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!
C). c0 , and (3CpCa) C0 , with type I error a(C0) = a , the chance of incorrectly judging an incapable process as a capable one. The calculations of p-value (rejection probability), critical value, and power of the test are provided below. Suppose the observed value of the statistic Cj = w , we then can calculate those values as follows (here 6 = 3Vn C): p - value = P{(5 7 > w | Ci < C) ^P\tn_i{8)>^
1.3 Estimating and testing Cp based on multiple samples 1.3.1
Estimation
of CP and its
properties
For cases where data are collected in the form multiple samples, Kirmani et al. (1991) consider m samples each of size n and suggest the following estimator of Cp (where af,- is the i-th sample mean, and st is the i-th sample standard deviation): ^
_
(USL-LSL)dp
16
Encyclopedia and Handbook of Process Capability Indices
Here
m(n - 1) - 1 £m(n-i)-i m(n — 1) sp
dp =
e
m(n-l)-l
= E
Xm(n-1)-1
yjm(n -1) - 1 m(n -1)
-1r m
s2 =
and
p
ra(n — 1)'
m n
(
' m(n — 1) — 1) ~ m
- 1 ) i=l
' " i=l
Recall that under the normality assumption the statistic sp / a is distributed as Xm(n-i)-i /[m(n - 1) - 1] 1/2 . Therefore, the estimator C„ is distributed as: yjm(n
- 1) - 1 e r a ( n _i)_i
Cn V^m(n-l)
Note that for m = 1 , the above estimator reduces to the ordinary estimator (UMVUE) of Cp . The estimator Cp is unbiased, and p.d.f., g(y) say, for y > 0 , can be obtained using the p.d.f. of the \2 distribution and denoting h = [m(n — 1) — 1] e^l(n_1)_1 C2V , (which can be expressed as a function of Cp ) becomes: 2flm(n-l)/2 j/-M«-i)+i] exp
2
T[m(n — 1J/2J
1_ 2
[y )
More recently, Pearn and Yang (2003) investigated some statistical properties of Cp and showed that Cp is asymptotically efficient and moreover (mnfl2(Cp — Cp) converges to N(0,Cl /2) in distribution. The variance of Cp can be calculated following (Kirmani et al. (1991)) as:
The CP Index
17
Var(C'r) = E\(C",?} - (E(C'r)f
f1}
UJ — Cp m(n - 1) - 1
2
— c„ m(n-1)-2J e -»(»-D-i Recall t h a t e m ( n _ 1 )_ 2 *s argument m(n — 1) / 2 .
a
cons
_
i _
M
2
I
,
°" 1 e2 (n-l)-2
-1
t a n t involving T functions in the
In addition to being unbiased, it can be shown t h a t Cp is consistent for Cp . For multiple samples with variable sample sizes, we can easily show t h a t the generalized estimator Cp obtained from m samples each of size n^, i = 1, • • •, m , is given by:
i/... x C, n„ — = b< P EK-l)
USL-LSL
, , where
6sv
y ^ (ni — i)si s„ = - ^ P m
1
, and
E^- )
m
r
2
l2^("i-i)
i=l / m
\
1=1
4
H
2
8= 1
»=i
This generalized estimator is unbiased.
1.3.2
Lower
confidence
bound
on
CP
Since the estimator Cp is subject to a sampling error, it is desirable to construct a confidence interval to provide a range which contains the true Cp with a desired high probability. For cases where multiple samples are available due to sampling at various times, Kirmani et al. (1991) used the lower confidence
18
Encyclopedia and Handbook of Process Capability Indices
bound corresponding to the prescribed minimum value of the capability (the precision requirement). A process is considered to be capable, if the precision requirement is greater than the lower confidence bound. The 100(1 — a) percent lower confidence bound can be expressed as:
CI = C M l
X-m(n—1),1—a
[m(n-l)-l]e2, L
1.3.3 Hypothesis
v
testing
'
n
,
' m(n—1)—1
with CP
To test whether the process meets the precision requirement, we shall consider the following testing hypothesis procedure with H0: Cp < C (the process is incapable), versus the alternative H^ Cp > C (the process is capable). Thus, we may consider the test of the form 4>'(x) = 1 if Cp > c0 , and
cr'
0
1 M
c2
= a.
Hence, given the value of precision requirement C , the critical value CQ of the a -level significance can be obtained by solving the equation 2£m(n-l)/2
Jo0_ '
2
m(n-\)/2 „
r
r
/
,y-[m(n-i)+i]exp - \
< ~
m(n-l)/2
dy = a,
19
The CP Index
where k = [m(n - 1) - 1] s^i^iyi Cp , or alternatively using Kirmani et al. (1991), to arrive at: ([mtn-l)-!]^.!).! ch=CA X-m(n—1),1—<*
Under the same conditions, the p-value (the risk of wrongly rejecting the null hypothesis H0: Cp < C ) corresponding to Cp , a specific value obtained via the sample data, denoted by cp is:
p(c; >c*p\cp = c) 2fcrn(n-l)/2
-
1
y-(m(n-l)+\]
0
" /
= P
9^-^/2
r[m(Vi-i;/2]
i"1^-1)-1]^™-!)-!C
2
dy
e x p
ly 2 1
= P
-^•m(n-l)
2 < _ •^m(n-l) — ~* Cp
Pearn and Yang (2003) developed an efficient test for Cp for the cases with multiple samples, and have shown that the proposed test is indeed the UMP test. The power of the UMP test (i.e. probability of correctly rejecting the null hypothesis Cp < C when Cp > C is true), can be computed for the alternative hypothesis, Hx: Cp = C\ > C. The power of the test, denoted as 7r(Cp) is obtained by calculating TT(Cp) = P(Cp>C*\ = p Xm(„-i) ^
Cp=C1)
S
I ° P — °i
20
Encyclopedia and Handbook of Process Capability Indices
1.4 Estimating and testing Cp based on (X, R) control chart samples For applications where the data are obtained in the form of a single sample, this problem has been discussed in an earlier work of Kane (1986). Chou et al. (1990) provide tables for lower confidence limit on Cp when a estimated by the sample standard deviation S. In this case, Pearn et al. (1998) introduced an unbiased estimator of Cp and showed that the unbiased estimator is also UMVUE. These authors also proposed an efficient test for Cp based on a single sample and showed that the test is a UMP (uniformly most powerful) test. Kirmani et al. (1991) considered the estimation of a and the precision index Cp using the data in the form of multiple samples. When a estimated by the sample range a\, Li et al. (1990) provide tables for the lower confidence limit on Cp . Pearn and Yang (2003) propose an unbiased estimator of Cp for multiple samples when a is estimated by pooled sample variance, and have shown that the unbiased estimator is the UMVUE of Cp , which is asymptotically efficient. Pearn and Yang (2003) also developed an efficient test for Cp when multiple samples are used and showed that the proposed test is indeed an UMP test. For applications where a routine-based data collection plans are implemented, a common practice for process control is to estimate the process precision by analyzing the past "in control" data. Consider m preliminary multiple samples (subgroups), each of size n taken from the control chart samples. To estimate a we usually use either the sample standard deviation or the sample range. The control chart can be used as a monitoring device or a logbook to show the effect of changes in the process performance. Observe that a process may be in control but not necessarily operating at an acceptance level. Thus, management intervention will be required either to improve the process capability, or to change the manufacturing requirements ensuring that the products
21
The CP Index
meet at least the minimum acceptable level. We emphasize again that the process must be stable in order to produce a reliable estimate of process capability. If the process is out of control in the early stages of process capability analysis, it will be undesirable and unreliable to estimate process capability. The priority action is to determine and eliminate the assignable causes in order to bring the process into an in-control state.
1.4.1 Estimation
of Cp based on (X, R)
samples
Let m samples each of size n, from a ( X , R) control chart be available, and Rin be the range of a sample of size n (i = 1, •••,m ) and Rm^n be the average range in m samples of each size n. Then the mean and variance of the relative standardized range Rm^ / a are given by Ei^/a) T/
/H
= E(R1
Var(Rm>n a) = '
Var R
( l,n)
(1.2) 4
— = — , ma m
(1.3)
where d^ and d3 are functions of n, widely tabulated in quality control books and literature based on the assumption of normality and independence (originally presented in Pearson's Table A (1932)). Thus, the estimated process capability precision using the range method can be expressed as: f, ° ^ =
USL - LSL „ 6aR .where g
J i
Rmn = — .
If m = 1, the c.d.f. of the range from a standard normal distribution is
22
Encyclopedia and Handbook of Process Capability Indices
[$(z + t) - #(*)]
-oo
where $(•) and !>(•) are the c.d.f. and the p.d.f. of the standard normal distribution JV(0,1). Using the first two moments of the average range, Patnaik (1950), over 50 years ago, has shown that Rm,n I o is distributed approximately as c\v / Vw , where xl is the chi-square distribution with v degree of freedom and c and v are constants which are functions of the first two moments of the sample average range. Specifically EiR^Ja) = ^^n^1)/^),
and
Var(Rmn/a)
= — v
[r(l±i)/r(|,
(1-4)
(1.5)
From the coefficients of the mean and variance of average range, and the values of d^ and d3 , the values of the mean (expected value) and variance in (1.2) and (1.3) are available. Equating (1.4) to (1.2) and (1.5) to (1.3), we can easily obtain the values of c and v, as unique solutions to the system of equations above. In the early days of control chart usage (the pre-computer era), the range method of estimating a was widely employed to simplify the arithmetic calculations associated with control chart operations. With the availability of modern computer software and efficient hand-held calculators for control chart operations, numerical calculations are not an issue any more, and other methods could be used. If the sample size is relatively small n < 8, the range method results in almost as good an estimator of
23
The CP Index
variance a2 as does the usual sample variance s2, which possesses attractive statistical properties. The relative efficiency (R.E.) of the range method to s2 is shown in Table 1.1 for various sample sizes (Montgomery (2001)). For moderate and large values of n, say n > 10, the range method rapidly loses its efficiency, since it ignores all the information presented in the sample between the largest and the smallest values. Nevertheless we emphasize that for the small sample sizes, which are often employed in various control charts (n = 4, 5, or 6), it is quite satisfactory (under approximate normality). Table 1.1. The relative efficiency of the range method to the sample variance S2 n
2
R.E.
1.000
1.4.2 Hypothesis
3 0.992
4
5
6
10
0.975
0.955
0.930
0.850
testing for Cp based on (X, R)
samples
When estimating the process capability precision by means of the range method from the (X, R) control chart samples, the critical value C0^ can be obtained by finding an appropriate value satisfying the following equation: P(Cm
> Cm ' _d_
~ P
\Cp=C)
= a
> C ,0(R) = P
*hn,n
..
Ofl
d
CQ{R)
3(7
xl<^c cC,0(R)
Note that d = (USL — LSL)/2 is unrelated to d^ and dP . In fact, the critical value Co(fl) f° r the range (R) method can be expressed as
24
Encyclopedia and Handbook of Process Capability Indices ~fva\
Co(ij) = —/—^— C ) "V Xi/,a where x?,a i s the lower a -th percentile of the chi-square distribution with v degree of freedom. Under the same conditions, the p-value corresponding to c ,m , a specific value obtained from the sample data, can be calculated as: p -value = P(Cp{R) > cp{R) \Cp = C)
= P
•Km,n ~ _
P(R)
"•
xl <
3a
^-c
ccp(R)
= G
-Jva\
C
cc'p(R)
where G(-) is the c.d.f. of the chi-square distribution with v degree of freedom. When multiple samples taken from the ( X , R) control chart at various times are available, using Patnaik's (approximate) distribution of the average range, the 100 (1 — a) % lower confidence bound CL^ can be constructed (see e.g. Li et al. (1990) and Pearn et al. (2004). The lower confidence bound satisfies P(CP > CL{R)) = 1 - a = P
aR
>
CL[R) C.p(R))
= P
Jva\CL{R) \ Xv —
cC,p(R)
Consequently, we have that:
Rm,n >
cp{R)
)
The CP
^^CL(R)
_
A
2
25
Index
, or the ratio 2m.
Av,a
- _ £ _ HJ~ ,
-
cCp{R)
r-i
Cm
\j Av,a
^
Observe that the ratio values of C^R) / CP(R) depends on v, c, d^, and a . The values of v, c, d? are determined from m and n. We shall call this ratio CL(R) /Cp(i?) a lower confidence factor.
1.5
Estimating and testing Cp based on (X, S) control chart samples
1.5.1 Estimation
of Cp based on (X, S)
samples
If m samples each of size n from a ( X , S) control chart are available, Kirmani et al. (1991) suggests measuring the process precision index Cp , by its natural estimator Cp defined, analogously to the case m = 1, as follows: -
USL-LSL
.
6<75
1
F
£„_! 1/2
where i=l
and the correction factor
F
, _
—
r[n/2]
n-ir[(n-l)/2]
Here Xt and 5j represent the sample mean and the sample standard deviation of the i-th sample, and the correction en^ is usually denoted in the general quality control literature by c4 . Kirmani et al. (1991) showed that under the normality assumption, the statistic S is approximately distributed according to the normal distribution. Hence with a reasonable approximation:
26
Encyclopedia and Handbook of Process Capability Indices
S - VnTTe,n - l
•N(O,I)
(n -1)(1 - el_x) m
This approximation is particularly appropriate in the situations where adequately tight control of the process variability is desirable so that at least moderately large subgroups (n > 10) are available. In this case, the S-chart is shown to be preferred over the R-chart. We note that the expressions for the distribution of Cp(s) presented in Kirmani et al. (1991), Kocherlakota (1992), and Kotz and Lovelace (1998) ought to be adjusted. In fact, these authors address the distribution of as to be:
&s ~ N
CT
2
>
Consequently,
aP(S)
1+ N
P2
1 1
~
£
n-l
C„.
The estimator Cp(s) i s biased. Its p.d.f., g(x) say, can be expressed as follows: q(x) = -===- x
2
exp
K/x-lf 2k2
for x > 0
g(x) is a function of the constant value Cp and a variant the inverse chi-square p.d.f.. Here
The CP Index
?i
melA
and e,B
27
T(n/2) -
1 _
V n - l r[(n-l)/2]
as above.
For the cases when m > 1 multiple samples of size n are available in view of sampling from the ( X , S) control charts at various time points. Kirmani et al. (1991) have also constructed the 100 (1 — a) % lower confidence bound C^s) to be e
CL{s) - CP(S) l + *„
n-l
me„_!
where as above za is the upper a quantile of the standard normal distribution.
1.5.2 Hypotheses
testing for Cp based on (X, S)
samples
If a ( X , S) control chart is available, then the critical value Co(s) can be obtained by determining the appropriate value satisfying the equation P(CP{S) > CQ{S) \Cp = C) = a
1
_ p —i==-x Cp
exp
(Cp/x-lf 2k2
dx
In fact, the critical value C^s) c a n D e found and expressed as follows, with za representing the lower 100a % percentage point of the standard normal distribution, N( 0,1):
28
Encyclopedia and Handbook of Process Capability Indices
c
Co(5) =
msn_i
where
e
— n-l —
r>/2]
n-lT[(n-l)/2]
Under the same conditions, the p-value corresponding to S(5) ' a s P ec ifi° value calculated from the sample data, can be expressed as: p-value = P(Cp(s) = 1-
> cp{s) | Cp = C)
r cp(s) C„ _„z / ,-- x exp F J 0 V27Tfc
{Cp/x-lf 2k2
dx
where k and e„_! are as given above.
1.6
A Bayesian approach to assessment of CP
The majority of the research works for testing capability indices have been focused on using the traditional distributional frequency approaches. An alternative is to use the Bayesian approach, which essentially specifies a prior distribution for the parameter of interest to obtain the posterior distribution of the parameter and then infers about the parameter by only using its posterior distribution given the observations. Cheng and Spiring (1989) proposed a Bayesian procedure for assessing the capability index Cp with a non-informative prior, n(/i, a) = 1/CT , —oo < ^ < oo , 0 < a < oo , and obtained the posterior probability:
29
The CP Index
p = Pv{process is capable | sample} = Pr{C p > w | x} which is equivalent to finding an one-sided credible interval for a. Furthermore, Pearn and Wu (2003) adopt a non-informative reference prior chosen by Cheng and Spiring (1989), given a prespecified precision level w > 0 (preset process capability requirement), and the posterior probability, based on the index Cp , that a process is capable is given as ..a-1
xexp(-y)dy
i
_ r(a,i/0
l-G(l/t,a,l),
T(a)
with
t =
(n* \2
27 771
1
\
Efa- )
W
1
Efa- )
)
7
~
i)
i=i
77!
E E (xij -x? 7= 1 j = l
where
E (n»~~ X)5P + E ^te - x ) ^2
7= 1
7= 1
^=[2/(n)]^ 2 r[(n)/2]/r[(n-i)/2] 2_^Z_^\xii
and
EK-i)
E( n * -V4
x
7= 1 j = l 777
v
ni
l^2_^\Xij
and
Wbm
7= 1
j=i m
C
27
„2 _ S
P
»=1 j = l 777
Efa-1) 7= 1
x
i)
30
Encyclopedia
and Handbook
of Process
Capability
Indices
Weakness of Cp Most industrial companies at the time dawn of the 21-st century are no longer relying solely on CP to quantify process capability in view of the perceived weakness of the index. The major weakness of this index lies in the fact that it measures potential capability as defined by the actual process spread and does not take into account the actual mean of the process (see Figure. 1.3.). Therefore, CP provides no indication of the actual process performance. It does not reflect the impact that shifting the process mean has on process's ability to produce product within the specifications (Chen et al. (1988, 1990) and Kane (1986)). For this reason a more refined index the Cpk was developed. The indices CP and Cpk, when used together, provide a proper indication of the process capability with regard to both of the process spread and of the process location.
LSL
Nominal
USL
Figure 1.3. Various processes with t h e same value of CV
Chapter 2
The Ca Index
2.1 2.2 2.3 2.4
Process accuracy and the Ca index Estimating and testing Ca based on a single sample Estimating and testing Ca based on multiple samples Bayesian-like estimator of Ca
2.1 Process accuracy and the Ca index
While the precision index Cp discussed in Chapter 1 measures the magnitude of the process variation, the index k, to be defined below, describes the process capability in terms of departure of the process mean // from the center (mid) point m and provides a quantified measure of the extent t h a t a process is off-center. T h e index k is one of the original Japanese indices and is defined as: |/x — m | k =
USL-
LSL
|/U — m | =
d
'
2 where fi is the process mean, d = (USL — LSL) / 2 is the half specification width, USL and LSL are, as above, the upper and the lower specification limits, and m is the midpoint between the upper and lower specification limits ( m = (USL + LSL) / 2 ). This index measures the departure of process mean from the mid-point
31
32
Encyclopedia and Handbook of Process Capability Indices
of specifications. For example, k = 0 indicates that the process is centered at the target (fi = m), k = 1 means that the process mean is located at one of the specification limits. When 0 < k < 1, the process mean is located between the target and one of the specification limits. Obviously, k > 1 indicates that \i fall outside the specification limits (i.e. /i > USL or fi < LSL ), the process is severely off-center and it needs an immediate troubleshooting. The complementary index Ca = 1 — k , referred to as the accuracy index, is defined to measure the degree of process centering relative to the manufacturing tolerance (Pearn et al. (1998)). The index is expressed as \ji — m |
It measures the degrees of process centering (i.e. the ability to cluster around the center), and alerts the user if the process mean deviates from the center m (which is often the target value). Table 2.1 displays various Ca values and the corresponding ranges of the "departure" magnitude of \i. Table 2.1. Ca values and ranges of \i. Range of fj,
Ca value Ca = 1.00 0.75 < Ca < 1.00 0.50
0.75
0.25 < Ca < 0.50 0.00 < Ca < 0.25
fj, =
m
0<|/j-m|<
d/4 < | n-m
d/4
|< d/2
d/2<\n-m\<
3d/A
3 d / 4 <\ fi-m\<
d
Ca = 0.00
H = LSL or
Ca < 0.00
\i < LSL or n > USL
M
= USL
33
The Ca Index
2.2 Estimation and testing Ca based on a single sample To estimate the accuracy index Ca , Pearn et al. (1998) proposed a natural estimator Ca . Let x = 2_j _i xi /n be the conventional estimator of the process mean JJ, , obtained from a stable process, then Cn = 1
\x — m
where, as above, d = {USL - LSL)/2
and m = (USL +
LSL)/2.
This estimator can alternatively be written as: \x — m \ ; = 1
Cn=l
\x — m d^fn a J' -In
where under the normality assumption, the statistic -In | x — m | / a has a folded normal distribution as originally defined by Leone et al. (1961) with the p.d.f.
f(y) = 4~a
l
cos
HwI
(y2 + M2) , y>o. z
V 7T
2a
Thus, the p.d.f. of Ca can be expressed as /(*) = 6C7j^cosh[9nC p 2 (l - Ca)(l - x)} 2TT
x exp
or
f(x) = 6 C p J — c o s h
-9nC2p[(l-x)2+(l-Ca)2} 2 (1 - x)6 -6_ (1 - xf 1+ exp2 1-C„ (i - ca)
34
Encyclopedia and Handbook of Process Capability Indices
for —oo < x < 1. 2.2.1 The first two moments
of
Ca
Using the p.d.f. of Ca , the first two moments of Ca can be calculated to be (Pearn et al. (1998)):
B.6.) - C. - ^ ^ e x p ( ^ )
B&) = cl
+
2(1 -
C.W-JI),
+ 4(1 c M
-^?r- irMM^)
]
" " -^ '
and the variance of Ca can be obtained by calculating Var(Ca) = E{C2a) - [E(Ca)f , where 6 = 9n(Cp - Cpk f . The estimator Ca is biased, but as the sample size increases and approaches infinity, while the terms l/(3C p ) , [2 /(mr)f'2 , exp(—6/2), and $[—3Vn (C p — Cpk )] all converge to zero. Hence, the estimator Ca is asymptotically unbiased. For the percentage bias to be less than 5% (i.e. | E(Ca) - Ca \ /Ca < 0.05) , it is required that n > 30 , and to be less than 1% (| E(Ca) -Ca \/Ca < 0.01) we must have n > 650. Since by definition MSE[Ca ) = Var{Ca) + Bias[Ca ) , we have explicitly:
+ 4(1-Ca)2$(-V£). In addition, the behavior of the natural estimator Ca has investigated by Pearn and Lin (2005) under general conditions (the behavior of the corrected estimator bn_iCa is the same as that of Ca). It is shown that under general conditions, Ca is a consistent estimator, which is asymptotically unbiased. It can be
35
The Ca Index
shown that Ca is asymptotically efficient provided the process characteristic is normally distributed.
2.2.2 Confidence
interval
on Ca
Under the normality assumption, Pearn et al. (1998) have shown that the natural estimator Ca = 1— | x — m \ / d of the process accuracy index Ca , is also the maximum likelihood estimator (MLE), consistent, asymptotically efficient estimator and that Vn(C a — Ca) converges to 7V(0, l/(9Cp)) in distribution. Since the distribution of -fn(Ca — Ca) converges to N(0,l/(9C2)), the distribution of 3VnCp(Ca — Ca) converges to N(0,1) (in distribution). An approximate 100(1 — a) percent confidence interval of Ca can thus be obtained to be: A L/ a
Zg/2
A ~—,^a
3-JnCp
, ~r
Z
a/2 =— •
3VnC p
Here as above za/2 is the upper a / 2 quantile for the standard normal distribution JV(0,1), Cp = b^Cj,, T(x) = J°° t^1 e~l dt, and bn_! = (2 /(n - 1)) 1/2 x T[{n - 1 ) / 2 ] / T[(n - 2) /2] . An 100(1 — a)% lower confidence limit on Ca can be constructed using only the lower limit Ca — za /(3Vn Cp).
2.3 Estimating and testing Ca based on multiple samples The past "in control" data consisting of m subgroups, with variable sample sizes n, are collected randomly from a stable process distributed in accordance with the normal distribution, N(fi,a2) . The collected data is then analyzed to compute the manufacturing capability. Let Xt = Y^._ xi j n and St — [(rij — 1)~ z2 '=i(xy ~ X^2]1'2 be the i-th sample mean and the sample standard deviation, respectively, and total number of
36
Encyclopedia and Handbook of Process Capability Indices
observations be N = J ^ n^ . As estimators of \i and a2 , we shall use the overall sample mean and pooled variance, which are unbiased estimators, i.e. _
1
Hi
ii = x = -jjJ2^i
a n d &2
i
ni
= 4 = w^2(ni - !)*? •
We shall thus consider the following estimator Ca to estimate the accuracy index Ca: £. IX — ml a \X — m n n C„ = 1 ; l = 1 - —-T= x where ViV | X — m | / a is distributed as in accordance with the folded normal distribution with parameters ViV \ \x — ra\ j a and 1 as given in 2.2. The corresponding p.d.f. is f(x) =
ry "l — <x =l-P(Y
fy(y)dy 1
T7=Uz{-4y)
+
fz{4y)]dy,
2^Fu
where fY(-) is the p.d.f. of the folded-normal distribution.
<{D(l-x)f)
37
The C„ Index
Hence, the c.d.f. of Ca can be expressed in terms of a mixture of t h e normal distribution:
Fd (x) = l -
fbmi~X)[cj>(t
+ £VAT) + ^ t -
t^N)]dt,
for x < 1 . Differentiating with respect t o x we arrive at t h e p.d.f. of Ca: fda ( i ) = D[fz ( - £ > ( 1 - x)) + j z (D(l - x))}, for x < 1 . Furthermore, the r-th moment of Ca can directly be obtained as: E{CTa)
VF1
= E
r /r
=E
(-lyzj-'^r'/ 2 ).
j=0
Now, following Pearn ei a/. (1998) we arrive at:
£(C0) = C
0
>/2e"A/2 - ±£>VTT ^ — +
2(l-Ca)$(-|«l)..
1 9 / 9 ^ A/ 2 2V2e~ E(C2a ) = C l + ^ - ~ ~ ^ - + 4 ( 1 - Ca
and
where $(•) is as above distribution N(0,1). T h u s , Var(Ca)
Lower
confidence
bounds
t h e c.d.f.
=
)$(-\6\),
of the s t a n d a r d
E(Cl)-E2{Ca).
on Ca
T h e index Ca m a y also be rewritten in t h e form:
normal
38
Encyclopedia and Handbook of Process Capability Indices
c.-i-i^-i- d / < r ' where £ = (/j, — M) / a . In the case when Ca = C , b = d /a can be expressed as b =| £ | /(l — C ) . Hence, given a sample of size N , the confidence level 7 , the estimated value Ca, and the parameter £, the lower confidence bounds C^ can be obtained using numerical integration with iterations, to solve the basic equation
j^(i-c.)^t
+
^ ^
+
^
_^
^
= l-
7
.
(2.1)
2.4 Bayesian-like estimator of CQ Pearn et al. (1998) investigated the process accuracy index Ca to measure the degree of process centering. The investigation, however, only restricted to cases where the process mean may be assumed as an unknown constant. Pearn and Lin (2002) pointed out that in real-world applications, the production may require: (1) multiple supplies with different quality characteristics on each shipment of the raw materials, (2) multiple manufacturing lines with inconsistent precision in machine settings and engineering effort for each manufacturing line, or (3) multiple workmanship shifts with unequal performance level on each shift. Those situations certainly reflect many industrial applications more closely, covering most manufacturing industries, including the manufacturing of the microelectronics, the IC chips, wafers, computer components, PC monitors, as well as most assembly lines. In such situations, the process mean, /j,, often shifts from one constant (with probability p ) to another constant (with probability 1 — p), and the assumption made in most capability research problems that the process mean is staying as a constant
39
The Ca Index
may not satisfied in real-world situations. Consequently, using the natural estimator Ca to measure the potential and performance of such a process would be inappropriate as the resulting capability measure and testing would not be accurate. In the following, a new estimator of Ca is described assuming that the knowledge of P(n > m) = p G [0, 1] can be obtained from historical information of a process that is demonstrably stable (being in statistical control). The new estimator Ca is defined as:
Ca 1
~
d
'
where A = {n\ \i > m} , and IA(-) is the indicator function. To evaluate the estimator Ca , one needs to determine the value of the indicator which requires additionally the knowledge of P(/i > m) , or that of P(n < m). If historical information of the process shows P(/tx > m) = p , one can then determine the value /^(/i) = 1, or — 1 using available random number tables. (For example, assume p = 0.62 is given, then IA(/J) = 1 if the generated 3-digit random number is no greater than 62, and /^(/x) = —1 otherwise.) Pearn and Lin (2005) showed that if the process obeys a normal distribution N( /tx, a2) with P(/x > m) = p G [0, 1] is given, then 3VnC p (C a - Ca) is distributed as JV(0,1). Let Xl,X2,...,Xn be a random sample of measurements from a process possessing the fourth central moment /x4 and LSL < JJ, < USL . Under general conditions, if the knowledge of P(fi > m) — p , where 0 < p < 1, is available, then as n —> oo , ~fn{Ca — Ca) converges in distribution to iV(0, (9Cp)_1) . Furthermore, Pearn and Lin (2005) have also shown that Ca is the UMVUE and asymptotically efficient estimator of Ca. Since the definition of Ca involves X , then the distribution of Ca must relate to a and hence CP appears in the normalized expression 3Vn • Cp (Ca — Ca).
Chapter 3
The Cpk Index
3.1 Process capability and the Cpk index 3.2 Estimating and testing Cpk based on a single sample 3.3 Estimating and testing Cpk based on multiple samples 3.4 The Bayesian approach to Cpk 3.5 The Bayesian-like estimator of Cpk 3.1 Process capability and the Cpk index Due to the simplicity of its design the index, Cp cannot reflect the tendency of process centering (targeting) and thus gives no indication of the actual process performance. In order to account for the deviations of process mean from the target value, several indices, similar in nature to Cp , have been proposed over the years. These indices attempt to account for the magnitude of process variance as well as for the process departures from the target value. One of such indices is Cpk defined as: _
.
USL — a
a — LSL
(3.1)
where as above USL is the upper specification limit, LSL is the lower specification limit, /i is the process mean and a is the process standard deviation. Utilizing the algebraic identity min{a, b} = (a + b)— \ a — b | / 2 , the definition of the index Cpk can alternatively be written as: 40
The Cpk Index
Cpk=
41
_ d— | fi — m | 3^ '
where as above d = (USL - LSL) / 2 is half of the length of the specification interval and m = (USL + LSL) / 2 is the mid-point between the lower and the upper specification limits. The Cp and Cpk indices are directly related by the process capability indices k or Ca, i.e.: ^pfc
=
Cp x (1 — k) = Cp x Ga .
Note that if /x were outside the specification range, Cv^ would be negative, and the process would then be inadequate for controlling the values of X. The index Cpj. was developed since Cp does not adequately deal in those cases when the process mean fi is not centered. However, Cpfc by itself still cannot provide adequate measure of process centering. In fact, a large value of Cpk does not really provides any information about the location of the mean /x inside the tolerance interval. Revealing relationships between the indices Cp and Cpk are also discussed by Gensidy (1985), Barnett (1988), Coleman (1991), Kotz and Johnson (1998), and Ramakrishnan et al. (2001). Cpk and the percentage of non-conforming (%NC) The process yield is at present defined as the percentage of the processed product units which pass the inspections. Production units are inspected in accordance with specification limits placed on various key product characteristics and are sorted into two categories: passed (conforming) and rejected (defectives). Consequently, yield is one of the transparent, basic criteria for interpreting the process capability. Suppose that a proportion of conforming items is of the primary concern. In this case a most natural measure of what is the above proportion is called the yield. This proportion p is defined as:
42
Encyclopedia and Handbook of Process Capability Indices
USL
/
LsLdF(x)
=
F(USL)-F(LSL),
where F(x) is the c.d.f. of the "measured" characteristic X, USL and LSL are as above the upper and the lower specification limits, respectively. If the process characteristic X follows the normal distribution, N(fi,a2), then the fraction of the nonconforming (%NC ) [defective in the old terminology] may be expressed as:
%NC = i-*(USL;-")
+ *("
~aLSL).
See Figure 3.1.
Figure 3.1. Normal distribution and the proportion of product units outside the specification limits.
Yield assurance based on Cpk The index Cvk has been viewed as an yield-based index since it provides bounds on the process yield for a normally distributed process with a fixed value of Cv\. . The bounds are 2$(3Cpk) -l
43
The Cpk Index
rate, a Cpk level of 1.67 is required. At a Cpk level of 2.0, the likelihood of a defective part drops to 2 parts per billion (ppb).
0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 Figure 3.2. The bounds on nonconforming units in PPM versus Cpk
These important bounds may be established by considering for a process with fixed Cpk the number of nonconformities (namely product items fall out the specification interval [ LSL, USL}). However the exact number of nonconformities will vary depending upon the location of the process mean and the magnitude of the process variation. Thus, we rewrite the definition of the index Cpk as follows. From (3.1) Cpk —
d— \ fj, — m\
3a 1-\S 37
1— \([i, — m)/
d
3(a/d) f 1+ 6 , LSL < fi < m 37 1-6 ,m< fj, < USL . 37
Here m = (USL + LSL) / 2 is the midpoint of the specification interval (the tolerance), d = (USL — LSL)/2 is the half-length of the specification interval, 6 = (/i — m) / d , and 7 = a / d .
44
Encyclopedia and Handbook of Process Capability Indices
For Cpk > 0 , we may obtain an exact expected proportion of nonconforming (%NCj as a function of Cpk and Ca or of Cpk and of Cp taken together to be: %NC = ${-3Cpk]
+ ${-3Cpk(2
-Ca)/Ca],
(3.2)
and %NC = ^[-3Cpk]
+ ^[~3(2Cp-Cpk)}.
(3.2)'
(under the normal assumption). As it is evident from Equation (3.2), under the normality assumption, when the process mean is located within the specification limits, i.e. 0 < Ca < 1 or Cpk > 0 , we have the two sided bounds on %NC to be $(-3C pfc ) < %NC < 2§(-3Cpk) since the standard normal c.d.f. $(•) is a non-decreasing function. This is of course equivalent to the two sided bounds on the process yield: 2<&{3Cpk) - 1 < p< $(3Cpfc) . Note that for Ca = 1.0, the process is perfectly centered (/U = m ) . For Ca = 0.0, the process mean is on the specification limits (fi = USL or /i = LSL ). For processes with fixed Cpk , the number of the nonconformities attains its maximum for a centered process (Ca = 1.0), and %NC reduces when the process mean moves away from the center (Ca decreases). Montgomery (2001) recommended some minimum capability requirements for processes runs under certain specified quality conditions. In particular, he suggests Cpk > 1.33 for existing processes, Cpk > 1.50 for new processes (to break in) ; Cpk > 1.50 also for existing processes on safety, strength, or critical parameters, and Cpk > 1.67 for new processes involving safety, strength, or critical parameters. Finley (1992) also found that the required Cpk values on all critical supplier processes should be 1.33 or higher and that Cpk values of 1.67 or higher to be preferred. Numerous companies have recently adopted criteria for
45
The Cpk Index
evaluating their processes that include process capability objectives more stringent than earlier. The widely popularized Motorola's "Six Sigma" program essentially requires the .process capability of at least 2.0 to accommodate the possible 1.5 a process shift, and that no more than 3.4 parts per million be defective. On the other hand, a process is currently called "Inadequate" if Cpk < 1.00 ; it indicates that the process is not adequate with respect to the production tolerances (specifications), and either process variation ( a 2 ) needs to be reduced or process mean (\i) needs to be shifted closer to the preset target value (often set to the mid-point m ). A process is called "Capable" if 1.00 < Cpk < 1.33 ; it indicates that some caution needs to be taken regarding to the process distribution, and certain process control is required. A process is called "Satisfactory" if 1.33 < Cpk < 1.50 ; it indicates that process quality is adequate, material substitution may be allowed, and no stringent quality control is needed. A process is called "Excellent" if 1.50 < Cpk < 2.00 ; it indicates that process quality exceeds "Satisfactory". Finally, a process is called "Super" if Cpk > 2.00. We emphasize that these designation ought to be taken with a grain of salt. They are based on the normality assumption and often are designated to attract new customers. 3.2 Estimating and testing Cpk based on a single sample The natural estimator Cpk is obtained by replacing the process mean \i and the process standard deviation a by their estimators x and s, which may be obtained from a process that is demonstrably stable (under statistical control): Cpk=
d— \ x — m\ ' ^ "M
17
L =
]
I
\x —m
_ M _ ^ I ^
!
(3 _ 3)
46
Encyclopedia and Handbook of Process Capability Indices
where x = 'Y2"=1xi In an<^ s = [X^"-i(x» — x ) 2 / ( n ~~ Vlf^2 a r e the sample mean and the sample standard deviation, respectively. We note that the terminology such as: demonstrably, very likely, etc. does not always conveys total unconditional confidence. 3.2.1 The r-th moment
of C,pk
Under the assumption of normality, Kotz, et al. (1993) derived the r-th moment of Cpk . The moment can be expressed as follows:
E(c;k) = ±E(±r)£(-i)
dr-JE(\x-m\j)
3=0
d\ln — 1 r E(xZi)J2(-iy "3a j=o
[—1 \d*Jnl
E
•fn(x
— m) a
For r = 1, we arrive at the expected value of Cpk :
Also for r = 2, we have the second moment E(Cpk) to be: \2
E(Cpk)
' dy/n — 1 E(Xn2l) 3^ ,
xE(-D and the variance is easily calculated as:
n (x — m)
47
The Cpk Index
Var(Cpk) =
2
9(n - 3)
+
a
-A/2
a
^(l-2$(-VA))] + ^ } - [ ^ p f c ) f
Here the non-centrality factor is A = n(/j, — m)2 / a2, and the correction factor is: K-i = (2/(« - 1)) 1/2 x T[(n - l ) / 2 ] / r [ ( n - 2)/2]. (See also Zhang et al. (1990), Pearn et al. (1992), and Kotz and Johnson (1993, 1998)). Kotz et al. (1993) have shown that Cpk is a biased estimator of Cpk . The resultant bias is positive when [x ^ m . When ^ = m , the bias is positive for n < 10 , but becomes negative for larger values of n. (For d/a values of 3.0 and 4.0 it is negative for all n > 20 , while for d / a = 5.0 it is negative for large n, n > 30 , and for larger rates d / a = 6.0 the negative bias starts at n > 40.) Evidently, this bias tends to zero as n approaches to infinity. Recall that large values of the ratio d/a indicate that the process has smaller relative variation. So far we don't have an intuitive explanation of the observed phenomenon that for large n the bias is negative for fi = m. Also the exact value of n for which in this case the bias is 0 seems to be not available yet. 3.2.2 Distributional
properties
of Cpk
Chou and Owen (1989) utilize the original definition given in Equation (3.1) as the starting point for deriving the distribution of Cpk . The authors express Cpk in the form Cpk = mm(Cpu,Cpi) >
48
Encyclopedia and Handbook of Process Capability Indices
where Cpu = (USL - x) /(3s) and Cpl = (x - LSL)/(Ss) The latter are natural estimators of Cpu and Cpi , respectively. The distributions of 3^/nCpu and that of 3^/nCpi are, under the normality assumption, the non-central t distribution with n — 1 degrees of freedom and the non-centrality parameters -Jn(USL — /x) / a and Vn(/Lt — LSL) /a respectively. Next, the Owen's (1965) formulas for the joint distribution of two dependent non-central t variables is used to derive the distribution of C.pk . The p.d.f. of Cpk , g^ (y) say, turns out to be rather complex
3Vn $>!;_,, «fe) 9cpk(y) =
3Vn^] n - 1 Qn+1 i=l
ify<0,
n +1 ij,^;0,i? 1
-Qn-i(ti,Si\Q,R)\
fory>0,
where ^ = —^ = 3Vny , <5i = —Z-J~nCpu , 82 = 3VnCp; , R — y/n — 1(<52 — &i)/(Pi ~ h) » a n d 9T„_1 {. (•) is the non-central t distribution with n — 1 degrees of freedom and the non-centrality parameter St. The function £?/ (t,6;0,Rj is given by eta;, Q, (t,6;0,R) = C(f)JoR\ $fe - <5J^VW
and the constant C(f) = V2^F/[r(//2)2^/ 2 - 1 ] . As above $(•) is the c.d.f. of standard normal distribution and
The Cpk Index
49
to the folded normal distribution with the parameter n 1,/2 |/u - m\/a (see e.g. Leone, et. al (1961)) Cpk becomes a mixture of the Xn-i a n d the folded normal distributions (Pearn, et. al. (1992)). The p.d.f. is given by (see Pearn, et. al (1999)): n +21
i =o xexp
fcJy) =
18a'
•(a2z2 + 9(1
-yzf)
y < o,
•dz,
AAnYJP^)B(x^fTTi'\l-yzrz^1 — 0n fi =
xexp
a
°U
• J^(aV+9(1-y*) 2 )}**
2/ > 0.
Here D = (n - l) 1 / 2 d/
W) = 5,=
(A/2/ , and > Ai — l\ 3n-12"/2r((n-l)/2)
2*r((2^ + i ) / 2 ;
Using the integration technique similar to the one suggested in Vannman (1997), Pearn and Lin (2002) obtained an exact and explicit form of the c.d.f. of the natural estimator Cpk , under the normality assumption. The c.d.f. of Cpk is expressed in terms of a mixture of the chi-square and the normal distributions:
o
G
\n-l)(bJn 9ny2
-t)
x [4>(t + £Vn) + 4>{t - £Jn)]dt
(3.4)
50
Encyclopedia and Handbook of Process Capability Indices
for y > 0, where b = d / a , £ = (fj, - m) / a , G(-) is the c.d.f. of the chi-square distribution with degree of freedom n — 1 , Xn-i > and (/>(•) as above is the standard normal p.d.f. Expression (3.4) is equivalent to the special case of the c.d.f. for the estimator Cp(u,v) with u = \ , v = 0 obtained originally by Vannman (1997). 3.2.3
Confidence
intervals
for Cpk
The construction of the exact confidence intervals for Cpj. is difficult due to the fact that the distribution of Cv^ now involves the joint distribution of a two non-central ^-distributed random variables. Chou and Owen (1989), Chou, Owen and Borrego (1990), Zhang et al. (1990), Kushler and Hurley (1992) and Nagata and Nagahata (1994) provide a through treatment of the construction of approximate confidence intervals for Cpk , which are relatively simple to calculate. Heavlin (1998) suggests the following approximation as a 100(1 — a)% confidence interval:
Cpk
| n-1 ' y9n(n-3)
Za 2
C
, Clk (, , 6 \ ' 2(n - 3)1* ' n - l j '
^ + ^ / 2 V 9 n ( n - 3 ) ' 2{n - 3)V ' n - l j
Zhang et al. (1990) obtained expressions for the mean and the variance of the sampling distribution of Cpk based on the normal approximation and determined a 100(1 — a)% confidence interval as:
51
The Cp/t Index
Franklin and Wasserman (1992) suggested an additional modification to produce remarkably accurate 100(1 — a)% lower confidence limits for n > 30 as C.pk
' 9n
2(n - 1) '
(3.5)
where za is the upper a quantile of the standard normal distribution. This approach was also proposed by Bissell (1990), who derived above formula using a Taylor series argument. Nagata and Nagahata (1994) have suggested modifying the above formulas by adding l/(30Vn) and obtained very good results for simulation experiments for the two-sided confidence intervals with this modification. Nagata and Nagahata (1992) also developed a two-sided confidence interval for CnJfe pk which is an extension of two-sided confidence intervals for Cpu and CPi developed by Nagata (1991). It is C.pk
z
a/2\
9n
ru 2
+ 2(npk-
1)
> Cpk + z, a/2
If the term l/(9n) is omitted, then an approximate lower confidence limit on Cpk can be reduced to
V^^TjJ c.
pk •
Kushler and Hurley (1992) suggested a simpler formula for the 100(1 — a)% lower confidence limit given by 1 - zn
U/2(^l)
C,pk '
52
Encyclopedia and Handbook of Process Capability Indices
Numerous investigations are available and have been presented in the literature, examining their results. Franklin and Wasserman (1992) carried out simulations to assess the properties of these limits and discovered that they are conservative, with actual coverage rate being from about 96 to 98%. Guirguis and Rodriguez (1992) have provided an explanation of why these limits are conservative. Kushler and Hurley (1992) described several methods for computing lower confidence bounds and evaluated performance of the various methods in terms of miss rate (percentage). Franklin and Wasserman (1991, 1992) proposed a nonparametric but computer-intensive bootstrap estimation to develop three types of bootstrap confidence intervals for Cpfc . The standard bootstrap interval (SB), the percentile bootstrap confidence interval (PB) and the bias-corrected percentile bootstrap confidence interval (BCPB). The results indicate that some of the bootstrap intervals provide 90% coverage when the process is normally distributed and provided 84-87% coverage when the process obeys a chi-square distribution. Practitioners, however, have to bear in mind that the practical interpretation of the index Cpk is questionable when the normality does not hold. Tang et al. (1997) developed an approximate method to derive the confidence limits for CpA. by graphically investigating the relationship between Cp , k and the percentage of nonconforming items. Hoffman (2001) obtained confidence intervals for Cv^ by using percentiles of the distribution of Cp , and compared them to bootstrap technique proposed by Franklin and Wasserman (1991, 1992) and the approximate method proposed by Tang et al. (1997) via simulation. The results indicated that Hoffman's intervals might be the preferred selection in most cases by providing the requisite coverage probability and possessing narrow interval widths.
The Cpk Index
53
Furthermore, based on the c.d.f. of Cpk expressed in Equation (3.4), Pearn and Shu (2003) developed an efficient algorithm with a Matlab computer program to obtain the lower confidence limits conveying critical information regarding the true process capability. In fact, given the sample of size n , the confidence level 7 , the estimated value Cpk and the parameter £, the lower confidence bounds Cpk can be obtained using a numerical integration technique with iterations, to solve the following equation (3.6) with bL = 3Cpk + |£|: brVn
/
G 0
(n - l)(fe£Vn - tf 9nC2pk
x[0(t + £Vn) + 0(t-£>/«)]<** = I - 7 .
(3.6)
We observe that development of a lower confidence bound on the actual process capability is essential for quality assurance. The lower confidence limit not only gives us a clue on the minimal level of the actual performance of the process which is closely related to the nonconforming units, but is also useful in making decisions for capability testing. Kotz and Johnson (2002) presented a thorough review of process capability indices during the past decade.
Approximate
sample size
determination
As noted previously, the distribution of Cpk is quite complex and rather intractable. Chou et al. (1990) derived 95% lower confidence limit and provided tables values for selected sample sizes and the Cp and Cpk values. Unfortunately, their work in obtaining the tables for Cpk values necessitated them to make certain assumptions in order to find the necessary percentile points for a bivariate non-central student's ^distribution. These
54
Encyclopedia and Handbook of Process Capability Indices
assumptions are, in general, not true for most of processes and this resulted in the 95% lower confidence limits table values actually performing at 96-98% coverage. This flaw, as well as the proposed recommended approximate formulas, have been noted and discussed by Kushler and Hurley (1992) and Franklin and Wasserman (1992). Moreover, Kushler and Hurley (1992) have noted that Bissell's approximation (1990) is equal to or superior to the five other approximate methods examined by means of numerical integration. Furthermore, Franklin and Wasserman (1992) showed that for n > 30 , Bissell's method performs at the desired 95% level for several values of Cpk in finding a 95% lower confidence limit via simulation. Bissell's approximation is given in (3.5) as
cT = c.pk
I 1
Za
i9n
,
+
Cpk
2(n-l)'
Accordingly, for n > 30 (Franklin (1999)) utilitized In - 2 ~ In and solving for n obtains 2 "I Cpk
l
(z*Y 9 n =
Cpk
i
+
r^ Lk(B) P
= (*«r
9C&
i +
2
(i-cw/c*)"
(3.7)
which is dependent on the value of C^ — Cpj, , Cpj. and a or and the value equivalent by depends on the ratio Cpl jCpk , of Cpk . For example, if we expect a 95% lower confidence limit for Cpk to be 0.80 of the Cpk (i.e. C$B)/Cpk = 0.80) and Cpk = 1.50, then the sample size recommended for estimation obtained by (3.7) is 38. (Franklin (1999) provides more details).
The Cpk Index
3.2.4 Hypothesis
testing
55
with Cpk
Remark: For completeness we shall briefly survey the basic concepts of classical hypothesis testing procedures. Critical values are usually used for making decisions in capability testing with a designated type I error a , [the risk of misjudging an incapable process (H0: Cpfc < C ) as a capable one (Hj: Cpk > C )]. The p-values are used for making decisions in capability testing, which present the actual risk of misjudging an incapable process (H0: Cpk < C) as a capable one (H^ Cpk > C). Hence, if p-value < a we then reject the null hypothesis and conclude that the process is capable with the actual type I error pvalue (rather than a). Based on the c.d.f. of Cpk , Pearn and Lin (2002) carried out the statistical theory of the hypotheses testing, and developed a practical procedure accompanied with convenient by tabulated critical values for engineers/practitioners, to be used for decision making in applications. Statistical hypothesis testing methodology used for examining whether the process capability meets the customers' demands can be stated as follows: H0: Cpk < C Hi: Cpk > C
(process is not capable), (process is capable).
As above we shall define the test (j>{x) - a decision making rule- as follows:
56
Encyclopedia and Handbook of Process Capability Indices
preassigned value), b = d/a can be expressed as b = 3C+ |£| . Given a value of C (the capability requirement), the Rvalue corresponding to c , (a specific value of Cpk , calculated from the sample data) is given by: p - value = Pr(C'pfc > c | Cpk = C)
•w. .ffc^B^I - /
o
9n(c )'
.(3.8)
J
with & = 3C+ |£| . Hence, given the values of capability requirement C , parameter £ , sample size n , and risk a , the critical value c0 can be obtained by solving the following equation:
((n-l){b^n-tf)
6Vn
G
/
o
9nco
x[>(* + £Vn) + 0(*-£Vn)]
(3.9)
Observe that we would obtain an identical equation if we substitute £ by —£ into equation (3.8) for fixed values of c , n and C . Therefore, given fixed values of n and C , the p-value corresponding to a specific value c for £ = £0 and £ = — £0 are the same. Similarly, we may obtain an identical equation if we substitute £ by —£ into equation (3.9) for fixed values of c 0 , C , n and a . Hence, given fixed values of C, n and a , the critical values c0 for £ = £0 and £ = — £0 are the same. The reader should carefully distinguish between C , C and c 0 . Critical value c0 and parameter £ Usually, the process parameters \i and a are unknown as well as the distribution characteristic parameter £ = (/i — m) / a . In real applications, these quantities are to estimated naturally by substituting fi and a by the sample mean x and the sample standard deviation s , respectively. This introduces additional
The Cpk Index
57
sampling errors from estimating £ when finding the critical values, and would render this approach to be less reliable. To avoid estimating the distribution characteristic parameter £ , Pearn and Lin (2002) carried out further examination of the behavior of the critical values CQ CLS Si function of the parameter £ . The authors performed extensive calculations to obtain the critical values c0 for the parameter values £ = 0(0.05)3.00, n = 10(50)300, C = 1.00, 1.33, 1.50, 1.67, 2.00, and a= 0.05., which cover pretty much range of applications for the process capability Cpk > 0 . The conclusion is that the critical value c0 is: (i) increasing in £, and decreasing in n (ii) attains its maximum at £ = 1.00 in all the cases, and (iii) remains constant for £> 1.00 for all C (with the accuracy up to 10~6 ). Moreover, the authors have discovered that the critical value CQ for sample sizes n > 30 reaches its maximum at £ = 0.50 and remains unchanged for £ > 0.50 (for n > 100 , £ = 0.35 with accuracy up to 10" 4 ). Hence, Pearn and Lin (2002) suggest that in practice one should solve equation (3.4) with £ = 1.00 in order to obtain the required critical values, without having to estimate directly the parameter £. This approach ensures that the decisions made based on these critical values are more reliable than other existing methods of estimating £ using only sample values. To the best of our knowledge, no work has been published which investigates the distributional properties of Cvk when a is estimated by the range coefficient R '
=
*
•
Li et al. (1990) have investigated lower confidence limits for Cph when a is used as an estimator of the standard deviation. These are similar in their structure to their lower confidence limits designed for Cp (See Chapter 1). From the definition oiCpk = mm{USL — fj,,/j, — LSL} /(3a) = mm{Cpu,Cpi}, the statement that a process is capable if
58
Encyclopedia and Handbook of Process Capability Indices
P(Cpk
> Ck) = a
is equivalent to the requirement that P\Cpu
> ck and Cpl > ck )
= P[Z-hiR<-hL
and Z + hjR > hx ) = a
where Z is the standard normal variable, h^ = (3-J~nCpk)/ a\ and hi = 3Snck . Evaluation of the last two quantities involves calculations of double integrals which calls for numerical integration. Li et al. (1990) have provided tables specifying the minimum values of Cpk for which the process is capable (i.e. Cpk > ck 95% of the time). 3.3 Estimating and testing Cpk based on multiple samples The majority of the results obtained so far related to the distributional properties of the estimated capability indices were obtained based on the assumption of having a single sample. However, as we have already mentioned a common practice is to estimate the process capability indices by using the past "incontrol data" from subsamples. To use estimators based on several small subsamples and then interpret the results as if they were based on a single sample is inappropriate and may result in incorrect conclusions. In order to use past in-control data from subsamples to make decisions regarding process capability, the distribution of the estimated capability index based on subsamples ought to be available. We have already mentioned that Kirmani et al. (1991) have studied the distribution of estimators of Cp based on the sample standard deviations of the subsamples. Almost concurrently Li et al. (1990) have studied the distribution of estimators of Cp and Cpk based on the ranges of subsamples.
59
The Cvk Index
Vannman and Hubele (2003) considered indices belonging to the class designated by Cp(u,v), alluded to in a previous section, and have derived the distribution of the estimators of Cp (u, v) , with the estimators of the parameters \i and a being based on subsamples. These authors also considered estimators that occur naturally when using an X -chart together with an S -chart in quality control. Furthermore, they apply the obtained distributions for studying the hypotheses testing procedure to assess the process capability. Numerical investigations were made to explore the effect of the size and number of subsamples on the efficiency of the hypotheses testing under consideration. Under the normality assumption we have m subsamples, with the sample size of the ith subsample being n^ . For each i , i = 1,2,..., m , let x^ , j = 1,2,..., n,, be a random sample (from a normal distribution) with mean /J, and variance a2 , measuring the characteristic under consideration. We shall assume that the process is in statistical control during the time period when the subsamples are collected. Furthermore, let the process be monitored using a X -chart with a S -chart. For each subsample, let Xj and st denote the sample mean and the sample variance, respectively, of the ith sample and let ./V denote the total number of observations, i.e:
x~i = —J2xv'
s =
i —rrZ]^' ~ ^)2
and N =
Sn» •
Furthermore, denote m
^=£(7^-1). i=l
When all the subsamples have the same size n, we have N = ran and N\ = m(n — 1) . As an estimator of \x the overall sample mean will be used; this is an unbiased estimator, i.e.:
60
Encyclopedia and Handbook of Process Capability Indices
—
1 . ,
The estimator of a2 considered by Vannman and Hubele (2003) is defined to be
- 2 =^I>-iK 2 Hence, one of the estimators of Cp(u,v) that will be studied below will be A ,
x
d- u |p - T | 3V<J2 + (/t - Tf '
The estimator a2 is not an unbiased estimator of a2 . To obtain an unbiased estimator based on the pooled variance, we define
*1
j=i
E
wnere me coernciem; g = —— — «__„,
m ._ l
Hi
.
For subsamples of the same size n we have g = n/(n — l) . Another estimator of Cp (u, v) is given by x, ,
,
d- u |p - T | 3yla2 +(p-
Tf
The distribution of Cp (u, v) can easily be derived once the distribution of Cp(u,v) is determined, since we have the simple relation
The Cpk Index
61
-P£
Cp(u,v) = CJu.
When all the subsamples are of the same size n we have n — 1 . In — 1
Vannman and Hubele (2003) observe that one can obtain X^"Li x l from Xi and sf via the relation ^ " ' {xy — xt f = 22 L, x% ~ n^1 without utilizing each individual observation x^. Thus, given the summary statistics x~t and sf , i = 1,2,...,m , an estimator of Cp(u,v) can be calculated as
c'(«,«) = pV
^-^1 A - y I 3Va'2 + (A - Tf
(310)
by using the unpooled estimator of a2 given by
,, 2 _ Z ^ i = :i 2 - o = i ( ^ V" = — "— TV
^
_ l = T T V 7 1 , y^™' -, xl - Nx N-
The estimator C*p(u, v) in (3.10) corresponds to the estimator of Cp(u,v) originally introduced by Vannmari (1995), based on a single sample of size TV and the unpooled variance estimator. Its distribution is presented in Vannman (1997).
3.4 The Bayesian approach to C,pk Cheng and Spiring (1989) proposed a Bayesian procedure for assessing the process capability index Cp . Shiau et al. (1999)
62
Encyclopedia and Handbook of Process Capability Indices
utilized the Bayesian method for testing the index Cpk under the restriction that the process mean /j, equals to the midpoint of the two specification limits m . This would seem to be a somewhat impractical assumption for many factory applications: in this case Cpk reduces to the basic index Cp . Consequently, Peam and Wu (2003) considered a Bayesian procedure for the capability index Cpk relaxing the restriction above. This yields results that are more general and practical for the real-world applications. The authors derive a Bayesian interval estimate for Cpk and propose a Bayesian procedure for testing the process capability. A 100p% credible interval is an Bayesian analogue of the classical I00p% confidence interval, where p provides the confidence level for the interval. A credible interval covers 100p% of the posterior distribution of the parameter under consideration (Berger (1980)). It is also assumed that the measures x = {xl,x2,- •-, xn} are a random sample taken from independent and identically distributed (i.i.d.) N(fi, a2)observations. Therefore, given a pre-specified capability level w > 0 , the posterior probability that a process is capable utilizing index Cpk is given as p = Pr{the process is capable| x}=Pr{Cpk exp a ~ Jo T(a)y T(a)va+1
+ l3
r r
$ 3VnC„i.
pk
V ~ Jo T(a)y
+$
eX
P
x
(n - \)y
y.
* Md*k+l6)xM(n-i)y T(a)i .a+l
> w | x}
3Vn (Cpk +^<5)x
— 3Vn w - 1 dy
$ 3Vn
y,
[n ~ 1)2/
3~/nw
w
CpkX
«(n-l)y
w
dy
The Cpk Index
- f'o
V
T(a)ya+1•exp
V,
Jo
where a = (n — 1) / 2 , <S = &i(y) = 3Vn
CpkX
62(2/, <5) = 3Vn
^
+
x{$[6!(y)] + * [ M y , « ) ] - l } dy:
\x — m\
«(n-l)y
(
63
w
35)XV(n-l), — w
and <&(•) as above is the c.d.f. of the standard normal distribution. Note that the posterior probability p depends on n, w , 6 (6 = | x — m | / s) and (7pfc . It may be rather complicated to compute the posterior probability p without certain advanced computer programming skills. However, by noticing that there is an one-to-one correspondence between p and C for given n and w, and using the fact that Cpk can be calculated from the original process data, we find that the minimum value of C required to ensure that the posterior probability p reaches a certain desirable level, can be useful in assessing the process capability. Thus, the value C (p) satisfies p = Pr{Cpk > w | x} = Pr{Cpk > C*(p) \ x} Vx. A 100p% credible interval for Cpk is therefore [C (p),oo), where p is a number between 0 and 1, say 0.95, for a 95% confidence interval, which means that the posterior probability that the credible interval contains the true value of Cpk is p. The credible interval (also called a credible set) is the Bayesian analogue of the classical confidence interval. Using the Bayesian approach we say that the process is capable in the Bayesian sense if all the points in this credible interval are greater than a pre-specified value of w , say 1.00 or
64
Encyclopedia and Handbook of Process Capability Indices
1.33. When this happens, we have p — Pi{Cpk > w | x} . In other words, to check whether a process is capable (with capability level w and confidence level p) one needs only to check whether Cpk > C (p) . Pearn and Wu (2003) have provided tables of the minimal values of Cpk and C (p) , with the commonly used capability requirements. (Those tables can be used for testing whether a process meets the preset capability requirements).
3.5 The Bayesian-like estimator of Cpk Bissell (1990) proposed a new and simple estimator C'pk to calculate Cpk , the parameter under the assumption that it is known whether the process mean is /J, < m or \x > m namely, either P({i > m) — 0 or 1. The estimator was defined as C'pk = {USL-x)/(3s) or C'pk = {x - LSL)/(3s) according to whether the process mean, /i , is greater or less than m = (USL + LSL) / 2 . An alternative form for C'pk may be written in the following manner: -, _ d — (x — m) sgn(/U — m)
where sgn(/i — m) — 1 if (i — m > 0 , and sgn(/i — m) = — 1 if \x — m < 0 . As Bissell (1990) point out in either case the distribution (under the assumption of normality) of 3nll2C'vk is the noncentral t distribution with n-1 degrees of freedom and the noncentrality parameter ?>nll2Cpk. Kotz et. al (1993) showed that for the index Cpk , the natural estimator gives values for the standard deviation that smaller than those of Bissell's. In real-world applications, the production may require: (1) multiple supplies with different quality characteristics in each single shipment of the raw materials (2) multiple manufacturing lines with inconsistent
The Cpk Index
65
variable precision in machine settings and engineering effort for each manufacturing line (3) multiple workmanship shifts with unequal performance level at each shift. These situations, no doubt, reflect many industrial applications more accurately, covering most manufacturing industries, including microelectronics, the IC chips, wafers, computer components, PC monitors, as well as most of the assembly lines. Another disadvantage of Bissell's estimator is that it is biased, and the location of the process mean with regard to the specification range midpoint must be known with exactitude. To cover these situations, Pearn and Chen (1996) proposed a new estimator Cpk , which relaxes Bissell's assumption concerning the mean. Evaluation of Cpk requires only the knowledge of P(fi > m) = p , where 0 < p < 1. This estimator is given by At, _ d-(x-
m)IA(n)
where A = {/J,\/J, > m}, and the indicator function IA (/z) = 1 for H > m , /J, £ A , IA(^) = —1 for y, < m , /J, ^ A . To utilize this estimator only P(/x > m) — p or P{JJL < m) = 1 — p must be known. If P((i > m) = 0 or 1, the estimator Cpk reduces to Bissell's Cpi~ . This Bayesian-like estimator, Cpk , is unbiased by adding a correction factor bn_i and its variance is smaller than Bissell's. Pearn and Chen (1997) investigated the asymptotic properties of the estimator Cpk • If the knowledge of P(M > m) = p (and P(/i < m) = 1 — p ) is provided, the asymptotic distribution of the estimator C^ corresponds to a contamination of two normal distributions. This estimator treats the process as a mixture of two manufacturing processes. Such situations occur, for example, when the raw materials or components come from two different conditions, or there are two different groups of workmanship involved in the process.
66
Encyclopedia and Handbook of Process Capability Indices
Weakness of Cpk (see Kotz and Johnson (1998) page 63-64)- Boyles (1991) emphasizes that Cpk fails to measure effectively the effect of process centering on process capability. In fact, it makes no clear distinction between on-target and off-target processes. More importantly, Cv^ by itself provides no indication of the direction in which the process is off-target. Both process location and process spread are employed in the structure of Cpk to quantify capability.
Chapter 4
The Cnm Index
4.1 Process capability and the Cpm index 4.2 Estimating and testing Cpm based on one single sample 4.3 Estimating and testing Cpm based on multiple samples 4.4 The Bayesian approach to Cpm 4.1 Process capability and the Cpm index The Cp and Cpk indices are appropriate measures of "progress" for quality improvement paradigms in which reduction of variability is the guiding principle and process yield is the primary measure of success. However, these indices are not taking into account the cost of failing to meet customers' requirements. A world-renown Japanese quality central expert, G. Taguchi, on the other hand, focus on the loss in a product's worth when one of its characteristics departs from the customers' ideal value T . To handle this situation, Hsiang and Taguchi (1985) introduced the index Cpm , independently proposed by Chan et al. (1988). The index is based on the idea of the squared error loss and it is sometimes called the Taguchi index. It concentrates on measuring the ability of the process to cluster around the target, which reflects the degrees of process targeting (centering). The index Cpm involves the variation of production items with respect to the target value and the specification limits that are preset in the factory. It is defined as:
67
68
Encyclopedia and Handbook of Process Capability Indices
Pm
USL - LSL ~ 6V<72 + (/* - Tf
_ USL -LSL ~ 6r
_ d "3r
(4J)
where as above USL — LSL is the allowable tolerance range of the process, d = (USL — LSL) / 2 is the half length of the specification interval, and r is a measure of the average product deviation from the target value T . The term r 2 = a2 + (fi - T)2 = E[(X — T)2] incorporates two variation components: (i) variation with respect to the process mean and (ii) deviation of the process mean from the target. Since E[(X — T)2] is the expected loss, where the loss of the characteristic X missing the target is often assumed to be well approximated by a symmetric squared error loss function, loss(X) = w(X — T)2 for some positive constant w, the capability index Cpm has been also referred to as a loss-based index. From the definition of Cpm , it is easy to note that if the process variance increases (decreases) then the denominator of Cpm will increase (decrease) and Cpm will decrease (increase). Also, if the process mean moves away from (towards to) the target value, then the denominator will increase (decrease) and Cpm will be decreased (increased). Evidently, Cpm adds an additional penalty for being off-target. It is often assumed that the target lies at the midpoint of the tolerance range (a symmetric tolerance). When this is not the case, there are serious disadvantages in the uncritical use of Cpm (the situation when T •*• m is sometimes called asymmetric tolerances). Note that Cpm differs from Cp , the first-generation index, only in the measure of the process variation. Simple algebra shows that: USL-LSL 2
Cv
6V^ + (^ - T)
2
^
+
(/£^T)2
69
The C„m Index
C.pk
1+
[i — m
h^>
For the symmetric case when T = m , Parlar and Wesolowsky (1998) have shown that Cp , Cpk and Cpm are related by the formula
( c„ CpJfc — Cp
c.pm J
-i-c,-|
H —m
or, equivalently,
Cp
c = pm
Jl + 9(Cp-Cpkf
Ramberg (2002) presents the relationship:
-P + Vi + P2cpm,
C,pk
where f3 =
ti-T
which is valid for T = m .
Also, it follows from the basic definitions, Cp >
max(Cpk,Cpm).
Boyles (1991) has provided a detailed analysis of Cpm and its usefulness in measuring process centering (targeting). He notes that both Cpk and Cpm coincide with Cp when n = T and decrease as /x moves away from T . However, Cpk < 0 for [i > USL or \i < LSL , whereas Cpm for a process with \n — T\> 0 is strictly bounded above by the Cp value of the process with a = |/z — T | . Namely, USL - LSL pm
<
6\fi-T\
(4.2)
70
Encyclopedia and Handbook of Process Capability Indices
The index Cpm approaches zero asymptotically as | \i — T \ tends to infinity. On the other hand, while Cpk = (d— | pL — m |)/(3cr) increases without bound for fixed fj, as a tends to zero, Cpm is bounded above by Cpm < d / ( 3 | / x — T\) . The right-hand side of above equation is the limiting value of Cpm as a tends to zero, and is equal to the Cp value of a process with a = \\i — T\. It follows from (4.2) that a necessary condition for Cpm > 1 is \fi-T\
H —m oC„
a
1+ M
i-^-mK/r+7< 3C
m
3C,
-lei
1 +
^
2
where as above £ =| yu — T \ / a Hence the relation Cpfc < C' pnmm is valid whenever 1 .... l t 2 ... . 2 I £1I > T T £ or | £ | < 3Cn 3C/" ' 2 Equivalently, whenever |/Li — m
2
k <
or Cn > 1 - ^ r r , where as above Ca = 1 9C 9Ci
C pm and percentage of non-conforming
(%NC)
-j—
71
The C„m Index
The motivation behind the index Cpm comes not from an examination of the percentage of NC product of a process, but rather from the process' ability to meet the target values. The Cpm index attempts to ignore conformance to specifications and refocuses on the optimal product's quality attained only when critical dimensions are conform the target. However, upper bounds on the number of NC pieces per lot can nevertheless be estimated for specific values of Cpm . Under the (common in practice) assumption that T = m , the definition of Cpm index (4.1) can be rewritten as a function of 7 = a / d and 6 = {JJL — m) / d to yield: Cpm
Zyju2 +(ji-
Hence, 7 2 =
-62
+6
+8
3C.pm
\d^pm)
3C.pm
mf
3C.pm
-6
3C,pm
3V~72 or
3C,pm
3C„pm
— \6
i.e., for 1-1/(3(7.pm 1 < ca < 1 holds for 0<|<5|<1/(3Cpmj Consequently, we have the following relationship between the exact expected proportion of nonconforming and the index Cpm and Ca for l-l/(3Cpm)
%NC
+
a $
a
2-C„ (3C p m )
2
(1-CJ2
(under the normal assumption)
c„
+$ (3Cpm) 2
(1 - Caf
72
Encyclopedia and Handbook of Process Capability Indices
Johnson and Kotz (1995) carried out further investigations to establish a more exact relationship between Cpm and percentage of NC. Assuming for convenience (without loss of generality) that USL + LSL = 0 , where in this case USL = -LSL = d = (USL — LSL) J 2, these authors found that the expected proportion of NC items associated with a given value of Cpm is %NC = $
—a — \i 2 2
UA -/i J
+ 1-$
V
{W^
where A = d(->pm
In fact, Ruczinski (1996) has shown that Yield > 2&(3Cpm) — 1 , or equivalently that the fraction of nonconformiting items %NC < 2$(—SCpm) . For example, if a process has capability with Cpm > 1.25 , then the production yield would be at least 99.9823%, or equivalently, the fraction of nonconformities no more than 0.0177%. For Cpm > 1 , the production yield would be at least 99.73%, or equivalently, the fraction of nonconformities will be no more than 0.27%. For larger values of Cpm , the expected proportion of nonconforming items is the largest when the process is centered, and moreover interpretations of % NC are applicable only when the target coincides with the midpoint of the specification interval. Loss Assurance based on Cpm If the loss at the specification limits (either USL or LSL) is AQ and the distance from the specification limits to the target T is USL — T = d (symmetric case), then the process loss at the specification limit. AQ = Loss(USL) = Loss(T ± d) = wd2 . Recall that the Loss(X) = WX(X- T)2 . Thus, we obtain that AQ /WT= d21 A 2 , and the expected relative loss Le can be rewritten as _ E[Loss(X)) _ E[Loss(X)] LP — A2
& ' •
( 4
'
3 )
73
The C„m Index
which provides a unitless measure of process performance in terms of lost value of the product (for industrial applications). Since the design of Cpm is based on the average process loss relative to the manufacturing tolerance, from the definition of expected relative squared loss in (4.3), the relationship between the expected relative squared loss and Cpm can be derived directly as
r
pm
2d
=
= I -*°-
6A JE[Loss{X)}
3WTLt
Thus, the expected relative squared losses Le based on the Cpm index can be rewritten as: Loss(Cvm) k
pm)
=
i
A) ^5—-
(4.4)
9WTC2pm
V
;
From the expected relative squared loss of Cpm expressed in (4.4), we note that Cpm is a larger-the-better index and hence, small values of Cpm may be contributed by high expected loss resulting in poorer process capability. Furthermore, Loss(Cpm) will be a constant AQ /{9WTClm) and is not affected by the values of Ca . Note that each newly proposed index is motivated by possible shortcomings of the precious ones, which for some reason is being objected by a number of prominent experts. (See e.g. Burke et at. (1991)) 4.2 Estimating and testing Cpm based on a single sample 4.2.1 Estimation
and distributional
properties
of
estimators
The index Cpm involves the unknown parameters n and a , which ought to be estimated from a sample. Chan et al. (1988)
74
Encyclopedia and Handbook of Process Capability Indices
(CCS) and Boyles (1991) propose respectively the following two estimators of Cpm: d Cpm — Cpm(cCS)
STCCS
d
ri ^ pm
r< ^pm(B)
3
(4.5) a
d o~
^EL(^- T ) 2 /n
(4.5) b
3V4 +(x- Tf '
where as above x = Y^._, x, / n ,s2 = /_). , (Xi — x)2 /{n — 1) and 5„ = 2__, . , (^ — z) / n — n"s • Assuming that the process measurements are normally distributed and that T = m , CCS (1988) derive the p.d.f. of the estimator Y = Cpm
fy(y) = n / 2 - y exp 2
+ A \y
«
V
A'(*)* + '' _ 1 T-,
y > o.
where a = C2pm(l + X/n)(n - 1) and A = n(/i - T)2 /a2 c.d.f. of Y — Cpm can be expressed as
. The
My) = i - expi-DE-nfr-y/o 2 " 2 it+"le_t *' y > °The reader familiar with non central x 2 distributions will recognize that Cpm given by (4.5)a can be shown to be a function of the inverse moments of a non-central chi-square distribution. An alternative, equivalent, formula is provided in Pearn et al.
75
The Cvm Index
(1992). The distributional properties of Cpm are quite complicated for asymmetrical specifications ( T ^ m ) and may serve as a challenging and revealing problem for our readers. When T — m , Cpm is a biased estimator of Cpm , but becomes asymptotically unbiased. Detailed descriptions and proofs of the properties of Cpm are provided in CCS (1988). Boyles (1991) is of the opinion that it would perhaps be more appropriate to replace the factor n — 1 by n in the denominator of (4.5)a since the expression appearing f| = si + (x — T)2 = (xi — T)2 / n whose expectation is
EL,(=
E[r%\ = ^+(p-
Tf
(4.6)
is the uniformly minimum variance unbiased estimator (UMVUE) of the term a2 + (/x — T)2 . In fact, the two estimators, Cpm and Cpm, are asymptotical equivalent. We note that x and s„ are the maximum likelihood estimators (MLEs) of ii and a2, respectively. Hence, the estimators Cpm is also the MLE of Cpm. It is therefore reasonable, for reliability purposes, to use the estimator Cpm = Cpm(B) for evaluation of the process performance. Under the normality assumption, Kotz and Johnson (1993) derived formulas for the r-th moment of Cpm , noting that for r > n , E(Cpm) is infinite. In particular, the first two moments, the mean, and the variance of Cpm are (see also Pearn, et al. (1992)): E{Cpm) =
Pexp(-A)f(i)i£ifiL±2} V 2 exp( ^ 3\ r ( t + J)
Cp
and Var(Cpm) = nC2p e x p ( - A ) ^ i i J ±^ j \ n + 2j 3=
[E(Cpm)f
76
Encyclopedia and Handbook of Process Capability Indices
Recall that the first generation index Cp = d /(3cr) and that A = n(/x — T)2 I a2 (the non-centrality or lack of "targetness"). Furthermore, Boyles (1991) proposes an accurate approximation of E(Cpm) and developes a normal approximation for n > 100 :
E{Cpm) —
n(n + A) 2(n + 2A)
(n-l+A) 2 + ( » - ! ) 2(n+2A)
r
r
(n+A) 2 2(n+2A)
c„
Consequently, Cpm is a biased estimator of Cpm and the bias of Cpm is positive in general since TB is negatively biased (cf. (4.6)). The bias of Cpm is given by Bias(Cpm) — E{Cpm) — Cpm — C,
M iW) r(f)
Readers familiar with the non-central \2 distribution will easily confirm that the quantity nr\ / a2 has the non-central chisquare distribution with n degrees of freedom and the noncentrality parameter A = n£2 , £ = (fi — T)/a. Boyles (1991) and Pearn et al. (1992) have shown that the distributed of Cpm is USL - LSI 6a
n Xn,A
= c,
Xn,A
which can alternatively be expressed as C, myjl Cpm ~ Cp
+
A / n n
(4.7)
\Xn,\
where, as above, Xn,\ denotes the non-central chi-square distribution with n degrees of freedom and the non-centrality parameter A = n£ 2 , £ = (\i — T )/a. Boyles (1991) also presents a moment approximation to the distribution of f| / r 2 in the form xl /v > where v =
77
The C n m Index
n(l + £ 2 ) 2 / ( l + 2£ 2 ). Numerical evidence on the adequacy of this approximation can be found in Patnaik (1949) and Tukey (1957). For large v , a related approximation based on the normal distribution: f B / r ~ N(l,(2v)~1) is quite appropriate. Vannman and Kotz (1995) derived the c.d.f. of a generalized process capability index: d — u\ C*p(u,v) =
3 ^
JJ,
— m\
V(H-TY
(to be discussed in a later chapter). The special case with u = 0 and v = 1 of an estimator of Cp(u,v) reduces to Boyles' (1991) estimator Cpm. The form for the c.d.f. of Y — Cpm given in Vannman and Kotz (1995) can be rewritten as: £ 2 /(93/ 2
L
My)
G
faWdt,
9y2
where D = nl'2d/a , G(-) is the c.d.f. of Xn-i > a chi-square random variable with n — 1 degrees of freedom, and fp(-) is the p.d.f. of a XI,A > a non-central chi-square variable with the noncentrality parameter A and 1 degree of freedom. Furthermore, by rewriting Cpm^ — D /(3JK + H) , where D = nll2d/a , K = ns2n /a2 ~ XLi , and H = n(x - T)2 / a 2 . Thus, the p.d.f. of Cpm that has been obtained by Chen et al. (1999) is: nl—n/2 r\n
fc„m(y)=
o»..n i Ty +
ex
P
(XD2) i=o
Uey2]
A 2
D:2
^
i8y"J
3\T\-
+j
, y > o,
78
Encyclopedia and Handbook of Process Capability Indices
which follows a non-central chi-square distribution with 1 degree of freedom and the non-centrality parameter X = 62 (Recall A = n(/j, - T)2 I a2). This formula is equivalent to the p.d.f. of Cpm presented in Vannman and Kotz (1995). We encourage the reader to compare the derivations of Chen et al. (1999) and that of Vannman and Kotz(1995). Adapting the integration technique presented in Vannman (1997), Pearn and Lin (2003) obtained an explicit form of the c.d.f. of the natural estimator Cpm , under the normality assumption. The c.d.f. of Cpm is expressed in terms of a mixture of the chi-square distributions and a standard normal distribution FA
(y) = 1 - I Jo
C VyJ
^
G
—5- - r {9y2
x[(j)(t + (^n) + (j)(t-^Vn)}dt,
(4.8)
for y > 0, where b = d/ a , and as above £ = (// — T) /a, G(-) is the c.d.f. of the chi-square distribution with degree of freedom n — 1, Xn-i J a n d >(•) is the p.d.f. of the standard normal variable iV(0,1). It is observed by Pearn and Lin (2003) that an identical equation is obtained if we replace £ by —£ in equation (4.8) for fixed values of x and n . An outline of the derivation can be found in Vannman and Kotz (1995).
4.2.2 Confidence
intervals
for Cpm
Several methods have been suggested in the literature for constructing approximate (lower) confidence bounds of Cpm . Marcucci and Beazley (1988) propose using the (ordinary) chisquare distribution to approximate the non-central chi-square distribution. This results in the following approximate lower confidence bound, denoted as Cp^B^:
The Cvm
79
Index
CLpmMB) = Cpm]xl{\
7)
, 0 < 7 < 1,
(4.9)
where Xn (1 — 7) i s the (1 — 7) th percentile of the ordinary central chi-square variable with n degrees of freedom. When the process is on-target (i.e. \i = T), this provides an exact confidence bound; otherwise the bound is conservative. Chan et al. (1990) derive a large sample normal approximation to the distribution of Cpm . The lower bound on CPm yields: CLP£XZ) = Cpm - zaam , where o\ = f | f *'(* ~ T) 2 + (a 4 /2) pm
pm
a m ,
™
\3J
l^
-T)2}
+(%
Moreover, Boyles (1991) used a moment approximation to a non-central chi-square distribution, Xn,x > to obtain an approximate lower confidence bound. It can be using the straight forward arguments shown using straightforward arguments that the distribution of
S2n+{x-Tf
f| 1S.
(T +((1-
T
2Z , ,
where
v =
7^I = —Z Z
T)
.
xl
approximately — , V
(l + 2£2)
Since Cpm/Cpm = TB / T , an approximate 100(7)% lower confidence bound for Cpm is given by Cpm^Xv (1 — l)/v • In practice, Boyles (1991) recommends to use the approximate lower Bo) confidence bound C£l of the form: 'pm C ^ S o ) = Cpm^'^
7)
,
(4.10)
80
Encyclopedia and Handbook of Process Capability Indices
where v = -^ ?-{- , and f = (l + 2£ 2 )
. sn
On the other hand, based on the expression (4.7), the 1007% lower confidence bound Cp^H^ is given as:
C^H)
= C^f^
(4-11)
(Zimmer and Hubele (1997)). Other approximating formulas have been developed for this confidence interval, such as for log(Cpm^B)) o r vC'pm(B) > e t c Subbaiah and Taam (1993) conducted a comparative Monte Carlo study of some of these functions. Zimmer and Hubele (1997) provide tables of exact percentiles for the sampling distribution of the estimator Cpm . Zimmer et al. (2001) proposed an useful graphical procedure to obtain exact confidence intervals for Cpm, where the parameter £ = (// — T) / a is assumed to be a known constant. It is noted that when (//— T) / a = 0 , the lower confidence bounds given by expression (4.11) (Zimmer et al. (2001)) are the same as the approximate lower confidence bounds obtained in equation (4.9) by Marcucci and Beazley (1988) and equation (4.10) by Boyles (1991). Kushler and Hurley (1992) compared several methods for computing lower confidence bounds for Cpm in terms of the miss rate (percentage) via simulations. Results indicate that Boyles' method maintains the nominal level. As for the consistently and quite closely Marcucci and Beazley method, it performs well provided \x is not too far from T . The method of Chan et al. (CXZ) results in a fairly stable performance but tends to be slightly "liberal" for small n . Overall, the Boyles approximation for a lower confidence bound on Cpm performs the best. As it was mentioned above, Lin and Pearn (2002) provided an explicit form for the c.d.f. of the maximal likelihood estimator
81
The Cvm Index
(MLE) of Cpm, which can be expressed in terms of a mixture of the chi-square and normal distributions. Noting that by setting Cpm = C , the term b = d / a can be expressed as b = SCyJl + £2 . We observe that the index Cpm can be expressed as a function of the parameter | = (/i — T)/a (a distributional characteristic): pm =
d
_
3W1 + e
=
d/a
_
b
Wi + e " 3>/i + e'
Consequently, given a sample of size n, a confidence level 7 , an estimated value, Cpm, and the parameter £, the lower confidence bounds (denoted CpmPS^) can be obtained by means of a numerical integration technique with iterations, and solving Equation (4.12) given below. In practice, the parameter £ is unknown, but it can easily be estimated from the sample data as £ = (x — T) / sn . It "regular type" should be noted that the l.h.s. of Equation (4.12) is an even function of £. Thus, for both £ = £0 and £ = — £0 we are to obtain the same lower confidence bound Cpm ' : ( b2n
rbSn/(3Cpm)
Jo
FK
2
^d^'
x[0(i + £Vn) + 0 ( i - £ > / n ) ] d i = l - 7 .
(4.12)
Lower confidence bound Cpm s> and the parameter £ Since the process parameters fi and a are unknown, so is the distribution characteristic parameter, £ = (^i —T)/CT which has to be estimated. In real world applications, £ has to be naturally estimated by replacing /J, and a by the sample mean x and the estimated sample standard deviation sn respectively. Such an approach evidently introduces additional sampling errors [from estimating £] in obtaining the lower confidence bounds, and thus would render this approach to be less reliable. Consequently, the conclusions deduced would result in a smaller production yield assurance to the factories, and also provide lesser quality
82
Encyclopedia and Handbook of Process Capability Indices
protection to the customers. To avoid the need for estimating the parameter £, Pearn and Shu (2003) examined the behavior of the lower confidence bounds C$£s' against the parameter £ , using the well known and time-honored approximation formula presented presented in Equation (4.13) which is due to Wilson and Hilferty (1931). This approximation of a x 2 variable by standard normal one is quite accurate for a wide range of percentile points even with sample sizes as small as 10:
X?(l-7) = v
V 9v
(4.13)
9v
here Z is the iV(0,1) and v is degrees of freedom of the chi-squire distribution. Substituting Equation (4.13) into (4.10), the approximate lower confidence bound suggested by Boyles (1991) can be obtained as 3/2 L
C
^(1-7)J
—C
2(1 + 2
S2 \ 2+ 1 -
9n(l + £
2 ( 1 +
^
9n(l + T•2\2 )
Define now the function /(s)
=
Cpm
z ( 1 _ 7 ) ,^a + 1 9n(l + r2 \)2
2(1 + 2£2 9n(l + C22\)2
3/2
whose first derivative is 2(1 + 2£2) 2(1 + 2£2) Z{1 - 7 ) + 1 9n(l + C2 \)2 9n(l + £ 2 ) 2
-i£3C
Z(l -
7)
2 v -1/2 2(1 + 2tj2^ )
{9n(l + e?
-1
1/2
The
CvrJl
Index:
S3
It is easy to verify that /'(0) = 0 , and that the second derivative /"(£) > 0 when either £ > 0 or £ < 0 , (an even function). Thus, the function /(£) takes its absolute minimum at £ = 0 . To demonstrate that this result is valid also for noncentral chi-square distributions, Pearn and Shu (2003) carried out extensive calculations and obtained the lower confidence bounds Cffs) for £ = 0(0.05)3.00, n = 10(5)200, Cpm = 0.7(0.1)3.0, and the confidence level 7 = 0.95. Note that the parameter values under investigation, [£ = 0(0.05)3.00], cover a quite wide range of applications involving process capability index Cpm > 0 . The results obtained indicate that (i) the lower confidence bound Cpm^ i s decreasing in £ , and increasing in n , (ii) the lower confidence bound C^s'1 attains its minimum value at £ = 0 (in all cases). Hence, in practice it is sufficient to solve Equation (4.12) for £ = 0 to achieve the required lower confidence bounds for given Cpm, n , and 7 , without estimating £. Adhering to this approach, the confidence level 7 will be ensured, and the decisions carried out based on such an approach would turned out to be more reliable. Unfortunately the cover age percentage using this method is not very clear and requires further investigation.
4.2.3 Sample
size determination
for Cpm
Sample size determination is an important step since it directly related to the cost of data collection plan. Thus, based on the several methods proposed above for computing the lower confidence bound on Cpm, approximate sample size required for a given estimation precision Rpm could be obtained in an alternative manner. Marcucci and Beazley (1988) suggested approximate formula for required sample size:
the
following
84
Encyclopedia
n
and Handbook
of Process
Capability
= X " f f ( (L ) 2 7 ) - w h e r e 4m B ) = ^LMB)/Cpm
Indices
.
(4.13)a
Based on the approximate lower confidence bound Cp4 So) proposed by Boyles (1991), an approximate sample size required for a given estimation precision Rffl can be obtained by solving for v , where
" -
r,(Bo)2 "pm
±l
•
Here analogously to R^B) we have Rpi°)2 = C^o)/6pm . Now the sample size n required with a given estimation precision RpmH^ can also be obtained using the formula: _ Xn,A(l-7) 71 =
z>{ZH)2 ^pm
^'
•L*"nrn.
where Rp^
=
C^/Cpm
Furthermore, using the lower confidence limit suggested by Boyles(1991), Franklin (1999) obtained the formula for determining sample size to estimate Cpm as:
" = »/,
, ^ 2
1
.
.o/, »
(4-14)
where zQ is the a percentile of JV(0,1). This formula involves the basic quantities and, in the sprit of the Hilferty and Wilson (1931) approximation, does not contain
The CDm Index
85
X2 variables. It is evident from the above that the sample size necessary to determine a desired lower confidence limit for Cpm depends on the ratio of Cpm / Cpm , and the characteristic of the process £ = QL* — T) j a. For example, if one requires a 95% lower confidence limit for Cpm to be 80% of the value of Cpm (i.e. Cpm/Cpm = 0.80), and £ = 0 then the sample size is determined (calculating using the formula (4.14)) to be 35. Note that the sample size is always rounded up to an integer since (at least) the desired ratio needed to be achieved. Additionally, Pearn and Shu (2003) have developed an efficient algorithm to compute lower confidence bounds to determine the sample sizes required for a specified precision of the estimation of Cpm . These authors also provided tables for the engineers/practitioners to be used in factory applications. Equation (4.12) can be rewritten as:
x [4>(t + £Jn) + c/)(t - £Jn)]dt
= 1- 7 .
(4.15)
Hence, given the desired estimation precision Rp^' , the confidence level y and the parameter £, the sample size n can be obtained by means of numerical integration technique with iterations, solving Equation (4.15). It should be noted, particularly, that the l.h.s. of (4.15) is an even function of £. Consequently, for both £ = £0 and £ = — £0 we shall obtain the same sample size n. In fact Franklin (1999) claims (perhaps without proper justification) that one can provide the largest sample size with £ = 0 , which coincides with the above results. We note that to maintain a confidence level of no less than 7 , the necessary sample size n , obtained using the approach described herein, cannot be reduced.
86
Encyclopedia and Handbook of Process Capability Indices
4.2.4 Hypothesis
testing
procedure
(using
Cpm )
To test whether a given process is capable, we may consider the following statistical testing hypotheses: HQ • Cpm < C (process is not capable), H\ : Cpm > C (process is capable). Based on a given a(c0) — a , the chance of incorrectly concluding an incapable process ( Cpm < C ) as a capable one ( Cpm > C )> the corresponding decision rule is to reject H$ ( Cpm < C ) if Cpm > c0 , and not to reject HQ otherwise. For processes with the target value located in the middle of the specification limits (T = m ) , which is a rather common situation, the index Cpm may be rewritten as: Cpm — 6/[3(1 +£ 2 ) 1//2 ] • Given Cpm = C , b — d / a can be expressed as b = 3C(1 + £ 2 ) 1 / 2 . Given a value of C (the capability requirement) r , c P (C pm > I Cpm = C) — a , the p-value corresponding to c (a specific value of Cpm calculated from the sample data) is: value = I Jo
6Vn/(3c*)„f b2n
G
(9(c)
,2
2
tz [)(t + £Vn) + (j){t - £~fn)]dt,
where G(-) is the c.d.f. of Xn-i - a n d >(•), as above, is the p.d.f. of the standard normal distribution. Given values of a and C , the critical value c0 can be obtained by solving the equation Pr(C' pm > c0 | Cpm = C) = a . Hence, given values of the capability requirement C , the parameter £ , the sample size n , and the risk a , the critical value c0 can be explicitly obtained by solving the equation: ^6Vn/(3co)
Jo
bjn__t2 [9c02
[cj)(t + £Vn) + cf>(t - £~Jn)}dt = a .
87
The C„m Index
Critical values are usually used for making decisions in capability testing with a designated type I error a (the risk of misjudging an incapable process (H0: Cpm < C) as a capable one ( # 1 : Cpm > C)). The p-values are used for making decisions in capability testing, which presents the actual risk of misjudging an incapable process ( H0 : Cpm < C ) as a capable one ( H^ : Cpm > C ) . Thus, if p-value < a then we reject the null hypothesis, and conclude that the process is capable with the actual type I error p-value (rather than a). Based on the c.d.f. of Cpm, Lin and Pearn (2002) implemented the statistical theory of the hypotheses testing, and developed a simple, but practical, procedure accompanied with conveniently tabulated critical values for engineers/practitioners to be used for decisions making in their factory applications.
4.3 Estimating and testing Cpm based on multiple samples Kirmani et al. (1991) indicate that a common practice of the process capability estimation in the manufacturing industry is to first implement a routine-basis data collection program for monitoring/controlling the process stability and then to analyze the past "in control" data. Thus, manufacturing information regarding a product quality characteristic is derived from multiple samples in this case, rather than being based on one single sample. Assuming that these samples are taken from the normal distribution N(/j,,a2) , we let xt and st denote the sample mean and the sample variance, respectively, of the ith sample and let N denote the total number of observations, i.e. ~Ylxv ' si I
1= 1
=
T^2(xv I
1= 1
- *i)2 ' and TV =
V j= l
Hi
88
Encyclopedia and Handbook of Process Capability Indices
As an estimator of fj, we shall use the overall sample mean, which is an unbiased estimator, i.e.: _
1
Hi
£=z= - J ^ A pooled variance estimator defined as
*2 = s? =N:i F, l > - i t f is used as the estimator of a2. For the cases with multiple samples several estimators of Cpm are avaiable as shown below (see Zhang (2001), Vannman and Hubele (2003), and Pearn and Shu (2003)): M c* ^ pm
(N
USL -LSL \2 -l)S2p + 7V(f- T)
JV " 6
c^ Mpm -
N
USL- -LSL
rii-iysj + Nfi -T)
JE> 6 V fiM
_
\2
N USL- -LSL
When all subsamples have the same size n , we have N = mn , p, = x = ^ . _ x"j / m and S2. = ^ . _ Sf / m . If the process follows the normal distributionN(fi,a) then
89
The Cvm Index
CM = -J~N v
^ pm
•"
USL - LSI 6a
NSl
+
N(X -
M)
2
a
, N(fi - T)
-1/2
+
The 7VS2 / a2 and JV(X - n) / a2 are distributed as the central chi-square distribution with N — m and one degree of freedom, XN-m a n d Xi > respectively. Therefore, N pm
~
6a
N
= C.
XN-m+1, A
V XN-m+l, A
where x%,\ denotes the non-central chi-square distribution with TV degrees of freedom and the non-centrality parameter A — JV[(/x — T)/af . The r-th moment (about zero) can be obtained to be: E{Cfmy
=
(^NCprE(xN-m+i,xT
(JNCA
(A/2) J
~7T
exp
]
j
j=0
72
r((2j + N-m X
T((2j + N-m
+
l-r)/2)
+ l)/2)
The p.d.f. of the natural estimator of Cpm can easily be attained as follows: 2(l-N")/2(jl{N"+l) /(*)
3(JV*+1)^+2)
xexp
A 2
C"2 18x2
E
Jj=0
AC/2 36x
2
j!r
AT*+l + 2j)
, x > 0.
Hence (7' = 3>/¥C p , and N - m = N*. Using the method similar to one presented in Vannman (1997), we could obtain an exact form of the c.d.f. of Cp^ . As
90
Encyclopedia and Handbook of Process Capability Indices
above, the c.d.f. of C^ can be expressed in terms of a mixture of a central chi-square distribution and the standard normal distribution: •»6VJV/(3a;) G
F6,m(x) = 1 - JQ
b2N
{^-
*2 p ( * + ^ ) +
where as above b = d /a , £ = (/i — T)/ a , G(-) is the c.d.f. of the central chi-square distribution xli-m w^ith TV — m degrees of freedom, and >(•) is the p.d.f. of the standard normal variable N(0,1). Consequently, the critical values, lower confidence bounds, sample size determination, and the manufacturing capability calculations can be performed here using the same technique as for cases with one single sample. The distribution of Cfm (with variable sample size) can easily be derived once the distribution of C^m (with fixed sample size) is determined due to a simple relationship between them (see Vannman and Hubele (2003)). We note that for cases with one single sample, m = 1 , which is a special case of multiple samples, statistical properties of estimator of Cpm , proposed over the last 14 years by Chan et al. (1988), Boyles (1991), Pearn et al. (1992), Kotz and Johnson (1993), Vannman and Kotz (1995), Vannman (1997) and Lin and Pearn (2002), were intensively investigated.
4.4 The Bayesian approach to Cpm For a single sample, Boyles (1991) has shown that f2 = si + (x — T)2 is an unbiased estimator of a2 + (/x — T)2 . Motivated by this result, for the cases where the data are collected as multiple samples, Wu and Pearn (2003) consider m samples, each of size n.^ and recommend the following estimator of Cpm :
The Cvm
91
Index
EE^-?1)2 T12
r* -
= i=l i = l
(4.16)
m
»=1
where xt is the i-th sample mean, and st is the i-th sample standard deviation. By taking the expectation of the numerator of f'2, Wu and Pearn (2003) obtain m
= £
E
Z v L^i
+ £
Xi
J
-2T
ni
\
xE Zv Z^ ^y
EE^2
= £ > (M2 + a2) - 2T x ]>> iM + ^ T i=l m
i=l
2
»=i
= ^ns[a2+(M-r)2]. i=l
Hence, the estimator f'2 satisfying E(T'2 ) = a 2 + (fi — T)2 is a unbiased estimator of a2 + (fj, — T)2 . However, for multiple control samples, we also need to consider variations between multiple control samples. For this purpose Wu and Pearn (2003) define the ratio of the total within the sample variation (SSW) and the total sum of squares variation (SST ) as:
ssw_V^{Xl3
7 = SST
m
Xl
m
EE(^ - f ) 2 8= 1 j = l
92
Encyclopedia and Handbook of Process Capability Indices
(4.17) n x
x
E f a - 1)4 + J2 i( i ~ f i=l
i=l
where s^ = E j = i (n^ — l)s| / E™-i (n^ — 1) is the pooled variance of the samples. The total sample variation about the target value T can be decomposed as:
EEto - T? = EE(^ - ^)2 + EE(^ - *? i=l
j=l j = l m ni
j—\
i—\
j=l
+ EE(^)2 j= l j = l
= £(»*- l)*2 + i - ^ f a - ^ + E ^ P ' i
m
m
- ^ ( n ? - l ) + ^n l ( 5 2 s„ , and o = J 2
I
c
•£
-*
Thus, the sample counterpart of Cpm for multiple samples, defined in (4.15), can be rewritten as: Clm =
d
,m
•
(4-18)
E^-1) oSp
»=1
iEn» !=1
For a single sample, (m = 1, 7 = 1, and s p = s) the estimator of Cpm i Cpm = d/(3syl(n — l ) / n + <52), which can be reduced to the estimator Cpm as defined in Boyles (1991).
93
The Cvm Index
A Bayesian procedure for assessing process capability was originally proposed in Cheng and Spiring (1989) for the basic, primitive index Cp under the assumption that the process mean \x equals to the target value T . However, this restriction (fi = T) turns out not to be a practical assumption for many industrial applications. Shiau et al. (1999) proposed a Bayesian procedure for the general situation with no restriction on the process mean. But, their proposal work focused on cases with a single sample. A common practice of process capability estimation in the manufacturing industry is first to implement a daily-based or weekly-based sample data collection plans for monitoring/ controlling the process stability, and next to analyze the past "in control" data. Along these lines, Wu and Pearn (2003) developed a Bayesian approach for assessing process capability in cases with multiple samples. As we have already pointed out, it is natural to consider the quantity p = Pr{process is capable | x} when applying the Bayesian approach. In the case of using the Cpm index we are required to determine the posterior probability p = Pr{C pm > u> | x} for some fixed positive number u> . Therefore, given a pre-specified precision level UJ > 0 , the posterior probability (based on the index Cpm) that a process is capable is given by the following expression (Wu and Pearn (2003)): t
P
/
i
o T(a)y T(a)va+1
exp
x[$(h(y) + h(y)) -$(h(y) - Uv))]dy,
(4.19)
where as above $(•) is the c.d.f. of the standard normal distribution, and 7 and 6 are as defined in (4.16) and (4.17).
Here the upper limit t =
1
Efa) »=1
apm
1 Efa) !=1
E n,8= 1
+ 7<S2
94
Encyclopedia and Handbook of Process Capability Indices m
27 E n * W(y) =
i=l
1
6,
|E(^-i>/ and
62(2/) =
't_
1/2
,y
Note that the posterior probability p depends on m , n,, 7 , w and C pm only via m , n,, 7 , 5 and (7pm / a; (see expression for t above). The quantity Cpm / UJ is often denoted by C . Expression (4.19) seems to be computationally cumbersome without some advanced computer programming skills. However, by noticing that there is a one-to-one correspondence between p and C for given m and n ; , and observing that 7 , 6 and Cpm (defined in equation (4.18)) can be calculated from the process data, we find that the minimum value of C necessary to ensure that the posterior probability p reaches a certain desirable level, can be useful for assessing the process capability. This minimum value is denoted by C (p). Thus, this value C (p) satisfies Pr{Cpm > u I x} = Pr{C pm >
apm C\p)
x}
For a single sample, the estimator of C.pm 'pm d/(3sy](n-l)/n + 52) , which can be reduced to the estimator Cpm defined in Boyles (1991) and the posterior probability (based on the index Cpm) that a process is capable is equivalent to one proposed by Shiau et al. (1999). * - " ' T17TJ
Chapter 5
The Loss Indices
5.1 Process loss and the Le index 5.2 Estimation of Lpe, Lot, and Le 5.3 Upper confidence bounds on Lpe, Lot, and Le 5.4 Testing process capability based on the process loss 5.1 Process loss and the Le index One of the most commonly understood basic primitive criterion for interpretation of the process capability is the so-called yield index, which is simply defined as the proportion of conforming items. However, the yield measure does not distinguish among the product that fall inside of the specification limits. To remedy this situation, a quadratic loss function is usually employed to distinguish between the products that fall inside of the specification limits by increasing "the penalty" as the departure from the target value increases. Hsiang and Taguchi (1985) were the first to introduce the loss function approach to quality improvement focussing on the reduction of variation around the target value. However, a quadratic loss function by itself does not provide a comparison with the specification limits and also depends on the unit of the characteristic involved. To address these issues, Johnson (1992) proposed the concept of relative expected squared error loss Le for symmetrical cases, by approaching capability in terms of loss functions. Actually, here
95
96
Encyclopedia and Handbook of Process Capability Indices
the opposite concept of worth is used. It was postulated that a characteristic achieves its maximal worth WT , when X = T , following by decreasing values of the worth as X moves away from the target value T (eventually the worth becomes zero, and then negative). The worth function can be described by WT - k(X - Tf , for WT > k(X - Tf , and becomes zero when | X-T | = (WT/kf/2 . Johnson (1992) viewed the pair (T + (WT /k)1'2, T-{WT/k)ll2) as specification limits for Cpm and defined A = (WT / kf/2 . The ratio of the worth (WT) to maximum worth is called the relative worth and can be expressed as:
w{x) = i - « x - r ' 2 = i WT
<*-2r > 2 A
'
where {X — T)2 / A 2 is the relative loss. The expected relative loss, Le = E[(X — T)2]/ A 2 , is now used to quantify capability. Effectively it is equivalent to Cpm, since Le — \TT) /Cpm • Suppose the product has zero worth outside the specifications by setting A = (USL — LSL) / 2 = d , the index Le is then defined as the ratio of the expected quadratic loss and the square of the half specification width as follows:
=L
(x - Tf d2
iF(l)
=
'-^Zll
where as above fi is the process mean, a is the process standard deviation, d = (USL — LSL) / 2 is the half specification width, USL and LSL are the upper and the lower specification limits, and T is the target value. This measure has a direct relationship with Cpm in view of Le = (3Cpm)~2 . Furthermore, defining
The Loss Index
97
Lpe =(a/ df and Lot = [(fi - T) / df , then Le can be expressed as Le = Lpe + Lot . Tsui (1997) interprets Le = Lpe + Lot , as providing an uncontaminated separation between the information concerning the potential relative expected loss ( Lpe ) and the relative off-target squared deviation (Lot). Since Lpe measures the process variation a relative to the specification tolerance ( d ) , it has been referred to as the potential relative expected loss index. On the other hand, Lot measures the relative process departure from the target value and has been referred to as the relative offtarget squared index. We note that the mathematical relationships Le = (3C p m )- 2 , Lpe = (SCpy2, and Lol = (1 - Caf can easily be established, where Cpm , Cp and Ca are the three basic process capability indices introduced by Chan et al. (1988), Kane (1986) and Peran et al. (1998), respectively. The advantage of using Le over Cpm is that the estimator of the former possesses superior statistical properties than that of the latter, as the former index does not involve a reciprocal transformation of the process mean and variance. It also provides an uncontaminated separation between the information concerning the process "precision" and process "accuracy". Process "precision" relates to product variation (uniformity). Process "accuracy" relates to the degree of process targeting (or centering). This separation provides a direction the practitioners may be considered to choose related to the process parameters in order to improve on the process quality. Some commonly used values of Le are, 1.00 (the process is incapable), 0.44 (the process is incapable), 0.11 (the process is referred as capable), 0.06 (the process is called satisfactory), 0.05 (process is classified as good), 0.04 (process is usually viewed as excellent), and 0.03 (the process is designated as super). The corresponding Cpm values are listed in Table 5.1.
98
Encyclopedia and Handbook of Process Capability Indices Table 5.1. Some commonly used Le and equivalent Cpm . Condition
4
Incapable
1.00
n 0.33
Incapable
0.44
0.50
Capable
0.11
1.00
Satisfactory
0.06
1.33
Good
0.05
1.50
Excellent
0.04
1.67
Super
0.03
2.00
The sub-index Lot measures the relative process departure, from the target value which has already been referred to as the relative off-target loss index. In turn, the sub-index Lpe measures process variation relative to the specification tolerance, which is sometimes referred to as the relative inconsistency loss index. Note that some commonly used values of Lpe, 0.11, 0.06, 0.05, 0.04 and 0.03, are equivalent to the Cp values = 1.00, 1.33, 1.50, 1.67, and 2.00, respectively, which covers a great variety of the precision requirements that are used in the majority of real-world applications. 5.2 Estimation of Lpe, Lot, and Le 5.2.1 Estimating
the process
relative
inconsistency
loss, Lpe
To estimate the process relative inconsistency loss, a natural estimator Lpe can be suggested: f
— ^ ~n-l{r(
pe
y ^ (Xj — X) _ Sn_i d2 ~ d2 '
is t h e Here 5B_i = E " = i ( * * _ *? ^n ~~ ^ "convention^" estimator of the process standard deviation a. The estimator Lpe can be rewritten as:
99
The Loss Index j pe
_ Lpe (n — l)Lpe _ Lpe y ^ (Xj — X) ~ n- 1 Lpe ~ n-lf^ a2
(5.1)
For a process following the normal distribution, Pearn et al. (2003) have shown that the natural estimator Lpe is distributed according to the distribution [Lpe /{n — l)]x^_i, where Xn-i i s a s above a chi-squared distribution with n — 1 degrees of freedom. The p.d.f. of Lpe can be derived as (n — l)x /(*) = 2L.pe xexp
(n-l)/2
—(n — l)x n-1 xT 2L,'pe
, for x > 0 .
The r-th moment, the expected value, the variance, and the mean squared error of Lpe are straightforwardly obtained as follows (recalling the mean value and the variance of the Xn distribution),
E(4e)r = Hence
E
(
Var(Zpe) =
T Ltnf,
n-1
(£pe) =
r{[(n-l)/2] + r} ( ^Lpe \ n-1
T[(n-l)/2]
( L,pe E(x„-i) = Lpe, n-1
"\2
V a r ( x ^ i ) = 2(n - 1)
•'pe
n-1
2IJ. pe n-1
(5.2)
and MSE(Lpe) = E(Lpe - Lpef=Vav(Lpe)
+ [E(Lpe) - Lpef |2 = _
?r2 Llj
pe
n-1
Pearn et al. (2003) has shown that the natural estimator Lpe is the uniformly minimum variance unbiased estimator (UMVUE)
100
Encyclopedia and Handbook of Process Capability Indices
of Lpe , which is also consistent, and asymptotically efficient. Moreover ~fn(Lpe — Lpe) converges to N(0,2I?pe) in distribution. Thus, in real-world applications, using Lpe having all the desired statistical properties as an estimate of Lpe , should be quite appropriate. We note that by multiplying the UMVUE Lpe by the constant cn — (n — 1) / n , one obtains the maximum likelihood estimate (MLE) of Lpe. Pearn et al. (2003) has shown that the MLE Lpe is consistent, asymptotically unbiased and efficient, and that the statistic Vn(Z pe — Lpe) converges to N(0,2I?pe) in distribution. Since cn < 1 , then Lpe = cnLpe underestimate Lpe but with smaller variance. In fact, the mean square error of Lpe can be calculated as MSE(Lpe) — [(2n — 1) / n2 ]I?pe and MSE(Zpe) - MSE(I P e) = {(3n - l)/[n 2 (n - l)]}L2pe > 0 for all sample sizes n . Thus, the MLE Lpe has smaller mean squared error than the UMVUE Lpe , and consequently is more reliable, particularly by short production run applications (such as accepting a supplier which provides short production runs in QS9000 certification). To summarize, for short run applications (with n < 35 ) we recommend using the MLE Lpe rather than the UMVUE Lpe . For applications with sample sizes n > 35 , the difference between the two estimators is quite negligible (less than 0.52%). The square root of the relative mean squared error is a direct measurement which provides the expected relative error of estimation from the true value Lpe . For the UMVUE Lpe , [MSE R (Z pe )]^ = [2/(n-l)f2, and is a function of the sample size n only. Therefore, [MSE R (Z pe )]^ values are identical for all the Lpe values. For example, with n = 300 we have [MSER(jLpe)]^ = 0.0818. Thus, for n = 300 , one can expect that the average error of Lpe would be not greater than 8.18% of the true value Lpe. For the MLE Lpe , [MSE R (4e)]^ = [(2n - l)/n2f which is also a
101
The Loss Index
function of the sample size n only. Consequently, [MSE R (L pe )]^ values are also the same for all Lpe values. 5.2.2 Estimation
of a process relative the off-target
the
loss, Lot
To estimate the relative off-target loss, we consider the natural estimator Lot defined as follows: , Lot
(X - Tf ~ d*
where X — jT] _i %i In 1S the conventional estimator of the process mean /j, . The estimator Lot can be also written as a function of Lpe: ^
=
Lpe nlot ^ Lpe n(X - T) 2 JX
Liftp -'pe
fl
For a normally distributed process characteristic, Pearn et al. (2003) have shown that the estimator Lot is distributed according to the distribution [Lpe / n] Xi (<5) , where Xi (<5) i s the non-central chi-squared distribution with one degree of freedom and the noncentrality parameter 6 = n(n — T) 2 /'a 2 = nLot /Lpe . Therefore, the p.d.f. of Lot can be expressed as:
<7(z) = £
nx)/(2Lpe)f+i exp[-(nx)/(2Lpe)} xT(k + 1)
(8/2)k exp(-<5/2) T(k + 1)
for x > 0. (cf. the p.d.f. of a Xi(^) variable) The r-th moment, the expected value, the variance, and the mean squared error of Lot, can therefore be calculated as:
102
Encyclopedia and Handbook of Process Capability Indices
E(Loty=i^\rE[xU6)r 2Lot V T(k + ± + r)
E
n I
fc=0
E(Lot)
ri
Vax(Z ot ) =
T(k + ±) '
r(fc + l)
f) E ^i = (¥)<1+« = ¥ + L-
fi»,i
Var[Xl2(«5)]
n ^LpeLot n
and
(6/2fexp(-6/2)
+
=
n
(2 + 46)
2D.pe
MSE(Lot) = Var(L ot ) + [E(Lot)
(5.4)
Lot? =
^LpeLpt
3-Lpe I
1
n
n Since Lpe is a function of Sn-i only (e./. (5.1)), and Lot is a function of X only ((5.2)), then the two estimators Lpe and Lot are statistically independent under the normality assumption. Because X is the MLE of \i, by the invariance property of MLE, the natural estimator Lot is the MLE of Lot . Noting that E(4rf) = Kt +(LPeln) , and E(Lpe) = Lpe , the corrected estimator Lot = Lot — (Lpe / n) should be unbiased for Lot . Pearn et al. (2003) have also shown that Lot is an UMVUE of Lot, being consistent and asymptotically efficient, and moreover Vn(Lol — Lot) converges to N(0,4LpeLot) in distribution. Thus, analogously to Lpe in real-world applications using the UMVUE Lot, which has all the desired statistical properties, as an estimate of Lot would be suitable. The MLE Lot has a smaller variance than the UMVUE Lot as it follows by comparing (5.3) and (5.4). However, it can be shown that MSE(Z oi ) = ALFeLot /n + {2/[n(n - l)]}(Lpe)2 , and thus the MSE(Lot) - MSE(Lot) = {(3 - n)/[n2(n - l)}}(Lpe)2 , which is greater than 0 for n = 2, equals 0 for n = 3 , and is less
103
The Loss Index
than 0 for n > 4 . It means that the UMVUE Lot has a smaller mean squared error than the MLE Lot , and is more reliable for applications with n > 4 , provided, of course, the normality and other "standard" assumption are satisfied. Observe that for n < 30 , the discrepancy between the two relative errors (percentage of deviations) is significant, and hence we would recommend using the UMVUE Lot instead of MLE Lot . For n > 30 , the difference between the two estimators is negligible (being less than 0.04%), and using either one of them is equally reliable for practical purpose.
5.2.3 Estimation
of process
expected
relative
loss, Le
To estimate the process expected relative loss (the combined measure of process relative inconsistency loss and process relative off-target loss), the natural estimator Le is defined as follows: f
X)2
l f ( I , - Tf J f f t nfr(
where X = Y^i=iXi
2
d
nfrt
2
d
(X - Tf +
d2
I'l
The estimator Le can be also written as a function of Lpe: LpenLe _ Lp " n Lpe n -r^i
(Xx-T)2 a2
If the process characteristic is normally distributed, the estimator Le possesses the distribution [Lpe / n] Xn (<5)) where Xn (<5) is above a non-central chi-squared distribution with n degrees of freedom and the non-centrality parameter 8 = n(n — T)2 / a2 = nLot I Lpe. Therefore, the p.d.f. of Le can be expressed as:
104
Encyclopedia and Handbook of Process Capability Indices
fc=0
[(nx)/(2Lpe)k+("^}exp[(-nx)/(2Lpe] xY(k + {n/2))
(8 J 2)k exp(-<5 / 2)
r(fc + 1)
, for x > 0
(Recall the p.d.f. of a Xn(6) random variable). Analogously to the Section 5.2.2 the r-th moment, the expected value, the variance, and the mean squared error of Le can be evaluated to be: ( LT
E(4)r =
V
pe
=E
fc=0
E(4) =
Var(4) =
(L.pe n (L.pe
nxi(6)r (2Lpe)rF(k
n J
+ % + r)
T(A; + f)
nxim =(L,n
pe
Var[X^)] =
{8/2)kexp(-6/2)
r(fc + i)
(n+6) = Lpe + Lot = Le,
(L.pe n
or
(2n + 46) =
2L„ and MSE(4) = V a r ( 4 ) + [E(4) - Lef = —^-{Kt
n
-^(Lot+Le), + 4) •
(under the normality assumption) Under this assumption, Pearn et al. (2003) have shown that Le is the MLE, and also the UMVUE of Le . As above the estimator Le is consistent, asymptotically efficient, and the statistic -Jn(Le — Le) converges to N(0,2LpeLot + 2LpeLe) in distribution. In view of all these attractive properties, in practice using Le to estimate the process expected relative loss should be quite appropriate.
105
The Loss Index
5.3 Upper confidence bounds on lipei li0ti and Le We shall now introduce the upper confidence bounds (UCBs) of the three process loss indices Lpe , Lot , and Le , respectively, described above. Evidently f
1 A ( I , - Tf _ l f ( I , - X)2
e
~n{^
2
I
~ n ^
2
J
+
(X - Tf 2
d
f
~^
+
f
^'
where as indicated above Lpe = S2 / d2 and Lot = (X — T)2 / d are the MLE of Lpe and Lot , respectively._ Estimating the unknown parameters mean \i by sample mean, X = z2_-,Xi In > 2 2 n X X n whic and the variance a by S n = Y^ i=i( i - f l \ > is l h e maximum likelihood estimator, the relationship Le = Lpe + Lot can easily be established. This expression provides an uncontaminated separation between the calculated information about the process relative inconsistency loss (L pe ) and the process relative off-target loss (Lot). Below, we shall use the estimators ( Lpe (MLE), Lot (MLE) and Le (MLE and also UMVUE)) to construct upper confidence bounds on Lpe Lot, Le, respectively. 5.3.1 An upper confidence
bound on
Lpe
Under the normality assumption, nLpe / Lpe is distributed as xl-i, a chi-squared variable with n — 1 degrees of freedom. Let Upe = Upe(Xi,X2,---,Xn) be a statistic, calculated from the sample data, satisfying ~P{Lpe < Upe) = 1 — a , where the confidence level 1 — a does not depend on Lpe . The Upe is an 100(1 — a)% upper confidence bound on Lpe. Also: P(£ pe < Upe) = F{nLpe/Lpe
>
nLpe/Upe)
= P(Xn-i > nLpe/Upe)
= 1 - a.
Thus, the quantity nLpe/Upe = Xn-i(cx) > where x L i ( « ) is the (lower) a t h percentile of the Xn-i distribution with n — 1
106
Encyclopedia and Handbook of Process Capability Indices
degrees of freedom. A 100(1 — a ) % upper confidence bound on Lpe can explicitly be expressed, in terms of Lpe, as: Upe
= (nLpe)/
5.3.2 An upper confidence
?&-i(ot).
bound on Lot
Under the normality assumption, nLot / Lpe is distributed as Xi(<5) , a non-central chi-squared variable with one degree of freedom and the non-centrality parameter 6 = n(/z — T) 2 / a2. Let U0t — Uot(Xl,X2,...,Xn) be a statistic calculated from the sample data satisfying P(Lot < Uot) = 1 — a , where as above the confidence level 1 — a does not depend on Lot . Then, Uot is an 100(1 — a)% upper confidence bound on Lot and moreover: P ( 4 t
6Lot/Uot)
= P(xUS)>SLot/Uot)
= l-a.
Consequently under the normality assumption, an 100(1 — a)% upper confidence bound on Lot can be expressed in terms of the estimator Lot as Uot =
6Lot/x}(a,6).
where xi(ai$) i s * n e (lower) a t h percentile of the Xi(^) distribution. Now capability testing can be conducted. In practice, however, the parameter 6 is usually unknown and should be estimated from the sample data. 5.3.3 An upper confidence
bound on Le
Under the normality assumption, (n + S)Le / Le is distributed as x\ (<5)) a non-central chi-squared distribution with n degrees of
107
The Loss Index
freedom and the non-centrality parameter 6 = n(/x — T)2 / a2 = nLot J Lpe . Analogously to the previous cases, let Ue = Ue(Xi,X2,...,Xn) be a statistic calculated from the sample data satisfying P(Le < Ue) = 1 — a , where the confidence level 1 — a does not depend on Le. The statistic Ue is then an 100(1 - a)% upper confidence bound on Le. Also note that P ( 4 < Ue) = P((n + 6)LjLt
>{n +
6)Le/Ue)
= P(->&(6)>(n + 6)Le/Ue) = l - a . Consequently, (n + S)Le / Ue = Xn(a^) > where Xn(a><5) is the (lower) ath. percentile of the Xn(^) distribution. An 100(1 — a) % upper confidence limit on Le can now be expressed, in terms of Le, as Ue =(n +
8)Le/xl(a,6).
5.4 Testing process capability based on the process loss It is becoming increasingly common in industry to require a supplier to demonstrate process capability as an integral part of the contractual agreement between the vendor and consumer. Suppose a customer informed his/her supplier that, in order to qualify for business with his/her company, the supplier ought to demonstrate that the process capability Le of producing the required items is less than C (a pre-assigned constant). This problem may be formulated as a hypothesis testing problem: Ho '• Le > C (incapable), Hi : Le < C (capable).
108
Encyclopedia and Handbook of Process Capability Indices
In statistical hypothesis testing a rejection of HQ is usually a severe conclusion. The supplier would of course, be interested to reject H0 , thereby demonstrating that his process is capable. Moreover, he/she desires to assure that if the process capability is below C the process will be judged as capable with a high probability (say, 0.95). Taking a random sample of size n , calculating the value of Le, we want to claim that the process is capable for at least 100(1 — a)% of the time as long as Le < c0 (a critical value). Pearn et al. (2003) have shown that the critical value is c0 = [xl(a,6) • C]/(n + 6) , where C is the preset capability requirement. In other words, c0 = [Xn(a, <5) • C]/(n + 6) is the maximum value of the estimated expected relative loss Le so that the process will be considered capable at least 1 0 0 ( 1 - a ) % of the time. Denoting (as above) £ = (/i — T) / a , we arrive at 6 = n(/x — T) 2 / a2 = n£ 2 . The formula for calculating the critical value c0 can now be written as c0 = [xl(ain£2) • C]/[n(l + £ 2 )], which is quite easy to interpret and is straightforward to apply. Unfortunately, since the process measurements /x and a are to be estimated from the sampled data (to obtain the characteristic parameter £), a great degree of uncertainty could be introduced to the capability assessments due to the inevitable sampling errors. Johnson (1992) suggests to estimate the parameters \i and a by X and Sn , respectively, to obtain upper confidence limit [(n + n£ 2 )/Xn( a ' n £ 2 )]-4 (which is easy to see to be equivalent to our expression U = Lot + [nLe / Xn(a^)]) f° r Le (see 5.3.2 and 5.3.3). If the estimated value Le is less than the critical value c 0 , we conclude that the process meets the preset capability requirement. Such an approach, of course, introduces additional sampling errors due to estimating £ , rendering it to be less reliable. Consequently, any decisions carried out in these situations would provide a lesser quality assurance to customers. To assess the need for additional estimation of the characteristic parameter £ = (JJ, — T) / a , we should examine the
109
The Loss Index
behavior of the critical value c0 versus the parameter £ . The results show that the critical value c0 is an increasing function of £ and reaches its minimum at £ = 0 (i.e. for /i = T ) in all the cases. Hence, for practical purposes we may calculate the critical value CQ by setting | = £ = 0 for given values of Le, n, and 7 , without the need to estimate the parameter £ further. Using such an approach, the 7 confidence level is ensured and the decisions become more reliable. Furthermore, for testing the hypothesis about Le , H0 : Le > C (incapable) versus H^ : Le < C (capable) where C is a pre-assigned value, one can define a classical test as
2%
L'J
X
where as above /. (x,Le) is the likelihood function and L' is another parameter. Since for Le > Le > 0 , the ratio X(x) is an increasing function of x , {/. (x,Le) \ Le > 0} has the monotone likelihood ratio (MLR) property in Le . Therefore, the test 0* ought to be the UMP test.
Chapter 6
The Cpmk Index
6.1 6.2 6.3 6.4
Process capability and the Cpmk index Estimating and testing Cpmk based on a single sample Estimating and testing Cpmk based on multiple samples Bayesian-like estimation of Cpmk
6.1 Process capability and the Cpmk index In t h e previous Chapters we have concentrated on t h e three basic indices Cp , Cpk and Cpm . T h e indices are interrelated but each one h a s it specific properties. Under t h e normal assumption the distributional properties of estimators of this indices are established without great difficulties and we have by now rather a complete picture of their properties. P e a r n et al. (1992) proposed the process capability index Cpmk , which combines t h e merits of three earlier indices. T h e index Cpmk alerts t h e user when t h e process variance increases and t h e process mean deviates from its target value (or both). T h e index Cpmk, sometimes referred to as the "third-generation capability index", has been defined as follows: USLCpmk = m i n
n
2
n-LSL 2
3V<x + (/x - Tf ' 3Va + (M - T) 2
110
The Cpmk Index
f-\»2
m
\
3V<x + (/x - T )
111
,
(6.1)
2
where as above C/5L and LSL are the upper and the lower specification limits, respectively, \i is the process mean and u is the process standard deviation, T is the target value, d = (USL — LSL) / 2 is the half length of the specification interval, m = (USL + LSL) / 2 is the mid-point between the lower and the upper specification limits. Comparing the pair of indices (Cpmk,Cpm) , similarly to (Cpk, Cp) , we have the relation Cpmk = Cpm x Ca and hence Cpmk — (Cpm x Cpk) JCp (recall Ca ). Furthermore, Vannman (1995) proposed a superstructure Cp(u,v) of capability indices for processes with normal distributions (to be discussed below), which include Cp , Cpk , Cpm and Cpmk as particular cases. By setting u, v = 0 and 1, we obtain the four indices, Cp(0,0) = Cp , Cp(l,0) = Cpk , Cp(0,l) = Cpm , and Cp(l,l) = Cpmk . Those indices are effective tools for process capability analysis and quality assurance. Two basic process characteristics the process location in relation to its target value, and the process spread (overall process variation) are used to establish the formula of these capability indices. The main tenehe of these constructions is the following observations: The closer the process output is to the target value and the smaller the process spread, the more capable is the process. Namely, the larger the value of a process capability index, the more capable is the process. The index Cpmk has been constructed (Pearn et al. (1992)) by appropriately combining the yield-based index Cpk and the lossbased index Cpm , accounting for the process yield (meeting the manufacturing specifications) as well as the process loss (variation from the target) (see Figure 6.1 and Figure 6.2). When the process mean n departs from the target value T, the reduction in the value of Cpmk is more substantial than the decrease in Cp , Cpk ,
112
Encyclopedia and Handbook of Process Capability Indices
and Cpm . Hence, the index Cpmk responds to the departure of the process mean [i from the target value T than the other three basic indices Cp , Cpk , and Cpm , more rapidly while remaining sensitive to the changes of process variation. We note that a process meeting the capability requirement " Cpk > C " for a preassigned C may not meet the capability requirement " Cpm > C ". On the other hand, a process that meets the capability requirement " Cpm > C " may not be meeting the capability requirement " Cpk > C " either. This discrepancy between the two indices may be due to the fact that the Cpk index primarily measures the process yield, while the index Cpm focuses mainly on the process loss. However, if the process meets the capability requirement " Cpmk > C ", then the process ought to meet both capability requirements " Cpk > C " and " Cpm > C " as Cpmk < Cpk and Cpmk < Cpm . Thus, the index Cpmk evidently provides more capability assurance with respect to process yield and process loss to the customers than do the other two indices Cpk and Cpm. According to current modern quality improvement theory, this is a desired property. Actually the reduction of process loss (variation from the target) is just as important as increasing the process yield (namely adhering to the specifications). Note that, the definition (6.1) can alternatively be expressed as Cpmk = Cpk / ( l + ef2 , Z = (M - T)/CJ and also Cpmk = (1— | \x — m | / d)Cpm . Admittedly the index Cpmk is worse than Cpk being associated with a certain percentage of nonconforming items, but we must keep in mind that one should not choose this index when the proportion p is of the main interest. Cpmk (and usually Cpm) is by far more sensitive than other capability indices to the movements in the process average relative to the target T . When JJL equals T , Cpmk is equal to Cpk . When the mean of a process moves away from T , Cpmk decreases faster than Cpk does (although both indices are 0 when /i is equal to one of the specification limits). Conversely, when [i is brought closer to T , Cpmk increases much rapidly than Cpk does.
The Cpmk Index
113
-
2000 C
-
1000
\..y 0.6
0.7
\ S
i
0.8 Ct
Figure 6.1 The actual NC curves for Cpk = 1.0 (top), Cpm = 1.0 (dash), and Cpmk = 1-0 (bottom) with various allowed values of C„.
Figure 6.2 The actual NC curves for Cvk = 1.5 (top), Cvm = 1.5 (dash), and Cpmk = 1 . 0 (bottom) with various allowed values of C„.
Additionally, Cpmk reveals the most of information about the location of the process average. On the assumption that T = m (which is quite common in many practical situations), given the Cpk index of 1.0, all that can be said about /x is that it is somewhere between the bounds USL and LSL , i.e., T-d
114
Encyclopedia and Handbook of Process Capability Indices
the location of /j,. On the other hand, Cpmk provides the most information about the location of /x and the least about the proportion (probability) p . In fact that the corporation's quality philosophy will dictate us which index is more appropriate for a particular situation. The four basic indices Cp , Cpk , Cpm , and Cpmk i when ranked in terms of the sensitivity to differences between the process mean /J, and the target value T are: (1) Cpmk (the most sensitive) (2) Cpm , (3) Cpk and finally (4) Cp which is totally insensitive (Pearn and Kotz (1994)). Yield comparisons between the PCIs As we have shown in Chapter 3, for a normally distributed process, the index Cpk provides a lower bound on the process yield: p > 2$(3Cpk) - 1 , or %NC < 2$(-3Cpk) for 0 < Ca < 1 , which is equivalent to LSL
115
The Cpmk Index
Derivations of this bound are presented in the Appendix of Pearn and Lin (2005). These results conform that the three indices, Cpk , Cpm , and Cpmk provide the same lower bounds on process yield, namely, p > 2$(3Cpk) - 1 = 2$(3C pm ) - 1 . For example, given Cpk = 1.00 the only information about the process yield is available via the upper bound %NC < 2699.796 ppm. However for Cpm = 1.00 the information of the process yield is available through the upper bound %NC < 2699.796 ppm and the process centering measure 0.667 < Ca < 1.00. Finally if Cpmk = 1.00 we are having the upper bound %NC < 2699.796 ppm and the process centering measure 0.750 < Ca < 1.00. These calculations illustrate the advantage of using the index Cpmk compared to the indices Cpk and Cpm when measuring the process yield, since Cpmk provides a better customer protection in terms of the quality yield of the produced items. Table 4.1 displays the bounds on %NC {P(NCJ) and Ca for Cpk = Cpm = CPmk = C(say), respectively. Table 4.1 Bounds on P(NC) and Ca for Cpk = C pm = Cpmk = C. nw
Cpk
c
p ^pmk
pm
Bound on Ca
Bound on P(NC)
1.00 2699.796
0
2699.796
1.33 66.334
0
66.334
0.750 < Ca< 1
66.334
0.812 < Ca
1.50 6.795
0
6.795
0.778 < C 0 < 1
6.795
0.833 < Ca< 1
1.67 0.554
0
0.554
0.800 < C„< 1
0.554
0.850 < Ca< 1
2.00 0.002
0
0.002
0.833 < Ca< 1 0.002
Bound on P(NC)
Bound on Bound on Bound on P{NC) Ca ca 0.667 < Ca< 1 2699.796 0.750 < Ca<\
0.875 < Ca< 1
Observe that the rate of increase in value of the lower bound on Ca decreases (as C increases from 1 to 2) when comparing Cpm a n d Cpmk •
116
Cpmk
Encyclopedia and Handbook of Process Capability Indices an
d the percentage of non-conforming (%NC)
Using the methodology for deriving formulas for an exact expected proportions of nonconforming items, and noting that Cpmk = Cpm x Ca , the following exact expected proportion of nonconforming expressed in terms of Cpmk and Ca are obtained:
2-<7„
%NC = $
3(7.pmk
,
+ $
-(1-Ca)2
ca
i
Observe that for Ca = 1 the %iVC becomes Cpmk and the expected
relative
squared
\ c« 1- ( 1 - -caf 2
on
2$(—3Cpmk). loss
For the index Cpmk , its relationship with the expected relative squared loss Le can directly be expressed as follows: n
—C
vC
1 AoC; 3 V WT4 '
=—
where AQ — Loss(USL) = k(USL - T)2 (if product adopts a specification limit, say USL), Le = [a2 + (fi-T)2]/d2 and WT is the worth of a product on target. Thus the expected relative squared losses based on Cpmk can be rewritten as: i AC2a Loss{Cpmk) = -ur 2 n2
(6.2)
$wTc pmk
From the expression (6.2) the expected relative squared loss °f Cpmk > Loss{Cpmk) , increases as Ca increases and reaches its
117
The Cpmk Index
maximum value for Ca = 1.00. For example, given the level of CVmk — 1-00, the Loss(Cpmk) attains a maximum value Ao/(9WT) at Ca=1.0. We have already mentioned on several occasion that according to current modern quality improvement philosophy a reduction in the process loss is as important as increasing the process yield. Although Cpk offers the most information about the yield, it provides the least insight about the location of \i. On the other hand, the index Cpmk provides the most information about the location of \x and the least about the yield. While the Cpk remains the more popular and widely used index, the index Cpmk is becoming to be an advanced and useful index for processes with two-sided specification limits. Once again, a corporation's quality philosophy will dictate us which index is more appropriate in a specific situation. 6.2 Estimating and testing Cpmk based on a single sample 6.2.1 Estimation
and the distribution
of the estimated
Cpmk
For a normally distributed process under statistical control, Pearn et al. (1992) suggested using the natural estimator of (j , defined as:
Cpmk = min
USL-x
x - LSL
3^/4 + (x - Tf ' 3V4 + (j _ Tf
\x-T\ / 2 + (j _ 3% s
Tf
where as above £ = ^ ™ £ i / n and s„ = ^T^™ (ZJ — x)2 / n are the maximum likelihood estimators (MLEs) of fi and a2 ,
118
Encyclopedia
and Handbook
of Process
Capability
Indices
respectively. Note that si + fa - T)2 = ^™_ fa - T)2 / n , which is in the denominator of Cpmk , is the uniformly minimum variance unbiased estimator (UMVUE) of a2 + (/i - T)2 = E[(X - T)2} appearing in the denominator of Cpmk . In fact, the natural a s of C.pmk can be expressed ^pmk — estimator 1 2 {D-VH)/(3yJK + H), where D = n l d/a , K = ns2n /a2 , and H = n(x — T)2 J a2 . Under the normality assumption, K is distributed as Xn-i> a chi-squared variable with n — 1 degrees of freedom, H is distributed as xi2\ a non-central chi-squared variable with one degree of freedom and the non-centrality parameter A = n(fi — T)2 / a2, and -Jll is distributed as a normal variable N(rj,l) with the mean 77 and variance 1. Here 77 = 1/2 fi — T I /(a j -Jn) . In other words, the natural estimator n Cpmk is distributed as a mixture of the chi-square and the noncentral chi-square variables expressed as: d-Jn a
C.pmk
, - Xl,A
3 ^ / X n - l + Xl 2 A
with Xn-i a n d Xi, A a r e a s defined above. In the general case (with T = m ) , Pearn et al. (1992) derived the r-th moment of Cpmk to be: -A/2 E C
( pmk)
= ~ 2= 0
00
(A\J
V I
r ( i±i
+ j) T{-^I + j)
x v i^L r(i + i) r ( ^ + ;) 3 = 0 J' Consequently for r = 1 and 2, we have
119
The Cpmk Index 3-A/2
EyCpmk)
3
oo / MJ j
(f) d m £ /=-„ r-[
n^ + j) r ( f + j)
i!r(f+ ;) r(i + J )r(^i + j)
(6.3)
and
Var(Cpmk) = E(C2pmk) -
[E(Cpmk)f
where
E(C2pmk)
9
^
d-J2n
i'i
dJnY a J ' n + 2j -2
J]T(l + j) n + 2 j - l
l + 2j + n + 2j
(6.3)
The bias and the mean square error (MSE) of Cpmk can be expressed as: BiasyCpmk) = E[Cpmk) MSE(Cpmk)
= Var{Cpmk)
— Cpmk , -
[Bias(Cpmk)}2,
Evidently Cpmk is a biased estimator of Cpmk (see (6.3) and (6.1)). Pearn et al. (1992) calculate the expected values and variances of Cpmk for d/a= 2(1)6, \n - T\ /a = 0.0(0.5)2.0 for the sample sizes n = 10(10)50. As \/J, — T\ /a increases, both the values of E(Cpmk) and Cpmk decrease for fixed d /a . Moreover, as \[i — T\/a increases, the variance of Cpmk is decreasing for a fixed sample size n and dja (see Pearn et al. (2003)). These results indicate that the bias of Cpmk is positive when fj, ^ T , i.e., Cpmk is generally overestimated by Cpmk . Furthermore, when
120
Encyclopedia and Handbook of Process Capability Indices
pi ^ T , both the bias and MSE of Cpmk decrease as |/x — T\ /a increases for a fixed sample size n and a fixed d / a . Constable and Hobbs (1992) have discovered that, even if the underlying distribution of the data is normal, small samples do not result in acceptable estimates of the mean and standard deviation of the process. A proper choice of sample sizes for capability estimation is of vital importance. The smaller the sample size, the higher Cpmk ought to be to d e m o n s t r a t e t h e t r u e process capability. However, one should keep in mind that when the sample size n is small or moderate (10 or even 50), the MSE of Cpmk is quite large as compared with the corresponding actual value of Cpmfc . For example, for |/z — T\ /a = 0 and d/a = 6, Cpmk = 2.00 while ^MSE(Cpmh) = 0.559 for n = 10 a n d •sJMSE{Cpmk) = 0.210 for n = 50. This shows that it is advisable to construct confidence intervals or carry out appropriate hypothesis tests. Chen and Hsu (1995) investigated the asymptotic sampling distribution of the estimated Cpmk as n —» oo and have shown that the estimator Cpmk is a consistent, asymptotically unbiased estimator of Cpmk and, provided the fourth moment of the distribution of the characteristic X is finite, Cpmk i s a l s o asymptotically normal. Vannman and Kotz (1995) obtained the distribution of the estimated "hyper index" Cp (u, v) for cases with T — m . By taking u = 1 and v = 1 , one arrives at the distribution of C p (1,1) = Cpmh . Vannman (1997) lateron derived a simplified form for the obtained distribution. Wright (1998) has obtained an explicit but rather complicated expression for the probability density function of the estimated Cpmk . Furthermore, Pearn et al. (1999) investigated an extension of Cpmk to handle processes with asymmetric tolerances (T ^ m). Under the normally assumption, Pearn et al. (2001) derived
121
The Cpmk Index
explicit forms of the c.d.f. and the p.d.f. of the estimator Cp'm& with asymmetric tolerances. If the manufacturing tolerance is symmetric (T = m), the c.d.f. of C ^ . reduces to
y
o,
-~r
poo
(D-JSD(y)tf
{
c
W -l-
xfn( SD(y)t)sD(y)dt, FfiAv)
y= o l-6)
+
\lFK
9y
Jo
x
$(.-D-6),
fH(SD(y)t)SD(y)dt,
y > o.
FK(-) is here the c.d.f. of a Xn-i variable. (Recall D = nl^2d/a, K = nsl/a2, H = n{x - T)2 /a2 .)
that
The corresponding p.d.f. of Cpmk becomes:
u*
9y
fa \SDW)
^-i
^ - 3 < y< °
fe Av) ''pink
]f«
9y2
x fH[sD(y)t)
SD(y)t
—$
dt
,
v > °>
122
Encyclopedia and Handbook of Process Capability Indices
where SD(y) = [D/(l + 3y)2} , FK{-) is the c.d.f. of K (a x«-i variable), fK(-) is the p.d.f. of K, and /#(•) is the probability density function of H (a non-central %2 variable). Using a variable transformation and the integration technique similar to those presented in Vannman (1997), the c.d.f. and the p.d.f. of the estimator C*pmfc may alternatively be expressed in terms of a mixture of the chi-square distribution and the normal distribution. The explicit form of the c.d.f. considerably simplifies the analysis of the statistical properties of Cpmk . The c.d.f. of Cpmj(. can be expressed as
Lpmk^'
Jo
I
9y2
x [4>{t + £Vn) + <j){t - £^n)]dt,
(6.4)
for y > 0, where b = d /a , £ = (/z — T)/ a, G(-) is the c.d.f. of a chi-squared random variable with n — 1 degrees of freedom. Xn-i and (/>(•) is the p.d.f. of the standard normal variable iV(0,1) (see Pearn and Lin (2002)). As in the case of the Cpmk p.d.f. (Chapter 4) we would obtain the same equation when we substituting £ by —£ in equation (6.4) for fixed values of x and n . Note that for it > USL or [x < LSL , the capability index Cpmk < 0 , and for li = USL or \i = LSL , it is 0. The restriction LSL < /j, < USL has been a minimum capability requirement which is valid in most of start-up engineering applications or of new processes. Jessenberger and Weihs (2000) studied the behavior of Cpmk for processes with asymmetric tolerances. Pearn et al. (2001) obtained an alternative but simpler form of the p.d.f. of Cpmk and investigated capability testing procedures in these cases. Pearn et al. (2002) studied statistical properties of the estimated Cpmj. . Pearn and Lin (2002a) have focused on a Bayesian-like estimator of Cpmk under a different manufacturing
123
The Cpmk Index
condition when the probability P(/i > m) is available. Pearn and Lin (2002b) also developed efficient SAS/Maple computer programs to calculate the critical values and the p-value for testing manufacturing capability based on Cpmk •
6.2.2 Confidence
intervals
on Cpmk
As mentioned above, Chen and Hsu (1995) examined the asymptotic sampling distribution of the estimated Cpmk and have shown that the estimator Cpmk is consistent, asymptotically unbiased estimator of Cpmk , and (provided the fourth moment of the distribution of the characteristic X is finite), Cpmk is also asymptotically normal. In particular, an asymptotic 100(1 — a)% confidence interval on Cpmk derived by Chen and Hsu (1995) is a n
^pmk
a
pmk A ~f=~i^pmk
z
a/2
pmk
z
af2
•Jn
•Jn
where
~2 Vpmk
72<52 + d
26
_
\V{l + 62)
+ 3(1 +
<52)3/2 J
C,pmk
+
ra4
2\2 72(1 + 6')
-i
r2 ^ pmk •
where a^mk is the asymptotic estimator of Var(Cpmk). za/2 is the upper a / 2 quantile of the standard normal distribution, and m4 = Y?i=Sxi
sn =Y^=i(xi
-x)4 /n , 8 = (x
~xf
-T)/sn,
/n.
Furthermore, based on the c.d.f. of Cpmk as expressed in (6.4), Pearn and Shu (2003) further developed an efficient
124
Encyclopedia and Handbook of Process Capability Indices
algorithm with a Matlab computer program to obtain the lower confidence limits that convey important information regarding the true process capability. For processes with the target value set at the mid-point of the manufacturing specifications (T = m , the symmetric case), the index Cpmk can be rewritten as _ Pmk
d-\fi-T\
_d/a - | £ |
" sja* + (M - Tf " SVTK 1 '
(
}
where the parameter £ = (/U — T) / a . For Cvmk = C , b = d / a can be expressed as & = 3Cv 1 + £2 + | £ | . Since the process parameters /J, and a are usually unknown, £ has to be estimated. Indeed for testing procedures for Cpm as presented in Cheng (1992) one can estimate £ by calculating the values of £ = (x — T) / sn from the sample. Hence, given the sample of size n , the confidence level 7 , the estimated values Cpmk and £ , the lower confidence bound Cpmk can be obtained using numerical integration technique with iterations be solving the equation bJn/(l+3C2pmk)
G
/ ,0
(ftVn - tf 9CU
2
^
x[<j>(t + t-Jn) +
=I - 7 .
(6.6)
It should be noted that the l.h.s. of the equation is an even function of £ . Consequently, for both £ = £0 and £ = —£0 we obtain the same lower confidence bounds.
6.2.3 Hypothesis
testing
with Cpmk
Using the index Cpmk engineers can access the process performance and to monitor the manufacturing processes on a
The Cpmk Index
125
routine basis. To test whether a given process is capable, we shall consider, as it was done on the previous occasions, the following statistical testing hypotheses: HQ '• Cpmk < C (process is not capable), H\ '• Cpmk > C (process is capable). where C as above is a given value of capability requirement. Based on a given a(c0) = a (explain c0 and a), the chance of incorrectly classifying an incapable process (Cpm): < C ) as a capable one ( Cpmk > C ), the decision rule is to reject H0 ( Cymk < C ) if Cpmk > c0 and not reject it otherwise. For symmetric cases with T = m , let Cpm/t = C then the b = d/a expressed in (6.5) is rewritten as b = 3C\Jl + £ 2 + | £ | . Hence, given values of the capability requirement C , parameter £ , the sample size n , and the risk a , the critical value CQ can be obtained by solving the equation P(Cpmk > c0 | Cp7n& = C) = a using standard but tedious numerical methods. In other words:
6Vn/(l+3co)
/ ,o
nx +"\2 _ {(b^n-tf t2
((h
9c 0 2
x [
(6.7)
where G(-), as above, is the c.d.f. of Xn-2 • Similarly, given values of C , n , and a , the critical values CQ for £ = £0 and £ = — £0 are the same since equation (6.7) is an even function of £ . In addition, given a value of the capability requirement C , the p-value corresponding to c , a specific value of Cpmk , calculated from the sample data, is (using equation (6.6)):
126
Encyclopedia and Handbook of Process Capability Indices
p - value = P(Cpmk > c \ Cpmk = C) 6Vn/(l+3c*)
= /,
{{b^n-tf { 9(cy
_t2
x[<£(i + £Vn) + (/>(£-£Vn)]d£.
(6.8)
[the readers ought to distinguish between the symbols C (related to Cpmk ) and CQ associated with Cpmk}. As we have already pointed out on several occasions, since the process parameters (i and a are unknown, so is £ which ought to be estimated in real-world applications. Such an approach would certainly be less reliable as the confidence level 7 may not be ensured. To eliminate the need for further estimation of the parameter £ , Peran and Lin (2002) also examined the behavior of the critical value c0 as a function of the process characteristic £ . They carried out extensive computations to calculate the critical value c0 for £ = 0(0.05)3.00, Cpmk = 0.7(0.1)3.0, and n = 5(5)200. Observe that the values £ = 0(0.05)3.00 cover a wide range of possible applications for the process capability Cpmk > 0. The results of these investigations indicate that the critical value c0 attains its maximum at either £ = 0.50 (for the most cases), or at 0.45 (in a few cases), and the difference between the two critical values is less than 5 x 10~4 . For instance, it follows from Figure 6.1., the critical value c0 first increases with £ attaining its maximum value at £ = 0.45 or 0.50, and then decreases within the range of £ G [0.5,3.0] . Hence, for practical purposes, we may solve equation (6.6) for £ = 0.50 to achieve the required critical values for given Cpmk > n a n d 7 , without the need to estimate further the parameter £ . Consequently, the confidence level 7 can be ensured and decisions carried out based on such an approach should be reliable.
127
The Cpmk Index
0
0.5
1
1.5
2
2.5
3
Figure 6.1. Plots of CQ versus | £ | for C = 1.00, a = 0.05, and n = 30, 50, 70, 100, 200, 300 (from top to bottom)
If the estimated value Cpmk is greater than the critical value Co ( Cpmk > cQ) or the calculated p-value is smaller than a (pvalue < a ), we conclude that the process meets the capability requirement ( Cpmfc > C ). Otherwise, there is no sufficient information to conclude that the process meets the specified capability requirement. In this case, we would tend to believe that Cpm). < C . In fact the approach described above ensures that decisions based on those critical values would be more reliable than on other available methods of estimating £ by means of the sample values only.
6.3 Estimating and testing Cpmk based on multiple samples From a practical perspective, manufacturing characteristic information about the process control may be obtained from estimating the process capability using the past in-control data which includes multiple samples (rather than a single sample) without recalculation on resampling. Many of modern manufacturing factories have implemented a daily-based production control plan for monitoring/controlling
128
Encyclopedia and Handbook of Process Capability Indices
process stability. A routine-basis data collection procedure is carried out to run the X and S control charts (for moderate sample sizes). The past "in control" data consisting of multiple samples of m groups, with variable sample sizes n^ = (xn,xi2,...,xin.), is then analyzed in order to compute the (overall) manufacturing capability. Thus, manufacturing information regarding the product quality characteristic is derived from multiple samples rather than from a single one. Under the assumption that these samples are taken from the normal distribution N(/j,,a2) , we consider the following estimators of the process mean and the process standard deviation:
for the i-th sample mean and the sample variance, respectively. Then, the overall sample mean X = ^ nj-Xj / N , n an N = ^2 •_, % d the pooled sample variance S2P = n i 2 i i 5 Ar m are the
Er=i( '- )^ /Er=i^- )=Er=i^- ) '/( - )
unbiased estimators of fi and a2 , respectively, used for calculating the manufacturing capability Cpmk . For cases with multiple samples the natural estimator of Cpmk is given below. The derivations of the sampling distribution, lower confidence bounds, and the manufacturing capability calculations for cases with multiple samples can be performed using the similar techniques for the situations with a single sample, although the derivations and calculations are usually more tedious and complicated (See e.g. Vannman and Hubele (2003) and Pearn and Shu (2003) for more details). The "natural" estimator of Cpmk is: Cpmk —
m m
USL-1
1
-LSL
2
^ 3yjS p + (X - Tf ' &Js* + ( f - T)2 '
Using the similar technique for cases with a single sample, the
129
The Cpmk Index
c.d.f. of Cpmk can be expressed in terms of a mixture of chisquared and normal distributions:
Fd, (x) = i - r
&V77/(l+3z) •if
9x2
[ # + ^v]v) + (/>(i-eVAr)]^ for x > 0 , where b = d/a , FK{.) is the c.d.f. of the (ordinary central) chi-square distribution XN-m > a n d >(•) is the p.d.f. of t h e s t a n d a r d n o r m a l d i s t r i b u t i o n 7V(0,1). As m e n t i o n e d above, under the condition with T = m , the index Cpmk may b e r e w r i t t e n a s Cpmk = (d/a— | £ | ) / ( 3 v l + £ 2 ) , w h e r e £ = (/i — T)/cr . T h a t i s , w h e n Cpmj. = C , 6 = dVer c a n be expressed as b = 3CyJl + £ 2 + | £ |. Namely, given t h e t o t a l sample size N with m subsamples, the confidence level 7 , the estimated value Cpmk , and the parameter £, the lower confidence bound Cpmk c a n be obtained using numerical i n t e g r a t i o n techniques with iterations, to solve t h e equation (6.7). We r e e m p h a s i z e t h a t t h e p a r a m e t e r £ = (// —T)/cr is usually unknown_in practice, but can be estimated from the sample data as £ = (X — T) J Sp. Again it should be noted that we have the same lower confidence bounds for both £ = £0 and £ = — £0 since t h e l.h.s. of e q u a t i o n ( 6 . 8 ) is a n even f u n c t i o n of £ : 6777/(1+3(7*pmk
/.
) •K
9CZk
[4>{t + £ V F ) +
=I - 7 .
(6.7)
Pearn and Shu (2003) also investigated the behavior of the lower confidence bound Cp^nk versus the process characteristic parameter £ based on multiple samples. The results indicate that
130
Encyclopedia and Handbook of Process Capability Indices
the lower confidence bound first decreases as £ increases, attains its minimum value at £ = 0.45 or 0.5 (with the difference of less than 0.001), and then, in all cases, increases again within the range of £ 6 [0.5, 3.0]. Hence, for practical purposes we may solve equation (6.7) with £ = 0.5 to obtain the required sample accuracy for given N , m and 7 , without the need to estimate the parameter £ . The decision making procedure just described ensures that the risk of a wrong decision will be not greater than the preset Type I error 1 — 7 .
6.4 Bayesian-like estimation of Cpmk We have always indicated that Pearn et al. (1992) proposed the capability index Cpmk , and also investigated the statistical properties of it's natural estimator Cpmk for stable normal processes with a constant mean \i. Chen and Hsu (1995) showed that under general conditions the asymptotic distribution of Cpmk is normal for /J, ^ m , and is a linear combination of the normal and the folded-normal distributions whenever \x = m , where m as is the mid-point between the upper and the lower specification limits mentioned above. In real-world applications the production may require multiple supplies with different quality characteristics in each single shipment of the raw materials, multiple manufacturing lines with sometimes inconsistent precision in the machine settings and engineering effort for each manufacturing line, or multiple workmanship shifts with variable performance level at each shift. Thus, the fundamental and common assumption that the process mean remains a constant may not be satisfied in real-world situations. Consequently, using the natural estimator Cvm-k to measure the potential and performance of such process may be inappropriate as the resulting capability measure may very well be inaccurate. For stable processes under these
131
The Cpmk Index
conditions, if we have knowledge about the process mean, P(/j, > m) — p , 0 < p < 1 , then we may consider the following novel estimator Cpmk : (Usually, the probability P([i > m) = p , 0 < p < 1 , can be obtained from historical information about a stable process.)
bn^[d-(X-m)IA(fi)] ^S2n+(X-T?
apmk
here &„_i = V2/(n - l){r[(n - l)/2]/T[(n - 2)/2]} is the correction factor, IA(-) is the indicator function defined as IA (M) = 1 f ° r M £ A, and IA (/J,) = — 1 for /i $ A , where the set A = {(i\n > m}. Observe that the new estimator Cpmk can be rewritten as: p
_ h-M ~{X - rn)IA(n)}
Lpmk
ZJSI +(Xd-(X_
m)IAM Sb^Sn
1+ t
where X = Y.',sxi
Tf
^ l l
_
Cpk
.k+ i l ^ l
I'" •*„ = E,",JX-
~ Xf
A B " *> '
And Cpk = bn_x[d — (X — m)IA(n)]/(3Sn) was introduced by Pearn and Chen (1996). If the process characteristic follows the normal distribution, N(fj,,a2) , then the new estimator Cpmk is distributed as:
132
Encyclopedia and Handbook of Process Capability Indices
bn-lt'n-l(f>)
c.pmk
3Vr 1 *i', -i(A) 1+ n n 1
where tn_x (6) is as above the non-central t distribution with n — 1 degrees of freedom and the non-centrality parameter 8 = 3VnCpfc , and iJl)„_i(A) is the non-central F distribution with (1, n —1) degrees of freedom, and the non-centrality parameter A = n((J, - Tf J a2 (Pearn and Lin (2003-Metrika)). Moreover, if the knowledge about the process mean, namely the probabilities P(fi > m) = p , and P(/x < m) = 1 — p , with 0 < p < 1 is known, then the asymptotic distribution of the new estimator Cpmk turns out to be a mixture of two normal distributions. The estimator Cpmk is consistent and asymptotically unbiased. Consequently, when either P(/x > m) = 1, or P(/j, > m) = 0 , the asymptotic distribution reduces to an (ordinary) normal distribution. Under center general conditions, namely if the fourth central moment /x4 = E(X — fj,)4 exists, the distribution of the standardized quantity Sn{Cpmk — Cpmk) converges to 2 2 in distribution, where the p • N(0,a pmkl) + (1 p) • N(0,apmk2) varmces: A2 _2
_
1+
^1
(M - Tf
( A i
3 ~ a3
a 1
A 2 "pmk2
_
^
4
a
2
,
^2 Q
<7
, 1 H4 - P <
.
A
a
A
°2 1 a
pmfcl
In (M- T2 ) <^ 1 2
2 M3
3 ~ a3
2 1 + (M - T)
-3/2
r2
i+fc^
/I
[ 1l l1
_T)2i
i i
1-2
c;pmk2
-3/2
c,pmk2
The Cpmk Index
_ 9Qz -
T)C2pmkl
133
d - ( / z - m) 2 3^/a 2 + (jz - T ) '
,
and
A
A2 =
9 ( M _ T)Cpmk2 , , ,
r
_
J-, ^pmk2
d + (/x - m) • — „ /—5
;
=75
Chapter 7
The Snh Index
7.1 Process capability and the Spk index 7.2 Estimating and testing Spk based on a single sample 7.3 Hypothesis testing with Spk 7.1 Process capability and the Spk index Process yield has for a long time been a standard criterion used in the manufacturing industry as a common measure of process performance. Process yield is currently defined as the percentage of processed product units passing inspection, which indicates that the product characteristic falls within the manufacturing tolerance. Product units that are rejected (nonconforming) result in additional costs incurred to the factory for scrapping or repairing the product. However a factory does not induce additional costs for a product that departs from the nominal specifications but is still within the tolerance. For processes with two-sided manufacturing specifications, the process yield can be calculated as p = F(USL) — F(LSL) , where USL and LSL are the upper and the lower specification limits, respectively, and F(-) is the c.d.f. of the process characteristic. If the process characteristic is normally distributed, the process yield can then also be expressed as p = $[(USL — //)/er] — $[(LSL — fj,)/ a] , where using the common notation ji is the process mean,
The S„k Index
135
As it was already mentioned in attempting to provide a common quantitative measure of process performance, Kane (1986) one of the pioneers of the process capability analysis utilized the capability index Cpk to determine process ability of reproducing items within the specified manufacturing tolerance. The index Cpk was specifically designed for processes with two-sided manufacturing specifications requiring knowledge of both USL and LSL , It is defined as: Cpk = min
USL-
n fi
-LSL
(c/. Chapter 4) We have also seen in Chapter 3 that for normally distributed processes, the relationship between the process yield p and the index Cpk is p > 2<£(3Cpfc) — 1. Thus, the index Cpk provides us with an approximate, rather than exact, measure of the actual process yield. This important observation allows us to develop the yield index Spk for normal processes, which can be interpreted as a smooth version of Cpk . This new index was proposed by Boyles (1991, 1994) allows a one-to-one transform to fraction nonconforming. The index Spk is defined as:
where $ - 1 (-) is the inverse function of standard normal c.d.f. $(•). The index Spk establishes a relationship between the manufacturing specifications and the actual process performance, which provides an exact measure of the process yield. If Spk = c, then the process yield can be expressed as p = 2<E>(3c) — 1. There is eventually a one-to-one correspondence between Spk and the process yield p . Consequently, the index Spk provides an exact (rather than approximate) measure of the process yield, (under the normal assumption).
136
Encyclopedia and Handbook of Process Capability Indices
We note that the Spk can be rewritten and expressed as a function of the basic indices Cp and Ca . Recall the index Cp defined as Cp = (USL — LSL) /(6a) which measures the overall process variation relative to the specification tolerance and reflects only the process precision (consistency). Similarly the Ca defined as Ca = 1— | ii — m | / d, measures the degree of process centering, where m = (USL + LSL) / 2 is the mid-point between the upper and the lower specification limits, and d = (USL — LSL) / 2 is the half of the length between the specification limits. The index Ca thus alerts the user if the process mean deviates from its target value. An informative mathematical relationship between 5pfc and Cv and Ca is easily derived to be: Spk =
^
V
2 $
\ [d-
+
V
m
a
¥
i.
fJ-
a
(i-
= ^-1[\^(3CpCa)
jj, — m
a/d
+
4*
LSLU a
(d+
Id)
+ £•
V a
m
' 1+ | n — m | / d a~fd
^(3Cp(2-Ca))Y
For a process where the nominal is equal to the process mean (T = m), the manufacturing tolerance is 3Spk standard deviations from the mean. For example, if Spk =1.00 , then USL = ji + 3(7 and LSL = JJ, — 3a , (this situation is sometimes called the natural tolerance). If Spk = 2.00 , we arrive at by the six-sigma principle: USL = [i + Qa and LSL = /j, — 6a , the well known 6a limits popularized, (and sometimes abused by the socalled six-sigma principle). 7.2 Estimating and testing Spk based on a single sample 7.2.1 Estimation
of Spk
In practice, the process parameters fi and a are unknown and have to be estimated from the sampled data. To estimate the
The S„t Index
137
yield measure Spk , we consider the following natural estimator involving the statistics x = ^ . _ xt / n and 'pfc 2 1 2 s = [Y^._ (XJ — S") /(n — l)] ' are the sample mean and the sample standard deviation being the conventional estimators of \i and a , respectively, obtained from a well-controlled (being definitely in statistical control) process. The estimator is evidently Spk = ~$~
2
\
+
s
l*
T5)
The exact distribution of Spk (even under the normal distribution) is mathematically intractable. Consequently, as it stands testing the process performance cannot be accomplished. On the other hand, Lee et al. (1999) obtained an approximate distribution of Spk using the Taylor expansion technique. Let Z = ~Jn(x — //) and Y = Vn (s 2 — a2), then Z and Y are independent since the first two moments of x and s2 exist, it can be shown that Z and Y converge by the Central Limit Theorem to N( 0, a2) and N(0,2a2) , respectively, as n goes to infinity. Accordingly, applying the first order Taylor expansion, the estimator o pfc can be expressed approximately as: Spk — Spk +
where
W =
\
6Vn
Op(n-1),
(7.1)
(1 Y (l + 6)
Z_ ,<7
(1 + 6
1-6
4>
7
->
7
(7.2)
and as above 6 = (fj, - m) / d and 7 = a / d . Note that the expression (7.2) can be written as W = sjn,/2[a(s2 — a2)/a2} — -Jn[b(x-^)/a}=[aY/(-/2a2)}-(bZ/a) for n < m , and W = 4^j2[a(s2 - a2)/a2} + Vn[6(z - n)/a] =[aY/(V2a2)] +(bZ/a) for \x > m , where a and b are functions of n and a (or
138
Encyclopedia and Handbook of Process Capability Indices
equivalently of 6 and 7 ) as defined below and (/>(•) is the p.d.f. of the standard normal variable N( 0,1). 1+6 d } a = y/2cr d + 6)
Here
and
b =
0
[1
+ a-s)
{ 7J
1 +
^ ) -
6
)
(7.3)
(7.4)
{ 7J
Consequently, the statistic W (by Central Limit Theorem) is asymptotically normally distributed with mean 0 and variance a2 + b2 , and the natural estimator Spk is approximately (asymptotically) distributed according to N([j,s,crg), where: E(Spk) = Ms = Spk Var(Spk) = a2 = (a2 + b2)/{36n[»(3^)]2} •
and
The asymptotic distribution of the estimate of Spk variance that decreases with n . Thus, the estimator 5 ^ is asymptotically unbiased. As the sample size increases both the sample mean and the sample variance approach to their expected values. The remaining term in (7.1), O p (n _ 1 ) , represents the error of the expansion having a leading term of order n^ 1 in probability. Hence, the approximation is also consistent. In fact, the parameters a and b expressed in (7.3) and (7.4) can also be rewritten as functions of the precision index Cp and the accuracy index Ca in the following manner: d (1 + S) a = V2cr (l + 6)
+ (!-<$)>
- ^ [ 3 C p ( 2 - Ca)4>{ZCp{2 - Ca)) +
(1 - / P
b = 4>
=
(1 + 6)}
-0 i 7 J { 7 JJ
1-6 7 3CpCa
139
The S„k Index
(The identity: 3CpCa —
1+6 7
is worth noting). Thus, the p.d.f. of the approximate distribution, N^fis,^), can easily be derived to be: (see Pearn, Lin and Wang (2004))
f{x) =
J—4a OO <
2
X <
2
+b
exp
18n(fl3S pfc ))' a2+b2
f {X
_
2 Pk)
OO .
The formula of the first-order approximation of Spk as given in (7.1), is somewhat clumsy, but it produces an adequate approximate distribution for a sample size large enough. Lee et al. (1999) indicate that the approximate and exact distributions are almost indistinguishable as the sample size n reaches 1000 or more. In fact, even for n = 100 the approximation is quite reasonable for practical purposes. However, the accuracy of the approximation is difficult to investigate since the parameters a and b are required to be further estimated from the sample, resulting in a loss of reliability. Thus, the analytical approximation would not be practically useful in practice until this difficulty has been resolved. Moreover, using rather complicated and tedious algebraic manipulations Pearn et al. (2005) have shown that also, the estimator Spk can be expressed in the form of: Spk = Spk + DXZ + D2Y + D3Z2 + D4ZY + D5Y2
+Op'
n~/n i
here Z and Y are distributed according to the joint bivariate normal distribution, and Di , i = l,2,---,5 , are functions of
140
Encyclopedia and Handbook of Process Capability Indices
6 = (fx — m) / d and 7 = a / d . (These are the basic functions for the specification bound and the mean.) Therefore, the distribution of Spk may alternatively be approximated by the following polynomial combination of the distributions of Z and Y: (Z, Y) -^—^
N((0,
0), S 2 ) , where S 2 =
0 1/2 '
with the bias approximated as: DXZ + D2Y + D3Z2 + D4ZY + D5Y2.
The formulae for the both approximations to the distribution of Spk are quite complicated, and the calculation is somewhat cumbersome. As mentioned above, since the parameters a and b in the approximate formula(s) ought also be estimated in real world applications, a substantial uncertainty may be introduced into the assessments of the performance due to the additional sampling errors resulting from the estimation of a and b . Furthermore, the accuracy of the approximation has not yet been investigated. Consequently, the approximation may not be useful in practice until these issues are successfully resolved. To analyze the actual accuracy of the natural estimator Spk, Pearn and Chuang (2003) investigate the accuracy computationally, using the simulation techniques to obtain the relative bias and the relative MSE for certain common performance requirements, where the relative bias is defined as BiasR(Spk) = [E(Spk) - Spk}/Spk = E(Spk / Spk) - 1 . It measures the average relative (percentage) deviation of Spk from the true value Spk . Moreover the relative MSE is defined as MSER(SSpk)=E[(Spk-Spk)/ Spk f =E[(Spk / Spk) - l ] 2 ; it measures the average of the squared relative deviation of Spk from the true
The S„k Index
141
value Spk . These authors have further considered the statistic [MSER(Spk)f/2 the square root of the relative MSE, which is a more direct measurement of the relative deviation (i.e. the percentage of deviation). Unfortunately either explicit or implicit mathematical formulae for BiasR(Spk) and for MSER(Spk) are analytically intractable. The simulation approach seems to be a best alternative for the accuracy assessment. For commonly used performance requirements which are Spk = 1.00, 1.33, 1.50, 1.67, and 2.00, each combination of the precision measure Cp , the accuracy measure Ca, and the (/J,, a) pair is first assigned generate random normal samples of size n . The sample data is then calculated to obtain the estimator Spk • A total of N =10000 replications are repeated for each sample size of n— 5(5)100, next the average value E(Spk) is computed and is compared with the preset true Spk to calculate the relative bias. The simulation error was also examined, and resulted in not being greater than 5 x 10~3 . These simulation results clearly indicate that the estimator Spk overestimates the true value of Spk in all cases that have been investigated. The amount of the overestimation, in terms of the relative bias, BiasR(Spk), appears to be increasing in Cp at the beginning, and then remained stable after roughly Cp = Spk + 0 . 2 . Particularly, this is valid in all cases for exceeding 15 . After the sample size n > 15 , the fluctuation of BiasR(Spk) remains less than 0.5%, and is no greater than 0.10.2% for n > 60 . This pattern is also apparent for [MSER(Spk)}1/2. Indeed in most cases, the magnitude of the deviation is increasing in Cp at the beginning and then becomes stable approximately after Cp = Spk + 0.1 , particularly, for n > 20 . (See e.g. Pearn and Chuang (2003)) The above described simulation results are useful for practitioners/engineers to measure the process performance for factory applications, especially, if the processes are controlled/monitored on a permanent routine basis.
142
Encyclopedia and Handbook of Process Capability Indices
7.2.2 Confidence
intervals
for Spk
Based on the first-order approximation as given by Lee et al. (1999-QREI), a 100(1 — a)% confidence interval can be expressed as:
, _ bpk
Za/2
Va 2 + b2
.
6^(3SpkY
bpk
JtfTb2 +Za/2
6Vncf>(3Spk)
where as usual za/2 is the upper a/2 quantile value of the standard normal distribution, and the parameters a and b are estimated from the sampled data. In addition, a 100(1 — a)% lower confidence bound of Spk can also be obtained from the lower (one-sided) confidence limit as: Spk -zax
(a2 + 6 2 ) 1 / 2 /[6Vn0(35 pfc )].
If the calculated lower confidence limit turns to be greater than the predetermined index value, the process is judged to be capable. Otherwise, the process is classified as incapable, and approximate quality improvement activities ought to be initiated. More recently, Pearn and Chen (2003) studied the natural estimator of Spk and investigated computationally its statistical properties, using the bootstrap simulation techniques to obtain the lower confidence bound in order that practitioners/engineers can utilize them to perform quality testing and to determine whether the process meets the preset quality requirements. These authors compared the four following bootstrap methods including (a) the standard bootstrap (SB), (b) the percentile bootstrap (PB), (c) the biased corrected percentile bootstrap (BCPB), and (d) the bias-corrected and accelerated (BCa) bootstrap methods based on the coverage proportion. For each one of these methods, they have calculated the proportion of the simulation runs for which the lower confidence bound is smaller than the corresponding Spk value. This "actual coverage proportion" (viewing the lower bound Spk as a one-sided confidence interval (Spk,oo)), could then be
The Sfk Index
143
compared with the expected value of 1 — a , say 0.95. The simulation results show that the SB method substantially outperforms the PB, BCPB and BCa approaches in the case of normally distributed processes, and is therefore recommended to be used when the assessing process performance Spk based on the yield. Furthermore, for the SB method the coverage fraction is greater than 0.90 in all the cases investigated when the sample size n exceeds 25, and it is also more stable than the other three methods. Therefore, the SB approach ought to be the one recommended for the Spk index, especially for sample sizes greater than 25.
7.3 Hypothesis testing with Spk Based on the asymptotic distribution of Spk derived by Lee et al. (1999-QREI), a hypothesis testing for Spk can be constructed in a straightforward manner. For example, the testing of the null hypothesis Ho : Spk < C , C is a specified value versus an alternative hypothesis Hi '• Spk > C
can be carried out by considering the test statistic: _ 6(Spk -
C)^n
where a and b are the estimates of a and b, and Cp , Ca, /i and a are replaced by Cp , Ca , x and s , respectively. The null hypothesis H0 is reject at a level if the statistic T > za , where as above za , is the upper 100a% point of the standard normal
144
Encyclopedia and Handbook of Process Capability Indices
distribution. The proposed testing procedure is quite simple to use based on the normal approximation, but a substantial uncertainty (sampling errors) could be introduced into the performance assessments since using the value of d and b calculated from the sample data could be misleading. Calculation of the critical values To assess further the accuracy of testing procedures, Pearn et al. (2003) investigated the critical values using a standard simulation technique to compare with the critical values obtained from the normal approximation as proposed by Lee et al. (1999), under the same performance requirements. The comparison of the critical values CQ calculated by means of the normal approximation and the simulation technique for various selected sample sizes from n = 5 to 150 and various Spk values = 1.00, 1.33, 1.50, 1.67, 2.00, with the risk a = 0.05 . The simulation results indicate that the critical values are more sensitive to the two indices Cp and Ca than to those that are obtained by the normal approximation. For example, given the fixed value Spk = 1.67 with n = 20 , c0 = 2.24 for Cp = 1.67 , and we have c0 = 2.30 for Cp = 1.90 . Thus, we could choose the maximal values of c0 (among these of Cp and Ca ) to obtain conservative bounds on the critical values for reliability purposes. Such an approach ensures that the decisions made based on the critical values have the risk of (wrongly) specifying an incapable process as a capable, to be not greater than the preset type I error a . It also have been noted that the normal approximation substantially under-approximates the critical values, especially, for small sample sizes n < 40 , when the amount of underestimation exceeds 0.10. Therefore, for short run applications (such as accepting a supplier that provides short production runs in the QS-9000 certification), one should definitely avoid using the normal approximation. It was also observed that the underestimation might be as large as 0.07 for n = 50 , 0.03 for n = 110 , and 0.02 for n = 150 .
The Srk Index
145
Therefore, in real-life applications we should strive towards sample sizes larger than 150. Furthermore, sample sizes which are necessary so that the normal approximation converge to 5pfc assuming the values 1.00, 1.33, 1.5, 1.67 and 2.0 within sampling errors less than 0.10, 0.09, 0.08, 0.07, 0.06, 0.05, 0.04, 0.03, 0.02, 0.01, 0.00 respectively (with the accuracy up to the second decimal point, i.e. 5 x 10~3 ). For example, for Sp^ = 1.33 with risk a = 0.05 , a sample size of n > 95909 ensures that the sampling error is no greater than 5 x 10"3, which is negligible. Thus, if Spk > 1.33 , we then may conclude that the actual performance S^ is also > 1.33 . This type of assessment investigation is not practical in applications. However, these computations illustrate the behavior and the rate of convergence using the normal approximation. [1] The comparison of the power analysis (the coverage rate) using a number of approaches (1 st order, 2nd order approximations, standard simulation, bootstrap) has not been carried out as yet. [2] For multiple samples, 1st order approximation by Taylor expansion + an accuracy analysis. [3] For multiple samples, critical values and lower (coverage rate) comparison of several approaches.
bounds
[4] For multiple samples, the MPPAC also can be applied in the same way.
Chapter 8
The CPU/CPL Index
8.1 Process capability and the index CPU/ CPL 8.2 Estimating and testing CPU/CPL: Based on single sample 8.3 Estimating and testing CPU/CPL: Based on multiple samples 8.4 Estimating and testing CPU/ CPL: Bayesian approach with single sample 8.1 Process capability and the index CPU/ CPL We hope that by now our readers are familiar with a number of process capability indices (including Cp , Cpj. , Cpm and Cpmk) that have been commonly used in the manufacturing industry to measure whether a process is capable of reproducing product items within the specified manufacturing tolerance. However the indices described earlier are appropriate measures only for processes with two-sided specifications only (which require specifying both USL and LSL). For the unilateral tolerance situation where only one specification limit is provided, Kane (1986) one of the early pioneers of process capability studies considered the following indices
Cpu
_ USL- n ~ 3a '
n CpL
146
_ fi-LSL ~ 3a
The CPUICPL Index
147
where USL and LSL are as above the upper and lower specification limits, respectively, \i is the process mean, and a is the process standard deviation. The index Cpu measures the capability of a "smaller-the-better" process with an upper specification limit USL, while the index CPL measures the capability of a "larger-the-better" process with a lower specification limit LSL. Process
yield measure
based on Cpu and Cpi
For normally distributed processes with specification limit USL, the process yield is given by
= P(Z <3CPU)=
one-sided
$(3CPU),
where the variable Z follows the standard normal distribution N(0, 1) with the c.d.f. $(z) = (2-K)1'2 P exp(-i 2 /2)dt. Similarly, J —oo
for normally distributed processes with one-sided specification limit LSL, the process yield is
P(X>LSL)=P(^<^=-^) = p(-iz
148
Encyclopedia and Handbook of Process Capability Indices
(2001)). These recommendations are summarized in Table 8.1 For existing manufacturing processes, the capability should be at least 1.25, and for new manufacturing processes, at least 1.45. For existing manufacturing processes dealing with safety, strength, or critical parameters, the capability must be no less than 1.45, whereas for new manufacturing processes dealing with the same issues, the capability ought to be no less than 1.60. In recent years, due to globalization various companies have adopted criteria for various evaluations of their processes based on capability objectives that are more stringent than the recommended minima. For instance, the popular "six-sigma" program pioneered by Motorola essentially requires that when the process mean is in control, it will not be closer than six standard deviations from the nearest specification limit. In effect, it requires that the process capability ratio will be at least 2.0. Table 8.1. Some minima capability requirements based on C/ for new and special processes. Ci Value 1.25
Production Process Type Existing Processes
1.45
New Processes; or Existing Processes involving Safety, Strength, or Critical Parameters
1.60
New Processes involving Safety, Strength, or Critical Parameters
8.2 Estimating and testing CPU/ CPL: Based on a single sample
8.2.1 Estimation
of CPU and CPL
In practice, sample data ought to be collected in order to calculate the indices CPU and CPL since the process mean fi and standard deviation a are usually unknown (as it was indicated in
The Cpu/ CPL Index
149
the previous chapters). To estimate the indices Cpy and CPL , Chou and Owen (1989) proposed Cpy and Cpi , the natural estimators of CPU and Cpi, defined as: Cpu
_ USL — x ~ 35 '
f, CpL
_x— LSL ~ 3s '
where as above x = Y^. xt / n is the sample mean, and s2 = (n — l ) - 1 ^ . _ A % i — x)2 is the sample variance, obtained from a process that is "demonstrably" stable (under statistical control). Under the normality assumption, the estimator Cpy is distributed as (3Vn) _1 tn_i(6) , where £n_i(<5) is a non-central t random variable with n — 1 degrees of freedom and the noncentrality parameter 8 = 3^/nCpu . The estimator Cpi has the same sampling distribution as Cpy with 6 = 3-fnCpi . However, both the estimators CPU and Cpi are biased. Pearn and Chen (2002) have shown that by adding the correction factor &„_! = [ 2 / ( n - l ) ] 1 / 2 r [ ( n - l ) / 2 ] / T [ ( n - 2 ) / 2 ] to CPU and CPL , one obtains unbiased estimators bn_i Cpy and hn_\ Cpi which have been denoted as Cpy and Cpi. That is, E(Cpu) = CPu , and E(Cpi) = CPL- Since 6n_! < 1, Var(CPU) < Var(CPU) and also Var{Cpi) < Var{Cpi) . The estimators CPU and Cp^ are based only on the complete and sufficient statistics (x,s2), consequently Cpy and Cp£ are the uniformly minimum variance unbiased estimators (UMVUEs) of Cpy and CPL, respectively.
8.2.2 r-th
moment
The r-th moment (about zero) and the variance of Cpy can be obtained as follows: set Z = -Jn(USL — x) / a , this variable is distributed as N(3*fnCPy,l). It is easy to verify that E(CPU) = Cpy under the normality assumption. It follows from the definitions of Z and that of Cpy that
150
Encyclopedia and Handbook of Process Capability Indices
^_(r[(n-l)/2])r-1r[(n-l-r)/2]
F(r
Var(CPU)
r[(n-l)/2] r[(n-3)/2]
=-
(r[(n-2)/2])
2
1
T
2
PU)
, 1 r[(n-l)/2] r[(n-3)/2] 9n (r[(rz-2)/2])2 The results for the r-th moment, the expected value, and the variance of the dual estimator CPL are analogous.
8.2.3
Distribution
We are changing the variables possessing the non-central t distribution with v = n — 1 degrees of freedom and the noncentrality parameter 8 = 3VnC; . Let X = Cj = 6„_i(3Vn)~ t„(6). The c.d.f. of Cj can be derived directly as: F(x) =
1 2(»-3)/2r[(n-l)/2] 2
poo
x I y Jo y
„ UlL n z 2
e
1
p3^/nxy/bn_l^/n-l
—f== | V2Wo
exp
-f(.-tf
dudy,
Here 6 = 3VnC 7 and 6n_i is the correction factor to CPu . Differentiating the equation of the c.d.f. with respect to y gives the p.d.f. of Cj to as
/(*)
3Vn/(n-l)-2-"/2 bn__l^fT[(n-l)/2]
L
(n-2)/2
x,o „
exp
2/ +
3xyfny [bn-iJn-1
dy,
The CPUICPL Index
151
where, as above, the correction factor is bn_i = 1/2 [2/(n - l)] r[(n - l)/2]/T[(n - 2)/2] and 8 = 3VnC 7 . Figures l(a)-l(b) display the p.d.f. of Ct for various CI =1.00, 1.25, 1.45, 1.60, 2.00 for n = 10 and 30, respectively. Figures 2(a)-2(b) display the c.d.f. of Cj for the same values of C/ and n.
8.2.4 Testing
hypothesis
with CPU and CPL
For processes with one-sided manufacturing specifications (requiring the existence of only one of USL or LSL, but not both), the index Cpy can be used to measure the capability of a smallerthe-better process for upper specification limit USL, whereas the index Cpi to measure the capability of a larger-the-better process with a lower specification limit LSL. To test whether a given process meets the capability requirement, we shall consider the following statistical testing hypothesis with HQ :Cj < C (the process is incapable), versus the alternative Hi :Cj > C (the process is capable). Consequently, one can utilize the test (p(x) = 1 if Cj > C0 , and ip(x) = 0 , otherwise. The test
\Cl
152
Encyclopedia and Handbook of Process Capability Indices
As above the critical value, C0, is then determined by:
p{c 7 >c 0 \cI = c)
a
= P Ci(«) >
3VnC0
Cj=C
Hence, we have 3VnC0
= tn-i,a(8)
or C0 =
fyI_id,a(tf)/(3Vn).
where, as above, tn__ia{6) is the upper a quantile of a non-central t distribution with n — 1 degrees of freedom satisfying P{tn-i(fi) > ^n-i,a(^)) = a • As indicated above, the sampling distribution of the UMVUE, (5/ , is bn^ (S^/n)"1 £„_i(6) , a noncentral t distribution with n — 1 degrees of freedom and noncentrality parameter 6 = 3VnC/ . Thus, the power of the test can be calculated as *(CI)
= P(CI
>C0\CJ)=P
1^(6)
>
3-JnC0) u
n-l
Under H0 for EC[ (ip(x)) = a , we obtain that the critical value is C0 = bn^i(3-Jn)~1tn_iQ(6), where 6 = 3~JnCj . Therefore, the test ip is a uniformly most powerful (UMP) test of size ECr(
The lot is accepted, if The lot is rejected,
Gt > ^ - t ^ 3vn otherwise.
a(6
= 3VnC 7 ).
The CPUICpL Index
8.2.5 Lower confidence
bound for
153
CPU/CPL
Critical values are used for making decisions in capability testing with designated type I error a which is the risk of misjudging an incapable process (H0 : Cj < C) as a capable one (Hi :Cj > C ). The Rvalues are used for making decisions in capability testing, which presents the actual risk of misjudging an incapable process (HQ :Cj < C) as a capable one (Hx :Cj > C). Thus, if p-value < a we then reject the null hypothesis, and conclude that the process is capable with the actual type I error pvalue (rather than a ) . Both approaches, the critical values and the p-values, do not convey any information regarding the minimal value (the lower confidence bound) of the actual process capability. Determination of the lower confidence bound on the actual process capability is essential for quality assurance. The lower confidence bound not only gives us a clue about the minimal level of the actual performance of the process which is closely related to the nonconforming units, but it is also useful for making decisions in the course of capability testing. Chou et al. (1990) established the lower confidence bounds on Cpu and Cpi based on the natural estimators Cpu and CPL. Let USL = x + kis and LSL — x — k^s , where kx = 3CPU and &2 = 3CPL • A 1007% lower confidence bound Cy for Cpu satisfies Pr (CPU > CJJ) = 7 . It can also be written as: 7 =
p r /^£> c \ =
= Pr
P r /£±^it>^
Z — 3-JnCu > -ki^n) sJa
or, equivalently, lower confidence It can be shown and in_1 (^2) a r e
= Pv(tn_i(6i) > -VnfcO,
Pr(f n _ 1 (6 1 ) < -Juki) = 7 . Similarly, a 1007% bound Ci for Cpi satisfies Pr (Cpi > Ci) = 7 . that Pr(i n _ 1 (5 2 ) < k^Vn) = 7 , where tn_i(8i) non-central t distributed random variables with
154
Encyclopedia and Handbook of Process Capability Indices
n — 1 degrees of freedom and the non-centrality 61 = 3-JnCu and 62 = 3VnCL , respectively.
parameters
Observe that these two estimators are biased. Pearn and Shu (2003) developed an algorithm (including a Matlab program) for calculating the lower confidence bounds on CPU and Cpi based on the UMVUEs and provide tables for the engineers/practitioners to be used in measuring their processes.
8.3 Estimating and testing CPU/ CPL: Based on multiple samples A common time honored practice of the process capability estimation in manufacturing industry is to implement first a routine-based data collection plans for controlling/monitoring the process stability, and then to analyze the past "in control" data. Kirmani et al. (1991) considered estimation of a and of the precision index Cp for cases where the data are given as multiple samples. Pearn and Yang (2003) proposed an unbiased estimator of Cp for multiple samples, and showed that the unbiased estimator is the UMVUE of Cp , which is asymptotically efficient. Li et al. (1990) have studied the distribution of estimators of Cp and Cpk based on ranges of multiple samples. More recently, under the assumption of normality, Vannman and Hubele (2003) investigated the distributional properties of estimators of a wide class of capability indices, containing the indices Cp , C^ , Cpm and Cpmj; based on multiple samples collected over time for an incontrol process.
8.3.1 Estimations
of CPU and CPL based on multiple
samples
Consider m preliminary multiple samples (subgroups) each of the same size n taken from control chart samples. We shall
The CPU/ CPL Index
155
assume that the sample of the characteristic under consideration, (Xii,Xi2,...,Xin) , follows a normal distribution N(/j,,a2) for i = 1,2,...,TO.The process mean is estimated by the grand average X and the process variance by the pooled sample variance S% from subsamples collected over time for an in-control process. Let x /n and Si 2 n 1 / 2 be the
* = EU v
= E"=i(*« - ^) /( -!)]
i-th sample mean and the sample standard deviation respectively. Then X = Yl™-\ %i I m and S2P = Y17=i ^ Im w i l 1 b e u n b i a s e d estimators of fi and a respectively. Pearn and Lin (2003) suggest an unbiased estimator of CPU and CPL of the following form: f,*
_ bm(n-i)(USL - X) ~ 3^ '
Cpu
^* CpL
~
_ bm(n-i){X - LSL) 35; '
where as above bg = (2 / gfl2T{g / 2)/T[(g - l ) / 2 ] is the correction factor already encountered earlier. Under the normally assumption, [3^/mn /fem(„_i)] x Cpu can be expressed as Zv /y/W Jm{n - 1) , where Zv = Z + 3JrrmCPU ~ N(6u,l) , Z ~ iV(0,l) , we have W = m{n - l)S2 /a2 ~ x ^ ^ , and Zv and W a r e independent. The estimator (3Vmn /& m ( n -i)) x Cpu is distributed as the non-central t variable with m(n — 1) degrees of freedom and the non-centrality parameter 8u — 3-JmnCPU . The estimator (3v?nn / bm(„_i)) x CPL has the same sampling distribution as {3-Jmn /6 m ( n -i)) x CPu but with the non-centrality parameter 8L = 3-Jm,nCPi . Furthermore, since CPU_a,nd CPL based solely on the complete and sufficient statistics (X,Sf,), one concludes that the estimators CPU and CPL are the UMVUEs of CPU and CPi respectively based on multiple samples. Thus, the c.d.f. of X = Cj (which denotes either Cpy or Cpi) can be derived as:
156
Encyclopedia and Handbook of Process Capability Indices
F(x) =
2^~2)l2T(g/2) poo
3Jg+mxy/bgJg
-JL
exp
yf2lrJo n Jo
-i(,-*j»
dudy.
We present a derivation in the Appendix (see Pearn and Wu (2003)). Differentiation with respect to x results in the p.d.f. of Cj given as:
m_
3y/(g + m)/g-2- (g+m)/2 bg^WT(g/2)
f °>-1)/2exPl *J 0
y+
3xy/(g + m)y bQJ~g
where as above bg = (2/g) 1/2 T{g/2)/F[(g and 6 = 3~JmnCj .
dy,
- l ) / 2 ] , 5 = m(n - 1)
From the plots of the p.d.f. of Cj for various sample sizes, we observe that (i) for fixed sample sizes m and n , the variance of Cj increases as Cj increases, (ii) for a fixed n and Cj , the variance of Cj decreases as m increases, and also (iii) for fixed m and Cj, the variance of Cj decreases as n increases. For cases with unequal sample sizes, the unbiased estimators of Cpu and Gpi can straightforwardly be generalized to be:
^PU —
T?
3S„
' ^PL —
d&r,
where X = '^2-_,iT'iXi /N is t h e grand mean of t h e overall sample, N = ]>J--i n » is the number of observations in the total sample and S% = X ^ w ( n » — ^)Sf /{N — m) is the pooled sample variance. From the representations
157
The Cpuj CPL Index
[2>^W^i/bN_m] x CpV = and where
Zu/ylW/iN-m),
[ 3 V F ^ T / bN_m] x C*PL = ZL / JW/(N - m) , Zv
and Zj, are as above Zv / yJW /(N — m) and are distributed as tv(6v) ( tv(6L) ), namely according to the non-central t distribution with v = N — m degrees of freedom and the non-centrality parameters 8u = 3VlV Cpu ( 6L = SJNCpi), respectively. We remind our readers that the process ought be stable in order to produce a reliable estimate of the process capability. If the process is being out of control in the early stages of process capability analysis, it would then be unreliable for estimating the process capability.
8.3.2 Testing
CPU and CPL based on multiple
samples
To test whether a given process meets the preset capability requirements and runs according to the desired quality condition, we can examine the following hypothesis: H0 :d
where g = N - m , bg = (2/g)ll2T{g/2)/T[(g - l)/2] , and tg^a{8) denotes the upper a -th percentile of a non-central t
158
Encyclopedia and Handbook of Process Capability Indices
variable with N — m degrees of freedom and the non-centrality parameter 6 = 3y/~NC[ . The power function may be expressed in the standard manner: (3{Cr,
=PT{C*J
>
C0\Cj}
Pv{tg(6)>3^NC0/bg}.
The size of the test is given by a(C0)=
sup P{Cj,
CreH0
Moreover, Pearn and Wu (2003) have shown that the test statistic
C*x > ^j^tN_m^a{8
= 3VJVC/).
otherwise.
The derivations of the lower confidence bounds, the p-values and the capability calculations for situations with multiple samples can be performed using the same techniques as in the cases with a single sample. However these calculations may be more involved. (See Pearn and Lin (2003), Pearn and Wu (2003) and Pearn et al. (2003) for details).
8.4 Estimating and testing CPU/ CPL: Bayesian approach with single sample 8.4.1 Bayesian approach to the assessment CPL based on single sample
of CPU and
A Bayesian procedure for assessing process capability was proposed in the early stages of process capability studies by Cheng
The CPU/ CPL Index
159
and Spiring (1989) for the initial primitive basic index Cp under the assumption that the process mean \i is equal to the target value T . However, the restriction that fi = T may not be a practical assumption for numerous industrial applications. (Bayesian approach has not been emphasized in the nineties of the 20-th century the formation and consolidation period of the PCI "technology".). In 1999 Shiau et al. suggested a Bayesian procedure for the general situation without restrictions on the process mean applied to the index Cpm. More recently, Pearn and Wu (2003) utilized a similar Bayesian approach for the capability index Cpk . For the one-sided specification limit, Pearn and Wu (2003) have derive a Bayesian interval estimate for Cpu and Cpi and proposed a Bayesian procedure for testing the process capability (relaxing the restrictions). We shall assume that the observations x = {xi,x2,- •-,£„} are random samples taken from independent and identically distributed (i.i.d.) normal variables with the mean /j, and variance a2 N(^,a2). Under these assumptions, given a pre-specified capability level w > 0, the posterior probability p based on the index Cj that a process is capable can be expressed as: p = Pr{the process is capable| x}=Pr{C/ > w | x}
= /.
o
1 T(a)ya+1 exp
x $ 3Vn
' Cj %
K~\
r V,
dy V (n - l)y— w
(8.1)
where a — (n — l ) / 2 and, as above, $(•) is the c.d.f. of a standard normal variable. Note that the posterior probability p depends on n , w and Cj . When n and w are given, and since Cj can be calculated from the process data, there is a one-to-one correspondence between p and Cj. Therefore, it follows that the minimal value of Cj = C (p) , necessary to ensure that the
160
Encyclopedia and Handbook of Process Capability Indices
posterior probability p reaches a certain desirable level, can be useful for assessing process capability. In other words, to see if a process is capable (with capability level w and confidence level p), we only need to check whether Cj > C*(p). Pearn and Wu (2003) have provided tables of the minimal values of Cj , C*(p), for various confidence level p = 0.99, 0.95, 0.90, and commonly used capability requirements C / = 1.25, 1.45, 1.60.
8.4.2 A Bayesian approach to assessment CPL based on multiple samples
with CPU and
Posterior Probability for Multiple Control Chart Samples For cases where the data are collected as multiple control samples, we have considered m samples each of size nt and have suggested the unbiased estimators of Cpu and CPL , where Xj is the i-th sample mean, and St is the i-th sample standard deviation, and N = Y^Z.\ ni i s the number of items in the total sample. These estimators are: n* <^PU
_ bN-m(USL - X) a n a x,* _ bN_m(X - LSL) — w > ^PL — w >
where as above bg = (2/ g)1/2T(g /2) /T[(g - l)/2] , g = N-m, n X = Yl™- iXi IN is the overall sample mean, and S2P = ^2™- (n> ~ 1 )^' 2 IW ~ m ) *s * n e P°°l e d sample variance and with Sf = y^m_ (xtj —XifKrii —1). However, for multiple control samples, it is necessary to take into account the variations between them. To assess these variations Pearn and Wu (2003) suggest the ratio of total within sample variation ( SSW ) and total sum of squared variation (SST ) defined as:
161
The Cpu/ CPL Index
ssw _ 5 § ( ^ ^ ^ 2
7
ssr
E E (**• - ^)2 i=i j=i
2= 1
: E fo !)4 + E «ite *) i=i i=i
Similarly in these situations, the extension of the posterior probability p for multiple control samples can be expressed as: noo
1
Jo T(a)ya+1 6XP x $
3VF
27
^tf-m '^(N
dy, -m)y — w
(8.2)
where a — ()~]._ n, — 1) / 2 . Compare with (8.1) for a single sample, where the ratio 7 = 1, N — m = n — 1 , a = (n — 1) / 2 . In this case the estimator of Cj, (7/ , is reduced to the estimator Cr discussed before.
Chapter 9
Multi-Process Performance Analysis Charts (MPPAC)
9.1 9.2 9.3 9.4 9.5
Introduction The modified Cpk MPPAC The Cvm MPPAC The Spk MPPAC The Le MPPAC
9.1 Introduction In this short Chapter we shall describe a number of new practical procedures for improving the analysis of process capability based on the indices CP, Cpk, Cpm, Spk and some others. Ever since Shewhart (1944) introduced his pioneering control charts, it has become a common practice for practitioners to use various control charts for monitoring different processes on a routine basis. As an example, when dealing with a variable data, the control chart technique usually employs a chart (such as a X chart) to monitor the process center and a chart (such as an R chart or an S chart) to monitor the process spread. These charts are easy to comprehend, and they effectively communicate the critical process information without necessarily employing words and formulas. Unfortunately, they are applicable only for a single process (one process at a time). T h u s , using t h e m in multiprocess environment could be a clumsy and time-consuming task for supervisors or shop engineers since it may require to analyze each
162
Multi-Process Performance Analysis Chart (MPPAC)
163
individual chart in order to evaluate the overall status of shop process control activities. We have already pointed out that process capability index measures the ability of the process to reproduce products that meet certain specifications. When dealing with the capability of a process, there are mainly two characteristics of importance involved, the process location in relation to its target value and the process spread. The closer the process output is to the target value and the smaller is the process spread, the more capable will the process be. However, the fact that process capability indices combine information about closeness to target and the process spread, expressing the capability of a process by a single number, may in some instances be viewed as one of their major drawbacks. When a process is found to be "non-capable", the operator might be interested in knowing whether the non-capability is caused by the fact that the process output is substantially off target or that the process spread is too large, or is a result of a combination of these two factors. In order to circumvent this defect some current researchers in the last 15 years were suggesting that different graphical methods be used to ascertain the improvement initiatives aimed at generating more capable processes (see e.g. Gabel (1990), Boyles (1991), Tang et al. (1997) and Deleryd and Vannman (1999)). A Multi-process Performance Analysis Chart (MPPAC), proposed by Singhal (1990), which evaluates the performance of a multi-process product with symmetric bilateral specifications, provides priorities among multiple processes for capability improvement and reveals whether reducing the variability or the departure of the process mean should be the main task for improvement. While Cpu and Cp\ present the X-axis and Y-axis, respectively in a MPPAC, Cp is the average of Cpu and Cpi , namely, Cp = (Cpu + Cpi) / 2 . Moreover, the Cpk MPPAC provides an efficient route to process improvement by comparing the locations on the chart of the processes before and after an improvement effort has been carried out. Singhal (1991) also provides a MPPAC with several welldefined capability zones by using the process capability indices Cp
164
Encyclopedia and Handbook of Process Capability Indices
and Cpk for grouping the processes in a multiple process environment, into different performance categories on a single chart (see Figure 9.1). This is very helpful when process performance is measured in terms of capability indices. Different capability zones describe a status of each process which is easy to interpret and assist in grouping the processes into performance categories for the purpose of quality improvement operations. The American giant corporation Motorola Inc. introduced their six sigma (6-cr) program (Mikel (1988)) which is equivalent to a defect rate of 3.4 parts per million (PPM). This program corresponds to a Cp value of 2.0 or more and a Cp^ value of 1.5 or more. All this is done under an implicit normal assumption. A Cp value of 2.0 tells us that the specification width is twice as wide as the distribution of the process (with a defect rate of 0.002 PPM) as long as the process mean is equal to the target value. Under the six sigma program, the process mean is allowed to shift as much as 1.5 sigma and still obtain a Cp^ value of 1.5 or more. (See Chapter 2 for the relation between the PPM defect rate and the values of PCI's)
Figure 9.1. MPPAC with capability and six sigma zones (Reproduced from Singhal (1991))
In practical applications, when a product has several models with required different specifications needs to monitor and control,
Multi-Process Performance Analysis Chart (MPPAC)
165
it may be difficult or very time-consuming to carry out factory control activities. As it was mention above, a MPPAC not only evaluates the performance of a multi-process product with symmetric bilateral specifications but also sets the priorities among multiple processes for capability improvement and indicate whether reducing the variability, or adjusting the departure of the process mean should be the focus of improvement operations. This renders a MPPAC to be an efficient tool for communicating between product designer, manufacturers, quality engineers, and among various management departments.
9.2 The modified Cpk MPPAC Pearn and Chen (1997) proposed a modification to the Cpk MPPAC combining the more-advanced process capability indices, Cpm or Cpmk, to identify the problems causing processes' failings to center around the target value. Under this modification the performance of an entire product, composed of the-nominal-thebest, the-larger-the-better, and the-smaller-the-better characteristics, could be measured. The modified Cpk MPPAC is shown in Figure 9.2, with five capability zones corresponding to the five standard process conditions for Cpk = 1.00, 1.33, 1.67, and 2.00. We mark these process capability zones with bold lines as the critical values of individual process capabilities for the multi-process product on MPPAC chart, and the corresponding fractions of nonconformities (in PPM) are summarized in Table 9.1 (see also Montgomery (2001)) among other sources. In this modified MPPAC, Cpu and Cp\ represent the X-axis and Y-axis, respectively. Recall that Cp is the average of Cpu and Cpi > [ CP = (C pu + Cpi) / 2 ] and Cpk is the minimum value of the X- and Y-axes [ Cpk = min{C pu ,C p ;} ]. Hence, based on Cpk MPPAC, the vertical and horizontal axes of the chart are geared to evaluate the-larger-the-better and the-smaller-the-better characteristics.
166
Encyclopedia and Handbook of Process Capability Indices
Combining Singhal's MPPAC with asymmetric process capability index Cpa , and considering unilateral characteristics, Chen et al. (2001) introduced a Process Capability Analysis Chart (PCAC) to evaluate the process potential and performance for an entire product composed of the-smaller-the-better, the-larger-thebetter, symmetric and asymmetric specifications. As mentioned above, the process yield of a multi-process product is lower than of any individual process yield. Similarly, when the entire product capability is preset to satisfy the required level, the individual process capability should exceed the preset standard for the entire product. Therefore, the process capabilities are recognized to be capable if all of the individual process capability indices are located within the process capability zone. Conversely, processes must be upgraded when some of the process capability indices are located outside this zone. It is straightforward to distinguish process performance with respect to the locations of process capability indices on a modified Cpk MPPAC. Hence, the modified Cpk MPPAC not just distinguishes process capabilities, but also reveals the degree of quality accuracy for multi-process products. This makes the modified Cpk MPPAC to be an effective and efficient tool for evaluating multi-process products, composed of numerous unilateral and bilateral specifications. C, =• 0.50
C , = 0.75
--7—
/ /
/
/
/
C, = 0.875
—r
/
. /
/
/
, - • "
/ • / /
/
, - • "
•
/ / ,• / / , • • - ' . , -
, • • • • -
/
.*'••
. • • *
" " " ' , - • • - " " " "
I ///sO*
/M*< 0.50
1.00
1-33
167
i m
C„
Figure 9.2. The modified Cpk MPPAC with capability zones for Cpk = 1.00, 1.33, 1.67, and 2.00.
Multi-Process Performance Analysis Chart (MPPAC)
167
Recall (Chapter 5) that the loss function utilized in the G. Taguchi method has the property that the closer the process mean is to the process target the better quality and fewer process losses are encountered. Conversely, the farther the process mean is located from the process target the worse are process capabilities. Keeping the process on-target is crucial for smooth operation. A few subsidiary lines of Ca = 1 — |/x — M\j d can be added on a MPPAC for controlling accurately the process centering. The index Ca measures what is known to be the relative distance of the shift from process mean to the preset target. (The defmition of relative distance is (/J, — M) / d or (M — fi) / d .) Evidently, equal relative distances result in the same values of Ca . Note that the preset target value coincided with the midpoint of the specifications (i.e. T = M). Off-diagonal subsidiary lines are plotted for values of Ca 0.500, 0.750 and 0.875 in Figure 9.2. Ca < 0.875 indicates that the process is not accurate and some actions to shift the process mean closer to the process target are required. Equivalently, Ca > 0.875 indicates that a process possesses quite a good accuracy. In general, Ca should not be too small, since it would imply that the process mean shifts further away from the process target which results in a substantial process loss. Set r = | [i — M | / d , then the values of [i will be M + r x d and M — r x d for each Ca . The slope of the corresponding subsidiary line is (1 + r ) / ( l — r) when the process mean is greater than the process target, and analogously the slope of the corresponding subsidiary line is (1 — r ) / ( l + r) when the process mean is greater than the process target, and analogously the slope of the corresponding subsidiary line is (1 — r) /(l + r) when the process mean is greater than the process target, and analogously the slope of the corresponding subsidiary line is (1 — r ) / ( l + r) when the process mean is smaller than the process target. Table 9.1 displays
168
Encyclopedia and Handbook of Process Capability Indices
five values of Ca with the corresponding fx, r , and the slopes of the subsidiary lines. Table 9.1. The values of Ca with the corresponding /J,, r , and the slope of lines. Ca 1.000 0.875 0.750 0.500 0.250 0.000
M M M M M M M M
+ + + + -
P M 0.125 0.125 0.250 0.250 0.500 0.500 0.750 0.750 USL LSL
xd xef xd xd xd xd xd xd
r
Slope
0.000 0.125 0.125 0.250 0.250 0.500 0.500 0.750 0.750 1.000 1.000
1.000 1.286 0.778 1.667 0.600 3.000 0.333 7.000 0.143 00
0.000
Concerning the modified Cpk MPPAC, it is useful to observe that: [a]
The parallel and perpendicular lines through the plotted point intersecting the vertical axis (y-axis) and the horizontal axis (x-axis) at the points represented by Cpu = (USL — ^)/(3<x) and Cpi = (/i — LSL)/(3a), respectively.
[b] The 7r/4 subsidiary line (i.e. the line with the slope 1) represents the points where the process mean equal to the target (// = M = T) and the values of the indices Cpu, Cpt and Cpk are equal. [c]
For the points inside the area to the right of the 7r/4 subsidiary line (i.e. with the slope < 1) represents processes with the process mean shifted towards the lower specification limit (i.e. the process mean is lower than the target value), or Cpk = Cpi < Cpu. On the other hand, for the points inside the area to the left of the 7r/4 subsidiary line (i.e. with the slope > 1) represents processes with the process mean shifted
Multi-Process Performance Analysis Chart (MPPAC)
169
towards the upper specification limit (the process mean is higher than the target value), or Cpk = Cpu < Cv\. [d] The origin point represents a process with Cpk = Cpu = Cpi — 0 which means that the standard deviation of the process is infinite or that the process mean lies at USL or LSL. In quality improvement, reduction of variation from the target is of the same importance as increasing the process yield and reduction of the process spread. The modified Cpk MPPAC combines the two concepts - closeness to the target and narrow spread - in more efficient manner than using a process capability index alone. If the cases when exact values of \x and a are known, the modified C^ MPPAC can easily be applied. In other words, if the corresponding value of {Cpu,Cpi) is inside the capability region, the process is then defined to be capable, and if the value is outside, the process is defined as a non-capable one. In practice, however, we never know the true values of Cpu and Cv\ • Pearn and Wu (2004) developed a procedure to construct the lower confidence bounds on the indices Cpk , Cpu , and Cpi for each characteristic type. These lower confidence bounds can simultaneously be plotted on a single chart to check whether the process output is off target or that the process spread is too wide (or if a combination of these two situations takes places).
9.3 The Cpm MPPAC
Since the index c simultaneously measures process variability and centering (see Chapter 5), a Cpm MPPAC provides a convenient way to identify problems in process capability after statistical control has been established. Based on the definition, set Cpm = h , for various h values, then a set of (n,a) values satisfying the equation:
170
Encyclopedia and Handbook of Process Capability Indices 2
a (d/3)
+
f(M-r)l
2
d/3
can be plotted on the contour (indifference curve) of Cpm = h. These contours are semicircles centered at (T, 0) of radius 1/ h . The more capable the process, the smaller is the semicircle. Six contours of the Cpm MPPAC for the six common values for Cpm = 1/3, 1/2, 1, 1.33, 1.67, and 2, are shown in Figure 9.3. Concerning the Cpm MPPAC, it is worth observing that: [a]
The parallel and perpendicular lines through the plotted point intersect the vertical and horizontal axes at the points represented by (a / (d / 3) )2 and ( 0 - T) / (d / 3) f , respectively.
[b] The distance between T and the point at which the perpendicular line through the plotted point intersects the horizontal axis, corresponds to the departure of the process mean from the target. [c]
The distance between 0 and the point at which the parallel line through the plotted point intersects the vertical axis represents the process variance.
[d] For the points inside of the semicircle of the contour (the indifference curve) Cpm = h , the corresponding Cpm values are larger than h . For the points outside of the semicircle of the contour Cpm — h the corresponding Cpm values are smaller than h. [e]
As the point gets closer to the target values the value of the Cpm becomes larger, and the process performance improves.
[f]
For processes with a fixed value of Cpm , the points within the two 7r/4 lines envelop, the process variability is due mainly to the process variance.
Multi-Process Performance Analysis Chart (MPPAC)
171
[g] For processes with a fixed value of Cpm , the points outside the two 7r/4 lines envelop, the process variability is due mainly to the process departure. (o/(d/3))2
9
Target Line
4
1
0.SS5 (U59 0.25
0,359 0.MS
(c/(d/3)f
1
4
Figure 9.3. The contours of a Cpm MPPAC.
The Cpm MPPAC conveys critical information about a multiple processes concerning the process departure and the process variability using a single chart. Current research about MPPAC is restricted to obtaining quality information from single sample of each process ignoring the sampling errors. The information provided from the existing MPPAC chart can therefore be unreliable and misleading resulting in incorrect decisions. Pearn et al. (2004) investigate the accuracy of estimation as a function of the process characteristic parameter £ = (/x — T) / a , given a group of multiple control chart samples. Data regarding the true capability values and the maximum of nonconformities (in PPM) are provided for production quality control. Suitable sample sizes are then recommended to the proposed Cpm MPPAC for multiple processes production quality control. This approach ensures that the critical information obtained from the Cpm MPPAC based on multiple control chart samples, is more reliable.
172
Encyclopedia and Handbook of Process Capability Indices
9.4 The Spk MPPAC Based on the yield index Spk (see Chapter 7), Boyles (1994) developed the so-called Spk contour plot that is a contour plot of the index Spk as a function of the process parameters (//, a), to be used for monitoring and controlling process performance. In fact, the Spk contour plot is a handy tool for evaluating multiple processes since it allows to obtain the process yield and the process departure ratio by checking the location of the index value appearing on the contour plot. For multiple processes with the same specification limits, the contour plot not only shows the process capability for multiple processes simultaneously, but also provides a quick reference about the parameters that should be targeted for process improvements. We note, however, that the Spk contour plot is applicable only for multiple processes with same specification limits on each single process which may not be appropriate for processes with multiple characteristics when the characteristic specifications are not equal. To enable the applicability of the contour plot for processes with multiple characteristics, Chen et al. (2003) applied the method developed by Deleryd and Vannman (1999) who introduced a process capability plot, called the (6,7) -plot, which is an adjustment of Boyles' (/i, <j)-plot where 6 — (/J, — T)/d , and 7 = a I d . The authors readjusted the definition of Spk in the following manner: bpk —
+ •.
M
3
12
I*"1 [5*
\
a 1 — Qr
,
Cdp
+
.
> .
LSL
^
[1 + C«*r]l ,
C
dp
J.
where Cdr = (// — T) / d , Cdp — a j d and $ is the standard normal c.d.f. Note that Cdr measures the departure ratio and Cdp measures the variation relative to the specification tolerance.
Multi-Process Performance Analysis Chart (MPPAC)
173
If /j, < T (/x > T), we have Cdr < 0 (Cdr > 0 ) , and when/i = T , Cdr = 0 (process is on target). Evidently, if zx = £SX , Cdr = — 1; and for /x = £/5X , C^ = 1. The new expression of Spk is for the case where the specification interval is two-sided with the target value T at M , a most common practical situation. Using Cdr as the x-axis and Cdp as the y-axis, Chen et al. (2003) have plotted the following point set forming the curve of Spk (bold-shape curve in Figure 9.4) on the (Cdr, Cdp) coordinates:
{cdr,cdp)\^-1
1 — Cdr
!« ,
C
dv
+
,
i + cdr
> . ,
C
dp
bpk ,
We note that the process capability plot is invariant to the value of the specification limits. Processes with multiple characteristics having different characteristic specification limits can thus be plotted simultaneously on a single chart. Chen et al. (2003) call this control chart a Multi-Characteristic Process Capability Analysis (MCPCA) chart. In Figure 9.4, the top curve is the capability plot for processes with Spk = 1.00, and the bottom one is the plot for processes with S^ = 1.33. As we have mentioned above (Chapter 7), Spk is a yieldbased index. According to the modern quality improvement theories, reduction of the variation from the target is as important a step increasing the process yield (i.e. meeting the specifications). Therefore, three pairs of control limits I1( I2, and I3 are drawn on the MCPCA chart (in Figure 9.4) to monitor the variation from the target of each one of the characteristics. Corresponding to the three pairs of control limits I1( I2, and I3, the six values of Cdr =(n-T)/d are -1.0, -0.5, -0.25, 0.25, 0.5, and 1.0 from the left to the right, respectively.
174
Encyclopedia and Handbook of Process Capability Indices
0.35-
0.30.25-
0.20.150.10.05-
0-1
-0.5
-0.25
0
_ 0.25
0.5
1
dr
Figure 9.4. A Sp^ MCPCA chart with several departure control zones.
Table 9.2. Various control regions for a process departure and improvement assessments. Control Region
C^ Value
Process Improvement Assessment
I
±0.25
characteristic departure is tolerable
I
±0.50
characteristic investigation desirable
I
± 1.00
characteristic departure is serious, the overall process should be reviewed.
departure is abnormal, and improvements are
In terms of the six-sigma quality improvement program (as formulated by Motorola, see, e.g., Noguera and Nielsen (1992)) setting d = 6cr , the three pairs of the control limits I t , I2, and I3 correspond to \ y, — T \= 1.5a, 3cr and 6
Multi-Process Performance Analysis Chart (MPPAC)
175
PPM of non-conformities, assuming that the specification limits are 6cr away from the target. Under the assumption of normality and stationarity of the process (see Chapter 2), the three pairs of control limits Il5 I2, and I3 provide various process accuracies (the degree of centering) of the control zones. Various control regions (zones) for process departure and the improvement assessments are summarized in Table 9.2. Practitioners can judge the degree of centering of the characteristic j by checking the location of the corresponding plotted point on a MCPCA chart. For example, the departure measure C^r value is equal to -0.5 for the characteristic A in Figure 9.4. This departure ratio is considered to be significant which calls for an immediate check to determine the abnormal changes of the parameter settings. Using a MCPCA chart, the practitioners can thus effectively control and monitor both the variation and the departure from its target for each single process characteristic.
0.350.30.250.20.150.10.050". -1
, -0.5
-0.25
, 0
. 0.25
. 0.5
. ~dr 1 dr
Figure 9.5. A MCPCA chart with contours of Spk = 1.00, 1.33, 1.50, 1.67, 2.00 (from top to bottom).
In a MCPCA chart (see Figure 9.5), we use standardized measures of process departure and process variation. Consequently, a MCPCA chart can be used for processes with multiple
176
Encyclopedia and Handbook of Process Capability Indices
characteristics where the individual characteristic specifications are not identical. The MCPCA chart displays all characteristic index values on one chart, and provides the characteristic yield based on the Spk contours. It can present instantly the information about the locations and spreads of all the characteristics under study by their (Cjr, Cdp) - coordinate values. The MCPCA chart also displays the relative magnitudes of the process variation and process departure in terms of the standardized measures (Crfr, C
9.5 The Le MPPAC As we have already discussed in Chapter 7 it were Hsiang and Taguchi (1985) who first introduced the loss function approach to quality improvement which focuses on reduction of variation around the target value. Johnson (1992) introduced the relative expected loss Le for symmetric cases, to provide numerical measures of process performance for industrial applications. Tsui (1997) interpret Le = Lpe + Lot , which results in an uncontaminated separation between the information concerning the process relative inconsistency loss ( Lpe ) and the process relative off-target loss (Lot). The index Le is defined as a ratio of the expected quadratic loss and the square of the half specification width as follows:
=/:
(x - Tf d2
dF(X) = r-\ + dl
ti-T \ d
2
Multi-Process Performance Analysis Chart (MPPAC)
177
where F(-) is the c.d.f. of the measured characteristic. The first term (cr / d)2 is denoted Lpe , and the second [(/i — T) I df is denoted Lot , and Le can thus be rewritten as Le = Lpe + Lot. The mathematical relationships: Le = (3C p m )" 2 , Lot = (1 - Ca)2 , and Lpe = (3C p )~ 2 can easily be established as indicated in Chapter 5. The advantage of using an equivalent index Le over Cpm is that an estimator of the former has superior statistical properties than that of the latter, since the former does not involve a reciprocal transformation of the process mean and variance. (See Chapter 4, Section 2 for a definition of Cpm.) The index Le also provides an uncontaminated separation between information concerning the process precision and the process accuracy. This separation indicates a direction practitioners may follow in modifying the process parameters to improve the process quality. Some commonly encountered values of Le are: 1.00 (a process is incapable), 0.44 (a process is incapable), 0.11 (a process is usually referred to as capable), 0.06 (process is called satisfactory), 0.05 (process is normally called good), 0.04 (process is normally called excellent), and 0.03 (process is designated as super). The corresponding Cpm values are also listed in Table 9.3. Table 9.3. Some commonly used Le values and the equivalent Cpm . Condition
Le
Cpm
Incapable
1.00
0.33
Incapable
0.44
0.50
Capable
0.11
1.00
Satisfactory
0.06
1.33
Good
0.05
1.50
Excellent
0.04
1.67
Super
0.03
2.00
The sub-index Lot measures the relative process departure, which as mentioned above, has been referred to as the relative offtarget loss index. The sub-index Lpe, as it was already pointed out,
178
Encyclopedia and Handbook of Process Capability Indices
measures process variation relative to the specification tolerance, and has been referred to as the relative inconsistency loss index. Some commonly used values of Lpe are: 0.11, 0.06, 0.05, 0.04 and 0.03. The corresponding quality conditions are indicated in Table 9.4. Evidently, theses 5 values of Lpe are equivalent to the Cp values of 1.00, 1.33, 1.50, 1.67, and 2.00, respectively, thus covering a wide range of the precision requirements needed for numerous real-world applications. Table 9.4. Some commonly used precision requirements. Quality Condition
Precision Requirement
Incapable
Lpi = 1 00
0.11 <
Lpe
Capable
0.06 < Lpe
< 0.11
Satisfactory
0.05 < Lpe
< 0.06
Good
0.04 < Lpe
< 0.05
Excellent
0.03 < Lpe
< 0.04
Super
Lpe
Target
LSL
U = 1.00 Ui = 0AA i-o,= 0.11
J
< 0.03
USL
tof=t).11
Ut = 0.AA ^ = 1.00
Figure 9.6. The Le MPPAC.
Pearn et al. (2003) using the expected loss index Le introduced a novel control chart, called the Le MPPAC. Based on the definition, Le = (// - T) 2 / d2 + a2 / d2 , we set Le = k for various k values, and then determine a set of (/i, a) values satisfying the equation: (/U - T)2 + a2 = kd2 which can be
Multi-Process Performance Analysis Chart (MPPAC)
179
plotted on the contour (a curve) of Le = k . These contours are semicircles centered at fi = T with the radius yfkd . Analogously to the situation encountered in the case of the Cpm index, the more capable is the process, the smaller is the semicircle. The seven contours on the Le MPPAC for the seven Le values listed in Table 9.3 are shown in Figure 9.6. Note a certain clustering of the semicircles for smaller values of k . A Le MPPAC has the following properties worth of mentioning: (1)
As a point gets closer to the target, the value of Le becomes smaller, and the process performance improves.
(2)
For the points inside the semicircle of the contour Le — k, the corresponding Le values are smaller than k. Analogously, for the points outside the semicircle of contour the Le = k, the corresponding Le are larger than k.
(3)
For processes with fixed values of Le, for the points within the two 7r/4 lines envelop, the variability is due mainly to the process variance.
(4)
For processes with fixed values of Le, for the points outside the two 7r/4 lines envelop, the variability is due mainly to the process departure.
(5)
The perpendicular and parallel lines passing through a plotted point intersect the horizontal and vertical axes at the points that represent its Lot and Lpe, respectively.
(6)
The distance between T and the point, at which the perpendicular line through the plotted point intersects the horizontal axis, denotes the departure of the process mean from the target.
(7)
The distance between T and the point, at which the parallel line through the plotted point intersects the vertical axis, denotes the departure of the process variance.
180
Encyclopedia and Handbook of Process Capability Indices
In addition, Pearn et al. (2003) obtained the exact upper confidence bound for the Le to provide a measure of the maximal expected loss of a process based on the sample data. The proposed decision - making procedure ensures that the risk of making a wrong decision would not exceed the preset Type I error. The Le MPPAC displays multiple processes with the mean departure and the process variability relative to the specification tolerances on a single chart, which is very appropriate for manufacturing capability control of a group of processes in a multiple process environment. Also the Le MPPAC prioritizes the order in which a further capability improvement effort should be accomplished: either to move the process mean closer to the target value or to reduce the process variation.
Chapter 10
PCIs with Asymmetric Specification Limits
10.1 10.2 10.3 10.4 10.5 10.6
Introduction The Cpk index for asymmetric tolerances The Cpm index for asymmetric tolerances The Cpn index for asymmetric tolerances The Crmk index for asymmetric tolerances The Loss index for asymmetric tolerances
10.1 Introduction
The concert of symmetric tolerance was used tacitly in this beak on numerous occasions. A process is said to have a symmetric tolerance if the target value T is set to be the mid-point of the specification interval [LSL, USL], i.e. T = M = (USL + LSL) / 2 . Most research in quality assurance literature has been focused so far on cases in which the manufacturing tolerance is symmetric. Examples include the earlier paper by Kane (1986), Chan et al. (1988), Choi and Owen (1990), Boyles (1991), Pearn et al. (1992), V a n n m a n (1995), V a n n m a n and Kotz (1995), Spiring (1997), P e a r n et al. (1998), P e a r n et al. (2003a, 2003b) among many others. Granted t h a t cases with symmetric tolerances are quite common in practical situations, cases with asymmetric tolerances ( T vt M) (see Figure 10.1) also often occur in the manufacturing industry. F r o m the customer's point of view, asymmetric tolerances reflect t h a t deviations from the target t h a t are less
181
182
Encyclopedia and Handbook of Process Capability Indices
tolerable in one direction than the other (see Boyles (1994), Vannman (1997), and Wu and Tang (1998)). Usually they are not directly related to the shape of the supplier's process distribution. Nevertheless, asymmetric tolerances can also arise in those situations where the tolerances are symmetric to start with, but the process distribution is skewed or follows a non-normal distribution. To deal with this the data are usually transformed to achieve approximate normality. A prominent example of this approach is Chou et al. (1998) who have used the well known, time - honored Johnson (1949) curves to transform the non-normal process data. Unfortunately, there has been so far comparatively few research publications dealing with asymmetric tolerances. (Exceptions are Choi and Owen (1990), Boyles (1994), Vannman (1997), Chen (1998), Pearn et al. (1998, 1999), Chen et al. (1999), and the more recent Jessenberger and Weihs (2000)).
OS
0*
Loss 04
0]
-n
t
35
Figure 10.1 Assymmetric tolerance.
10.2 The Cpk index for asymmetric tolerances To account for the asymmetry, Kane (1986) seems to be the first to modify Cp , Cpk , Cpu and Cpt indices for the situations
PCIs with Asymmetric Specification
Limits
183
where USL — T ^ T — LSL . These modifications are denoted by Cp , C*pk , Cpu and Gv\, and define to be Cp = min
USL-T 3a
Cpk -=
T - LSL ' 3a
min[C*pu,C*pi } ,
(10.1) (10.2)
where
r —
USL-T 3a
1-
I V-T\ USL-T
)
> Cpi
—
T
-LSL 3a
\H-T\ T -LSL,
respectively. Observe that the index Cp represents the relative size of the smaller (one-sided) semi-tolerance, rather than the relative size of the entire specification range involved in Cp . The semi-tolerances are, by definition, the distances between the target and the specification limits. In these cases, the potential of a process is interpreted in terms of the ability of the distribution's half-width to fit within the (smaller) semi-tolerance. These indices attain their maximum at [i = T . For example, the upper limits for Cpt a n d o pU are (T — LSL) / 3a and (USL — T) / 3a , respectively, while the original indices Cpu and Cpi do not possess upper bounds. The values (T - LSL) / 3a and (USL - T) / 3a are the arguments of the min function in Cp (equation (10.1)). As the mean deviates from the target, the value of Cp decreases linearly. When the mean shifts by a distance equal to or exceeding the corresponding semi-tolerance, the index value reduces to zero. Consequently C*pt = 0 if | ji - T \ > T - LSL , while C*pu = 0 if | /i — T | > USL — T . These properties can be used for an onesided specification limit with a target value. The index C ^ is also defined to be zero for | /x — T | > T — LSL or | \i - T \ > USL — T and can be represented as:
184
Encyclopedia and Handbook of Process Capability Indices
Cpk — Cp (1 — k J ,
where k* =\fi-T |/min{T - LSL,USL - T} . Therefore C*pk has a maximum value equal to Cp , which is attained when fj, = T corresponding to k = 0 . Also, C*pk will have zero value whenever the mean is shifted away from the target by an amount equal to or exceeding the smaller semi-tolerance (i.e. Cpk = 0 for k > 1 ). Recall that the index Cpk vanishes only when the mean is located at or outside the specification limits. The initial generalization developed for processes with asymmetric tolerances simply shifts one of the two specification limits, so that the new (shifted) specification limits become symmetric with respect to the target value (Kane (1986); Chan et al. (1988)). In other words, this generalization replaces the original specification limits (T — DhT + Du) by the new symmetric limits (on occasion admittedly unjustified) T ± d , where d = min{D u ,D,} , Du =USL - T , Dl =T - LSL , in order to apply the standard definition of Cpk . The generalization may be rewritten as:
c;k = d*~l£a~Tl.
(io.3)
Evidently, if Du = Dt , the specification tolerance becomes symmetric and the generalization as expressed by equation (10.3) reduces to the original Cpk defined in (3.1). Boyles (1994) notes that this generalization can underestimate the process capability by restricting the process to a proper subset of the actual specification range. Indeed consider a process with n = T — d / 2 and
PCIs with Asymmetric Specification
185
Limits
l_i = T - d/2 = 30, and a = d/3 = 2 0 / 3 is depicted in Figure 10.2.
Figure 10.2. An example for USL = 50, LSL = 10 , T = 40 , fi = 30 , ]and cr = 2 0 / 3 .
An alternative generalization suggested for processes with asymmetric tolerances shifts both specification limits to arrive at one that is symmetric (Franklin and Wasserman (1992), Kushler and Hurley (1992)). This modification replaces the original specification limits (T — DhT + Du) with the new symmetric limits T ± (D[ + Du)/2 (which, as in the previous situation, may sometimes be unjustified), and next applies the standard definition of Cpk . For this generalization, the index Cpfc (defined in equation (3.1)) can be rewritten as:
CL pk
d'-\n-T 3a
(10.4)
where d' = (Dt + Du) / 2. Evidently if Du = Dt , the specification tolerance becomes symmetric and C'pk defined in equation (10.4) reduces to Cp£ . Boyles (1994) also observes that this approach can
186
Encyclopedia and Handbook of Process Capability Indices
either understate or overstate the process capability, depending on the position of /x relative to T . Indeed, consider two processes A and B with fiA = T - d , aA = d/6 , fiB = T - 3d / 4 , aB=d/U and T = (3USL + LSL)/4 . For the process A we have A Cpk = 0 . While the expected proportion nonconforming is approximately 0.135%. Consequently, in this case C'pk understates the capability of the process A. In contrast, it is straightforward to be verity that for the process B, BC'pk = 1. Here however, the expected proportion nonconforming is approximately 99.865% rather than (See Fig 10.3). Obviously g C ^ overstates the process capability. To remedy the situation, Pearn and Chen (1998) proposed the index Cpk , - another generalization of Cpk - for processes with asymmetric tolerances. The motivation for the new index Cpk is based on the general criteria stipulated by Boyles (1994), Choi and Owen (1990) and Pearn et al. (1992) when analyzing and comparing the existing capability indices dealing with (a) process yield; (b) process centering; (c) other process characteristics. The generalization Cpk (the Pearn-Chen index) is defined formally as d* — A* L
c;k = -^-,
(io-5)
where A* = m a x ^ V - T)/Du, d*(T - n)/Dt} , DU=USL-T , Dt =T - LSL, d* = min{A,,I>,} . Observe that d*(/z - T)/Du = \mm{Du,Dl}(n-T)]/(USL-T). Obviously, if T = M (a symmetric tolerance), then d = Du = Dt = d , A =\ fj, — M \ and Cpk reduces to the original index Cpfc . We note that the index Cpk attains the maximal values at fj, = T , regardless of whether the preset specification tolerances are symmetric or not. Table 10.1 provides numerical values of Cpk , C'pk and Cpk for several selected values of the "parameters": T, USL, LSL, ji and
PCIs with Asymmetric Specification
187
Limits
a. Pearn and Chen (1998) provide a thorough comparison among the three indices, Cpk , C'pk and Cpk .
Table 10.1. Numerical values of Cpk , C'pk and Cpk for various values of ix and fixed a = 1 0 / 3 , with {USL, T, LSL) = (10, 40, 50). V 10 13 16 19 22 25 28 31 34 37 41 42 43 44 45 46 47 48 49 50
Cpk 0.000 0.300 0.600 0.900 1.200 1.500 1.800 1.900 1.600 1.300 0.900 0.800 0.700 0.600 0.500 0.400 0.300 0.200 0.100 0.000
Cpk
Cpk
0.000 0.000 0.000 0.000 0.200 0.500 0.800 1.100 1.400 1.700 1.900 1.800 1.700 1.600 1.500 1.400 1.300 1.200 1.100 1.000
0.000 0.100 0.200 0.300 0.400 0.500 0.600 0.700 0.800 0.900 0.900 0.800 0.700 0.600 0.500 0.400 0.300 0.200 0.100 0.000
For processes with asymmetric tolerances, the corresponding loss function is also asymmetric with respect to T , which can be defined as
Loss(x) —
[(T - x)/(T - LSL)]2,
LSL
[(x - T)/(USL -T)f,
T < x < USL,
1
<x
otherwise.
188
Encyclopedia and Handbook of Process Capability Indices
Figure 10.3. An asymmetric loss function corresponding to the asymmetric tolerance (LSL, T, USL) = (10, 40, 50).
Figure 10.4. processes with
values for 10 < fj, < 50 and a = 2 0 / 3 (LSL, T, USL) =
Figure 10.3 displays a typical loss function for processes with an asymmetric tolerance (LSL, T, USL) = (10,40,50) . Figure 10.4 displays the plots of C'^ for processes with 10 < /J, < 50 , CT = 1 0 / 3 (top) and o = 2 0 / 3 (bottom), where (LSL,T,USL) = (10,40,50) is tilted towards USL. It is evident that index C'^ takes into account the asymmetry of the loss function. Thus, given two processes E and F with fiE > T and fiF < T , satisfying (/j,E -T)/Du = (T - fiF) /Di (i.e., processes E and F have equal "departure ratios"), C'^ values given to the processes E and F are the same if aE = aF . For example, consider the processes E and F with \xE = 45 > T and \iF = 25 < T. Clearly, the corresponding departure ratios are 1/2 for both processes: ((45 - 40)/10 = (40 - 25)/30 = 1/2). We thus have Cpk = 0.50 for both processes with aE = aF = 1 0 / 3 , and C'^ = 0.25 for both processes with aE = oF = 2 0 / 3 (see Figure 10.4.). In addition, the index C"k decreases as the mean fi shifts away from target T in either direction. In fact C ^ decreases faster when ytx moves away from T to the closer specification limit than to the further specification limit. This is an advantage since the index would respond faster to the shift towards "the wrong side" of T
PCIs with Asymmetric Specification
Limits
189
as compared with the shifts towards the middle of the specification interval. To estimate the index Cpk , Pearn and Chen (1998) propose the natural estimator: -A* 35
r" — where
A*
=max{d*(x-T)/Du,d*(T-x)/Dl},
(Compare with A , in Equation (10.5)). As usual, x=^2. Xi /n and s = [^._, (XJ — x)2/(n—l)]1'2 are the sample mean and the sample standard deviation, respectively. These can be obtained for a process that is "demonstrably" stable (well in control). In the case when the production tolerance is symmetric A reduces to | x — T | and the estimator Cpk becomes: Cpk = mm{(USL -x)/
3s, (x - LSL) / 3s},
i.e. the natural estimator of Cpk discussed in Kotz et al. (1993). Therefore, we may view the estimator C'v\ as an extension of Cpk . Assuming that the process is normally distributed, the estimator Cpk can be rewritten as
Cnk
W 2 # - i / 2 c[_ (n - i 1) a
,1/2
max
K = (n — l)s 2 / o2 is distributed according to the distribution, and Z = nl'2(x — T) / a is distributed as XnN(6,l) with 6 = nl/2(/j,-T)/a . Furthermore, since x and s are mutually independent (in the normal case), so are the Z and K. To derive the expected value and the variance of Cpk , we first need to evaluate:
where
190
Encyclopedia and Handbook of Process Capability Indices
-Z A
E
-
I
—
—
1 2
~ Du + A ; (27T) /
and E max*
U
ID.' A J)
+
A
exp
' 2
f <5
/ <5 1
vA.
A <S2^
2
Dl
A2
$(- |6
c2 ^
+
A2
A2;
(27T) 1 / 2
exp
+2
U
[1 - 2 * ( - | 8 |;
2
{^}) (!-»(-!*
max<
The rth moment (about zero) of C'v\ can then be expressed as
E(c;ky = (n x£
1)r/2
max
£(^ r / 2 )E Du ' A J,
r-j
cr
n
1/2
(10.6)
Note that C'v\ is a biased estimator of C ^ . Some numerical values of E(Cpk) and Var{Cpk) are presented in Table 3-5 of Pearn and Chen (1998). A few representative values are given below. The resultant bias is negative for all the cases in the Pearn and Chen (1998) investigation as long as /J, ^ T . For \x = T , the bias is positive for n — 10 but becomes negative for larger values of n . For d j a = 2.0 it is negative for all n > 20 ; for d Io = 3.0 it is negative for n > 30 ; for d /a = 4.0 , it is negative for n > 40 and for d / a = 5.0 , it is negative starting with n > 60 . As n becomes very large, the bias becomes 0. Note that when the tolerance is symmetric, the results are the same as those provided by Kotz et al. (1993). (c.f. Chapter 3)
PCIs with Asymmetric Specification
191
Limits
We shall now denote B = nl,2{d* / a) and Y = 2 [m&x{(d / DU)Z,— (d / Di)Z}] . Using these notations C'^ can be expressed as
C';k = [ V ^ I ( 5 -
<JY)\/{MnK).
Let, as above, <&(•) and (/)(•) be the c.d.f. and the p.d.f. of the standard normal distribution N(0,1), respectively. The c.d.f. and the p.d.f. of the variable Z distributed as N(S, 1) can be expressed as: Fz(z) = $(z — 6) and cj){z — 6) , respectively. Hence, the c.d.f. of Y is given by: FY{y) = $[(£>„ /d*)4y -6}-
$[-(Dt /d*)Jy
- 6} ,
(10.7)
and the p.d.f. is fr(y) = —^-(Du4>[{Du 2d V^ + Dt4[(Dt/d*)Jy
Id*)Jy
- 6} (10.8)
+ 6]).
Therefore, the c.d.f. of the estimatior C'^ can be expressed as follows (Pearn et al. (2003)).
f FK (L(x,y)) fY(y)dy, FA„
(£) =
l-Fy(B2),
x < 0, x = 0,
JFK(L(x,y))fY{y)dy,
(Pearn et al. (2003)), and the p.d.f. of C'^ becomes
x>0
(10.9)
192
Encyclopedia and Handbook of Process Capability Indices
/ / * ( £ ( * , BH))fy(BH)
k;Sx) =
'dt,
)
X<0,
,
JfK(L(x,
BH))fY{BH) —
(10.10)
-dt, x>0, x
. o
where, L(x, y) = (n - 1){B - Jyf /(9nx2) , i ^ ( - ) and fK(-) are c.d.f. and p.d.f. of K, respectively, FY(-) is the c.d.f. of Y given by (10.7), and /y(-) is the corresponding p.d.f. (10.8). In the cases symmetric manufacturing tolerance (T = M ),DU = Dt = d = d*, B = n 1//2 (d / a ) = D ; consequently the c.d.f. of Cpk given by (10.9) is reduced to (3.1) A decision making rule for testing Cpk Using the index C'v\ , the engineers can access to the process performance and monitor the manufacturing processes on a routine basis. To obtain a decision making rule one can set a testing hypothesis with the null hypothesis Cpk < C for a given constant value C (the process is incapable) and the alternative C'v'k > C (the process is capable). The null hypothesis will be rejected if Cpk > cQ J where the constant ca (called the critical value) is determined so that the significance level of the test is a , i.e., P{Cpk > ca | Cpk = C) = a . The decision making rule will then be: for given values of the risk a and the sample size n , the process will be declared capable if Cpk > ca and incapable otherwise. By using the familiar designation £ = (// — T) / er and b = d I a, the index C'v\ can be rewritten as: C;k =[& + £min{l,r}]/3,for £ < 0 and
Clk =
[b-£ min{l, r}} / 3 for £ > 0,
where r = Dt / Du . Hence, the value of C ^ can be calculated from the values of £, b, and r . Specifically, if the triple ( £ , b, r) = (-1, 3, 3/2), C'lk = [3 + (-l)/max{l, 3/2}]/3 = 7/9 = 0.7778. Ssetting C^ = C , we arrive at b = 3C — £ / max{l, r} for £ < 0
PCIs with Asymmetric Specification
Limits
193
and b = 3C + £/max{l,r} for £ > 0. Recalling that B = n 1 / 2 x (d* /CT) and 6 = d j a we have 5 = n&2. Therefore, for Cp'k = C ,
2
B
n ( 3 ( 7 - £ / m a x { l , r}) 2 , £ < 0 (10.11)
= rc(3C + £
2
min{l, r}) , £ > 0.
Using the definitions of £ and n we use the (central) chisquare and the normal distributions to determine the critical value ca satisfying P(Cpk > ca \ C'p\ = C) = a . The last equation is equivalent to 1 — F^,„ (ca) = a given C'p'k = C . Observe that ca is in general larger than zero, hence we can find ca from Equation (10.9) B2
JFK(L(ca,y))fY(y)dy
= a,
(10.12)
o where B2 is given in (10.11) and L(ca,y) = (n — 1)(B - J~y)2 x (97JC 2 )- 1 in ( 1 0 . 1 0 ) .
As it is the case in the previous cases discussed above the critical values ca for £ = £0 and £ = — £0 are the same, provided T = M . To test if the process meets a capability (quality) requirement, we shall first determine the value of C and the a -risk. Since the process parameters fi and a are unknown, so is the parameter £ = (fj, — T) / a. However, £ can be estimated by calculating £ = (x — T) / s from the sample. If the estimated value Cpk is larger than the critical value ca ( C'p'k > ca ), we conclude that the process indeed meets the capability requirement ( Cpfc > C ). Otherwise, there is no sufficient information to conclude that the process meets the present capability requirement. In such cases, we would tend to believe that C'p\ < C (the process is incapable).
194
Encyclopedia and Handbook of Process Capability Indices
Following the standard time-honored approach we calculate the p-value, i.e. the probability that Cpk exceeds the observed estimated index given the values of C , £ , r = Dl / Du , and sample size n . This probability is then compared with the significance level a . If the (estimated) C'p' value is c* (given the values of C , £, r , and n) the p-value can be calculated as p - value = P(C'Jk > c \ Cp'k = C) = 1 - %sph( c * ) B2
+ ^4&Jy a a
+ tJn]]dy,
(10.13)
where Du / d = 1 / min{l, r} , Dt / d = max{l, r} (recall that d = mm(Du, De) ), B2 is given in (10.11) and L(c ,y) in (10.10)). The rather cumbersome calculations in (10.13) can be carried out by means of numerical integration. If the j>-value turns out to be smaller than the a -risk, we than conclude that the process meets the capability requirement (C'v\ >C). Otherwise, as above, no sufficient information is available to conclude that the process meets the current capability requirement. In this case, we would tend to declare that Cpk < C (namely that the process is incapable). A simple example could clarify the procedure described above. Consider a normally distributed process with asymmetric specification tolerances LSL = 20, T = 26.5, and USL = 32. Hence d =(USL - LSL)/2 = 6 , D, =T - LSL = 6.5 , Du =USL - T = 5.5, d* = min{D 0 , A } = 5.5 , and r = Dt / Du = 1.18 . To test whether the process meets the capability (quality) requirement, we first assign C = 1.33 , i.e., we define a process with Cpk > 1-33 as capable. With the sample size n = 100 , the sample mean x = 27, and the sample standard deviation s = 1.10, we obtain A* = m&x{d*(x - T)/Du,d*(T - x)ID{\ = 0.5 , |
PCIs with Asymmetric Specification
Limits
195
= (x — T) / s = 0.45 and finally Cpk = 1.515 . By using numerical integration to calculate the r.h.s. of equation (10.13) we arrive at the p-value of 0.055. We thus conclude that the process meets the capability requirement provided the a -risk is set to be larger than 0.055. If the a -risk is set smaller than 0.055, we again do not have sufficient information to assert that the process meets the current capability requirement.
10.3 The Cpm index for asymmetric tolerances Recall (Chapter 4, Section 1) that Chan et al. (1988) proposed the capability index Cpm using in the denominator r 2 = E(X - T)2 instead of a2 = E(X - nf . This index takes into account the proximity of the process mean to the target value and as well as the process variation. For asymmetric cases {USL-T ^T - LSL ), Chan et al. (1988) modified Cpm as follows. Simply shift appropriately one of the two specification limits, so that the new (shifted) specification limits are now become symmetric with respect to the target value T : C*pm = min
USL-T
3r
T - LSL
'
d* 3V
3r (10.14)
T)
where as above d = min{Du,Di} , Du =USL — T , and A =T - LSL . Obviously, if Du =Dt , then T = M = (USL + LSL)I'2 and d* = d =(USL - LSL)/2 , the specification tolerance becomes symmetric and the index C*m defined in (10.14) reduces to the original index Cpm defined in Equation (4.1). Boyles (1994) also notes that this generalization could understate process capability by restricting the process to a proper subset of the actual specification range. For the two processes E
196
Encyclopedia and Handbook of Process Capability Indices
and F with —fj,E < T , and jiF > T - satisfying the relationship | fJ-F — T | = | T — nE I (an equal absolute departure) - the index values calculated for the processes E and F are the same. Consider now specifically the two processes E and F with the target value T = (3USL + LSL)/A , /j,E = T - 0.5d = M , fiF = T + O.bd = USL , and aE = aF = d / 6 . In this case, we have | MF — T | = | T — nE | = 0.5c? and obtain the same C*m = 0.316. However, process E is eventually substantially better than process since the expected proportions of non-conforming items are approximately 0% and 50% for the processes E and F, respectively. An example for USL = 50 , LSI = 10 , T = (3USL + LSL) / A = 40 , ^ =T - 0.5d = 30 , fiF = T + 0.5d = 50 , and aE = aF = 1 0 / 3 is illustrated in Figure 10.5. Consequently, the index C*pm in appropriately measures process capability in these cases.
Figure 10.5. A example ioiUSL = 50 , LSL = 10 , T = 40 , [iE = 30 , UP = 50 , and aE = o> = 10/3 .
Definition (10.14) shows that Cpm assesses the distance of JJ, from T and the asymmetric specification limits. A natural estimate of Cpm can be defined as
PCIs with Asymmetric Specification
C*pm =
USL-T
m i n
197
Limits
T-USL 3f
3r
where f2 = [^"_ fa - T)2 /{n - 1)] 1/2 . When the process measurements follow a normal distribution, then, for given values of the parameters T , USL , and LSL , the p.d.f. of Y = Cpm is a variant of the inverted non-central x 2 distribution (note f2 appears in the denominator) and is given by:
My) =
9n/2
-exp
(("-1)3pra
y
y
. -/(n-l)C^sfl+,-
7 V(-
+A
E j!r(f+
, y > 0,
J)2 2J '
where A — n(fi — T)2 / a2 . The Cpm index can also be used to assess the process capability of the unilateral specification limit cases by letting LSL = — oo when only an upper specification limit exists and correspondingly letting USL = oo when only a lower specification limit is finite (see Chan et al. (1988), Stoumbos (2002)). Motivated by G. Taguchi's (1986) strong recommendation to apply the loss function to quality assessment procedures, Boyles (1990) generalized the index Cpm for asymmetric tolerances as follows: +
pm
_ USL - LSL
~
(10.14)a
6jE(L^sj
where E(Loss) denotes the expectation of the loss function
Loss(X,T)
k^X -Tf,X
< T,
kziX -Tf,X
> T.
where fc, (i = 1,2) are "standardizing" constants. We have:
(10.15)
198
Encyclopedia and Handbook of Process Capability Indices
E(Loss) =a2 [(l+e ){h ( l - * ( 0 ) + * & * ( £ ) } - ( * a - W ( 0 ] . e = (/i-T)/(7,
h = fa/ri)ko, h={ri/r2)k>, n=(T
(10.16)
- LSL) /(USL - LSL),
r2 = (USL - T) /{USL - LSL), k0 = max(ri /r2,r2/
n)/[2(r^ + r 2 2 )], and
$(•) and >(•) are the c.d.f. and the p.d.f. of a standardized normal variable, respectively. In the symmetric case when r i — r2 = 1/2 , E(Loss) — T2 and hence Cpm reduces to Cpm . Note that k0 is determined in a manner so that Cp = Cpm when \i = T (see Choi and Owen (1990)). As we have mentioned several times the indices Cpm , Cpm and Cpm consider not only the process variation but also the proximity of the process mean to the target value. However, they still do not take into account the location of the process mean in the basic interval [LSL, USL]. The Cpm index has an additional advantage in that it uses an asymmetric loss function to measure differently the capability at /U = T — h and // = T + h for a constant h > 0 in the case when USL — T ^ T — LSL . A drawback is that the calculations of E(Loss) in (10.16) are somewhat complicated, even when utilizing current truly "mighty" computer "capabilities". (Not to be confused with the capabilities in the sense used in this look!) The new generalization Cpm, which assumes a larger value for a more capable process (as the traditional indices Cp , Cpk , Cpm, and Cpm do), have been developed by Chen et al. (1999) to be:
PCIs with Asymmetric Specification
Limits
199
where d*=min{£>u, A } , A* = max{d\n - T)/Du,d*(T - n)/ Dt} , Du =USL-T and Dt =T - LSL (Compare with C'v'k in (10.5)). Obviously, if T = M (a symmetric tolerance), then d = d , A = | n — T | and Cp'm reduces to the original index Cpm . The index C"m is > 0 when the process mean falls within the tolerance limits. However, according to the modern quality improvement theories, reduction of variation from the target is as important as meeting the specifications requirements. The factor A in the definition ensures that the generalization Cpm achieves its maximal value at /j, = T (the process is on-target) regardless of whether the tolerances are symmetric ( T = M ) or not ( T ^ M ). Furthermore, for the processes E and F with a E = &F) MB > T , /JLF < T , satisfying the equal departure ratio relationship (/j,E — T)/ Du = (T — fip)/ Di , the index values assigned to the processes E and F are the same. In fact, the value of Cpm decreases faster, when fi shifts away from T , towards the closer specification limit than that to the farther one. In particular, for processes E and F with <JE = aF, /z# = LSL, and UP = USL, the index values of C'v'm assigned to the processes E and F are the same. In this special case, for normal processes with fiE = LSL and fip = USL , the same index values indicate the same expected proportions of nonconforming items. Furthermore, given Cpm > c , we can obtain a bound on | fi — T \: T - [(1 - R)/3c]Di < ii < T + [(1 - R)/3c]Du
,
(10.18)
where R = \ 1 - r | / ( l + r) and r = Dt / Du . As we have mentioned above, the index C*m cannot differentiate between processes capabilities in a accurate manner because the index values of C*m for two processes with equal variance and equal absolute departure are the same. This measure is therefore not very reasonable, especially for processes with asymmetric tolerances. On the other hand, the index values of Cpm assigned to processes with equal variance and equal departure ratio are the same. In this case, both processes have an
200
Encyclopedia and Handbook of Process Capability Indices
equal average loss. Therefore, it can be concluded that Cpm is superior to C*m based on the criteria of process targeting which is related to process loss. Let Xi,x2,- •-, xn be a random sample taken from a normal distribution N(fi, a2). To estimate the new generalization C" , we propose the natural estimator given by <%»=
,f
>-,
(10-19)
where x = ^ . _ xi / n , A = max{d(x — T)/ Du, d(T — x) / Dt} and s\ — ] P _i (xi ~ %)2 In • This estimator should be obtained from a process that is "demonstrably" stable (in-control). If the production tolerance is symmetric, A is simplified to | x — T \ and the estimator C'pm becomes
which is the natural estimator of Cpm discussed in Boyles (1991). The natural estimator C'p'm (10.19) can be rewritten as:
^-iznr-
(ia20)
where R = n1/2d* j a, K = nsl/a2, and Y = [max.{{d/Du)Z,-(d/Dl)Z}f with Z = nll2{x - T)/ a. Under the normality assumption, the statistic K is distributed as Xn-i > % as N(6,l), 6 = nll2(n — T)/ a and the p.d.f. of Y is expressed in terms of /y(-) as:
+ >0 ^)=2^(^(-^/*) ' 2Vy [ 4 " V ^ / - / • ^^ ( ^ / * ) ] ^
(1 21)
°-
where d\ = d / Dl and (4 — d / Du regardless of the specific distributions. Hence, the p.d.f. of Cpm can be expressed as:
PCIs with Asymmetric Specification II 3
fd" 0*0
c.pm •1 1 /if = 27z 4 Jo Jo7T VI X
dy
[
Cpm (1
—
201
Limits
*)
2
9x
k' „pm m V £
+ -rfz
3xdi
3x^2 J
dt, x > 0
(10.22)
(See Chen ei a/. (1999) for details.) Since the statistic Z2 follows a non-central chi-square distribution with one degree of freedom and the non-centrality parameter 62 , Chen (1998) defines the distribution of Y as a weighted non-central chi-square distribution with one degree of freedom and the non-centrality parameter ( 62 under the assumption of normality). This author also presents the p.d.f. of Y , in an alternative form (Equation (10.21)):
fr(y)
e~x'2
(V2gy r /i + j
2V7T
, i=l
y>0.
(10.23)
a
i
(Recall the connection between the central and non-central x 2 distributions.) Here A = <52 and Yj is distributed as Xi+j • Hence, the p.d.f. of Cpm can also be explicitly expressed as: 2^-n12
fc- (*) = d-U+V
rjnx~(n+l)
3T((n-l)/2) r\1_y){n-3)/2yU-D/2 J0
-A/2
uscy
2SKJ^J\\SX}
U
e x p f " ^2 (1 - y + d~2y) 18x
a; > 0. (see Appendix A of Chen et al. (1999))
•dy
(10.24)
202
Encyclopedia and Handbook of Process Capability Indices
If the production tolerance is symmetric (i.e. dx = d± = 1) the p.d.f. of Y reduces to the ordinary non-central chi-square distribution with one degree of freedom and the non-centrality parameter A = 82, and the p.d.f. of C"m becomes nl—n/2/~
/» =
-exp
AC 2 ) 2 k36x
jlT
3"n x„n+l oo
C2 18z2
A 2
[2+j
n
•, x > 0.
(10.25)
3=0
which coincides with the p.d.f. of Cpm (as given by Vannman and Kotz (1995)). Observe that the estimator Cpm is biased. The magnitude of the bias is Bias(Cpm) = E(Cpm) — C'p'm . The mean square error of Cp'm can be expressed as MSE{Cv\n) = Var(C'pm) + [Bias(Cpm)f . The bias and the mean square error of C'pm where calculated by Chen et al. (1999) for various values of £ = (fi — T) / a , d j a , di = d j Di, dq = d j Du , and the sample size n. The results show that as | £ | increases, the MSE decreases. As the sample size n increases, both the bias and the MSE decrease. Furthermore, when the value of d J a increases, both the bias and the mean square error increase for fixed d\, dq, £ and n .
10.4 The Cpn index for asymmetric tolerances Choi and Owen (1990) proposed a process capability index Cpn that takes into account all the three factors, the process variation, the proximity of the process mean to the target value and the location of the mean in the interval [LSL, USL] . The index is given by Cpn = mm{CPNU,
CPNL] ,
PCIs with Asymmetric Specification
where CPNU =
USL
~ M , and CPNL =
n
M
3T
203
Limits
~ LSL or
Also, we define Cpn = 0 for CPNU < 0 for CPNL < 0 . It can easily be verified that Cpn
=
Cpm ( 1 — K) ,
where k = 2 | T - \x \/{LSL +USL) as defined in Chapter 2. Similarly to Cpk and Cpfc , Cpn = 0 when k > 1. Choi and Owen (1990) also express Cpn in terms of the loss function as follows
C
"» = lS^TLS\ '
(10 26)
-
where E(Loss) denotes the expectation of the loss function Loss(X, T) = w(^)(X - T) 2 ,
(10.27)
with w(n) — 1/(1 - kf for A; < 1 and w(fi) — oo for A; > 1. The definition of C pn as given in (10.26) is the same as that of Cpm defined in (10.14a) except for the difference between the loss functions as defined in (10.15) and (10.27). Observe that the weight function W(/JL) reflects the centering of the process mean via the quantity \ M — fi \ while the indices Cpm , Cpm and Cpm do not account for | M — [i \ in their definitions. Choi and Owen (1990) provided an estimator of Cpn in the usual manner as: _ . \USL-x x-LSL] Cpn = mm | — , — J,
(10.28)
where, as above, f is the estimator of r obtained by replacing /j, and a2 by their UMVUE estimators x = ^ ajj / n and 2 12 s = [^._ (xi — x) /(n — I)] ' , respectively. The estimator C ^ i s
204
Encyclopedia and Handbook of Process Capability Indices
evidently consistent. For a given positive constant probability that Cpn > w can be expressed as:
w , the
P(Cpn > V>)
P
Z + Vn(Z\ 3yl^K + (Z + yR)2
> w, • , ^-^lK
-Z + -Jndv + {Z + 4<)
>w
(10.29)
where as above dj = (/z — LSL) / a , d? = (USL — fi) / a and £ = (/J, — T) I a . Also the random variables Z and K have the standard normal and the chi-square distributions with n — 1 degrees of freedom, respectively, and are independent. After some algebraic manipulations, for w > 1 / 3 , the probability in (10.29) can be expressed as P(CP„ >w) = P{0
D,A(K)
B(K)},
(10.30)
where D=
A(K) =
-Jn l-9w2
{n-l)
K +
(dlj - rfi + 2£)2
36wJ
•{(dw^-di)
+ V(9w2£ - dx)2 - (1 - 9«/2)[d? - 9^ 2 {ir/(n - 1) + e 2 }]},
and fl(tf)
Vn •{(Q^-dx) l-9w2
-V(9tw2C " ^i) 2 - (1 - ^2)[di
- 9w2{K/(n
- 1) + e 2 }]}.
(see Choi and Owen (1990) for details). Since the joint density function of the independent variablese K and Z is g(k,z) =
2 1 l .(n-3)/2 e -(ft+z )/2 ) V2^r[*j*]2(n- l)/2
random
PCIs with Asymmetric Specification
Limits
205
where 0 < k < oo , and —oo < z < oo , the probability in (10.30) can be represented as P(Cpn >w) = J o j =
m
g(k,z)dzdk
T[^-^fo\^m)-*{A{k))}k^l\-*l*dk.
(10.31)
There seems to be no closed form for the integral in (10.31). For a given a , however, the values of w can be obtained by means of numerical integration satisfying P(Cpn > w) = a . 10.5 The Cpmk index for asymmetric tolerances Recall that the index Cpmk is more sensitive to departure of the process mean from the target value T than the other three indices Cp , Cpk , and Cpm . Along these lines Pearn and Chen (1998) considered a generalization of Cpk for processes with asymmetric tolerances. The generalization takes into account the asymmetry of the tolerance, which would reflect the process capability more accurately than the original Cpk . Applying the same idea Pearn et al. (1999) proposed a capability index Cpmk , which is a generalization of Cpmk , to deal with processes possessing asymmetric tolerances. The generalization Cpmk is defined as Cpmk —
; 0 -') 3V
(10.32)
where d = {USL - LSL)/2,A = max{d(/x - T)/Du, d(T - /i)/D t } , Du =USL-T , Dt = T - LSL , d* = mm{Du,Dt} , and A* = Taax{d*(n-T)/Du,d*(T-fi)/Di} . Evidently, if T = M (a symmetric case), A = A* =\ ^ — T \ and Cpmk reduces to the
206
Encyclopedia and Handbook of Process Capability Indices
original index Cpmk. We note that Cpmk > 0 for a process with mean fi falling within the tolerance limits, analogously to the indices Cpmk , Spmk , and the yield-based index Cpk . However, as we have emphasized above, according to the modern quality improvement theories, reduction of variation from the target is as important for an index as meeting the manufacturing specifications. In fact, being on-target would be a highly desired condition for a process. The factors A and A ensure that Cpmk obtains its maximal value at [i = T (process is on-target) regardless of whether the tolerances are symmetric (T = M ) or not (T ^ M). Furthermore, for processes E and F with aE = aF, HE > T , fiF < T , satisfying the relationship (/j,E — T) / Du = (T — fxp)IDi (i.e., having the same departure ratio), the index values for the processes E and F are identical. Actually, the value of Cpmk decreases faster when n moves away from T to the closer specification limit, but decreases slower when fi moves away from T to the farther specification limit. Since under the normal assumption the Yield > 2Q(3Cpk) — 1 and C'p'mk < Cpmk < Cpk , given the value of Cpmk , one can calculate a lower bound of the process yield as 2Q(3Cp'mk) — 1 (Boyles (1991)). For example, given a process with capability Cpmk — 1 ) the process yield is guaranteed to be no less than 2<1>(3) — 1 = 99.73% . On the other hand, an upper bound on the departure ratio can be calculated as A/d
PCIs with Asymmetric Specification
C';mk =
207
Limits
d* -A* » ^ A ^ 2,
(10.33)
zjst + i
where d = min(Du, D{) , A = max{d(x — T)/Du,d(T — x)/D{\ , J4 = max{d (x — T)/Du, d (T — x)/D[} , and x and s 2 as above. If the manufacturing tolerance is symmetric, d = d , A = A =\x—T\ , and the estimator Cpmk reduces to Cpmk = (d-\x -m |)/{3[s 2 + (x - T) 2 ] 1 / 2 }, which is the natural estimator of Cvmk considered by Pearn et al. (1992) for symmetric case. Define now D* = n1/2d* /a , D = nl/2d/a , K = ns2n /a2 , Z = nll2{x-T)/a , Y = [max{(d / DU)Z, - (d/ D^Z})2 , 1 2 2 8 = n / (/i — T) j a , and A = 8 . Using these definitions, the estimator Cvmk can be rewritten as: &,
Cpmk -
_D*-(d*/d)JY ^K + Y •
(10-34)
It remains to recall that under the normality assumption, K is distributed as Xn-i ( a chi-squared variable with n — 1 degrees of freedom), and Z obeys the normal distribution N(6,l) with mean 6 and variance 1. Let as above <&(•) and (/>(•) be the c.d.f. and the p.d.f. of the standard normal distribution JV(0,1) , respectively. Then, the c.d.f. and the p.d.f. of Z can be expressed as: Fz(z) = $(z — 6) and fz(z) = 4>{z - 8) respectively. Consequently, the c.d.f. of Y is given by: FY(y) = P(Y
®{-d{l4v
^Jy)
- S).
where dx = d / Dx and d^ = d / Du (Don't confuse with dif i = 1,2 , defined in 10.4). The p.d.f. of Y is then given by
208
Encyclopedia and Handbook of Process Capability Indices
fy{y) =
f-FY{y) 2Vy
(^""VWr1 w -s) + df Wi'JV
T h e c.d.f. of C'^ FA„
+ s)).
(10
can now be expressed as:
=
(X)
o,
x < --
(D* - u< JS(x)if •K /
Ox1
•
-S(x)t
fY(S(x)t)S(x)dt,
< x < 0, (10
1
1
l-^(d^ D-6)+^(-dr D-6),
I-]FK
(D* - u' JS(x)if
i = 0,
-S(x)t
fy(S(x)t)S(x)dt,
(where if1 = — , d — min(Du,
x > 0,
Dj))
(See Appendix to Pearn et al. (1999)) and the p.d.f. of Cpmfc becomes:
PCIs with Asymmetric Specification
Limits
209
/ <Spink? " » =
JIK
b{X)t
9^
/ctr^2S(x)(D*-u<JS(x)i) fv (S(x)t) i — -9x6
'— dt,
< x < 0,
(10.37)
JfK
(D* -u<JS(xjif - S(x)t 9a? 2S(x)(D*
fy (S(x)t)
,{
9 g
-u'JS(xji) 3 **>
x > 0,
where D* = n1/2d* / a , u' = d* / d , S{x) = [D* /{u' + 3a;)f , F^-(-) is the c.d.f. of K , fx{-) is the p.d.f. of K , and /y(-) is the p.d.f. of Y given by (10.35). Pearn et al. (1999) derive the r-th moment of Cpmk without utilizing the distribution of Cp'mk • Note that the estimator Cpmk is biased. The magnitude of the bias is Bias(C';mk) = E{C';mk) - d;mk , and the MSE is MSE{C';mk) = VariC^ + iBiasiC'^f ^ where Var(C'p'mk) = E{C';mkf [E(Cpmk)f is the variance of Cpmk . The results indicate that the bias of Cpmk is positive when H ^ T . That is, C'p'mk is generally overestimated by Cpmk . On the other hand, when \i = T , we have A = A — 0 and Cpmk — d /(3cr), the bias of Cpmk tends to be negative for certain values (see Pearn et al. (1999)). Thus, C'p'mk may be smaller than Cvmk and the bias be negative when n = T . This is partially due to the fact that both A and A* are calculated to be positive even for \i — T while A = A = 0 . Clearly, the presence of A and A* decreases the value of the calculated C'p'mk . As the sample size n increases, the mean square error of Cpmk decreases. Consequently appropriate sample sizes for capability estimation are essential.
210
Encyclopedia and Handbook of Process Capability Indices
The smaller the sample size is, the higher value of C'v'mk will be required to ascertain the true process capability. We have mentioned above that Vannman (1994) investigated a general class of capability indices for processes with asymmetric tolerances. Her generalizations have been defined as
3yja2 + v(ju - Tf where u, v > 0 . Vannman (1994) showed that among numerous (u,v) values, the pairs (u, v) = (1,3) and (it, v) = (0,4) generate two indices which are most sensitive to process departures from the target value. For u > 1, the indices Cpa(u,v) decrease when mean fi shifts away from the target T in either direction. In fact, Cpa(u)w) decrease faster when ^ shifts away from T to the closer specification limit than that to the farther one. This is quite an advantage since the index would respond faster to the shift towards "the wrong side" of T rather than towards the middle of the specification interval.
10.6 The loss index for asymmetric tolerances Johnson (1992) developed the relative expected loss Le (discussed in Chapter 5 Section 1) for symmetric case, which is defined as the ratio of the expected quadratic loss and the square of the half specification width: = /.
(x - Tf dF(x) 2 d
( M - T ) 2 + a 2 _ / M - T \ 2 , lu\ + 7 , d2 \ d j \dl
(10.38)
PCIs with Asymmetric Specification
Limits
211
where F(-) is the c.d.f. of the measured characteristic. Under asymmetric tolerances conditions would be risky and it is quite possible that the results obtained may be misleading. Consider the following classical example with the asymmetric tolerance, (LSL, T, USL) where T = (3 USL + LSL)/A and a = d/3. For the processes A and B with \i^ = T — d / 2 = M (the midpoint of the specification interval) and \IQ = T + d / 2 = USL both processes have the index value Le = 13/36 and an equal degree of clustering around the target, namely, | fi — T | = d/2 for both processes A and B. However, the expected proportions nonconforming are approximately 0.27% for process A and 50% for process B. Evidently, Le measures process capability inconsistently in this case, and is inappropriate for the situations with asymmetric tolerances. To overcome this consistency, Pearn et al. (2003) proposed a modification of the expected loss index, which is referred to as L'J , to deal with processes in both symmetric and asymmetric tolerances cases. In developing the generalization, the term | fi — T | in Le have been replaced by A. Adding the terms (d J Du) and {—d/D{) to (fi — T) according to whether /i is greater or less than T, the asymmetry of the loss function is taken into account. Regardless of whether the tolerances are symmetric or not, these modifications ensure that the new index attains the minimal value at /i = T . Additionally, the half specification width d is substituted by d where d is given below. Actually Pearn et al's (2003) generalization of the expected loss index can be expressed as L" =
(A)
2
/
x2
+ [Y) >
( 10 - 39 )
where A = max{(/x - T)d/ Du, (T - fi)d / £>,} , Du = USL - T , Di = T — LSL and d = vnm{Du, D{\ . The quantity d will appear in numerous expressions related to the estimation of Le" . Note that Le is sensitive to target value T and attains larger
212
Encyclopedia and Handbook of Process Capability Indices
value when T is away from the mid-point between the upper and the lower specification limits (M = (USL + LSL)/2). Denote (A/dJ by L':t , (a/d*)2 by L'Je, obtaining l£ = C + L^ . The terms (d/ Du) and (—d/D{) ensure that if the processes A and B with fiA > T and \xB < T , respectively, satisfy (fiA — T) / Du = (T — fiB)/Di, then the index values obtained for A and B are the same. Evidently for symmetric tolerances ( T = M ), A = \ n - T \ , Du=Dl=d , and d* = d = (USL - LSL)/2 . Accordingly, in such a case the generalization L" reduces to the original index Le. To estimate a loss index L" Pearn et al. (2003) proposes a natural estimator of the form:
L'I = A_ d*
' a
+ |4
\2
(10.40)
where A = max{(X — T)d/Du, (T — X)d/Di} , the mean // is estimated by the sample mean X = ^ J . Xt j n and the variance a2 by Si = YTi=SXi -X? /n, which is the maximum likelihood estimator (MLE). In the case where the production tolerance is symmetric, A becomes \ X — T \ . Therefore, the estimator L" reduces to Le = (n~ld~2)'^2l. (JQ — T)2 , the natural estimator of Le proposed by Johnson (1992) (Chapter 5 Section 3). Consequently, we may view the estimator L" as a direct extension of Le.
The rth moment about zero of L" has also been derived by Pearn et al. (2003) using the properties of a non-central Xv distribution:
E(L':Y r X
E i=0
a nd
e-V2 2VTF
2rr E(^) j=0
r([(l + j ) / 2 ] + i) T([(n + j)/2] +
n+ j
\(di-iy +
+ r
{-iy{df-iy\
(10.41)
PCIs with Asymmetric Specification
Limits
213
where />• = W / O ' O , <4 = d/Du , dh = d/Dx , yu = (y/d2u) , Vi — (y / df) , X = 62 and 6 = V7T(/x — T) / a . In particular, the expected value and the variance of Le are:
E(W) =
({n-l)a nd
2
2 \
+
a n
i+i
U-A/2 2V?f
(I + J K + H W
(10-42)
3=0
and
Var(W) -A/2
(7 2J*4
2V¥
n^d
2 O"
A/2
^
2V?f
nd
(2(n-l)a4)
+
£(W
2/4
l + j (i + i)(3 + j ) [ ^ + (-iy^ 4 ]
°°
E(W
l + j
(i +
j)[dl+(-iydf}
3=0
(10.43)
nd
The estimator Le" is biased as are many of the estimators discussed in this Chapter. The bias of I/e" may be computed in a standard manner to the Bias(L") = E(L") — L" , and the MSE, which is more relevant to an analysis of process quality, is given by MSE{L'^ ) = Var(L? ) + [Bias(W )f . The results in Table 2 of Pearn et al. (2003) indicate that as | a | increases, so do the bias and the mean squared error. Furthermore, as the sample size increases, the bias and the MSE decrease. In those cases when the production tolerance is symmetric, which implies du = du — 1, L" is an unbiased estimator of L" , ( Bias(L") — 0 ). The unbiased estimator depends only on the complete, sufficient statistic (X,S2) for (/i, a2) , by the Lehmann-Scheffe Theorem (Hogg and Craig (1989)) L" is an uniformly minimum variance unbiased estimator (UMVUE) of L".
214
Encyclopedia and Handbook of Process Capability Indices
Let Xi,X2,...,Xn be a random sample of size n from a normally distributed process N(fi,a2) . Then the p.d.f. and the c.d.f. of Le can be expressed as:
fL^)=J^^fK(Bx(l-t)) o x (di1(f>(di1^Bxi
+ 6) + d^1(l)(d-1yfBxi-6))dt,
(10.44)
FL„(x)= f FK(Bx - y)-L= x {d^dj1 Jy + 6) V2/ o + c ^ V K " 1 ^ - S))dy , for x > 0,
(10.45)
Bx
and
respectively where fK(-) and F^Q denote the p.d.f. and the c.d.f. of K, distributed as %n-i.
A decision
making rule for testing
L"
Based on the sampling distribution of the L" . (Equations (10.44), (10.45)), we shall develop a decision making rule, which can be used for testing capability of a process. Using the index L" , the engineers can access the process performance and monitor and control the manufacturing processes on a routine basis. Thus, we can consider a testing hypothesis with the null hypothesis and the alternative hypothesis to be: H0 : L" > C (incapable), versus Hl : L" < C (capable). The null hypothesis H0 will be rejected if Le" < ca , where constant ca , called the critical value, is determined so that significance level of the test is a , i.e., P(L" < ca \ L" = C) = The decision making rule to be used is then stated as follows:
the the a . for
PCIs with Asymmetric Specification Limits
215
given a-risk (the probability of wrongly rejecting the null hypothesis when it is true) and sample size n, the process will considered to be capable if L" < ca and incapable otherwise. Note that, by defining the characteristic parameters a = (/J, — T) J a and b — d / a , the indices L„t and Lpe can be rewritten as L"t — (dia/b)2 for a < 0, Lpe = (dua/b)2 for a > 0, and Lpe = (1/6) 2 , where du = d/Du, and dt = d/Dt. Hence, the value of L'J = L"t + Lpe can be calculated given values of the four tuple (a, b, du, df). For example, if (a, b, du , d{) = (1, 3, 2/3, 2) 4 " = (2x 1/3)2 + (1/3) 2 = 5/9. If 4 " = C, using 4 " = L^ + L%e, we obtain: (deaf 1 . ^-V 2 - + -o2 for 6 6
^ M , „ (d„a)2 1 a < 0 and C = v V + -5£r ir
In this case b2 = [(dea)2 +1] / C for a < 0 and 62 = [(dua)2 + 1 ] / C for a > 0. Moreover, we have B = nd*2 J a2 = nb2 . Hence, if L" = C, the term B becomes:
and
B = n[(dea)2 +1]/C
for a < 0
B = n[(dua)2 +1]/C
for a > 0.
(10.46)
Furthermore, we have the equality 6 = n 1 ' 2 (/i — T)/a = n1'2a (see (10.41)). With the result of (10.45), we can use the central chi-square and the normal distributions to find the critical value ca satisfying P(L" < ca | L" — C) = a , i.e., F^,(ca) = a given Lg = C, or Bca
J FK{BCU - y)^-={dil(t>{dil4y + -Jna) + d~l(l>{d~ljy - -Jna))dy = a,
(10.47)
216
Encyclopedia and Handbook of Process Capability Indices
where <j> is the standard normal p.d.f. and B is given in (10.46). Evidently determination of ca may require tedious calculations. Note that the distribution characteristic parameter a = (jU — T) J a in the expression (10.47) is usually unknown, and has to be estimated in real-world applications, substituting [i and a by the sample mean X and the maximum likelihood estimator (the sample variance) S% = ^ ™ ( J Q — X ) 2 / n . To reveal explicitly the relationship between a and ca, we shall examine the behavior of the critical value ca against the parameter for a = -3(0.05)3. Note that a = (/i - T) / a = -3(0.05)3 covers a wide range of real world situation involving process capability analysis. The results of these calculations indicate that the critical value reaches its minimum at a = 0.5 (corresponding to a small deviation between /J, and T or large value of a) in all the cases with accuracy up to 10~3 . Making
decisions
by means
of a critical
value
Whenever the calculated value of L" is smaller than the critical value ca ( L" < ca ), we shall conclude that the process meets the capability requirement (L"
decisions
by means
of a
p-value
An alternative decision rule can be stated as follows. After calculating the estimated statistics via the sample information, we determine the corresponding p-value, i.e., the probability that Le does not exceed the observed estimated index given the values of C, du, dt, a, and the sample size n, and then compare it with the significance level a . If the estimated index value is 1® , (given the values of C, du , dt, a, and the sample size n), the p-value can then be calculated using the:
PCIs with Asymmetric Specification
Limits
217
p-value = J FK(Bl0 - y)^-=(^V(^ _ 1 >/y + >/*"») + (d^id^Jy
- Vna))dy ,
(10.48)
where B is given in (10.46). Compare with (10.47) the numerical calculations can easily be carried out using the computer software, to integrate the function based on the chi-square distribution and the normal distribution. If the p-value is smaller than the a -risk, we shall conclude that the process meets the capability requirement ( L"
Chapter 11
Supplier Selection Based on PCIs
11.1 11.2 11.3 11.4
Introduction Tseng and Wu's MLR selection rule based on Cv Chou's approximate selection rule based on Cpu and Cp[ Huang and Lee's approximate selection rule based on Cpm
11.1 Introduction
In an initial stage of production setting, the decision maker often faces the problem of selecting the best manufacturing supplier from several available candidates. There are many factors, such as quality, cost, service and so on, which ought to be considered when selecting the best supplier. Several selection rules have been proposed for selecting the means or variances in the analysis of variance (ANOVA) (see Gibbons, Olkin and Sobel (1977), G u p t a and Panchapakesan (1979), G u p t a and Huang (1981) more details). T h e vast majority all of the selection rules is based on the ordered sample variances. A common persistent drawback of these selection rules is t h a t the information available sample d a t a cannot be efficiently used. As it was mentioned in the earlier Chapters of this book, process capability analysis is applied to determine whether the process capability of a supplier conforms to customer's requirements, by evaluating the process capability index (PCI) to a controlled process. 218
Supplier Selection Based on PCIs
219
In the first 10 chapters of this book we have studied the most common PCIs. In the nutshell, PCIs provide common quantitative measures of the manufacturing capability and production quality, to be used by both producer and supplier as a guiding reference when signing a contract. Purchasing personnel may use the PCI to decide whether to accept or reject products provided by suppliers. As we have indicated on numerous (perhaps too numerous!) occasions in this book in practice the process mean and the process variance are usually unknown. In order to calculate the index value sample data ought to be collected, and a great degree of uncertainty could be introduced into the capability assessments if the sampling errors are ignored. A multitude of methods have been developed to evaluate whether a single supplier's process conforms to customer's requirements. However, few studies have been addressed to the problem of selecting between two or several suppliers' PCIs. There are two common methods that are available to determine the better suppliers' PCI. First 100% inspection is performed to calculate separately the PCI for each supplier, the suppliers can then be compared according to their respective true PCI values; this approach is very expensive and time-consuming and is therefore rarely used in practice. The second method involves sampling implementation, and statistical testing is then used to assess suppliers' process capabilities. 11.2 Tseng and Wu's MLR selection rule based on Cp In a situation when the manufacturing process is in control we are going to assume that the quality characteristic X is normally distributed, USL and LSL are usually fixed and are determined in advance. Tseng and Wu (1991) consider the problem of selecting the best manufacturing process from k available processes based on the "precision" capability index Cp (see Chapter 1) and have proposed a modified likelihood ratio (MLR)
220
Encyclopedia and Handbook of Process Capability Indices
selection rule. Suppose that there are k available manufacturing processes, 7^ , 7r2 , ..., ^ , and that the product quality characteristic data of each process 7Tj follows a normal distribution N(fa, of) , i = l,---,k . Without loss of generality, it may be assumed that each process has the same process mean (i.e. fa = • • • — fa = fj, ). Kane (1986) proposed the statistic, Cp(i) = (USL — LSL)/(6ai) representing the process capability ratio (PCR) index 7Tj , where, as above, USL and LSL are the upper and the lower specification limits of the process, respectively. Let (Xn, Xi2,- • •, Xin.), i = 1,2,- ••, k , be a random sample of size r^ , taken from the population of the process 7r,. The estimated PCR (to be denoted by Cp(i)) of the process TT{ is given by A ,.*
6p(j)
=
USL-LSL 6s,
where Si
—
Based on the estimated PCR values, the MLR selection rule S = (6i,62,- • ;6k) f° r selecting the best manufacturing process is as follows: For each i = 1,2,- --,k the rule St states: select the process 7r, if and only if
(n^V/lCM2
<
di,
Z(nt-l)/[Cp(j)? i^i
where d^ is a constant to be determined below. This MLR selection rule is then completely specified provided the sample sizes (ni,n2-l- • -, nk) and the critical values (d^d^,- • •, dk) are known. In order to determine these values Tseng and Wu (1991) propose the following two criteria.
Supplier Selection Based on PCIs
221
1. Probability of correct selection: The selection of rule <5j is called a correct selection (CS) if 7Tj is the best process and it can be correctly selected by 6t. Usually, it is required that the probability of CS exceeds a minimal value P , referred to as the P -condition. 2. Error probability: The selection of rule <% is called an error selection if there is no significant difference in Cp index values among these k processes and 7Tj is selected arbitrarily. Usually, it is required that the probability of the error selection be less than a maximum value a , referred to as the a -condition. It is reasonable to propose the ratio r^ = [Cp(i)]/[Cp(j)] as a measure of separation between the processes 7Tj and TTJ . For a fixed i , we consider T = (TH,- •;Tii_i,Tii+i,- • -,Tik) as the separation vector of process 7Tj from the remaining processes Ttj , j vt %. Evidently, process 7Tj would be called the best process if and only if r^ > 1 for all j ^ i. If f^ represents the Ith preference region and Q0 represents the indifference region, then they can be suitably defined as follows: fij = { r | Tij > A, for all j ^ i}, A > 1, and ^o = { T I Tij = 1, for all j ^ i} . Now for i = 1,2,- • •, k the P -condition and the a -condition can be stated as follows: P r ^ C S I ^ ] > P*
for all r G ^ t ,
Pr T [CS| 6t] < a*
for all r e fi0 ,
and
where Pr,.[CS | <5j] represents the probability of correct selection of rule <5j under r .
222
Encyclopedia and Handbook of Process Capability Indices
More details related to theoretical derivations of the MLR selection rule are given in Tseng and Wu (1991). Some Tables of the sample size and of the critical values for selecting the best manufacturing are also computed by controlling the probability of a correct selection (CS) and the error probability based on the proposed MLR selection rule. Furthermore, when the product quality characteristic data for each process follows a non-normal symmetric distribution, a simulation study was carried out to examine the robustness of the selection rule. The results indicate that the proposed MLR selection rule is insensitive to a number of non-normal symmetric process distributions (such as logistic and uniform).
11.3 Chou's approximate selection rule based on Cpu and Cpl As mentioned above, we cannot compare two suppliers directly since the 100% inspection is not practical to use. We are compelled to sample some products made by these two suppliers, and then use a statistical analysis to determine which one has the superior process capability. Chou (1994) developed an approximate method for selecting a better supplier based on onesided capability indices Cpu and Cp\ when the sample sizes are the same, which deals with comparing two one-sided processes and selects a better one with higher process yield. Let X 1 1 ,X 1 2 ,...,X l n and X2i,X22,...,X2m be the measurements of two samples independently drawn from two suppliers 7Tj following the normal distributions N^i^af) , for i = 1,2 . Based on the hypothesis testing for comparing the two Cpu values, H0 : Cpui > Cpu2 versus Ht : C p u l < Cpu2 • Chou (1994) has carried out a variable transformation so that Wj = USL — X, will follow normal distributions N(USL — fa,erf), for i = 1,2 . The sample mean and the sample standard deviation, Wi and 5, , are calculated for suppliers i , i = 1,2 . Thus, the test is equivalent to testing H0 : Wi J ox > w2 J o2 versus Hi : w^ / (Ji < w2 / a2 • Furthermore
Supplier Selection Based on PCIs
223
under the assumption that n = m , Miller and Karson (1977) showed that the likelihood ratio test statistic is given by A =
2YY {(W? + 2Y1 f/ (W22 + 2Yi)1'2 2 2
- WXW7
where Y2 = (n — 1)S? / n is the maximum likelihood estimator of the variance a2 , for i — 1,2 . Using the likelihood ratio test, the null hypothesis H0 is rejected if
Wl/S1
<W2/S2
and A < c
for some constant c satisfying 0 < c < 1. The above is equivalent to Cpui < CPu2 and A < c , where A =
[(aCPui + 2)
{o-CpU2 + 2)
— aCpuiCpu2
and a = 9 n / ( n — 1). Since yl is a function of (7pnl and C ^ > it's cumbersome to derive the exact distribution of A which cannot to the best of our knowledge be simplified to a closed form. Observing however that — 2\nA is the log-likelihood ratio statistic approximating a chi-square distribution with one degree of freedom under the stipulations of large-sample theory and that HQ is true we can provide a test with the critical value c that can be expressed as
exp
Xi 2 (l-2a)1
0 < c < 1.
224
Encyclopedia and Handbook of Process Capability Indices
Here —xi(l — 2a) is the (1 - 2a) -th quantile of a chi-square distribution with one degree of freedom. Consequently, the rule rejects H0 provided C pul < Cpu2 and the condition A < c = exp{-xi(l - 2 a ) / 2 } is satisfied. The selection method proposed by Chou (1994) utilizes some approximating results but provides no indication on how one could further proceed with selecting the better supplier by testing process capability index Cpu or Cpi . Pearn et al. (2004) investigated the selection power analysis of the method via simulation. The accuracy analysis provides useful information regarding the sample size required for designated selection power. To simplify the presentation, we shall use Cpu only. When only the lower specification LSL is given, we calculate Cp[i , i — 1,2 for the zth supplier and a selection procedure also can be constructed in the same way. That is, if the minimum requirement of Cpu values for two candidate processes, and the minimal difference of Cpus between these two suppliers, 8 = Cpu2 — Cpui, are determined in advance how many sample sizes are required such that the suppliers can be differentiated with the designated selection power. The selection power corresponds to calculating the probability of rejecting the null hypothesis H0 : Cpui > Cpu2 , when actually Cpu\ < Cpu2 is true, by using simulation technique with 20,000 replications. To render this method practical for inplant applications, a two-phase selection procedure is developed by Pearn et al. (2004) to select the better supplier and examine further the magnitude of the difference between the two suppliers.
11.4 Huang and Lee's approximate selection rule based on Cpm We have already mentioned on several occasions that, according to modern quality improvement theory, reduction of the process loss is as important as increasing the process yield. The use of loss functions in quality assurance settings has been invigorated with the introduction of the G. Taguchi's philosophy. The index Cpm incorporates the variation of the production items
Supplier Selection Based on PCIs
225
with respect to the target value and specification limits preset in a factory. Based on the Cpm index a some what mathematically complicated approximation method is developed by Huang and Lee (1995) for selecting a subset of processes containing the best supplier from a given set of processes. Under these circumstances, a search for the larger Cpm which are used to provide unitless measure of the process performance is equivalent to a search for the smaller 7 2 . We emphasize again that the term 2 = a2 + (/j, - Tf = E[(X - T)f is the major part of the 7 denominator of Cpm , which incorporates two variation components: (i) variation with respect the process mean and (ii) deviation of the process mean from the target, (c.f. Chapter 4, Section 1) The method essentially compares the average loss of a group of candidate processes, and selects a subset of these processes with a small process loss 7 2 , which, with certain level of confidence, contains the best process. Since the specification limits are usually fixed and determined in advance, searching for the largest Cpm is equivalent to searching for the smallest j 2 . The selection rule of Huang and Lee (1995) is that one retains the population i in the selected subset if and only if, j 2 < c x m i n ^ ^ j , . , j 2 where the value of c is determined by a function of parameters, which can be in turn determined by calculations from the obtained samples. Let 7Tj be the population with mean /ij and variance a2 , i = 1,2,'••, k , and Xiy,Xi2,...,Xin. be the independent random samples from -ni. If the populations are ranked in terms of 7 2 , our interest would be to select the best process with the smallest value of 7 2 . A correct selection is coded as CS, and the ordered 7 2 in the form of 7r2, < 7,22| < • •• < 7^, are assumed. Now denote by 7 ^ the population associated with 7(21 , i = 1,2,- ••,k. In this case, the best population is ir^. We wish to define a procedure with a selection rule R such that the probability of a correct selection be no less than a pre-assigned number p * where 0.5 < p* < 1 , i.e. Pr(CS | R) > p * . This is
226
Encyclopedia and Handbook of Process Capability Indices
requirement is referred to as the p *-condition. The selection rule R based on the unbiased and consistent estimators •yf of 7? , i = 1,2,- •-, k , defined as follows:
E(^-r)2
(rii -l)Sf
+n,(xi
-Tf
h
where
Sf =
n,; -
-. /__j \Xij 1
X
i)
)
x
i
2—4 xij •
— Ui
j=l
j=l
Comparing the estimators Cpm (c.f. Chapter 4 Section 2) among k populations is equivalent to comparing the values of jf, i = 1,2,-••, k . Hence, using the result of Pearn, et al. (1992), concerning the distribution of Cpm :
if ~— xl,(\), \
(IM-T)
m
where, as above, Xn W) i s the non-central chi-squared distribution with n^ degrees of freedom and the non-centrality parameter A,. Huang and Lee (1995) consider the problem of selecting the populations with the smallest 7 2 . The selection rule R is here as follows: Consider 7Tj as the better supplier if and only if 7I < c • min1<J<j.i j^i 7j , where c > 1 . In order that the p * - condition, be satisfied, the appropriate c ought to be q = exp- -2 A
T
+
1_
_1
I Til]
l[k})
227
Supplier Selection Based on PCIs
Choose now the value of c which is larger than 1 and choose the value as small as possible, so that c = min{c 1 ,c 2 }, if q > 1 and <% > 1, c= q
if q > 1 and c^ < 1;
c = q,
and c = Oi if c? > 1 and q < 1;
where the factors are
A =
-dz + yj4 ~ 4*443
24 fc-i
A, -
-£fe - Vdf - 4di^3 2di
Jfc-1
tij = a x y ] 1 + O/t
+a
i=i
fc-i
£-" <:-i
ak
ab T-^ yjdi +
£i = l a>i +a U7= iE * = »xEJi+? fc-l
d, =
4a
ak^/aj
(k-l
.
\
\2
.
In[2p 2* _ W2a •
1 V ffl,
a = 0.5 — a
E
\ 2= 1
a
k »
6 = -0.513277 Mi
a = -0.085514.
«i +2Aj
,and \ = Hi
Here ^ is an estimator of vt, i = 1,2,- • •,fc, and the ordered «j are denoted by v^ < i\2} < " < %] • The method, however, provides no indication on how one could further proceed with selecting the best population among the chosen subset of populations. Pearn et al. (2004) have investigated this method for the cases with two candidate processes. In practice, if a new supplier II wants to compete for
228
Encyclopedia and Handbook of Process Capability Indices
the orders by claiming that its capability is better than that of the existing supplier I, he (she) must furnish convincing information justifying his (her) claim with a prescribed level of confidence. Thus, the sample size required for a designated selection power must be determined in order to collect actual data from the factories. If the minimum requirement of Cpm values for the two candidate processes, Cpm0 , say and the minimal difference $ = Cpm2 ~ Cpmi a r e determined, the sample size required need to be such that the suppliers could be differentiated with designated selection power. Thus, based on the proposed selection procedures, if 72 < c ' 7i a n d 7i > c • 7I we shall conclude that the Cpm of 7r2 is better than that of 7rx. Otherwise, we would assume that the existing supplier I is better than the new one II since we don't have sufficient information which would allow as to reject the null hypothesis. The accuracy of the selection method is investigated using simulation technique with simulated 10,000 numbers. For users' convenience in applying this procedure in practice, the sample size required for various designated selection powers = 0.90, 0.95, 0.975, 0.99 with various capability requirements Cpm = 1.00, 1.33, 1.50, 1.67 and the differences 6 = 0.05(0.05)1.00 under the p * -condition = 0.95 are tabulated in Pearn et al. (2004). The selection power which amounts to calculate the probability of rejecting the null hypothesis HQ : Cpm\ > Cpm2 > while actually Cpmi < Cpm2 is true, using a simulation technique. (See above). It should be noted that the required sample size is a function of Cpm , the difference 8 between the two suppliers capability indices and the designated selection power. From the tables of Pearn et al. (2004), it can be seen that the larger the value of the difference 6 , the smaller the sample size will be required for a fixed selection power. For fixed 6 and Cpm the required sample size increases as the designated selection power increases. This phenomenon can easily be explained, since the smaller is the difference and the larger is the designated selection power, the larger sample size will be required to account for a smaller uncertainty in the estimation.
Supplier Selection Based on PCIs
229
The following two-phase selection procedure was developed by Pearn et al. (2004) to select a superior supplier and further to examine the magnitude of the difference between the two suppliers. Phase I - Supplier Selection In most applications, the supplier selection decisions would solely be based on the hypothesis testing, comparing the two Cpm values: H0 : Cpml > Cpm2 versus Hx : C p m l < Cpm2 . If the test rejects the null hypothesis H0 : C p m l > Cpm2 , there is then sufficient information to conclude that the (new) supplier II is superior to the (original) supplier I, and the decision "to replace" would be suggested. Phase II - Magnitude Outperformed
Measurement
We have just seen that In the Phase I of supplier selection problem, the decision would be solely based on the hypothesis testing comparing the two Cpm values (without a further investigation of the magnitude of the difference between the two suppliers). In other applications, supplier selection decisions would be based on the hypothesis testing comparing the two Cpm values: HQ • Cpmi + h> Cpm2 versus Hx : CpmX + h < Cpm2 , where h > 0 is a specified constant. If the test rejects the null hypothesis H0 : CpmX + h > Cpm2 , there is sufficient information to conclude that supplier II is significantly better than supplier I by a magnitude of h , and the replacement would then be made. In this case, however, one would have to compare the test statistic 7i , i = 1,2 , with the selection value c corresponding to the preset capability requirement for a given sample and designated selection power to ensure that the magnitude of the difference between the two suppliers exceeds h . Note that CpmX must be greater than the preset capability requirement, and = Cpm2 Cpmi + h , where h = max{ h! \ test rejects Cpmi + h' > Cpm2}. The basic problem of checking whether or not the two suppliers meeting the preset capability requirement could
230
Encyclopedia and Handbook of Process Capability Indices
be accomplisher by finding the lower confidence bounds on their process capabilities. We recommend consulting Huang and Lee (1995) and Pearn et al. (2004) for further clarifications and details.
Chapter 12
Acceptance Sampling Plans Based on PCIs
12.1 12.2 12.3 12.4 12.5
Introduction Acceptance sampling Acceptance sampling Acceptance sampling Acceptance sampling
plans plans plans plans
based based based based
on on on on
Cpk Cpm Cpmk C pu and Cp[
12.1 Introduction A reader having diligently studied the previous 11 chapters may find some of the material in Chapter 12 to be repetitious. T h e reason is t h a t we strive to have this chapter to be self contained. Beginners a n d / o r practitioners may be inclined to the start reading the book with this chapter. Acceptance sampling plans have been for the last 50 years one of the most practical tools in classical quality control applications, which deal with a quality contracting of product orders between t h e factories and their customers. Such a plan provides the vendor and the buyer with a general criterion for lot sentencing while meeting their preset requirements for the quality of product orders. A well-designed sampling plan can substantially reduce the difference between the required (expected) and the actual supplied product quality. Acceptance sampling plan, however, cannot avoid the risk of accepting unwanted poor product lots, nor can it avoid the risk of rejecting good product lots unless 100% inspection is implemented. Acceptance sampling plan is a statement regarding the required
231
232
Encyclopedia and Handbook of Process Capability Indices
sample size for product inspection and the associated acceptance or rejection criterion for sentencing each individual lot. The criteria used for measuring the performance in an acceptance sampling plan, are usually based the operating characteristic (OC) curve which quantifies the risks for vendors and buyers. The OC curve plots the probability of accepting the lot against actual lot fraction defective, which provides the discriminatory power of the sampling plan. That is, the OC curve shows the probability of accepting a lot submitted with a certain fraction of defectives (nonconformities) which results in the producer and the buyer having a common base for judging whether the sampling plan is appropriate. For product quality protection and company's profit, both the vendor and the buyer would focus on certain points on the OC curve to reflect their benchmarking risks. The vendor (supplier) usually would focus on a specific level of product quality, traditionally referred to as AQL (average quality level), which would yield a high probability for accepting the lot. The AQL also represents the lowest level of quality for the vendor's process that the consumer would consider acceptable in the form of a process average. Therefore, a preferred sampling procedure would be one which gives a high probability of acceptance at the AQL that is usually specified in the contract. The consumer would also focus at another point at the other end of the OC curve, called LTPD (lot tolerance percent defective). The LTPD is the lowest level of quality that the consumer is willing to accept for an individual lot. Note that the LTPD is a level of quality specified by the consumer, corresponding to the specified low probability of accepting a lot with the defect level as high as the LTPD. For completeness we shall now briefly survey the fundamental concepts of Type I and Type II errors. Acceptance sampling plan consists of a sample size ( n ) and an acceptance criterion (C 0 ). Sampling involves risks that the sample will not adequately reflect the totality of quality conditions of the product. Type I error (a)
Acceptance Sampling Plans Based on PCIs
233
is the probability, (for a given (n,C0) sampling plan), of rejecting the product that has the defect level equal to the AQL. The producer suffers when this occurs because here a product with acceptable quality is rejected. Thus, a is called the producer's risk with values commonly ranging from 0.01 to 0.10. In most cases, the values of Type I error a is set somewhere between 0.010.10. The Type II error (/?) is the probability for a given (n,C0) sampling plan accepting the product with defect level equal to LTPD. Here the consumer suffers when this occurs, since a product of unacceptable quality is accepted. Thus, j3 is called the consumer's risk with the values typically ranging also from 0.01 to 0.10. There are a number of papers in the statistical literature providing historical justification for using these values (which are often chosen to be 0.05). While theoretically any two points on the OC curve could be used to construct a sampling plan, it is customary in the industry to use the AQL and LTPD for this purpose. When the levels of lot quality specified are p± = AQL and p2 = LTPD, the corresponding points on the OC curve are usually referred to as the producer's risk point and the consumer's risk point, respectively. Consequently, a well-designed sampling plan must provide the probability of at least 1 — a of accepting a lot if the lot fraction of defectives is at the contracted value AQL. Analogously, the sampling plan must also provide the probability of acceptance no more than (3 if the lot fraction of defectives is at the LTPD level which is an undesired level designated by the buyer. Thus, the acceptance sampling plan must have its OC curve passing through those two designated points (AQL, 1 — a) and (LTPD, {3). There are a number of different ways to classify acceptance-sampling plans. Two major approaches are by attributes or by variables. Assuming that the product quantity N is large, the distribution of the number of defectives d in a random sample of size n of inspected items are Binomially distributed with parameters n and p , where p is the fraction of defectives in the lot. The sample size
234
Encyclopedia and Handbook of Process Capability Indices
n and the acceptance number c of an attribute sampling plan are the solution to the two nonlinear simultaneous equations:
(These equations must be solved simultaneously, for c and n which appear in these equations in the most unexpected positions!) There is no simple direct method for solving them. A graphical procedure presented in Montgomery (2001) can be utilized. Two lines are drawn on a nomograph, one connecting pi and 1 — a , and the other connecting p2 and 0 . The intersection of those two lines gives the region of the nomograph in which the desired sampling plan is located. In addition to the graphical procedure described above, tabular procedures are also available for the same purpose. When a quality characteristic is measurable on a continuous scale and is known to have a distribution of a specified type, it may be possible to use as a substitute for an attributes sampling using instead the based on sample measurements such as the mean and the standard deviation of the sample. These variables sampling plans possess an advantage that the same OC curve can be obtained with a smaller sample than it is required by an attributes plan. The precise measurements required by a variables plan would probably cost more than the simple classification of items associated with by an attributes plan, but the reduction in sample size may offset this constant expense and result in savings. Such savings are especially substantial if the inspection is destructive and moreover the items are expensive (see e.g. Schilling (1982) and Montgomery (2001)).
Acceptance Sampling Plans Based on PCIs
235
We shall now present a few comments of historical nature. Duncan (1986) provides a good description of these available techniques. His book served for many years as a comprehensive reliable compendium. Guenther (1969) has developed a systematic search procedure, which can be used along with Binomial, Hypergeometric, and Poisson tables to obtain the desired acceptance sampling plans. Stephens (1978) provides a closed form solution for single sample acceptance sampling plans using the normal approximation to the Binomial distribution. Hailey (1980) has written a computer program to obtain single sampling plans (with a minimal sample size) based on either the Poisson or Binomial distributions. Hald (1981) gave a systematic exposition of the existing statistical theory for the sampling inspection and provided careful tables for sampling plans which are still quite useful after some 25 years. The basic concept and models of statistically based sampling plans were described in Jennett and Welch as early as (1939). Since 15 years later, Lieberman and Resnikoff (1955) obtained tables and operating characteristic (OC) curves for various AQLs for the popular MIL-STD-414. Owen (1967) considered sampling plans based on the normal distribution, and developed sampling plans for various values of the probabilities of the Type I error when the standard deviation is unknown. Das and Mitra (1964) investigated the effect of nonnormality on the performance of the sampling plans. Bender (1975) considered sampling plans for assuring the percent age of defective in the case of the product quality characteristics obeying a normal distribution with unknown standard deviation, and presented a procedure using a iterative computer program calculating the somewhat cumber some non-central ^-distribution. Govindaraju and Soundararajan (1986) developed sampling plans that match the OC curves of another popular MIL-STD-105D plan. More recently, Suresh and Ramanathan (1997) developed a sampling plan based on a more general symmetric family of distributions.
236
Encyclopedia and Handbook of Process Capability Indices
As the rapid advancement of manufacturing technology manifests itself the customers are starting to demand that the fraction of nonconforming products to be quite low (in parts per million, PPM). In such situations, the required number of inspection items must be quite large. Here the process capability enters into our discussions. Briefly a process capability index is a function of process parameter and manufacturing specifications. It measures the ability of a process to reproduce product units that meet these specifications. Consequently, process capability indices can be used as a quality benchmark for product acceptance. Consider a variables sampling plan to control a lot or a process fraction defective (or nonconforming) items. Since the quality characteristic is a variable, there will exist either an upper specification limit (USL) or a lower specification limit (LSL), or both, that define the acceptable values of this parameter. As indicated earlier, selection of a meaningful critical value for a capability test requires specification of an acceptable quality level (AQL) and a lot tolerance percent defective (LTPD). The AQL is simply a standard against which one judges the lots. It is envisioned that the vendor's process will operate at a fallout level that is considerably better than the AQL. In practice, both the vendor and the buyer will lay down their requirements in a contract: the former demands that not too many "good" lots shall be rejected by the sampling inspection, while the latter requests that not too many "bad" lots will be accepted. A sampling plan will be chosen to reconcile and meet these somewhat opposing requirements. Let (AQL, 1 — a ) and (LTPD, (3 ) be the two points on the OC curve under consideration. To determine whether a given process is capable, we could first consider the following testing hypothesis procedure. H0 : p = AQL (process is capable) Hi : p = LTPD (process is not capable)
Acceptance Sampling Plans Based on PCIs
237
Hence, as we have already alluded above the required inspection sample size n and the critical acceptance value C 0 for the sampling plans are simply the solution to the following two nonlinear simultaneous equations. P {Accepting the lot| proportion defective p = AQL} > 1 — a and P {Accepting the lot| proportion defective p = LTPD} < (3 .
12.2 Acceptance sampling plans based on the Cpk index If the Cpf, index (see Chapter 3) is used as a quality benchmark for acceptance of a product lot. Then the null hypothesis about proportion defectives H0 : p = AQL is equivalent to testing process capability index with •^o : Cpfc > CAQL > where C^QL is the level of acceptable quality for Cpk index corresponding to the lot or the process fraction of nonconformities AQL. For instance, if the proportion of defectives p = AQL of vendor's product is less than 66 NCPPM (nonconforming percentile per million), then the probability of consumer accept the lots will be larger than 100(1 — a)% . On the other hand, if the proportion of defectives of vendor's product, p = L T P D , is more than 2700 NCPPM, then the probability that consumer would accept the lot would not exceed 100/3% . Thus, from the relationship between the index value and the fractions of defectives, we could obtain the equivalent values CAQL = 1-33 and the C L P T D = 1 . 0 0 based on the capability index Cpk . For processes with the target value set at the mid-point of the specification limits (i.e. T = M), the index may be rewritten as: Cpk = (d I a— \ £ | ) / 3 , where £ = (/J, — M) / a (see Chapter 3). As noted earlier, the sampling distribution of Cpk is expressed in terms of a mixture of the chi-square and a normal distributions. Given Cpk = C , the ratio b = d j a can be expressed as b = 3C+ | £ |. Thus, the probability of accepting a product can be expressed as:
238
Encyclopedia and Handbook of Process Capability Indices
•XA(Cpk) = P{Cpk
> C0 |
Cpk)
-L
2
9nco x (^(t + §Vn) + ^ ( « - § V n ) ) d t .
(12.1)
(c/. Chapter 3 Equation (3.8)). Recall that G(-) is the c.d.f. of chisquared distribution and (/>(•) is the standard normal density. The required inspection sample size n and the critical value Co of Cpk for the sampling plan are obtained as the solution to the following two nonlinear simultaneous equations: >6i -Jn
Jo
G '{n-l){b\Jn
-if
9nco
x {
and
P>
n
rG Jo
(12.2)
'(n - l)(&2Vn - t)2 "i 9nco
x (
c\Jn))dt,
(12.3)
where ^ = 3CAQL + I ^ I a n d h = 3CLTPD + I £ I > noting that CAQL > CLTPD • Observe that the required sample size n is the smallest possible value of n satisfying equations (12.2) and (12.3), and we shall use [n] as the sample size, where mi denotes the least integer greater than or equal to n . However, since the process parameters fj, and a are unknown, so is the distribution characteristic parameter, £ = (// — M) /a , which has to be estimated in real-world applications. As it has been already mentioned above such approach introduces additional sampling errors from estimating £ in finding the critical acceptance values and the required sample sizes. To eliminate the need for this additional estimation procedure of £, Pearn and Wu
Acceptance Sampling Plans Based on PCIs
239
(2004) carried out extensive calculations and have obtained the critical acceptance values CQ and the sample size n for £ = 0(0.05)3.00, for various values of CAQL a n d CLTPD • Observe that the parameter values £ = 0(0.05)3.00, cover a wide range of applications with process capability Cpk > 0 and provides the same values whenever £ is replaced by — £ . The results of the calculations show that the critical value Cg is: (i) increasing in £, (ii) reaches its maximum at £ = 1.00 in all cases, and (hi) remains the same for £ > 1.00 for all CAQL (with the accuracy up to 10~ 6 ). The critical acceptance value c0 reaches its maximum at £ = 0.50 and remains unchanged for £ > 0.50 as the sample size n exceeds 30 (and for n > 100 , the maximum is at £ = 0.35 with the accuracy up to 10~ 4 ). Hence, one could solve equations (12.2) and (12.3) with £ = 1.00 to obtain the critical acceptance value and the sample size n without having to estimate the parameter £. This approach ensures that the decisions which are carried out based on these critical acceptance values may be more reliable. To solve the two nonlinear simultaneous equations (12.2) and (12.3) given above, let /•»<wn
Si(n>Cb)= / *J 0
'(n-l)(6iVn-i)2' G 9nco
x (cj)(t + ^Vn) + <j>(t - Z,Jn))dt - (1 - a) (12.4) and 5 2 (n,c 0 )= /
G
'(n-l)(6j2Vra - £ ) 2 ' 9rcco
x ((t>(t + ^^n) + (f)(t-^^n))dt-p
•
(12-5)
Figures 12.1-12.2 display the surface and contour plots of the equations (12.4) and (12.5) jointly under CAQL = 1-33 and
240
Encyclopedia and Handbook of Process Capability Indices
=1-00 with the a-risk = 0.05 and the /?-risk = 0.10, respectively. From Figure 12.2, one observes that the interaction of ^ ( n , c0) and S2(n,c0) contour curves at level 0 is (n,^) = (61.74, 1.1477), which is evidently also the solution to the nonlinear simultaneous equations (12.2) and (12.3). Namely, in this case, the minimum required sample size rni equals to 62 and the critical acceptance value c0 is 1.1477 for a sampling plan based on the capability index Cpk . We observe in conclusion that in practice to bring the value of the PCI Cpfc to acceptable level is to design equipment and develop a process in which the process variable is robust to external disturbances or alternatively to use in situ sensors and alternative feedback control. See El-Awady et al. (1996) for a detailed and lucid discussion. CLTPD
50
Figure 12.1. Surface plot of S\ andS^ .
100
110
n Figure 12.2. Contour plot of S\ and£2 •
12.3 Acceptance sampling plans based on Cpm As described above, due to the increasing importance in applications of clustering around the target, rather than adherence to specification limits, and a deeper understanding of loss functions there appears to be a necessity for an alternative to PCIs Cp and Cpk . The Cpm index discussed in Chapter 4 can be used as a quality benchmark for acceptance of a productions lot. Recall
Acceptance
Sampling Plans Based on PCIs
241
the definition of this index as given in Chapter 4, and remark that the index could be rewritten as Cpm = d /[3a x (1 + £ 2 ) x ' 2 ], where, as usual, £ = (/J, — T) / a . Hence, given Cpm = C , the ratio b = d J a can be expressed as b = 3C(1 + £2)*' 2 . The probability of accepting the lot is then becomes ^A{Cpm) — P(Cpm
> C 0 I Cpm — C)
6Vn/(3C 0 ) 0
G
(b2n 9Cn2
[<j>(t + £Vn) + (j)(t - £Vn)]d£-
(12.6)
The required inspection sample size n and the critical acceptance value C0 of Cpm (the natural estimator of Cpm , see Equation (4.5)b) for a given sampling plan will be the solution to the following two nonlinear simultaneous equations (12.7) and (12.8): &,Vn/(3C„)
1-a
<
bin
i:
t2
9Cn2
x [cj)(t + £Vn) + <j)(t -
and
&2Vn/(3C0) J 0
' ljn
G
£~Jn))dt
(12.7)
2
9Cn2
x [<j)(t + £Vn) + (f>(t -
$,Vn)]dt,
(12.8)
where b, = 3C AQL (1 + £ 2 ) 1 / 2 and 62 = 3C LTPD (1 + £ 2 ) 1 / 2 • (Observe that CAQL > CLTPD •) I n order to solve the above two nonlinear simultaneous equations (as above), we set
Si(n,CQ)=
f
G
bfn _
t2
9Cn2
x [
(12.9)
242
and
Encyclopedia and Handbook of Process Capability Indices
s2M)
f^G[^-S
= x
[4>{t + £Vn) + <j>{t - £Vn)]dt
- (3 • (12-10)
Figures 12.3-12.4 display the surface and the contour plots of equations (12.9) and (12.10) jointly for C A Q L = 1.33 and CLPTD =1-00 with both the a and the /3 -risks = 0.05. From Figure 12.4, one deduces observe that the interaction of the Si(n,CQ) and the S2(n,C0) contour curves at the level 0 is (n,C0)= (68, 1.1668), which serves as the solution to nonlinear simultaneous equations (12.7) and (12.8). Namely, in this case, the minimal required sample size n is 68 and critical acceptance value C0 become 1.1668 for the sampling plan which is based on the "new" capability index Cpm.
50
Figure 12.3. Surface plot of Si a n d ^ .
„
100
1J0
Figure 12.4. Contour plot of Si and 52 •
The sample size n and the critical acceptance value C0 based on Cpm for sampling plans with various CAQL , CLPTD > a -risks and /?-risk are provided in Pearn and Wu (2004). One observes that the larger are the risks (a and/or /?) which producer or customer could tolerate, the smaller will be the required sample size n. This phenomenon has an intuitive explanation, since if we
Acceptance Sampling Plans Based on PCIs
243
insist that the chance of wrongly classifying a deficient process as good one (or specifying good lots as inadequate ones) be smaller then more sample information is needed to asses the lots. Furthermore, for fixed a-risk, CAQL and C LPTD > the corresponding critical acceptance values become smaller as the /?-risk increases. On the other hand, for fixed (3 risk, CAQL a n ( i C'LPTD > the corresponding critical acceptance values become larger as the a-risk increases. This later fact can also be explained by the same reasoning as above. Also the required sample size becomes smaller as the difference between the CAQL a n d CLPTD increases since such a decision will be more appropriate.
12.4 Acceptance sampling plans based on Cpmk We have already emphasized above that although Cp^ contains the most of the information concerning the yield, it, unfortunately provides the least insight about the location of the parameter fj, . Conversely, the index Cpmk (Chapter 6) provides the most information about the location of // but the least amount concerning the yield. Although Cp^ index is still serves a the more popular and widely used index on the floor, the index Cpmk is now increasing is applicability as to be a refined measure and a more sensitive index for processes with two-sided specification limits. For the processes with target values set at the mid-point of the specification limits, this index may be rewritten as Cpmk=(d/a-\t\)/[3(l
+
e)1/2},
where, as usual, £ = (/j, — T) / a (see the definition of Cpmjt index in Chapter 6, Equation (6.1)). Furthermore, given Cpmi, = C , the ratio b = d / a can be expressed as b = 3C(1 + £ 2 ) 1 ' 2 + | £ |. The probability of accepting a lot is consequently can be expressed as
244
Encyclopedia
and Handbook of Process Capability
Indices
KAyCpmk) — P{Cpmk > C0 | Cpmk • 6Vn/(l+3x) J 00
G
[(bJn-tf 9x2
t2)
x [4>{t + (Vn) + (f>(t -
£Jn)]dt.
(12.11)
where, as above, Cpmk is the "natural" estimator of Cpmk . It thus follows that the required inspection sample size n and the critical acceptance value C0 of Cpmk for acceptance sampling plans are expressed as the solution to the following two nonlinear simultaneous equations (12.12) and (12.13): biJn/(l+3x)
1
o
G
{{b^n-tf 9z2
2)
x [(f>(t + £Vn) + cp(t - £-Jn)]dt 62Vn/(l+3x)
and
G
J0
{{bJn-tf
(12.12)
t2)
2
9a;
[4>{t + £Vn) + <j>(t -
£-Jn)]dt,
(12.13)
where h = 3<7AQL(1 + £ 2 ) 1 / 2 + | £ I and fe = 3C LTPD (1 + ^ 2 ) 1 / 2 + | £ | (noting that CAQL > CLTPD)- To solve these two nonlinear simultaneous equations, we set as above &!Vn/(l+3x) 0
((bVn-Q2 G 9a;2
f2)
x [(j)(t + £Vn) + cj)(t - £Sn)]dt - (I - a)
h^fn/(l+3x) t w n ni+ix)
/ 0
G
{{b^n-tf
(12.14)
^
2
9a;
x [0(t + £V7i) + 0(i - i^n)\dt
- p.
(12.15)
Acceptance Sampling Plans Based on PCIs
245
Figures 12.5-12.6 display the surface and contour plots of the equations (12.14) and (12.15) jointly for CAQL = 1-33 and CLPTD =1-00 with the a-risk = 0.10 and the /?-risk = 0.05, respectively. From Figure 12.6 one concludes that the interaction of S^n^o) and S2(n,C0) contour curves at level 0 is (n,C0) = (82, 1.1870), which is therefore the solution of the nonlinear simultaneous equations (12.12) and (12.13). Namely, in this case, the minimal required sample size n is 82 and the critical acceptance value C0 equals 1.1870 for the sampling plan based on the capability index Cpm),.
Figure 12.5. Surface plot of 5! andS 2 •
Figure 12.6. Contour plot of 5V and 5;
Pearn and Wu (2004) investigate the behavior of the critical acceptance values and the necessary sample sizes for various parameters by solving simultaneously equations (12.12) and (12.13). The results indicate that the larger is the risks that producer or customer could tolerate, the smaller is the required sample size n. As in the situation with the Cp\. index this allows an intuitive explanation, since if we expect that the chance of incorrectly assenting that a defective process satisfactory or classifying good lots as inadequate ones is smaller, the more sample information will be require to decide about the quality of
246
Encyclopedia and Handbook of Process Capability Indices
the lots under inspection. Furthermore, for fixed a , j3 risks and CLPTD I the required sample sizes decrease as CAQL increases.
12.5 Acceptance sampling plans based on Cpuand Cpi For processes with one-sided specification limits, the indices C/ (Cpu o r CPL) c a n D e applied in similar manner (see Chapter 8 Section 1). Since the sampling distribution of Cj is given as a multiple of a non-central t distribution, specifically as 6n_1(3Vn)~1 tn_i(6) , where £„_i(<5) is a non-central t distribution with n — 1 degrees of freedom and the non-centrality parameter 6 = 3VnC/ , the probability of accepting a lot can thus be expressed as: *A(CI)
= P{Cj >C0\d=C)=P
t^S)
>
3VnC0
(12.16)
"n-l
Therefore, the required inspection sample size n and the critical acceptance value C0 of Cj (a natural estimator of Cj) for the sampling plan under consideration are the solution to the following two nonlinear simultaneous equations (12.17) and (12.18): P and
P
3VnC 0 ' K-i , 3VnC 0 )
K-\
>\-CL
(12.17)
(12.18)
where ^ = 3Vn CAQL and 62 = 3Vn C LTPD , with CAQL > CLPTD • The required sample size n is the smallest possible value of n satisfying equations (12.17) and (12.18), and we use the rnl being least integer greater than or equal to n. To illustrate the manner in which the above two nonlinear simultaneous equations, are solved, we set:
Acceptance Sampling Plans Based on PCIs
2—F[n^l]J
247
0 3VnCox
V2IF 1
exp - ^ - < y
0
POO
„ Z±.
2Xpptl ^ 0
1
,-
2
ducfc - a
(12.19)
3VnCn:r "—Q"
V27T J 0
xexp - - ( u - < 5 2 ) 2 dvdx
-{1-15).
(12.20)
Figures 12.7 and 12.8 display the surface and contour plots of equations (12.19) and (12.20) jointly with the a-risk = 0.10 and the /?-risk = 0.10 for C A Q L = 1.50 and C LPTD = 1.00, respectively. From the Figure 12.8, one concludes that the interaction of 5i(n,C 0 ) and 5 2 (n,Co) contour curves at level 0 is (n,C0) = (24.49, 1.2200), which is also the solution to nonlinear simultaneous equations (12.17) and (12.18). Namely, in this case, the minimal required sample size is [n] = 25 , and the critical acceptance value C0 equals 1.2200 for the sampling plan based on the one-sided capability index Cj . 1
1
!
,.
1.3-
C„
"-^, —--i»
. -
-
-
•
-"=4^-__m—1—
Figure 12.7. Surface plot of Si andSj,-
—
•
-
"
M^-
Figure 12.8. Contour plot of Si andS 2
248
Encyclopedia and Handbook of Process Capability Indices
An acceptance sampling plan in principle consists of a sample size (n) and an acceptance criterion (CQ). Since the sampling cannot guarantee that every defective item in a lot will be inspected, the sampling involves risks of not adequately reflecting the quality conditions of a lot. Risks are becoming to be more significant as rapid advancements in the manufacturing technology occur and a stringent customers demand is enforced. In particular, when the fraction of defectives is very low, more rigorously measured in parts per million (PPM), the required number of inspection items could be exceedingly large in order to reflect adequately the actual lot quality. To summarize process capability indices are function of process parameter and manufacturing specifications. It measures of the ability of a process to (re)produce product units that we meet the specifications. Consequently, utilizing these acceptance sampling plans based on PCIs, practitioners all able to determine the number of inspection units required assess the critical acceptance value, and to make more sound decisions.
Chapter 13
Process Capability Measures in Presence of Gauge Measurement Errors
13.1. 13.2. 13.3. 13.4. 13.5. 13.6.
Introduction Estimating and testing Cp in presence of gauge measurement errors Estimating and testing Cpk in presence of gauge measurement errors Estimating and testing Cpm in presence of gauge measurement errors Estimating and testing Cpmk in presence of gauge measurement errors Estimating and testing Cpuand Cpl in presence of gauge measurement errors
13.1 Introduction
T h e inevitable variations in process measurements comes from two sources: the manufacturing process and the gauge. Gauge capability reflects the gauge's precision, or lack of variation, but is not the same as calibration, which assures the gauge's accuracy. As it was emphasized on the numerous occasions, process capability measures the ability of a manufacturing process to meet preassigned specifications. Nowadays, many customers use process capability to judge supplier's ability to deliver quality products. Suppliers ought to be aware of how gauges affect various process capability estimates. The gauge capability includes two parts: repeatability and reproducibility. Repeatability, (actually the lack of it), is the gauge's experimental or random error. This means t h a t , when
249
250
Encyclopedia
and Handbook
of Process
Capability
Indices
measuring the same specimen several times, the gauge will never return exactly the same measurement. Reproducibility (actually the lack of it) is the (often annoying) inability of several inspectors or gauges to arrive at the same measurement from a given specimen, i.e. the variability due to different operators using the gauge (or different time periods, or different environments, or in general, different conditions). To summarize we have: 2 "Measurement error
2 "Gauge
2 ""repeatability
2 > ^reproducibility •
Estimates for ^repeatability a n d "reproducibility come from a gauge study, or an GR&R (gauge repeatability and reproducibility) study. Hradesky (1988), Barrentine (1991), Levinson (1995) and Montgomery (1997, 2001), among others, provide procedures for gauge studies. Gauge capability is a gauge's ability to repeat and reproduce measurements. Its measurement is the % of tolerance consumed by (gauge) capability (PTCC). Montgomery (1991) refers to it as the precision-to-tolerance (or P / T ) ratio. It is the ratio of the gauge's variation to the specification width; its smaller numbers are of course preferable. Denoting the gauge's standard deviation by °Gauge ,
w e
n a v e :
PTCC
=
usT^fsLxlmy-
Some authors and practitioners use the coefficient 5.15 instead of 6 (see e.g. Barrentine (1991) and Levinson (1995, 1996)). This formula uses 6 a as the natural tolerance width for the gauge based and motivated by the normal distribution assumptions. A substantial majority of appeared in the literatures do measurement errors. However significant effect on process inaccurate measurement system
capability research works that not take into account gauge the gauge capability has a capability measurement. An can thwart all the benefits of
Process Capability Measures in Presence of Gauge Measurement Errors
251
improvement endeavors resulting in poor quality. Analyzing process capability without considering gauge capability may often lead to unreliable decisions. It could result in a serious loss to producers if gauge capability is not being considered in process capability estimation and testing. On the other hand, improving the gauge measurements and properly trained operators can reduce the measurement errors. Since measurement errors unfortunately can not be avoided, using appropriate confidence coefficients and power becomes an essential task. However, the reality is that no measurement is free from error or uncertainty even if it is carried out with the aid of highly sophisticated and precise measuring instruments. Montgomery and Runger (1993) point out that the quality of the data related to the process characteristics relies substantially on the gauge. Any variation in the measurement process has a direct impact on the ability to execute sound judgment about the manufacturing process. Conclusions about capability of a process based only on the single numerical value of the index are not reliable. Analyzing the effects of measurement errors on PCIs, Mittag (1993) and Levinson (1995) developed very definitive techniques for quantifying the percentage error in process capability indices estimation in the presence of measurement errors. Gauge repeatability and reproducibility (GR&R) studies focus on quantifying the measurement errors. Common approaches to GR&R studies, such as the Range method (Montgomery and Runger (1993)) and the ANOVA method (Mandel (1972), Montgomery and Runger (1993)) assume that the distribution of the measurement errors is normal with a mean error of zero. Let the measurement errors be described by a random variable M ~ N(0, a2M) , Montgomery and Runger (1993) present the gauge capability A by means of the formula:
X
= UsBlSL*m%-
<'«>
252
Encyclopedia and Handbook of Process Capability Indices
For the measurement system to be deemed acceptable, the variability in the measurements due to the measurement system ought to be less than a predetermined percentage of the engineering tolerance. The Automotive Industry Action Group recommends the following guidelines for gauge acceptance presented in Table 13.1 right below. Table 13.1. Guidelines for gauge capabilities Gauge Capability A <10% 10%< A <30%
30%< A
Result Gauge system is O.K. May be acceptable based on the importance of applications, cost of the gauge, cost of repairs, and so on. Gauge system needs improvement; make every effort to identify the problems and have them corrected.
In the Sections 13.2-13.6, we shall survey estimation and testing procedures for various PCIs in presence of gauge measurement errors. 13.2 Estimating and testing Cp in presence of gauge measurement errors Denote by X~N(fi,a2) the relevant quality characteristic of a manufacturing process and consider this process, capability in the measurement error system. Because of measurement errors, the observed random variable Y ~ N(fiy = /i, oY =
Process Capability Measures in Presence of Gauge Measurement Errors
C¥ = • ° \
2-
253
(13.2)
Since the variation of the observed data is larger than the variation of the original one, the denominator of the index Cp = (USL — LSL)/(6a) becomes larger, and the variable Y will underestimate true capability of the process. Pearn and Liao (2004a) listed a number process capabilities with A = 0.05(0.05)0.50 for various true process capability indices Cp = 0.50, 1.00, 1.33, 1.50, 1.67, 2.00, and 2.50. It is obvious that the gauge becomes more crucial as the true capability improves (Levinson (1995)). If A = 0.50 (50%), CYP = 0.49 with the true process capability is Cp = 0.50, while Cp — 1.56 when the true process capability being Cp = 2.50. Utilizing a "perfect" measuring instrument will aid substantially in the case of processes with higher capability. Suppose that { Xt, i — 1, 2, • • -, n } denote the random sample of size n from the quality characteristics X . To estimate the precision index Cp , the natural estimator CP = (USL — LSL) / 6s is used, where S = [ZliWCXj — X)/(n — 1)]1//2 is the conventional estimator of a , which should be obtained from a stable process. Adding to eliminate bias the well-known correction factor bn^i = [ 2 / ( n - l ) ] 1 / 2 r [ ( n - l ) / 2 ] / r [ ( n - 2 ) / 2 ] to CP , and denoting CP = bn_i Cp (see Chapter 1), it was shown by Pearn et al. (1998) that Cp is the UMVUE (uniformly minimum variance unbiased estimator) of CP. However, the sample observations are not { Xi, i = 1, 2,- ••, n } but we actually { Y{ , i = 1, 2, • ••, n } . We thus obtain an estimator of CP to be -Y ^P - A - H
USL - LSL r^
)•
Note that Cp estimates Cp , using Sy = 1/2 [E"=i(li - F ) / ( n - l ) ] . Utilizing the arguments of _Chou and Owen (1989) and Pearn et al. (1998), the p.d.f. of Cp can be expressed as
254
Encyclopedia and Handbook of Process Capability Indices
f(y) = 2
( V ( n - l ) / 2 C P / V l + A 2 C|)"- 1 (yTn T[(n-l)/2]
X exp
-(n-l)CP(2y2y 2 1 + X2nlCp
The expected value and the variance of the estimator Cp are easily obtained using E(Xn-i) a n d Var(Xn-i) to be: CP
E(CYP) =
yfl + ^C]
, Var{CYP) = ( ^ - U' n2- l - 1 ) n —3
cl
z 1 + X2r<2 Cp
For A > 0, it is obviously that Cp is a biased estimator of CP = (USL - LSL) 16a , and the bias is (1/Jl +\2Cp -l)CP which decreases as A increases. Since n is a finite positive integer, [(n — l ) / ( n — 3)](&n_x)2 — 1 is positive (recall that bn_i > 1 ), we have Var(Cp) < Var{CP). The reader should differentiate between CP , CP , (7P , C^ and (7^ . The mean square errors of the two estimators CP = bn^Cp and Cp are MSE(CP) = MSE{Cl)
n-l n
bn-i
i c:p' 6'?n - l
l 1 + \2n2 C
Vl + A2C|
G„
To determine whether a process under consideration meets the present capability requirement and runs according to the desired quality condition, statistical testing hypothesis can be employed by setting H0: CP < c versus Ex:CP>c. A process fails to meet the capability requirements if CP < c , and meets then for CP > c. The critical value c0 can be determined with a given a -risk (the chance of incorrectly judging an incapable process as capable). In the presence of measurement errors, the true capability of the process will unfortunately be underestimated if one calculates
Process Capability Measures in Presence of Gauge Measurement Errors
255
capability index using the estimator Cp instead of Cp. Thus, the a -risk using Cp to estimate CP is less than the a -risk using CP to estimate the same parameter, and the power using Cp to estimate Cp is likewise less than the power using CP as an estimator of Cp . Pearn and Liao (2004a) indicate that both the a-risk and the power of the test decrease when the gauge measurement error increases. If the producers do not take into account the effects of the gauge capability on process capability estimation and testing, it may cause substantial and serious loss. In that case producers are unable affirm anymore that their processes will meet the capability requirements even if they are indeed sufficiently capable. The producers may incur substantial costs as (large) quantities of acceptable product units will be incorrectly rejected. Improving the gauge measurement accuracy and training the operators to proper by execute the production procedures at each step are essential for reducing the measurement errors. Even so the stubborn measurement errors may be unavoidable in many manufacturing processes. Pearn and Liao (2004a) adjusted the confidence intervals and the critical values in order to ensure the intervals possess the desired confidence coefficients and will improve the power of the test with an appropriate «-risk. For the desired confidence coefficient 1 — a , the adjusted confidence interval of CP , with the confidence interval bounds L* and U*, turns out to be: P{L*
= P L* < = P
r*2
*2
C„
V(n - l)bU{KYY'
- {\CYvf
(n-l^i^r 1 -^) 2 (cYp"i2 (n - ! $ _ ! ( # {cYY\1 v)
Y\2 {\CD
< 1
256
Encyclopedia
= P
and Handbook
r*1
(n -
of Process
Y l)bU{KY\~l )
+ P 1 + £T2A2 < U*2
Indices
< 1 + L** Z2X\ 2
Y\2
K)
Capability
(n-l)bU(KY) (cYPf
Y\-l
,2
= P
(«- m-x2 Y
r*2
(C f(l
+ L* X2)
(n - 1K2_! (C ) (l + U*2\2)
< U *2
=
Y 2
1-a.
(13.3)
Expression (13.3) can be verified by carrying out appropriate substitutions. Here KY = (n — l)Sy / Oy is distributed as x\-\ • Further simplifications show that the adjusted 100(1 — a)% confidence interval bounds can be written as L* =
y Xn-i,i-a
/ 2^;
^n-l^U-iXClfxl^
a/2
V^-l,a/20f
U*
J(n - \)bU - (\CYfxUa/2
where X^_n_ a /2 ^s the 1 — <*/2 percentile of the chi-squre distribution with n-1 degrees of freedom. With the adjusted confidence interval bounds, the desired confidence coefficient can be achieved. Moreover, to improve the power of the test, Pearn and Liao (2004a) obtain adjusted critical values (denoted by CQ ) to be Y a = P{C >c*0\cp = c) = p
= p K-i-Jn ~
1C
2
\_i>/n
VF
,S- > >IKY I C = n
{ co*Vl + X C:
-ICY
> co* I Cp = c
Process Capability Measures in Presence of Gauge Measurement Errors
2
2
c >K* = P ( t f2 _ i ( n - l )2„2\ U ( l + A2c2) —
257
^ &g-i(n - l ) c 2 '
v = P Xn — 1 — *2/-i , \2 2\ co (1 + A V ) ,
Note that the last probability statement involves both c and CQ . To ensure that the a;-risk is within the preset magnitude, we set a* = a , and from the last probability statement we have 6n-i>/n - \c Co
=
V(l + A 2 c 2 )x2-i,i-a '
and the power of the test (denoted by IT* ) can be calculated as K*{CP) = P{CYP >4\CP)
= P
= P
{bn_xyJn
VF
'CPN/(I + A2C2)X2-I,I^ 2
Xn-1 <
> co I Cp
_pgr
2
Cyjl + X C
Cp
-lCp
p
2„2 \ 1 + X'c
1 + A2C2
Xn—1,1—a
The reader is advised the compare the expressions for the adjusted critical values and the power. 13.3 Estimating and testing Cpk in presence of gauge measurement errors The well-known process capability index Cpfc (discussed in some detail in Chapter 3) has been widely used in the manufacturing industry for measuring process quality reproduction capability. Most research works related to Cpk are carried out under the assumption of no gauge measurement errors. Unfortunately, such an assumption does not accommodate adequately real world situations even with the modern highly sophisticated measuring instruments and devices. Let
258
Encyclopedia and Handbook of Process Capability Indices
X ~ N(n,cr2) represent the relevant quality characteristic of a manufacturing process and Cpk measures the "true" process capability. However in practice, we deal with the observed variable Y (the measurement error) rather than with the "true" variable X . Assume that X and M (the measurement error) are stochastically independent. In this case we have Y ~ N(HY = M) °Y — v2 + 0M) , and the empirical process capability index Cpk will be obtained after substituting aY for a . The relationship between the true process capability Cpk = min j (USL — /x) / 3
pk
_ d— I Y — m \ 3Sy
obtained by replacing the process mean /i and the process standard deviation a by their conventional estimators Y = E"=i Yi/n and SY = [E?=i(*i ~ Y) /(n - 1)] 1 / 2 , from a "bona fide" stable process. Accordingly, applying the technique used in Kotz and Johnson (1993) and Pearn and Lin (2003), the c.d.f. of CYpk is obtained as:
Process Capability Measures in Presence of Gauge Measurement Errors
•ACl-Sn F*Y(X) C
P*V
= 1 '
\ JO
G
259
(n-l)(3CjVn-i) 2 ^ 9nx2
x $[* + 3(CY - C^kyjfi] + *[t - 3(<# - C p K ,)Vn]^ , where CYp = Cp / Jl + \2C2p , is CYvk = Cpk / yjl + X2C2p , $(•) is the c.d.f. of standard normal distribution and, as above, A = 6aM /(USL - LSL) x 100% . To determine if a given process meets the preset capability requirement, we could consider the statistical testing with the null hypothesis H0 : CPK < c (the process is not capable) and alternative hypothesis H0 : Cp% > c (the process is capable), where c is the required critical value. If the calculated process capability is greater than the corresponding critical value, we reject the null hypothesis and conclude that the process is capable. Discussions in Pearn and Liao (2004b) indicated that the true process capability would be severely underestimated if Cpk is used. The probability that C\\ is greater than c0 would be less than when using Cpk . Thus, the a-risk using Cpk is less than the a -risk of using Cpk when estimating Cpk . The power of the test based on CYk is also less than that based on Cpk . That is, the a risk and the power of the test decrease with the measurement error (as in the case of the Cp index (Section 13.1)). Since the lower confidence bound is severely underestimated and the power becomes small, the producers cannot firmly state that their processes meet the capability requirement even if their processes are indeed sufficiently capable. Adequate and even superior product units could be incorrectly rejected in this case. To remedy the situation, Pearn and Liao (2004b) consider the adjustment of the confidence bounds and of the critical values to provide a superior capability assessment. Suppose that the desired confidence coefficient is 9, the adjusted confidence interval of Cpk with the lower confidence bound L*, can be shown to be:
260
Encyclopedia and Handbook of Process Capability Indices
e =
p(cpk>L*) G 0
(n - l)(6*Vn - i),2: 1 9n(Cj,) 2
x [$(* + £ y Vn) +$(t - iYJn)\
dt,
where b* = 3L* /Jl + \2C2p + | £Y \ , and £ y = 3(
18L* + yj324L*2 - 4(9 - X2)(9L*2 - 1) 2(9 - A2)
= C.pi
To improve the power of the test, revised critical values CQ can be determined from a* = P(CYpk > 4 I Cpk = c) nzc* vn
~ Jo
((n - 1) {3CY Vn - tf )
{
9n(c0*)2
JX
X $[* + 3(CY - CYk)Vn) + $[< - 3(C y - C^)V7T]^ , where a* is the a -risk corresponding to the test using Cpk and CQ , CYp = Cp/Jl + \2C\ , and CYpk = c/^l + \2C2p and c is the capability requirement. To eliminate the need for further estimation of the characteristic parameter Cp , we set Cp = CYk + 1/3 as suggested by Pearn and Lin (2003) and find a reliable adjusted critical values CQ , where Cp can be obtained by using the equation Cp / yjl + \2C2p = c/^l + X2C2p + 1/3 to be:
Process Capability Measures in Presence of Gauge Measurement, Errors
Cp
_ 18c + V324c2 - 4(9 - A2)(9c2 - 1) _ ~ 2(9^) -
G
261
^
To ensure that the a -risk is within the preset magnitude, we let a* = a and solve the above equation to obtain CQ . The revised power (denoted by IT* ) can be calculated as 7r
*(Cpfc) = P\Cpk > c0 | Cpk,Cp = Cp2) 3CT>/n
= /. o
'(n-i)(3C^Vn-*)2' [$(£ + Vn) + $ ( i - V n ) ] d t . G 9n(c0*)2
2 where C^ = C^fc + 1/3, and C ^ = Cpk / Jl + X2C p2
When estimating the capability, the estimator C ^ in the case of contaminated data, severely underestimates the true capability in the presence of measurement errors. Consequently, if a statistical test is used to determine whether the process meets the capability requirement, the power of the test would decrease substantially. Consequently, lower confidence bounds and critical values ought to be adjusted to improve the accuracy of capability assessment. Example 13.1 The LM 2576 series of regulators, as depicted in Figures 13.1(a)-(b), are monolithic integrated circuits, which provide active functions for a step-down (buck) switching regulator, capable of driving 3A load with excellent line and load regulation. Those devices are available in fixed output voltages of 3.3V, 5V, 12V, 15V and an adjustable output version. Requiring a minimum number of external components, those regulators are simple to use and they include internal frequency compensation and a fixed-frequency oscillator. The LM 2576 series offers a highefficiency replacement for popular three-terminal linear regulators. It substantially reduces the size of the heat sink, and in some cases
262
Encyclopedia and Handbook of Process Capability Indices
no heat sink is required. A standard series of inductors optimized for use with the LM 2576 are available from several different manufacturers. This feature greatly simplifies the design of switchmode power supplies. Other features include a guaranteed ±4% tolerance on output voltage within specified input voltages and output load conditions, and ±10% tolerance on the oscillator frequency. External shutdown is included, featuring a (typical) 50 /J, A standby current. The output switch includes cycle-by-cycle current limiting, as well as a thermal shutdown for full protection under fault conditions.
Top View
0
0
TT™i5-cs/CFF "1 :c::::u-fadbaek 1 1 "13-OfciufliJ " E l 2-Output 31-Vh,
Figure 13.1 (a). LM2576 series stepdown voltage regulator product (top
Side View ^ = 3 PIUS 1 , 3 , 4 3
Figure 13.1 (b). LM2576 series stepdown voltage regulator product (side
Following this (positive) description of the LM 2576 series of regulators, Consider a supplier manufacturing step-down voltage regulator products in Taiwan, the producing LM 2576-3.3 type with specifications of output voltage: T = 3.3V, USL = 3.366V, LSL = 3.234V for conditions of VIN = 12V (input voltage), ILOAD — 0.5A (load current), and Tj = 25°C (temperature). A total of 70 observations are collected and displayed in Table 13.2. histogram and a normal probability plot show that the collected data follows the normal distribution. The Shapiro-Wilk test is applied to further justify this assumption. (The reader is urged to carried out the normal probability plot and the Shapiro-Wilk test). Now to determine whether the process is "excellent" (Cpfc > 1.50) with the unavoidable measurement errors A = 0.25, we first determine that c = 1.50 and a = 0.05. Then, based on the
Process Capability Measures in Presence of Gauge Measurement Errors 263
sample data of 70 observations, we obtain the sample mean Y = 3.299, the sample standard deviation SY = 0.013, and the point estimator Cjfc = 1.632. From Table 4(c), we obtain the critical value Cfl = 1.595 based on a , A and n . Since Cpk > CQ , we conclude that the process is "excellent". Moreover, by inputting Cpk , A , n , and the desired confidence coefficient 0 = 0.95 into the appropriate computer program we obtain the 95% lower confidence bound on the true process capability to be 1.542. We can see that if we ignore the measurement errors and evaluate the critical value without any correction, the critical value will be calculated to be as CQ — 1.758. In this case we would reject that the process is "excellent" since Cpk is not the estimator greater than the uncorrected critical value 1.758. Table 13.2 observations for output voltage (unit: V) 3.292
3.289
3.293
3.323
3.279
3.304
3.306
3.288
3.287
3.319
3.299
3.269
3.300
3.294
3.308
3.285
3.292
3.278
3.285
3.282
3.297
3.278
3.311
3.295
3.319
3.303
3.305
3.326
3.315
3.298
3.321
3.315
3.284
3.319
3.302
3.314
3.308
3.303
3.294
3.312
3.297
3.305
3.306
3.295
3.286
3.293
3.288
3.314
3.318
3.295
3.309
3.296
3.296
3.305
3.293
3.298
3.305
3.289
3.288
3.315
3.308
3.279
3.292
3.293
3.265
3.283
3.307
3.314
3.303
3.305
Estimating and testing measurement errors
Cpm
13.4
in
presence
of
gauge
We have already mentioned that most of research works (both theoretical and applied) related to Cpm have been carried out under the assumption that no gauge measurement errors are present. See Marcucci and Beazley (1988), Boyles (1991), Pearn et al. (1992), Zimmer and Hubele (1997), Zimmer et al. (2001), and Pearn and Shu (2003) among other publication. Such an assumption, however, does not accommodate satisfactorily realworld situations by even using the advanced top-line measuring
264
Encyclopedia and Handbook of Process Capability Indices
instruments. Quite likely conclusions concerning the process capability that are based on empirical indices are not fully reliable. The relationship between the true process capability Cpm and the empirical process capability C^pmm IisD
£>L =
^1 + ?
(i3 4) l j
cpm vr+^cj+F' where £ = ( / i - T ) / c r .
Since the variation of the observed data is larger than the variation of the original one, the denominator of the index Cpm = (USL - LSL)/^a2 + (/i - T) 2 ] increases, and the "true" capability of the process is therefore understated if the calculations of the process capability index are based on the empirical data Y . In practice, sample data must be available to estimate the empirical process capability Cpm . For a stably normally distributed process, the empirical data (namely the observed measurements contaminated by errors) Yt for i = 1,2,...,n is collected and then the maximum likelihood estimator (MLE) of Cpm is defined as ~Y ^pm
USL - LSL
Gjsl+iY-Tf
d
3^+(f-T)
2
where Y = ] T } - - i ^ / n > "^« — Xy--i(^» ~ ^ ) l n ' anc * ^ u s u a , l d = (USL - LSL)/2 is used. We note that F and S 2 are the maximum likelihood estimators (MLEs) of \i and a\ respectively. The term S\ + (Y - T)2 = ^ " = 1 ( K i - T) 2 In in the denominator of Cpm is the uniformly minimum variance unbiased estimator (UMVUE) of cr2y +{n-Tf = E\(Y-T)2) appearing in the denominator of Cpm.
Process Capability Measures in Presence of Gauge Measurement Errors
265
Evidently, if the a^ — 0 , the estimator of process capability (C m) results in the basic estimator Cpm introduced by originally Boyles (1991). Observe that Cpm is distributed as Y
C
d 3ov"^A/<«
Pm
where OQ — a2 + a\
which can be alternatively expressed as CY ~
d I n 3oy 1
CY
P
I n
~ V<«
V*iU
= c
^ v 1 + ^I \
n
(see (13.5)),
X
n,SY
where %2 ^ denotes the non-central chi-square distribution with n degrees of freedom and the non-centrality parameter 6Y = n^y and £y = (M ~~ T)/<JY • The c.d.f. of CYvm can also be expressed in terms of a mixture of a chi-square and a normal distributions: F^y
-x
(x)
&rVn/(3z) ?K
(6yVn) 9x2
t2
x [»(* + £yVri) + >(* -£y>/ra)]di, for £ > 0 , where by = d/ay = 3 C j , FK(-) is the c.d.f. of the ordinary central chi-square variable Xn-i a n d
(Y
=
M
A*
aY
2 (C Y \ w
pm
1 > C„ —
c„
vr+A^
(see (13.2)),
266
Encyclopedia and Handbook of Process Capability Indices
rY
and
-
(c.f. (13.5)).
^ pm
We again remind our readers to differentiate between Cpm, f
j
a l l u
Cpm,
pG ^ pm •
Pearn and Shu (2004) note that using the described procedure, the lower confidence bound is underestimated and the a -risk and the power of the test decrease with the measurement errors. Suppose that the desired confidence coefficient is 7 then the adjusted confidence interval of Cpm with the adjusted lower confidence bound LA can be obtained by solving the equation
1-2
f
MJ[ZC\
•K
Jo
where bA = 3LA/yjl + X2Cp and the value of Cp can be obtained by solving Cp = Cpm . To improve the power of the test, then we shall revise the critical values CQ to satisfy c0 < c0 . Thus, the probability P(Cpm > CQ J will be greater than P(C^m > c0) . Both the a -risk and the test power increase when CQ is used as a new critical value in the testing. Suppose that the a -risk using the revised critical value CQ is aA ; the revised critical values CQ can then be determined by P(Cpm > c0 \Cpm < c) = aA . Consequently 6 G Vn/(3c^) J 0
?K
(bG^nf [9(^) 2
-r
aA,
where bG = 3c(1 + A 2 Cp) -1 / 2 and Cp is obtained as stated above; thus Cp = c . To ensure that the a-risk is within the preset magnitudes, we let aA = a and solve the above equation to
Process Capability Measures in Presence of Gauge Measurement Errors
267
obtain CQ . The power of the test (denoted by irA ) can be calculated as the follows: KA(Cpm)
= P{CGpm
2 ^Vn/(3c£)
= f JO
>ct\Cpm>c) . ?K
(by^nf
iw
_t2
\
"If
where by =
VT+AV
Example 13.2 The product investigated is the pH sensor combining process-hardened pH electrodes, double junction reference electrode, temperature compensation element, and a solution ground in a durable, reliable, high performance design. It is ideal for process control applications in the most aggressive streams found in the traditional processing industries, including chemicals, paper, metals and mining, utilities, food, pharmaceutical, etc. For wide range measurements where high accuracy is required at either or both extreme ends of the pH scales, a spherical glass with a minimal sodium error is considered good quality. For applications involving abrasive processes, a rugged glass with a thicker membrane is recommended. The rugged glass is most accurate in the range of 1 to 12 pH. The high temperature construction allows the sensor to be used for the process pH measurements at the temperatures up to 120° C (250°F). An integral lOOplatinum RTD (resistance temperature detector), compatible with many common pH transmitters and monitors, is a standard feature of this device. The reference electrode utilizes a double junction design to inhibit silver ions from contacting the process solution, thereby preventing junction fouling from silver compound precipitates. To help facilitate a noise-free signal, the sensor incorporates a solution ground post of titanium metal. 3/4 NPT threads are provided at both sensor ends to allow connection to insertion or submersion type mountings.
268
Encyclopedia and Handbook of Process Capability Indices
The measuring accuracy of the pH sensor is an important factor have a significant effect on the pH sensor quality. A type of the pH sensors has the specification limits, T = 0.0 pH, USL = 0.05 pH, and LSL = -0.05 pH. A total of 70 observations were collected and displayed in Table 13.3. A histogram and a normal probability plot show that the collected data follows the normal distribution. The Shapiro-Wilk test is applied to further justify this assumption. To determine whether the process is "excellent" ( Cpm > 1.33 ) with the unavoidable measurement errors A =0.30, we first determine that c = 1.33 and a = 0.05 . Then, based on the sample data of 70 observations, we obtain the sample mean G = 0.0200, the sample standard deviation Sn =0.0109, and the point estimator C$m = 1.4629 . From the Table 5(b), we obtain the critical value CQ =1.436 based on a , A and n . Since Cpm > c0 we conclude that the process is "excellent". However if we ignore the measurement errors and evaluate the critical value without any correction, the critical value will be c0 =1.547. In this case we would reject that the process is "excellent" since Cpm is not greater than the uncorrected critical value 1.547. Input now T = O.OpH, USL = 0.05pH, and LSL = -0.05pH, the 70 observations, A =0.30 (provided by the gauge manufacturing factory), and the desired confidence coefficient 7 = 0.95 into the Matlab computer program (available upon request) the 95% lower confidence bound on the true process capability is to be obtained 1.415. We thus can be assured that the production yield is 99.9978%, and the number of the nonconformities is less than 21.78 PPM (Parts Per Million). Table 13.3 observations for the measuring accuracy (unit: pH) 0.0236 0.0257 0.0304 0.0289 0.0057 0.0188 0.0204
0.0433 0.0281 0.0198 0.0227 -0.0018 -0.0046 0.0153
0.032 0.0246 0.0172 0.0186 0.0385 0.0322 0.0154
0.0157 0.0198 0.0234 0.0066 -0.0086 0.0136 0.0413
0.0129 0.0335 0.0311 0.0265 0.0178 0.0187 0.0089
0.0085 0.0253 0.0278 0.0399 0.0130 0.0249 0.0344
0.0406 0.0115 0.0035 0.0168 0.0240 0.0157 0.0273
0.0047 0.0306 0.0289 0.0240 0.0317 0.0250 0.0236
0.0047 0.0097 0.0044 0.0040 0.0200 0.0167 0.0158
0.0254 0.0194 0.0188 0.0235 0.0137 0.0083 0.0185
Process Capability Measures in Presence of Gauge Measurement Errors 269
13.5
Estimating and testing measurement errors
Cpmk in presence
of gauge
For the third-generation process capability index Cvm^ , we assume that X ~ N ([i, a2 ) represents the quality characteristic of the manufacturing process under consideration. We know that in practice, the observed variable Y (in the presence of gauge measurement errors) is actually measured instead of the true variable X. Assuming that X and M are stochastically independent, we then have Y ~ N(/x,o\ = a2 + <j2M) , and the empirical process capability index Cpmf. is obtained by substituting aG for a. The relationship between the actual process capability Cpmk and the empirical process capability Cpmk can be expressed as
cPmk ~~ v i + \ 2 c \ + e ' where, as above, £ = (/J, — T)h . To estimate the empirical process capability CGmk sample data must be available. For a bona-fide stable normal process, the empirical data (the observed data contaminated by errors) (7j, for i = 1, 2,.., n , is collected. The maximum likelihood estimator (MLE) of Cpmk is defined as USL-Y 2
^S
Y -LSL 2
j 2
+(Y- Tf ' 3yJS + (Y - T) J '
a n d § = are where Y = E"=i^/n " ELi(^ " Vf/n respectively the MLEs of [x and a\ • Note that the statistic S2n + (Y - Tf = E " = 1 ( i i -T)2 /n in the denominator of CYpmk is the uniformly minimum variance unbiased estimator (UMVUE)
270
Encyclopedia
and Handbook
of Process
Capability
Indices
of uY + (fi - Tf = E[(Y - T)2] . For processes with symmetric manufacturing tolerance (T — M), the estimator CYmk reduces to
rY
-
^ pmk
d-\Y-M\ 3^S2n+{Y-Tf,
where as above d = (USL — LSL )/2 . Evidently, if aM = 0 then the empirical process capability CYmk The c.d.f. of CYmk can be becomes the basic index C,pmk expressed in terms of a mixture of the chi-square and the normal distributions as
pmk
bY-fnj{l+'ix)
W-i-f
{by-Jn-t) ?K
2
9x2
x [
£YVri)]dt
for x > 0 where by = dloY = 3 C j , and, as above, FK(-) is the c.d.f. of the (central) chi-square distribution Xn—i and
fi-M
CY p -
+
()
+
3
{Cpmk ) ~ Q {Cp )
{Cpmk j
(CY ? -\
where CY =
p n
2
n
pmk J
, and C\mk =
Vl + A ^
Q
pmk
Vl + A2C72 + £2
Process Capability Measures in Presence of Gauge Measurement Errors 271
Assume that the desired confidence coefficient is 7 . Then the adjusted confidence interval for CYmk with the adjusted lower confidence bound LA can be expressed as bAJn/(l+3C%mk)
{bAVn -tf •K
/.
x [0(* + O.5Vn) + ^ ( t - O . 5 V n ) ] d i = 7 ,
(13.6)
where bA = 3.75LA/yjl. 25 + X2C2P + 0.5 and Cp can be obtained by solving the equation CYp = JT?XCYpmk + 1 / 6 .
Thus
C„
1.25L 1 2 2 + T7Vl.25 + X C p 6
2 2
Vl + X C p
Given A , L (the 95% lower confidence bound using Cpmk ), bisection and interpolation methods can be used to find the Cp values. In order to improve the test power, we shall revise the critical value CQ to satisfy eg < c0 . Thus, the probability P\C^mk > co) w m De greater than P[CYmk > CQJ. Both the a risk and the test power increase when CQ is used as the new critical value in testing. Let the a -risk using the revised critical value CQ be aA , the revised critical values CQ determined from P (Cpmk > c^ Cpmk < c J = aA, can be expressed in explicit form as: byJn/{l+3c^)
L
(byJn-t)
_t2
^K
X [4>(t + 0.5VTT) + (p(t ~0.5Jn)}dt
= aA,
272
Encyclopedia and Handbook of Process Capability Indices
where analogously to (13.6) by = obtained
by
solving
3 75c , + 0.5 and C„ is 2 2 Vl.25 + X C p
equation
Cp = y/l.25Cpmk + 1/6
Therefore , °p = , °5c +1. 2 2 Vl + X C p Vl.25 + X2C2P 6 To ensure that the a-risk is within the preset magnitude we let aA = a and solve the equation to obtain c^ . The power of the test (denoted by nA) is calculated as *A {Cpmk ) = P{C^
> <* | Cpmk > C) , (pG^/n-t)
r 6 K Vn/(l+3c 0 ^)
i.e.
TtA{Cpmk) = j
1
_t2
^
bK x [<j>{t + 0.5Vn) + 4>(t-
0.5-Jn)}dt,
O • I D O yjrfik
where bG =
. + 0.5 . Vl.25 + X2C2p
The results imply that employing the above estimator for sample data contaminated by measurement errors causes severe a underestimation of the actual capability, resulting in imperceptibly smaller test power. To measure the actual process capability, one must use modified confidence bounds and critical values that are suitable to practitioners for their factory applications. 13.6 Estimating and testing Cpu and Cpi in presence of gauge measurement errors Bordignon and Scagliarini (2002) performed a statistical analysis related to estimating Cp and Cpk . The empirical process capability indices Cpu and C$ are obtained by substituting oY for a. (Cpu or Cpi were discussed in Chapter 8) The degree of
Process Capability Measures in Presence of Gauge Measurement Errors
273
error contamination r has been defined to be (Mittag (1997)).
r = ^L a Thus the relationship between the empirical process capability index Cj and the actual process capability index Cj: Cj _
1
where as above CYU or CYt is denoted here as Cj and Cj denotes Cpu or Cpi. Since the variation of the observed empirical data is greater than the variation of the original data (with no measurement errors), the denominator of the index C/ increases, and we would have underestimated the true capability of the process when calculating the process capability based on the empirical data from Y . Since the process parameters (i and a are unknown, we cannot evaluate the actual process capability. However, given (independent) sample data {Yi,i = l,...,n} taken from the process, the estimators of Cpu and Cpi become ^y ^pu
T UoLi — Y f^y , = "n-1 To ' ^ p i = 0n_i
i
— L/oLi — .
Following the arguments in Chou and Owen (1989) and in Pearn and Chen (2002), we can show that the estimator Cj (Cpu or C^i) is distributed as d-tn^i(6Y), where d = 6„_ 1 (3Vn) _1 (don't confuse it with d =(USL — LSL)/2\) and tn_1(6Y) is a non-central t distribution with n-1 degrees of freedom and the non-centrality parameter 6Y = 3VnC7 / >/l + r 2 , with T = aM / a . Let the desired confidence coefficient be 6, then the adjusted confidence interval on Cpu with the confidence interval bound C*u , can be shown (using the methodology employed in this book on numerous occasions) to be
274
Encyclopedia and Handbook of Process Capability Indices
e = p{cpu >c:) = p
(USL-
, {Y + k[SY - HY > = P 3<7y
= P
= P
nY
vr+^j > -fc/Vn
Sy I Oy Z - 3VnC:/^/^T7?
_ 3(7pu
Sy / Gy
= P *n-l($/ =
Vi + r 2 > c u * |
y
n-l
-3VnC,
VT+ r
r) > *T
where Z~N(0,1) *F = - ( 3 C £ lK_^Jn
and k? = 3 Cju / bn^ ,
Analogously, the adjusted confidence interval on Cpi with the confidence interval bound C*, can be obtained as
9 = P tn-l({>L =
3VnCf
}
< ^
VI + T<
where tf = (3C;,/^_i)Vn . Note that in this case, the probability that the interval with the bound C\ contains the actual Cpu value is greater than the corresponding probability for the interval with the bounds Cu or Cu , while the probability that the interval with the bounds Cu or Cu contains the actual CPU value is the constant 0.95. In order to improve the power of the test, the revised critical q$ ought to satisfy a* = P(C]
>cl\CI=c)
= P{?>JnCj > 3VncS | C1 = c)
Process Capability Measures in Presence of Gauge Measurement Errors
{K-\
K-\
= p[tn^(6Y = P
275
= SVnCj)
> ^-ct
tn_^ =^-1J=)>^2
Vi + r '
| Cj = c) A
K-\
CO
To ensure that the a-risk is within the preset magnitude, we let a* = a. Then CQ can be obtained as Co
=
hn l - t ^ {*Y 3Vnn a
= 3Vn
C
4i + ^-
and the power it* becomes 7r*(C/) = P(Cj > 4 | Ci) = P(3^nCf = P T I °n-l
W
> 3Vnco* | C 7 )
> -7 Co | W °n-l
= P "n-l
= P tn-\(6
= 3Vn
C/
VI +
f) > * n - l , a ( ^ = 3Vn TZ
VTT^J
Example 13.3 TFT-LCDs (thin-film-transistor liquid crystal display) consist of a lower glass plate on which the T F T is formed, an upper glass plate on which the color filter is formed, and the injected liquid crystal between both glass plates (see Figure 13.2 (a)). The T F T plays a critical role in transmitting and controlling electric signals, which determines the amount of voltage applied to the liquid crystal. The liquid crystal controls light permeability using different molecular structures that vary in accordance with the voltage. In this way, the desired color and image is displayed as it passes through the color filter (see figure 13.2 (b)). The TFT-
276
Encyclopedia and Handbook of Process Capability Indices
LCD consumes less energy compared to a CRT (cathode-ray tube), is slimmer and weighs less. The TFT-LCD becomes the most widely used display solution, because of its high reliability, viewing quality and performance, compact size and environmentfriendly features. Because of the heat resistance, non-conductance, and simple processing steps. Non-alkali thin film glass is the major material of manufacturing TFT-LCD. While manufacturing nonalkali thin film glass, flatness is one of the critical quality characteristics. If the flatness of glass is not in control, the TFTLCD products may result in a certain chromatic aberration.
AA » » » » *
M f f ColorRtterArroy
Liquid Ovsto!
r—:
Back light Plasma' PelaVlIlPtJj '
IFTArfOY
, .
5?Z.««vCT\ Cotern,t8r
Back plate with discharge eiectrode pattern
LC transparent electrode
Bock Light
Figure 13.2(a). Structure of a TFT- Figure 13.2(b). Structure of a TFT-LCD. LCD.
The measuring accuracy of a pH sensor is an important factor which has significant effect on the pH sensor quality. A type of the pH sensors has the specification limits T = 0.0 pH, USL = 0.05 pH, and LSL = -0.05 pH. A total of 70 observations were collected and displayed in table 6. A histogram and a normal probability plot indicate that the collected data follows the normal distribution. The Shapiro-Wilk test is applied to further justify the stated assumptions. To determine whether the process is "excellent" ( Cpm > 1.33 ) with unavoidable measurement errors A =0.30, we shall first determine that c = 1.33 and a = 0.05 . Then, based on the sample data of 70 observations, we arrive at
Process Capability Measures in Presence of Gauge Measurement Errors
277
the sample mean G = 0.0200, the sample standard deviation Sn =0.0109, and the point estimator Cpm = 1.4629 . From table 13.4, we calculate the critical value c$ =1.436 based on a, A and n. Since Cpm > CQ , we thus conclude that the process is "excellent". Also one can verify that if we ignore the measurement errors and evaluate the critical value without any correction, the critical value becomes c0 =1.547. In this case we would reject the hypothesis that the process is "excellent" since C^m is not greater than the uncorrected critical value 1.547. Moreover, inputing T = O.OpH, USL = 0.05pH, and LSL = -0.05pH, 70 observations, A =0.30 (provided by the gauge manufacturing factory), and the desired confidence coefficient 7 = 0.95 into the Matlab computer program (available upon request), the 95% lower confidence bound of the true process capability is calculated to be 1.415. We thus can assure that the production yield is 99.9978%, and the number of the nonconformities is less than 21.78 PPM (Parts Per Million). Table 13.4 observations for the measuring accuracy (unit: pH) 0.0236
0.0433 0.032 0.0157 0.0129 0.0085 0.0406 0.0047 0.0047 0.0254
0.0257
0.0281 0.0246 0.0198 0.0335 0.0253 0.0115 0.0306 0.0097 0.0194
0.0304
0.0198 0.0172 0.0234 0.0311 0.0278 0.0035 0.0289 0.0044 0.0188
0.0289
0.0227 0.0186 0.0066 0.0265 0.0399 0.0168 0.0240 0.0040 0.0235
0.0057 -0.0018 0.0385 -0.0086 0.0178 0.0130 0.0240 0.0317 0.0200 0.0137 0.0188 -0.0046 0.0322 0.0136 0.0187 0.0249 0.0157 0.0250 0.0167 0.0083 0.0204
0.0153 0.0154 0.0413 0.0089 0.0344 0.0273 0.0236 0.0158 0.0185
To summarize our discussion: when estimating capability the confidence bounds are substantially underestimated in the presence of measurement errors. When we use statistical testing to determine if the process meets the capability requirements, we find that the power of the test decreases with the measurement errors. Since the measurement errors are, unfortunately, unavoidable in many branches of manufacturing industry, to obtain a more
278
Encyclopedia and Handbook of Process Capability Indices
accurate confidence bound and to improve the power with an appropriate a-risk, adjusted confidence bounds and the critical values should definitely be applied.
Chapter 14
Process Capability Assessment with Tool Wear
14.1 Introduction 14.2 A review of various approaches 14.1
Introduction
Manufacturing systems are nowadays geared up to meet the challenges of a quality-based competition. Process capability studies and analyses have become critical issues in process control; indeed a number of guidelines are available for process capability assessment. Moreover, certain conditions such as normally distributed output, statistical independence of observed values and the existence of only random variation resulted from chance causes are stipulated for this assessment. All these conditions may not be fully satisfied in a practical set-up and departures are quite likely to occur. Tool wear is, evidently, constitutes a dominant and inseparable component of variability in many "machining" processes, and hence it represents a systematic assignable cause. Process capability assessment in such cases may become a bit tricky since the standard procedure may not provide accurate results. Process capability studies play an important role in the process control since they assist to decide whether a manufacturing process is suitable and the applications meet the necessary quality standards. The intensified quality-based competition has shifted the focus from the product control to the process control in manufacturing 279
280
Process Capability Assessment with Tool Wear
industries. The assessment of process capability, which now appears to be quite simple, involves significant dimensions in a practical set-up. This is due to the fact that some of the conditions necessary to establish process capability are not fully satisfiable. These conditions stipulate that the process has to be under statistical control, that the process output has to follow the normal distribution, be free of assignable causes, and the observed values of the quality characteristics be statistically independent. Since these conditions are not fully satisfied in many manufacturing situations, the process capability analysis is becoming to be a critical issue. It just does not make sense to proceed with a full-scale production without first conducting process capability studies: This situation was of course realized by many researchers and resulted in numerous publications in the literature. A process capability analysis is therefore valid only when the process under the investigation is free of any special or assignable causes (i.e., is in-control). Not much has been done in the area of process capability for the situations where the presence of an assignable cause is known and tolerated. A process is said to have a "tool wear problem" when a variation due to a systematic cause is present. There are, in fact, two areas of interest when studying a process: process stability and process capability. It is important to have clear guideless about control before developing the plan for a tool wear process. Specifically the intent of the plan to detect changes in the process or is the goal just to monitor the tool? An action to be taken in an out-of-control situation should be determined by the intent of the plan. Statistical process studies and the ongoing control are quite complicated when dealing with machine processes possessing a tool wear. Indeed such a wear is a fact, and it is essential for processes that exhibit tool wear to be controlled, to maintain high part quality and to maximize the tool life. In its simplest and most common form, tool wear data tend to have an upward or a downward slope over time. To determine this trend, a best-fit line to the data ought to be generated. For standard control charts, the grand average and the control limits are
Process Capability Assessment with Tool Wear
281
usually horizontal. In contrast, when the tool wear is involved, the control limits will be parallel to the tool wear slope. Once control has been assessed, the capability of the process can be then determined. 14.2
A review of various approaches
First of all it is necessary to ensure that the process at hand is under statistical control in order to assess its capability. Porter and Oakland (1991) pointed out that the two specific conditions which make the process capability assessment to be difficult are: 1) ensuring stability of the mean and of the standard deviation; and 2) an absence of any special causes. It is also expected that observations be statistically independent, which is not always the case. Indeed, processes with uncontrollable but acceptable trend are common in practice. The causes such as tool wear are responsible for inducing autocorrelations and are not therefore physically removable from the process. The issue of a correlation among the samples and its effect on the control chart limits has been examined by numerous authors in the last 25 years. Vasilopoulos and Stamboulis (1978) have shown that the limits for an X chart can vary by a wide margin especially if the correlation effect is not taken into account. These authers have suggested factors that should be used with an S chart or an X chart, provided the data can be fitted within either a first or second order autoregressive model. However, the effect of correlation in estimating standard deviation, a, has not been considered. Burr (1979) has analyzed the problem of estimating a from the aspect of constructing control limits. Initial setting of a tool has also been studied to prolong the utility of the tool before its replacement or resetting. Accordingly in these situations a is estimated as R / d^, where R refers to the range of values over a short run and ^ is a constant based on the sample size. The control limits are then calculated in the usual manner. Montgomery (1985) suggested fitting the first order autoregressive model to the correlated data,
282
Process Capability Assessment with Tool Wear
which could be of the form: xt = £>+cj>xt_l +et,
(14.1)
where xt is the value of the variable at time t, £ and
here cf> refers to the correlation coefficient. The effect of correlated subsamples in statistical process control was also observed by Neuhardt (1987), who commented that the correlated measurements within subgroups tend to increase Type I error rate in the case of X charts. Accordingly, the actual variance is given by, a2 (1 — r + kr), where r= the pairwise correlation among the subgroups, and A; is the common subgroup size. Time series modeling trend data had been also suggested by Alwan and Roberts (1988), who recommend using residuals in monitoring the process. As the residuals do not have any engineering significance, it is necessary to plot the individual values. An important aspect in deciding on the control limits for the individual observations chart is proper estimation of a. The commonly accepted procedure is to from ranges by taking differences of successive values. Then the average of these moving ranges of size two is applied to estimate a using the standard formula:
Process Capability Assessment with Tool Wear
.
283
MR
where MR is the average of the moving ranges and c^is a factor based on the subgroup size used extensively in control charts analysis. Ryan (1989) object to the above procedure on the grounds that a moving range of two will smooth out any trend presented in the data, which is inadvisable if the trend is a permanent part of the data. Furthermore, it is demonstrated that when successive values do not differ much and are strongly correlated, the moving average approach substantially underestimates a. The alternative suggestion is to estimate a as: s o = —, c4 where s is the sample standard deviation for at least 50 (calculated from an unbiased correction factor involving Gamma function) observations that are independent and c4 is another factor based on sample size. It is also stated that s / c4 gives a better estimate of a than that MR / d^ even in the situations when the observations are not independent as long as at least 50 observations represent the data generated by the process. Furthermore, Ryan (1989) has also suggested that an appropriate method to estimate a would be to fit a time series model and then perform the estimation. For instance, the variance a\ of an autoregressive process of the first order, is given by _2 _
°z
where ^ = is the correlation coefficient between successive values, and a\ = 1 is the variance of residuals from the fitted model. Although several authors have looked at the issue of correlation from the point of control charts, the process capability aspects did not received due attention. There are some situations when
284
Process Capability Assessment with Tool Wear
assignable causes are systematic, such as tool wear, so that their effects can be decomposed before the capability is estimated. This approach is also referred to as a constant or consistent process drift, and other examples which include accumulation of contaminants and temperature change drift must be quantified and removed before the remaining variability can be analyzed for statistical control (Kotz and Lovelace (1998)). Some approaches attempt to remove the variability associated with the systematic cause. For example, Yang and Hancock (1990) recommended that in computing the basic Cp index, an unbiased estimator of a can be obtained as a/(\ — pfl2, where p is defined as the average correlation factor and is given by, P = er^/[fc(A;-l)], where {r^} , i, j = 1 to k is the correlation matrix, with the entries r- • —
1
for all i — j
r\j
I fij \< 1, for i * j
V i,j = 1 to k ,
Some other authors make a general assumption of linear degradation in the tool. For instance, Quesenberry (1988) suggests that tool wear can be modeled over an interval of tool life by a regression model and assumes that the tool wear rate is known or a good estimate of it is available, and that the process mean can be adjusted after each batch without an error. However, the task of model-building does not appear to be neither very easy, nor directly applicable to realistic conditions. The procedure suggested by Long and De Coste (1988) first removes the linearity by regressing on the means of the subgroups and then determines the process capability. Long and De Coste (1988) discuss the techniques for obtaining the best-fit line through the data, calculating the control limits, comparing the slopes to determine different tools, and calculating the capabilllty of the process. These techniques are based on the
Process Capability Assessment with Tool Wear
285
assumption that tools are "consistent" within their tool groups. As the data are recorded over several tools, the subgroup averages are plotted over time. A best-fit line is then determined using standard linear regression analysis. The standard equations for calculating the best-fit line are:
y = b0 +b1x, E(s - x)(y - y)
where ^ = — ^ ^ —
2 —-
Yl(x — x)
and b0 = y — }\x .
The best-fit regression line for every tool is estimated by the points that are the averages of each subgroups, comprised of five consecutive parts. Once the best-fit lines have been determined for each of the tools under investigaton, the slopes of the lines are tested for a statistically significant difference. One possible method for determining significant difference is the "£" test. First, the largest and the smallest slopes are tested for significant difference. The procedure of testing hypothesis is here: HQ : bi — b} = 0 versus Hj : ^ — 65 ^ 0, and the test statistic
JMSE/J2(x - xf ' where b\ and bi are the slopes of best-fit lines 1 and 2, respectively, and MSE is the residual mean square error for line 1. The calculated test statistic is compared with a critical " t" value. If the test statistic is larger than this critical " t " value, the null hypothesis is rejected. This shows that a statistical difference exists between the slopes. If there is no significant difference between the largest and smallest slopes, then evidently "statistically", no difference exists between the remaining intermediate slopes. However, if a statistical difference does exist
286
Process Capability Assessment with Tool Wear
between the largest and smallest slopes, one then repeats the procedure with the next largest or smallest slopes. In summary, we repeat this procedure until no statistical difference is evident between the slopes. When two or more slopes are found to be statistically equivalent, the data can then be considered to have come from the same population and can be combined and studied together. In particular, the average slope is calculated and the equations ought to be adjusted so that the y-intercept, b0, becomes zero. After the equations have been adjusted (to have all zero intercepts) the slopes are being averaged. Those slopes that turn out to be statistically different indicate that their data cannot be considered to have come from the same population and ought to be studied separately. The next step involves calculation of the control limit around the average best-fit line. Quite a few methods exist for determining the so-called "three-sigma" control limits. The method suggested by Long and De Coste (1988) incorporates the average range and the subgroup size to estimate the three sigma limits. The procedure for calculating the control limits is: Upper Control Limit = Y + (A2)(R)(^Jk / m) Lower Control Limit = Y - (A2)(R)(->Jk/ m), where Y is average best-fit line for i = 1 to the maximum number of subgroups A% = a constant based on a subgroup size R = an average of subgroup ranges k = subgroup size used to calculate the average slope line and m = subgroup size used on the control chart. While calculating the control limits, we compute the average range over all the tools that were found to be statistically equivalent. Separate control limits must be calculated for tools that are found to
Process Capability Assessment with Tool Wear
287
be statistically different, and control limits are plotted around the average best-fit line. Calculations for the range chart follow the standard procedure. The spread of these control limits serves as the guideline for the ongoing control. For a tool wear processes, the ongoing control plan incorporates the control limit spread calculated from historical data along with the starting and the ending points of the control tool. An estimate of the starting point must be provided for each new tool. The starting point is approximated by sampling three successive subgroups. From these subgroups, the mean is calculated and the best-fit line is estimated using the slope of the average best-fit line. Once the estimated best-fit line is plotted, the control limits, calculated from the sample data, are drawn around the line. Next one has to develop an appropriate sampling procedure taking all sources of variation into account. The minimum or maximum starting and ending values should be developed to ensure that the material produced is within engineering specification limits at a certain confidence level. The starting and ending values are calculated by adding or subtracting a confidence factor from the specification limits. The confidence factor is determined by multiplying the standard deviation of the best-fit line by a constant to ensure a specific level of confidence. A constant of three or four is usually used. If the process being studied has a positive slope, then the sample starting point must be above the minimum starting value. The last sample should not exceed the maximum ending value. For a negative slope, a maximum starting value is of interest as is the minimum ending value. It should be noted that the specification limits are used only as a reference in determining the tool life and are not drawn on the actual control chart. Once the charting commences the stability of the process or simply the tool wear behavior can be checked. A reaction to an out-of-control situation depends upon the purpose of the chart. If the chart is solely for tool monitoring, the samples are taken as
288
Process Capability Assessment with Tool Wear
scheduled and the points are recorded until the actual number of tool cycles reaches the maximum ending point. At this point the tool is replaced. However, the process knowledge is also of importance, the control limit spread, based on the expected data, is centered around the best-fit line of the current tool. This allows for detection of variation other than tool wear. This procedure was suggested by Long and De Coste (1988) and it requires that the cycle of the tool be identical over the entire lifetime. This is usually not the case in situations where a tool is retooled several times or even when an old tool is discarded at the end of each life cycle. The tool must therefore be considered as a potential source of variability as well. When systematic assignable causes are present and tolerated, the overall variation of the process ( a 2 ) is composed of the variation due to random causes ( a 2 ) and the variation due to assignable causes (cr2,)- That is, a2 = a2 + a\ . The traditional PCI measures fails to acknowledge that portions of the overall variation, (in the presence of tool wear), will be due to assignable causes. Hence any estimates of the process capability will confound the true capability with these two sources. In order to get a true measure of process capability, any variation due to an assignable cause must be removed from the measure of process capability. However, the above approaches tacitly assume a static process capability over a cycle. By allowing the process capability to be dynamic within a cycle, as well as from a cycle to cycle, one could circumvent some of the problems encountered. Spiring (1989, 1991) viewed this as a dynamic process which is in a constant change as a process, tool, age, etc. In this dynamic model, the capability of the process may vary, possibly in a predictable manner. Spiring has devised a modification of Cpm index for this dynamic process under the influence of systematic assignable causes. In this scenario the goal is to maintain some minimum level of capability at all times. As a result, the capability will be cyclical in nature, its period being defined by the frequency of the process/tolling adjustments. Even
Process Capability Assessment with Tool Wear
289
when the assignable cause variation is not systematic, as is in the case with tool or die wear, one ought to be able to deal with random fluctuations of the process mean over time. Often in practice, deviations from the target value are due to easily determined assignable causes, such as shift-to-shift changes, differences in the raw material batches, environmental factors, etc. The proposed measure of process capability for dynamic process proposed by Spiring (1991) is
mm{uSL-T,T Cpm =
- LSL}
3 ^ + (* - Tf
'
where USL , LSL and T are the usual standard entities used in the assessing of Cpm and other PCIs but where /xt represents the mean and o>(-the variation (due to random causes only) of the process at time period t. As we have already remarked (perhaps too often!) the actual value of m and oft a r e rarely known, hence in order to get some assessment of process capability these values ought to be estimated. Analogously to the population values, the estimators will have to incorporate various existing sources of variation. Monitoring process's capability will thus require obtaining the value of Cpm or of a suitable estimate at various times t over each cycle during the lifetime of the tool. A sampling scheme that is used to monitor the capability of a process ought to reflect an array of factors associated with the process. Several factors, such as the frequency and the magnitude of other assignable causes, can be dealt with in a manner similar to the approach when dealing with the control chart sampling schemes. See, e.g. Montgomery (1985) for a discussion of "rational subgroups". However other (new) factors, such as the length of the sampling "window", will be influenced by the factors that are normally not encountered in control chart sampling schemes. The proposed sampling scheme is similar to the schemes used in monitoring a process for control charting procedures. The general
290
Process Capability Assessment with Tool Wear
format is to gather k subgroups of size n from each cycle (e.g., the period from t0 to tx) over the lifetime of the tool. The value of k will be unique to each process and, in fact, may change from cycle to cycle within the process. On the other hand, sample size of less than five (i.e., n < 5) are cautioned against, while larger samples (e.g., n > 25) may also pose problems. The optimal sample size for assessing process capability in the presence of systematic assignable causes will vary for each process under consideration. The manner in which the samples are taken is also of importance. In most situations consecutive sampling is suggested (i.e., the samples are collected as they come off the line). However, if desired, a "systematic 1 in R sampling scheme" can be used (i.e., within a period every second unit off the line is collected). However as R increases, problems may be encountered. Assuming that the effect of the tool deterioration is linear over the sampling window, estimates of Cpm are available which would not involve contribution of the assignable causes. Such an estimate is:
mm{uSL-T, Pm
~
T - LSL]
3ylMSEt+^(Xt-Tf
This measure of process capability considers only the proximity to the target value T and the variation associated with random causes as the linear effect of the tool wear is effectively removed by using n
MSEt = ^ * ra-2 of sequentially selected points (i.e., t^,^,^,---) rather than the common estimator s2. The MSEt is the mean square error associated with the regression equation xa. = aa + /3ta. + ea. and where ta. is the sequence number of the sampling unit and
291
Process Capability Assessment with Tool Wear
ea. ~ N(0,1). The coefficient 0 denotes the linear change in the tool wear given a unit change in time/production. Using the ordinary least square (OLS) estimates of a and f3 and assuming the sampling scheme to be sequential, the computational formula for Cvm becomes
h
' pm *-* rym.
—
. \USL-T min | 3
12
E<
'
ix 2
Y,( 0
T-LSL] i=l 3 J (n - 1) 12
2
i-i
^E(^) »=i
n(n - l)(n - 1) + - (n - 1) 2
2n(2n + 1) _ X (n - I)2~ k
+
n(xta - Tf n-l
where n denotes the subgroup sample size, T the target value and xt representes the i th value of the quality characteristic in the sampling period ta , while 3c(<j = Z"=i xta. n • The above method was proposed by Spiring (1991) who have suggested (as it was already mentioned above) that the problem can be tackled by viewing the process capability as being dynamic rather than static. This entails calculating a new index and monitoring it as the process advances. When the index reaches a preset minimum value, processing is terminated and resetting/replacement is carried out. The method has unfortunately some drawbacks: •
In addition to the regular charts, a new chart showing various values of the process capability will have to be drawn. This necessitates deciding on the sample size, the number of samples, the associated confidence level and the frequency of sampling which may involve a substantial number of calculation.
•
Great effort may be required for a graphical analysis and interpretation.
292
•
Process Capability Assessment with Tool Wear
The usefulness of the procedure described ought to be ascertained in a typical practical set-up.
Thus it would seem that there is no generally agreed upon procedure for assessment of process capability in the presence of assignable causes that affect the process in the case of a tool wear situations. Wallgren (1996) has also studied the properties and implications of the index Cpm when the consecutive measurements represent observations of dependent variables stemming from a Markov process in discrete time. This occurs, for example, when consecutive measurements from a process are serially corrected. The author had also developed an augmentation of the index Cpm , denoted Cpmr based on the first order autoregressive model (AR(1)). Unfortunately this report to the best of our knowledge has been not published as yet.
Chapter 15
Process Capability Assessment for Non-normal Processes
15.1 Introduction 15.2 A brief review of various approaches 15.1 Introduction We shall very briefly sketch the basic assumptions and properties of process capability indices. Process capability indices have been widely investigated as a means of summarizing process performance relative to a set of specification limits. These indices are effective tools for both process capability analysis and quality assurance. The proper use of process capability indices, which are statistical measures of process capability, is based on several assumptions. One of the most essential is that the process monitored is supposed to be stable and the output is approximately normally distributed. When the distribution of a process characteristic is non- normal, PCIs calculated using conventional methods could often lead to erroneous and misleading interpretation of the process's capability. A number of quality control experts have provided useful and insightful information regarding the errors in interpretation of the values of the PCIs that occur due to the misapplication of the indices to non-normal data (see, e.g., Sarkar and Pal (1997), Chou et al. (1998) among others). English and Taylor (1993) have examined the effect of the non-normality assumption on PCIs and
293
294
Encyclopedia and Handbook of Process Capability Indices
have concluded that Cpk is more sensitive to departures from normality than Cp . Somerville and Montgomery (1996) studied the errors that can occur in calculating Cp or Cpj. for a nonnormal distribution and making inferences about the PPM nonconforming under normality assumption. They have reached a conclusion for the four non-normal distributions, most common in engineering and reliability applications: the t, the gamma, the lognormal, and the Weibull distributions, that although the magnitude of an error can vary substantially depending on the true (unknown) distribution parameters, the errors are of consequence in nearly all cases. Hence, if the capability indices based on the normal assumption concerning the data are used to deal with non-normal observations, the values of the capability indices may, in a majority of situations, be incorrect and quite likely misrepresent the actual product quality. Actually, a leading investigator of quality control problemsnon-normal processes occurs frequently in practice. Gunter (1989) - a leading investigator of quality control problems - has mentioned the special case of the hole diameter of a drilled hole. Due to the nature of the process a minimum hole size is available in the vicinity of which the majority of hole diameters will cluster, however some very large holes occasionally occur due to such factors as vibrations, dull tools, or crooked setups among others. Pyzdek (1995) also informs us that distributions of certain chemical processes such as zinc plating thickness of a hot-dip galvanizing process are very quite often skewed. Choi (1996) presents an example of a skewed distribution in the "active area" shaping stage of the wafer's production processes. Bittanti et al. (1998) address the problem of defining and computing reliable estimates for process capability indices (in particular for Cpk) for non-normal processes; specifically, a curve-fitting approach to the estimation problem is employed and the problem of providing confidence intervals for the estimates of the "non-normal" PCI's is considered. Applications examples of non-normal process
Process Capability Assessment for Non-normal Processes
295
capability indices in various manufacturing industries are provided in Pearn and Chen (1995, 1997). For these skewed processes, the proportion of nonconforming items for fixed values of standard PCIs tends to increase as skewness increases, i.e. the standard PCIs simply ignore the skewness of the underlying population. For example, if the underlying distribution is Weibull with the shape parameter (3 = 2.0 or 1.0 (i.e. the skewness is 0.63 or 2.00, respectively) and LSL = —3.0 and USL = 3.0 , then the expected proportions of non-conforming items below and above the lower and upper specification limits are 0.56% and 1.83%, respectively, for the same value of |U = 0 and a = 1 . Hence Cp = Cpk = 1.0 , whereas the expected non-conforming proportion for a normal population is 0.27% being. Therefore, a method of adjusting the values of a PCI in accordance with the expected proportion of non-conforming items by considering the skewness of the underlying population would be very desirable (see, e.g., Gunter (1989), Pyzdek (1992), English and Taylor (1993), Somerviile and Montgomery (1996), and Kotz and Lovelace (1998)). In the recent years, several approaches to the problems of PCIs for the non-normal populations have been suggested. One suggestion is to use data transformation techniques such as the Box-Cox power transformation, Johnson's transformations and quantile transform techniques (see e.g. Somerviile and Montgomery (1996), Kotz and Lovelace (1998), Tang and Than (1999)). In general, however, practitioners may feel uncomfortable working with transformed data and may have some difficulty in reversing the results of the calculations back to the original scale. A second method is to replace the unknown distribution by an empirical distribution or by a known three or four-parameter distribution. Examples include Clements (1989), Pearn and Kotz (1994), Franklin and Wasserman (1991, 1992), Shore (1998) and Polansky (1998). Castagliola (1996) introduced a non-normal PCI calculation method by estimating the proportion of non-conforming items using the Burr's distribution. This method is complicated and
296
Encyclopedia and Handbook of Process Capability Indices
requires proivitably large samples, so that most of the quality engineers inclined towards simple methodology may hesitate to use this approach. A third way is to modify the standard definition of PCIs in order to increase their robustness. This approach makes the PCI as insensitive as possible to the shape of the distribution (see, e.g., Pearn et al. (1992) and Rodriguez (1992)). Practice convey to us that the performance of this approach, is however, often unsatisfactory. A fourth method is to use heuristic arguments to develop new indices. Bai and Choi (1995) proposed weighted variance (WV) PCIs based on Choobineh and Branting (1986) ideas, and Wu et al. (1999) suggest new WV PCIs by modifying the WV method of Choobineh and Branting (1986). Chang et al. (2002) developed a novel heuristic method of constructing simple PCIs for skewed populations based on a weighted standard deviation (WSD) method. This method adjusts the values of PCIs in accordance with the degree of skewness in the underlying population by employing different factors when computing the deviations above and below of the process mean. When the underlying population is symmetric, these indices are reduced to the standard PCIs. Tang and Than (1999) review several methods and provide a comprehensive evaluation and comparison of their ability to handle non-normal situation. For more details about process capability indices, the two books, (Kotz and Johnson (1993) and Kotz and Lovelace (1998)), each have devoted a chapters for dissecting such problems. The recent review paper of Kotz and Johnson (2002) also discusses the problems and consequences of non-normality. As a consequence, capability indices are very powerful, but, like many powerful tools, they could inflict heavy damage if incorrectly used. Properly calculated, they provide a whale of vital information concerning how the current output of a process satisfies customer requirements. Incorrectly applied and/or interpreted, these indices can generate an abundance of
Process Capability Assessment for Non-normal Processes
297
misinformation that may confuse practitioners and cause them to carry out incorrect decision. 15.2 A brief review of various approaches Below we shall summarize several different methods for measuring process capability indices appearing in the literature. 15.2.1 Probability
plotting
approach
Probability plotting is an alternative to a histogram that can be used to determine the shape, center, as well as the shape of the distribution. It has the advantage that it is unnecessary here to divide the range of the variable into class intervals, and it often produces reasonable results when using moderately small samples. A probability plot is a graph of the ranked data versus the sample cumulative frequency on a special paper with a vertical scale chosen so that the c.d.f. of the assumed type becomes a straight line. A widely accepted approach for PCI computation is to use the popular normal probability plot (1994) so that the normality assumption can be verified simultaneously. Analogous to the normal probability plot, where the "natural" process width is between the 0.135 percentile and the 99.865 percentile, surrogate PCI values may be obtained via appropriately selected probability plots c
USL - LSL USL - LSL ~ (upper 0.135% point) - (lower 0.135% point) ~ Up - Lp '
= v
where Up ( Lp ) is the 99.865 (0.135) percentile of observations respectively. These percentile points can easily be obtained from a simple computer code performing probability plots. Since the median is usually the preferable central value for a skewed distribution, the corresponding Cpu and Cpi are defined as:
298
Encyclopedia and Handbook of Process Capability Indices
~
__
*~/ pu nil
_ p
USL — median (upper 0.135% point) — median
USL — median — median '
-XQ.99865
median — LSL _ median — LSL median — (lower 0.135% point) median — X0.00135
The index Cpk will then be calculated as the minimum of Cpu and Cpi, namely: Cpk = rnin | C p u , Cpl j USL — median median — LSL = mm' ^0.99865 - median ' median - X . 35 0 001 In the non-normal case, if we are able to find a better distributional form for the data, which provides a very satisfactory fit (this can be tested by means of goodness-of-fit tests), we can obtain more accurate measures of the three quantiles under consideration. This involves modeling the process data with alternative probability models, such as the Weibull or gamma ones (see e.g. Dudewicz and Mishra (1998), Kotz and Lovelace (1998)). However, an obvious disadvantage of probability plotting is that it is not a truly objective procedure. It is quite possible for two analysts to arrive at different conclusions using the same data. For this reason, it is often desirable to supplement probability plots with goodness-of-fit tests which possess more formal statistical foundations (see, e.g., Shapiro (1980)). A good introduction to these tests is in Shapiro (1980). Choosing the distribution to fit the data is also an important step in probability plotting. Sometimes one can use the available knowledge of the physical phenomenon or the past experience to suggest a choice of the distribution. Several earlier authors recommend to employ data transformations. Kane (1986) suggested transforming the data to
Process Capability Assessment for Non-normal Processes
299
create an approximately normal situation. Gunter (1989), among numerous other researchers and applied statisticians, has found empirically that in this general situation, results for transformed data are typically much better than for the original raw data! The well-known quantile transformation techniques were developed by Clements (1989), who utilized the Pearson curves to provide better estimates of the relevant quantiles. 15.2.2
Clements'
approach
For non-normal Pearsonian distributions (which include a wide class of "populations" with non-normal characteristics), Clements (1989) has proposed a novel method for calculating estimators of Cp and Cpk , which uses Pearson curves to provide more accurate estimates of -XQ.00135 i -^0.50 (= median) and •^0.99865 • His modifications of Cp and Cpk are appealing since they do not require mathematical transformation of the data, have a transparent meaning which is easy to comprehend, for nonstatisticians and also can be estimated manually with miniature hand-held calculators. Furthermore, Pearn and Kotz (1994) applied Clements' method to obtain estimators for the other two basic indices: Cpm and Cpmk. To refresh the memory, we cite the relevant expressions: - _ USL - LSL ~ U -L. P n V
Cp
Cpk = min r>
—
USL-M Up-M
M-LSL ' M - Lp
USL - LSL Up-Lp
+ (M -Tf
300
Encyclopedia and Handbook of Process Capability Indices
c,pmk
min
Up
USL-M -M? + (M
M M-Lp]
-Tf
-LSL + (M
-Tf
where Up is the 99.865 percentile and Lp is the 0.135 percentile determined from the "universal" tables of Gruska et al. (1989) for the particular values of mean, variance, skewness, and kurtosis calculated from the sample data. Clements' estimators for Cp and Cpm regardless of a specific distribution are obtained by replacing the 6cr by Up — Lp . For the indices Cpk and Cpmk , the process mean \x is estimated by the median M , and the two 3a s are estimated by Up — M and M — Lp respectively for the right and left-hand sides. Clements (1989) uses the classical estimators of skewness (/? 3 ) and kurtosis ( (3$ ) based on the third and fourth moments respectively, which are often unreliable for small sample sizes. To improve the accuracy, Pearn and Chen (1995) consider the following modification replacing the a by (Up — Lp) / 6 in all cases regardless of whether it is on the right-hand or left-hand sides. Thus, Clements' estimators become: o„ —
Cpk
USL - LSL U„ — L„
USL - M M - LSL = mm [Up-Lp}/r\Up-Lp}/2\
Process Capability Assessment for Non-normal Processes
r' ^pm
Cpmk —
m
i
n
'
-
301
USL - LSL Up — Lp
+ (M
-Tf
USL-M Up — Lp
+ {M
M Up — Lp
-Tf
-LSL + (M
-Tf
See also the discussions of the estimator of C„ C C ^pk 5 ^pm and Cvmk for various particular situations presented in previous chapters. To illustrate how the modified estimators outperform the original Clements' estimators, Pearn and Chen (1995) consider an example of three processes: one on-target and the other two offtarget. In their example original Clements' estimators show little sensitivity to the departure of the process median from the target value, while the modified estimators clearly differentiate the ontarget process from the other two (severely) off-target ones. 15.2.3 Box-Cox
power transformation
approach
Box and Cox (1964) (B.C) proposed a useful and by now widespread family of power transformations of a (necessarily) positive response variable X . The B.C. times transformation is given as:
302
Encyclopedia and Handbook of Process Capability Indices vA _ -i
x^ = •
, for A * 0 A
lnX,
for A = 0.
This (continuous) family depends on a single parameter A which can easily be estimated by the method of maximum likelihood. Indeed first, a value of A from a pre-assigned range is collected. Next, we evaluate Anax = ~ 2 l n ( ? 2 +ln(A,X)
= -iln
where J(A, X) = II ^ i=l OX
= II X?-1
for all A .
i=l
(Thus lnJ(A,X) = (A - l ) ^ " = i l n Z i •) T h e e s t i m a t e o f &2 f o r a fixed A is a 2 = 5(A)/n , where 5(A) is the residual sum of squares in the analysis of variance of X^ . After calculating Anax (A) for several values of A within the range, we plot Lmax(X) against A . The maximum likelihood estimator of A is then obtained as the value of A that maximizes Lmax(\) . With the optimal A value, each of the X data specification limits are transformed to a normal variate using the above equation. Thus, the corresponding PCIs can be calculated from the mean and standard deviation based on the transformed data. 15.2.4 Johnson
transformation
approach
Over 50 years ago, N.L. Johnson (1949) proposed-by now a classical system of distributions based on the method of moments,
Process Capability Assessment for Non-normal Processes
similar to the Pearson system. transformation is given by
The
general
form
303
of
the
z = 7 + T]T(X; e, A)
= 7 + Sg(x) T) > 0 , A > 0, —oo < 7 < oo, —oo < e < oo , X = 9~1
or
(Z-l\ V
Here z is the variable to be fitted by a Johnson distribution and Z is a standard normal variate. The four parameters, 7 , rj , £ and A are to be estimated and g is a function which may take one of the following three forms: the lognormal system (Si), the unbounded system (SJJ) and the bounded system (SB) (see, e.g., Tang and Than (1999)). For example, the Sg (bounded system) involves transformation of the form g(x) = \og{x/(I — x)} . (Z is a standard normal random variable with the p.d.f. /(z) = (l/V2^)e-* 2 /2 ,
zeR,)
and the system Sg results in the p.d.f. of X :
^l7,n) = ^ ^ 15.2.5 Other quantile
r
^ e x
P
transform
-H
7 + 7?1 g
\2
° ((T^
approaches
Gilchrist (1993) has also developed a quantile transform technique that is based on the same philosophy as Clements' method but uses the so-called standard distributions for the transform rather than Pearson's curves. Gilchrist's technique utilizes an inverse transform function to define Xp in terms of p for the modeled distribution (see, e.g., Banks (1996)). As an
304
Encyclopedia and Handbook of Process Capability Indices
example, suppose the process data are "reasonably" exponentially distributed with the p.d.f. f(x) = Xe~Xx , x > 0, A > 0. In this case, the function Xp = —(1/A)ln(l — p) can be used to estimate ^0.00135 , -Xo.50 a n d -^Q).99865 > where 1/A is the mean of the exponential distribution and also 1/A = X . Large sample sizes are required in order to ensure a satisfactory distributional fits. Chang and Lu (1994) have also extended Clements' technique by proposing a percentile method for calculating PCIs without assuming normality of the distribution, using appropriate statistical tables, or calculating skewness and kurtosis of the data. This approach guarantees that both the indices Cp and Cpk > 1 if all the observations in the sample fall within the specification limits (and the process is admissible). The technique assumes that the sample dataset reflects the entire variation range of the population, in other words, it is a "perfect" sample (which rarely appears in real-world situations). The algorithm developed by Hill et al. (1976) is used to match the first four moments of X to the above distribution families. Then the values of PCIs can be calculated using the transformed data. 15.2.6 Distribution-free
tolerance
intervals
approach
Chan et al. (1988) adopted a distribution-free tolerance interval approach to compute Cp and Cpk for non-normal processes using _ USL - LSL _ USL - LSL _ USL - LSL ~ 6a ~ f (4a) ~ 3(2a) "
v
Process Capability Assessment for Non-normal Processes
305
This results in p
_ USL - LSL _ USL - LSL _ USL - LSL w | ^2 3u; 3 min | USL — n, n — LSL [
where w is the width of the tolerance interval with the 99.73% coverage 95% of the time (corresponding to 3cr in each direction in the normal case), w2 is the width of the tolerance interval with the 95.46% coverage 95% of the time(corresponding to 2a in each direction) and w3 is the width of the tolerance interval with 68.26% coverage 95% of the time(corresponding to 1
15.2.7 Flexible
index Cjkp
Johnson et al. (1994) introduced a "flexible" PCI Cjkp (the letters jkp indicate the first letter of the authors names) which takes into account possible differences in the variability above and below the target value. The C ^ index also provides an advantage of changing its value (slightly) when the process distribution changes its shape (i.e. becoming skewed), something that the other process capability indices do not accomplish. CU^ and CL^ are defined as the capability of the upper and lower halves of the process data, respectively, CjAp being the minimum of the two. That is
306
Encyclopedia and Handbook of Process Capability Indices
Cjkp = min|C77#p, CLjkp j = min
USL-T T -LSL 3W+ V2 ' 3W_V2
where W2 = EX>T[(X-T)2] and Wl = EX
= E[(X-T)2 -T)2} = E[(X -T)2
\X>T]PT(X>T),
\ X < T]Pr(X
< T).
Johnson et al. (1994) have used the multiplier 1/(3V2) instead of 1/3 (used in other PCIs) because, for a symmetrical distribution with variance a2 and the expected value T EX>T [(X - Tf} = EX
Note that if we have T = (USL + LSL) /'2 = m , so that USL-T = T -LSL = d, where d = (USL -LSL)/ 2, then Cjkp reduces to: Cjkp = ^ = m a x { %
T
{(X - T)2\, EX
T)2}}^.
A natural estimator of C^p in the general case is then Cjkp = r - ^ m i n
3V2
where s+ = J2Xl>T^ that
~ T?
USL-T T -LSL\ yjs+ / n ' yjs_ I n J '
and s
~ = T.Xi
~ T? •
E[8+] = nE[(X - Tf | X > T] P r ( X > T)
Note
Process Capability Assessment {or Non-normal Processes
307
and E[s+] = nE[(X
-Tf
| X > T ] P r ( X > T)
Consequently s+ / n and s_ / n are unbiased estimators of EX>T [(X - Tf ] and EX
15.2.8 The Wright's
Cs index
The index Cpmk provides warnings about the increase of the process variation and the process departure from the target, but is sensitive to the changes in the shape of the distribution, particularly its skewness. A modification of Cpmk proposed by Wright (1995) incorporates a skewness term in the denominator to
308
Encyclopedia and Handbook of Process Capability Indices
reduce the index value when non-symmetry is present. Utilizing the third central moment /x3 = E(X — /x)3 as a measure of skewness, Wright's index Cs possesses the Cpm^ characteristic of applicability to processes whose mean may not be centered between the specification limits and moreover may not be centered at the target. The index Cs is defined as follows: min{t/SL-/x, (i - LSL] 3yja2 +(n-Tf
+\nz/°\
d-\fi-T\ 3Vc72+(M-r)2+|/x3/a|
(d-\»-T\)/a 3jl+
[(fi-T)/a}2
+1^'
where as usual d = (USL — LSL) / 2 is the half-interval of the specification range, T = (USL + LSL) / 2 is the target value, and fcj = /i 3 / a 3 is the skewness coefficient. The term / j 3 is divided by a to ensure that the skewness term is expressed in the same units as the other terms in the denominator and the absolute value of this term guarantees that a negative skewness will also incur a penalty. In the numerator, the well known identity min(x, y) = (x + y)/2 — \x — y\/2, x, y e R is utilized. In practice, it is required to estimate the unknown process mean fj,, the process standard deviation a, and the third central moment fj,3 , to calculate Wright's Cs. Usually these are estimated from a sample of size n. A natural estimator of Cs is proposed by Wright (1995) to be: d-\X C, =
Tf +
n2m3 (n - l)(n - 2)
n
x
n —if 1
rri2 ~2~ ci
-1/2
where r ^ = Y^i=SXi ~ x? I n , m3 = ^ " = i ^ ~ Xf In a r e the sample central moments and c4 = [2/(n — l)] 1 ' 2 x
Process Capability Assessment for Non-normal Processes
309
r ( n / 2 ) r [ ( r a - l ) / 2 ] " 1 is the corrector for the bias Wright (1995) investigated this estimator of Cs and studied its bias and variance using simulation for normal distributed processes. By evaluating the percentage bias in Cs and Cpm\., he has discovered that Cs is a more accurate estimator than C,prafc for small samples from offtarget processes, when significant shifts in the mean have been occurred. For skewed distributions, Pearn and Chang (1997) examined the bias of Cs, which turn out to be quite substantial. Chan and Kotz (1996) investigate the distributional properties of Wright's index and reach the conclusion that the asymptotic behavior of the estimator Cs is actually sensitive to skewness, and the contribution to overall variability of the skewness correction factor is additive and non-interactive. These authors also substantiate Wright's conclusion that the asymptotic distribution of Cs is normal when /i ^ T and /ix3 ^ 0 , which makes it a natural choice for process capability assessment for large sample sizes, regardless of the underlying distribution of the data. In general the Cs index provides a statistically sound approach to process capability analysis under the three (common) conditions: (1) USL-^^fi-LSL (i.e. n * M = (USL + LSL)/2 ); (2) \i ^ T ; (3) /i 3 ^ 0 . In our discussion, the process is assumed to be exactly centered ( \x = T ) which implies that Cs = Cpk = Cpmk for symmetric distributions.
15.2.9 A superstructure
capability
indices
CNp(u, v)
As it was mentioned in previous Chapter, Vannman (1995) constructed a superstructure of indices Cp(u,v) which includes the four basic indices, Cp , Cpk , Cpm and Cpmk as special cases. However, the indices Cp(u,v) seems to be appropriate only for the normal and near-normal processes. To accommodate those cases where the underlying distributions may not be normal, Pearn and Chen (1997) considered a generalization of Cp(u,v) defined below which is called CNp (u, v), suitable for application to processes with
310
Encyclopedia and Handbook of Process Capability Indices
arbitrary distributions Recall that d — u\ fi — m\ • '
CP(u, v) =
Sja2
+v(fi-T)2
,
(15.1)
Pearn and Chen (1997) modification is: d - u\M - m\ ^
CNp(u,v) = —f= r
99.865
—
^0.135
(15.2) 2
+ v(M - T)
Here Fa is the a th percentile of the underlying distribution, M is the median of the distribution, m = (USL + LSL) / 2 is the mid-point between the upper and the lower specification limits, JJL and a are the process mean and the process standard deviation, respectively and u, v > 0 . In developing this generalization C?fp(u,v), Pearn and Chen (1997) replaced the process mean /i by the process median, M (a more robust measure for process central tendency, particularly for skewed distributions with long-tails), and the process standard deviation a by (F 99865 — F0,135)/6 , calculated from the distributional percentiles in the definition of the original index Cp(u, v). The idea behind such replacements is to mimic the property of a normal distribution that the tail probability that the process is outside the ±3
Process Capability Assessment for Non-normal Processes
311
which can explicitly be expressed as: USL - LSI CfJr, 'Np —
•^99.865
—
(15.3)
-^0.135
USL-M
CNVk = min
M - LSI r
•99.865 ~~ ^0.135
99.865
—
, (15.4)
^0.135
to mimic the property of a normal distribution that the tail probability that the process is outside ±3
^Npm
and C Npmk
~
~
4
w
[ ^99.865
—
^0.135
(18.8)
I2 + (M - Tf
6
USL-M -^99.865
—
^0.135
2
6
+ (M - Tf
M -LSL
si
-^99.865 ~ -^0.1 35
6
(15.6)
2
+ (M - Tf
Zwick (1995) and Schneider et al. (1995) consider two generalizations of Cp and Cpk , which are similar to C^p and Cpfpk but using process mean /z rather than process median M in the basic definitions. Extending their definitions to include the indices Cpm and Cpmfc , a superstructure:
CiNp{u,v) =
d — u\fi — m\ ^99.865 ~ ^0.135
6
+ «(/x - T)
(15.7) 2
312
Encyclopedia and Handbook of Process Capability Indices
could be constructed, designated C'Np(u,v) (Chen and Pearn (1997)). Note that (15.3) is a variant of the original CP(u,v). Chen and Pearn (1997), Pearn and Kotz (1994), and Pearn and Chen (1995) propose estimators for calculating the indices Cp(u,v) assuming that the underlying distributions are of the Pearson types. These estimators essentially apply Clements' method (1989) (described above) by replacing the unknown 6CT distance by Up — Lp , which is calculated based on the available sample data obtained from a stable process using the estimates of the mean, standard deviation, skewness and kurtosis. These four parameters determine the type of the Pearson distribution curve; hence the Fa percentiles of the Pearson curves as a function of skewness and kurtosis can be calculated utilizing, for example, the tables provided by Gruska et al. (1989). These estimators are of the form CNp(u,v) =
d — u\M — m Up — Lp
l2
+ v(M - Tf
where Up estimates the 99.865 percentile F 99 865, Lp — the 0.135 percentile F0 135, and M is an estimator of the median. To obtain the values of Up , Lp and M Gruska et al. (1989) tables supplemented by certain interpolation are required. Based on sample percentiles, Chang and Lu (1994) consider a different method for calculating the -F99.865, ^0.135 > a n d the median M for non-normal distributions. Their method is essentially based on sample percentiles which can be calculated using only interpolations, and do not require the use of in Gruska et al. (1989) tables. Applying this method one can obtain the percentile estimators for C^p(u,v)&s
Process Capability Assessment for Non-normal Processes
CNp(u,v)
d — u\M — m\
=
•^99.865 ~~ -^0.135
+ v(M - Tf
6 where
313
F 99 . 865 = X{[Rl]) + {fy - [i^]} x {X{[Ri]+1) - X{[Rl]) } , ^0.135 = ^([i%]) + {-^2 - [-^2]} X {-^([iJj]+l) - X(.[Ri])}> M
=
X
([R3]) + iR3
_ /99.865n + 0.135 \ \ 100 /
~ [^3]}
x
{^([iJ3]+i) -
_ /0.135n + 99.865 \ \ 100 }'
x
([i? 3 ])} '
_/n + T ° \ 2 I
The notation [R] signifies the greatest integer less than or equal to the number R , and X^ is as usual defined as the i-th order statistic.
15.2.10
The Cpc
index
Luceno (1996) introduced the Cpc index, which is designed as to consider both the process location and the spread, but the other indices also provides for the confidence bounds that are insensitive to departures from normality. This index is defined as „
_
USL-
LSL
' pc
6j^E\X~m\ where m = (USL + LSL) / 2 . Luceno (1996) mentions that the second subscript in Cpc stands for confidence, to indicate that the confidence intervals based on Cpc are reliable. This statement however ought to be interpreted rather cautiously. The author also uses the factor 6-7^/2 = 7.52 in the denominator, to equate
314
Encyclopedia and Handbook of Process Capability Indices
it to 6cr in the case of the normal distribution N(fi,a2) . Note, however, that the Q-JTT/2 factor does not depend on the normality. Luceho (1996) developed a 100(1 — a)% confidence interval for Cpc by first evaluating | X — m | , whose expected value is estimated by the sampling analog: _ 1 " c = - Y] I Xi - m |, thus leading to the estimator n
—
USL - LSL < #
A 100(1 — a)% confidence interval for E\X — m\ is given by T +t u
-1- ua, n - 1
A. /— )
where ta/2,n~i i s the a/2 quantile (i.e. the value of that is exceeded with probability a/2 by the Student's t distribution with n — 1 degrees of freedom) and
;
l
A„
, _,,
l
= ^rE(l^- m l- c ) 2 =^ri ElXi — m
|2
—nc2
Therefore, a 100(1 — a)% confidence interval for Cpc is given by n
Process Capability Assessment for Non-normal
Processes
315
Lucefio (1996) comments that the confidence intervals are insensitive to departures from normality since they are based on the Central Limit Theorem. In addition, approximations are available for this index for n large. For instance, when X is distributed as N(n, a2) , the quotient sc/~c is close to ( T T / 2 - 1 ) 1 / 2 » 0.7555 , so that the 100(1 - a)% confidence interval for Cpc can be compactly approximated by Cpc/(l±0.7555iQ/2,n_1Vn). A critical region for the hypothesis H0 : Cpc = Cpc0 versus Hi : Cpc < Cpca at a level of significant is ~c-Jn C,pc -1 CBCO
15.2.11 The (general)
Weighted
-^ ''a, n—1 •
Variance
(WV)
method
The weighted variance (WV) method, already alluded to in an earlier Chapter, was first introduced by Choobineh and Ballard (1987) to construct control charts when the underlying population is skewed. The idea is based on the "semivariance" approximation of Choobineh and Branting (1986). Their method provides asymmetric control limits on X and R charts for skewed distributions. Bai and Choi (1995) also proposed a simple heuristic method to construct X and R charts using the weighted variance method with no assumptions on the population. This method provides asymmetric control limits in accordance with the direction and the degree of skewness estimated from the sample data by using different variances in calculating upper and lower control limits for skewed or asymmetric populations. When the population is symmetric, these charts are reduced to the Shewhart control charts. Choobineh and Ballard (1987) conclude that the weighted variance method gives the same limits as the Shewhart method in the case of a symmetric underlying distribution. When
316
Encyclopedia and Handbook of Process Capability Indices
the underlying distribution is skewed, however, the new limits are adjusted in accordance with the direction of the skewness. The study of Bai and Choi (1995) also indicates that for symmetric populations these control charts are equivalent to the Shewhart charts. However, when the underlying population is Weibull or Burr, the "heuristic charts" are found to be superior over Shewhart charts especially as the skewness increases. Bai and Choi (1997) have utilized the weighted variance method to adjust capability index values in order to account for the degree of skewness of non-normal process data. This technique computes separately the standard deviations above and below the mean. The Cp and Cpk indices based on the weighted variance method, C^v and C^, are defined as: C?v
=min\
USL - LSL 6axy[2Px
USL - LSL -mm 6cr T
and
^Pk
— min-j
USL - LSL 6^V2(1 - Px) 1
1
m^2{\-px)\
USL — nx 3ax^2K'
nx — LSL 3axj2(l-Px)
where Wx = 42 max (4TX, 41 - Px ) = J1+ \1-2PX\ Px=Pi(X<^).
where
Similarly, the Cpm and C*m indices (defined by Chan et al. (1988)) can be modified as CpVf = min •
USL - LSL USL - LSL 6aTj2fy ' 6aTj2(l - PT)
USL - LSL -min V 2 ^ ' 6(7T
c,pm yl2(l-PT)
Wr
Process Capability Assessment for Non-normal Processes
and
Cpm
= min-
USL-T 6<JTJ2F£,
317
T-LSL §oTJ2(l-PT)y
where, now PT = P r ( X < T ) , WT = Jl+ | 1 - 2PT | and aT — yjal + (n — T)2 . To evaluate these indices, the parameters \ix , ax and
318
Encyclopedia and Handbook of Process Capability Indices
a similar formula to the one used to calculate the sample variance in this situations as: 2^/(Xl-Xf Si = - ^ 2nx - 1
2^(Xt-Xf and Sf = -i=i ' 2«2 - 1
On the other hand, the sample standard deviations from the target T, STi and ST2 , can be expressed by the following equations using the Taguchi's unbiased estimator ( ST = Y^l=l(Xi -Tf /n) as given in Boyles (1991):
2Yj{Xi-Tf S k = - ^
2
=
2ni
-^Sl+{X-T)\
2rii
2J2(X1-Tf
and
Sh- «
=^SUtX-T?.
For the original weighted variance method (as used by Bai and Choi (1995)), the expression for Sf , 5 | , STi a n d ST2 can be modified to be:
5X2(B&C)
E(^-^)2
= 2(1 - Px)^
lb
E(^-^)2
,Sf (B&C) = 2PX ^
—-
i
X
E(^-r)2 $ i ( B & C ) = 2(1 - PT)-1^ n n
E(^-T) 2 ST2 (B&C) = 2P T i=i
l
l
.
Process Capability Assessment for Non-normal Processes
319
where, as above, Px = P r ( X < /x) and PT =?x{X
=
UoL — LSL
fiwv
3(5 +s)
pk
= mm
USL-X
X - LSL
002
ob\
Moreover, the denominator ST in the estimators Cpm and Cpmk is replaced by Sn and ST2 using the expressions of the WV method. Then, the estimators C^ and C 1 ^ c a n be expressed as:
^wv
. \USL-T
C„ = mm — J m pm I
and
C^k = min
OOx2
USL-X 0&T2
T-LSL]
,——
OOTI
)
X - LSL ODyj
The advantage of using the WV method to modify normally-based process capability indices is that the weighted variance-based capability indices reflect the values of the skewness and kurtosis of a distribution, since their sample standard deviations, Si and S2 ,
320
Encyclopedia and Handbook of Process Capability Indices
and the sample standard deviations from the target, S^i and ST2, are adjusted by means of different skewness and kurtosis values. Moreover, the samples of size nx and no. vary as the skewness and kurtosis undergo changes. On the other hand, when the distribution is symmetric, S\ and S2 are equivalent to S and so are Sx\ and 5 , r 2 , while n\ and no, are approximately equal to n / 2 . Consequently for symmetric distributions the weighted variance-based capability indices turn out to be the original process capability indices. 15.2.12
The Weighted
Standard
Deviation
(WSD)
method
Chang and Bai (2001) suggested a WSD method to construct control charts in accordance with the degree of skewness of the underlying population. Similarly to the WV method (Section 15.2.1) the WSD method is based on the idea that ox can be divided into upper and lower deviations, ofj and oY , which represent the degrees of the dispersions of the upper and lower sides from \ix , respectively. An asymmetric p.d.f. f(x) can be approximated by two normal p.d.f. s
fu(x) = irw Zov
Vx
2ov
and
fdx) =
W
4,
%
l-t>x
~2aT
with the same mean \ix but different standard deviations lay and loY i i-e-j t n e upper and lower sides of f(x) are approximated by the upper side of fu(x) and the lower side of fi(x) , respectively. The WSDs off and oY are obtained as o$ = Pxox and oY = (1 — Px)ox , where, as above, Px = P r ( X < \ix) . Details can be found in Chang and Bai (2001).
Process Capability Assessment for Non-normal Processes
321
The WSD method can be used to adjust the values of PCIs according to the degree of skewness of the underlying distribution by using different standard deviations when computing the upper and lower PCIs. Namely, 2ajj and 2oY can be used to construct PCIs for skewed populations. The Cp based on the WSD method, CfSD, is defined as (see, e.g., Chang et al. (2002)): C™D = min
~
USL - LSL 6-2^
USL - LSL '
. [ USL -LSL l Q-2Pxax '
m m
_ USL - LSL Qax
6-2af USL - LSL 6-2(1-Px)ax
. |J_ 1 \ _CP l 2 P / 2(1 - Px)\ ~ D,'
m m
where Dx = 1+ | 1 — 2PX \. In this definition, the quantities 2ojf and 2aY are used in place of ax to reflect the degree of skewness. If the underlying distribution is symmetric, Px = 0.5 and Dx = 1.0 , i.e. C^SD = Cp . For skewed distributions, however, Dx > 1.0 and thus C^0 < Cp. Chang et al. (2002) have also considered the situation when only a single specification limit exists. In this case using the WSD method, the upper and lower capability indices stemming from Cpk are defined as: n
WSD pku
_ USL — [ix _
3 • 2aJf
USL — [ix 6PX"xaX
and riWSD _ Vx ~ LSL _ \ix — LSL Pkl
O-W 3Q • 2<
6Pxax
The capability index for the case of two-sided specifications becomes
322
Encyclopedia and Handbook of Process Capability Indices
Cpk
f USL — fJLx fix — LSL
- min(C p f a , Cpkl ) - m m |
Qp^
,
_
For a symmetric population, CJiJt'51' reduces to Cpj. , and whenever \ix is located at the center of the two specification limits, C^kD reduces to CfSD . To apply these indices in practice, \ix, ax , and Px ought to be estimated. Let a random sample Xx, X2 , ..., X„ of size n be available, /i^ and ax can then be estimated by the sample mean X and the sample standard deviation Sx . Since Px is the probability that X will be less than or equal to \xx (i.e. Px — Pr (X < /J,X ) ) , it can naturally be estimated by using the number of observations less than or equal to X : 1 "
n J=I ., where the characteristic function I(x) = 1 for x > 0 and I(x) = 0 for x < 0 . Thus, the proposed PCIs C\fD and Cf OT can be estimated as WSD
C.
USL — LSL 6 • DXSX
and ^Pk
— rnin|Opfo, , o p H j — m m
USL6PXSX
X
X - LSL 6(1 - PX)SX
where D, = 1+ I 1 - 2PIt is well known that Sx underestimates ax , hence the estimate of a PCI overestimates the true value, and the bias correction factor will be (as noted in earlier Chapters):
Process Capability Assessment for Non-normal Processes
323
r[(n-l)/2]
^ = £?Tr [ ( n - 2 ) / 2 ] which is applicable whenever n is small. Note that b(n) is obtained under the normality assumption, and thus cannot be directly used for skewed populations. Therefore, a bias correction factor that also reflects the degree of skewness ought to be determined for skewed populations. The WSD method approximates an asymmetric distribution by means of two normal distributions, and it is assumed that nPx observations among the n observations are obtained from the lower side of the p.d.f., fi{x) , and n(l — Px) observations from the upper side of fu(x) . Hence, if 2nPx and 2n(l — Px) observations are viewed as random samples from fL (x) and fv (x), respectively, the bias correction factors bYSD = b(2nPx) and ^WSD _ fr(2n(l — Px)) could be used to reduce the bias. Using these correction factors, the estimated WSD PCIs finally become: nWSD
USL — LSL
r<WSD _ ___•
^pk
WSD
WSD
2PX
2(1-PJ
[CWSDfiWSD
— mill 10(7
Opfoj , 0L
iWSDr<WSD\
Opfc;
|,
where bfSD = b(2nPx) and bJ[SD = 6(2n(l - Px)). As Px increases, so does bYSD, while bjjSD decreases. Consequently the skewness is reflected in the correction factors. In addition, the performance of these WSD PCIs compared with the various PCIs, was analyzed, and finite sample properties of these estimates were also investigated by Chang et al. (2002) utilizing Monte Carlo simulations. Numerical results show that substantial improvements over the existing methods can be achieved by using the WSD method when the underlying distributions are skewed.
324
Encyclopedia and Handbook of Process Capability Indices
Ahmed and Abdul Kader capability index
(2004) consider the
following
„ _ allowable process spread or tolerance actual process spread or variation where, as above, LSL = lower specification limit and USL = upper specification limit and o is the standard deviation associated with the process. Let R be the specification range, i.e., R = USL — LSL. The formula above can be rewritten as Ct=^,
to
o>0,
where t is a positive constant. This constant is called the tuning factor and is chosen so that the probability Pr{^i — to < Y < JJL + to} will be close to 1. This index is less sensitive to departure from normality. For t — 6 , Ct = Cp. The number 6 associated with the assumption of a normal distribution for Y , provides a satisfactory approximation. For t = 6 , a value of Ct greater than or equal to 1 implies that the tolerance band (USL — LSL) is at least 6a of the process distribution. If \i is the location parameter of the process population, then for t = 6 , by the Chebyshev inequality at least 89% of the products are within the specification limits. On the other hand, if the process distribution is normal, then by the properties of the normal distribution approximately 99.73% of the products will be within the specification limits. Pearn et al. (1992) discuss the problem of dealing with the above index whose interpretation (unlike that of the other standard indices such as Cp). ) is less sensitive to departures from normality. The capability index as defined above, is a multiple of Cp and has a quite similar interpretation provided the samples are drawn from a normal population. For example, for t = 5.15 , C5.15 = 1.17Cp . Thus, one can paraphrase the interpretation of Cp as given by Ekvall and Juran (1974) and assert that the values of C 5 1 5 less than 1.17 should be classified as inadequate, the values between 1.17 and 1.56 are designated as
Process Capability Assessment for Non-normal Processes
325
adequate, and the values greater than 1.56 are considered to be even better than adequate. However, a practical interpretation of C515 is not fully apparent for data populations that are substantially skewed.
Chapter 16
Multivariate Process Capability Indices
16.1 Introduction 16.2 Multivariate PCIs 16.3 Concluding remarks 16.1 Introduction This Chapter deals with some more recent developments in the field of Process Capability when several characteristics of a process are involved. For completeness we shall present again some basic facts related to the subject matter under consideration, in our attempt to render this Chapter to be as self-contained as possible. Various quality measures have been proposed to evaluate a process's performance. In general, there are several issues to be considered when assessing product quality. First and foremost is the issue of process capability indices. A process capability index (PCI) is a numerical summary that compares the behavior of a product or processes characteristics related to engineering specifications. Its convenience is due to the reduction of complex information about the process to a single number. Specifically, process capability is universally defined as the range over which the measurements of a process vary when the process variation is due only to random causes. In this case, process capability indices do provide an effective measure of process capability. By using a process capability index, we may obtain a conclusion whether the product meets its specifications and is stable. Currently, numerous 326
Multivariate Process Capability Indices
327
customers request their suppliers to record capability indices for product characteristics on a regular basis. In the majority of companies it became a key index to evaluate product quality. A large number of quality engineers and statisticians have proposed methodologies for assessing product/process quality. However, the bulk of the studies associated with analyzing the quality and efficiency of a process, are so far limited to discussing one single quality specification. In most processes, the products possess multiple quality characteristics rather than a single one. Multiple characteristics processes are by now so common that our studies to capability indices can't be restricted to the univariate domain. The multivariate relationship among the quality characteristics may or may not be reflected in the engineering specifications. For instance, USL 's and LSL 's may be given separately for each quality characteristic. In two-dimension cases, these tolerance ranges compose a rectangular tolerance region. In higher dimensions, they form a hyper cube. For more complex engineering specifications, the tolerance region could be quite complicated. For processes with multiple characteristics, Bothe (1992) proposed a simple measurement of a tolerance region by taking the minimum of the measure for each single characteristic. Consider a v -characteristic process with v yield measures (percentage of conformities) Pu P 2 , ..., and v . The overall process yield is measured as P = minjPx, P 2 , ..., u}. However this approach does not accurately reflects the real situation. Suppose the process has five characteristics ( v = 5) with equal characteristic yield measures P1 = P2 = P3 = Pi = Pt = 99.73%. Using the approach considered by Bothe (1992), the overall process yield is calculated as P = min{i\, P2, P 3 , P 4 , P 5 } = 99.73% (or equivalently 2700 PPM of non-conformities). Assuming that the five characteristics are mutually independent, the actual overall process yield should be calculated as P = Px x P 2 x ... x P 5 = 98.66% (or 134273 PPM
328
Encyclopedia and Handbook of Process Capability Indices
of non-conformities), which is significantly less than calculated by Bothe (1992) method. When these variables are related characteristics, the analysis ought to be based on a multivariate statistical technique. Chan et al. (1991), Taam et al. (1993), Pearn et al. (1992), Chen (1994), Karl et al. (1994), Shariari et al. (1995), Boyles (1996), Wang and Du (2000), Wang et al. (2000) and others have developed and presented multivariate capability indices for assessing capability. Wang and Chen (1998) and Wang and Du (2000) proposed multivariate extentions for Cp , C^ , Cpm and Cpmk based on the principal component analysis, which transforms numbers of original related measurement variables into a set of uncorrected linear functions. A comparison of three novel multivariate methodologies for assessing capability are illustrated (and their usefulness is discussed) in Wang et al. (2000). 16.2 Multivariate PCIs Some basic knowledge of the classical multivariate analysis based on multivariate normal distribution is required to fully appreciate the theory behind the Multivariate PCIs. For the multivariate case, it is even more dangerous to use a single number as a measure for the capability of process than in the univariate case. In such cases, it has been usually assumed (up till very recently) that the observations X posses a multivariate normal distribution Nv([i, E) , where v is the dimension of the variables, fi is the mean vector and E represents the variancecovariance matrix of X . Chan et al. (1991) introduced a version of the multivariate index Cpm . They have considered measurements of a set of v characteristics Xi, x2, ,xv . For specification region, all points lying on or within the specification boundary satisfy the inequality (X -T)'A-\X
-T)
< c2,
(16.1)
Multivariate Process Capability Indices ./
__
r r r
,
_ ,
329
j
where X = [x\, ,xv] , T = [Tlf ,TV] and A is a « x « positive define matrix. Here T and ^4 completely determine the shape and center of the ellipsoid, while c determines the coverage of the ellipsoid. The coverage is the percentage of observations expected to be within the boundaries. Chan et al. (1991) proposes a multivariate process capability analog of Cpm defined by Cpm
'^[(Xi-TyA-^Xi-T)]'
(16 2)
'
For a known A, an unbiased estimator of the denominator of C2pm is
\Y.{Xi-TfA-\Xi-T), n .,
and a natural estimator of the multivariate Cpm becomes: r
_ ,
nv
fciXi-Tj'A-^Xi-T) *=i
where Xt denotes the i th vector of dimension v ( v is the number of variables used in assessing capability), n is the sample size, T is the target value and A is a matrix representing the covariance structure determined from the specification limits. The numerator of Cpm is the product of the sample size n and the numbers of variables v (used in evaluating capability). This value indicates the degrees of freedom associated with the denominator representing the sum of the observed the so-called Mahalanobis distances from the target. For processes with observations clustered around the target, the denominator of Cpm will be smaller in magnitude than that of the same process, with more scattered and with a center of mass located not at the target value. For a fixed nv observations, the value of Cpm will diminish when
330
Encyclopedia and Handbook of Process Capability Indices
the denominator grows. Smaller values of Cpm indicate that the process is unable to meet the specifications, while larger values indicate that the process is indeed capable of meeting its specifications. Based on equation (16.2), Pearn et al. (1992) have introduced two multivariate PCI which are viewed to be more natural generalizations of Cpm proposed by Chan et al. (1991). Setting H = T and £ = A , the value of c2 in the specification requirements leading to the proportion 0.0027 of non-confirming items becomes the upper 0.9973 quantile of a chi-squared distribution with v degrees of freedom. This is a basic property of the multivariate normal distribution presented in any text on multivariate analysis (e.g. Johnson and Wichern (2004)). Indeed: let T2 = XTS^X, where X and S are independently distributed random variables with X ~ Nv([i,Yl) and fS ~ Wv(/,]C), / > p • Then XT £ ~ : X ~ X2(P,T2) with r 2 = fiT £ ~ 1 fi . Denoting this 2 2 value by c (c = xi 0.9973 ) one defines 2 VC P
=4
(16.3)
cv as a generalization of Cp where c2 is determined by (16.1). Analogously one can define C:2 vCpm 1
<ji-T)'A-\n-T) v
as a generalization of Cpm , where Vol. ((X - ii)'A-\X v
p
Vol. ((X - n)'A~\X
- /x) < K2) - M) < xl 0.9973)
(16 4)
"
331
Multivariate Process Capability Indices
For /J, = T , E = A and „ C2p = 1, the expected proportion of NC items is 0.0027 as in the univariate case. It should be noted that for [i •*• T and E ^ A, the value cl satisfying P[(X - T)'A-\X
-T)
= 0.9973
— under the multivariate normal assumption — depends on /i, T, E, and A in a rather complicated manner and /z and E ought to be estimated. Taam et al. (1993) proposes a multivariate capability index defined as a ratio of two volumes n MCpm
_ Vol. (Rx) _ Vol. (modified tolerance region) ~ Vol. (ifc) _ Vol. ((X - p)W(X -fi)< K{q))'
, (
. j
where R1 is a modified tolerance region (to be defined below) and Rq is a scaled 99.73% percent region. Rq is an elliptical region under the normality assumption. A modified tolerance region is the largest ellipsoid that is centered at the target completely within the original tolerance region. When the process of multiple characteristics is in statistical control so that the process mean is on target and the process variation is small relative to the engineering specifications, the multivariate MCpm should be close to 1. An estimator for MCpm can be expressed as: Q
MCpm = -J- , where G„v —
Vol. (tolerance region) Vol. (estimated 99.73% process region)
(16.6)
332
Encyclopedia and Handbook of Process Capability Indices
Vol. (tolerance region) ISI 1 / 2 (7rK(v)y/2[T(v J2 + l)]- 1 ' and
77
-
,
-
l1/2
D = 1 + — — (X - T) S-\X - T) n —1 Note that | S | 1 / 2 (?rii'(7j))''/2[r(7j/2 + l)]^ 1 is a generalization of the volume of a sphere and from elementary geometry where K(v) is the 99.73% percentile of the x 2 distribution with v degrees of freedom and |-| denotes the determinant. When the process mean vector is equal to the target vector and the index has the value 1, then 99.73% of the process values lie within the modified tolerance region. The numerator of Cp is analogous to Cp in the univariate case. Namely value of Cp greater than 1 implies that the process has a smaller variation than it is allowed by the specification limits with a certain confidence level; on the other hand, a value less than 1 implies a larger variation. Analogously, 0 < 1/ D < 1 measures the closeness between the process mean and the target; a larger 1/ D would imply that the mean is closer to target. Chen (1994) proposes a multivariate process index using the concept of a tolerance zone. A tolerance zone is expressed as V = {XeRv
: h(X - ^) < r0},
(16.7)
where h(-) is a nonnegative homogeneous scalar function satisfying the condition h(tx) = th(x) for all t > 0 and r0 is a positive number. A process is considered to be capable if P(X e V) > 1 - a , i.e., P(h(X - /x0) < r0) > 1 - a , where a is the allowable expected proportion of nonconforming production from a process (conventionally, a — 0.27% (about 1/4 of percent)). Let now r = min{c : P(h(X — fi0) < c) > 1 — a} . It should be noted that if the c.d.f. of h(X — /j,0) is increasing in a neighborhood of r , then r is simply the unique root of the equation P(h(X - /i0) < r) = 1 - a. The process is deemed to be
333
Multivariate Process Capability Indices
capable if and only if r < r0 , that is, (r0 / r) > 1. This allows us to express the multivariate process capability in the form MCV=^.
(16.8) r
Here, the value of MCP corresponds to the expected proportion of conforming products of 1 — a . Namely, MCp = 1 indicates that the expected proportion of nonconforming products is exactly a . The index MCp is well-defined, since a tolerance zone V (Equation (16.7)) is specified by the function h uniquely up to a scale constant and MCP is scale-invariant. The definition of MCp provides the following advantages: (i) the specifications are general as given by V (ii) there is no assumption about the underlying particular distribution, such as the normal one, and (iii) the arbitrariness of a permits flexibility in setting a criterion for the capability of a process. In the case when the function h is of the form: h(X -T)
= [(X - TJA~\X
-
T)fl2
with a v x v positive definite matrix A, V (the tolerance zone) is an ellipsoidal specification given by { X e Rv : {X - T)'A-\X
-T)
}.
where r0 determines the coverage of the ellipsoid. For this tolerance zone and a = 0.0027 , MCp reduces, under the normality assumption to the multivariate PCI proposed by Pearn et al. (1992). Setting ft = T and E = A , MCp becomes in this case. r0 /(o"Xu,a) > (where Xv,a is the 100(1 — a ) t h percentile of the X2 distribution with v degrees of freedom). Here a2 =(n-l)-1J2ni=l(X-X)'A~\X-X) so that MCp can be estimated by r0 /{aXv,a) •
334
Encyclopedia and Handbook of Process Capability Indices
Furthermore, considering that a rectangular solid tolerance zone: V = {XeRv:\Xi-^ii\
,i = 1,
2,...,v},
where /% is the process mean f i = 1, 2,...., v) and rt is a positive constant, we arrive at the multivariate process index MCp 1 / r , where r is such that: P(max{|Xj -/Mal/n,
i = 1, 2,....,v} < r) = 1 - a.
Let F be the c.d.f. of the variable h(X - fi0) = max{|Xj - ^ o l / n , * = 1, 2,...,«}, Setting r = F _ 1 ( l — a ) , it follows immediately that for any y > 0 F(y) < mm{P\Xi
-fio\/rty,
i = 1, 2,...,v}.
If the value of MCp is greater than or equal to 1, the process is capable at a certain confidence level. Thus, a necessary condition for a process to be capable over a rectangular solid zone is that each individual univariate process is capable with the corresponding specification limits. This seems to be an appropriate condition. Indeed, this fact is the starting point of the Kocherlakota and Kocherlakota (1991) bivariate generalizations of Cp . Karl et al. (1994) introduce applications of multivariate capability index in geometric dimensioning and tolerance (GDT). Their paper presents three cases of GDT drawings of the same product and the corresponding capability indices evaluations. Shariari et al. (1995) proposes a multivariate capability vector, consisting of three components. The first one is a ratio of areas or volumes, analogous to the ratio of lengths in the univariate Cp .
335
Multivariate Process Capability Indices
The numerator is the area (in the two dimensional case) or the volume (in three or more dimensions) defined via the engineering tolerance region. The denominator is the area or the volume defined by the "modified process region", which is the smallest region, similar in its shape to the engineering tolerance region, circumscribed about a specified probability contour. The number of dimensions of the process data is accounted for by taking the u t h root of the ratio (see also Wang et al. (2000)). Thus, the first component, labeled CPM , is defined as Vol. of engineering tolerance CPM
Vol. of modified process region
1/v
(16.9)
This approach yields a "modified process region" by drawing the smallest rectangle around the ellipse. The edges of the rectangle are defined as the lower and upper process limits {LPLt and UPLi where i = 1, 2,..., v) and are determined by solving the equations of first derivatives, with respect to each xt , of the quadratic forms such as (X-Lio)'m-l^) = xla,
(16.10)
where X is a v x n sample matrix, v being the number of product quality characteristics measured on a given part and n is the number of parts measured (see also Nickerson (1994)). Namely, each column in the matrix represents the v measurements recorded for a sampled part. These n observations are assumed to be independent and represent a sample drawn from a multivariate distribution with a correlation among the v variates. The two solutions to this equation, for each dimension, provide the upper and lower limits
UPLi = m+ A
—,-—
V det^" 1 )
336
Encyclopedia and Handbook of Process Capability Indices
and
LPLi = ^
det^" 1 )
'
, i = l, 2,...,v,
(16.11)
where Xv,a is the upper a quantile of a \2 distribution with v degrees of freedom associated with the probability contour and det(S^ 1 ) is the determinant of E^ 1 (a matrix obtained from E^ 1 by deleting the ith row and column). As a result, the CPM can now be expressed as i/t)
HiUSL, - LSh) CpM —
i=\
(16.12)
Y[(UPLt - LPL,) j=i
If this value is greater than 1, this indicates that the circumscribed "modified process region" is smaller than the engineering specified region, namely the part is "acceptable". Evidently, the modified process region is influenced by the shape of the elliptical contour (represented by the variance-covariance matrix) and the size of the contour (based on the chosen probability level). The second component of the Shariaried al.(1995) index is the significance level of the observed value with the Hotelling's T2 statistic T2 = n(X - n)'S-l{X
- fi)
(16.13)
expressed as PV = Pr
F
>
n ~ v T2 v(n — 1)
(16.14)
Multivariate Process Capability Indices
337
where Fv
if modified process region is contained within the tolerance regions
LI = 0
otherwise.
Consequently, the multivariate capability vector has three components, [ CPM , PV, LI } providing a comparison of the volumes of the regions, locations of the centers and location of the regions. Wang and Du (2000) proposes an useful method using the principal component analysis (PCA) for describing the process performance for a multivariate data. As a rule, multivariate PCIs assume the multivariate normal distribution. But for an application of the PCA this assumption is not required. Advantages of this approach are that the principal component analysis is able to not only to transform high-dimensional problems into lower dimensional ones, but also can transform correlated data into independent data, which simplifies calculations of a multivariate PCI. In addition, the corresponding confidence intervals are also derived. We shall expand on the structure of this method below: Assume that X is a v x n sample data matrix, where as usual in this chapter v is the number of product quality
338
Encyclopedia and Handbook of Process Capability Indices
characteristics from one part and n is the sample size of the measured part. X is the sample mean vector ( v x 1 ) of the observations and S is a v x v symmetric matrix representing the covariance between the observations. Engineering specifications are given for each quality characteristic. Here LSL and USL are the v vectors of the lower and upper specification limits, respectively. The T vectors (v xl) represent the target values of the v quality characteristics. By using the spectral decomposition theorem, we can obtain a matrix D = UTSU , where D is a diagonal matrix. The elements of D , Aj, A2, A„ , are the eigenvalues of S and the columns of U , ux, it?, •••,uv are its eigenvectors. Consequently, the i th principal component (PCi) is expressed as PCi = uTx, i = 1,2,...,«,
(16.15)
where x s are u x l vectors of the original observations on the variables. The engineering specifications and target values of PCt s are LSLPC. = ujLSL • USLpC. = ufUSL,
i = 1, 2, ...,v ,
(16.16)
TPCi = ujT. Similarly, the relevant sample estimators, S and X , of PC^s can defined as SpCi
XpCl = ufX,
=
\)
i = 1,2,...,«.
It is known (see e.g. Johnson and Wichern (2004)) that the ratio of each eigenvalue to the sum of the eigenvalues is the proportion
Multivariate
Process
Capability
Indices
339
of variability associated with each principal component variable. Namely, the proportions of variability are Ai/j>,
Vi = l,2,...,V.
(16.17)
j=i
Usually, only a few principal components contribute to the bulk of the total variability (about 70%~90%). By using just this subset of the components, the dimension of the multivariate quality characteristic problem can be reduced. Jackson (1980) proposes a X2 test for identifying the significant components. Using this method, we can easily choose the suitable number of PC^s. There are of course more recent method to identity significant components (see, e.g. Johnson and Wichern (2004)) but Jackson's method seems to quite straightforward. Subsequently in the paper, the procedure of obtaining multivariate capability indices for the normal data (as well the non-normal data) is described in the following manner (see Wang, and Du (2000)). a. Multivariate normal data Define
MCV
m
MCpk
(16.18)
P;PC,
«=i I v
and
l/v
l/v
C
Yl Pk\PCi Vi=l
where ~
a
vk\PCi
USLPC. — LSLPC.
min[USL PCi - XPCi, XPCi - LSLPCi }
(16.19)
340
Encyclopedia and Handbook of Process Capability Indices
Hence an approximate 100(1 — a)% confidence interval for MCP is —
\l/v
Xl-g/2,n-l
IK
p-PC,
n-1
i=l
<MCp
<
n
l/v
A
jXa/2,n-l
(16.20)
Analogously an approximate 100(1 — a)% confidence interval for MC„i. turns out to be l/v
1 1 Cpk;PCi
1-Z,
all
1 1 9nC.pk;PC-t + 2(n - 1)
< MC,p i l/u
<
C
Yl pk;PC, l + Za / 2
1 1 2(n - 1) 9nC,pk\PCi- +
(16.21)
where in (16.20) X^-i, a / 2 a n d X^-i,i- Q /2 a r e t h e u P P e r a / 2 a n d the 1 — a j 2 quantiles of a chi-squared distribution with n — 1 degrees of freedom, ZQ/2 is the upper a/2 quantile of the standard normal distribution and the Hilfiey-Wilson approximation to the x2 distribution is utilized (c./. Section 3.2.3). The estimators Cpm-pct and Cpmk.PCi can be developed in a similar fashion. b. Non-Multivariate normal data As it was mentioned earlier, Luceno (1996) proposed a process capability index, Cpc , designed to account for both the process location and spread, as well as to provide for confidence bounds that are insensitive to departures from a normal data. Wang and
Multivariate Process Capability Indices
341
Du (2000) suggest that for this type of data the process capability could be estimated as l/v
(16.22)
MCVc = [ [ Cpc-Pd
where Cpc.
C.pc;Pd
jM
< MCpc < 1
+ tcn-iSci / (CiJii) d <
l/v
C, ^pc;FCj
rr
(16.23)
where
and
_
1 "
Sn. =
1 n-1
PC;,
(USLPCi
E rL/ii
—
+LSLPCi)
(USLPCi
+LSLPCi) nc:
Here tan_\ is the upper a / 2 quantile of the Student t distribution with n — 1 degrees of freedom. Wang et al. (2000) compare the three process capability indices: a multivariate capability vector [ CpM , PV , LI ] proposed by Shariari et al. (1995), a multivariate capability index MCpm proposed by Taam et al. (1993), and a multivariate capability index MCp proposed by Chen (1994). They conclude that in general, the multivariate indices can be obtained from (a) the ratio of a tolerance region to a process region (b) the probability of the nonconforming product, and (c) other approaches using loss functions or vector
342
Encyclopedia and Handbook of Process Capability Indices
representation. The purpose of Wang et al. (2000) paper is to illustrate the distinctions among the various meanings of capability in the multivariate domain. Through the use of several graphical and computational examples, the information summarized by these methodologies is illustrated and their usefulness is discussed. 16.3 Concluding remarks It is evident that in the multivariate case specification limits and the actual process spread is more difficult to define than their univariate counterparts. A difficult issue involving multivariate PCI is also the interpretation of the modified tolerance transferred from the tolerance region. This is an important issue and worth discussing. In general, the specification region is assumed to be ellipsoidal in most multivariate process capability indices due to the assumption of the multivariate normality. Chen (1994) proposes a multivariate process index over a tolerance zone, which makes the definition of the engineering specifications more general and does not require a particular distribution. However, one should note that if the observation are not from the multivariate normal distribution, calculations of these multivariate PCIs will be quite involved and even may not be obtained. Looking back at the Cpm as defined in equation (16.2), this capability index considers proximity to the target and dispersion around the target while will retaining some reasonably attractive statistical properties. Using these statistical properties, one is able to carry out a further analysis and to test (using a hypothesis testing) whether the process capability is capable. Here a benchmark of 1 for assessing the process is suggested. However, we know that the process should be capable but can't as yet determine what the nonconforming percentage will be when this index is equal to 1. The two multivariate PCIs, mentioned in equations (16.3) and (16.4) respectively, seem to be more natural generalizations of the univariate Cp and Cpm. When T = \x and
Multivariate Process Capability Indices
343
A = £ , equations (16.3) and (16.4) are equivalent, and the value 1 in equation (16.3) indicates that the nonconforming portion in the process is 0.27%. The MCpm presented in equation (16.5) has the same property as a univarite PCI, when the process is centered at the target and the capability index equals 1, it should indicate that the 99.73% of the process values lie inside the tolerance region. The definition of engineering specification of MCp as presented in (16.8) provides some flexibility in setting a criterion for the capability of a process. When the values of MCP is 1, it corresponds to the expected proportion of conforming product being 1 — a . Recall that the three components of a multivariate capability vector, CPM , PV and LI, provide a comparison of volumes of the regions, locations of the centers, and location of regions. These generate additional versions of the process capability indices. When the Cpy[ is higher than 1, it implies that the circumscribed modified process region is smaller than the engineering specified region, that is, the index is "capable". The value 1 of PV indicates that the center of the process is close to the engineering target value. And finally the value 1 for LI implies that the modified process region is contained within the tolerance region. Wang as we have mentioned above, and Du (2000) propose an application of the PCA to the process capability indices for multivariate data. This approach does not require the assumption of normality. It reduces the dimension of the data and also transfers correlated data into independent data. Consequently calculation of the indices becomes much easier. Although the use of PCA when calculating the multivariate process indices may result in some loss of information, it is nevertheless an adequate approach for assessing the multivariate process capability. No index is the best to cover all the aspects of the structure of multivariate process capability indices. Some methods may possess letter properties than others in some fields, but the others may
344
Encyclopedia and Handbook of Process Capability Indices
have good properties that are superior to those enjoyed by the original method. To summarize, the proposed approaches all have some reasonable qualities, but we still can't obtain a fully consistent method for assessing the multivariate PCIs. This requires a concentrated further study of the issues related to multivariate PCIs. Up untie very recently, the research in the theory and practice of multivariate process capability indices has been very sparse in comparison to the research dealing with the univariate case. At present, for the multivariate capability indices, consistency in the methodology for evaluating this capability is still absent. Moreover, it is quite difficult to obtain the relevant statistical properties needed for a more detailed inference about multivariate PCI. Moreover there are, of course, still intrinsic difficulties in trying to assess the value of multivariate systems by mean of a single number. Obviously, further investigations in this field are strongly desirable. This completes our surrey of the process Capability Indices. (up dated up to 2004). The dynamic area-inspite of obstacles has grown tremendously during the 20 years of its existence. The unprecedented technological advances, we are witnessing at present, will no doubt provide a fertile grown for new developments in the theory, methodology and application of the PCIs.
Bibliography
Ackermann, C. S. (1993). Supplier improvement via SPC applications workshops. IEEE Transactions on Semiconductor Manufacturing, 6(2), 178-183. Adler, Ya. and Shper, V. (1994). Some remarks on Bayesian approach to process capability estimates. 5 International Meeting on Bayesian Statistics, Alicante, Spain. Ahmed, S. E. (1998). Test and confidence intervals for process capability indices in k arbitrary populations, MS Thesis, Department of Mathematical Statistics, University of Regina, Saskatchewan, Canada. Ahmed, S. E. and Rohbar, M. H. (1997). Inference concerning a process capability index. MS Thesis, Department of Mathematical Statistics, University of Regina, Saskatchewan, Canada. Ahmed S. E. and W. Abdul- Kader (2004). Inference Procedures for Process Capability Indices in Non-Normal Distribution From Various Sources. Technical Report, Dept of Math and Stat., University of Windsor, Ont. Canada. Anjard Sr, R. P., Hagerty, D., Griffith, G. K., Liu, S. T., Mustonen, E. M. and Pasfield, D. A. (1991). Cpj. applications - uses and abuses. Microelectronics Reliability, 31(6), 1123-1125. Bai, D. S. and Choi, S. S. (1997). Process capability indices for skewed populations. MS Thesis, Dept. Industrial Engineering, Advanced Institute of Science and Technology, Taejon, South Korea. Banks, D. (1993). Is industrial statistics out of control? Statistical Science, 8(4), 356-377. Banks, D. (1993). Is industrial statistics out of control? Rejoinder. Statistical Science, 8(4), 402-409. Bar, R. (1999). Statistical evaluation of stability data for assessing variability of an analytical method. Accreditation and Quality Assurance, 4(6), 235-239. Barnett, N. S. (1996). Process control and product quality: the Cp and Cpk revisited. International Journal of Quality and Reliability Management, 7(5), 34-43. 345
346
Encyclopedia and Handbook of Process Capability Indices
Bates, T. E. and English, J. R. (1991). The robustness of modern process capability analysis. Computers and Industrial Engineering, 21(1-4), 45-49. Behara, R. S., Fontenot, G. F. and Gresham, A. (1995). Customer satisfaction and analysis using six sigma. International Journal of Quality and Reliability Management, 12(3), 9-18. Benneyan, J. and Kaminsky, F. (1995). Answers regarding process capability index Cpk. IEE Solutions, 27(7), 8. Bernardo, J. M. and Irony, T. Z. (1996). A general multivariate Bayesian process capability index. The Statistician, 45(3), 487-502. Besterfield, Dale H. (1999). Total Quality Management, Prentice Hall, 2nd edition. Bissell, A. F. (1990). How reliable is your capability index? Journal of the Royal Statistical Society, C 39(3), 331-340. Bittanti, S., Lovera, M. and Moiraghi, L. (1998). Application of non-normal process capability indices to semiconductor quality control. Semiconductor Manufacturing, IEEE Transactions on Reliability, 11(2), 296 -303. Bordignon, S. and Scagliarini, M. (2002). Statistical analysis of process capability indices with measurement errors. Quality and Reliability Engineering International, 18(4), 321-332. Borges, W. and Ho, L. L. (2000). On the sampling distribution of Clements' capability index. Communications in Statistics: Simulation and Computation, 29(1), 121-138. Borges, W. and Ho, L. L. (2001). A fraction defective based capability index. Quality and Reliability Engineering International, 17(6), 447-458. Bothe, D. R. (1992). A capability study for an entire product. ASQC Quality Congress Transactions, Nashville Tenn., 46, 172-178. Bothe, D. R. (1997). Measuring Process Capability. McGraw-Hill, New York, N.Y. Bothe, D. R. (1999), A capability index for multiple process streams. Quality Engineering, 11(4), 613-618. Bothe, D. R. (2002). Statistical reason for the 1.5CT shift. Quality Engineering, 14(3), 479-487. Boyles, R. A. (1991). The Taguchi capability index. Journal of Quality Technology, 23(1), 17-26. Boyles, R. A. (1994). Process capability with asymmetric tolerances. Communications in Statistics: Simulation and Computation, 23(3), 615-643. Boyles, R. A. (1996). Exploratory capability analysis. Journal of Quality Technology, 28(1), 91-98. Boyles, R. A. (1996). Multivariate process analysis with lattice data. Technometrics, 38(1), 37-49. Buckland S. T. (1981). Monte Carlo Intervals. Department of Statistics Technical Report No. 1, University of Aberdeen, Scotland.
Bibliography
347
Buckland S. T. (1982). A note on the fourier series model for analysing line transect data. Biometrics, 38(2), 469-477. Buckland S. T. (1983). Monte Carlo methods for confidence interval estimation using the bootstrap technique. Bias, 10(2), 194-212. Buckland S. T. (1984). Monte Carlo confidence intervals. Biometrics, 40(3), 811817. Buckland S. T. (1985). Algorithm AS 214: Calculation of Monte Carlo confidence intervals. Applied Statistics, 34(3), 296-301. Burke, R. J., Davis, R. D. and Kaminsky, F. C. (1991). Responding to statistical terrorism. J^lth Annual Conference, Rochester Section of the American Society for Quality Control, March 12. Carr, W. E. (1991). A new process capability index: parts per million. Quality Progress, 24(8) 152. Case, K. E., Brooks, D. H. and Bigelow, J. S. (1987). Proper use of process capability indices in SPC. IEE Integrated Systems Conference Proceedings, Nashville, Tennessee, 107-111. Castagliola, P. (1996). Evaluation of non-normal process capability indices using Burr's distribution. Quality Engineering, 8, 587-593. Chan, L. K., Cheng, S. W. and Spiring, F. A. (1988). A graphical technique for process capability. Quality Congress Transactions, Dallas, 268-275. Chan, L. K., Cheng, S. W. and Spiring, F. A. (1988). A new measure of process capability: Cpm. Journal of Quality Technology, 20(3), 162-175. Chan, L. K., Cheng, S. W. and Spiring, F. A. (1988). The robustness of the process capability index, C„, to departures from normality. Theory and Data Analysis II (edited by K. Matusita), 223-239, North Holland. Chan, L. K., Cheng, S. W. and Spiring, F. A. (1991). A multivariate measure of process capability. Journal of Modeling and Simulation, 11(1), 1-6. Chan, L. K., Spiring, F. and Xiao, H. (1988). An OC curve approach for analyzing the process capability index C^. Technical Report, Department of Statistics, University of Manitoba, Canada. Chan, L. K., Xiong, Z. and Zhang, D. (1990). On the asymptotic distributions of some process capability indices. Communications in Statistics: Theory and Methods, 19(1), 11-18. Chang, P. L. and Lu, K. H. (1994). PCI calculations for any shape of distribution with percentile. Quality World, technical section (September), 110-114. Chang, P. L. and Lu, K. H. (1995). The construction of the stratification procedure for quality improvement. Quality Engineering, 8(2), 237-247. Chang, Y. S., Choi, I. S. and Bai, D. S. (2002). Process capability indices for skewed populations. Quality and Reliability Engineering International, 18(5), 383-393.
348
Encyclopedia and Handbook of Process Capability Indices
Chen, H. F. (1992). A multivariate process capability index over a rectangular parallelpiped. Technical Report, Mathematical Statistics Department, Green State University, Bowling Green, Ohio. Chen, H. F. (1994). A multivariate process capability index over a rectangular solid tolerance zone. Statistica Sinica, 4, 749-758. Chen, H. F. (1996). Asymptotic analysis of a class of process capability indices. (Manuscript). Chen, H. F. and Kotz, S. (1996). An asymptotic distribution of Wright's process capability index sensitive to skewness. Journal of Statistical Computation and Simulation, 55, 147-158. Chen, J. P. (2000). Re-evaluating the process capability indices for non-normal distribution, International Journal of Production Research, 38(6), 13111324. Chen, J. P. and Ding, C. G. (2001). A new process capability index for nonnormal distributions, International Journal of Quality and Reliability Management, 18, 762-770. Chen, J. P. and Pearn, W. L. (2002). Testing process performance based on the yield: an application to the liquid crystal display module (LCDM). Microelectronics Reliability, 42, 1235-1241. Chen, K. S. (1995). Process capability indices. Unpublished Ph.D. Thesis, Department of Industrial Engineering and Management, National Chiao Tung University. Chen, K. S. (1996). Process capability indices for skew populations. Journal of Management and Systems, 3(2), 207-216. Chen, K. S. (1997). A new process capability index for asymmetric tolerances. Journal of the Chinese Institute of Industrial Engineers, 14(4), 355-362. Chen, K. S. (1998). Estimation of the process incapability index, Communications in Statistics - Theory and Methods, 27(5), 1263-1274. Chen, K. S. (1998). Incapability index with asymmetric tolerance, Statistica Sinica, 8, 253-262. Chen, K. S. (1998). Estimation of the process incapability index. Communications in Statistics: Theory and Methods, 27(5), 1263-1274. Chen, K. S. (1998). Incapability index with asymmetric tolerances. Statistica Sinica, 8, 253-262. Chen, K. S. and Pearn, W. L. (1997). An application of the non-normal process capability indices. Quality and Reliability Engineering International, 13, 355-360. Chen, K. S. and Pearn, W. L. (2001). Capability indices for processes with asymmetric tolerances. Journal of the Chinese Institute of Engineers, 24(5), 559-568.
Bibliography
349
Chen, K. S., Huang, M. L. and Li, R. K. (2001). Process capability analysis for an entire product. International Journal of Production Research. 39(17), 4077-4087. Chen, K.S., Chen, S. C. and Li, R. K. (2002). Process quality analysis of products. International Journal of Advanced Manufacturing Technology, 19(8), 623-628. Chen, K. S., Pearn, W. L. and Lin, P. C. (1999). A new generalization of the capability index C„m for asymmetric tolerances. International Journal of Reliability, Quality and Safety Engineering, 6(4), 383-398. Chen, K. S., Pearn, W. L. and Lin, P. C. (2003). Capability measures for processes with multiple characteristics. Quality and Reliability Engineering International, 19(2), 101-110. Chen, S. M. (1996). Erratum. The asymptotic distribution of the process capability index Cvmk. Communications in Statistics: Theory and Methods, 25(5), 1151-1154. Chen, S. M. (1996). Response to the editor. Communications in Statistics: Theory and Methods, 25(6), 1407-1408. Chen, S. M. and Hsu, N. F. (1995). The asymptotic distribution of the process capability index Cpmk. Communications in Statistics: Theory and Methods, 24(5), 1279-1291. Chen, S. M. and Pearn, W. L. (1997). The asymptotic distribution of the estimated process capability index Cpk. Communications in Statistics: Theory and Methods, 26(10), 2489-2497. Chen, S. M., Hsu, Y. S. and Pearn, W. L. (2000). Estimating process capability index Cp(a,b) for m-dependent stationary processes. Far East Journal of Theoretical Statistics, 4(1), 45-64. Chen, S. M., Hsu, Y. S. and Pearn, W. L. (2003). Capability measures for m dependent stationary processes. Statistics, 37(2), 145-168. Cheng, S. W. (1992). Is the process capable? Tables and graphs in assessing Cpm. Quality Engineering, 4(4), 563-576. Cheng, S. W. (1994). Practical implementation of the process capability indices. Quality Engineering, 7(2), 239-259. Cheng, S. W. and Spiring, F. A. (1989). Assessing process capability: a Bayesian approach. IEE Transactions, 21(1), 97-98. Choi, B. C. and Owen, D. B. (1990). A study of a new capability index. Communications in Statistics: Theory and Methods, 19(4), 1231-1245. Choi, K. C , Nam, K. H. and Park, D. H. (1996). Estimation of capability index based on bootstrap method. Microelectronics Reliability, 36(9), 1141-1153. Chou, Y. M. and Owen, D. B. (1984). One-sided confidence regions on the upper and lower tail areas of the normal distribution. Journal of Quality Technology. 16(3), 150-158.
350
Encyclopedia and Handbook of Process Capability Indices
Chou, Y. M. and Owen, D .B. (1989). On the distributions of the estimated process capability indices. Communications in Statistics: Theory and Methods, 4549-4560. Chou, Y. M., Owen, D. B. and Borrego, S. A. (1990). Lower confidence limits on process capability indices. Journal of Quality Technology, 22(3), 223-229. Chou, Y. M. and Polansky, A. M. (1993). Power of tests for some process capability indices. Communications in. Statistics: Simulation and Computation, 22, 523-544. Chou, Y. M., Polansky, A. M. and Mason, R. L. (1998). Transforming nonnormal data to normality in statistical process control. Journal of Quality Technology, 30, 133-141. Chou. Y. M. (1994). Selecting a better supplier by testing process capability indices. Quality Engineering, 6(3), 427-438. Christofferson, J. (1999). Process capability indices and block bootstrap for serially correlated data. Research Report, Division of Quality Technology and Statistics, Lulea University, Sweden. Clements, J. A. (1989). Process capability calculations for non-normal distributions. Quality Progress, September, 95-100. Coleman, D. E. (1991). Relationships between loss and capability indices. Applied Math. Comp. Techn., ALCOA Technical Center, PA. Collins, Alan J. (1995). Bootstrap Confidence limits on process capability indices, Statistician, 44(3), 373-378. Constable, G. K. and Hobbs, J. R. (1992). Small samples and non-normal capability. Transactions, ASQC Quality Congress, 1-7. Corbett, C. J. and Pan, J.-N. (2002). Evaluating environmental performance using statistical process control techniques. European Journal of Operations Research, 139, 68-83. Craig, R. J. (1993). Six-sigma quality, the key to customer satisfaction. ASQC Quality Congress Transactions - Boston, 206-215. Davis, R. D. and Kaminsky, F. C. (1989). Statistical measures of process capability and their relationship to nonconforming product. Proceedings of the Third Biannual International Manufacturing Technology Research Forum, Illinois, August. Davis, R. D. and Kaminsky, F. C. and Saboo, S. (1992). Process capability analysis for process with either a circular or a spherical tolerance zone. Quality Engineering, 5(1), 41-54. Deleryd, M. (1998). On the gap between theory and practice of process capability studies, International Journal of Quality and Reliability Management, 15, 178-191. Deleryd, M. (1998). Enhancing the industrial use of process capability studies. PhD Thesis, Lulea University of Technology, Sweden.
Bibliography
351
Deleryd, M. (1999). A pragmatic view on process capability studies. International. Journal of Production Economics, 58, 319-330. Deleryd, M. (1999). The effect of skewness on estimates of some process capability indices. International Journal of Appl. Qual. Marat., 2(2), 153186. Deleryd, M., Deltin, J. and Klefsjo, B. (1997). Critical factors for successful implementation of process capability. Deleryd, M. and Vannman, K. (1998). Process capability studies for short production runs. International Journal of Reliability, Quality and Safety Engineering, 5(4), 383-401. Deleryd, M. and Vannman, K.(1999). Process capability plots - A quality improvement tool. Quality and Reliability Engineering International, 15, 213-227. Ding, C. G. (1994). An efficient algorithm for computing the noncentrality parameters of chi-squared tests. Communications in Statistics: Simulation and Computation, 23(3), 861-870. Ding, C. G. (1999). An efficient algorithm for computing quantiles of the noncentral chi-squared distribution. Computational Statistics and Data Analysis. 29, 253-259. Duffy, Diane E. (1993). Is industrial statistics out of control?. Comment, Statistical Science, 8(4), 380-384. Dumitrescu, M. and Hubele, N. F. (1999). An entropic framework for the normal distribution in capability analysis. Communications Statistics - Theory Math, 28(6), 1361-1337. Dyson, F. J. (1943). A note on kurtosis. Journal of the Royal Statistical Society. 106(4), 360-361. Efron, B. (1979). Bootstrap methods: Another look at the jackknife. Annals of Statistics, 7(1), 1-26. Efron, B. (1982). The Jackknife, the Bootstrap and Other Re-sampling Plans, SIAM, Philadelphia, Society for Industrial and Applied Mathematics. Efron, B. (1992). Jackknife - after - Bootstrap Standard errors and influence functions. Journal of the Royal Statistical Society. Series B, 54 (1), 83-127. Efron, B. and Tibshirani R. (1986). Bootstrap methods for standard errors, confidence intervals, and other measures of statistical accuracy. Statistical Science, 1 (1), 54-75. EI-Awady, K., Schaper, C. and Kailath, T. (1996). Improvements in using realtime feedback control. IEEE Transactions on Semiconductor Manufacturing, 9(1), 87-94. Ekvall, D. N. and Juran J. M. (1974). Manufacturing planning. Quality Control Handbook, Third Edition, (J. M. Juran ed.) McGraw-HilhNew York.
352
Encyclopedia and Handbook of Process Capability Indices
English, J. R. and Taylor, G. D. (1993). Process capability analysis - a robustness study. International Journal of Production Research, 31(7), 1621-1635. Farnum, N.R. (1996-97). Using Johnson Curves to Describe Non-Normal Process Data. Quality Engineering, 9(2),329-336. Fieler, P. E. and Loverro, N. Jr. (1991). Defects tail off with six-sigma manufacturing. IEEE, 18-20/48. Fink, J. and Wolff, H. (1995). A new approach for describing and controlling process capability. Proc. Osaka Conf. Qual. Control, Osaka, Japan, 1-25. Finley, J. C. (1992). What is capability or what is Cp and Cpk. ASQC Quality Congress Transactions - Nashville, 186-192. Finucan, H. M. (1964). A note on kurtosis. Journal of the Royal Statistical Society. Series B (Methodological), 26(1), 111-112. Flaig, J. J. (1996). A new approach to process capability analysis. Quality Engineering, 9(2), 205-211. Flaig, J. J. (1999). Process Capability Sensitivity Analysis. Quality Engineering, 11(4), 587-592. Foguth, R. C. and Motwani, F. (1998). A method of including material variation in capability studies. The TQM magazine, 10(2), 104-108. Franklin, L. A. Cooley, B. J. and Elrod, G. (1999). Comparing the importance of variation and mean of a distribution. Quality Progress, 32(10), 90-94. Franklin, L. A. (1999). Sample size determination for lower confidence limits for estimating process capability indices, Computers and Industrial Engineering, 36, (3), 603-614. Franklin, L. A. and Wasserman, G. S. (1991). Bootstrap confidence interval estimates of Cpk: an introduction. Communications in Statistics: Simulation and Computation, 20(1), 231-242. Franklin, L. A. and Wasserman, G. S. (1992). Bootstrap lower confidence limits for capability indices. Journal of Quality Technology, 24(4), 196-210. Franklin, L. A. and Wasserman, G. S. (1992). Bootstrap lower confidence limits estimates for C,h, (the new flexible process capability indices), Pakistan Journal of Statistics, 10(1), 33-45. Franklin, L. A. and Wasserman, G. S. (1992). A note on the conservative nature of the tables of lower confidence limits for Cpk with a suggested correction. Communications in Statistics: Simulation and Computation, 21(4), 11651169. Frey, D. D., Otto, K. N. and Wysocki, J. A. (2000). Evaluating Process Capability During the Design of Manufacturing Systems. Journal of Manufacturing Science and Engineering, 122(3), 513-519. Fun, Y. P. (1995). A second thought on the quadratic loss function of Taguchi's quality engineering and its application. Unpublished Ph.D. Thesis,
Bibliography
353
Department of Industrial Engineering, National Tsing Hwa University, Taiwan. Fun, Y. P. and Li, C. C. (1991). Some discussion and proposition on process capability index. Journal of the Chinese Institute of Industrial Engineers, 8(2), 83-89. Gabel, S. H. (1990). Process performance chart, ASQC Quality Congress Transactions -San Francisco, CA., 683-688. Gensidy, A. (1985). Cp and Cpk. Quality Progress, 18(4), 7-8. Gielen, G.; Debyser, G.; Leyn, F.; Sansen, W. (1998). Efficient yield estimation within automated analog IC design. Statistical Metrology, 1998. 3rd International Workshop, 118 -121 Gilchrist, W. (1993). Modeling capability. Journal of the Operational Research Society, 44(9), 909-923. Goh, T. N. (1994). Some practical issues in the assessment of nonconforming rates in a manufacturing process. International Journal of Production Economics, 33, 81-88. Goh, T. N. (2002). The role of statistical design of experiments in six sigma: perspectives of a practitioner. Quality Engineering, 14(4), 659-671. Grant, E. L. (1988). Statistical quality control, 6-th edition, McGraw-Hill, New York. Greenwich, M. and Jahr-Schaffrath, B. L. (1995). A process incapability. International Journal of Quality and Reliability Management, 12(4), 58-71. Gronwich, U. (1990). The concept of capability index: assumptions and conclusions, Res. Rep. 1990:2, Div. Qual. Technol., Lulea Univ. Sweden. Gruner, G. (1991). Calculation of Cpk under conditions of variable tolerances. Quality Engineering, 3(3), 281-291. Gruska, G. F., Mirkhani K. and Lamberson L. R. (1989). Non-normal Data Analysis, Multiface Publishing Co, Michigan. Guirguis, G. H. and Rodriguez, R. N. (1992). Computation of Owen's Q function applied to process capability analysis. Journal of Quality Technology, 24(4), 236-246. Gunter, B. H. (1989). The use and abuse of Cpk, Parts 1-4, Quality Progress, 22(1), 72-73; (3), 108-109; (5), 79-80; (7) 86-87. Gunter, B. H. (1991). Process capability studies, Parts 1-5, Quality Progress, 24 (2), 97-100; (4), 102-106; (6), 77-85; (8), 123-132; (10), 109-113. Gunter, B. H. (1991). Bootstrapping: How to Make Something from Almost Nothing and Get Statistically Valid Answers. Part I: Brave New World. 24(12), 97-103. Part II: The Confidence Game. 25(2), 83-86. Part III: Examples and Enhancements. 25(4), 119-122. Part IV: Bias-Corrected Intervals. 25(6), 79-83. Gupta, A. K. and Kotz, S. (1997). A new process capability index. Metrika, 45(3), 213-224.
354
Encyclopedia and Handbook of Process Capability Indices
Hall, Peter (1988). Theoretical comparison of Bootstrap confidence intervals. Annals of Statistics, 16(3), 927-953. Hallin, M. and Sech, M. (1999). When does Edgeworth beats Berry and Esseen? Numerical evaluation of Edgeworth expansion, Journ. Statist. Plann. Inf., 63 19-38. Han, Chien-Pai and Hawkins, D. L. (1994). A smooth adaptive estimator of the mean of a symmetric or asymmetric distribution. Communications in Statistics: Theory and Methods, 23(1), 1 -10. Han, J. H., Cho, J.J., and Leem, C. S. (2000). Bootstrap confidence limits for Wright's Cs. Communication in Statistics: Theory and Methods, 29(3), 485505. Hansen, J. L., Zirk, W. E. and D'Esposito, L. (1993). What is the capability of your measurement process? ASQC Quality Congress Transactions - Boston, 676-680. Harry, M. J. (1988). The nature of six sigma quality. Technical Report, Motorola, Schaumburg, Illinois. He., Zhen; Qi, Er-shi and Zhang, S. H. (2000). Some problems in process capability analysis and evaluation. Industrial Engineering Journal, 3, 25. Heavlin, W. D. (1988). Statistical properties of capability indices. Technical Report No. 320, Library, Advanced Micro Devices, Sunnyvale, CA. Herman, J. T. (1989). Capability index - enough for process industries? ASQC Quality Congress Transactions - Toronto, 670-675. Hoffman, L. L. (1993). A general approach for testing the process capability index. Quality and Reliability Engineering International, 9, 670-675. Hoffman, L. L. (1999). Hypothesis testing of Cp. MS, Dept. Statistics, Univ. Central Florida, Orlando, FL. Hoffman, L. L. (2001). Obtaining confidence intervals for Cpk using percentiles of the distribution of Cp. Quality and Reliability Engineering International, 17(2), 113-118. Hogg, R. V. (1967). Some observations on robust estimation. Journal of the American Statistical Association, 62, 1179-1186 Hogg, R. V. (1984). A review of some adaptive statistical techniques. Communication in Statistics: Theory and Methods, 13, 1551-1579. Hogg, R. V. and Craig, A. T. (1989). Introduction to Mathematical Statistics. 4th edition, New York. Holmes, D. (1980). A quality portfolio management chart, Stochas, Schenectady, N. Y. (Also in Quality, 25 (12), 67(1986)). Holmes, D. S. and Mergen, A. E. (1988). A managerial tool for process capability analysis. Quality and Reliability Engineering International, 4, 63-66. Hoskins, J., Stuart, B. and Taylor, J. (1988). A Motorola commitment: a six sigma mandate. The Motorola Guide to Statistical Process Control for
Bibliography
355
Continuous Improvement Towards Six Sigma Quality, Motorola, Schaumburg, Illinois. Hossain, A. (2000). Statistical analysis and process capability determination of an industrial process. Industry Applications Conference 2000, IEEE, (2), 1178 -1185. Hsiang, T. C. and Taguchi, G. (1985). A tutorial on quality control and assurance - the Taguchi methods. ASA Annual Meeting, Las Vegas, Nevada Huang, D. Y. and Lee, R. F. (1995). Selecting the largest capability index from several quality control processes. Journal of Statistical Planning and Inference, 46, 335-346. Hubele, N. F. and Zimmer, L. (1994). Average run lengths for a Cpm control chart. ASQC 48th Annual Quality Congress Proceedings, 577-582. Hubele, N. F., Montgomery, D. C. and Chih, W. H. (1991). An application of statistical process control in jet-turbine engine component manufacturing. Quality Engineering, 4(2), 197-210. Hubele, N. F., Shahriari, H. and Cheng, C. S. (1991). A bivariate process capability vector, in Statistical Process Control in Manufacturing (Edited by Keats, J.B. and Montgomery, D.C.), Dekker, 299-310, New York, N. Y. Hubele, N. F. Beaumariage, T., Bawaja, G., Hong S.C. and Chu, R. (1994). Using experimental data to access capability of a system. J. Quality Technology, 26, 1-11. Hubele, N. F., Zimmer, L.S., Kunjura R., and Dimitrescu, M. (1999). Estimation for Cpm -hypothesis tests, MS. Arizona State Univ., Tempa, AZ. Hung, K. and Hagen, D. (1992). Quality computation using GAUSS: examples in process capability research. 24-th Symposium on the Interface.? Huttunen, E. K. (1992). Quality planning for a precision machining process. ASQC Quality Congress Transactions, Nashville, 405-411. Ingle, S. (1985). Cv and Cpk. Quality Progress, 7. Jackson, J. E. (1980). Principal component and factor analysis: Part I - principal components. Journal of Quality Technology, 12, 201-213. Jagadeesh, R. and Babu, A. S. (1994). Process capability assessment with tool wear: an investigative study. International Journal of Quality and Reliability Management, 11(2), 51-62. Janakiram, M. and Keats, B. (1995). The use of FMEA in process quality improvement. Journal of Reliability, Quality and Safety Engineering, 2(1), 103-115. Jeang, A. (1997). An approach of tolerance design for quality improvement and cost reduction. International Journal of Production Research, 35(5), 11931211. Jessenberger, J., and Weihs, C. (2000). A note on the behavior of Cpmk with asymmetric specification limits. Journal of Quality Technology, 32, 440-443.
356
Encyclopedia and Handbook of Process Capability Indices
Jessup. P. (1987). Continuing process control and process capability improvement. Ford Motor Company, Dearborn, MI. Johnson, F. C. and Smith, G. S. (1997). Process stability and process capability in public schools. Quality Engineering, 9(3), 503-520. Johnson, N. L. (1996). The asymptotic distribution of the process capability index Cpmk. Communications in Statistics: Theory and Methods, 25(6), 1405. Johnson, N. L, Kotz, S. and Pearn, W. L. (1994). Flexible process capability indices. Pakistan Journal of Statistics, 10(1)A, 23-31. Johnson, T. (1991). A new measure of process capability related to Cpm. Department of Agricultural Resource Economics, North Carolina State University, Rayleigh, North Carolina Johnson, T. (1992). The relationship of Cpm to squared error loss. Journal of Quality Technology, 24(4), 211-215. Jung, H.-W. and Hunter, R. (2001). The Relationship Between ISO/IEC 15504 Process Capability Levels, ISO 9001 Certification and Organization Size: An Empirical Study. International Journal of Systems and Software, 59, 4355. Juran, J. M. (1974). Juran's Quality Control Handbook, 3rd edn., McGraw-Hill, New York, N.Y. Juran, J. M. (1988). Juran On Planning For Quality, The Free Press, New York, N. Y. Juran, J. M. and Gryan, F. M. (1980). Quality Planning and Analysis, McGrawHill, New York, N. Y. Kalyanasundaram, M. and Balamurali, S. (1997). A superstructure capability index for short run production processes and its bootstrap lower confidence limits, Exon. Qual. Control., 12 (2), 85-93. Kaminsky, F. C , Dovich, R. A. and Burke, R. J. (1998). Process capability indices: now and in the future. Quality Engineering, 10(3), 445-453. Kane, V. E. (1986). Process capability indices. Journal of Quality Technology, 18(1), 41-52. Kane, V. E. (1986). Corrigenda. Journal of Quality Technology, 18(4), 265. Kaplansky, I. (1945). A common error concerning kurtosis. Journal of the American Statistical Association. 40(230), 259. Karl, D. P., Morisette, J. and Taam, W. (1994). Some applications of a multivariate capablity index in geometric dimensioning and tolerancing. Quality Engineering, 6(4), 649-665. Kazmer, D. and Roser, C. (1999). Evaluation of product and process design robustness. Research in Engineering Design, 11(1), 20-30. Kazmerski, T. J. (1995). Statistical Problem Solving in Quality Engineering. McGraw-Hill, New York.
Bibliography
357
Kirmani, S. N. U. A., Kocherlakota, K. and Kocherlakota, S. (1991). Estimation of a and the process capability index based on sub-samples Communications in Statistics: Theory and Methods, 20(1), 275-291. Kirmani, S. N. U. A., Kocherlakota, K. and Kocherlakota, S. (1991). Correction to: Estimation of cr and the process capability index based on sub-samples Communications in Statistics: Theory and Methods, 20(12), 4083. Kiska, D. (1991). Cp and Cpk statistics should be discontinued. Quality Progress, March, 8. Knowles, Graeme, March, Guy and Anthony, Jiju (2002). Evaluating process capability for geometrically toleranced Parts: A practical approach. Quality Engineering, 14(3), 365-374. Kocherlakota, S. (1992). Process capability index: recent developments. Sankhya, 54(B), 352-369. Kocherlakota, S. and Kocherlakota, K. (1991). Process capability index: bivariate normal distribution. Communications in Statistics: Theory and Methods, 20(8), 2529-2547. Kocherlakota, S. and Kocherlakota, K. (1994). Confidence intervals for the process capability ratio based on robust estimators. Communications in Statistics: Theory and Methods, 23(1), 257-276. Kocherlakota, S., Kocherlakota, K. and Kirmani, S.N.U.A. (1992). Process capability indices under non-normality. International Journal of Mathematical and Statistical Sciences, 1(2), 175-210. Kocherlakota, S. and Kocherlakota, K. (1994). Confidence intervals for the process capability ratio based on robust estimators. Communications in Statistics: Theory and Methods, 23(1), 257-276. Kotz, S. and Johnson, N. L. (1993). Process capability indices for non-normal populations. International Journal of Mathematical Statistical Science, 1, 35-44. Kotz, S. and Johnson, N. L. (1993). Process Capability Indices. Chapman and Hall, London, U. K. Kotz, S. and Johnson, N. L., (1999). Delicate relations among the basic process capability indices Cp, Cpk, Cpm and their modifications. Communications. Statistics- Theory and Method, 26, 849-861. Kotz, S. and Johnson, N. L. (2002). Process capability indices - a review, 19922000. Journal of Quality Technology, 34(1), 2-53. Kotz, S., Johnson, N. L. and Balakrishnan, N. (2002). Distributions in Statistics: Continuous Multivariate Distributions, (second edition) Wiley, New York. Kotz, S. and Lovelace C. R. (1998). Process Capability Indices in Theory and Practice, Arnold, London, U.K. Kotz, S., Pearn, W. L. and Johnson, N. L. (1993). Some process capability indices are more reliable than one might think. Journal of the Royal Statistical Society, Series C: Applied Statistics, 42(1), 55-62.
358
Encyclopedia and Handbook of Process Capability Indices
Ku, Li-Lin (1997). Robust Inference about Means using Smooth Adaptive Estimators. Ph.D. dissertation, The University of Texas at Arlington, Department of Industrial and Manufacturing Systems Engineering, Anlington, Texes. Kushler, R. H. and Hurley, P. (1992). Confidence bounds for capability indices. Journal of Quality Technology, 24(4), 188-195. Lee, J. C , Hung, H. N., Pearn, W. L. and Kueng, T. L. (2002). On the distribution of the estimated process yield index Spt. Quality and Reliability Engineering International, 18(2), 111-116. Lee, S. J. and Amin, R. W. (2000). Process tolerance limits. The TQM magazine, 11(3), 267-280. Lee, Yeong-Hoang, Wei, Chiu-Chi and Chang, Ching-Liang (1999). Fuzzy design of process tolerances to maximise process capability. International Journal of Advanced Manufacturing Technology. 15(9), 655-659. Lehrman, K. H. (1991). Capability analysis and index interval estimates: quantitative characteristics (variables). Quality Engineering, 4(1), 93-129. Leone, F. C , Nelson, L. S. and Nottingham, R. B. (1961). The folded normal distribution. Technometrics, 3, 543-550. Levinson, W. A. (1997). Exact confidence limits for process capabilities. Quality Engineering, 9(3), 521-528. Levinson, W. A. (1997). Approximate confidence limits for Cpk and confidence limits for non-normal process capabilities. Quality Engineering, 9(4), 635640. Lewis, S. S. (1991). Process capability estimates from small samples. Quality Engineering, 3(3), 381-394. Li, C. C. and Fun, Y. P. (1993). Using expected loss of quadratic loss function with specification limits. Consideration to compare various process capability indices. Proceedings of the IASTED International Conference, Cambridge, Mass. USA, October 18-20, 5-8. Li, H., Owen, D. B. and Borrego A., S. A. (1990). Lower confidence limits on process capability indices based on the range. Communications in Statistics: Simulation and Computation, 19(1), 1-24. Li, Y. and Lin, C. (1996). Multivariate Cp value, Chinese J. Appl. Prob. Statist, 12(2), 132-138. Lin, G. H. and Pearn, W. L. (2002). On the interval estimation of Cpk with asymmetric tolerances. Journal of Nonparametric Statistics, 14(6), 647-654. Lin, G. H., Pearn, W. L. and Chen, S. M. (2002). A note on the distribution of the estimated Cpk with asymmetric tolerances. Far East Journal of Theoretical Statistics, 6(1), 25-37. Lin, G. H. and Pearn, W. L. (2003). Distributions of the estimated process capability index Cpk. Economic Quality Control, 18(2), 67-91.
Bibliography
359
Lin, P. C. and Pearn, W. L. (2002). Testing process capability for one-sided specification limit with application to the voltage level translator. Microelectronics Reliability, 42(12), 1975-1983. Lin, P. C. and Pearn, W. L. (2005). Testing process performance based on the capability index Cpm. International Journal of Advanced Manufacturing Technology. Forthcoming. Lin, P. C. and Pearn, W. L. (2005). Assessing process performance based on incapability index Cpp. International Journal of Industrial Engineering, 12(2), 143-156. Lin, P. C , Pearn, W. L. and Chen, K. S. (2001). Distributional properties and implications of the estimated process incapability index Cpp . American Journal of Mathematical and Management Sciences, 23(2), 75-92. Lin, T.Y.; Yak, A.S. (2000). Reliability by Robust Design for optical alignment in a CD pickup. Mechatronics, IEEE/ASME Transactions, 5(4), 386 -393. Littig, J. L., Lam, C. T. and Pollock, S. M. (1992). Process capability measurement for a bivariate characteristic over an elliptical tolerance zone. Technical Report No. 92-42, Department of Industrial and Operations Engineering, University of Michigan, Ann Arbor. Littig, J. L. and Lam, C. T. (1993). Case studies in process capability measurement. ASQC Quality Congress Transactions, Boston, 569-575. Little, T. A. and Harrison, C. S. (1993). Short run capability assessment using the t distribution. ASQC Quality Congress Transactions, Boston, 181-191. Liu, X. Q. (1995), Analytical analysis of effect of sampling plan on Cpk. Engineering Management Conference, 1995. Global Engineering Management: Emerging Trends in the Asia Pacific, Proceedings of the 1995 IEEE Annual International , 257 -259. Long, J. M. and De Coste, M. J. (1988). Capability studies involving tool wear. ASQC Quality Congress Transactions, Dallas, 590-596. Loon Ching Tang; Su Ee Than (1999). Computing process capability indices for non-normal data: a review and comparative study. Quality and Reliability Engineering International, 15, 339-353. Lovera, C , Moiraghi, L., Montefusco, A. and Valentini, D. (1997). Estimation of industrial process capabilities: Some proposals and studies to overcome the obstacle of non-normal distributions, Abstract III, Cong.-Naz. SIMAI, Pavia, Italy, 627-631. Lu, K. H. (1994). A study of the recommended minimum values of process capability indices and the testing procedure of process capability. Unpublished Ph.D. Thesis, Institute of Management Science, National Chiao Tung University, Taiwan. Luceno, A. (1996). A process capability index with reliable confidence intervals. Communications in Statistics: Simulation and Computation, 25(1), 235-245.
360
Encyclopedia and Handbook of Process Capability Indices
Luceno, C , (1999). Discrete approximations to continuous univariate distributions - an alternative to simulation, J. R. Statist. Soc, B, 61, 345352. Lupienski, J. A. (1989). Taguchi quality loss function and its relationship to process capability centering (Cpk). 45-th Annual ASQC Quality Control Conference Transactions - Rochester, N.Y. Lyth, D. M. and Rabiej, R. J. (1995). Critical variables in wood manufacturing's process capability: species, structure, and moisture content. Quality Engineering, 8(2), 275-281. Mahalanobis, P. C. (1936). On the generalized distance in statistics. Proceedings of the National Institute of Sciences of India, 12, 49-55. Marcucci, M. O. and Beazley, C. C. (1988). Capability indices: process performance measures. ASQC Quality Congress Transactions, Dallas, 516523. MATLAB version 6.1 R12, The Mathworks Inc., 2001. Mazzuchi, T. (1997). Notes on Cp and Cpk estimation. Report, School of Engineering, George Washington Univ., Washington, D. C. McCormack, D. W. Jr., Harris, I. R., Hurwitz, A. M. and Spangon, P. D. (2000), Capability indices for non-normal data. Quality Engineering, 12(4), 489495. McCoy, P. F. (1991). Using performance indexes to monitor production processes. Quality Progress, February, 49-55. McFadden, F. R. (1993). Six-sigma quality programs. Quality Progress, June, 3741. Mcneese, W. H. and Klein, R. A. (1991). Measurement systems, sampling, and process capability. Quality Engineering, 4(1), 21-39. Melloy, B. J. and Chandra, M. J. (1992). Consequence of using the Cpk index to estimate the proportion of non-conforming items. Quality Engineering, 4(1), 21-39. Mirabella, J. (1991). Determining which capability index to use. Quality Progress, 24(8), 8-10. Mittag, H.- J. (1997). Measurement errors effects on the performance of process capability indices. In Frontiers in Statistical Quality Control (H. J. Lenz and P. M. Wilrich, eds.) Montgomery, D. C. (1985). Introduction to Statistical Quality Control, John Wiley, New York. Montgomery, D. C. and Mastrangelo, C. M. (1991). Some statistical process control method for autocorrelated data. Journal of Quality Technology, 23(3): 179-199. Mukherjee, S. P. (1995). Process capability indices and associated inference problems, Proc. Internat. Conf. Statist. Method, Statist. Compt. Seoul, South Korea, 243-249.
Bibliography
361
Mukherjee, S. P. and Singh, N. K. (1994). Sampling properties of an estimate of the bivariate process capability index. Economic Quality Control, 9, 73-78. Mukherjee, S. P. and Singh, N. K. (1997/8). Sampling properties of an estimator of a new process capability index for Weibull distributed quality characteristics, Quality Engineering, 10, 291-294. Munichika, M. (1986). Evaluation of process capability for skew distributions. 30th EOQC Conference, Stockholm, Sweden. Munichika, M. (1992). Studies on process capability in machining processes, Rep. Statist. Appl. Res. JUSE, 39, (4), 14-29. Nagata, Y. (1991). Interval estimation for the process capability indices. Journal of Japanese Society of Quality Control, 21, 109-114. Nagata, Y. (1995). Point estimation of the process capability index, Okayama Econ. Rev. 26, Nos. 3-4. Nagata, Y. and Nagahata, H. (1992). Approximation formulas for the confidence intervals of process capability indices, Rep. Statist. Appl. Res, JUSE, 39, (3), 15-29. Nagata, Y. and Nagahata, H. (1994). Approximation formulas for the lower confidence limits of process capability indices. Okayama Economic Review 25, 301-314. Nahar, P. C ; Hubble, N. F. and Zimmer, L. S. (2001). Assessment of a capability index sensitive to skewness. Quality and Reliability Engineering International, 17, 233-241. Nam, K.H.; Dae Kyung Kim; Dong Ho Park (2002). Large-sample interval estimators for process capability indices. Quality Engineering, 14, 213. Nelson, L. S. (1999) The calculation of process capability indices, J. Quality Technology, 31, 249-250. Nelson, P. R. (1992) Editorial, J. Qual. Technol, 24, 175. Newnes, L.B.; Mileham, T.R.; Doniavi, A. (2001). A systems approach to semiconductor optimization. Electronics Packaging Manufacturing, IEEE Transactions on, 24(3), 171 -177. Newton, D. (1994). Process Capability In: Gower Handbook of Quality Management ( D. Lou, ed.) 2nd edition, Gower, Aldershot, Hampshire, U.K.. Niverthi, M. and Dey, D. K. (2000). Multivariate process capability: A Bayesian perspective. Communications in Statistics: Simulation and Computation, 29(2), 667-687. Noguera, J. and Nielsen, T. (1992). Implementing six sigma for interconnect technology. ASQC Quality Congress Transactions, Nashville, TENN., 538544. Noorossana, R. (2002). Process capability analysis in the presence of autocorrelation. Quality and Reliability Engineering International, 18(1), 75-77.
362
Encyclopedia and Handbook of Process Capability Indices
Olorunniwo, F. and Knight, J.E. (1995). Monitoring process parameters with precontrol. Journal of Operations Management, 12, 75-87. Orchard, T. (2000) Quality improvement, RSS News, 27 (8), 8. Owen, D. B. (1965), A special case of a bivariate non-central ^-distribution. Biometrika, 52(3/4), 437-446. Padgett, M. M., Vaughn, L. E. and Lester, J. A. (1995). Statistical process control and machine capability in three dimensions. Quality Engineering, 7(4), 779-796. Padgett, W.J. and Sengupta, A. (1996). Performance of process capability indices for Weibull and Lognormal distributions on autoregressive processes. International Journal of Reliability, Quality and Safety Engineering, 3, 217229. Page, M. (1994). Analysis of non-normal process distributions. Semiconductor International, 10, 88-96. Palmer, K and Tsui, K. L. (1999). A review and interpretations of process capability indices, Ann. Oper. Res., 87, 31-47. Pan, J. N. (1995). Process capability analysis. Quality Control Journal, 9, 50-53. Pan, J. N. and Wu, S. L. (1996). A study on process capability indices for nonnormal data. Journal of Quality, 2(2), 79-97. Pan, N. J. and Wu, L. S.(1997). Process capability analysis for non-normal relay test data, Microelectronics and Reliability, 37(3), 421-428. Papp, Z. (1992). Two-sided tolerance limits and confidence bounds for normal populations. Communications in Statistics: Theory and Methods, 21(5), 1309-1318. Parlar, M. and Wesolowsky, G. O. (1999). Specification limits, capability indices, and process centering in assembly manufacture. Journal of Quality Technology, 31(3), 317-325. Parlar, M. and Wesolowsky, G. O. (1999). Specification limits, capability indices and centering in assembly manufacture, J. Quality Technology, 31, 317-325. Pearn, W. L. and Chang, C. S. (1997). The performance of process capability index Cs on skewed distributions. Communications in Statistics: Simulation and Computation, 26(4), 1361-1377. Pearn, W. L. and Chang, C. S. (1998). An implementation of the precision index for contaminated processes. Quality Engineering, 11(1), 101-110. Pearn, W. L. and Chang, C. S. (2005). Precision measures for processes with multiple manufacturing lines. International Journal of Advanced Manufacturing Technology. Forthcoming. Pearn, W. L., Chang, Y. C. and Wu, C. W. (2004). A quality-yield measure for production processes with very low fraction defective. International Journal of Production Research, 42(23), 4909-4925.
Bibliography
363
Pearn, W. L., Chang, Y. C. and Wu, C. W. (2004). Distributional and inferential properties of the process loss indices. Journal of Applied Statistics, 31(9), 1115-1135. Pearn, W. L., Chang, Y. C. and Wu, C. W. (2005). Bootstrap approach for estimating process quality yield with application to light emitting diodes. International Journal of Advanced Manufacturing Technology, 25, 560-570. Pearn, W. L. and Chen, K. S. (1995). Estimating process capability indices for non-normal Pearsonian populations, Quality and Reliability Engineering International, 11(5), 386-388. Pearn, W. L. and Chen, K. S. (1996). A Bayesian like estimator of Cpk. Communications in Statistics: Simulation and Computation, 25(2), 321-329. Pearn, W. L. and Chen, K. S. (1997). A practical implementation of the process capability index Cpk. Quality Engineering, 9(4), 721-737. Pearn, W. L. and Chen, K. S. (1997). Capability indices for non-normal distributions with an application in electrolytic capacitor manufacturing. Microelectronics and Reliability, 37(12), 1853-1858. Pearn, W. L. and Chen, K. S. (1997). Multi-process performance analysis: a case study. Quality Engineering, 10(1), 1-8. Pearn, W. L. and Chen, K. S. (1998). A generalization of the process capability index Cpk for asymmetric tolerances, Journal of Applied Statistics, 25(6), 801-810. Pearn, W. L. and Chen, K. S. (1998). New generalizations of the process capability index Cpk. Journal of Applied Statistics, 25(6), 801-810. Pearn, W. L. and Chen, K. S. (1999). Making decisions in assessing process capability index Cpk. Quality and Reliability Engineering International, 15, 321-326. Pearn, W. L. and Chen, K. S. (2002). One-sided capability indices CPU and CPL: decision making with sample information. International Journal of Quality and Reliability Management, 19(3), 221-245. Pearn, W. L., Chen, K. L. and Chen, K. S. (2002). A practical implementation of the incapability index Cpp. International Journal of Industrial Engineering. 9(4), 372-383. Pearn, W. L., Chen, K. S. and Lin, G. H. (1999). A generalization of Clements' method for non-normal Pearsonian processes with asymmetric tolerances. International Journal of Quality and Reliability Management, 16(5), 507521. Pearn, W. L., Chen, K. S. and Lin, P. C. (1999). On the generalizations of the capability index Cpmk for asymmetric tolerances. Far East Journal of Theoretical Statistics, 3(1), 47-66. Pearn, W. L., Chen, K. S. and Lin, P. C. (1999). The probability density function of the estimated process capability index Cpk. Far East Journal of Theoretical Statistics, 3(1), 67-80.
364
Encyclopedia and Handbook of Process Capability Indices
Pearn, W. L. and Chuang, C. C. (2004). Accuracy analysis of the estimated process yield based on Spk. Quality and Reliability Engineering International, 20, 305-316. Pearn, W. L., Chung, S. H. and Yang, M. H. (2002). Minimizing the Total Machine Workload for the Eafer Probing Scheduling Problem. HE Transactions. 34, 211-220. Pearn, W. L., Hsu, Y. C. and Wu, C. W. (2005). Tool replacement for production with low fraction defective. International Journal of Production Research. Forthcoming. Pearn, W. L., Kotz, S. and Johnson, N. L. (1992). Distributional and inferential properties of process capability indices. Journal of Quality Technology, 24(4), 216-231. Pearn, W. L., Ko, C. H. and Wang, K. H. (2002). A multi-process performance analysis chart based on the incapability index Cpp with an application to the chip resistors. Microelectronics Reliability, 42(7), 1121-1125. Pearn, W. L. and Liao, M. Y. (2005). Measuring process capability based on CPU with gauge measurement errors. Microelectronics Reliability, 45, 739-751. Pearn, W. L. and Liao, M. Y. (2005). One-sided process capability assessment in the presence of measurement errors. Quality and Reliability Engineering International. Forthcoming. Pearn, W. L. and Liao, M. Y. (2005). Estimating and testing process precision with presence of gauge measurement errors. Quality and Quantity. Forthcoming. Pearn, W. L. and Lin, G. H. (2000). A note on estimating process capability indices Cv(u,v). Far East Journal of Theoretical Statistics, 4(1), 1-12. Pearn, W. L. and Lin, G. H. (2000). Estimating capability index Cpk for processes with asymmetric tolerances. Communications in Statistics: Theory and Methods, 29(11), 2593-2604. Pearn, W. L. and Lin, G. H. (2001). On the reliability of the estimated incapability index, Quality and Reliability Engineering International, 17(4), 279-290. Pearn, W. L. and Lin, G. H. (2001). A note on the estimated incapability index: reliability and decision making with sample information. Quality and Reliability Engineering International, 17, 279-290. Pearn, W. L. and Lin, G. H. (2002). Estimated incapability index: reliability and decision making with sample information, Quality and Reliability Engineering International, 18, 141-147. Pearn, W. L. and Lin, G. H. (2003). A reliable procedure for testing linear regulators with one-sided specification limits based on multiple samples. Microelectronics Reliability, 43, 651-664.
Bibliography
365
Pearn, W. L. and Lin, G. H. (2003). A Bayesian like estimator of the process capability index Cpmk. Metrika, International Journal of Theoretical and Applied Statistics, 57(3), 303-312. Pearn, W. L., Lin, G. H. and Chang, C. S. (2003). On the distributions of the estimated process capability index Cpmk. Economic Quality Control, 18(1), 5-30. Pearn, W. L. and Lin, G. H. and Chen, K. S. (1998). Distributional and inferential properties of the process accuracy and process precision indices. Communications in Statistics: Theory and Methods, 27(4), 985-1000. Pearn, W. L., Lin, G. H. and Wang, K. H. (2004). Normal approximation to the distribution of the estimated yield index S pk . Quality and Quantity, 38(1), 95-111. Pearn, W. L. and Lin, P. C. (2002). Computer program for calculating the pvalue in testing process capability index Cpmk. Quality and Reliability Engineering International, 18(4), 333-342. Pearn, W. L. and Lin, P. C. (2004). Testing process performance based on the capability index Cpk with critical values. Computers and Industrial Engineering, 47, 351-369. Pearn, W. L. and Lin, P. C. (2004). Measuring process yield based on the capability index C m. International Journal of Advanced Manufacturing Technology, 24(7), 503-508. Pearn, W. L. and Lin, P. C. (2005). Process yield measure based on capability index Cpmk. IEM-0501 Working Paper. Pearn, W. L., Lin, P. C. and Chen, K. S. (2004). Distributional properties and implications of the estimated process capability index C"pk. Metrika., 60(2), 119-136. Pearn, W. L., Lin, P. C , and Chen, K. S. (2001). Estimating process capability index C"vmk for asymmetric tolerances: Distributional Properties, Metrika, 54(3), 261-279. Pearn, W. L., Lin, P. C , Chang, Y. C. and Wu, C. W. (2005). Quality yield measure for manufacturing processes with asymmetric tolerances. HE Transactions. Forthcoming. Pearn, W. L. and Shu, M. H. (2003). Lower confidence bounds with sample size information for Cpm with application to production yield assurance. International Journal of Production Research, 41(15), 3581-3599. Pearn, W. L. and Shu, M. H. (2003). An algorithm for calculating the lower confidence bounds of Cpu and CPL with application to low-drop-out linear regulators. Microelectronics Reliability, 43(3), 495-502. Pearn, W. L. and Shu, M. H. (2003). Manufacturing capability control for multiple power distribution switch processes based on modified Cpk MPPAC. Microelectronics Reliability, 43, 963-975.
366
Encyclopedia and Handbook of Process Capability Indices
Pearn, W. L. and Shu, M. H. (2004). Measuring manufacturing capability based on lower confidence bounds of Cpm£ applied to current transmitter process. International Journal of Advanced Manufacturing Technology, 23, 116-125. Pearn, W. L., Shu, M. H. and Hsu, B. M. (2004). Cpm MPPAC for manufacturing quality control applied to precision voltage reference process. International Journal of Advanced Manufacturing Technology, 23, 712-719. Pearn, W. L., Shu, M. H. and Hsu, B. M. (2004). Lower confidence bounds for Cpu and CPL based on multiple samples with application to production yield assurance of HSBA. International Journal of Production Research, 42(12), 2339-2356. Pearn, W. L., Shu, M. H. and Hsu. B. M. (2005). Monitoring manufacturing quality for multiple Li-BPIC processes based on capability index Cpmk. International Journal of Production Research, 43(12), 2493-2512. Pearn, W. L., Shu, M. H. and Hsu, B. M. (2005). Testing process capability based on Cpm in the presence of random measurement errors. Journal of Applied Statistics. Forthcoming. Pearn, W. L. and Wu, C. W. (2005). A Bayesian approach for assessing process precision based on multiple samples. European Journal of Operations Research, 165(3), 685-695. Pearn, W. L. and Wu, C. W. (2005). An effective modern approach for measuring high-tech product manufacturing process quality. Journal of the Chinese Institute of Industrial Engineering, 22(2), 119-133. Pearn, W. L. and Wu, C. W. (2005). Process capability assessment for index Cpk based on Bayesian approach. Metrika, International Journal of Theoretical and Applied Statistics. Forthcoming. Pearn, W. L. and Wu, C. W. (2005). Critical acceptance values and sample sizes of a new variables sampling plan for very low fraction of defectives. OMEGA, International Journal of Management Science. Forthcoming. Pearn, W. L. and Wu, C. W. (2005). Variables sampling plans with PPM fraction of defectives and process loss consideration. Journal of the Operational Research Society. Forthcoming. Pearn, W. L. and Wu, C. W. (2005). Production quality and yield assurance for processes with multiple independent characteristics. European Journal of Operations Research. Forthcoming. Pearn, W. L. and Wu, C. W. (2005). An effective decision making method for product acceptance. OMEGA, International Journal of Management Science. Forthcoming. Pearn, W. L., Wu, C. W. and Chuang, H. C. (2005). Procedures for testing manufacturing precision Cp based on (X, R) or (X, S) control chart samples. International Journal of Advanced Manufacturing Technology, 25, 598-607.
Bibliography
367
Pearn, W. L., Wu, C. W. and Lin, H. C. (2004). A procedure for supplier selection based on Cpm applied to STN-LCD processes. International Journal of Production Research, 42(13), 2719-2734. Pearn, W. L.„ Wu, C. W. and Wang, K. H. (2005). Capability for asymmetric tolerance non-normal processes applied to speaker driver manufacturing. International Journal of Advanced Manufacturing Technology, 25, 506-515. Pearn, W. L., Yang, S. L., Chen, K. S. and Lin, P. C. (2001). Testing process capability using the index Cpmk with an application. International Journal of Reliability, Quality and Safety Engineering, 8(1), 15-34. Pearn, W. L. and Yang, Y. S. (2003). Distributional and inferential properties of the estimated precision Cp based on multiple samples. Quality and Quantity, 37, 443-453. Pearson, E. S. and Tukey, J. W. (1965). Approximate means and standard deviations based on distances between percentage points of Frequency Curves. Biometrika, 52(3/4), 533-546. Perakis, M. and Xekalaki, E. (1998). On a refinement of certain process capability indices, Tech. Rep. No. 56, Dept. Statistics, Athens Univ. Econ. Bus., Greece. Persjin, M. and Nuland, Y.V. (1996-97). Relation between measurement system capability and process capability. Quality Engineering, 9(1), 95-98. Phillips, G. P. (1994). Target ratio simplifies capability index system, makes it easy to use Cpm. Quality Engineering, 7(2), 299-313. Pignatiello, J. J. and Ramberg, J. S. (1993). Process capability indices: just say "no". ASQC Quality Congress Transactions, Boston, 92-104. Pillet, M., Rochon, S. and Duclos, E. (1997). SPC - generalization of capability index C • case of unilateral tolerances. Quality Engineering, 10(1), 171176. Polansky, A. M. (1998). A smooth nonparametric approach to process capability, Qual. Rel. Eng. Intern., 14, 43-48. Polansky, A. M. (1998). A smooth nonparametric approach to process capability. Quality and Reliability Engineering International, 14, pp.43-48. Polansky, A. M. (2000). An algorithm for computing a smooth non-parametric capability estimate, J. Qual. Technol., 32, 284-289. Polansky, A. M., Chou, Y.-M. and Mason, R. L. (1998-99). Estimating process capability indices for a truncated distribution. Quality Engineering, 11(2), 257-265. Porter, L. J. and Oakland, J. S. (1990). Measuring process capability using indices - some new considerations. Quality and Reliability Engineering International, 6, 19-27. Porter, L. J. and Oakland, J. S. (1991). Process capability indices - an overview of theory and practice. Quality and Reliability Engineering International, 7, 437-448.
368
Encyclopedia and Handbook of Process Capability Indices
Prasad, S. and Calis, A. (1999). Capability indices for material balance accounting. Eur. J. Op. Res., 114, 93-104. Prasad, Sameer and Bramorski, Tom (1998). Robust process capability indices. Omega, 26(3), 425-435. Price, B. and Price, K. (1992). Sampling variability in capability indices. Technical Report 92-23, Wayne State University, Detroit, Michigan. Price, B. and Price, K. (1992). Estimating Cpk accuracy. ASQC Quality Congress Transactions, Nashville, Tenn, 9-15. Price, B. and Price, K. (1993). A methodology to estimate the sampling variability of the capability index Cpk. Quality Engineering, 5(4), 527-544. Pyzdek, T. (1992). Process capability analysis using personal computers. Quality Engineering, 4(3), 419-440. Quesenberry, C. P. (1988). An SPC approach to compensating a tool wear process. Journal of Quality Technology, 20(4), 220-229. Rado, L. G. (1989). Enhance product development by using capability indexes. Quality Progress, April, 22(4), 38-41. Ramakrishnan, B., Sandborn, P. and Pecht, M. (2001). Process capability indices and product reliability. Microelectronics Reliability, 41(12), 2067-2070 Rao, S. S. and Xia, W. (1999), Measuring quality by process capability indices and the effects of non-normality. Journal of Statistics and Management Systems. 2(2-3), 127-142. Raouf, A. and Ali, Z. (1992). On developing optimal process capability index. International Journal of Quality and Reliability Management, 9(7), 17-22. Rinne, M. and Mittag, H.-J. (1999). Prozessfahigkeitsmessung Fur die Industrielle Praxis (with English summary), Germany. Rivera, L. A. R., Hubele, N. F. and Lawrence, F. P. (1996). Cpk index estimation using data transformation. Computers and Industrial Engineering, 29 (4), 55-58. Rodriguer, R. N. (1992). Recent developments in process capability analysis. Journal of Quality Technology, 24, 176-187. Rosander, A. C. and Gunter, B. (1991). Letter to editor. Quality Progress, March (3), 9. Ruczinski, I. (1996). The Relation Between Cpm and the Degree of Includence. Doctoral Dissertation, Univ. Wurzberg, Germany. Runger, G. C. (1993), Designing process capability studies. Quality Progress, 26, 31-32. Sarkar, A. and Pal, S. (1997). Estimation of process capability index for concentricity. Quality Engineering, 9(4), 665-671. Schneider, H., Pruett, J. and Lagrange, C. (1995). Uses of process capability indices in the supplier certification process. Quality Engineering, 8(2), 225235.
Bibliography
369
Scholz, F. and Vengel, M. (1998). Tolerance bounds and Cpk confidence bounds under batch effects. In: Advances in Stochastic Models for Reliability, Quality and Safety, (W. Kahle, Evon Collani, J. Franz and U. Jensen, eds.) Birkhauser, Basel, Chapter. 24, 361-373. Shariari, H., Hubele, N. F. and Lawrence, F. P. (1995). A multivariate process capability vector, Proc. Jfii Industr Eng. Res. Conf., Nashville, TN. Shiau, J.-J. H., Chiang, C.-T. and Hung, H.-N. (1999). A Bayesian procedure for process capacity assessment. Qual. Rel. Eng. Internat., 15, 369-378. Shiau, J.-J. H. et al. (1999). A note on Bayesian estimation of process capability indices. Statistics and Probability Letters, 45: 215-224. Shore, H. (1997). Process capability analysis when data are auto-correlated. Quality Engineering, 9(4), 615-626. Shore, H. (1998). A new approach to analyzing non-normal quality data with application to process capability analysis. International Journal of Production Research, 36(7), 1917-1933. Singhal, S. C. (1990). A new chart for analyzing multi-process performance. Quality Engineering, 2(4), 379-390. Singhal, S. C. (1991-92). Multiprocess performance analysis chart (MPPAC) with capability zones. Quality Engineering, 4(1), 75-81. Singpurwalla, N. D. (1998). The stochastic control of process capacity indices. Test, 7(1), 1-33 (Discussion 33-79). Smith, B. (1993). Stalking six sigma, Business Month, 42-46. Smith, R. L. (1987). Estimating tails of probability distributions. Ann. Statist., 15, 1174-1207. Somerville, S. E. and Montgomery, D. C. (1996-97). Process capability indices and non-normal distributions. Quality Engineering, 9(2), 305-316. Spiring, F. A. (1989). An application of Cpm to the tool wear problem. ASQC Quality Congress Transactions, Toronto, 123-128. Spiring, F. A. (1991). An alternative to Taguchi's loss function. Quality Congress Transactions, Milwaukee, 660-665. Spiring, F. A. (1991). Assessing process capability in the presence of systematic assignable cause. Journal of Quality Technology, 23(2), 125-134. Spiring, F. A. (1991). The Cpm index: an alternative measure of process capability. Quality Progress, February, 57-61. Spiring, F. A. (1993). The reflected normal loss function. Canadian Journal of Statistics, 21, 321-330. Spiring, F. A. (1995). Process capability: a total quality management tool. Total Quality Management, 6(1), 21-33. Spiring, F. A. (1997). A unifying approach to process capability indices. Journal of Quality Technology, 29(1), 49-58.
370
Encyclopedia and Handbook of Process Capability Indices
Spiring, F., Leung, B., Cheng, S. and Yeung, A. (2003). A bibliography of process capability papers. Quality and Reliability Engineering International. 19(5), 445-460. Steenburg, T. J. and Gunter, B. H. (1991). The use and abuse of Cpk revisited. Quality Progress, 24(1), January, 90-94. Steenburgh, T. J. (1989). A method to study process capability for manufacturing strike teams. Xerox Corporation Technical Report. Steenburgh, T. J. (1991). Process capability indices for targets. Xerox Corporation Technical Report. Stoumbos, Z. G. (2002). Process capability indices: overview and extensions. Nonlinear Analysis: Real World Applications, 3(2), 191-210. Subbaiah, P. and Taam, W. (1993). Inference on the capability index: Cpm. Communications in Statistics: Theory and Methods, 22(2), 537-560. Sullivan, L. P. (1984). Reducing variability, a new approach to quality. Quality Progress, 17, 15-21. Sullivan, L. P. (1985). Cp and Cpk (Letters). Quality Progress, 18(4), April, 7-8. Sun, F.; Laramee, J. and Ramberg, J. (1996). On Spiring's inverted normal loss function. The Canadian Journal of Statistics, 24, 241-249. Sundaraiyer, V. H. (1996). Estimation of a process capability index for inverse Gaussian distribution. Communications in Statistics: Theory and Methods, 25(10), 2381 -2393. Taam, W., Subbaiah, P. and Liddy, J. W. (1993). A note on multivariate capability indices. Journal of Applied Statistics, 20(3), 339-351. Tang, L. C , Than, S. E. and Ang, B. W. (1997). A graphical approach to obtaining confidence limits of Cpk. Quality and Reliability Engineering International, 13, 337- 346. Tang, P. F. and Barnett, N. S. (1998). Capability indices for multivariate processes. Tech. Rep., Div. Comput. Math. Sci., Victoria Univ., Melbourne, Australia. Tong, Y. L. (1990). The Multivariate Normal Distribution. Springer-Verlag, New York. Tong, L.-I. and Chen, J.-P. (1998). Lower confidence limits of process capability indices for non-normal process distributions. International Journal of Quality and Reliability Management, 15(8/9), 907-919. Tsui, K. L. (1997). Interpretation of process capability indices and some alternatives. Quality Engineering, 9(4), 587-596. Vannman, K. (1995). A unified approach to capability indices. Statistica Sinica, 5, 805-820. Vannman, K. (1997). A general class of capability indices in the case of asymmetric tolences. Communication in Statistics: Theory and Methods, 26(8), 2049-2072.
Bibliography
371
Vannman, K. (1997). Distribution and moments in simplified form for a general class of capability indices. Communication in Statistics: Theory and Methods, 26(1), 159-179. Vannman, K. (1998). Families of capability indices for one-sided specification limits. Statistics, 31, 43-66. Vannman, K. and Hubele, N. F. (2003). Distributional properties of estimated capability indices based on subsamples. Quality and Reliability Engineering International, 19(5), 445-460. Vannman, K. and Kotz, S. (1995). A superstructure of capability indices distributional properties and implications. Scandinavian Journal of Statistics, 22, 477-491. Vannman, K. and Kotz, S. (1996). A superstructure of capability indices asymptotics and its implications. International Journal of Reliability, Quality and Safety Engineering, 2(4), 343-360. Vasilopoulos, A. V. and Stamboulis, A. P. (1978). Modification of control chart limits in the presence of data correlation. Journal of Quality Technology, 10(1), 20-30. Veevers, A. (1995). A capability index for multiple response, CSIRO Math. Statist. Rep. DMS-095, Australia. Veevers, A. (1998). Viability and capability indices for multi-response processes. J. Appl. Statist. , 25, 545-558. Veevers, A. (1999). Capability Indices For Multiresponse Processes. In: Statistical Process Monitoring and Optimization (S. H. Park and G. G. Vining, eds.), 241-256. Vrieland, B. and Chin, E. (1990). Bellcore variables rating plan to assist in SQC at the Atlanta Workshop. Presented paper, ATT Bell Laboratories. Walker, E. (1995). Review of Kotz, S. and Johnson, N. L. (1993), J. Amer. Statist. Assoc., 90, 1127. Wallgren, E. (1996). Properties of the Taguchi capability index for Markov dependent quality characteristics. Technical report, Dept of Mathematics University of Orebro, Sweden. Wang, C. M. and Lam, C. T. (1992). Confidence limits for proportion of conformance. Journal of Quality Technology, 28(4), 439-445. Wang, F. K and Chen, C. Y. (1998). Confidence intervals for multivariate capability indices using bootstrap. Journal of the Chinese Institute of Industrial Engineering, 15(5), 429-438. Wang, F. K and Chen, J. C. (1998). Capability indices using principle components analysis. Quality Engineering, 11(1), 21-27. Wang, F. K. and Du, T. C. T. (2000). Using principle component analysis in process performance for multivariate data. Omega, 28(2), 185-194.
372
Encyclopedia and Handbook of Process Capability Indices
Wang, F. K. and Hubele, N. F. (1999). Quality evaluation using geometric distance approach. International Journal of Reliability, Quality and Safety Engineering, 6(2), 139-153. Wang, F. K. and Hubele, N. F. (2000). Quality evaluation of geometric tolerance regions in form and location. Quality Engineering, 14(2), 203-209. Wang, F. K., Hubele, N. F., Lawrence, F., Miskulin, J. and Shahriari, H. (2000). Comparison of three multivariate process capability indices. Journal of Quality Technology, 32(3), 263-275. Wasserman, G. S., and Franklin, L. A. (1992). Standard bootstrap confidence interval estimates of Cpk. Computers and Industrial Engineering, 22(2), 171176. Wasserman, G. S., Mohsen, H. A. and Franklin, L. A. (1991). A program to calculate bootstrap confidence intervals for the process capability index, Cpk. Communications in Statistics: Simulation and Computation, 20, 497-510. Wasserman, G. S. and Franklin, L. (1992). Standard bootstrap confidence interval estimate of Cpk. Computers and Industrial Engineering. 22(2): 171176. Wheeler, D. J. (1999). Beyond Capability Confusion. SPC Press, Knoxville, TN. Wierda, S. J. (1992). Multivariate quality control-estimation of the percentage of good product. Tech. Rep. Univ. Groningen Netherlands. Wierda, S. J. (1993). A multivariate process capability index. ASQC Quality Congress Transactions-Boston. 47, 342-348. Wierda, S. J. (1994). Multivariate Statistical Process Control. (Ph.D. Thesis in Economics), Wolters-Noordhof, Groningen Netherlands. Woodall, W. H. (1999). Research issues and ideas in statistical process control. Journal of Quality Technology. 31(4), 376-386. Woodall, W. H. (2000). Controversies and contradictions in statistical process control. Journal of Quality Technology. 32(4), 341-350. Wright, P . A. (1995). A process capability index sensitive to skewness, Journal of Statistical Computation and Simulation, 52, 195-203. Wright, P. A. (1998). The probability density function of process capability index Cpmk. Communications in Statistics: Theory and Methods, 27(7), 1781-1789. Wright, P. A. (2000). The cumulative distribution of process capability index Cvm, Statist. Probab. Letters, 47, 249-251. Wright, P. A. (1998). The probability density function of process capability index Cpmk. Communications in Statistics: Theory and Methods, 27(7), 1781 -1789. Wu, C. C. and Tang, G. R. (1998). Tolerance designed for products with asymmetric quality losses. International Journal of Production Research, 36, 9, 2529-2541. Wu, C. C , Chen, K. S. and Kuo, H. L. (2001). Decision making in assessing process capability indices. International Journal on Information and Management Sciences, 12, 29-42.
Bibliography
373
Wu. C. W. and Pearn, W. L. (2005). Bayesian approach for measuring EEPROM process capability based on one-sided indices CPU and CPL. International Journal of Advanced Manufacturing Technology. Forthcoming. Wu, C. W. and Pearn, W. L. (2005). Measuring manufacturing capability for coupler and wavelength division multiplexer. International Journal of Advanced Manufacturing Technology, 25, 533-541. Wu, C. W. and Pearn, W. L. (2005). Capability testing based on Cpm with multiple samples. Quality and Reliability Engineering International, 21(1), 29-42. Wu, H. H. and Swain, J. J. (2001). A Monte Carlo comparison of capability indices when processes are non-normally distributed. Quality and Reliability Engineering International, 17(3), 219-231. Wu, H-H., Swain, J. J., Farrington, P.A. and Messimer, S. L. (1999). A weighted variance capability index for general non-normal processes. Quality and Reliability Engineering International, 15, 397- 402. Xie, M. and Goh, T. N. (1999). Statistical techniques for quality. The TQM Magazine, 11(4), 238-241. Xie, M., Tsui, K. L., Goh, T. N. and Cai, D. Q. (2002). Process capability indices for a regularly adjusted process. International Journal of Production Research, 40(10), 2367-2377. Yang, K. and Hancock, W. M. (1990). Statistical quality control for correlated samples. International Journal of Production Research, 28(3), 595-608. Yeh, A. B. and Bhattacharya, S. (1998). A robust process capability index. Communications in Statistics: Simulation and Computation, 27(2), 565-589. Yeh, A. B. and Chen, H. (2001). A nonparametric multivariate process capability index. International Journal of Modeling and Simulation, 21, 218-223. Zhang. N. F. (1998). Estimating process capability indexes for autocorrelated data. Journal of Applied Statistics, 25, 559-574. Zhang, N. F. (2001). Combing process capability indices from a sequence of independent samples. International Journal of Production Research, 39(13), 2769-2781. Zhang, Y. and Feng, X. D. (1999). Target allocation for maximizing wear allowance of running fits based on process capability. Quality Engineering, 12(2), 169-176. Zhang, B. and Ni, J. (2001). Relative probability index C^,: an alternative process capability index. Quality Engineering, 14(2), 267-278. Zang, N. F., Steinback, G. A. and Wardrop, D. M. (1990). Interval estimation of process capability index Cpk. Communications in Statistics: Theory and Methods, 19(12), 4455-4470. Zimmer, L. S. and Hubele, N. F. (1997/8). Quantiles of the sampling distribution of Cpm. Quality Engineering, 10, 309-329.
374
Encyclopedia and Handbook of Process Capability Indices
Zimmer, L.S.; Hubele, N. F. and Zimmer, W. J. (2001). Confidence intervals and sample size determination of Cpm. Quality and Reliability Engineering International, 17(1), 51-68. Zwick, D. (1995). A hybrid method for fitting distributions to data and its use in computing process capability indices. Quality Engineering, 7(3), 601-613.
Index
Clements' method, 299 process yield, 41 sampling plan, 237 weakness, 66
Assignable cause variation, 289 Asymmetric tolerances, 181-217 Autoregressive model, 281, 292 Beazley, 78, 80, 262 Binomial distribution, 235 Bissell's approximation, 54 Boyles, R.A., asymmetric, 181-91 Cpk, 42, 66 Cpm, 69, 73, 75-94 Spk, 135, 163
confidence intervals, 160 estimator, 161 process yield, 147 n asymmetric tolerances, 197-202 confidence intervals, 78-80 estimators, 88 expected loss, 68 hypothesis testing, 86 measurement errors, 263 multivariate, 329 non-normality, 299-301 Clements' method, 301 Capability, 1-6 Capability Potential Index, see Cp Capability ratio (CR), 8, 144, 220 Central Limit Theorem, 137-8, 315 Chebyshev inequality, 324
confidence intervals, 13 estimator, 9-10 hypothesis testing, 12-3 measurement errors, 252 Clements' method, 299 weakness, 30 chi-square distribution, 50 confidence intervals, 51-4, Bayesian interval estimate, 62-4 Bayesian-like, 64 hypothesis testing, 55 measurement bias and error, 258-62 375
376
Encyclopedia and Handbook of Process Capability Indices
Chi-square distribution, 21, 24, 49, 52, 89-90, 122, 129, 202-4, 218, 223-5, 270 non-central, 78, 264 Clements, J.A., 295, 299 Clements' technique, 304 Column vectors, 224 Contaminants, accumulation of, 284 Control charts, 23, 26, 128, 281, 315, 320 Shewhart, 162, 284, 315-6, 320 Correlated data, 282, 337, 343 Cumulative distribution function (CDF), 9 Data collection, 14, 20, 83, 87, 93, 128, 154 Data transformation techniques, non-normal process, 295 Deleryd, M., 163, 172 Drift, process, 284 Expected relative loss, 72, 96, 1034, 108 Expected values, 120, 138 Exponential distribution, 304 Folded normal distribution, 33, 36, 48,49 Ford Motor Company, 8 Franklin, L.A., 14, 50, 84, 185, 295, Gamma function, 11, 283 Gauge capability, 249-252, 255, 258 Gilchrist, W.G., 303 Guirguis, G., 48 Gunter, B.H., 294
Gupta, A.K., 218 Heavlin, W.D., 13, 50 Hobbs, J.R., 120 Hoffman, L.L., 52 Hsiang,. T.C., 57, 95, 177 Hsu, N.-F., 120 Huang, D.-Y., 218 Hubele, N.F., 58, 80, 129, 154, 263 Hurley, P., 50, 80, 185 Hypothesis testing, Cp, 14, 18, 23 Cpk, 54 Cpm> 86 Cpm/b 125 Spk, 143
Johnson, N.L., 11, 41, 47, 70, 96, 176, 182, 258, 295, 303 Kane, V.E., 7, 97, 134, 146, 181, 220, 298 Kirmani, S.N.U.A., 15, 58, 87, 154 Kotz, S., 11, 41, 70, 114, 181, 258, 284, 295 Kurtosis, 300, 304, 312, 319-20 Kushler, R, 50, 80, 185 Lee, R.F., 137, 218 Leone, F.C, 33, 48, Li, H., 20, 57, 154 Lovelace, C.R., 26, 284, 295 Lu, K., 305, 313 Marcucci, M.O., 78, 263 Mean square error (MSB), 100, 115, 202, 210, 254, 285, 291 Measurement system, 250-1 Median, 297, 299-301, 310-2 Montgomery, D.C. Cp, 8, 23
377
Index Cpk, 44, 132 CPU/CPL, 148
MPPAC, 165 non-normal 293, 295 gauge error, 250-1 sampling plans, 234 tool wear, 282, 289 Nagahata, H., 50-1 Nagata, Y., 50-1 Non-central chi-square distribution, 74, 76, 78-9, 83, 89, 201, 202, 265 Nonconformities (NC), 43, 72, 114, 147, 165, 172, 232, 268, 277 Normal distribution, 9, 34, 42, 77, 99, 111, 134, 148, 182, 219, 235, 249, 280, 294, 328 folded, 33, 36, 49, 131 multivariate data, 329-30, 338, 343 process spread, 30 specification limits, 10 Oakland, J. S., 281 One-sided specifications, 148 Operating characteristic (OC) curve, 232, 235 Owen, D.B., 10, 47, 149, 181, 198, 203, 235, 273 Pearn, W.L., Cp, 11 C 0 ,32 Cpk, 47 Cpm, 74 C-pmk: H O
Pearson curves, 299, 303, 312 Pearson system, 303
Percentage nonconformities, see Nonconformities Porter, L.J., 281 Probability density function (pdf), Cp, 10 G„, 33 Cpm, 74
Process capability analysis, 7, 21, 111, 157, 172, 216, 280 Process capability indices (PCIs), 1,2,5 basic structure, 7-8 non-normal populations, 296 Process Capability Ratio, 8, 148, 220 Process capability studies, 8, 146, 158, 279-280 Process drift, 284 Process yield, 41-44 p-value, 19, 28 Quadratic form, 335 Quantiles, 12, 298-9, 340 Relative loss, 72, 96, 103-4, 108 Relative worth, 96 Rodriguez, R.N., 48, 51, 296 Sample sizes, 14, 22, 238, 245, 300 Sampling plan, 231-248 Shapiro, S.S., 298 Shapiro-Wilk test, 262, 268, 277 Shewhart, W.A., 162, 315-6 Shewhart control charts, 315 Skewness, 294, 296, 307-9, 312, 315-6 Special cause, 281 Spiring, F.A., 28-9, 61, 92, 109, 181, 288, 191
378
Encyclopedia and Handbook of Process Capability Indices
Standard deviation, 8-10, 15, 20 Statistical control, 8, 39, 46, 59, 117, 169, 280, 331 Statistical process control (SPC), 9, 282 Subbaiah, P., 80 Systematic assignable causes, 28890 Taam, W., 80, 237, 331, 341 Taguchi, G., 67, 95, 176, 198 Taguchi method, 167 Target values, 71, 170, 243, 338 Taylor expansion, 137, 145 Temperature change, 284
Tolerances, 45 Tool wear, 279-292 Tukey.J.W., 76 Unbiased estimators, 36, 128, 149, 160 Vannman, K., 76, 78 Vectors, 337 Wallgren, E., 292 Wasserman, G.S., 51, 185, 295, 307 Weibull distributions, 294 Wesolowsky, G.O., 69 Wright, P.A., 120, 307-9 Zhang, N.F., 47, 50, 88
Encyclopedia and Handbooh of Process Capability. Indices R Comprehensive Exposition of Quality Control Measures
'. i
tt
This unique volume pro\ ides an up-to-date and detailed description ol the various process capabilitv indices widely (and sometimes misleadingly) used in the applications at production sites. The authors. who are internationally recognized experts in this area with numerous contributions to the field, provide a lucid exposition, which
covers a l l
the m a i n a s p e c t s , d e v e l o p m e n t s
and
advances.
The concept of Process Capability Index (PCI) is barely 20 years o l d . but the multitude of available versions can overwhelm even the most seasoned practitioner. The organized and self-contained presentation of the material starling from I980's primitive indices (Cp and Cpk) up to the newly proposed indices for the cases of multiple dependent characteristics results in an authoritative and indispensable reference. A proper balance between theoretical investigation and "rule-of-lhumb" practical procedures is maintained in order to eliminate the tensions among various methodologies of assessing the capabilitv of industrial processes.
www.worldscientmc.com