Convolution
Process Operation: z(t) = Function Folding (Faltung) Displacement Ordinate by ordinate multiplication of function and displaced function
/!£oy(*)*(f-T)dr
!^000y(r)x{t
+ x)dz
X(T)
y(r)
NA JC(-T) x(t — r) shift positive r by t jc(f + T) shift negative r by t y{t) x x(t — T) y(t) x x(f + r)
integral in Eq. (2.97a) in Chapter 2.5.3 (where x = t and § = x for the comparison in Table 9.14) is that the past history of the loading / ( £ = r) is convolved with the future response of the convolution kernel and the future loading is convolved with the past response of the convolution kernel in order to determine the response at the present time x (= t in the comparison in Table 9.14.) Convolution input-output in the time domain The input-output relation for a linear system may be represented by the Duhamel convolution integral from Eq. (2.97a) in Chapter 2.5.3 by [TR/2
fout(t)=
lim
/
TR^OO
Kut(.t-r)fi„(r)dr
[TR/2
=
lim TR^OO
(9.227a)
J-TR/2
/
hout(T)fin(t-T)dz,
(9.227b)
J-TR/2
where fin(t) = an input function; fout(t) = an output function and &out(t — T) = hout{x) = the unit impulse response function or, equivalently, the convolution kernel that may be computed from an homogenous boundary value or initial value problem. The convolution operation defined in Eqs. (9.227a, b) is often abbreviated in the literature by the following asterisks *
858
Waves and Wave Forces on Coastal and Ocean Structures
product notation: foutiO = houtit - T) * fin(T) = Kutij)
* fin{f - t ) .
(9.227c,d)
Applying the Fourier transform to Eqs. (9.227a, b) with the positive sign option for the exponential term given by Eq. (9.2b) in Sec. 9.2 and letting \TR\ ->• oo gives
— [ oo fout(t)exp JinJ-
icotdt /-OO
OO
\fljt J-C /
expicotdt hout(t-r)fin(r)dt. J —OO Applying the following-00change of variables to Eq. (9.228):
(9.228)
t — T = [M, dt = d/X
and applying the definition for the complex-valued Fourier coefficient in Eq. (9.2b) with the positive sign option for the exponential term, transforms Eq. (9.228) to i
-oo roo
Fout(co) = - = l
roo roo
expico(fi + r)dfi
V2TT J- OO
hout(n)fin(r)dr
./—00
OO
/-OO
V27T ^-oo / how? (/^) exp icofidii j fin (T) exp iwxdx = Hout(co)F-OOin(co), (9.229) J— OO where Hout (co) = Fourier transform of the unit impulse response function houtit) (or> equivalently, the convolution kernel); and where Eq. (9.229) demonstrates that convolution in the time domain is equivalent to multiplication in the frequency domain. To motivate that frequency domain transfer functions for spectra may also be related to a time domain operation called cross-covariance (-correlation) or auto-covariance (-correlation), note that multiplying both sides of Eq. (9.229) by their complex conjugate quantities (denoted by a superscript asterisks •*) gives Fout{co) x F*ut(o)) = {Hout(a>) x H*ut{oo)) (Fin(a>) x F*n(co)), \Fout{co)\2 = \Hout(co)\2\Fin(co)\2,
859
Real Ocean Waves
that may be related to frequency domain multiplication of transfer functions with ocean wave spectra as demonstrated below (cf., Fig. 9.38 in Sec. 9.6). Cross-covariance and auto-covariance input-output in the time domain The following review of cross- and auto-covariance input-output operations may also be extended to cross- and auto-correlation operations by simply normalizing the dimensional variables in accordance with Eq. (9.14c) in Sec. 9.3. In order to demonstrate that computing cross-covariance functions in the time domain is equivalent to multiplication of frequency domain functions in the frequency domain analogous to Eq. (9.229), note that the cross-covariance function in Eq. (9.17a) in Sec. 9.3 may be expressed as TR -> oo by 1
rTR/2
CfinJouM') = lim — / TR^OO
1R
fin(t)fout(t pOO
OO
/
finit) / -oo oo
/
+ x)dt
J-TR/2
hout(v)fin(t
+
r-v)dtdv
J—oo poo Kutiy)
-00
I
finif)finif
+ T-
v)dtdv
J — OO
oo
/
Kutiv)CfinfiniT-v)dv,
(9.231)
-00
where the auto-covariance definition for the input function /;„(•) in Eq. (9.18a) in Sec. 9.3 has been substituted into Eq. (9.231). Applying the Fourier transform with the positive sign option for the exponential term given by Eq. (9.2a) in Sec. 9.2 to Eq. (9.231) yields 1 f°° ~7E= / C/^/^CiOexpiam/T V Lit J—oo 1
/-OO
= —7= / V 2 7 T J—oo
pOO
expicor
Kutiv)Cfintfinir-v)dvdr.
(9.232)
J—oo
Applying the following change of variables in Eq. (9.232): T — v = ii, dx = djx and the definition for the complex-valued Fourier coefficient in Eq. (9.2b) with the positive sign option for the exponential term, transforms
860
Waves and Wave Forces on Coastal and Ocean Structures
Eq. (9.232) to i
poo
G
fin,fou,(w) = -j=
poo
\
exp ieo(fi + v) /
\l Lit J—oo 1
pOO
= —=
hoUt(y)Cfinifln(n)dvdii
J—oo pOO
\
hout(v)expicovdv
Cfin>fin(iJ,)exp ico/idfi
VLTC J—oo
J—oo
= Hout(co)Gfinjin(co),
(9.233)
where G (»,•)(&>) = complex-valued, two-sided cross-spectral density function defined by the Wiener-Khinchine Fourier transform pair in Eq. (9.15b) in Sec. 9.3; and where the complex-valued result in Eq. (9.233) demonstrates that cross-correlation in the time domain is equivalent to multiplication in the frequency domain. Similarly, the auto-covariance function (9.18a) in Sec. 9.3 may be expressed as TR —> oo by 1 c
fou,,fouW
= Tlim TR^OO
oo
/
J-TR/2
^r / 1R
fout(t)fout{t + r)dt J-TR/2
poo
/
ho«,Mfin(t-$)dS
-OO J — OO
oo Kut(v)fin(t
+ T
-V)dvdt
/ -oo /oo
poo
Kut(S)d$ / -oo
J—oo
hout(v)dv
oo
fin(t-$)fin(t / /oo
T-v)dt
poo
hout($)d$ -oo
+
-oo
/
hout(v)CfinJin(v
- v - $)dv.
(9.234)
J—oo
Applying the Fourier transform to Eq. (9.234) with the positive sign option for the exponential term given by Eq. (9.2b) in Sec. 9.2 and letting \TR\ -> oo
861
Real Ocean Waves
gives
1
-/=
f°°
/
CfoutJou,(r)exVicordx
\l Lit J—oo 1
pOO
= -—= \2TV
/-OO
/
/>00
hout(%)d%
J—OO J—OO
h0Ut(v)Cfinjin(T-v)expicordvdr.
J—OO
(9.235) Applying the following change of variables to Eq. (9.235): r — v — § = s,
dx = ds
and applying the definition for the complex-valued Fourier coefficient in Eq. (9.2b) with the positive sign option for the exponential term, transforms Eq. (9.235) to oo
/ G
/0«r,/ou,(w) =
hout(%)expico%d% oo
/
poo
hout(v)expicovdv -OO
I
Cfjnjin(s)expicosds,
•/— 0 0
n (co)\2GfinJin(co), (9.236) that demonstrates thatout auto-correlation in the time domain is equivalent to multiplication in the frequency domain between the square of the modulus of a complex-valued transfer function \Hout(a>)\2 and the two-sided autospectral density function G (.,.)(&>). The frequency domain multiplication in Eq. (9.236b) may be transformed to multiplication of one-sided auto-spectral density functions by Eq. (9.15h) in Sec. 9.3 to obtain SfoulJout(co) = \nout(u)\2Sfin,fin(co),
co>0.
(9.237)
The frequency domain multiplication in Eq. (9.237) may be illustrated graphically by Fig. 9.38 in Sec. 9.6. In Fig. 9.38, interchange the target spectral density functions Snr](o)m) on the second row of graphs with the white noise spectral density functions W{com) on the first row of graphs; replace these interchanged white noise spectral density functions that are now on the second row with the squared modulus of a desired frequency domain transfer functions \Hout (co)\2 and the third row of graphs will now be product of the frequency domain multiplication in Eq. (9.237) and represent the one-sided spectral density functions required.
862
Waves and Wave Forces on Coastal and Ocean Structures
Examples of transfer functions from linear wave theory (LWT) A number of kinematic and dynamic fields from LWT illustrate the frequency domain multiplication between transfer functions and wave spectral densities.
Horizontal water particle velocity at a vertical elevation z The horizontal water particle velocity at a vertical elevation z in a linear progressive surface gravity wave is given by Eq. (4.37b) in Chapter 4.4 as u(x, z, t) =
-d$(x, z, t) ox H gkcosh.k(z + h) cos(kx — cot + ao) 2 co cosh kh
(4.37b)
provided that koh = kh tanh kh and the deep-water wave number &o = co2/g. The water surface elevation is given by Eq. (4.39b) in Chapter 4.4 as H r](x, t) = — cos(kx — cot + ao)
(4.39b)
so that Eq. (4.37b) becomes at a fixed horizontal location xo, say,
(
gk cosh k(z + h)\ — r\ (kxo -cot + ao). co cosh kh J
(9.238a)
Computing the auto-covariance function Cuu{x) by Eq. (9.234) with Eq. (9.238a) and then applying the Fourier transform, the one-sided spectral density function for the horizontal water particle velocity at a vertical elevation z may be computed from a target spectral density function for the water surface elevation Snr](co) from Table 9.5 in Sec. 9.3.1 at a fixed horizontal location {XQ,Z} according to SUu(co) = \Bu(co,z)\2 S^ico), Uu(co,z) =
gkcoshkiz + h) co
T-— cosh kh
.
co>0,
(9.238b)
,„„.,
(9.23 8c)
863
Real Ocean Waves
Some care must be exercised when evaluating Eq. (9.238c) as co ->• 0 because the transfer function does not approach zero as the radian wave frequency approaches zero. The limit as co ->- 0 represents the long wave, shallowwater approximations to the LWT kinematics and shallow-water waves are non-dispersive. Accordingly, the shallow water wave celerity C = co/k & *Jgh and the small frequency limit for the transfer function in Eq. (9.238c) is, approximately Uu(eo,z)^ — = - = J->0, h
co
co^O.
(9.238d)
Vh
Dynamic wave pressure field p(x, z, t) at vertical elevation z The dynamic wave pressure field p(x, z, t) at a vertical elevation z in a linear progressive surface gravity wave at a fixed horizontal location XQ, say, is given by Eq. (4.50f) in Chapter 4.5 as p(x0, z, t) = y
coshk(z + h)H ———— cos(kx0 -cot + or0)
coshk(z + h)\ — t](kx0 -cot + a0) cosh kh ) = yK(z)T](kx0 -cot+ ao), y
(9.239a)
where y — pg and K(z) = the pressure response function from Eq. (4.48m) in Chapter 4.5. Computing the auto-covariance function Cpp{z) by Eq. (9.234) with Eq. (9.239a) and then applying the Fourier transform, the one-sided spectral density function for the dynamic wave pressure field at a vertical elevation z may be computed from a target spectral density function for the water surface elevation Sm(co) from Table 9.5 in Sec. 9.3.1 at a fixed horizontal location {xo,z} according to I
|2
SpP(co) = \Hp(co,z)\ Sm{co), co>0, cosh k(z + h) Hp(co,z) = y c^hkh =yK(z).
(9.239b)
864
Waves and Wave Forces on Coastal and Ocean Structures
An example of an application of spectral transfer functions to compute the wave-induced hydrodynamic pressure forces/moments by the modified wave force equation (WFE) on a proto-type steel-j acketed offshore platform in 100 ft of water in the Gulf of Mexico is reviewed in Sec. 9.8.2. Dynamic wave pressure and horizontal water particle velocity at vertical elevation z Directional wave spectra in Sec. 9.3.4 may be computed from gauges that record the dynamic wave pressure p(xo, z, t) and the horizontal water particle velocity w(;co, z, t) at a fixed horizontal location xo, say. Computing the crosscovariance function Cpu (r) by Eq. (9.17a) in Sec. 9.3 with the dynamic wave pressure from Eq. (9.239a) and the horizontal water particle velocity from Eq. (9.238a) and then applying the Fourier transform, the one-sided crossspectral density function for the dynamic wave pressure field and the horizontal water particle velocity field at a vertical elevation z may be computed from a target spectral density function for the water surface elevation Sm{co) from Table 9.5 in Sec. 9.3.1 at a fixed horizontal location according to SpuUo) = H.p(a),z)Uu((0,z)Sr,r)(a>), a>>0, = (Y8Kc{Z))sr]11(a>y,
co>0.
(9.240)
9.10. Problems 9.1.
Synthesize a discrete time sequence of length N = 32 and At = 0.5 sec with the following harmonic components: IW
9.2.
Am
1m
ttm
1
4
\6At
JT/3
4
2
4 At
-2JT/3
Select an FFT algorithm and compute the complex-valued FFT coefficients of the time sequence synthesized in 9.1 above with the FFT algorithm selected. Complete Table 9.1 from Sec. 9.2 with these FFT coefficients and determine where your FFT algorithm places the normalizing constant CN defined in Eq. (9.10) in Sec. 9.2.
865
Real Ocean Waves
9.3.
The vertical z motion of a "FLIP" type spar buoy is governed by the following ordinary differential equation of motion: *-
V7
R "
h-D-H 2
dz , ,nD2 dz m d^ = iPd-yz) — - ^ where m = mass of the buoy, pd = dynamic progressive wave pressure, y = pg = specific weight of sea water, and /3 = damping coefficient. Derive the following frequency domain transfer function from the procedures given in Sec. 9.9: Fz(f) F,(f)
9.4.
where Fz(f) — Fourier coefficient for the vertical heave motion of the buoy at frequency / ; and Fn(f) = Fourier coefficient for the water surface elevation at frequency / . Assume deep-water conditions and linear wave theory. A Rayleigh distribution for wave amplitudes A is given by Eq. (9.87a) in Sec. 9.4.2. as p(A) =
A
-2cxP-[^)U(A);
a > 0.
Prove that the average of the p = \/n highest amplitudes Ap = \/n is given by Eq. (9.91a): i.e.
a
Vi^Mpj + J-^= VV-lnQ?) + ^
l
Erf(7-ln(/>))]
(^) Erfc (V-ln(p))
and that n
866
9.5.
Waves and Wave Forces on Coastal and Ocean Structures
Simulate an NSA digital simulation from an FFT algorithm by the procedures given in Sec. 9.6 from the generic two-parameter wave spectrum in Eq. (9.29c) in Sec. 9.3.1 with the following parameters: N 2048
At (sec) 0.5
mo (ft2) 25.0
a>o (rad/s) 0.5
From your NSA digital simulation, compute the following: a) Smoothed spectrum using block averaging of blocks of length iV = 512. Determine if the sampling rate of At = 0.5 sec is adequate for blocks of 512 data values. b) Assume that the errors in your smoothed spectrum are Chi-squared distributed and place confidence intervals on the 10 most energetic spectral estimates near the peak of your smoothed spectrum. c) Compute the wave height function A (t) by the Hilbert transform inEqs. (9.135) in Sec. 9.5 and superimpose the wave height function on r)(t) and fj2(t) computed by Eqs. (9.135) in Sec. 9.5 for the first block of your blocks of 512 values in a). d) Compare the distribution of your NSA time sequence simulation for the water surface elevation with the Gaussian distribution.
Bibliography
Ablowitz, M.J. (1971) Applications of slowly varying nonlinear dispersive wave theories, Studies Applied Mathematics, L(4):329-344. Ablowitz, M.J. and D.J. Benney (1970) The evolution of multi-phase modes for nonlinear dispersive waves, Studies in Applied Mathematics, XLIX(3):225-238. Abraham, R.H. and J.E. Marsden (1978) Foundation of Mechanics, Benjamin/ Cummings, Reading, MA. Abramowitz, M. and I. A. Stegun (1965) Handbook of Mathematical Functions, NBS Applied Mathematics Series 55, Reprinted by Dover Publications, Inc., New York. Airy, G.B. (1845) Tides and Waves, Encyclopedia Metropolitana, § 192: 241-396. Alber, I.E. (1978) The effects of randomness on the stability of two-dimensional surface wave trains, Proceedings of the Royal Society of London, Series A, 363:525-546. Allen, J. and D.H. Gibson (1959) Experiments on the displacement of water waves of various heights and frequencies, / Inst of Civil Engineers, London, England, 13:363-386. American Petroleum Institute (1987) Recommended Practice for Planning, Designing and Constructing Fixed Offshore Platforms, API RP2A, Washington, D.C. Anand, S. and S.L. Agarwal (1980) Field and laboratory studies for evaluating submarine pipeline frictional resistance, Proc 12th Offshore Technology Conference, 11:371-382. Andrews, L.C. (1985) Special Functions for Engineers and Applied Mathematicians, Macmillan Pub Co., New York. Apelt, C.J. and A. MacKnight (1976) Wave action on large off-shore structures, Proc 15th Conference on Coastal Engineering, Honolulu, Hawaii, 111:2228-2247. Arfken, G. (1985) Mathematical Methods for Physicists, 3rd Ed, Academic Press, New York: 189-193. Aris, R. (1962) Vectors, Tensors, and the Basic Equations of Fluid Mechanics, Prentice Hall Pub. Arnold, V.I. (1978) Mathematical Methods of Classical Mechanics. Second Edition. Springer-Verlag. 867
868
Bibliography
Ashford, R.A. and W.L. Wood (1978) Numerical integration of the motions of a tethered buoyant platform, Int. J. Numerical Methods in Engineering, 13:165-180. Atkinson, K.E. (1989) An Introduction to Numerical Analysis, 2nd Ed, John Wiley & Sons, NY:529-540. Averbeck, P.J. (1993) The boundary value problem for the rectangular wavemaker, MS Thesis, Oregon State University. Bai, K.J. (1972) A Variational Method in Potential Flows with a Free Surface, University of California at Berkeley, College of Engineering, Report NA 72-7, 137 p. Bai, K.J. (1975) Diffraction of oblique waves by an infinite cylinder, J. Fluid Mech., 68:513-535. Baliga, B.R. and R.T. Hudspeth (1981) Evaluation of sand waves in an estuary, Journal Hydraulics Div, 107(HY2):161-178. Banaugh, R.P. and W. Goldsmith (1963) Diffraction of steady acoustic waves by surfaces of arbitrary shapes, Journal of Acoustics Society of America, 35: 1590-1601. Barcilon, V. (1990) Special functions, Ch. 7, Handbook of Applied Mathematics, Second Ed., C.E. Pearson, editor, Van Nostrand Reinhold Company, New York: 344-377. Barnard, B.J.S. and W.G. Pritchard (1972) Cross-waves, Part 2, experiments, J. Fluid Mech., 55(2):245-255. Barnard, B.J.S., J.J. Mahony, and W.G. Pritchard (1977) The excitation of surface waves near a cut-off frequency. Phil. Trans. R. Soc. Lond. A 286:87-123. Batchelor, G.K. (1983) An Introduction to Fluid Dynamics, Cambridge Press: 84-87. Battjes, J.A. (1974) Surf similarity, Proc of the 14th International Conference on Coastal Engineering.466-4%$. Battjes, J.A. and G.Ph Van Vledder (1984) Verification of Kimura's theory for wave group statistics. Proc, 19th Int. Conference on Coastal Engrg.642-648. Beale, J.T. (1977) Eigenfunction expansion for objects floating in an open sea, Communications on Pure and Applied Mathematics, XXX:283-313. Becker, J.M. and J.W. Miles (1992) Progressive radial cross-waves, /. Fluid Mech., 245:29^16. Bendat, J.S. and A.G. Piersol (1980) Engineering Applications of Correlation and Spectral Analysis, Wiley-Interscience, New York, 302 p. Bendat, J.S. and A.G. Piersol (1986) Random Data: Analysis and Measurement Procedures, 2nd Ed., Wiley-Interscience, New York, 407 p. Benettin, G.L., L. Galgani and J.M. Strelcyn (1976) Lyapunov characteristic exponents for smooth dynamical systems and for Hamiltonian systems; A method for computing all of them. Part 2: Numerical application. Phys. Rev. A14: 2338-2347.
Bibliography
869
Benjamin, T.B. (1967) Instability of periodic wave trains in nonlinear dispersive systems, Proceedings of the Royal Society of London, Series A, 299:59-75. Benjamin, T.B. and J.E. Feir (1967) The disintegration of wave trains in deep water, Part 1. Theory, /. FluidMech., 27:417^130. Benjamin, T.B. and J.C. Scott (1979) Gravity-capillary waves with edge constraints, J. Fluid Mech. 92(2):241-267. Benney, D.J. (1967) The asymptotic behavior of nonlinear dispersive waves, Journal of Mathematics and Physics, LXVI(2): 115-132. Benney, D.J. (1969) Instabilities associated with forced nonlinear waves, Studies in Applied Mathematics, 60:27-41. Benney, D.J. and RG. Saffmann (1966) Nonlinear interactions of random waves in a dispersive medium, Proceedings of the Royal Society of London, Series A, 289:301-320. Benney, D.J. and A.C. Newell (1967a) The propagation of nonlinear wave envelopes, Journal of Mathematics and Physics, XLVI(2): 133-139. Benney, D.J. and A.C. Newell (1967b) Sequential time closures for interacting random waves, Journal of Mathematics and Physics, XLVI(4):363-393. Benney, D.J. and A.C. Newell (1969) Random wave closures, Studies in Applied Mathematics, XLVIII(l):29-53. Benney, D.J. and G. Roskes (1969) Wave instabilities, Studies in Applied Mathematics, 48:377-385. Benton, E.R. (1990) Ordinary differential and difference equations Ch 6, Handbook of Applied Mathematics, Second Ed., C.E. Pearson, editor, Van Nostrand Reinhold Company, New York:316-323. Benzi, R., E. Salusti and A. Sutera (1979) Variational approach to gravity waves in terms of stream-function, Journal of Physics in the Ocean, 9:619-620. Berge, VI. (1978) Math Methods of Classical Mechanics, 2nd Ed. Springer-Verlag. Berge, P., Y. Pomeau, and C. Vidal (1984) Order Within Chaos Towards a Deterministic Approach to Turbulence, John Wiley & Sons. Berhault, C. (1980) An integro-variational method for interior and exterior free surface flow problems, Applied Ocean Research, 2(l):33-38. Bettess, P. and O.C. Zienkiewicz (1977) Diffraction and refraction of surface waves using finite and infinite elements, Int. Jour. For Numerical Methods in Engineering, 11:1271-1290. Bharuchua-Reid, A.T. (1972) Random Integral Equations, New York, Academic Press. Biesel, R and R Sequet (1953) Laboratory Wave Generating Apparatus, St. Anthony Falls Hydraulic Lab, Univ. of Minnesota, Report No. 39, Minneapolis, MN. Bijker, E.W., J.P. Th. Kalkwijk, and T. Pieters (1974) Mass transport in gravity waves on a sloping bottom, Proc 14th Conf Coastal Engrg, Copenhagen, Denmark: 447-465.
870
Bibliography
Bily, M. and J. Bukoveczky (1976) Digital simulation of environmental processes with respect to fatigue, J. Sound and Vibration, 49(4):551-568. Birkhoff, G., J. Bona, and J. Kampe de Feriet (1973) Statistically well-set Cauchy problems, in Probabilistic Methods in Applied Mathematics, ed. by A.T. Bharuchua-Reid, New York, Academic Press, 3:2-121. Black, J.L. (1975a) Wave forces on large bodies, a survey of numerical models, Proc Civil Engineering in the Oceans III Conference, University of Delaware: 924-933. Black, J.L. (1975b) Wave forces on vertical axisymmetric bodies, /. Fluid Mech., 67(2):369-376. Black, J.L. and C.C. Mei (1969) Scattering and radiation of surface waves by rectangular obstacles in water of finite depth, J. Fluid Mechs., 38:499-511. Black, J.L., C.C. Mei, and M.C.G. Bray (1971) Radiation and scattering of water waves by rigid bodies, /. Fluid Mech., 46:151-164. Blackman, R.B. and J.W. Tukey (1959) The Measurement of Power Spectra, Dover Publications Inc., New York, N.Y., 190 p. Booij, N. (1978) Refraction and diffraction of water waves in regions with variable depth and current, Manuscript, Delft University of Technology. Boreel, L.J. (1974) Wave action on large offshore structures, Proceedings of the Institution of Civil Engineers Conference on Offshore Structure, London, England. Borgman, L.E. (1965) The spectral density for ocean wave forces, Hyd. Eng. Lab Report HEL 9-8, Berkeley, CA. Borgman, L.E. (1969a) Directional spectral models for design use, Proc. Offshore Tech Conf, Houston, Texas, OTC Paper No. 1069:721-746. Borgman, L.E. (1969b) Ocean wave simulation for engineering design, Journal of Waterways and Harbors Division, 95(WW4):557-583. Borgman, L.E. (1972a) Confidence intervals for ocean wave spectra, Proc 13th Int. Conference on Coastal Engineering I(10):237-250. Borgman, L.E. (1972b) Statistical models for ocean waves and wave forces, Advances in Hydroscience, 8, New York: Academic Press. Borgman, L.E. (1973) Computer simulation of multivariate normal vectors applied to the reliability of spectral-based calculations, MIMEO No. 2007, University of Wyoming. Borgman, L.E. (1982) Techniques for computer simulation of ocean waves. Topics in Ocean Physics. A.R. Osborne and P.M. Rizzoli ed., North-Hollan Pub Co: 387-417. Borgman, L.E., W.A. Bartel, and D.R. Shields (1993) SIMBAT Theoretical Manual, TN R-941. Naval Civil Engineering Laboratory, Port Hueneme, CA. Bowers, E.C. (1977) Harbor resonance due to set-down beneath wave groups, J. Fluid Mech., 79:71-92. Bowline, CM., R.T. Hudspeth and R.B. Guenther (1999) Are cross waves chaotic? Applicable Analyses, 72(3^l):287-326.
Bibliography
871
Box, E.P. and G.M. Jenkins (1976) Time Series Analysis: Forecasting and Control, 2nd edition, Holden-Day, Inc., San Francisco, California. Boyce, W.P. (1968) Random eigenvalue problems, Probabilistic Methods in Applied Mathematics, ed. by A.T. Bharuchua-Reid, Academic Press, NY:75-198. Bracewell, R.N. (1986) The Fourier Transform and Its Applications, 2nd Ed., McGraw-Hill Book Co., 474 p. Brebner, A. and J.I. Collins (1961) The effect on mass transport of the onset of turbulence at the bed under periodic gravity waves, American Society of Mechanical Engineers, Engineering Institute of Canada Hydraulics No. 61-EIC-8, Montreal, Canada. Brebner, A., J.A. Askew, and S.W Law (1966) The effect of roughness on the mass-transport of progressive gravity waves, Proc 10th Int. Conf on Coastal Engineering, Tokyo, Japan, 12:175-184. Bretschneider, C.L. (1958) Revisions in wave forecasting; deep and shallow water, Proc of the 6th Conf on Coastal Engineering Chapter 3, Houston, Texas: 30-67. Bretschneider, C.L. (1959) Wave variability and wave spectra for wind-generated waves, TM 118 U.S. Army Corps of Engineers, Beach Erosion Board. Bretschneider, C.L. (1961) A one-dimensional gravity wave spectrum, Ocean Wave Spectra, Prentice-Hall, Englewood Cliffs, N.J.:41-65. Bretschneider, C.L. (1966) Wave generation by wind, deep and shallow water, Estuary and Coastline Hydrodynamics, A.T. Ippen, Editor, 3:133-196. Bretschneider, C.L. (1968) Significant waves and wave spectrum, Ocean Industry: AQ-A6.
Bretschneider, C.L. (1969) Wave forecasting, Handbook of Ocean and Underwater Engineering, Myers, J.J. et al., Eds, McGraw-Hill Book Co. Inc., NY, 11:97-109. Brigham, E.O. (1974) The Fast Fourier Transform, Prentice-Hall Pub. Brillinger, D.R. (1975) Time Series Data Analysis and Theory, Holt, Rinehart and Winston, Inc., New York. Brillinger, D.R. and M. Rosenblatt (1967a) Asymptotic theory of estimates of kth order spectra, Spectral Analysis of Time Series, Ed. B. Harris, John Wiley and Sons:153-188. Brillinger, D.R. and M. Rosenblatt (1967b) Computation and interpretation of kth order spectra, Spectral Analysis of Time Series, Ed. B. Harris, John Wiley and Sons: 189-232. Brown, L. J. (1967) Methods for the analysis of non-stationary time series with applications to oceanography, TR HEL 16-3, Hyd Engrg Lab, Univ of California. Buhr Hansen, J.B. and LA. Svendsen (1974) Laboratory generation of waves of constant form, Proc. 14th Int. Conf. Coastal Engrg, Copenhagen, Denmark: 321-339.
872
Bibliography
Bullock, G.N. and I. Short (1982) Characteristics of laboratory generated regular waves, Report 82/167, Dept of Civil Engr, University of Salford, Salford, England. Butkov, E. (1968) Mathematical Physics, Addison-Wesley Pub Co, 537 p. Cacko, J. and M. Bily (1979a) Simulation of non-stationary stochastic process with respect to its probability density function, J. Sound and Vibration, 62(2):293-299. Cacko, J. and M. Bily (1979b) Simulation of non-stationary stochastic process with respect to its power spectral density, / . Sound and Vibration, 66(2): 277-284. Caligny, A.F.H. de C.R. (1878) Experiences sur les mouvements des molecules liquides des ondes courantes, considerees dans leur mode d'action sur la marche des navires, C. R. Acad deScl, Paris, 87:1019-1023. Carnahan, B., H.A. Luther and J.O. Wilkes (1965) Applied Numerical Methods, John Wiley and Sons, New York, 604 p. Carrier, G.F., M. Krook and C.E. Pearson (1966) Functions of a Complex Variable, Theory and Technique, McGraw-Hill Book Co. Inc., New York. Carrier, G.F. and C.E. Pearson (1968) Ordinary Differential Equations, Ginn (Blaisdell), Boston, MA, 229 p. Carrier, G.F. and C.E. Pearson (1976) Partial Differential Equations, Academic Press, New York, NY, 320 p. Carry, C. (1953) Clapotispartiel, La Houille Blanche, No. 4, Aout:482^194. Carter, T.G., PL-F. Liu, and C.C. Mei (1973) Mass transport by waves and offshore sand bedforms, J. Waterways, Harbors and Coastal Engrg, 99(WW2): 165-184. Cartwright, D.E. and M.S. Longuet-Higgins (1956) The statistical distribution of the maxima of a random function, Proc Roy Soc London, Series A, 237:212-232. Case, K.M. and W.C. Parkinson (1957) Damping of surface waves in an incompressible liquid, J. Fluid Mechs, 2(2): 172-184. Chakrabarti, S.K. (1972) Nonlinear wave forces on vertical cylinder, Journal of Hydraulics Division, 98(HY11):1895-1909. Chakrabarti, S.K. (1973) Wave forces on submerged objects of symmetry, Journal of Waterways, Harbors and Coastal Engineering Division, 99(WW2):147-164. Chakrabarti, S.K. (1975) Second-order wave force on large vertical cylinder, Journal of Waterways, Harbors and Coastal Engineering Division, 101(WW3): 311-317. Chakrabarti, S.K. (1977a) Discussion of Forces due to nonlinear waves on vertical cylinder, by H. Raman and P. Venkatanarasaiah, J. Waterway, Port, Coastal and Ocean Engrg, 103(WW3):404-^06. Chakrabarti, S.K. (1977b) Discussion of Nonlinear wave forces on large offshore structures, by M. de St. Q. Isaacson, J. Waterway, Port, Coastal and Ocean Engrg, 103(WW4):568-570.
Bibliography
873
Chakrabarti, S.K. (1978) Comments on second-order wave effects on large-diameter vertical cylinder, Journal of Ship Research, 22(4):266-268. Chakrabarti, S.K. (1987) Hydrodynamics of Offshore Structures, SpringerVerlag, N.Y. Chakrabarti, S.K. (1994) Offshore Structure Modeling, Vol. 9, World Scientific Pub, New Jersey, 470 p. Chakrabarti, S.K. and W.A. Tam (1973) Gross and local wave loads on a large vertical cylinder - theory and experiment, Proc Offshore Tech Conf Houston, Texas. Chakrabarti, S.K. and R.A. Naftzger (1976) Wave interaction with a submerged open-bottom structure, Proc Offshore Tech Conf, Houston, Texas, 11:109-123. Chang, M.S. (1969) Mass transport in deep-water long-crested random gravity waves, Journal of Geophysical Research, 74(6): 1515-1536. Chang, M.S. and RC. Pien (1976) Velocity potentials of submerged bodies near a free-surface application to wave excited forces and motions, Proceedings of the 11th Symposium on Naval Hydrodynamics, London, England. Chaplin, J.R. (1980) Developments of stream-function-wave theory, Coastal Engineering, 3:179-205. Chertock, G. (1953) The flexural response of a submerged solid to a pulsating gas bubble, Journal of Applied Physics, 24(2): 192-197. Chertock, G. (1962) General reciprocity relation, Journal of the Acoustical Society of America, 34(A):989. Chertock, G. (1964) Sound radiation from vibrating surfaces, Journal of the Acoustical Society of America, 36(7):1305—1313. Chertock, G. (1970) Solutions for sound radiation problems by integral equations at the critical wave numbers, Journal of the Acoustical Society of America, 47:387-388. Chester, W (1978) A general theory of resonance between weakly coupled oscillatory systems, Institute of Mathematics and Its Applications, 22:383-400. Chu, V.H. and C.C. Mei (1970) On slowly varying Stokes waves, J. Fluid Mech., 41:873-887. Chu, V.H. and C.C. Mei (1971) The nonlinear evolution of Stokes waves on deep water, / . Fluid Mech., 47(2):337-351. Chung, J.S.(1975)A note on the two force equations for afloatingplatform, Journal of Hydronautics, 9(4):170-172. Chung, J.S. (1976) Motion of a floating structure in water of uniform depth, Journal of Hydronautics, 10(3):65-73. Churchill, R.V. (1960) Complex Variables and Application, 2nd Ed, McGraw-Hill Book Co., Inc., New York. Churchill, R.V, J.W. Brown and R.F. Verhey (1976) Complex Variables and Application, 3rd Edition, McGraw-Hill Book Co, Inc., New York.
874
Bibliography
Clifton, H.E., R.E. Hunter and L.R Phillips (1971) Depositional structures and processes in the non-barred high-energy nearshore, J. Sedimentary Petrology, 41(3):651-670. Clough, R. W. and J. Penzien (1975) Dynamics of Structures, McGraw-Hill Co, N. Y. Cokelet, E.D. (1977) Steep gravity waves in water of arbitrary uniform depth, Philosophical Transactions of the Royal Society ofLondon, Series A, 286:183-230. Cole, R.J. and D.C. Pack (1975) Some complementary bivariational principles for linear integral equations of Fredholm type, Proc Royal Soc of London, Series A, 347:239-252. Collin, R.E. (1960) Field Theory of Guided Waves, McGraw-Hill Book Company, N.Y. Collins, J.I. (1963) Inception of turbulence at the bed under periodic gravity waves, Journal of Geophysical Research, 68:6007-6014. Copely, L.G. (1968) Fundamental results covering integral representations in acoustic radiation, Journal of Acoustical Society of America, 44:28-32. Count, B.M. (1978) On the dynamics of wave-power devices, Proceedings of the Royal Society of London, Series A, 363:559-579. Courant, R. and D. Hilbert (1966) Methods of Mathematical Physics, I, John-Wiley and Sons (Interscience), New York, NY. Courant, R. and D. Hilbert (1962) Methods of Mathematical Physics, II, John-Wiley and Sons (Interscience), New York, NY. Cox, C.S. and W.H. Munk (1954a) Measurement of the roughness of the sea surface from photographs of the sun's glitter, J. Optical Soc. Am., 44(ll):838-850. Cox, C.S. and W.H. Munk (1954b) Statistics of the sea surface derived from sun glitter, / . Marine Research, 13(2): 198-227. Craik, A.D.D. (1978) Evolution in space and time of resonant wave triads, II. A class of exact solutions, Proc Royal Society of London, Series A, 363:257-269. Craik, A.D.D. (1982) The drift velocity of water waves, /. Fluid Mechs., 116:187-205. Craik, A.D.D. (1985) Wave Interactions and Fluid Flows, Cambridge Univ Press, NY. Craik, A.D.D. and J. A. Adam (1978) Evolution in space and time of resonant wave triads, I. The pump-wave approximation, Proc Royal Soc London, A363:243-255. Crandall, S.H., D.C. Karnopp, E.F. Kurtz, Jr. and D.C. Pridmore-Brown (1968) Dynamics of Mechanical and Electromechanical Systems, McGraw-Hill Book Co., NY. Crawford, D.R., P.G. Saffmann, and H.C. Yuen (1980) Evolution of a random inhomogeneous field on nonlinear deep water waves, Journal of Wave Motion, 2:1-16. Crawford, D.R., B.M. Lake, P.G. Saffman and H.C. Yuen (1981) Stability of weakly nonlinear deep water waves in two and three dimensions, /. Fluid Mechs, 105:177-191.
Bibliography
875
Curie, N. (1979) Analysis of certain slowly converging series, Journal of the Institute of Mathematical Applications, 23:265-275. Daily, J.W. and D.R.F. Harleman (1966) Fluid Dynamics, Addison-Wesley Pub., 454 p. Dalrymple, R.A. (1974) Models for non-linear water waves on shear currents, Proceedings of the 6th Offshore Technology Conference.843-856. Dalrymple, R.A. (1976) Wave-induced mass transport in water waves, Journal of Waterways, Harbors and Coastal Engineering Division, 102(WW2): 255-264. Dalrymple, R.A. (1989) Directional wavemaker theory with side wall reflections, Journal Hydraulic Research, 27(1): 23-34. Dalrymple, R.A. and R.G. Dean (1972) The spiral wavemaker for littoral drift studies, Proc 13th Confon Coastal Engineering :689-705. Dankowicz, H. (1996a) Looking for chaos. An extension and alternative to Melnikov method. International Journal of Bifurcation and Chaos. 6(3):485—496. Dankowicz, H. (1996b) Analytical expressions for stable and unstable manifolds in higher degree of freedom Hamiltonian systems. International Journal of Bifurcation and Chaos. 6(11): 1997-2013. Darbyshire, J. (1959) A further investigation of wind generated waves, Deutsche Hydrographische Zeitscrift, 1:1-13. Daugaard, E. (1972) Generation of Regular Waves in the Laboratory, doctoral dissertation. Ins Hyd Engr, Technical University of Denmark, Copenhagen, Denmark. Davenport, W.B. and W.L. Root (1963) Random Signals and Noise, McGraw-Hill BookCo.,N.Y.:50-52. Davey, A. (1972) The propagation of a weak nonlinear wave, /. Fluid Mechs., 53:769-781. Davis, A.M.J. (1976) A relation between the radiation and scattering of surface waves by axisymmetric bodies, J. Fluid Mechs., 76:85-88. Davis, D.A. and J.B. Ciani (1976) Wave forces on submerged pipeline - a review with design aids, TR R844, Civil Engrg Lab, Port Hueneme, CA, 55 p. Dean, R.G. (1965) Stream function representation of nonlinear ocean waves, Journal of Geophysical Research, 70(18):4561^1572. Dean, R.G. (1972) Application of stream function wave theory to offshore design problems, Proc 4th Offshore Technology Conference, OTC Paper No. 1613. Dean, R.G. (1974) Evaluation and Development of Water Wave Theories for Engineering Application, U.S. Army Corps of Engineers, Coastal Engineering Research Center, Ft. Belvoir, Virginia, Special Report No. 1, 534 p. Dean, R.G. (1976) Methodology for evaluating suitability of wave and wave force data for determining drag and inertia coefficients, Proceedings of BOSS '76: 40-64.
876
Bibliography
Dean, R.G. and F. Ursell (1959) Interaction of a Fixed, Semi-Immersed Circular Cylinder with a Train of Surface Waves, T.R. #37, Massachusetts Inst. Tech. Dean, R.G. and D.R.F. Harleman (1966) Interaction of structures and waves, Estuary and Coastline Hydrodynamics, A.T. Ippen, Ed., McGraw-Hill Book Co.: 341-403. Dean, R.G. and RM. Aagaard (1970) Wave forces: data analysis and engineering calculation method, Journal of Petroleum Technology, 22(3):368-375. Dean, R.G., R.A. Dalrymple and R.T. Hudspeth (1974) Analysis of Wave Project I and Wave Project II Data, Coastal and Oceanographic Engineering Lab, Univ. Fla., 75 p. Dean, R.G. and R.A. Dalrymple (1991) Water Wave Mechanics for Engineers and Scientists, Prentice-Hall, Englewood Cliffs, New Jersey. De Best, A. and E.W Bijker (1971) Scouring of a sand bed in front of a vertical breakwater, Communications on Hydraulics, Delft Univ Tech, Report 71-1. Delves, L.M. (1977) A fast method for the solution of Fredholm integral equations, Journal of the Institute of Mathematical Applications, 20:173-182. DeSilva, S.J., R.B. Guenther and R.T. Hudspeth (1996) Irregular points in 2-D free surface flows with surface tension for the wavemaker boundary value problem, Applied Ocean Res., 18:293-302. Desloge, E.A. (1982) Classical Mechanics. Volume 2. John Wiley & Sons. Diez, J.J. and J.R. Medina (1983) Problemas associados a la utilizacion de los simulladores numeicos de oleaje, Revistas de Obras Publicas, Spain:89-109. Donelan, M. and W.J. Pierson (1983) The sampling variability of estimates of spectra of wind-generated gravity waves, Journal of Geophysical Research, 88(C7):4381-4392. Dore, B.D. (1970) Mass transport in layered fluid systems, J Fluid Mechs. 40(1):113-116. Dore, B.D. (1973) On steady particle motion in circular gravity waves, Acta Mechanica, 7:227-249. Dore, B.D. (1974) The mass transport velocity due to interacting wave trains, Acta Meccanica, 9(3): 172-178. Dore, B.D. (1975) Wave-induced vorticity in free-surface boundary layers: Application to mass transport in edge waves, J. Fluid Mechs., 70(2):257-266. Dore, B.D. (1976) Double boundary layers in standing surface waves, Pure and Applied Geophysics, 114:629-637. Dore, B.D. (1977) On mass transport velocity due to progressive waves, Quarterly Journal of Mechanics and Applied Mathematics, 30:157-173. Dore, B.D. (1978) Some effects of the air-water interface on gravity waves, Geophysics, Astrophysics Fluid Dynamics, 10:215-230. Dore, B.D. (1982) On the second approximation to mass transport in the bottom boundary layer, Coastal Engineering, 6:93-120.
Bibliography
877
Dugundji, J. (1958) Envelopes and Pre-Envelopes of Real Waveforms, IRE Transactions of Information Theory, March 1958:53-57. Eagleson, RS. (1959) The Damping of Oscillatory Waves by Laminar Boundary Layers. TR. 32, M.I.T. Hyd Lab, Cambridge, MA and Beach Erosion Board TM117. Eagleson, RS. and R.G. Dean (1959) Wave-induced motion of bottom sediment particles, Journal of Hydraulics Division, HY10, 85(2):53-79. Eagleson, RS. and R.G. Dean (1966) Small amplitude wave theory, Estuary and Coastline Hydrodynamics, A.T Ippen, Ed., McGraw-Hill Book Co.: 1-92. Elgar, S., Guza, R.T. and Seymour, R.J. (1984) Groups of Waves in Shallow Water, Journal of Geophysical Research, 89(C3):3623-3634. Elgar, S., Guza, R.T. and Seymour, R.J. (1985) Wave Group Statistics from Numerical Simulations of a Random Sea, Applied Ocean Research, 7(2):93-96. Erdelyi, A. (1956) Asymptotic Expansions, Dover Publications, Inc., New York. Eshbach, O.W (197'4) Handbook of Engineering Fundamentals, Third edition, John Wiley & Sons, New York, 480 p. Evans, D.V. (1976) A theory of wave power absorption by oscillating bodies, /. Fluid Mechs., 77:1-25. Evans, D.V. (1981) Power from water waves, Annual Review Fluid Mechs, 13: 157-187. Fadel, S. (1998) Application of the generalized Melnikov method to weakly damped parametrically excited cross waves with surface tension, Ph.D. Dissertation, Oregon State University. Falnes, J. and K. Budal (1978) Wave-power conversion by point absorbers, Norwegian Maritime Research, 6(4):2-l 1. Faltinsen, O.M. and F.C. Michelsen (1974) Motions of large structures in waves at zero Froude number, Proc Int Sym Dynamics of Marine Vehicles and Structures in Waves, Boss '74, London:91-106. Faltinsen, O.M. and A.E. Loken (1979) Slow drift oscillations of a ship in irregular waves, Applied Ocean Research l(l):21-31. Fassardi, CD. (1993) Analysis of Wave Groups Effects on Rubble Mound Breakwater Stability, MSOE thesis, Oregon State University. Fenton, J.D. (1978) Wave forces on vertical bodies of revolution, J. Fluid Mechs, 85:241-255. Fenton, J.D. (1985) A fifth-order Stokes theory for steady waves, Journal of Waterway, Port, Coastal and Ocean Engineering, 111(2):216-234. Fenton, J.D. (1988) The numerical solution to steady water wave problems, Computers and Geophysics, 14(3):357-368. Flick, R.E. and R.T. Guza (1980) Paddle generated waves in laboratory channels. Journal of Waterway, Port, Coastal and Ocean Division, 106(WWl):79-97.
878
Bibliography
Fontanet, P. (1961) Theorie de la generation de la houle cylindreque par un batteur plan, La Houille Blanche, 16(1):3—31. Foster, E.T. (1970) Model for nonlinear dynamics of offshore towers, Journal of Engineering Mechanics, 96(l):41-67. Frank, W. (1967) Oscillation of Cylinders In or Below the Free Surface of Deep Fluids, Naval Ship Research Development Center, Report 2375, 40 p. Friedman, B. (1956) Principles and Techniques of Applied Mathematics, John Wiley & Sons, Inc. New York, 158 p. (Reprinted by Dover Publications, 1990). Frisch, U. (1968) Wave propagation in random media, Probabilistic Methods in Applied Mathematics, ed. by A.T. Bharuchua-Reid, Academic Press:75-198. Funke, E.R. (1974) Random wave signal generator by minicomputer. Proc 14th International Conf on Coastal Engrg, Chap. 19, Copenhagen, Denmark, 1:353-371. Funke, E.R. (1978a) SPLASH-program for synthesis of episodic waves in a laboratory wave flume, HLR Report LTR-HY 65, National Research Council of Canada, Ottawa, Canada. Funke, E.R. (1978b) On the Synthesis of Realistic Sea States in a Laboratory Flume, HLR Report LTR-HY 66, National Research Council of Canada, Ottawa, Canada. Funke, E.R. (1978c) On the Meaning of Phase Spectra in Fourier Transform of Random Wave Trains, HLR Report LTR-HY 68, NRC of Canada, Ottawa, Canada. Funke, E.R. and E.P.D. Mansard (1979) Synthesis of Realistic Sea States in a Laboratory Flume, HLR Report LTR-HY-66, NRC of Canada, Ottawa, Canada. Funke, E.R. and E.P.D. Mansard (1983) The NRCC Random Wave Generation Package, HLR Report TR-HY 002, Div of Hydraulics Laboratory, Ottawa, Canada. Gallagher, B. (1971) Generation of surf beat by nonlinear wave interactions, /. Fluid Mechs, 49:21-31. Galvin, C.J. (1964) Wave-height Prediction for Wave Generators in Shallow Water, Tech. Memo 4, U.S. Army, Coastal Engineering Research Center. Galvin, C.J. (1968) Breaker type classifications of three laboratory beaches, Journal of Geophysical Research, 73(12):3651-3659. Garabedian, P.R. (1964) Partial Differential Equations, Wiley, Inc., New York. Garrett, C.J.R. (1970) On cross-waves, /. Fluid Mechs, 41(4):837-849. Garrett, C.J.R. (1971) Wave forces on a circular dock, /. Fluid Mechs., 46:129-139. Garrison, C.J. (1973) Discussion of Nonlinear wave forces on vertical cylinder, by S.K. Chakrabarti, Journal of Hydraulics Division, 99(HY7):1196-1198. Garrison, C.J. (1974) Hydrodynamics of large objects in the sea, Part I: Hydrodynamic analysis, Journal of Hydronautics, 8(1):5—12. Garrison, C.J. (1975) Hydrodynamics of large objects in the sea, Part II: Motion of free-floating bodies, Journal of Hydronautics, 9(2):58-63.
Bibliography
879
Garrison, C.J. (1976) Second-order theory of wave/structure interaction, Abstracts 15th Coastal Engineering Conference, Honolulu, Hawaii: 180-181. Garrison, C.J. (1978a) Discussion of Nonlinear wave forces on large offshore structures, by M. Isaacson, /. Waterway, Port, Coastal and Ocean Engrg, 104(WWl):93-94. Garrison, C.J. (1978b) Hydrodynamic loading of large offshore structures, threedimensional source distribution methods, Numerical Methods of Offshore Engrg, Eds. O.C. Zienkiewicz, R.W. Lewis, and K.G. Stagg, Wiley, Chichester, England: 97-140. Garrison, C.J. (1980) A review of drag and inertia forces on circular cylinders, Proceedings of the 12th Offshore Technology Conference, Houston, Texas, 11:205-218. Garrison, C.J. (1982) Forces on semi-submerged structures, Proc. Ocean Structural Dynamics Symp '82, Oregon State Univ., Corvallis, OR, 28-64. Garrison, C.J. (1990) Drag and inertia forces on circular cylinders in harmonic flow, Journal of Waterway, Harbors, and Coastal Engineering Division, 116(2): 169-190. Garrison, C.J. and V. Rao Seetharama (1971) Interaction of waves with submerged objects, / Waterway, Harbors, and Coastal Engineering Division, 97(WW2):259-277. Garrison, C.J. and RY. Chow (1972) Wave forces on submerged bodies, Journal of Waterways and Harbors Division, 98(WW3):375-392. Garrison, C.J., A. Torum, C. Iverson, S. Leivseth and C.C. Ebbesmeyer (1974) Wave forces on large volume structure - a comparison between theory and model tests, Proceedings of the Offshore Technology Conference, Houston, TX. Garrison, C.J. and G.W. Smith (1977a) The Second-Order Solution for Wave Interaction With a Submerged Cylinder, Naval PG School, Monterey, CA. Garrison, C.J. and R. Stacey (1977b) Wave loads on North Sea gravity platforms: a comparison of theory and experiment, Proceedings of the Offshore Technology Conference, Houston, TX, 1:513-524. Garrison, C.J., J.B. Field and M.D. May (1977c) Drag and inertia forces on cylinder in periodic flow, Journal of Waterways and Harbors Division, 103(WW2): 193-204. Gaster, M. and J.B. Roberts (1977) The spectral analysis of randomly sampled records by a direct transform, Proc of the Royal Society of London, Series A, 354:27-58. Gilbert, G., D.M. Thomson and A.J. Brewer (1971) Design curves for regular and random wave generators, Journal of Hydraulic Research, 9(2): 163-196. Gilbert, G., D.M. Thomson and A. J. Brewer (1972) Closure to design curves forregular and random wave generators, Journal of Hydraulic Research, 10(2):233-236.
880
Bibliography
Glaz, H.M. (1981) Statistical behavior and coherent structures in two-dimensional inviscid turbulence, SIAM Journal of Applied Mathematics, 41(3):459—479. Glowinski, R. (1984) Numerical Methods for Nonlinear Variational Problems, Springer-Verlag, 493 p. Goda, Y. (1967) Traveling secondary wave crests in wave channels, Report of the Port and Harbour Research Institute, Ministry of Transport, Japan, 13(32). Goda, Y. (1970) Numerical experiments on wave statistics with spectral simulation, Report of the Port and Harbour Research Institute, Ministry of Transport, Japan, 9(3). Goda, Y. (1977) Numerical experiments on statistical variability of ocean waves. Report of the Port and Harbour Research Institute, Ministry of Transport, Japan, 16(2):3-26. Goda, Y (1981) Wave simulation for the examination of directional resolution, ASCE andECOREIntSym Directional Wave Spectra Applications '81, Berkeley, CA. Goda, Y (1983) Analysis of wave grouping and spectra of long-travelled swell, Report of the Port and Harbour Research Institute, 22(1), 56 p. Goda, Y (1985) Random Seas and Design of Maritime Structures, Univ. Tokyo Press, 223 p. Goda, Y (1998) An overview of coastal engineering with emphasis on random wave approach, Coastal Engineering Journal, 40(1): 1-21. Goda, Y. and Y Abe (1968) Apparent coefficient of partial reflection of finite amplitude waves, Report of the Port and Harbour Res Ins, Ministry of Transport, Japan, 7(3). Goda, Y and Y Suzuki (1975) Computation of refraction and diffraction of sea waves with Mitsuyasus's directional spectrum, TN Port and Harbour Res Inst, 230,45p. Goda, Y and Y Suzuki (1976) Estimation of Incident and Reflected Waves in Random Wave Experiments. Proc. 15th Int. Conf on Coastal Engrg:. 828-845. Goldstein, H. (1980) Classical Mechanics, 2nd Edition, Addison-Wesley. Gradshteyn, I S . and I.M. Ryzhik (1980) Table of Integrals, Series, and Products. Academic Press. Graham, D.S. and A.J. Mehta (1981) Burial design criteria for tidal flow crossings, Journal of Transportation Engineering Division, 107(TE2):227-242. Gray, E.P. (1978) Scattering of a surface wave by a submerged sphere, Journal of Engineering Mathematics, 12(1): 15-44. Grecco, M. and R.T. Hudspeth (1983) Stochastic Response of a Prototype Ocean Platform, Journal of Structural Engineering, 109(5)dll9-1138. Greenspan, H.P. (1968) The Theory of Rotating Fluids, Cambridge Univ Press, Cambridge, England. Grimshaw, R. (1981) Mean flows generated by a progressing water wave packet, Journal of Australian Mathematical Society, B(22):318-347.
Bibliography
881
Guckenheimer, J. and P. Holmes (1983) Nonlinear Oscillations, Dynamical Systems, and Bifurcations of Vector Fields. Springer-Verlag. Guenther, R.B. and Schwerdtfeger (1985) Herglotz. Vorlesungen uber die Mechanik der Kontinua. BSB B.G. Teubner Verlagsgesellschaft, Leipzig. Guenther, R.B. and J. Lee (1996) Partial Differential Equations of Mathematical Physics and Integral Equations, Dover Publications, Inc, New York, 562 p. Guenther, R.B., CM. Guenther and J.A. Gottsch (1997) The Herglotz lectures on contact transformations and Hamiltonian systems, Nonlinear Analysis, Vol 1, Juliusz Schauder Ctr for Nonlinear Studies, Nicholas Copernicus University, Torun, Poland. Hansen, A.G. (1964) Similarity Analyses of Boundary Value Problems in Engineering, Prentice-Hall, Englewood Cliffs, New Jersey. Harms, V.W. (1976) Diffraction of Water Waves by Cylindrical Structures of Arbitrary Shape, University of California at Berkeley, Ph.D. thesis. Hasimoto, H. and H. Ono (1972) Nonlinear modulation of gravity waves, Journal of Physical Society of Japan, 33:805-811. Haskind, J.D. (1962) The exciting forces and wetting of ships in waves, (in Russian) Isvestra Akademic Nauk SSSR, Otdelenic Teknicheskikh Nauk, No. 7, 1957, and David Taylor Model Basin Translation, No. 30, Washington, D.C. Hasselmann, K. (1962) On the nonlinear energy transfer in a gravity-wave spectrum, Part 1, general theory, / . FluidMechs, 12:481-500. Hasselmann, K. (1963a) On the nonlinear energy transfer in a gravity-wave spectrum, Part 2, conservation theorems, wave-particle correspondence, irreversibility, / . Fluid Mechs, 15:273-281. Hasselmann, K. (1963b) On the nonlinear energy transfer in a gravity-wave spectrum, Part 3, computation of the energy flux and swell-sea interaction for a Neumann spectrum, J. Fluid Mechs, 15:393-398. Hasselmann, K. (1963c) On the nonlinear energy transfer in a wave spectrum, Ocean Wave Spectra, Englewood Cliffs, New Jersey, Prentice-Hall: 191-197. Hasselmann, K. (1963d) A statistical analysis of the generation of microseisms, Review of Geophysics, 1(2):177-210. Hasselmann, K. (1966a) Feynman diagrams and interaction rules of wave-wave scattering processes, Review of Geophysics, 4(1): 1-32. Hasselmann, K. (1966b) On nonlinear ship motions in irregular waves, /. Ship Res.: 64-68. Hasselmann, K. (1967) Nonlinear interactions treated by the methods of theoretical physics (with application to the generation of waves by wind), Proceedings of the Royal Society of London, Series A, 299:77-100. Hasselmann, K., W. Munk and G. MacDonald (1963) Bispectra of ocean waves, Time Series Analysis, ed. by M. Rosenblatt, John Wiley and Sons, Inc., NY: 125-140.
882
Bibliography
Hasselmann, K., T.P. Barnett, E. Bouws, H. Carlson, D.E. Cartwirght, K. Enke, J.A. Ewing, H. Gienapp, D.E. Hasselmann, P. Kruseman, A. Meerburg, P. Muller, D.J. Olbers, K. Richter, W. Sell and H. Walden (1973) Measurements of Wind-Wave Growth and Swell Decay During the Joint North Sea Wave Project (JONSWAP), UDC 551.466.31, ANE German Bight, Deutsches Hydrographisches Institut, Hamburg, Germany. Haubrich, R.A. (1965) Earth noise, 5 to 500 millicycles per sound, Journal of Geophysical Research, 70(6): 1425-1427. Havelock, T.H. (1929) Forced surface-wave on water, Phil Mag., F(8):569-576. Havelock, T.H. (1940) The pressure of water waves on a fixed obstacle, Proceedings of the Royal Society of London, Series A, 175:409-421. Havelock, T.H. (1963) The Collected Papers of Sir Thomas Havelock on Hydrodynamics, Office of Naval Research, ONR/ARC-103. Hayashi, K. and T. Takenouchi (1979) The fundamental study of flow field around a vertical cylindrical pile subjected to waves, Proc 26th Japanese Conference on Coastal Engineering, Japan Society of Civil Engineers, Nov:406-410 (in Japanese). Hayashi, K. and T. Takenouchi (1985) Characteristics of flow around a vertical circular cylinder in a wave, Coastal Engineering in Japan, 28:207-222. Heideman, J.C., O.A. Olsen and P.I. Johansson (1979) Local wave force coefficients, Proc. Civil Engineering in the Oceans IV, ASCE, San Francisco, CA, 11:684-699. Helland, K.N., K.S. Lii and M. Rosenblatt (1978) Bispectra of atmospheric and wind tunnel turbulences, Applications of Stats., Ed. P.R. Krishniah, North Holland: 223-248. Helland, K.N. and M. Rosenblatt (1979) Spectral variance estimation and the analysis of turbulence, Physics of Fluids, 22(5):819-823. Helleman, R.H.G. (1980) Self-generated chaotic behavior in nonlinear mechanics. Reprinted in Universality in Chaos. 2nd edition. Ed. P. Cvitanovic, AIP Pub: 420^188. Herring, J.R. (1980) Theoretical calculations of turbulent bispectra, J. Fluid Mech, 97:143-204. Hildebrand, F.B. (1965) Methods of Applied Mathematics, 2nd edition, Prentice-Hall, Inc., Englewood Cliffs, New Jersey, 362 p. Hildebrand, F.B. (1976) Advanced Calculus for Applications, 2nd edition, PrenticeHall, Inc., Englewood Cliffs, New Jersey, 733 p. Hinch, M.J. and C.S. Clay (1968) The application of the discrete Fourier transform in th estimation of power spectra, coherence and bispectra of geophysical data, Review of Geophysics, 6(3):347-363. Hino, M. (1969) Spectrum of wave drag on a pile, Coastal Engrg in Japan, 12:17-27. Hochstrasser, Urs. W. (1965) Orthogonal Polynomials, Handbook of Mathematical Functions, ed. by Abramowitz, M. and LA. Stegun, Dover Pub., NY:774-775.
Bibliography
883
Hogben, N. and R.G. Standing (1974) Wave loads on large bodies, Proc Int Sym on Dynamics of Marine Vehicles and Structures in Waves, London, England: 273-292. Hogben, N. and R.G. Standing (1975) Experience in computing wave loads on large bodies, Proceedings of the Offshore Technology Conference, OTC No. 2189, Vol. 1. Hoffman, D., E.S. Geller and C.S, Niederman (1973) Mathematical simulation and model tests in the design of data buoys, SNAME Annual meeting, New York, NY: 243-273. Hoffman, D. and O.J. Karst, (1975) The theory of the Rayleigh distribution and some of its applications, Journal of Ship Research, 19(3):172—191. Holmes, P. (1986) Chaotic motions in a weakly nonlinear model for surface waves. J. FluidMech. 162:365-388. Horton, T.E. (1977) A Study Advancing Wave-Force Methodology, TR RT-TH-77-6, Dept of Mechanical Engineering, Univ of Mississippi, University, Mississippi, 27 p. Houmb, O.G. (1974) Spectra and bispectra of ocean waves, Proceedings of the 14th Conference on Coastal Engineering, I(16):301-320. Hsu, F.H. and K.A. Blenkarn (1970) Analysis of peak mooring forces caused by slow vessel drift oscillations in random sea, Proc Offshore Tech Conf Houston, 43:135-146. Hsu, J.R.C., R. Silvester and Y. Tsuchiya (1980) Boundary-layer velocities and mass transport in short-crested waves, J. FluidMechs, 99(2):321-342. Huang, N.E. (1970) Mass transport induced by wave motion, Journal of Marine Research, 28(l):35-50. Huang, N.E., S.R. Long, C-C. Tung, Y. Yuen and L.F. Bliven (1981) A unified two-parameter wave spectral model for a general sea state, J. Fluid Mechanics, 112:203-224. Huang, M-C. and R.T. Hudspeth (1982) Pipeline stability under finite-amplitude waves, Journal of Waterway, Port, Coastal and Ocean Div., 108(WW2): 125-145. Huang, M-C, J.W. Leonard and R.T. Hudspeth (1983) Hydrodynamic Interference Between Floating Cylinders in Oblique Seas, Applied Ocean Research, 5(3):158-166. Huang, M-C. and R.T. Hudspeth (1984) Stream function solution for steady waves, Continental Shelf Research, 3(2): 175-190. Huang, M-C, J.W. Leonard and R.T. Hudspeth (1985a) Wave interference effects by finite element method, / . Waterway, Port, Coastal & Ocean Engrg, 111(1): 1-17. Huang, M-C, J.W. Leonard and R.T. Hudspeth (1985b) FEM solution of 3-D wave interference problems, J. Waterways, Port, Coastal & Ocean Engrg, 111(4): 661-677. Hudspeth, R.T. (1974) Prediction of Wave Forces From Nonlinear Random Sea Simulations, Ph.D. thesis, University of Florida.
884
Bibliography
Hudspeth, R.T. (1975) Wave force predictions from nonlinear random sea simulations, Preprints, Offshore Technology Conference, Paper No. OTC 2193,1:471-486. Hudspeth, R.T. (1977) Effects of Nonlinearities on Wave Power Estimates, Journal of the Power Division, 103(POl):51-64. Hudspeth, R.T. (1991) Significance of the Keulegan-Carpenter parameter, Journal of Hydraulics Division, 117(HY12):1626-1638. Hudspeth, R.T. (1994) Nonlinear laboratory waves, Ch. 1 in Fluid Structure Interaction in Offshore Engrg, Adv FluidMech, VI, Ed. S. Chakrabarti, Boston: 1—41. Hudspeth, R.T. and L.S. Slotta (1978) Linear and nonlinear wave action estimates, Journal of the Engineering Mechanics Division, 104(EM2):319-334. Hudspeth, R.T. and M-C. Chen (1979) Digital simulation of nonlinear random waves, Journal of the Port, Coastal and Ocean Division, 105(WWI):67-85. Hudspeth, R.T. and L.E. Borgman (1979) Efficient simulation of digital time series by FFT, Journal of Engineering Mechanics Division, 105(EM2):223-235. Hudspeth, R.T. and J.W. Leonard (1980) Dynamic covariance equations for hinged wavemakers, Engineering Structures, 2:217-224. Hudspeth, R.T. and M-C. Chen (1981) Design curves for hinged wavemakers: theory, Journal of Hydraulics Division, 107(HY5):533-552. Hudspeth, R.T., J.W. Leonard and M-C. Chen (1981) Design curves for hinged wavemakers: experiment, Journal of Hydraulics Division, 107(HY5):553-574. Hudspeth, R.T. and Medina, J.R. (1988) Wave groups analyses by the Hilbert transform, Proc. 21st Int. Conference on CoastalEngrg.:884-%9&. Hudspeth, R.T., J.H. Nath and P.K. Khare (1988) Wave phase/amplitude effects on force coefficients, / Waterway, Port, Coastal, and Ocean Engineering, 114(1): 34-49. Hudspeth, R.T. and W Sulisz, (1991) Stokes drift in 2-D wave flumes, / . FluidMechs, 230:209-229. Hudspeth, R.T., J.M., Grassa, J.R. Medina and J. Lozano (1994) TYPE E doubleactuated wavemakers, IAHR J. Hydraulic Res., 32(3):387-400. Hudspeth, R.T., T. Nakamura and C-K. Pyun (1994) Convergence criteria for axisymmetric green's function with application to floating bodies, J Ocean Engrg, 21(4):381-400. Hudspeth, R.T., L.E. Borgman and B. Samorsorn (1999) Conditional simulation of laboratory waves, / Waterways, Port, Coastal and Ocean Engrg, 125(6): 311-321. Hudspeth, R.T., R.B. Guenther and S. Fadal (2005) Chaotic analyses of weakly damped parametrically excited cross waves with surface tension, Acta Mechanica., 175(1-4): 139-179. Hughes, S.A. (1993) Physical Models and Laboratory Techniques in Coastal Engineering, V7, World Scientific Publishing, New Jersey. Hunt, J.N. (1952) Viscous damping of waves over an inclined bed in a channel of finite width, La Houille Blanche, 7:836-842.
Bibliography
885
Hunt, J.N. (1964) The viscous damping of gravity waves in shallow water, La Houille Blanche, 6:685-691. Hunt, J.N. (1979) Direct solution of wave dispersion equation, Journal of Waterway, Port, Coastal and Ocean Division, Technical Note, 105(WW4):457^159 Hunt, J.N. and S.K.A. Massoud (1962) On mass transport in deep water waves, Pure and Applied Geophysics, 53:65-76. Hunt, J.N. and B. Johns (1963) Currents induced by tides and gravity waves, Tellus, 15:343-351. Hunt, J.N. and R.E. Baddour (1980) Second-order standing waves bounded by circular cylinders, / Waterway, Port, Coastal and Ocean Div, Tech Note, 106(WW1): 122-127. Hunt, J.N. and R.E. Baddour (1981) The diffraction of nonlinear progressive waves by a vertical cylinder, Quarterly Journal of Applied Mathematics, XXXIV(l):69-87. Hyun, J.M. (1976) Theory for hinged wavemakers of finite draft in water of constant depth, Journal of Hydronautics, 10(l):2-7. Hyun, J.M. (1977) Theory for articulated double-flap wavemakers in water of constant depth, Journal of the Society of Naval Architects of Japan, 142:85-92. IAHR Working Group on Wave Generation and Analyses (1989), List of SeaState Parameters, J. of Waterway, Port, Coastal and Ocean Engrg, 115(6): 793-808. Ince, E.L. (1956) Ordinary Differential Equations, Dover Publications, Inc. Ingard, K.U. (1990) Fundamentals of Waves and Oscillations, Cambridge University Press, 595 p. Inman, D.L. (1957) Wave-generated ripples innearshore sands, Technical Memorandum 100, Beach Erosion Board, Washington D.C. Inman, D.L. and A.J. Bowen (1962) Flume experiments on sand transport by waves and currents, Proc 8th Confon CoastalEngr, Mexico City, Mexico: 137-149. Ippen, A.T. (1966) PART I: Estuaries of rectangular section, Estuary and Coastline Hydrodynamics, A.T Ippen, Ed., McGraw-Hill Book Co.:505-507. Iribarren, R. (1965) Formule pour le Calcul des Diques en Enrochements Naturels on Elements Artificieles, XXIInt. Navigation Cong, Sect. II-I, Stocolmo: 15-46. Isaacson, M. de St. Q. (1976a) Mass transport in the bottom boundary layer of cnoidal waves, / . Fluid Mechs, 74:401^113. Isaacson, M. de St. Q. (1976b) The second approximation to mass transport in cnoidal waves, J. Fluid Mechs, 78:445-457. Isaacson, M. de St. Q. (1977a) Nonlinear wave forces on large offshore structures, Journal of Waterway, Port, Coastal and Ocean Division, 103(WW1): 166-170. Isaacson, M. de St. Q. (1977b) Discussion of Forces due to nonlinear waves on vertical cylinders, by H. Raman and P. Vankatanarasaish, Journal of Waterway, Port, Coastal and Ocean Division, 103(WW2):299-300.
886
Bibliography
Isaacson, M. de St. Q. (1978a) Mass transport in shallow water waves, Journal of Waterway, Port, Coastal and Ocean Engineering, 104:215-225. Isaacson, M. de St. Q. (1978b) Vertical cylinders in arbitrary section in waves, Journal of Waterway, Port, Coastal and Ocean Engineering, 104(WW4): 309-324. Isaacson, M. de St. Q. (1978c) Closure to Nonlinear wave forces on large offshore structures, Journal of Waterway, Port, Coastal and Ocean Div, 104(WW4): 457-459. Isaacson, M.de St.Q. (1982) Fixed and Floatmg Axisymmetric Structures in Waves, J of Waterway, Port, Coastal and Ocean Engineering, 108(WW2): 180-199. Isaacson, E. and M. Isaacson (1975) Dimensional Methods in Engineering and Physics, Edward Arnold, London, England. Isaacson, M. de St. Q. and N.G. MacKenzie (1981) Long-term distributions of ocean waves: a review, / of Waterway, Port, Coastal and Ocean Engineering, 107(WW2):93-109. Ishida, A. (1972) Transformation of power spectra of wind-generated waves caused by reflection, Coastal Engineering in Japan, 15:25-33. ISSC, (1964) Proc. Second International Ship Structures Conf, Delft, The Netherlands. ITTC, (1966) Proc. 11th International Towing Tank Conf, Tokyo, Japan. Iwagaki, Y. and Y. Tsuchiya (1966) Laminar damping of oscillatory waves due to bottom friction, Proc 10th Int Conf Coastal Engineering, Tokyo, Japan: 149-174. Jazwinski, A.H. (1970) Stochastic Processes and Filtering Theory, Academic Press, NY. Jen, Y. and L. Skjelbriea (1973) Discussion of Nonlinear wave forces on vertical cylinder, by S.K. Chakrabarti, Journal of Hydraulics Division, 99(HYll):21-35. Jenkins, G.M. and D.C. Watts (1968) Spectral Analysis and Its Applications, San Francisco, Holden Day. John, F. (1949) On the motion of floating bodies, I., Communications on Pure and Applied Mathematics, 2:13-57. John, F. (1950) On the motion of floating bodies, II., Communications on Pure and Applied Mathematics, 3:45-101. Johns, B. (1968) A boundary-layer method for the determination of the viscous damping of small amplitude gravity waves, Quarterly J. Mechs and Applied Maths, 21:93-103. Johns, B. (1970) On the mass transport induced by oscillatory flow in a turbulent boundary layer, / . Fluid Mechs, 43(1): 177-185. Jones, A.F. (1984) The generation of cross-waves in a long deep channel by parametric resonance. J. FluidMech. 138:53-74. Jones, D. (1964) The Theory of Electromagnetism, MacMillan, New York. Jonsson, I.G. and O. Skovgaard (1979) A mild slope equation and its application to tsunami calculations, Marine Geodesy, 2:41-58.
Bibliography
887
Jordan, D.W. and P. Smith (1987) Nonlinear Ordinary Differential Equations. Second Edition. Clarendon Press, Oxford. Joseph, D.D. (1973) Domain perturbation: the higher order theory of infinitesimal water waves, Arch. Rational Mech. Analysis, 51:295-303. Kampe de Feriet, J. (1954) Introduction to the statistical theory of turbulence, SIAM Journal of Applied Mathematics, Part I, 2:1-9; Part II, 2:143-174; Part III, 2:244-271. Kampe de Feriet, J. (1961) Statistical fluid mechanics: two-dimensional linear gravity waves, In Partial Differential Equations and Continuum Mechanics, ed. by R.E. Langer, University Wisconsin Press, Madison, Wisconsin. Karim, S. M. (1953) The second viscosity coefficient of liquids, / . Acous. Soc. Amen, 25:997-1002. Keating, T. and N.B. Webber (1977) The generation of periodic waves in a laboratory channel: a comparison between theory and experiment, Proc Inst Civil Engs, 63(2):819-832. Kellog, O.D. (1953) Foundations of Potential Theory, Dover Publ, Inc., NY: 211-216. Kelman, R.B. Least squares Fourier series solutions to boundary value problems, SIAM Review, 21(3):329-338. Kendall, M.G. and A. Stewart (1961) Advanced Theory of Statistics, 2, Hafner Pub., N.Y. Kennard, E.H. (1949) Generation of surface waves by a moving partition, Quarterly Journal of Applied Mathematics, 7(3):303-312. Kershaw, D.S. (1978) The incomplete Cholesky-conjugate gradient method for the iterative solution of systems of linear equations, / . of Computational Physics, 26:43-65. Keulegan, G.H. (1959) Energy dissipation in standing waves in rectangular basins, J. Fluid Mechs, 6:33-50. Keulegan, G.H. and L.H. Carpenter (1958) Forces on cylinders and plates in an oscillating fluid, Journal of Research, National Bureau of Standards, No. 2857, 60(5):423^140. Kevorkian, J. (1980) Resonance in a weakly nonlinear system with slowly varying parameters, Studies in Applied Mathematics, 62:23-67. Kim, C.H. (1969) Hydrodynamic forces and moments for heaving, swaying and rolling cylinders on water of finite depth, Journal of Ship Research, 13:137-154. Kim, WD. (1964) On Oscillating Ships in Waves, Technical Memorandum No. 25, Flight Science Laboratory, Boeing Scientific Research Laboratories. Kim, W.D. (1965) On the harmonic oscillations of a rigid body on a free surface, J. FluidMechs, 21:429-451. Kim, W.D. (1966) On a free-floating ship in waves, Journal of Ship Research, 10(3).
888
Bibliography
Kim, Y.Y. and H.C. Hibbard (1975) Analysis of simultaneous wave force and water particle velocity measurements, Proc Offshore Technology Conf., Proc paper 2192. Kimura, A. (1985) The decomposition of incident and reflected random wave envelopes, Coastal Engineering in Japan, 23:59-69. Kimura, A. and Y. Iwagaki (1976) Random wave simulation in a laboratory wave tank, Proc 15th International Conf Coastal Engrg, Honolulu, Hawaii, Chap. 21, 1:368-387. Kinsman, B. (1965) Wind Waves, Prentice-Hall, Inc., Englewood Cliffs, New Jersey: 107-110:424^188. Kirby, J. T. (1986a) A general wave equation for waves over rippled beds, J. Fluid Mechs, 162:171-186. Kirby, J. T. (1986b) On the gradual reflection of weakly nonlinear Stokes waves in regions with varying topography, /. Fluid Mechs, 162:187-209. Kit, E., Shemer, L. and Miloh, T. (1987) Experimental and theoretical investigation of nonlinear sloshing waves in a rectangular channel. J. FluidMech. 181:265-291. Kitaigorodskii, S.A., V.P. Krasitskii and M.M. Zaslavskii (1975) On Phillips theory of equilibrium Range in the spectra of wind-generated gravity waves, Journal of Physical Oceanography, 5:410-420. Kobayashi, N., D.T. Cox and A. Wurjanto (1990) Irregular wave reflection and run-up on rough impermeable slopes, / . Waterway, Port, Coastal and Ocean Engrg Div, 116(6):708-726. Komar, P.D. (1976) Beach Processes and Sedimentation, Prentice-Hall, Inc., Englewood Cliffs, N.J., 429 p. Komar, P.D. and M.K. Gaughan (1972) Airy wave theory and breaker height prediction, Proc 13th International Conf Coastal Engrg, Chap. 20,1:405—418. Kravtchenko, J. (1954) Remarqures sur le calcul des amplitudes des houle linaire engendres par un batteur, Proc. 5th Conf. Coastal Engrg., 5:50-61 (in French). Kreisel, G. (1949) Surface waves, Quarterly of Applied Mathematics, 7:21—44. Krishnamacker, P. (1972) Discussion on design curves for regular and random wave generators, by Gilbert, G., D.M. Thompson and A.J. Brewer, / Hyd Res, 10(2):227-233. Kritis, Ir. B. (1979) Heaving motions of axisymmetric bodies, The Naval Architect, 1:26. Lake, B.M., H.C. Yuen, H. Rungalkier and W.E. Ferguson (1977) Nonlinear deepwater waves: theory and experiment, Part 2. Evolution of a continuous train, J. Fluid Mechs, 83(l):49-74. Lake, B.M. and H.C. Yuen (1977) A note on some nonlinear water-water experiments and the comparison of data with theory, / . Fluid Mechs, 83(1):75-81. Lake, B.M. and H.C. Yuen (1978) A new model for nonlinear wind waves: Part 1, Physical model and experimental evidence, /. Fluid Mechs, 88:33-63.
Bibliography
889
Lamb, H. (1932) Hydrodynamics, 6th Ed., Dover Pub. Inc., New York, 738 p. Lambrakos, K.F. (1981a) The extended velocity potential versus Stokes wave representation, Journal of Geophysical Research, 86(C7):6473-6480. Lambrakos, K.F. (1981b) Extended velocity potential wave kinematics, Journal of Waterway, Port, Coastal and Ocean Division, 107:159-174. Landau, L.D. and E.M. Lifshitz (1987) Fluid Mechanics. Pergamon Press. Langhaar, H.L. (1951) Dimensional Analysis and the Theory of Models, Chapman and Hall, New York. Lau, J. and B. Travis (1973) Slowly varying Stokes waves and submarine longshore bars, Journal of Geophysical Research, 78(21):4489-9497. Le Mehaute, B. (1976a) Similitude in coastal engineering, Journal of Waterway, Port, Coastal and Ocean Division, 102(WW3):317-335. Le Mehaute, B. (1976b) Hydrodynamics and Water Waves, Springer-Verlag. Le Mehaute, B., D. Divorky and A. Lin (1968). Shallow water waves: a comparison of theory and experiment, Proc 11th International Conf Coastal Engrg, London: 86-107. Lee, CM. (1960) The second-order theory for nonsinusoidal oscillations of a cylinder in a free surface, Proceedings of the 8th Symposium on Naval Hydrodynamics: 905-951. Lee, CM. (1968) The second-order theory of heaving cylinders in a free surface, Journal of Ship Research, 12(4):313-327. Lee, CM. and J.N. Newman (1971) The vertical mean force and moment of submerged bodies under waves, Journal of Ship Research:231-245. Leonard, J.W., M-C Huang and R.T. Hudspeth (1983) Hydrodynamic interferences between floating cylinders in oblique seas, Applied Ocean Research, 5(3):158-166. Levi-Civita, T. (1977) The Absolute Differential Calculus (Calculus of Tensors), Dover Pub, Inc., 452 p. Lewy, H. (1946) Water waves on sloping beaches, Bulletin of the American Mathematical Society, 52(9):737-775. Lewy, H. (1950) Development at the confluence of analytic boundary conditions, University of California Publication on Mathematics, New Series, l(7):601-605. Lichtenberg, A.J. and M.A. Lieberman, (1992) Regular and Chaotic Dynamics. 2nd Ed, Springer-Verlag. Lieberman, L. N. (1949) The second viscosity of liquids, Phys. Rev., 75:1415-1422. erratum, 76:440. Lighthill, J. (1964) Fourier Analysis and Generalised Functions, Cambridge University Press, New York, N.Y.
890
Bibliography
Lighthill, J. (1979) Waves and hydrodynamic loading, Proc 2nd International Conference on the Behavior of Offshore Structures, London, England, I: 1-29. Lii, K.S., M. Rosenblatt and C. Van Atta (1976) Bispectral measurements in turbulence, J. Fluid Mechs, 77:45-62. Lin, WM. (1984) Nonlinear motion of the free surface near a moving body, Ph.D. Thesis, Mass. Inst. Tech., Boston, MA., 127 p. Lin, C.C. and L.A. Segel (1974) Mathematics Applied to Deterministic Problems in the Natural Sciences, MacMillan Publishing Company, Inc., New York, 604 p. Littlewood, R.K. and V. Zakian (1976) Numerical evaluation of Fourier integrals, Journal of the Institute of Mathematical Applications, 18:331-339. Liu, A-K. and S.H. Davis (1977) Viscous attenuation of mean drift in water waves,J. Fluid Mechs, 81(l):63-84. Liu, P. L-F. (1977) Mass transport in water waves propagated over a permeable bed, Coastal Engineering, The Netherlands, 1:79-96. Longuet-Higgins, M.S. (1950) A theory for the origin of microseisms, Philosophical Transactions of the Royal Society of London, A, 245:1-35. Longuet-Higgins, M.S. (1953) Mass transport in water waves, Philosophical Transactions of the Royal Society of London, A, 245:535-581. Longuet-Higgins, M.S. (1956) The statistical analysis of a random, moving surface, Proceedings of the Royal Society of London, A, 249:321-386. Longuet-Higgins, M.S. (1958) The mechanics of the boundary-layer near the bottom in a progressive wave, (Appendix to Russell and Osorio (1958)), Proceedings of the 6th International Conference on Coastal Engineering, Miami, Florida: 184-193. Longuet-Higgins, M. S. (1960) Mass transport in the boundary layer at a free oscillating surface, J. Fluid Mechs, 8(2):293-306. Longuet-Higgins, M.S. (1963) The effects of nonlinearities on statistical distributions in theory of sea waves, J. Fluid Mechs, 17:459—480. Longuet-Higgins, M.S. (1964) Modified Gaussian distribution for slightly nonlinear variables, Radio Science Journal of Research, 680(9): 1049-1062. Longuet-Higgins, M.S. (1969) On wave breaking and the equilibrium spectrum of wind-generated waves, Proceedings of the Royal Society of London, A, 310: 151-159. Longuet-Higgins, M.S. (1975) On the joint distribution of the periods and amplitudes of sea waves, Journal of Geophysical Research, 80(18):2688-2694. Longuet-Higgins, M.S. (1978a) The instabilities of gravity waves in Finite amplitude in deep water, II Subharmonics, Proc of the Royal Society of London, A, 360:489-505.
Bibliography
891
Longuet-Higgins, M.S. (1978b) On the nonlinear transfer of energy in the peak of a gravity-wave spectrum: a simplified model, Proc Royal Soc London, A, 347:311-328. Longuet-Higgins, M.S. (1978c) The orbital motion in steep water waves, (Appendix to Nath (1978)), Proc 16th International Conf Coastal Engrg, Hamburg: 874-877. Longuet-Higgins, M.S., D.E. Cartwright andN.D. Smith (1961) Observations of the directional spectrum of sea waves using the motions of a floating buoy, Ocean Wave Spectra, Prentice-Hall, Inc.:lll-132. Lorentz, H.A. (1926) Verslag staatscommissie Zuidezee 1918-1926, Staatsdrukkerij, The Hague, The Netherlands. Luke, J.C. (1967) A variational principle for a fluid with a free surface. J. Fluid Mechs., 27:395-397. Luke, Y.L. (1975) Mathematical Functions and Their Approximations, Academic Press, Inc., London:337-359. Lumley, J.L. (1970) Stochastic Tools in Turbulence, Academic Press, New York. Lyons, C.G. (1973) Soil resistance to lateral sliding of marine pipelines, Proceedings of the 5th Offshore Technology Conference, OTC No. 1876,11:479-484. Ma, Y.-C. (1978) The complete solution of the long-wave short-wave resonance equations, Studies in Applied Mathematics, 59:201-221. MacCamy, R.C. and R.A. Fuchs (1954) Wave Forces on Piles: A Diffraction Theory, T.M. No. 69, Beach Erosion Board. MacDonald, G.J.F. (1963) The bispectra of atmospheric pressure records, Proceedings of the IBM Scientific Symposium on Statistics, White Plains, New York: 247-269. Madsen, O.S. (1970) Waves generated by a piston-type wavemaker, Proceedings of the 12th Conference on Coastal Engineering, Washington, D.C.:589-607. Madsen, O.S. (1971) On the generation of long waves, Journal of Geophysical Research, 76(36):8672-8683. Madsen, O.S. (1974) A three dimensional wavemaker, its theory and application, IAHR, Journal of Hydraulic Research, 12(2):205-222. Madsen, O.S., C.C. Mei and R.P. Savage (1970) The evolution of time-periodic long waves of finite amplitude, J. Fluid Mechs, 44:195-208. Magnuson, A.H. (1981) Spectral theory of water waves, Journal of Ship Research, 25(1): 1-7. Mahoney, J.J. (1972) Cross waves, Part I, theory, J. Fluid Mechs, 55(2):229-245. Mardia, K.V. (1972) Statistics of Directional Data, Academic Press, London, 239 p. Marquardt, D.W. (1963) An algorithm for least squares estimation of nonlinear parameters, Journal of the Society of Industrial Applications of Mathematics, 11(2):431-441. Maruo, H. (1960) The drift of a body floating on waves, / . Ship Research, 4:1-10.
892
Bibliography
Mase, M. and Iwagaki, Y. (1986) Wave group analysis from statistical viewpoint, Proc. Ocean Structural Dynamics Symposium '86, Corvallis, OR: 145-157. Maskell, S.J. and F. Ursell (1970) The transient motion of a floating body, / . Fluid Mechs, 44:303-313. Maull, D.J. and M.G. Milliner (1979) The forces on a circular cylinder having complex periodic motion, in Mechanics of Wave-Induced Forces on Cylinders, ed. T.L. Shaw, Pitman, London:490-502. McComas, C.H. and M.G. Briscoe (1980) Bispectra of internal waves, J. Fluid Mechs, 97:205-213. McCowan, J. (1894) On the highest wave ofpermanent type, Phil. Mag., 38:45-5 8. McDougal, W.G. and R.T. Hudspeth (1983a) Wave setup/setdown and longshore current on non-planar beaches, Coastal Engineering, 7(2):103—117. McDougal, W.G. and R.T. Hudspeth (1983b) Longshore transport on non-planar beaches, Coastal Engineering, 7(2):119-131. McDougal, W.G. and C.K. Sollitt, (1984) Geotextile stabilization of seabeds: theory, Journal ofEngrg Structures, Butterworth Scientific-Journals Div, Surrey, United Kingdom, 6:211. McDougal, W.G. and R.T. Hudspeth (1986) Influence of lateral mixing on longshore currents, Journal of Ocean Engineering, 13(5):419-433. Mclver, D. and D.V. Evans (1984) The occurrence of negative added mass in free surface problems involving submerged oscillating bodies, /. Engineering Mathematics, 18:7-22. Medina, J.R. (1992) Deterministic computer-aided optimum design of rock rubble-mound breakwater cross-sections, (Discussion) Coastal Engineering, 17:279-282. Medina, J.R. and R.T. Hudspeth (1987) Sea states defined by wave height and period functions, Proc. IAHR Seminar on Wave Analysis and Generation in Laboratory Basins, 22nd IAHR Congress, Lausanne, Switzerland:249-259. Medina, J.R., C. Fassardi and R.T. Hudspeth (1990) Effects of wave groups on the stability of rubble mound breakwaters, Proc, 22nd Int. Conf Coastal Engrg.: 1552-1563. Medina, J.R. and R.T. Hudspeth (1990) A review of the analyses of wave groups, Coastal Engineering, 14:515-542. Medina, J.R. and W.G. McDougal (1990) Deterministic and probabilistic design of breakwater armor layer, (Discussion), / Waterway, Port, Coastal and Ocean Engrg, 116(4):508-510. Medina, J.R., R.T. Hudspeth and C. Fassardi (1994) Breakwater armor damage due to wave groups, J Waterways, Port, Coastal and Ocean Engrg, V120(2): 179-198. Mei, C.C. (1969) Weak reflections of water waves by bottom obstacles, Journal of Engineering Mechanics Division, 95(EM1): 183-194.
Bibliography
893
Mei, C.C. (1973) Shoaling of spiral waves in a circular basin, Journal of Geophysical Research, 78(6):977-980. Mei, C.C. (1976) Power extraction from water waves, J. Ship Research, 20(2):63-66. Mei, C.C. (1978) Numerical methods in water-wave diffraction and radiation, Annual Review of Fluid Mechanics, 10:393—413. Mei, C.C. (1989) The Applied Dynamics of Ocean Surface Waves, World Scientific Pub., Singapore, 740 p. Mei, C.C. and J. Black (1969) Scattering of surface waves by rectangular obstacles in waters of finite depth, / . Fluid Mechs, 38:499-511. Mei, C.C, P.L.-F. Liu and T.R. Carter (1972) Mass Transport in Water Waves, Report 146, R. M. Parsons Lab, Massachusetts Institute of Technology, Cambridge, Massachusetts. Mei, C.C. andU. Unluata, (1972) Harmonic generation in shallow water waves, Waves on Beaches and Resulting Sediment Transport, R.E. Meyer ED, Academic Press: 181-202. Mei, C.C. and P.L.-F., Liu (1973) The damping of surface gravity waves in a bounded fluid, / . Fluid Mechs, 59(2):239-256. Melnikov, V.K. (1963) On the stability of the center for time-periodic perturbations. Trans. Moscow Math. Soc. 12:1-57. Miche, A. (1951) Le pouvoir reflechissant des our rages maritimes a expose Taction de la houle, Annates desponts et chaussees, annee 121(3):285-319. Michell, J.H. (1893) On the highest waves in water, Phil. Mag, 36(5):430-435. Middleton, D. (1960) An Introduction to Statistical Communication Theory, McGraw-Hill Book Company, Inc., New York. Mihram, G.A. (1962) Simulation: Statistical Foundations and Methodology, Academic Press, NY: 5 8-64. Mikhlin, S.G. (1957) Integral Equations and Their Applications to Certain Problems in Mechanics, Mathematical Physics and Technology, Pergamon Press, New York. Mikhlin, S.G. (1965) Multidimensional Singular Integrals and Integral Equation, Pergamon Press, New York. Miles, J.W. (1962) Stability of Forced Oscillations of a Spherical Pendulum, Qtly. Appl. Math., 20:21-32. Miles, J.W. (1967a) Surface wave scattering matrix for a shelf, J. Fluid Mechs, 28(4):755-767. Miles, J.W. (1967b) Surface wave damping in closed basins, Proceedings of the Royal Society of London, Series A, 297:459-^175. Miles, J.W. (1971) A note on variational principles for surface wave scattering, J. FluidMechs, 46:141-149. Miles, J.W. (1974) Harbor seiching, Annual Reviews of Fluid Mechanics, 6: 17-35.
894
Bibliography
Miles, J.W. (1976) Nonlinear surface waves in closed basins, J. Fluid Mechs, 75(3):419^148. Miles, J.W. (1977) On Hamilton's principle for surface waves, J. Fluid Mechs, 83(1):153-158. Miles, J.W. (1981a) Hamiltonian formulations for surface waves, Applied Scientific Research, 37:103-110. Miles, J.W. (1981b) Nonlinear Helmholtz oscillations in harbors and coupled basins, J. Fluid Mechs, 104:407^118. Miles, J.W. (1984a) Strange Attractors in Fluid Dynamics, Advances in Appl. Mech., 24:189-214. Miles, J.W. (1984b) Resonant Motion of a Spherical Pendulum, Physica, 11D: 309-323. Miles, J.W. (1984c) Resonantly Forced Motion of Two Quadratically Coupled Oscillators, Physica, 13D:247-260. Miles, J.W. (1984d) Nonlinear Faraday resonance, J. Fluid Mechanics, 146:285-302. Miles, J.W. (1984e) Internally resonant surface waves in a circular cylinder, /. Fluid Mech, 149:1-14. Miles, J.W. (1984f) Resonantly forced surface waves in a circular cylinder, J. Fluid Mechs, 149:15-31. Miles, J.W. (1985a) Note on a parametrically excited, trapped cross-wave, /. Fluid Mechs, 151:391-394. Miles, J.W. (1985b) Surface waves in basins of variable depth, J. Fluid Mechs, 152:379-389. Miles, J.W. (1988a) Parametrically excited, standing cross-waves, /. Fluid Mechs, 186:119-127. Miles, J.W. (1988b) Parametrically excited, progressive cross-waves, J. Fluid Mechs, 186:129-146. Miles, J.W. (1988c) The evolution of a weakly nonlinear, weakly damped, capillarygravity wave packet, / . Fluid Mechs, 187:141-154. Miles, J.W. (1988d) Guided surface waves near cutoff, /. Fluid Mechs, 189:287-300. Miles, J.W and J. M. Becker (1988) Parametrically excited, progressive cross-waves, J. Fluid Mechs, 186:129-146. Miles, J.W. and D. Henderson (1990) Paramerically forced surface waves. Annu. Rev. FluidMech. 22:143-165. Milgram, J. (1970) Active wave absorbers, /. Fluid Mechs, 43:845-859. Milgram, J.H. and J.E. Halkyard (1971) Wave forces on large objects in the sea, Journal of Ship Research, 15(2): 115-124. Miller, J.C.P. (1965) Parabolic Cylinder Functions, Handbook of Mathematical Functions, ed. by M. Abramowitz and LA. Stegun, Dover Pub Inc., NY:686-720.
Bibliography
895
Miller, B.L.P. (1975) Discussion of Forces on cylinders and spheres in a sinusoidally oscillating fluid, by T. Sarpkaya, Trans ASME, J Applied Mechanics, E, 42:744-745. Milne-Thomson, L.M. (1965a) Chapter 16. Jacobian elliptic functions and theta functions, ed. by M. Abramowitz and LA. Stegun, Dover Publications, Inc., NY: 567-586. Milne-Thomson, L.M. (1965b) Chapter 17. Elliptic integrals, ed. by M. Abramowitz and LA. Stegun, Dover Publications, Inc., NY:587-625. Milne-Thomson, L.M. (1968) Theoretical Hydrodynamics, 5th Ed., The Macmillan Co., 743 p. Miloh, T. (1980) Irregularities in solutions of nonlinear wave diffraction problem by vertical cylinder, / Waterway, Port, Coastal and Ocean Div, TN, 106(WW2): 279-284. Mitsuyasu, H. (1971) On the form of fetch-limited wave spectrum, Coastal Engrg in Japan, 17:7-14. Mitsuyasu, EL, T. Suhaya.S. Muzuno, M. Ohkuso, T. Honda and K. Rikiishi (1975) Observations of the directional spectrum of waves using a cloverleaf buoy, /. Phys. Oceanography, 5:750-760. Molin, B. (1979a) Second-order diffraction loads upon three-dimensional bodies, Applied Ocean Research, 1 (4): 197-202. Molin, B. (1979b) Second-order drift forces upon large bodies in regular waves, Proc 2nd Int Conf on Behavior of Offshore Structures, Imperial College, London: 363-370. Monacella, V.J. (1966) The disturbance due to a slender ship oscillating in waves in a fluid of finite depth, Journal of Ship Research, 10(4):242-252. Mook, D.T., L.R. Marshall and A.H. Nayfeh (1974) Subharmonic and superharmonic resonances in the pitch and roll modes of ship motions, J Hydronautics, 8(1):32^10. Moon, EC. (1992) Chaotic and Fractal Dynamics. An Introduction for Applied Scientists and Engineers, John Wiley & Sons. Morales, C. de C. (1970) Experiments of wave reflection on impermeable slopes, Proc 12th Conference on Coastal Engineering, Washington, D.C., 31:509-521. Morison, J.R., M.P. O'Brien, J.W. Johnson and S.A. Schaaf (1950) The force exerted by surface waves or piles, Petroleum Transactions, American Institute ofMin. Metal Engineering, 189:149-157. Morse, P.M. and H. Feshbach (1953) Methods of Theoretical Physics, McGraw-Hill Book Company, New York. Morse, P.M. andK.V. Ingard(1968) Theoretical Acoustics, McGraw-Hill Book Co. Muga, B.J. and J.F. Wilson (1970) Dynamic Analysis of Ocean Structures, Plenum Press, NY:92-98.
896
Bibliography
Multer, R.H. and C.J. Galvin (1967) Secondary waves: Periodic waves of nonpermanent form, (abstract) EOS Transaction of the American Geophysical Union, 48. Multer, R.H. (1970) A Nonlinear Numerical-Hydrodynamical Model of a Mechanical Water Wave Generator, Doctoral dissertation, University of Iowa, Ames, Iowa. Multer, R.H. (1973) Exact nonlinear model of wave generator, Journal of the Hydraulics Division, 99(HY1):31^16. Munk, W.H. (1949) The solitary wave theory and its application to surf problems, N.Y.Acad. Scl Ann., 51:376-424. Murnaghan, F.D. (1963) Introduction to Applied Mathematics, Dover Publications, Inc., NY:205-304. Murphy, J.E. (1978) Integral equation failure in wave calculation, Journal of Waterway, Port, Coastal and Ocean Division, 104(WW4):330-334. Muskhelishvili, N.I. (1953) Singular Integral Equations, Noordhoff, Holland. Nath, J.H. (1978) Drift speed of buoys in waves, Proceedings of the 16th Conference on Coastal Engineering, Hamburg, Germany:859-873. Nath, J.H. (1981) Hydrodynamic coefficients for macro-roughness, Proceedings of the Offshore Technology Conference, Houston, Texas, OTC 3989:180-208. Nath, J.H. (1982) Heavily roughened horizontal cylinders in waves, Proc 3rd International Conf on the Behavior of Offshore Structures, Boston, MA: 387^107. Natvig, B.J. and J.W. Pendered (1977) Nonlinear motion response of floating structures to wave excitation, Proceedings of the Offshore Technology Conference, OTC 2796. Natvig, B.J. and J.W. Pendered (1980) Motion response of floating structures to regular waves, Applied Ocean Research, 2(3):99-106. Nayak, LB. (1970) Equilibrium profiles of model beaches, Hydraulic Engineering Laboratory, Technical Report HFL-2-25, University of California, Berkeley, California. Nayfeh, A.H. (1973) Perturbation Methods, John Wiley & Sons, Inc., New York. Nayfeh, A.H. (1981) Introduction to Perturbation Techniques, John Wiley & Sons, N.Y. Nayfeh, A.H. (1985) Problems in Perturbation, John Wiley & Sons, New York, 556 p. Nayfeh, A.H., D.T. Mook and L.R. Marshall (1973) Nonlinear coupling of pitch and roll modes in ship motions, Journal ofHydronautics, 7(4):145—152. Nayfeh, A.H., D.T. Mook and L.R. Marshall (1974) Perturbation energy approach for the development of the nonlinear equations of ship motion, / Hydronautics, 8(4): 130-136. Nayfeh, A.H. and D.T. Mook (1979) Nonlinear Oscillations, John Wiley & Sons, N.Y.
Bibliography
897
Nekrasov, A.I. (1921) On waves of permanent type I, Izv. Ivanovo-Vosnosonk. Politehn. Inst, 3:52-65. Nekrasov, A.I. (1922) On waves of permanent type II, Izv. Ivanovo-Vosnosonk. Politehn. Inst, 3:155-171. Neshyba, S. and E.J.C. Sobey (1975) Vertical cross coherence and cross bispectra between internal waves measured in a multiple-layered ocean, J Geophysical Res, 80(9):1152-1162. Neumann, G. (1943) On ocean wave spectra and a new method of forecasting windgenerated sea, TM43, U.S. Army Corps of Engineers, Beach Erosion Board. Newell, A.C. (1968) The closure problem in a system of random gravity waves, Review in Geophysics, 6(1): 1-31. Newman, J.N. (1959) The damping and wave resistance of a pitching and heaving ship, Journal of Ship Research, 3:1-19. Newman, J.N. (1961) The linearized theory for the motion of a thin ship in regular waves, Journal of Ship Research, 34. Newman, J.N. (1962) The exciting forces on fixed bodies in waves, J. Ship Res, 6(3):10-17. Newman, J.N. (1964) A slender-body theory for ship oscillations in waves, J. Fluid Mechs, 18(4):602-618. Newman, J.N. (1965a) Propagation of water waves past long 2-D obstacles, / . Fluid Mechs, 23:23-29. Newman, J.N. (1965b) Propagation of water waves over an infinite step, /. Fluid Mech, 23:249^115. Newman, J.N. (1967) The drift force and moment of ships in waves, / . Ship Res, 11:51-60. Newman, J.N. (1974) Second-order, slowly-varying forces on vessels in irregular waves, Proceedings of the International Symposium on the Dynamics of Marine Vehicles and Structures in Waves, University College, London, England: 193-197. Newman, J.N. (1975) Interaction of waves with two-dimensional obstacles: a relation between the radiation and scattering problems, J. Fluid Mechs, 71(2): 273-282. Newman, J.N. (1977) Marine Hydrodynamics, MIT Press, Massachusetts: 285-321. Newman, J.N. (1985) Algorithms for the free-surface Green's function, / . of Engineering Mathematics, 19:57-67. Noda, H. (1968) A study on mass transport in boundary layers in standing waves, Proc 14th Conference on Coastal Engineering, London, England:227-247. Norris, J.W. (1994) The nonlinear Mathieu equation. International Journal of Bifurcation and Chaos 4:71-86.
898
Bibliography
Oates, G.C. (1990) Ch. 3, Vector Analysis, Handbook of Applied Mathematics, Second Ed., C.E. Pearson, editor, VanNostrand Reinhold Company, NY: 176-7. Oberhettinger, F. (1965) Hypergeometric Functions, Handbook of Mathematical Functions, Dover Publications, Inc., NY:556-565. Oberhettinger, F. (1972) Tables ofBessel Transforms, Springer-Verlag, New York. Oberhettinger, F. and L. Badd (1973) Tables of Laplace Transforms, Springer-Verlag. O'Brien, M.P. and J.R. Morison (1952) The forces exerted by waves on objects, Transactions of the American Geophysical Union, 33(l):32-38. Ochi, M.K. (1978) Generalization of Rayleigh probability distribution and Its application, Journal of Ship Research, 22(4):259-265. Ochi, M.K. and E.N. Hubble (1976) Six-parameter wave spectra, Chapter 18 Proceedings of the 15th Coastal Engineering Conference, Hawaii:301-328. Ockendon, J.R. and H. Ockendon (1973) Resonant surface waves, J. Fluid Mechs, 59:397^113. Ogilvie, T.F. (1963) First- and second-order forces on a cylinder submerged under a free surface, / . Fluid Mechs, 16:451^172. Ogilvie, T.F. (1973) The Chertock formulas for computing unsteady fluid dynamic force on a body, ZAMM 53:573-582. Olver, W.J. (1990) Asymptotic methods, Ch. 12, Handbook of Applied Mathematics, Second Ed., C.E. Pearson, editor, Van Nostrand Reinhold Company:631-632. Orum, J.C, R.T. Hudspeth, W. Black and R.B. Guenther (2000) Extension of Herglotz algorithm to nonautonomous canonical transformations, SIAM Review, 42(l):83-90. Otnes, R.K. and L. Enochson (1972) Digital Time Series Analysis, Wiley Interscience, New York, 467 p. Papoulis, A. (1968) Systems and Transforms with Applications in Optics, McGrawHill Book Company, New York. Papoulis, A. (1984) Probability, Random Variables, and Stochastic Processes, Second Edition, McGraw-Hill International Editions, McGraw-Hill Book Company, New York, 576 p. Parker, T.S. and L.O. Chua (1989) Practical Numerical Algorithms for Chaotic Systems, Springer-Verlag. Patel, M.H. and P.A. Ionnaou (1980) Comparative performance study of paddle- and wedge-type wave generators, Journal ofHydronautics, 14(l):5-9. Pearson, C.E. (1990) Handbook ofApplied Mathematics, VanNostrandReinholdCo. Pearson, H.L. (1990) Elements of analysis, Chapter 2, Handbook of Applied Mathematics, Second Ed., C.E. Pearson, editor, Van Nostrand Reinhold Co:83-128. Penzien, J., C.F. Scheffey and R.A. Parmelee (1964) Seismic analysis of bridges on long piles, Journal of Engineering Mechanics, 90(3):223-254.
Bibliography
899
Peters, A.S. and J.J. Stoker (1957) The motion of a ship as a floating rigid body in a seaway, Communications on Pure and Applied Mathematics, 10:399—490. Phillips, O.M. (1977) The Dynamics of the Upper Ocean, 2nd edition, Cambridge University Press, Cambridge, England, 336 p. Phillips, O.M. (1981) Wave interactions - the evolution of an idea, / . Fluid Mechs, 106:215-227. PIANC-PTC II (1985) The stability of rubble mound breakwaters in deeper water, Permanent Int Assoc Navigation Congresses, Brussels, Supplement to Bulletin No. 48. Pierson, W.J., Jr. (1955) Wind generated gravity waves Advances in Geophysics, Academic Press, New York, 2:93-178. Pierson, W.J. (1962) Perturbation analysis of the Navier-Stokes equations in Lagrangian form with selected linear solutions, Journal of Geophysical Research, 67(8):3151-3160. Pierson, W.J., Jr., G. Neumann and R.W James (1955) Observing and forecasting ocean waves by means of wave spectra and statistics, Hydrographic Pub, No 603, U.S. Navy. Pierson, W.J., Jr. and K. Moskowitz (1964) A proposed spectral form for fully developed wind seas based on the similarity theory of S.A. Kitaigorodskii, Journal of Geophysical Research, 69(24):5181-5190. Ploeg, J. and Funke, E. (1980): A survey of random wave generation techniques. Proc. of the 17th Coastal Engineering Conference, Sydney, Australia: 137-154. Porter, W.R. (1960) Pressure Distribution, Added-Mass and Damping Coefficients for Cylinders Oscillating in a Free Surface, T R 82(16), Univ California, Berkeley, CA. Potash, R.L. (1971) Second-order theory of oscillating cylinders, J Ship Res, 15(4): 295-324. Potynody, J.G. (1961) Skin friction between various soils and construction materials, Geotechnique, 11:339-353. Press, W. H., B. P. Flannery, S. A. Teukolsky and W. T. Vetterling (1986) Numerical Recipes. Cambridge: Cambridge University Press:43-52. Price, W.G. and R.E.D. Bishop (1974) Probabilistic Theory of Ship Dynamics, John Wiley & Sons, Inc. Priestly, M.B. (1965) Evolutionary spectra and nonstationary processes, Journal of the Royal Statistical Society, Series B, 27:204-236. Priestly, M.B. (1966) Design relations for nonstationary processes, Journal of the Royal Statistical Society, Series B, 28:228-240. Priestly, M.B. (1967) Power spectra analysis of nonstationary random processes, Journal of Sound Vibration, 6(l):86-97. Proceedings Civil Engineering in the Oceans IV (1979) American Society of Civil Engineers, San Francisco, CA.
900
Bibliography
Radder, A.C. (1979) On the parabolic equation method for water wave propagation, / . Fluid Mechs, 95:159-176. Rahman, M. and I.C. Chakravartty (1981) Hydrodynamic loading calculations for offshore structures, SIAM Journal of Applied Mathematics, 41(3): 445-458. Raman, H., G.V. Rao Prabhakara and P. Venkatanarasiah (1975) Diffraction of nonlinear surface waves by circular cylinder, Acta Mechanica, Springer-Verlag, 23:145-158. Raman, H. and P. Venkatanarasiah (1976) Forces due to nonlinear waves on vertical cylinders, Journal of Waterway, Port, Coastal and Ocean Div, 102(WW3): 301-316. Raman, H., N. Jothishankar and P. Venkatanarasaiah (1977) Nonlinear wave interaction with vertical cylinder of large diameter, Journal of Ship Research, 21(2): 120-124. Ramberg, S.E. and J.M. Niedzwecki (1979) Some errors and uncertainties in wave force computations, Proc 11th Offshore Tech Conf Houston, TX, 111:2091-2102. Ramsey, K.A. (1976) Effective measurements for structural dynamics testing, Part II, Sound and Vibration, 10(4): 18-31. Rasband, S.N. (1990) Chaotic Dynamics of Nonlinear Systems. John Wiley & Sons. Rector, R.L. (1954) Laboratory study of equilibrium profiles, Technical Memorandum No. 48, Beach Erosion Board, Government Printing Office, Washington, D.C. Reid, R.O. and C.L. Bretschneider (1953) Surface wave forces on offshore structures, Texas A. andM. Research Foundation Technical Report. Rellich, F. (1969) Perturbation Theory of Eigenvalue Problems, Gordon and Breach Science Publishers, New York. Resch, F.J. and R. Abel (1975) Spectral analysis using Fourier transform techniques, International Journal for Numerical Methods in Engineering, 9:869-902. Rice, S.O. (1954) Mathematical analysis of random noise, Bell System Technical Journal, Vol. 23, 1944, and Vol. 24, 1945. Reprinted in Selected Papers on Noise and Stochastic Processes, ed. by N. Wax, Dover Publishers, Inc., NY:133-204. Richtmyer, R.D. (1978) Principles of Advanced Mathematical Physics, SpringerVerlag, New York, 1:190-221. Rienecker, M.M. and J.D. Fenton (1981) A Fourier approximation method for steady water waves, /. Fluid Mechs, 104:119-137. Riley, N. (1967) Oscillatory viscous flows: Review and extension, Journal of Institute of Mathematics and Its Applications, 3:419-434. Riley, N. (1975) The steady streaming induced by a vibrating cylinder, J. Fluid Mechs, 68:801-812.
Bibliography
901
Roache, G.F. (1970) Green's Function, Van Nostrand Reinhold Company. Roberts, J.B. and M. Gaster (1980) On the estimation of spectra from randomly sampled signals: a method of reducing variability, Proc Royal Soc London, A, 371:235-258. Rodenbush, G. and C.A. Gutierrez (1983) Forces on cylinders in two-dimensional flows, Tech. Progress Report, BRC 13-83, Shell Development Co., Houston, TX. Rosenblatt, M. (1957) A random model of the sea surface generated by a hurricane, Journal of Mathematics and Mechanics, 6(2):235-246. Rosenblatt, M. (1962) Random Processes, Oxford University Press, New York. Rosenblatt, M. and J.W. Van Ness (1965) Estimation of the bispectrum, Annual Mathematical Statistics, 36:1120-1136. Rudnick, P. (1951) Correlograms for Pacific Ocean waves, Proc 2nd Berkeley Symp Math Statistics and Probability, Univ California Press, Berkeley, CA:627-638. Russell, R.C.H. and J.D.C. Osorio (1958) An experimental investigation of drift profiles in a closed channel, Proc 6th Conf Coastal Engrg, Univ Cal, Berkeley, CA: 171-193. Rye, H. (1974) Wave group formation among storm waves, Proc 14th Int. Conf on Coastal Engineering J64-1 %3. Rye, H. (1977) The stability of some currently used wave parameters, Coastal Engineering, 1(1):17—30. Rye, H. (1982) Ocean Wave Groups, Dept. Marine Tech., Norw. Inst. Tech., UR-82-18. Saffman, P.G. (1967) Discussion to nonlinear interactions treated by the methods of theoretical physics (with application to the generation of waves by wind), by K. Hasselmann, Proceedings of the Royal Society of London, Series A, 299:101-103. Santon, L. (1954) Enregistrement graphique de profil d'une houle de laboratoire analyse harmonique, Proc 5th Int. Conf on Coastal Engineering, Grenoble, France: 189-206. Santon, L. and C. Marcou (1953) Enregistrement graphique de profil d'une houle produite en laboratoire, La Houille Blanche, 3:411-417. Sarpkaya, T. (1975) Forces on cylinders and spheres in a sinusoidally oscillating fluid, Journal of Applied Mechanics, Al'.'H-'il. Sarpkaya, T. (1976) Vortex Shedding and Resistance in Harmonic Flow About Smooth and Rough Circular Cylinders at High Reynolds No., Naval PG Sch., Monterey, CA. Sarpkaya, T. (1977) In-line and transverse forces on cylinders in oscillatory flow at high Reynolds number, Journal of Ship Research, 21(4): 200-216. Sarpkaya, T. (1979) Vortex-induced oscillations, J. Applied Mechs., June 46:241-258.
902
Bibliography
Sarpkaya, T. (1989) Computational methods with vortices-the 1988 Freeman scholar lecture, J. Fluid Engineering, Mar., 111:5-52. Sarpkaya, T. and M. Isaacson (1981) Mechanics of Wave Forces on Offshore Structures, Van Nostrand Reinhold Company, 651 p. Sayer, P. (1980) An integral-equation method for determining the fluid motion due to a cylinder heaving on water of finite depth, Proc Royal Society of London, A, 372:93-110. Scheck, F. (1990) Mechanics, Springer-Verlag. Schenck, H.A. (1967) Improved integral formulation for acoustic radiation problems, Journal of the Acoustical Society of America, 44(2). Scheuer, E.M. and D.S. Stoller (1967) On the generation of normal random vectors, Technometrics, 18:278-281. Schlichting, H. (1979) Boundary Layer Theory, 7th Ed., McGraw-Hill Book Co., 817 p. Scott, J.R. (1965) A sea spectrum for model tests and long-term ship protection, Journal of Ship Research: 145-152. Serrin, J. (1959) Mathematical Principles of Classical Fluid Mechanics, Encyclopedia of Physics, Ed. S. Flugge, 8, Springer-Verlag, Berlin, Germany. Seshadri, R. and T.Y. Na (1985) Group Invariance in Engineering Boundary Value Problems, Springer-Verlag, New York. Sharma, J.N. and R.G. Dean (1979) Development and evaluation of a procedure for simulating a random directional second order sea surface and associated wave forces, Ocean Engineering Report No. 20, University of Delaware, Newark, Delaware. Shaw, D.C. (1985) Some asymptotic results in the theory of wave-energy devices, Journal of Engineering Mathematics, 19:79-103. Shemer, L. and E. Kit (1988) Study of the role of dissipation in evolution of nonlinear sloshing waves in a rectangular channel. Fluid Dynamics Research 4:89-105. Shen, H.T. (1977) Discussion of Forces due to nonlinear waves on vertical cylinders, by H. Raman and P. Venkatanarasaiah, / Waterway, Port, Coastal & Ocean, 103(3):406-407. Shenk, N. and D. Thoe (1971) Eigenfunction expansions and scattering theory for perturbations of —A, Journal of Mathematical Analysis and Applications, 36:313-351. Shinozuka, M. (1971) Simulation of multivariate and multidimensional random processes, Journal of Acoustical Society of America, Part 2, 49(l):357-367. Shore Protection Manual (1984). Coastal Engineering Research Center, Department of the Army, Waterways Experiment Station, Vicksburg, Miss, 7:202-7:242. Simiu, E. (1996) Melnikov Process for Stochastically Perturbed, Slowly Varying Oscillators: Application to a Model of Wind Driven Coastal Currents. /. Applied Mechs, 63:429-435.
Bibliography
903
Skjelbreia, L. (1958) Gravity waves: Stokes' third-order approximation, tables of functions, The Engrg Foundation Coun Wave Research, Berkeley, CA:231-251. Sleath, J.F.A. (1972) A second approximation to mass transport by water waves, Journal of Marine Research, 30:295-304. Sleath, J.F.A. (1973) Mass-transport in water waves of very small amplitude, Journal of Hydraulic Research, 11:369-383. Smale, S. (1963) Diffeomorphisms with many periodic points. Differential and Combinatorial Topology. Edited by S. S. Cairns. Princeton University Press: 63-80. Smith, R. and T. Sprinks (1975) Scattering of surface waves by a conical island, J. FluidMechs, 72:373-384. Smithies, F. (1958) Integral Equations, Cambridge University Press. Sneddon, I.N. (1957) Elements of PartialDifferentialEquations, McGraw-Hill Book Company, New York. Sneddon, I.N. (1966) Mixed Boundary Value Problems in Potential Theory, John Wiley & Sons, Inc., New York. Sneddon, I.N. (1972) The Use of Integral Transforms, McGraw-Hill Book Co., N.Y. Solkolnikoff, I.S. and R.M. Redheffer (1966) Mathematics of Physics and Modern Engineering, 2nd Edition, McGraw-Hill, Inc, 752 p. Somerfeld, A. (1949) Partial Differential Equations in Physics, Academic Press, N.Y. Soong, T.T. (1973) Random Differential Equations in Science and Engineering, Academic Press, New York. Spring, B.H. and P.L. Monkmeyer (1974) Interaction of plane waves with vertical cylinders, Proc of the Coastal Engineering Conf, Copenhagen, Denmark, 111:1828-1847. Srokosz, M.A. (1979) The submerged sphere as an absorber of wave power, / . Fluid Mechs, 95:717-741. Srokosz, M.A. (1980) Some relations for bodies in a canal, with an application to wave power absorption, J. Fluid Mechs, 99:145-162. St. Denis, M. (1980) Some comments on certain idealized variance spectra of the seaway currently in fashion, Journal of Ship Research, 24(4):271-278. Stacy, E.W. and G.A. Mihram (1965) Parameter estimation for a generalized gamma distribution, Technometrics, 7(3):349-358. Stakgold, I. (1967) Boundary Value Problems of Math, Phys, I & II, MacMillan Co. Stakgold, I. (1979) Green's Functions and Boundary Value Problems, John Wiley & Sons, New York. Standing, R.G. (1978) Applications of wave diffraction theory, International Journal for Numerical Methods in Engineering, 13:49-72.
904
Bibliography
Stoker, J.J. (1947) Surface waves in water of variable depth, Quarterly of Applied Mathematics, 5(1): 1-54. Stoker, J.J. (1957) Water Waves, Interscience Publications, New York. Stokes, G.G. (1845) On the theories of internal friction of fluids in motion, Transactions of the Cambridge Philosophical Society, London, England, 8:287-305. Stokes, G.G. (1847) On the theory of oscillatory waves, Transactions of the Cambridge Philosophical Society, London, England, 8:441—455. Stokes, G.G. (1880) Supplement to a paper on the theory of oscillatory waves, Mathematical and Physical Papers, Cambridge University Press, 1:314-326. Straub, L.G. and J.B. Herbich (1956) Experimental Studies of Wave Filters and Absorbers, Project Report No. 44, David Taylor Model Basin, Washington, D.C. Streeter, V.L. (1948) Fluid Dynamics, McGraw-Hill Book Co. Stuart, J.T. (1966) Double boundary layers in oscillatory viscous flow, /. FluidMechs, 23(4):673-687. Sulisz, W and R.T. Hudspeth (1993) Second-order wave loads on planar wavemakers, Journal of Waterways, Port, Coastal and Ocean Engineering, 119(5):521-536. Suyehiro, K. (1920) Yawing of ships caused by oscillation amongst waves, Transactions of the Institution of Naval Architects, 62:93-101. Takayama, Y. and Y. Goda (1973) Wave forces on a semi-submerged vertical cylinder of elliptical shape, Coastal Engineering in Japan, 16:147-164. Tanaka, Y and R.T. Hudspeth (1988) Earthquake Response of Circular Cylindrical Structures in Water, Earthquake Engineering and Structural Dynamics, 16:99-119. Tanaka, Y. (1988) Irregular Points in Wavemaker Boundary Value Problem, Ph.D. Thesis, Oregon State University. Tayfun, M.A. (1980) Narrow-band nonlinear sea waves, Journal of Geophysical Research, 80(24): 1548-1552. Tayfun, M.A. (1983) Frequency analysis of wave heights based on wave envelopes, Journal of Geophysical Research, 88(12):7573-7587. Tayfun, M.A., C.Y. Yang and G.C. Hsiao (1972) Nonstationary spectrum analysis of ocean waves, Proc 13th Coastal Engineering Conference, Chapter 11, 1:251-269. Tayfun, M.A., C.Y. Yang, V. Klemas and H. Wang (1975a) Analysis of inhomogeneous wave number spectra, Journal of Geophysical Research, 80(24): 3469-3473. Tayfun, M.A., C.Y. Yang and G.C. Hsiao (1975b) Optimal design for wave spectrum estimates, Journal of Geophysical Research, 85(3):3469-3473. Taylor, R.E. and J.B. Waite (1978) The dynamics of offshore structures evaluated by boundary integral techniques, Int J for Numerical Methods in Engineering, 13:73-92.
Bibliography
905
Theocaris, P.S. (1981a) Numerical solution of singular integral equations: methods, Journal of the Engineering Mechanics Division, 107(EM5):733-752. Theocaris, P.S. (1981b) Numerical solution of singular integral equations: applications, Journal of the Engineering Mechanics Division, 107(EM5):753-771. Theocaris, P.S. and N.I. Ioakimidis (1979) A method of numerical solution of Cauchytype singular integral equations with generalized kernels and arbitrary complex singularities, Journal of Complex Physics, 30:309-323. Theocaris, P.S., N.I. Ioakimidis and J.G. Kazantzakis (1980) On the numerical evaluation of two-dimensional principal value integrals, Short Communications, John Wiley:529-534. Thomas, G.B. Jr. (1965) Calculus and Analytic Geometry, Addison-Wesley, Reading, Massachusetts:473^182 & 490-500. Thomas, J.R. (1981) The absorption of wave energy by a three-dimensional submerged duct, / . FluidMechs, 104:189-215. Thompson, E.F. and W.N. Seelig (1984) High wave grouping in shallow water, Journal of Waterway, Port, Coastal and Ocean Division, 110(2): 139-157. Thornton, E.B. and R.J. Calhoun (1972) Spectral resolution of breakwater reflected waves, Journal of Waterway, Port, Coastal and Ocean Division, 98(WW4): 443-460. Tick, L.J. (1959) Differential equations with frequency-dependent coefficients, Journal of Ship Research, 3(2):45-46. Tick, L.J. (1963) Nonlinear probability models of ocean waves, Ocean Wave Spectra, Englewood Cliffs, New Jersey, Prentice-Hall, Inc.: 163-169. Titchmarsh, E.C. (1962) Eigenfunction Expansions Associated With Second-Order Differential Equations, Part One, Oxford University Press. Tricomi, F.G. (1957) Integral Equations, Interscience Publishers, Inc., New York. Tuah, H. and R.T. Hudspeth (1982) Comparisons of numerical random sea simulations, Journal of Waterway, Port, Coastal and Ocean Division, 108(WW4): 569-584. Tuah, H. and R.T. Hudspeth (1985a) Finite water depth effects on nonlinear waves, Journal of Waterway, Port, Coastal and Ocean Engineering, 111(WW2): 401-416. Tuah, H. and R.T. Hudspeth (1985b) Nondeterministic Wave Force on Fixed Small Vertical Piles, Probabilistic Offshore Mechanics, Progress in Engineering Science; Vol. 7, Ed. P.D. Spanos, CML Publications, Southampton, England: 49-54. Umeki, M. and T. Kambe (1989) Nonlinear dynamics and chaos in parametrically excited surface waves. Journal of Physical Society of Japan. Fluid Mech, 212:140-154. Unluata, U. and C.C. Mei (1970) Mass transport in water waves, Journal of Geophysical Research, 75(36):7611-7618.
906
Bibliography
Ursell, F. (1950) Surface waves on deep water in the presence of a submerged circular cylinder, Proceedings of Cambridge Philosophical Society, 46:141-152. Ursell, F. (1952) Edge waves on a sloping beach, Proceedings of the Royal Society of London, London, England, Series A, 214:79-97. Ursell, F. (1953a) Mass transport in gravity waves, Proceedings of the Cambridge Philosophical Society, London, England, 49:145-150. Ursell, F. (1953b) The long-wave paradox in the theory of gravity waves, Proceedings of the Cambridge Philosophical Society, 49:685-694. Ursell, F. (1959) Forced small-amplitude water waves: a comparison of theory and experiment, J. FluidMechs, 7:33-52. Ursell, F. (1964) The decay of the free motion of a floating body, J. Fluid Mechs, 19:305-319. Ursell, F. (1981) Irregular frequencies and the motion of floating bodies, J. Fluid Mechs, 105:143-156. Ursell, F., R.G. Dean and Y.S. Yu (1960) Forced small amplitude water waves: A comparison of theory and experiment, J. Fluid Mechs, 7(3):33-52. U.S. Beach Erosion Board (1941)^4 Study ofProgressive Oscillatory Waves in Water, Technical Report No. 1, Government Printing Office, Washington, D.C. Valent, RJ. (1979) Coefficients of Friction Between Calcareous Sand and Some Building Materials, and Their Significance, Technical Note, No. N-1542, Civil Engineering Laboratory, Naval Construction Battalion Ctr, Port Hueneme, CA, 25 p. Van Atta, C.W. (1979) Inertial range bispectra in turbulence, Phy of Fluids, 22:1440-1442. Van Dorn, W.G. (1966) Boundary dissipation of oscillatory waves, /. Fluid Mechs, 24(4):769-779. Van Dyke, M. (1964) Perturbation Methods in Fluid Mechanics, Academic Press, N.Y. Vanmarcke, E.H. (1972) Properties of spectral moments with applications to random vibrations, Journal of Engineering Mechanics Division, 98(EM2): 425—446. Vanmarcke, E.H. (1983) Random Fields: Analysis and Synthesis, MIT Press, Cambridge, Massachusetts, 382 p. Van Oorschot, J.H. and F. Koopmans (1976) Random wave generation for research on immersed marine vehicles, 11th Symposium Naval Hydrodynamics, London, England. Venezian, G. (1980) Discussion of Direct solution of wave dispersion equation, by J.N. Hunt, Journal of Waterway, Port, Coastal and Ocean Division, 106(WW4): 501-502. Venkatanarasaiah, P. (1978) Discussion of Nonlinear wave forces on large offshore structures, by M. de St. Q. Isaacson, J Waterway, Port, Coastal and Ocean, 104(WW2).
Bibliography
907
Verhulst, F. (1990) Nonlinear Differential Equations and Dynamical Systems, Springer-Verlag. Von Schwind, J.J. and R.O. Reid (1972) Characteristics of gravity waves of permanent form, Journal of Geophysical Research, 77(3):420-433. Vugts, J.H. (1968) The hydrodynamic coefficients for swaying, heaving, and rolling cylinders in a free surface, International Ship-Building Progress, 15(167): 251-276. Wang, C.Y. (1968) On high-frequency oscillatory viscous flows, / . Fluid Mechs, 32(l):55-68. Wang, H. and W-C Yang (1980) A similarity model in the surf zone, Proceedings of the 17th Conference on Coastal Engineering, Sidney, Australia. Wang, H., T. Sunamura and RA. Huang (1982) Drift velocity of the wave breaking point, Coastal Engineering, The Netherlands, 6:121-150. Webber, N.B. and CD. Christian (1974) A programmable irregular wave generator, Proceedings of the 14th Conference on Coastal Engineering, I(17):321—339. Weggel, J.R. (1972) Maximum breaker height, Journal of Waterway, Port, Coastal and Ocean Engineering, 98(4):529-548. Wehausen, J.V. (1971) The motion of floating bodies, Annual Review of Fluid Mechanics, 3:237-268. Wehausen, J.V. (1980) Perturbation methods in diffraction, Journal of Waterway, Port, Coastal and Ocean Engineering, Technical Note, 106(WW2):290-291. Wehausen, J.V. and E.V Laitone(l 960) Surface waves in Handbook of Physics, Fluid Dynamics III, Springer-Verlag, IX:446-757. West, B.J., K.M. Watson and J.A. Thomson (1974) Mode coupling of ocean wave dynamics, Physics of Fluids, 17:1059. West, B.J., J.A. Thomson and K.M. Watson (1975) Statistical mechanics of ocean waves Journal ofHydronautics, 9(l):25-31. Wheeler, J.D. (1969) Method for calculating forces produced by irregular waves, Preprints 1969 Offshore Technology Conference, 1:83-94. Whitham, G.B. (197'4) Linear andNonlinear Waves, John Wiley & Sons, New York. Whittaker, E. and G. Watson, (1927) A Course in Modern Analysis, 4th ed., Cambridge University press, 608 p. Wiegel, R.L., K.E. Beebe and M. Moon (1957) Ocean wave forces on circular cylindrical piles, Journal of the Hydraulics Division, 83(HY2):89-116. Wiener, N. (1958a) The Fourier Integral and Certain of its Applications, Dover Publications, Inc., New York. Wiener, N. (1958b) Nonlinear Problems in Random Theory, MIT Press, Cambridge, MA. Wiener, N. (1964) Time Series, M.I.T Press, Cambridge, MA. Wiggins, S. (1988) Global Bifurcations and Chaos, Springer-Verlag. Wiggins, S. (1990) Introduction to Applied Nonlinear Dynamical Systems and Chaos, Springer-Verlag.
908
Bibliography
Wiggins, S. and P. Holmes (1987) Homoclinic orbits in slowly varying oscillators, SIAMJ. Math Anal. 18:3:612-629. Williams, A. N. and W. G. McDougal (1989) Hydrodynamic analysis of variable draft spiral wavemaker, Ocean Engineering, 16(4):401—410. Wilson, EX., K.J. Bathe and W.P. Doherty (1974) Direct solution of large systems of linear equations, Computers and Structures, 4:363-372. Wylie, C.R. and L.C. Barrett (1982) Advanced Engineering Mathematics, McGrawHill Book Co., New York, 1103 p. Yamamoto, T. and J.H. Nath (1976) High Reynolds number oscillating flow by cylinders, Proceedings of the 15th Conference on Coastal Engineering, 111:2321-2340. Yih, C.S. (1969) Fluid Mechanics, McGraw-Hill, New York. Young, I.R. and R.J. Sobey (1980) Discussion of direct solution of wave dispersion equation, by J.N. Hunt, J. Waterway, Port, Coast and Ocean Div, 106(WW4): 499-501. Yu, Y.S. and F. Ursell (1961) Surface waves generated by an oscillating cylinder in water of finite depth, / . Fluid Mechs, 11:529-551. Yuen, H.C. and B.M. Lake (1975) Nonlinear deep water waves: theory and experiment, Physics of Fluids, 18(8):956-960. Yuen, H.C. and WE. Ferguson, Jr. (1978a) Relationship between Benjamin-Feir instability and recurrence in the nonlinear Schrodinger equation, Physics of Fluids, 21:1275-1278. Yuen, H.C. and WE. Ferguson, Jr. (1978b) Fermi-Posti-Ulam recurrence in twospace dimensional nonlinear Schrodinger equation, Physics of Fluids, 21: 2116-2118. Yuen, H.C. and B.M. Lake (1979) A new model for nonlinear wind waves, Part 2, theoretical model, Studies in Applied Mathematics, 60:261-270. Yuen, H.C. and B.M. Lake (1980) Instabilities of waves in deep water, Annual Review Fluid Mechanics, 12:303-334. Zelen, M. and N.C. Severo (1968) Probability functions Handbook of Mathematical Functions, ed. by M. Abramowitz and LA. Stegun, Dover Pub Inc., NY: 925-995. Zenkovich, V. (1967) Processes of Coastal Development, Oliver and Boyd Company, London, England. Zimmerman, J.T.F. (1979) On the Euler-Lagrange transformation and the Stokes drift in the presence of oscillatory and residual currents, Deep-Sea Research, 26A:505-520. Zimmerman, M., R. Hudspeth, J. Leonard, J. Tedesco and L. Borgman (1986) Dynamic behavior of deep-ocean pipeline, / Waterway, Port, Coast & Ocean Div, 112(2): 183-199. Zopf, D.O., H.C. Creech and W.H. Quinn (1976) The wavemeter: a land-based system for measuring nearshore oceans waves, MTS Journal, 10(4): 19-25.
Author Index
Aagaard, 523, 535, 537 Abraham, 465 Abramowitz, 19, 24, 31, 214 Agarwal, 608, 610 Airy, 85 Anand, 608 Andrews, 8 API RP2A, 503, 846 Arfken, 7, 286 Arnold, 423, 425, 468, 473, 474 Atkinson, 301, 306, 540, 555, 557 Averbeck, 284, 294, 295, 297, 298,300
Badd, 835, 841 Barcilon, 9, 20, 26, 28, 29, 30 Barnard, 268 Barrett, 17, 23 Batchelor, 66 Battjes, 792 Becker, 372, 428, 429 Bendat, 720, 730, 731, 738, 759, 760,792,797,825,831 Benettin, 425, 485 Benjamin, 432, 437 Benton, 48, 50, 165, 261, 274 Benzi, 395
Berge, 424, 426 Bettess, 711 Biesel, 154 Bily, 809 Black, 688, 694 Bogoliubov, 846 Booij, 251 Borgman, 247, 755, 756, 759, 808, 809, 813, 814, 822, 832, 838, 842 Bowline, 268, 423, 424, 467, 483 Bracewell, 792, 794 Bretschneider, 400, 401, 750 Brigham, 738, 856, 857 Bukoveczky, 809
Cacko, 809 Calhoun, 803 Carnahan, 301 Carpenter, 78, 81-83, 498, 517, 521,523,528,532,533, 538, 539, 543, 545, 546, 551,558,600,601,602, 610 Carrier, 96, 99, 159, 260, 263, 271,273,279,284,286, 288, 295, 296 909
910
Cartwright, 748, 749, 780, 787, 789, 790, 792, 809 Chakrabarti, 149, 251, 538, 539, 545,603,604,752 Chaplin, 395 Chen, 154, 770, 769 Chertock, 698 Chow, 82 Chu, 372, 374 Chua, 484 Churchill, 296 Clough, 843, 844, 848,850,851 Cokelet, 280, 395 Courant, 292, 688, 693, 700 Cox, 758 Crandall, 128, 645, 646, 659
Daily, 65 Dalrymple, 140, 143, 181, 204, 230,251,252,255,260, 262,264,266,310-312, 320,361,394,395,543, 544,642, 643 Darbyshire, 750 Davenport, 833, 837 Davey, 372 Dean, 60, 71, 104, 121, 126, 140, 143,145,181,204,230, 310-312,320,361,394, 399,401,492,502,517, 518,523,535-538,540, 542-544,547,548,601, 605, 612, 642 Delves, 698 DeSilva, 280, 301, 306
Author Index
Desloge, 487, 488 Dugundji, 792
Eagleson, 60, 71, 104, 121, 126, 605 Elgar, 792 Enochson, 759, 823 Eshbach, 610
Fadel, 425, 446, 450,457 Fassardi, 791, 801 Fenton, 315, 325, 332, 395, 517, 693 Feshbach, 15, 48, 49, 54, 55, 142, 153, 165,211,255,260, 261,272,274,279-281, 284,286,292,319,688, 689,690,691,693,694, 706 Foster, 846 Friedman, 31, 326 Fuchs, 669, 672, 677, 679 Funke, 242, 797
Galvin, 155,403 Garabedian, 280, 281, 284, 292, 377 Garrison, 82, 83, 513-515, 521-524,530-534,601, 602,715,716 Gaughan, 404 Gilbert, 192, 195, 199, 201 Glowinski, 301,398
911
Author Index
Goda, 751, 752, 753, 757, 792, 797, 803-806 Goldstein, 442, 446 Gradshteyn, 476, 775 Grecco, 809, 832, 843, 844, 848-850, 854-856 Guckenheimer, 470 Guenther, 48-51, 280, 283, 292, 300,377,425, 429, 444, 487, 488, 523, 688, 694, 695 Gutierrez, 523
Harleman, 65, 502, 601 Havelock, 149, 154, 201, 405 Hayashi, 539, 545, 546, 558, 601, 602 Heideman, 524 Helleman, 466 Henderson, 426 Hibbard, 525 Hilbert, 292, 683, 693, 700 Hildebrand, 13-15, 18, 21, 22, 24-26,31,34,35, 39-42, 45, 48, 49, 55, 69, 96, 98, 99, 159, 165, 177, 180, 204, 208,210,214,216,255, 258, 260, 261, 271-274, 285,286,292,395,398, 430, 638, 654, 668, 676, 690, 693-695, 697, 698 Hino, 832, 838, 842 Hochstrasser, 31
Hoffman, 715, 716, 772 Holmes, 423, 444,454, 456, 461,470 Huang, M.-C. 76, 396, 397, 399, 517,605,606,611-615, 706,707,715 Huang, N. E. 750 Hubbard, 523 Hudspeth, 154, 155, 161-164, 242, 247, 249, 268, 364, 365, 367, 370, 395, 396, 397, 399, 405, 410, 422, 423, 471, 486, 487,517,534,539,540, 545, 548, 550, 555, 605, 606,611-615,662,693, 706,715,745,769,770, 791,792,796-800,808, 809,812,813,814, 819-822, 832, 833, 837-839, 841-844, 848-850, 854-856 Hughes, 149 Hunt, 108 Hyun, 154
Ince, 96, 98, 99, 159,208,210, 260, 271, 273 Ingard, 142,211,311,317,319, 672 Ioakimidis, 698 Ionnaou, 155 Ippen, 530, 531
912
Isaacson, 79, 497, 538, 539, 545, 662, 677, 693, 719, 787, 790 ISSC, 750 ITTC, 750 Iwagaki, 799
John, 170, 178, 280,624, 637, 639, 640, 642, 688, 690-693, 703 Jones, 440,441,444 Jonsson, 251 Jordan, 424 Joseph, 280, 301
Kambe, 461 Kambi, 485 Karim, 65 Karst, 772 Keating, 155 Kendall, 780 Kennard, 154 Kershaw, 306 Keulegan, 78, 81-83, 498, 517, 521,523,528,532,533, 539,543,545,546,551, 558, 600, 602, 610 Kim, 523, 525 Kimura, 803, 805, 806 Kinsman, 719, 733, 750 Kirby, 372 Kit, 268 Kobayashi, 806 Komar, 404
Author Index
Kravtchenko, 281 Krylov, 846 Laitone, 280, 366, 404,406, 432 Lamb, 59-61, 65, 69, 70, 510, 688, 693, 695-698 Landau, 70 Lee, 48, 280, 283, 292, 300, 377, 688, 694, 695 Leonard, 154, 706 Lewy, 281 Lichtenberg, 424, 441, 442, 468, 486 Lieberman, L.N., 65 Lieberman, M.A., 424, 441, 442, 468,486 Lighthill, 7, 338, 721,723 Lin, 281 Longuet-Higgins, 337, 361, 748, 749,751,757,780,787, 789, 790, 792, 809 Lorentz, 530, 531,846 Luke, 372, 424, 428, 429, 837 Lyons, 610 MacCamy, 672, 669, 677, 679 MacKenzie, 719 Mansard, 797 Mardia, 757 Marquardt, 108, 398, 399, 540, 744 Marsden, 465 Mase, 799 Maull, 556,559, 601,602 McCowan, 404
913
Author Index
Medina, 791, 792, 796-801, 803, 806
Newman, 637, 688 Norris, 426
Mei, 89, 140, 204, 315, 317, 323, 325,332, 372, 374, 375, Oates, 48 383,497, 637, 639, 643, Oberhettinger, 835, 841 672,677,681,688
Melnikov, 423 Miche, 402 Michell, 402 Mihram, 780 Miles, 372, 426, 428, 429, 437 Miller, B. L. P., 532 Miller, J. C. P., 836 Milliner, 556, 559, 601,602 Milne-Thompson, 69, 279, 280, 284-286, 288, 294, 296-298, 402 Mitsuyasu, 750, 756, 757 Moon, 484 Morison, 497, 523, 525 Morse, 15, 48, 49, 54, 55, 142, 153, 165,211,255,260, 261,272,273,279-281, 284,286,292,311,317, 319,672,688-691,693, 694, 706 Moskowitz, 750 Munk, 404, 758
Nath, 522, 523 Nayfeh,5,310,316,320,322, 371,405,408 Nekrasov, 280 Neumann, 750
OCEAN IV proceedings, 523 Ochi, 753, 779, 780 Ochi-Hubble, 753 Olver, 5 Orum, 425, 446, 487 Otnes, 759, 823 Pade, 108-110 Papoulis, 760, 761, 764, 765, 772, 835 Parker, 484 Patel, 155 Pearson, 96, 99, 159,260, 271,273 Penzien, 843-845, 848, 850, 851 Peters, 637 Phillips, 70, 156, 363 Piersol, 720, 730, 731, 738, 759, 760,792, 797, 823, 825, 831 Pierson, 750 Potynody, 608, 609 Radder,251,259 Rasband, 425, 468, 484 Redheffer, 108 Reid, 394, 400, 401 Rice, 749, 780, 792, 809, 813, 814 Rienecker, 395
914
Rodenbush, 523 Root, 833, 837 Rye, 792 Ryzhik, 476, 775
Sarpkaya, 497, 522, 523, 532, 534, 538, 539, 545, 600, 787,789 Scheck, 428, 446 Schlichting, 53, 65, 498, 508 Schwerdtfeger, 425, 429, 444,487 Scott, J. C , 432, 437 Scott, J. R., 750 Seelig, 797 Sequet, 154 Severe, 760, 767-769, 810 Shemer, 268 Simiu, 479 Skovgaard, 251 Slotta, 395 Smale, 479 Smith, P., 424 Smith, R., 251 Sobey, 109, 110 Sokolnikoff, 108 Sprinks, 251 Stacy, 780 Stakgold, 32,33,99, 184, 261,292,688,700,703 Stegun,20, 24, 31,214 Stewart, 780 Stoker, 281, 632, 637, 688
Author Index
Stokes, 82,310-313,315-318, 323, 324, 345, 346, 402, 418,427 Sulisz, 164, 364, 365, 367, 370, 405, 410, 422 Suzuki, 757, 803-806
Takenouchi, 539, 545, 546, 558, 601,602 Tanaka,281,282,284 Theocaris, 698 Thomas, 542, 544 Thompson, 797 Thornton, 803 Tuah, 792, 809, 812-814, 832, 833, 837-839, 841-843
Umeki,461,485 Ursell, 82, 155,688
Valent, 608, 609 Vanmarcke, 741, 742, 746, 747, 749,751 Vaughan-Makela, 371 Venezian, 109 Verhulst, 484 Vledder, 792 Von Schwind, 394
Watson, 296 Webber, 155 Weggel, 404
915
Author Index
Wehausen, 149, 280, 366, 406, 432, 637, 688 Whitham, 372, 429 Whittaker, 296 Wiener, 734 Wiggins, 423, 425,444, 463,465, 468, 472^174, 476^182 Wylie, 17, 23
Yamamoto, 522, 523 Young, 109, 110
Zelen,760, 767-769, 810 Zienkiewicz, 706, 711, 715 Zimmerman, 605 Zopf, 337
Subject Index
2-D Cartesian coordinates, 86, 155,251,271,372,509, 544,576, 663, 667 2-D channel, 364 3-D circular cylindrical coordinates, 228, 231, 233, 236, 287, 298, 663, 667,673,689 accelerating cylinder in still fluid, 510,513 action/angle transformation, 449 added mass, 189, 231, 497, 509, 510,562,569,597 added mass coefficient, 192, 199, 235,237,239,241,510, 514,652 added mass/mass moment of inertia, 561 amplitude/phase method, 548 angle of rotation, 542 angle sum identities, 19 Archimedes buoyant force, 574, 577 articulated tower, 563, 615 associated Laquerre polynomials, 29 auxiliary function, 34
auto-correlation function, 734, 738,771,859,861 auto-covariance function, 734, 816, 839, 842, 858, 859, 862 auto-spectral density function, 734,861 average rate of work, 185 average or mean radian wave frequency u>, 745 average value, 761 average wave amplitude, 774 averaged system, 461 averaging method, 479 band pass filter, 806 Battjes surf similarity parameter, 403 beach slope, 607 Bernoulli equation, 70, 122, 506, 510 Bessel functions, 21, 25, 668 asymtotic approximations, 24 derivatives, 25 first kind, 21 second kind, 21, 23 third kind, 22, 214 Wronskians, 26 917
918
bilinear transformations, 288 binomial coefficient, 18 binomial expansion, 17, 316, 346, 432, 483 body forces, 69 body weight force/moment, 656 boundary conditions, 91, 96, 99, 156, 161, 163-168, 171, 178,203,210,287,299, 317,624,638,643,708 break water damage, 802 breaking waves, 400 kinematic parameter 7TK, 401 dynamic parameter JTD, 401 Buckingham Pi theorem, 84 buoyant moment, 569 canonical transformation, 447, 451,454,487,488 Cauchy parameter, 81 Cauchy-Riemann equations 286 CEDEX-CEPYC, 249 central moment, 763 characteristic equation, 45—47 characteristic function irx(oo), 764,834 characteristic or eigenfunction, 15, 50,52 characteristic wave frequency Qk, 797 chaos, 423 Chebyshev polynomials, 28 chi-squared variable, 759, 792, 866 circular cylinder, 509, 833
Subject Index
circular cylindrical coordinates, 10,12,24,201,287,298, 667,673,689 circular wavemakers, 201 coherence function, 732 coincident spectral density function, 731 complementary error function, 9, 775 completeness, 30, 31 conditional wave simulations, 813 numerical instabilities, 821 confidence intervals, 759, 829-830 conformal mapping, 280 conic section equation, 542, 544 conjugate gradient method, 301, 306 conservation of mass, 55 conservation of wave period, 104 conservation of wave phase, 140 conservation principle, 140 contact line boundary condition, 430, 432, 437 continuity field equation, 55 continuum, 55, 57, 88, 159, 317, 510,629 convolution kernel, 37, 378, 699, 857 copolar half-period trio, 297 copolar trio, 298 Coulomb friction, 605, 607, 609 covariance function, 731, 832 Crank-Nicholson, 301 critical damping, 37, 41, 561, 571 critically damped system, 47
919
Subject Index
cross-correlation function, 734 cross-covariance function, 733, 739,839,852,859 cross-spectral density function, 731, 734, 852, 860 cross waves, 423 cumulative distribution function (cdf) P(x), 760, 766, 772, 843 curl vector operator, 10 current, 375, 396, 574, 576, 607
d'Alembert's paradox, 511 damped dynamic systems, 46 critically, 47 over, 47 under, 46 damped harmonic oscillator, 37, 41,46,74,571,844,848, 851 damped natural frequency, 37, 41, 46,572,583,587,851 damping critical, 37,41,561,571,582, 597 ratio, 37,41,47, 561,573, 587, 851 viscous, 560, 570 damping forces, 444, 459 data condition, 538, 540 Dean eccentricity parameter, 538, 544,551 Dean error ellipse, 540, 547, 601 Dean stream function, 394, 612
deep-water, 103, 111, 117, 121, 124, 138, 338, 345, 402, 690,692,713 definition fluid, 53 Fourier coefficients, 720 DeMoivre's theorem, 19 depth-integrated drag force/moment, 504, 516, 519,520 depth-integrated inertia force/moment, 504, 516, 519 depth-intergrated total force, 504, 515,517 deterministic spectral amplitude simulation (DSA), 811 detuning parameter, 456 deviatoric stress, 66 diffraction boundary value problem, 627, 632, 672, 686 diffraction wave potential, 676, 682, 707 dimensional analyses, 81 dimensional boundary value problem (BVP) for LWT, 87, 92, 94, 97 dimensionless parameter, 81 dipole source density, 698 Dirac delta distribution, 7, 184, 755 directional wavemaker, 251 discus buoy, 715 displacement potential, 272 dissipation, 429
920
dissipative phase space, 477 distribution of the maxima, 780 domain mapping, 279, 301 drag force, 503, 505, 506, 516, 519,530,551,577,608, 680 drag force coefficient, 507, 518, 533,537,538,552,568, 606 Duhamel convolution integral, 35, 851,857 dynamic body boundary condition (DBBC), 625, 635, 643 dynamic magnification factor, 39, 562, 573, 584, 588, 599, 717,852 dynamic pressure, 87, 186, 189, 221,223,231,236,323, 496,511,579,677,709 dynamic response, 579, 620, 710, 716 dynamic similitude, 81 dynamic viscosity, 82, 496 dynamical constraint, 436
edge constraint boundary condition, 432, 437 eigenfunction solutions, 50, 96, 102,165,170,178,261, 365,367,411,675 eigenvalues, 96 evanescent, 176 propagating, 106 elliptical cylinders, 506
Subject Index
energy density, 121, 127-137, 140, 363 energy flux conservation principle, 121, 135, 138, 187,228,243,248,633, 797, 798 ensemble average, 719, 790, 828 envelope exceedance coefficient a(O,791,801 equivalent linearization, 846, 848 error function, 9, 767, 775 Euler parameter, 81 Euler's constant, 20, 22, 677 Euler's equation, 70, 512 Euler's formula, 19, 150 Eulerian field, 11, 55, 76, 86, 113, 121, 135, 151,345,566, 580, 620, 625, 630, 643 Eulerian fluid field, 151, 499, 624 Eulerian velocity field, 113, 369, 521,541 excess of kurtosis, 767, 838 exciting force, 192,497, 582, 583, 592,619 exciting force/moment, 435, 560-562, 595, 626, 635, 647, 649, 684, 687, 707 exciting heave load, 586, 587, 596 exciting load moment, 569, 571, 573,594, 596-598 expectation operator £ [ • ] , 719, 771,781,834,839-841 expected value, 761, 802 exponential function, 18, 19 external constraining forces/moments, 657, 711
921
Subject Index
FFT coefficients F{m), 722, 725-729, 732, 736-740, 759,805, 813-830 FFT normalizing constant CN, 721,723-729,732,736 finite element method (FEM), 706,711,715 finite Fourier transform (FFT), 161, 204, 245, 252, 269, 566,725, 803 first order linear ordinary differential equations, 34, 39,263, 462 fixed cylinder in accelerating flow, 509, 510, 513 fixed-free beam, 515 fixed tower, 569, 571 Floquet oscillator, 424, 426 Floquet parametric forcing, 426, 437, 441^144, 449, 457, 487 Floquet parametric resonance, 454, 456 Floquet stability, 424 flow separation, 495^497, 511, 600 fluid control volumes, 54 fluid definition, 53 flux of momentum, 499 force/moment load exciting, 560-562 fixed body, 560, 569, 571, 582, 586, 594, 647, 649, 684, 709 restoring, 647, 650, 652, 710
form drag, 506 Fourier analyses, 524, 528, 616, 618,720 Fourier transform, 721, 731, 764, 842,858 Fredholm alternative, 377 Fredholm integral, 292, 688, 693-700, 706 free boundary Eulerian fluid field, 624, 625 Lagrangian solid body, 624, 625 two Eulerian fluid fields, 624 frequency dispersion equation, 102, 105, 107, 137, 166, 795,815 frequency domain input-output transfer functions, 854, 858, 861-865 frequency parameter, 81 Froude parameter, 72, 78, 81-83, 616 Froude-Kriloff, 497, 566-567, 662-663, 667, 669, 674, 679,680,682
Galilean transformation, 280, 311, 317,319,389,396 Gamma functions, 8, 9, 29, 742, 747,756, 775, 841 generalized incomplete, 9, 751, 775 Gamma pdf, 780 Gauss divergence theorem, 48, 49,431
922
Gaussian (Normal) probability distribution, 748, 752, 759, 760, 763-770, 765, 780,781,786,792,809, 810,813,816,824,831, 837,838,843, 847 generalized Herglotz algorithm, 446,450,451,453,456, 457, 487^191 generalized incomplete Gamma function, 9, 751,775 generalized Melnikov method (GMM), 421, 423, 426, 463,473-476, 479, 482, 483 generalized Rayleigh distribution, 748, 760, 770-780, 786-790 generating function, 23, 24, 27, 28,30,31,446,451,455, 457, 485, 487, 490^192 generic characteristic wave frequency Qk, 747 generic planar wavemaker, 155 geometric similitude, 81 Goda dimensionless spectral peakedness parameter Qp, 751 Goda-JONSWAP variance-preserving spectrum, 752, 797, 820 gradient theorem, 48 gradient vector operator, 10 Gram-Schmidt process, 395 Green's identities first, 49, 430
Subject Index
second, 49, 292 third, 49 Green's function, 292, 688, 699, 715 2D, 689-691 2D wavemaker, 699-706 3D, 691-696
Hamiltonian, 86, 128, 424, 426, 441^44,446-449,452, 457, 461, 464-474, 482, 490,492 Hamilton's equations, 426, 446, 461 Hamilton's principle, 427,429, 433,435, 444 Hamilton-Jacobi transformation, 454,472 Hankel functions, 22, 214, 676 first kind, 22, 214, 676, 693 second kind, 22, 214, 678 hard-rock, 613 Haskind's theorem, 682 heave mode, 574, 585, 588, 637 heave mode load, 585, 589, 591, 595 heave-roll moment load, 591, 593, 595,596 heave stiffness, 585, 590 Heaviside step function, 7, 161, 196, 210, 226, 245, 254, 287,294,407,411,732, 750, 772-774, 801 Helmholtz equation, 255, 259 Herglotz algorithm, 425, 487
Subject Index
Herglotz auxiliary functions, 450, 455^157, 489, 491 Hermite polynomials, 30 heteroclinic orbits, 425^27, 444, 463, 465^170, 483 Hilbert transform, 832, 792-795, 797 analytic function, 793 envelop function, 793, 832 envelope spectral density function, 796 mean celerity, 796 mean group velocity, 796 orbital velocity, 794 phase function, 793 wave height function, 793 Hooke's law, 53 hurricane CARLA, 745, 760, 794, 823-832, 843 hyperbolic identities, 18-20 hyperbolic invariant manifold, 463,465-468,479,481, 483 hyperbolic saddle points, 425, 427,464-465, 467, 469 hypergeometric function, 841
impulse response function, 37, 378,688,689 incomplete Gamma function, 9, 751,755,776 indicial equation, 40-42 inertia, 53, 72, 73, 499, 501, 515, 518,541
923
inertia coefficient, 199, 496, 509, 511,513,514,523,534, 537,541,550,553,586, 587,601,606,679 inertia forces, 58, 59, 62, 72-74, 76,81,194,496,498,503, 505,508,514,515,519, 521,526,551,565,605, 607,608 inertia matrix, 659 inertial coordinate axis, 102, 103, 169,280,311,317,363, 372,574,576,621,623, 638,640 initial conditions, 34, 36, 302, 305, 323-324, 355, 465, 485,486 in-line force, 600, 636 instantaneous phase function, 793 integrating factor p(x), 34, 35, 39, 264 Iribarren number, 403 irregular boundary points, 280-282, 284, 293, 295
Jacobi-Anger expansion, 23, 668, 674 Jacobi symbol, 7 Jacobian, 288, 450, 455, 489, 491, 733,771,782 Jacobian elliptic function, 294-298 Jacobian elliptic integral, 296 joint probability density function, 781,834
924
KAM nondegeneracy, 425 KAM theorem, 425,473^74,480 KAM tori, 473, 474, 476 Kelvin function, 23, 214, 219, 693 Keulegan-Carpenter data, 523, 528, 532, 539, 545 Keulegan-Carpenter parameter, 78-79,81,498,517,521, 528, 532, 533, 538, 539, 543,545,551,559, 600-602, 610 kinematic body boundary condition (KBBC), 625, 627-636, 638, 642-643, 672 kinematic radiation boundary condition (KRBC), 157, 168, 169, 176, 177,205, 244, 270, 275-277, 280, 283,287-291,298,366, 406, 410, 620, 632, 635, 702,709,711 kinematic similitude, 81 kinematic wavemaker boundary condition (KWMBC), 156,182,203,265,270, 287 kinetic energy, 127, 130-134 Kronecker delta function, 7, 52, 178,184,262,275,300, 368, 448,488 kurtosis /X4(x), 762, 766, 767, 838
Lagrange identity, 51 Lagrangian constraints, 394
Subject Index
Lagrangian coordinate axes, 501 Lagrangian density function, 424, 428^133, 437^45 Lagrangian dynamic equations, 659 Lagrangian multipliers, 398 Lagrangian particle, 11 Lagrangian solid body, 72, 74-79, 151,162,198,210,233, 235,237,239,241,503, 560, 603, 620, 636, 640, 643, 646, 682, 845 Lagrangian water particle displacements, 113-115, 118-120 Lame constants, 65 Landau order symbols, 5 Laplacian operator, 11 Laplace transforms, 835, 836, 841 Laplace's equation, 49, 88, 94, 95, 98, 159,215,286,317, 510,629,634,700,708 lateral stability criterion, 605, 607,612 least-squares method, 524, 527, 529,531,535,540,618, 676,721,743,745 Legendre polynomials, 26 Legendre transform, 424, 441 Leibnitz's rule, 13 Levi-Civita permutation symbol eiJk, 7, 622, 645 Liapunov characteristic exponents, 423, 484-486 lift coefficient, 603, 604, 607
Subject Index
Lindstedt-Poincare perturbation, 309,314-316,320,322, 324, 334, 345, 404-409 linear differential operators, 32, 39, 326, 392 linear ordinary differential equations, 31, 34, 50 linear wave theory (LWT), 85 linearized drag coefficient, 568, 582,586,618 linearized drag force, 497, 530, 581,586,847 Liouville theorem, 484 Lorentz's method of equivalent work, 530, 846 low pass filters, 797 Hilbert transform, 792, 795 LVTS, 797 SIWEH, 797 LVTS, 797
MacCamy-Fuchs diffraction theory, 672, 679 Maclauren series, 15, 91, 93, 319, 326-332, 362, 387 marine pipelines, 605 mass moment of inertia, 74, 198, 199, 561-565, 569, 589, 590, 646, 660, 716 Mathieu equation, 424, 426 matrix condition numbers, 540, 555-559 maximum static-equivalent force/moment (Fixed-Free Beam), 515
925
MDOF Lagrangian solid body, 559,573-575,621,623, 636, 684 mean iu,\(x), 723, 730, 733, 735, 737, 761-763, 766, 767, 830 mean frequency of occurrence, 782, 784 mean run length, 799 mean square error, 184, 527, 535, 550,601,744,850 mean square value, 730 mean value, 722, 723, 761 mean wave height, 778 mechanical energy principle, 70 median value, 762, 763, 778 Melnikov integral, 425, 427, 474^76, 479, 482, 483 method of equivalent work, 531, 846 method of Frobenius, 40-^13 method of multiple scales (MMS), 371-374, 380 method of successive approximations, 310, 313, 315 method of undetermined coefficients, 45, 560-562, 571-573,584,588,599, 717,865 Miche formula, 402 Michell breaking wave theory, 400,402, 403 mild slope equation, 251, 255-259, 264 modal analysis, 849-851
926
mode value, 762 mode wave height, 778 modified Bessel function, 22, 214, 219,220,693,757,779, 836 modified wave force equation (WFE), 75, 79, 503, 559, 560, 843, 846, 864 linearized, 565 moment generating function *Cs), 764, 832, 833, 835, 837, 838 moment load, 515, 559-561, 589-598 moments of inertia, 564, 576, 589, 599 momentum principle, 57, 69, 70 mooring line stiffness Kj, 576, 581,590 Morison equation, 75, 497, 498, 503,505,514,520,522, 523,525,527,532,541, 559,565,581,605,608, 609,846, 854 most probable value, 762 most probable wave height, 778 moving body in still fluid, 509 multi-degree of freedom oscillator (MDOF), 504, 559, 572-575, 621, 623, 683
nabla, 10 natural frequency, 37, 41, 561, 571,851 damped, 37,41,46, 561,583
Subject Index
in air, 561 with hydrodynamic effects, 561,571,583 Navier-Stokes equations, 69, 72-74,513 Neumann boundary condition, 34, 165,283,286,413,435, 695-700 Newton's iterative method, 15 Newton's law of viscosity, 53,496 Newton-Raphson method, 106, 110,170 nonautonomous canonical transformation, 454, 455, 487 nondeterministic spectral amplitude simulation (NSA), 810, 812, 816 nondimensional parameters, 434 non-inertial coordinate axis, 102, 103,314,318,574,575, 638-640, 647 nonlinear planar wavemaker theory, 309, 405 normal mode superposition solution, 849 Nyquist folding frequency, 722, 723, 728, 729, 737, 810, 811,819
O'Brien force ratio, 540, 551-553 O'Brien parameter, 527 Ocean Test Structure (OTS), 523 ocean wave data, 535 ocean wave spectra, 730, 750
927
Subject Index
OCHI-HUBBLE six-parameter wave spectrum, 753 offshore structure, 843 O.H. Hinsdale-Wave Research Laboratory, 150, 154, 229, 251,554,754,801,807, 822 one-sided spectral amplitude, 732 one-sided spectral density function, 732, 740, 741, 746,762,781,791,792, 795, 830, 853, 861-864 operational calculus, 31 orbital velocity, 121, 797 orthogonality, 27, 29, 30, 51, 177, 183,232,849 orthonormal eigenfunctions, 52, 177, 180-187,212,215, 244,246,262,271,274, 275,365,410,411,420, 421,675,691,693,699, 703 oscillating U-tube, 523, 532, 534 over damped system, 47
parabolic cylindrical function, 836 parabolic pdf, 837, 838 parabolic series, 837 parallel axis theorem, 565, 589, 646 parametric dependency, 71, 82, 97, 403, 520, 538, 539, 543, 545, 552-554, 573, 611,612,679,756-758, 786, 790, 803
period parameter, 81, 83, 498, 533,534,600, 601 perturbation expansion, 160, 316, 321,322,375,405,639 phase method, 524, 525, 618 phase shift error, 548-555, 558 pile roughness, 602 pipeline stability parameter, 608 pitch mode, 563, 574, 621, 637, 662, 683, 844 planar wavemakers, 151, 152, 154, 155, 181, 188, 189, 243,248 Poisson bracket, 448^156, 488^192 polar mass moment of inertia, 576, 589, 590, 597, 716 polynomials (orthogonal), 26 associated Laquerre, 29 Hermite, 30 Legendre, 26 Tchebyschev, 28 potential energy density, 127-134 power series, 13-15, 18,40-44 pressure coefficient, 512 pressure force/moment, 648, 649, 678, 686 probability density function (pdf) p(x), 731, 760, 772, 779, 781, 832-835, 847 progressive waves, 118, 125, 129, 132 pseudo-direction cosines, 622, 631,671,683,686,709
928
quadrature spectral density function, 731 radiated wave potential, 627, 631, 633,682, 683,707, 712 radiation boundary dampers, 711 radiation damping, 189, 231, 654, 680 radiation damping coefficient, 191-194,199,200,201, 233-242,619,652,654, 681,710,717 radius of convergence, 14 random wave simulations, 770, 808-823 rate of decay of evanescent eigenmodes, 176 ratio test, 14 Rayleigh cumulative distribution (cdf), 772 Rayleigh parameter, 772-779 Rayleigh probability distribution (pdf), 748, 760, 770-780, 786-790 Rayleigh structural damping, 850, 884 reciprocity relationships, 682 recurrence formula, 27, 41, 42 recurrence relations, 28-31 refraction coefficient, 139-144, 610 regular boundary points, 283 regular Sturm-Liouville systems, 50, 51 relative amplitude parameter, 601
Subject Index
relative motion Morison equation (WFE), 503, 559,581, 586, 589-597 relative roughness, 533 resonance, 315, 316, 454,456, 457 restoring force/moment, 564, 626, 647,650,707,710 Reynolds parameter, 72, 78, 81-83, 507, 532-535, 559, 601,616 Reynolds transport theorem, 57, 69,140,430 Robin boundary conditions, 34, 165,283,286 Rodriques' formula, 27-31 roll mode, 589, 596-599 roll mode exciting moment, 596, 598 root-mean-square frequency a>s, 746 root-mean-square lift coefficient, 604 root-mean-square value, 764 root-mean-square wave height, 773,776 rotation of axes, 447 rotational stiffness, 569, 598 rotational stiffness excess buoyancy Kj-, 563 run lengths, 791,799, 800 length of run, 799 length of total run, 799 Runge-Kutta, 484
Subject Index
sand, 609, 610, 613-614 scaling, 58, 92-94, 111, 112, 192, 198,312,320,321,369, 372, 382, 434, 583, 587, 597,834 scaling of equations, 71-79, 122, 258 model and prototype, 79 scattered wave potential, 627, 631,673,682,684,687, 707 Schwarz-Christoffel transformation, 284, 295 SDOF damped harmonic oscillator, 37, 41, 46, 504, 559,563,570,571 segment averaging, 760, 825, 829, 830 separation of variables, 49, 92, 96, 98,160,164,211,260, 273,755 shoaling coefficient, 139-144, 610 Shore Protection Manual (SPM), 144, 802 significant wave height H5, 775 silty-soil, 610, 613, 614 signum function, 13 sine amplitude function, 296 single-degree-of-freedom oscillator (SDOF), 37, 41, 46, 504, 559, 562, 571 SIWEH, 797 skewness ^(x), 762, 838 skin drag, 502, 506
929
sloshing waves, 268, 270 Smale horseshoes, 310, 427, 477 Smale-Birkhoff theorem, 479 small body, 79,496-498, 503, 520 smooth boundary, 282, 284 solvability, 377-382, 388-390 source density a, 688, 697-698, 715 space frame structure, 832, 843, 854-856 spectral bandwidth parameters, 748-749, 782, 786, 790 spectral directional spreading functions, 754-758 uni-directional, 755 cosine, 756 Fourier series, 757 Von Mises circular normal, 757 SWOP, 758 wrapped normal, 759 spectral moment mn, 730, 742, 745-749, 782 spectral narrowness parameter v, 751 spectral peak frequency a>o, 741-744, 752, 753, 757, 828 spectral peakedness parameter y, 751-753 spectral shape parameter, 752 sphere, 507, 509, 691-694 stable manifold, 310, 426^127, 463, 465, 468, 470, 472^182
930
standard deviation ox, 730, 731, 733, 740, 759, 763, 764, 766, 769, 773, 833, 834, 838,843,848,854 static displacements, 37, 78, 79, 561,562 standing waves, 82, 101, 103, 115-118, 123-125, 129-132, 134, 423^124, 439,523,528 stationary ergodic hypothesis, 719, 730, 828, 848 steel-jacketed structure, 495, 831-832,864 stiffness, 37, 41, 75, 79, 561, 563, 569,573,576,581,585, 590, 598, 626, 635, 637, 654,657,660,711 stiffness matrix, 660, 662, 845, 849 Stirling's formula, 9 Stokes drift, 309, 315, 361-364, 368,371 Stokes hypothesis, 65 Stokes parameter, 81, 82, 95, 158-159,206,313,338, 348 Stokes law of viscous friction, 66 Stokes material derivative, 11, 12, 59,89,121,156,189,203, 271,312,317,429,535, 638,642 Stokes material surface, 11, 89, 162, 189,203,253,270, 624, 630, 638, 640, 641
Subject Index
Stokes wave, 20, 310, 316, 321, 339,385,402,414,418, 422 stream function wave theory, 401, 492,518,612 stress tensor, 65, 66, 68, 541 Strouhal parameter, 81, 601 Sturm-Liouville problem, 48-52, 96,97,102,108,165,177, 212,244,261,274,276 surf similarity parameter, 403 surface stresses, 57, 63, 67, 84 surface tension, 81, 301, 303, 424, 425,428, 429, 435, 449, 483 sway mode, 579, 581, 584, 717 sway mode load, 589, 591, 594 sway-heave-roll mode, 593 sway-roll moment load, 592, 594 SWOP wave spreading function, 758
Taylor differential correction, 398 Taylor series, 6, 15-17, 55, 62, 63, 107,114,172,255,475, 501,547,641,744,765 Tchebyshev polynomials, 28 temporal averaging operator, 186,248,362, 527 time sequences, 721, 735, 803, 828 time series, 719, 721-723, 733-735,749,781,803
931
Subject Index
Toeplitz matrix, 816, 818 torus, 466-470, 479-481 total depth-integrated drag force/moment, 504, 505, 516,517,519,520 transformation Jacobians, 733 translated axes, 542, 544, 618 transverse lift forces, 539, 546, 556, 599-604 trigonometric identities, 18, 19, 100-103, 475 two-sided spectral density function, 731-740, 744, 760,842, 860 type E double-actuated wavemakers, 242 type I and II planar wavemaker, 155, 161, 163, 164, 188, 194, 195,200,210,243, 245
U-tube, 515, 522-524, 532-534 under damped system, 46 uni-directional wave spectra, 755 uniform current, 396, 605, 607 unit disk, 281, 284, 288-294, 298-300 unit impulse response function, 37,851-853,857,858 unstable manifold, 310, 426, 427, 460, 463, 465, 468, 470, 472^182
variance a^, 721, 730, 733, 735, 753, 755, 762, 764, 766, 771,796,797,814,821, 850,852 variation of parameters, 35, 33, 48 variational principle, 428 velocity defect, 599 vertical circular cylinder, 203, 231,233,236,501,504, 505,515,534,546,554, 556, 573, 602, 637, 669, 672,680,832, 833 viscosity /x, 58, 65-66, 82-83, 513,599,620 viscous damping coefficient, 583, 588, 595, 598 Von Mises circular normal spreading function, 757 vortex wakes, 602 wake drag, 502, 506, 507 wake effects, 533, 534, 601 water particle velocity, 76, 82, 113,186,314,362,396, 504,517,579,580,607, 680,833,834, 862, 864 water plane stiffness, 585, 590, 660,626, 635, 637 wave action, 140 wave beats, 137, 373 wave breaking, 399^104 wave celerity, 93, 103, 105, 112, 138, 169,261,311,313, 322,324, 345, 401
932
wave density spectrum generic two-parameter, 743, 748, 750,751 four-parameter, 740, 741, 745 wave force equation (WFE), 78, 503,559-565,581,586, 846 wave group velocity, 137-140, 188,228,259,261,382, 733,796 wave groups, 790-792, 797, 799-803, 814, 820-824 wave power, 135-138, 797 Wave Projects I and II, 523, 535, 537, 538, 754, 844,854 wavemaker, 7, 149, 154, 185, 201, 242,251,365,439,699 wavemaker boundary value problem (WMBVP), 151, 155,160,182,252,269, 283, 302
Subject Index
wavemaker gain function, 188, 228, 248 wavemaker power, 185, 187, 191, 221, 225, 227, 228, 243, 247-249 Weber parameter, 81 Weierstrass M test, 15 Wentzel, Kramers, Brillouin and Jeffreys approximation (WKBJ), 260,263,264 Wiener-Khinchine Fourier transform, 731,739, 842, 860 Winkler springs, 845 wrapped normal spreading function, 759 Wronskians, 26, 36, 37, 51 zeroes of polynomials, 27, 29, 30 Zero-crossing d>z or root-mean-square a>s frequency, 746
Advanced Series on Ocean Engineering — Volume 21
WAVES AND WAVE FORCES ON COASTAL AND OCEAN STRUCTURES This book focuses on: (1) the physics of the fundamental dynamics of fluids and of semi-immersed Lagrangian solid bodies that are responding to wave-induced loads; (2) the scaling of dimensional equations and boundary value problems in order to determine a small dimensionless parameter s that may be applied to linearize the equations and the boundary value problems so as to obtain a linear system; (3) the replacement of differential and integral calculus with algebraic equations that require only algebraic substitutions instead of differentiations and integrations; and (4) the importance of comparing numerical and analytical computations with data from laboratories and/or nature.
Acknowledgment for the Cover Photo: Port Camelle, Galicia, Spain, courtesy of Jose Maria Grassa, Director CEPYC, CEDEX, Ministerio de Fomento, Spain
5397hc ISBN 981-238-612-2 'orld Scientific YEARS OF PUBLISHING
s
i
-
2 o o 6
9 789812
386120"
www.worldscientific.com