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!
) = m22(f>2 one has chaos. Each homogeneous domain in this scenario is much greater than the size of the observable part of the universe. 8p 8V37t m 2 4> • (6) The first models of chaotic inflation were based on the — p = —7— 5 theories with polynomial potentials, such as V{<j>) = Taking into account (3) and also the expansion of the ±(m2/2)(j)2 + (//4)0 4 . But the main idea of this scenario universe by about 1030 times after the end of inflation, is quite generic. One should consider any particular potential one can obtain the following result for the density V( Mp in bang loses its dominant position, being removed to the the theory (m2/2)cj)2). This result is based on the investi- indefinite past. gation of the probability of quantum jumps with amplitude From this new perspective many old problems of 5(j) » H/2n. cosmology, including the problem of initial conditions, look Until now we have considered the simplest inflationary much less profound than they seemed before. In many model with only one scalar field, which had only one mini- versions of inflationary theory it can be shown that the mum of its potential energy. Meanwhile, realistic models fraction of the volume of the universe with given properties of elementary particles propound many kinds of scalar (with given values of fields, with a given density of matter, fields. For example, in the unified theories of weak, strong etc.) does not depend on time, both at the stage of inflation and electromagnetic interactions, at least two other scalar and even after it. Thus each part of the universe evolves © Physica Scripta 2000 0.3MP [6]. Suppose way baryons are produced in the Affleck-Dine scenario [42], for definiteness that initially 4> is large and negative, and and in the electro weak theory [43]. inflation ends at <j> ~ — 0.3M p . After that the field (/> rollsUsually only a small fraction of the energy of the inflaton to 4> = 0, then it grows up to 10 _1 M P ~ 1018 GeV, and finally field ~ lO - 2 ^ 2 is transferred to the particles x when the field rolls back and oscillates about > = 0 with a gradually (j) approaches the point <j> = 0 for the first time [44]. The role decreasing amplitude. of the parametric resonance is to increase this energy We will assume that g > 10~5 [38], which implies exponentially within several oscillations of the inflaton field. gMp > 102m for the realistic value of the mass m ~ But suppose that the particles x interact with fermions \j/ with 10~6Mp. Thus, immediately after the end of inflation, when the coupling h^/^x- If this coupling is strong enough, then x cj) ~ Mp/3, the effective mass g\4>\ of the field / is much particles may decay to fermions before the oscillating field greater than m. It decreases when the field = 0. If the inflaton is completely decoupled, then once inflation id z , ; z = act>/H , (3) d\ ends it will oscillate about the minimum of the potential, 32S: dx2with the cycle-average of the energy density decreasing as a~3, i.e., as a matter-dominated universe. But at the end i.e., the action in conformal time x {dx2 = a2(t)dt2) for a of inflation the universe is cold and frozen in a low-entropy scalar field in Minkowski space, with mass-squared state: the only degree of freedom is the zero-momentum m\ = — z~1d2z/dx2. Here, the scalar field u is a combination mode of the inflaton field. It is necessary to "defrost" the of metric fluctuations Sg^ and scalar field fluctuations universe and turn it into a "hot" high-entropy universe with 5 cos <> / 2 2 i„ = M20 cos (j) + M02 sin 0 + 2M\ \ sin <j> cos 0
© Physica Scripta 2000
Inflationary Cosmology
171
of the scalar field in some parts of the universe. These regions fields exist. The potential energy of these scalar fields expand at a greater rate than their parent domains, and may have several different minima. This means that the same quantum fluctuations inside them lead to production of theory may have different "vacuum states," corresponding new inflationary domains which expand even faster. This to different types of symmetry breaking between fundamensurprising behavior leads to an eternal process of tal interactions, and, as a result, to different laws of low-energy physics. self-reproduction of the universe. To understand the mechanism of self-reproduction one As a result of quantum jumps of the scalar fields during should remember that the processes separated by distances inflation, the universe may become divided into infinitely / greater than H~i proceed independently of one another. many exponentially large domains that have different laws This is so because during exponential expansion the distance of low-energy physics. Note that this division occurs even between any two objects separated by more than H~l is if the whole universe originally began in the same state, corgrowing with a speed exceeding the speed of light. As a responding to one particular minimum of potential energy. result, an observer in the inflationary universe can see only If this scenario is correct, then physics alone cannot provide the processes occurring inside the horizon of the radius H"1. a complete explanation for all properties of our part of An important consequence of this general result is that the the universe. The same physical theory may yield large parts process of inflation in any spatial domain of radius H~x of the universe that have diverse properties. According to occurs independently of any events outside it. In this sense this scenario, we find ourselves inside a four-dimensional any inflationary domain of initial radius exceeding H~x domain with our kind of physical laws not because domains with different dimensionality and with alternate properties can be considered as a separate mini-universe. To investigate the behavior of such a mini-universe, with are impossible or improbable, but simply because our kind an account taken of quantum fluctuations, let us consider of life cannot exist in other domains. This consideration is based on the anthropic principle, an inflationary domain of initial radius H~l containing sufficiently homogeneous field with initial value (j> » M p . which was not very popular among physicists for two main Equation (3) implies that during a typical time interval reasons. First of all, it was based on the assumption that A? = H~l the field inside this domain will be reduced by the universe was created many times until the final success. Acj) = (Mp/4n(j)). By comparison this expression with Second, it would be much easier (and quite sufficient) to \8(J)(x)\ « (H/2n) = J(2V(c^)/3nM2) ~ (m0/3M p ) one can achieve this success in a small vicinity of the Solar system rather than in the whole observable part of our universe. easily see that if <j> is much less than cf>* ~ (M p /3) ^/(Mp/m), then the decrease of the field 4> due to its classical Both objections can be answered in the context of the motion is much greater than the average amplitude of the theory of eternal inflation. First of all, the universe indeed quantum fluctuations 5(j> generated during the same time. reproduces itself in all its possible versions. Second, if the But for 4> ^ 4>* o n e has 8
Physica Scripta T85
172
Andrei Linde
in time, but the universe as a whole may be stationary, and the properties of its parts do not depend on the initial conditions [12]. Of course, this happens only for the (rather broad) set of initial conditions which lead to self-reproduction of the universe. However, only finite number of observers live in the universes created in a state with initial conditions which do not allow self-reproduction, whereas infinitely many observers can live in the universes with the conditions which allow self-reproduction. Thus it seems plausible that we (if we are typical, and live in the place where most observers do) should live in the universe created in a state with initial conditions which allow self-reproduction. On the other hand, stationarity of the self-reproducing universe implies that an exact knowledge of these initial conditions in a self-reproducing universe is irrelevant for the investigation of its future evolution [12].
requires the universe to be exponentially large, flat and homogeneous prior to the stage of the pre-big-bang inflation [15,16]. Thus it would be very difficult to replace the usual post-big-bang inflation by the pre-big-bang one. And if the post-big-bang inflation is indeed necessary, then the pre-big-bang stage will have no observational manifestations. Therefore it is very desirable to find those versions of string theory and supergravity where a consistent post-big-bang inflationary theory can be developed. This is a difficult problem, but not hopeless. Perhaps the easiest way to find a consistent inflationary model in supersymmetric theories is based on the hybrid inflation scenario [17]. The simplest version of this scenario is based on chaotic inflation in the theory of two scalar fields with the effective potential
V{a, 0) = 1 (M2 - key2)2 + y 4>2 + y * V •
(8)
The effective mass squared of the field a is equal to —M2 +g2(j>2. Therefore for
Physica Scripta T85
© Physica Scripta 2000
Inflationary Cosmology
173
nor of the type of [26]. A consistent model of one-field open inflation was proposed only very recently [32]. Effective potential in this model has a sharp peak at large 4>. During the tunneling and soon after it the condition \V"\
© Physica Scripta 2000
Physica Scripta T85
174
Andrei Linde
arately in accordance with perturbation theory of particle particles produced each time will be proportional to the decay. However, recently it was understood that in many number of particles produced before. This leads to explosive inflationary models the first stages of reheating occur in a process of particle production [38]. regime of a broad parametric resonance. To distinguish this Bosons produced at that stage are far away from thermal stage from the subsequent stages of slow reheating and equilibrium and have enormously large occupation thermalization, it was called preheating [38]. The energy numbers. Explosive reheating leads to many interesting transfer from the inflaton field to other bose fields and effects. For example, specific nonthermal phase transitions particles during preheating is extremely efficient. may occur soon after preheating, which are capable of To explain the main idea of the new scenario we will con- restoring symmetry even in the theories with symmetry 16 sider first the simplest model of chaotic inflation with the breaking on the scale ~ 10 GeV [39]. These phase 2 2 effective potential V(<j>) = (m /2)(j) , and with the inter- transitions are capable of producing topological defects such action Lagrangian —{\/2)g2(j)2x1 — h}jj\jJx. We will take as strings, domain walls and monopoles [40]. Strong m = 10~ 6 M p , as required by microwave background ani- deviation from thermal equilibrium and the possibility of sotropy [6], and in the beginning we will assume for sim- production of superheavy particles by oscillations of a relaplicity that x particles do not have a bare mass, i.e. tively light inflaton field may resurrect the theory of GUT baryogenesis [41] and may considerably change the mx(4>) = g\)\. In this model inflation occurs at \
© Physica Scripta 2000
Inflationary Cosmology This is exactly what happens in the theories where the post-inflationary motion of the inflaton field occurs along a flat direction of the effective potential. In such theories the standard scenario of reheating does not work because the field > does not oscillate. Until the invention of the instant preheating scenario the only mechanism of reheating discussed in the context of such models was based on the gravitational production of particles [45]. The mechanism of instant preheating is much more efficient. After the moment when % particles are produced their energy density grows due to the growth of the field >. Meanwhile the energy density of the field cp moving along a flat direction of F($) decreases extremely rapidly, as a~6(t). Therefore very soon all energy becomes concentrated in the particles produced at the end of inflation, and reheating completes. 7. Conclusions During the last 20 years inflationary theory gradually became the standard paradigm of modern cosmology. But this does not mean that all difficulties are over and we can relax. First of all, inflation is still a scenario which changes with every new idea in particle theory. Do we really know that inflation began at Planck density 1094 g/cm 3 ? What if our space has large internal dimensions, and energy density could never rise above 1025 g/cm 3 [46]? Was there any stage before inflation? Is it possible to implement inflation in string theory/M-theory? Inflationary theory evolved quite substantially, from old inflation to the theory of chaotic eternal self-reproducing universe. We learned that in most versions of inflationary theory the universe must be flat, and the spectrum of density perturbations should be also flat ("flat paradigm"). But we also learned that there are models where the spectrum of density perturbations is not flat. If necessary, it is possible to obtain not only adiabatic perturbations, but also isocurvature nongaussian perturbations with a very complicated spectrum. It is even possible to have inflation with This situation makes some observers unhappy. Few years ago it seemed that one can easily kill inflationary theory if one finds, for example, that the density of the universe is not equal to the critical density. It would be an important scientific result. Now the situation changed. If we find that Q = 1, it will be a confirmation of inflationary theory because 99% of inflationary models predict that Q=\, and no other theories make this prediction. On the other hand, if observations will show that Q ^ 1, it will not disprove inflation. Indeed, the only consistent theory of a large homogeneous universe with Q ^ 1 that is available now is based on inflationary cosmology. Still inflationary models are falsifiable. Each particular inflationary model can be tested, and many of them have been already ruled out by comparison of their predictions with observational data. A new generation of precision experiments in cosmology are going to make our life even more complicated and interesting. However, it is difficult to disprove the basic idea of inflation. To make my position more clear, I would like to use an analogy from the history of the standard model of electroweak interactions [47]. Even though this model was developed by Glashow, Weinberg and Salam in the 60's, © Physica Scripta 2000
175
it became popular only in 1972, when it was realized that gauge theories with spontaneous symmetry breaking are renormalizable [48]. However, it was immediately pointed out that this model is far from being perfect. In particular, it was not based on the simple group of symmetries, and it had anomalies. Anomalies could destroy the renormalizability, and therefore it was necessary to invoke a mechanism of their cancellation by enlarging the fermion sector of the theory. This did not look very natural, and therefore Georgi and Glashow in 1972 suggested another model [49], which at the first glance looked much better. It was based on the simple group of symmetry 0(3), and it did not have any anomalies. However, after the discovery of neutral currents, which could not be described by the 0(3) model, everybody forgot about the issues of naturalness and simplicity and returned back to the more complicated Glashow-Weinberg-Salam model. This model has about twenty free parameters which so far did not find an adequate theoretical explanation. Some of these parameters may appear rather unnatural. The best example is the coupling constant of the electron to the Higgs field, which is 2 x 10~6. It is a pretty unnatural number which is fine-tuned in such a way as to make the electron 2000 lighter than the proton. It is important, however, that all existing versions of the electroweak theory are based on two fundamental principles: gauge invariance and spontaneous symmetry breaking. As far as these principles hold, we can adjust our parameters and wait until they get their interpretation in a context of a more general theory. We do not call it "fine tuning," we call it "fitting." This is the standard way of development of the elementary particle physics. For a long time cosmology developed in a somewhat different way, because of the scarcity of reliable observational data. Ten years ago many different cosmological models (HDM, CDM, Q=l, £ 2 « 1 , etc.) could describe all observational data reasonably well. The main criterion for a good theory was its beauty and naturalness. Now the situation is changing. Cosmology gradually becomes a normal experimental science, where the results of observations play a more important role than the considerations of naturalness. But in our search for a correct theory we cannot give up the requirement of its internal consistency. In particle physics the main principle which made this theory internally consistent was gauge invariance. It seems that in cosmology something like inflation is needed to make the universe large and homogeneous and resolve a pletora of other cosmological problems. It is encouraging that even if we find that the universe is open, we will be able to fit it into inflationary cosmology. But it seems equally important that so far we did not find any other theoretical framework apart from inflation which would explain the extraordinary homogeneity of an open universe. Thus, even if we find that Q ^ 1, we will rule out a very large class of inflationary models but we will be unable to rule out the idea of inflation. What's about microwave background anisotropy and the theory of large scale structure of the universe? Again, observational data may confirm inflationary theory because most of inflationary models predict perturbations with specific properties. However, if it so happens that inflationary perturbations are in a conflict with observational data, then one can easily propose inflationary Physica Scripta T85
176
Andrei Linde
models where inflation solves homogeneity, isotropy and other problems, but produces extremely small density perturbations. Then one will be free to use his best noninflationary mechanism for production of density perturbations (strings, textures, etc.). Perhaps there is a specific way to disprove inflationary cosmology by comparing its predictions with observational data: If we find that the universe rotates as a whole, or has any other anisotropy which is described by vector perturbations of metric, then it will be very difficult to make it compatible with inflation [50]. Indeed we know that scalar and tensor perturbations can be produced during inflation, but vector perturbations, just like other vector fields, can hardly be produced. But what if we find out a nontrivial way of producing vector perturbations, just like we found a way to produce an inflationary universe with Q ^ 1? It seems that the only sure way to kill inflationary cosmology (if one really wants to do it) is to suggest a better cosmological theory.
References 1. Starobinsky, A. A., J. Exp. Theor. Phys. Lett. 30, 682 (1979); Phys. Lett. 91B, 99 (1980). 2. Mukhanov, V. F. and Chibisov, G. V., J. Exp. Theor. Phys. Lett. 33,532 (1981). 3. Hawking, S. W., Phys. Lett. 115B, 295 (1982); Starobinsky, A. A., ibid. 117B, 175 (1982); Guth, A. H. and Pi, S.-Y., Phys. Rev. Lett. 49, 1110 (1982); Bardeen, J., Steinhardt, P. J. and Turner, M. S., Phys. Rev. D 28, 679 (1983); Mukhanov, V. F., J. Exp. Theor. Phys. Lett. 41, 493 (1985). 4. Guth, A. H., Phys. Rev. D 23, 347 (1981). 5. Kirzhnits, D. A., J. Exp. Theor. Phys. Lett. 15, 529 (1972); Kirzhnits, D. A. and Linde, A. D., Phys. Lett. 42B, 471 (1972); Sov. Phys. J. Exp. Theor. Phys. 40, 628 (1974); Ann. Phys. 101, 195 (1976); Weinberg, S., Phys. Rev. D9, 3320 (1974); Dolan, L. and Jackiw, R., Phys. Rev. D9, 3357 (1974). 6. Linde, A. D., "Particle Physics and Inflationary Cosmology" (Harwood, Chur, Switzerland, 1990). 7. Guth, A. H. and Weinberg, E. J., Nucl. Phys. B 212 321 (1983). 8. Linde, A. D., Phys. Lett. 108B, 389 (1982); 114B, 431 (1982); 116B, 335, 340 (1982); Albrecht, A. and Steinhardt, P. J., Phys. Rev. Lett. 48, 1220 (1982). 9. Linde, A. D., Phys. Lett. 129B, 177 (1983). 10. Steinhardt, P. J., in: "The Very Early Universe," (G.W. Gibbons, S.W. Hawking, S. Siklos, eds), (Cambridge U.P. Cambridge, England, 1982), p. 251; Linde, A. D., "Nonsingular Regenerating Inflationary Universe," (Cambridge University preprint, 1982); Vilenkin, A., Phys. Rev. D 27, 2848 (1983). 11. Linde, A. D., Phys. Lett. B 175, 395 (1986). 12. Linde, A. D., Linde, D. A. and Mezhlumian, A., Phys. Rev. D 49, 1783 (1994). 13. Veneziano, G., Phys. Lett. B 265, 287 (1991); Gasperini, M. and Veneziano, G., Astropart. Phys. 1, 317 (1993). 14. Brustein, R. and Veneziano, G., Phys. Lett. B 329, 429 (1994); Kaloper, N., Madden, R. and Olive, K. A., Nucl. Phys. B 452, 677 (1995); Copeland, E. J., Lahiri, A. and Wands, D., Phys. Rev. D 50 4868 (1994); Kaloper, N., Madden, R. and Olive, K. A., Phys. Lett. B 371 34 (1996); Easther, R., Maeda, K. and Wands, D., Phys. Rev. D 53 4247 (1996). 15. Turner, M. S. and Weinberg, E., Phys. Rev. D 56, 4604 (1997). 16. Kaloper, N., Linde, A. and Bousso, R. "Pre-Big-Bang Requires the Universe to be Exponentially Large from the Very Beginning," hep-th/9801073. 17. Linde, A. D., Phys. Lett. B 259, 38 (1991); Phys. Rev. D 49, 748 (1994).
Physica Scripta T85
18. Copeland, E. J., Liddle, A. R., Lyth, D. H., Stewart, E. D. and Wands, D., Phys. Rev. D, 49, 6410 (1994); Dvali, G., Shafi, Q. and Schaefer, R. K., Phys. Rev. Lett. 73, 1886 (1994); Linde, A. D. and Riotto, A., Phys. Rev. D 56, 1841 (1997). 19. Binetruy, P. and Dvali, G., Phys. Lett. B 388, 241 (1996); Halyo, E., Phys. Lett. B 387, 43 (1996). 20. Lyth, D. H. and Riotto, A., hep-ph/9807278. 21. Coleman, S. and De Luccia, F., Phys. Rev. D 21, 3305 (1980). 22. Gott, J. R. Nature 295, 304 (1982); Gott, J. R. and Statler, T. S., Phys. Lett. 136B, 157 (1984). 23. Bucher, M., Goldhaber, A. S. and Turok, N., Phys. Rev. D 52, 3314 (1995); Yamamoto, K., Sasaki, M. and Tanaka, T., Astrophys. J. 455, 412 (1995). 24. Linde, A. D., Phys. Lett. B 351, 99 (1995); Linde, A. D. and Mezhlumian, A., Phys. Rev. D 52, 6789 (1995). 25. Hawking, S. W. and Moss, I. G., Phys. Lett. HOB, 35 (1982). 26. Hawking, S. W. and Turok, N., Phys. Lett. B 425, 25 (1998); Hawking, S. W. and Turok, N., Phys. Lett. B432 271 (1998), hep-th/9803156. 27. Linde, A. D., Phys. Rev. D58, 083514 (1998), gr-qc/9802038. 28. Barvinsky, A. O., hep-th/9806093, gr-qc/9812058. 29. Hartle, J. B. and Hawking, S. W., Phys. Rev. D 28, 2960 (1983). 30. Linde, A. D., Zh. Eksp. Teor. Fiz. 87, 369 (1984) [Sov. Phys. J. Exp. Theor. Phys. 60, 211 (1984)]; Lett. Nuovo Cim. 39, 401 (1984); Zeldovich, Ya.B. and Starobinsky, A. A., Sov. Astron. Lett. 10, 135 (1984); Rubakov, V. A., Phys. Lett. 148B, 280 (1984); Vilenkin, A., Phys. Rev. D 30, 549 (1984). 31. Vilenkin, A., hep-th/9803084; Unruh, W., gr-qc/9803050; Bousso, R. and Linde, A. D., gr-qc/9803068; Garriga, J., hep-th/9803210, hep-th/9804106; Bousso, R. and Chamblin, A., hep-th/9805167. 32. Linde, A. D., Phys. Rev. D 59, 023503 (1999). 33. Linde, A. D., Sasaki, M. and Tanaka, T., astro-ph/9901135. 34. Garcia-Bellido, J., hep-ph/9803270. 35. Garcia-Bellido, J., Garriga, J. and Montes, X., Phys. Rev. D 57, 4669 (1998). 36. Garcia-Bellido, J., Garriga, J. and Montes, X., hep-ph/9812533. 37. Dolgov, A. D. and Linde, A. D., Phys. Lett. 116B, 329 (1982); Abbott, L. F., Fahri, E. and Wise, M., Phys. Lett. 117B, 29 (1982). 38. Kofman, L. A., Linde, A. D. and Starobinsky, A. A., Phys. Rev. Lett. 73, 3195 (1994); Kofman, L., Linde, A. A. and Starobinsky, A. A., Phys. Rev. D 56, 3258-3295 (1997). 39. Kofman, L. A., Linde, A. D. and Starobinsky, A. A., Phys. Rev. Lett. 76, 1011 (1996); Tkachev, I., Phys. Lett. B 376, 35 (1996). 40. Tkachev, I., Kofman, L. A., Linde, A. D., Khlebnikov, S. and Starobinsky, A. A., in preparation. 41. Kolb, E. W., Linde, A. and Riotto, A., Phys. Rev. Lett. 77,4960 (1996). 42. Anderson, G. W., Linde, A. D. and Riotto, A., Phys. Rev. Lett. 77, 3716 (1996). 43. Garcia-Bellido, J., Grigorev, D., Kusenko, A. and Shaposhnikov, M., hep-ph/9902449; Garcia-Bellido, J. and Linde, A., Phys. Rev. D 57, 6075 (1998). 44. Felder, G., Kofman, L. A. and Linde, A. D., hep-ph/9812289. 45. Ford, L. H., Phys. Rev. D 35, 2955 (1987); Spokoiny, B., Phys. Lett. B 315, 40 (1993); Joyce, M., Phys. Rev. D 55, 1875 (1997); Joyce, M. and Prokopec, T., Phys. Rev. D 57, 6022 (1998); Peebles, P. J. E. and Vilenkin, A., astro-ph/9810509. 46. Arkani-Hamed, N., Dimopoulos, S., Kaloper, N. and March-Russell, J., hep-ph/9903224. 47. Glashow, S. L., Nucl. Phys. 22, 579 (1961); Weinberg, S., Phys. Rev. Lett. 19, 1264 (1967); Salam, A., in: "Elementary Particle Theory," (edited by N. Svartholm), (Almquist and Wiksell, Stockholm, 1968), p. 367. 48. 't Hooft, G., Nucl. Phys. B 35, 167 (1971); Lee, B. W., Phys. Rev. D 5, 823 (1972); Lee, B. W. and Zinn-Justin, J., Phys. Rev. D 5,3121 (1972); 't Hooft, G. and Veltman, M., Nucl. Phys. B 50, 318 (1972); Tyutin, I. V. and Fradkin, E. S., Sov. J. Nucl. Phys. 16, 464 (1972); Kallosh, R. E. and Tyutin, I. V., Sov. J. Nucl. Phys. 17, 98 (1973) [Yad. Fiz. 17, 190 (1973)]. 49. Georgi, H. and Glashow, S. L., Phys. Rev. Lett. 28, 1494 (1982). 50. Ellis, J. and Olive, K. A., Nature 303, 379 (1983).
© Physica Scripta 2000
Physica Scripta.Vol. T85, 177-182, 2000
Quintessence and the Missing Energy Problem Paul J. Steinhardt Department of Physics, Princeton University, Princeton, NJ 08650, USA Received November 29, 1998; accepted August 2, 1999 PACS
Ref: 98.80.-k, 98.70Vc, 98.65.Dx, 98.80.Cq
Abstract Recent evidence suggests that the total matter density of the universe is significantly less than the critical density. The shortfall may be explained by curvature (an open universe), vacuum energy density (a cosmological constant), or quintessence (a time-evolving, spatially inhomogeneous component with negative pressure). In all three cases, a key problem is to explain the initial conditions required to have the energy density nearly coincident with the matter density today. A possible solution is "tracker fields," a form of quintessence with an attractor-like solution which leads to the same conditions today for a very wide range of initial conditions. Tracker field make quintessence a more viable candidate for the missing energy component and produce more distinctive predictions.
region lies far from Qm = 1. Explaining the shortfall constitutes the missing energy problem. As convincing as the case for Qm < 1 may appear to be, the conclusion should be accepted reluctantly. Every resolution of the missing energy problem entails introducing a profound and bizarre addition to our cosmological model compared to the Qm = 1 case. Especially for those whose research is infieldsoutside cosmology, a reasonable reaction is to ignore the problem until yet further evidence is gathered. At the same time, the strength of the current evidence summarized in Figs. 1 and 2 may explain why some cosmologists, including myself, have felt compelled to look ahead and explore solutions to the problem [4].
1. The problem The missing energy problem is emerging as one of the most profound puzzles of cosmology. The problem is best illustrated by displaying the current observational constraints in a plot of Qm versus Ho, where Qm is the ratio of the baryonic plus dark matter density to the critical density and Ho is the the Hubble constant HQ = 100/z km/s/Mpc [1,2]. See Fig. 1. The observational constraints cut a series of swaths through this plane. In Fig. 2, the swaths have been removed and all that remains is the region which is in concordance with all current observations [1,3]. Given more than a dozen independent constraints, the existence of a region of concordance is remarkable. Equally significant is the fact that two or more independent constraints limit each boundary of the concordance region. That is, the concordance region can move by a significant amount in any direction only if two or more independent constraints are incorrect. The most notable feature is that the concordance
2. The Solutions One possible explanation for the missing energy problem is that the universe is open - space is curved and there is no missing energy needed to explain the difference between Qm and unity. A theoretical difficulty with this option is that it contradicts inflationary cosmology.[5-8] To be sure, cosmological models can be rigged so that there is just the right amount of inflation to make an Qm = 0.3 open universe today [9]. However, inflation is a rapid, runaway process which removes the curvature exponentially fast. As a result, open inflation requires extraordinary tuning
1.0
0.8
0.6 f"Pfmova e ;
n„ 0.4
0.2
0.0 0.4
cluster abundance & evolution: baryon fraction 0.5
0.6
0.7
0.8
0.9
1.0
Hubble Constant (x 100 km/sec/Mpc) Fig. 2. A plot showing the range of Qm and the Hubble constant which are in concordance with all current observations (the shaded region between the vertical dashed lines). Along the boundary are named the constraints that fix the given boundary section. Shown separately with dashed lines are current Fig. 1. Some of the current observational constraints in the Qm-h plane, illus- constraints on the Hubble constant (65 ± 10 km s_1 Mpc-1). Note that each boundary is fixed by two or more constraints. trating how the constraints cut swaths through parameter space. Hubble Expansion Rate (x 100 km/sec/Mpc)
© Physica Scripta 2000
Physica Scripta T85
178
Paul J. Steinhardt
of parameters to control the amount of inflation precisely; with this tuning, inflation loses its simplicity and power as a scientific theory. If inflation is to remain a compelling, explanatory theory, then it is essential that the flatness be verified observationally. In fact, the flat universe is somewhat preferred by current observations. Evidence from Type IA supernovae [25] and from the cosmic microwave background anisotropy [26] disfavor the open option. Both sets of results are inconclusive, but near-future developments should be definitive. A second possibility is that the universe is flat, but there is a non-zero vacuum density or cosmology constant (A). A cosmological constant is, by definition, static and spatially uniform. At present, cosmological models with a mixture of A and cold dark matter (ACDM) fit all current observations well. A third possibility, the focus of this paper, is that the universe is flat and the missing energy consists of quintessence, [10] a dynamical, evolving, spatially inhomogeneous component. The term "quintessence" derives from the ancient term for the fifth element after earth, air, fire and water, an all-pervasive, weakly interacting component. In the current context, quintessence refers to a weakly interacting component in addition to the four already known to influence the evolution of the universe: baryons, leptons, photons, and dark matter. More specifically, quintessence refers to a component with negative pressure p and an equation-of-state (w = p/p) with 0 > w > — 1. (A cosmological constant has w precisely equal to —1.) Unlike a cosmological constant, the quintessential pressure and energy density evolve in time, and w may also. Furthermore, because the quintessence component evolves in time, quintessence is necessarily spatially inhomogeneous. The quintessential concept is related to the notion of a time-dependent cosmological constant, which dates back to the early days of relativistic cosmology, when it was ascribed political as well as scientific significance [27]. In these and some later treatments, an energy component was assumed whose density decreases more slowly than the matter density but which is smoothly distributed. However, it is important to note that a time-dependent cosmological constant, in the sense of a spatially uniform but temporally varying energy, is inconsistent with the equivalence principle [10]. Quintessence, which has non-negligible spatial fluctuations on large scales (> 100 Mpc), is the closest approximation that is physically consistent.
describe the inflaton field in inflationary cosmology [6,7]. The major difference is that the energy scale associated with quintessence is much lower so that it remains a subdominant contribution to the Friedmann equations that describe the expansion of the universe until recently. In this framework, quintessence can be viewed as the onset of a low-energy, late inflationary epoch. The pressure of the scalar field, p — 1 Q2 ~ ^ ( 2 ) is negative if the field rolls slowly enough that the kinetic energy density is less than the potential energy density. The ratio of kinetic-to-potential energy is determined by equation-of-motion for the scalar field: Q + 3HQ+V(Q)
=0
(1)
where
*-$-¥«•+>*
<2)
a(t) is the Robertson-Walker scale factor, and G is Newtons' constant. Depending on the detailed form of V(Q), the equation-of-state w can vary between 0 and —1. For most potentials, w evolves slowly over time. The field is assumed to couple only gravitationally to matter. In all cases, the Q-energy density decreases with time as l/a 3 ' 1+H, e) 5 so negative pressure corresponds to a density which decreases more slowly than 1/a3. Spatial fluctuations in Q evolve over time due to the gravitational interaction between Q and clustering matter. The perturbations are important because they can leave a distinguishable imprint on the CMB and large-scale structure. To determine how the perturbations evolve, specifying w is insufficient. One must know the response of the component to perturbations. This can be defined by specifying the sound speed cs as a function of wavenumber k or, alternatively, by specifying the equations-of motion (from which the perturbative equations can be derived). Note that it is possible, in principle, to have two fluids with the same w but different cs, which would lead to distinct observational predictions. For a scalar field, the equation-of-motion for the perturbations SQ in synchronous gauge is: SQ + 3HSQ + (k2 + a2 V"(Q))5Q = -\hQ,
(3)
where the dot represents the derivative with respect to conformal time, the prime represents the derivative with respect to Q, and hk is the kth fourier mode of the perturbed metric. The source term in Eq. (3) has several important properties. First, any realistic cosmological model has clustering 3. Quintessential basics matter, so hk must be non-zero. Also, except for the limit The general description of quintessence given in the previous of a constant V(Q) (which corresponds to a true section incorporates many proposals for the missing energy cosmological constant), Q is non-zero. Hence, the source that have been proposed over the past decade or so [10-19]. term must be non-zero overall for any p > — p (w > — 1). For example, under some conditions, light, nonabelian This is significant because it ensures that Q cannot be cosmic defects, strings or walls, evolve into a frozen network smoothly spread. Even if SQ is non-zero initially, the source whose equation-of-state corresponds to w = —1/3 and term causes perturbations to grow. (If the universe consisted w = —2/3, respectively [12]. The example considered here initially of a uniform distribution of quintessence, hk could and in most papers is the energy density associated with remain zero and Q could remain smoothly spread. However, a scalar field Q rolling down the potential V(Q), which once there is a mixture of components and at least one can have an equation-of-state anywhere between zero and clusters, smoothness is no longer possible.) A further conse— 1. The equations are nearly identical to the equations that quence of the source term is that the perturbations in Q Physica Scripta T85
© Physica Scripta 2000
Quintessence and the Missing Energy Problem 1.0 r-
0.8
0.6
0.4
/
0.2
~"
-»v___——"""^
o.o L — • 1.0
i 0.8
•
' 0.6
.
<
•
0.4
' 0.2
.
1 0.0
equation-of-state w Fig. 3. For QCDM models, the range of Qm and w in concordance with all current measurements. The full region represents the combination of COBE constraints from measurements of the microwave background anisotropy on large angular scales and other astrophysical observations. The right end represents the parameter range cut off by current measurements of the microwave background anisotropy at small angular scales.
observed today are extremely insensitive to the initial conditions for 8Q. Assuming that SpQ/pQ is comparable to the perturbations in other energy components, the transient solution to the perturbation equation is negligible today compared to the particular solution set by the source term.
4. Quintessential motivation Why consider quintessence if its effect on the expansion of the universe is similar to a cosmological constant? The reasons are that: (a) quintessence has different implications for fundamental physics; (b) quintessence may fit the observational data better than cosmological constant; and (c) quintessence may solve cosmoological problems which a cosmological constant cannot. Distinguishing quintessence from a cosmological constant is important both for cosmology and for fundamental physics. A vacuum density or cosmological constant (A) is static and spatially uniform. Its value is set once and for all in the very early universe. Hence, A is tied directly to quantum gravity physics near the Planck scale. Quintessence is new dynamics at ultra-low energies (energy scale ~ 1 meV today), perhaps a harbinger of a whole spectrum of new low-energy phenomena. At present, both models with cosmological constant and cold dark matter (ACDM) and models with a mixture of quintessence and cold dark matter (QCDM) agree with current data [3]. However, QCDM fits marginally better. Figure 3 illustrates the range of Qm and WQ which fit all current observations. Note that the allowed region only includes negative pressure whether or not recent supernovae observations are included. The allowed region encompasses cosmological constant w = — 1 and a substantial region with w > — 1. The best-fit model [3] corresponds to a quintessence model with w & —2/3, although the difference in fit between this model and the best-fit A model is small. © Physica Scripta 2000
179
For quintessence composed of scalar fields, there is the added observational constraint that the coupling to ordinary matter be sufficiently suppressed to evade fifth force and other constraints on light fields [20]. This constraint does not project into the Qm-w plane. Whatever form the missing energy takes, two new cosmological problems arise. First, the component must be comparable in density today to the critical density, 10~47 GeV 4 . This small value is forced by the observation that QQ and Qm are comparable today. We will refer to the puzzle of explaining this tiny energy as the "fine-tuning problem." A second problem arises when the cosmological model is extrapolated back in time to the very early universe, at the end of inflation, say. The curvature, the cosmological constant, and, in general, quintessence decrease at different rates than the matter density. At the end of inflation, it appears that the ratio of missing energy to matter and radiation energy must be specially fixed to an extraordinarily small value (of order 10~100 for the case of vacuum density) in order to have the ratio evolve to be of order unity today. Accounting for the special ratio in the early universe will be referred to as the "coincidence problem [23]." The coincidence problem is a generalization of the flatness problem pointed out by Dicke and Peebles [24]. The fine-tuning and coincidence problems are often treated as one and the same because they occur together for models with vacuum density or curvature. Vacuum density and curvature are fixed by a single, unvarying constant set at the beginning of the universe. The constant must be tiny compared to particle physics scales (the fine-tuning problem) and must be set to an infinitesimal fraction of the initial total energy density to have the correct density today (the coincidence problem). For quintessence, though, the two problems separate [21,22]. Because quintessence evolves temporally and can interact with matter gravitationally or directly, the quintessential field might be able to begin with natural initial conditions and, through its dynamics, evolve to the correct value today. Indeed, recent progress in resolving the coincidence problem along these lines has been made, as described in the next section. The fine-tuning problem requires a separate solution, as described in Section 5.2.
5. A quintessential solution to the cosmic coincidence problem Recently, Zlatev et al. have introduced the notion of tracking fields and tracking potentials as a quintessential solution to the coincidence problem. Tracker fields have an equation-of-motion with attractor-like solutions in which a very wide range of initial conditions rapidly converge to a common, cosmic evolutionary track. The initial value of PQ can vary by nearly 100 orders of magnitude without altering the cosmic history. The acceptable initial conditions include the natural possibility of equipartition after inflation - nearly equal energy density in Q as in the other 100-1000 degrees of freedom (i.e., QQI % 10~3). Furthermore, the resulting cosmology has desirable properties. The equation-of-state WQ varies according to the background equation-of-state WB- When the universe is radiation-dominated (WB = 1/3), then WQ is less than or Physica Scripta T85
180
Paul J. Steinhardt
equal to 1 /3 and pQ decreases less rapidly than the radiation density. When the universe is matter-dominated (WB = 0), then WQ is less than zero and pQ decreases less rapidly than the matter density. Eventually, pQ surpasses the matter density and becomes the dominant component. At this point, Q slows to a crawl and WQ -> — 1 as QQ -*• 1 and the universe is driven into an accelerating phase. These properties seem to match current observations well [3]. 5.1. Trackers and initial conditions The key property of tracker fields is that they rapidly converge to the tracker solution for an extraordinarily wide range of initial conditions. The tracker solution differs from a classical dynamics attractor because it is time-dependent: during tracking, QQ increases steadily. This contrasts with the "self-adjusting" solutions based on exponential potentials recently discussed by Ferreira and Joyce [14,15,18] in which QQ is constant during the matter dominated-epoch. For the self-adjusting solution, QQ is constrained to be small (QQ < 0.15) at the beginning of matter domination in order that large-scale structure be formed, but then it cannot change thereafter. Hence, the quintessence remains a small, subdominant component. The tracker solution is more desirable, though, because it enables the g-energy to eventually overtake the matter density and induce a period of accelerated expansion, which produces a cosmology more consistent with measurements of the matter density, large scale structure, and supernovae observations [3,25].
Z+1 Fig. 4. Energy density versus red shift for the evolution of a tracker field. For computational convenience, z = 1012 (rather than inflation) has been arbitrarily chosen as the initial time. The white bar on left represents the range of initial PQ that leads to undershoot and the grey bar represents overshoot, combining for a span ofmore than 100 orders of magnitude if we extrapolate back to inflation. The solid black circle represents the unique initial condition if the missing energy is vacuum energy density. The solid thick curve represents an "overshoot" in which p e begins from a value greater than the tracker solution value, decreases rapidly and freezes, and eventually joins the tracker solution.
oscillating around and then rapidly converging to the tracker solution. Next, let's consider the "overshoot" case in which pQ is To study how the tracker solution is approached, it is initially much greater than the tracker value pQ. This case useful to rewrite the equation-of-motion Eq. (1) in the form: includes equipartition where, for 100 — 1000 degrees of freedom, say, pQ « 10~(2~3)pB. Now, the initial Q is at a high I d lnx V (4) point in the potential where V/ V is much greater than the V 1 + WQ 6d In a tracker solution. At this steep point in the potential, the field where x = (1 4- WQ)/(1 - WQ) = \ Q^/V is the ratio of the is rapidly accelerated to the point where its energy is kinetic to potential energy density for Q and prime means dominantly kinetic and the potential is irrelevant. The field derivative with respect to Q. The ± sign depends on whether overshoots the tracker by a calculable amount [21,22] that V > 0 or V < 0, respectively. The tracking solution (to does not depend on the potential. Eventually, it freezes when which general solutions converge) has the property that its kinetic energy is red shifted away. From that point on, it WQ is nearly constant and lies between WB and — 1. For behaves just as in the undershoot case above. An example 1 + WQ = 0(1), Q2 « QQH2 and the equation-of-motion, of an overshoot solution is illustrated in Fig. 4. Eq. (4), dictates that Only if Q dominates over the radiation density initially, PQ » pB, does tracking fail. In this case case, Q rolls rapidly H 1 (5) down hill and overshoots the attractor solution to an ~v" /Or. ~Q extraordinary degree and is "deep-frozen." The field freezes for a tracking solution; we shall refer to this as the "tracker at a value so far downhill that V'/V « : H/Q today and condition." In order to maintain this condition, it is necess- the attractor solution has not been reached yet. However, ary that V'/V be decreasing as Q rolls downhill and QQ exempting the initial condition pQ » pB still leaves a range of over 100 orders of magnitude of viable initial conditions increases. Solutions to the equation-of-motion are drawn towards a including the most likely case, equipartition. common solution, Q(t) whose energy density is Pg(t). First, let's consider the "undershoot" case where pQ is initially much smaller than the tracker value pQ. See Fig. 4. In this 5.2. Testing for trackers case, the initial Q is at a low point on the potential where Hence, testing for the existence of tracking solutions reduces V'/V is much smaller than the tracker value. This value to a simple condition on V(Q) without having to solve the is so small that the Hubble damping term dominates the equation-of-motion directly. It suffices to check whether >\ and is nearly constant equation-of-motion and the field is nearly frozen in place. T =V"V/(V')2 The frozen field cannot evolve until the universe ages (and (|d(r - \)/Hdt\ «; \r - 11) over the range of plausible initial H decreases) to the point where V'/V = H/Q; that is, Q Q. Here we are assuming WQ < wB; that is, the energy density is now at the ideal point on the potential required by the in Q decreases more slowly than the background density. tracker solution for the given H. The field begins to move, The range of initial Q extends from 2max> where K(Q max ) Physica Scripta T85
© Physica Scripta 2000
Quintessence and the Missing Energy Problem is equal to the initial background energy density p B , to Qmm where V{Qm\n) is equal to the background density at matter-radiation equality. The energy density spans more than 100 orders of magnitude over this range. The condition that r« constant can be evaluated by testing whether
r
r(v/v)
« i
(6)
over the range between Q max and Qm\n as V decreases. These conditions encompass an extremely broad range of potentials, including inverse power-law potentials (V(Q) = M 4+a IQ* for a > 0) and combinations of inverse power-law terms {e.g., V(Q) = M 4 e x p ( M / 0 ) . Some potentials of this form are suggested by particle physics models with dynamical symmetry breaking or nonperturbative effects [29-34].
181
V(Q). Now consider the entire family of tracker solutions for each given V(Q) and consider whether QQ is more likely to dominate late in the universe for one V(Q) or another. In general, QQ is proportional to a3(-WB~w^ oc t2(WB-WQ)K\+WB)^ w h e r e i t c a n b e s h o w n t hat [22] WB -
WQ
=
2(r - \)(wB +1) i + 2(r-i)
(7)
Hence, we find QQ OC tp where P =
4(r-i) l + 2(r - i) "
(8)
A satisfying explanation as to why QQ dominates at late times, rather than early, would be if QQ automatically changes behavior at late times. This is not the case for two special cases (which have the been the focus of earlier papers): K ~ l / 2 a and V ~ exp(/?£?). As the universe evolves, T — 1 and, hence, P are constant; consequently, 5.3. Tracking and the fine-tuning problem Some tracker solutions do not require small mass par- QQ grows as the same function of time throughout the ameters in the scalar potential to obtain a small energy den- radiation- and matter-dominated epochs. However, these sity needed today to explain Qm, which is relevant to the are the exception, rather than the rule. For more general potentials, P increases as the universe fine-tuning problem. The one free dimensional parameter in V(Q), M, is determined by the observational constraint, ages. Consider first a potential which is the sum of two QQ « 0 . 7 today. To have Qn«0.7 today requires inverse power-law terms with exponents ai < o<2. The term 47 4 V(Q « Mp) « pm, where pm « 10" GeV is the current with the larger power is dominant at early times when Q matter density. The condition imposes the constraint is small, but the term with the smaller power dominates M«(pmAf«)1/("+4). For low values of a or for the at late times as Q rolls downhill and obtains a larger value. exponential potential, this forces M to be a tiny mass, as Hence, the effective value of a decreases and T — 1 oc 1/a low as 1 meV for the exponential case. However, we note increases; the result is that P increases at late times. For that M > 1 GeV — comparable to particle physics scales more general potentials, such as V ~ exp(l/g), the effective — is possible for a > 2. This case exemplifies how a small value of a decreases continuously and P increases with time. density may be achieved by functional forms combining a Fig. 5 illustrates the comparison in the growth of P. An increasing P means that QQ grows more rapidly as the typical particle physics mass scale with a large expectation universe ages. Figure 6 compares a tracker solution for a value for Q. Hence, while this is not the real aim of tracking, pure inverse power-law potential (V ~ l/Q6) model with it is interesting to note that the tracker solution does not require the introduction of a new mass hierarchy in funda- a tracker solution for V ~ e x p ( l / 0 , where the two solutions have been chosen to begin at the same value of QQ. (The start mental parameters. time has been chosen arbitrarily at z = 1017 for the purposes of this illustration.) Following each curve to the right, there 5.4. The QQ-WQ relation is a dramatic (10 orders of magnitude) difference between An important consequence of the tracker solutions is the the time when the first solution (solid line) meets the backprediction of a relation between WQ and QQ today [22]. ground density versus the second solution (dot-dashed line). Because tracker solutions are insensitive to initial con- That is, beginning from the same QQ, the first tracker solditions, both WQ and QQ only depend on V(Q). Hence, ution dominates well before matter-radiation equality and for any given V(Q), once QQ is measured, WQ is determined. the second (generic) example dominates well after In general, the closer that QQ is to unity, the closer WQ is matter-domination. The difference is less dramatic as a to —1. However, the fact that Qm > 0.2 today means that increases for the pure inverse power-law model; also, as a. WQ cannot be so close to —1. We find that WQ > - 0 . 8 for increases, the value of WQ today (given Q > 0.2) approaches m practical models.[21,22] This QQ-WQ condition on the zero and the universe does not enter a period of acceleration equation-of-state forces WQ to be sufficiently different from by the present epoch. Hence, a significant conclusion is that — 1 that the tracker field proposal can be distinguished from the pure exponential (exp(/J0) and inverse power-law the cosmological constant, e.g., by measurements of the (M 4+(X /2 o: ) models are atypical; the generic potential expansion history of the universe using Type IA (V(M/Q)) has properties that make it more plausible that supernovae. QQ dominates late in the history of the universe and induces a recent period of accelerated expansion. 5.5. Why now? Why should QQ only begin to dominate and initiate a period of accelerated expansion late in the history of the universe? The tracker field notion offers some explanation. Thus far, 6. Conclusions we have imagined fixing M to guarantee that Any resolution of the missing energy problem requires a QQ = 1 — Qm has the measured value today. The choice dramatic departure from the simplicity of the standard cold amounts to considering one tracker solution for each dark matter picture. Within a few years, improved obser© Physica Scripta 2000
Physica Scripta T85
182
Paul J. Steinhardt not require a small mass parameter. This progress makes the notion of quintessence as the missing matter a much more interesting and compelling concept. In addition, tracking, like inflation, is a general idea which may find other applications in cosmology not yet anticipated.
10
~ tp
Q
£6
Acknowledgements
QL
V - exp(1/Q) - 1
The work described in this paper is the result of collaborations with Limin Wang, Ivaylo Zlatev and Rob Caldwell. This research was supported by the US Department of Energy grant DE-FG02-95ER40893 (Penn).
/
V-1/Q a References 10'
10'
10"
10°
10 J
101
z+1 Fig. 5. A plot of P/P0 versus /, where QQ a tp and P0 is the initial value of P. The plot compares pure inverse power-law (V ~ l/Q") potentials for which P is constant with a generic potential (e.g. K~exp ( 1 / 0 ) for which P increases with time.
\ V~1/Q6
10" \
s^
10°
<\ J^
>
o
O
\
10'1 -
-
io"20 -
' ^ % \
10"
~ N
\ V - exp(1/Q) - 1
10" 10"'
- \ 1016
1013
101°
10 7
10*
X
10'
z+1 Fig. 6. A plot comparing two tracker solutions for the case of a V ~ \IQ potential (solid line) and a V ~ e x p ( 1 / 0 potential (dot-dash). The dashed line is the background density The two tracker solutions were chosen to have the same energy density initially. The tracker solution for the generic example ( K ~ e x p ( l / 0 ) reaches the background density much later than for the pure inverse-power law potential. Hence, QQ is more likely to dominate late in the history of the universe in the generic case.
vations will determine whether cosmologists are forced to this consideration. In the meantime, the imaginative exploration of possible solutions to the missing energy problem may lead to new concepts in cosmology. Quintessence serves as an instructive example. As a general notion, quintessence has many degrees of freedom and no clear advantage compared to the cosmological constant solution. Now, however, the tracker field form of quintessence introduces a new approach to two seemingly intractable issues: the problem of initial conditions and the issue of why pQ is dominating today rather than at some early epoch. It leads to a new, unanticipated prediction a relation between Qm = 1 - QQ and WQ today that makes tracker fields distinguishable from a cosmological constant. And, it may Physica Scripta T85
1. See, for example, Ostriker, J. P. and Steinhardt, P. J. Nature 377, 600 (1995), and references therein. 2. Turner, M. S., Steigman, G. and Krauss, L. Phys. Rev. Lett. 52, 2090 (1984). 3. Wang, L., Caldwell, R. R., Ostriker, J. P. and Steinhardt, P. J., in preparation. 4. See paper by Turner, M. in this volume for a fuller discussion of the current observational evidence. 5. Guth, A. H., Phys. Rev. D23, 347, (1981). 6. Linde, A. D., Phys. Lett. 108B, 389 (1982). 7. Albrecht, A. and Steinhardt, P. J., Phys. Rev. Lett. 48, 1220 (1982). 8. Guth, A. H. and Steinhardt, P. J., "The Inflationary Universe" in "The New Physics," (ed. by P. Davies), (Cambridge U. Press, Cambridge, 1989) pp. 34-60. 9. See. for example, Hawking, S. W. and Turok, N., Phys. Lett. B425, 25 (1998); and Bucher, M., Goldhaber, A. S. and Turok, N., Phys. Rev. Lett. Phys. Rev. D52, 3314 (1995). 10. Caldwell, R. R., Dave, R. and Steinhardt, P. J., Phys. Rev. Lett. 80, 1582 (1998). 11. Turner, M. S. and White, M., Phys. Rev. D 56, R4439 (1997). 12. Spergel, D. and Pen, U.-L., Astrophys. J. 491, L67, (1997). 13. Weiss, N., Phys. Lett. B 197, 42 (1987). 14. Ratra, B. and Peebles, P. J. E., Phys. Rev. D 37, 3406 (1988); Peebles, P. J. E. and Ratra, B., Astrophys. J. 325, L17 (1988). 15. Wetterich, C , Astron. Astrophys. 301, 32 (1995). 16. Frieman, J. A., et al., Phys. Rev. Lett. 75, 2077 (1995). 17. Coble, K. Dodelson, S. and Frieman, J., Phys. Rev. D 55, 1851 (1995). 18. Ferreira, P. G. and Joyce, M. Phys. Rev. Lett. 79, 4740 (1997); Phys. Rev. D 58, 023503 (1998). 19. Copeland, E. J., Liddle, A. R. and Wands, D. Phys. Rev. D57, 4686 (1998). 20. Carroll, S., Phys. Rev. Lett. 81, 3067 (1998). 21. Zlatev, I. Wang, L. and Steinhardt, P. J., Phys. Rev. Lett. 82, 896 (1999). 22. Steinhardt, P., Wang, L. and Zlatev, I., submitted to PRD (1998). 23. Steinhardt, P., in "Critical Problems in Physics," (ed. by V.L. Fitch and D.R. Marlow), (Princeton U. Press, 1997). 24. Dicke, R. H. and Peebles, P. J. E. in "General Relativity: An Einstein Centenary Survey," (ed. by S.W. Hawking & W. Israel), (Cambridge Univ. Press 1979). 25. Perlmutter, S., etal., Nature 391, 51 (1998); A. G. Riess, et al, Astron. J. 116, 1009 (1998). 26. Tegmark, M., Ap. J. 514, L69 (1999) and Ap. J. 519, 513 (1999). 27. See Ref. 33 in Overduin, J. M. and Cooperstock, F. I., Phys. Rev. D58 043506 (1998); Bronstein, M., Phys. Z. Sowjet Union 3, 73 (1933). 28. Binetruy, P., Phys. Rev. D 60, 063522 (1999). 29. Anderson, G. W. and Carroll, S. M., astro-ph/9711288. 30. Barreiro, T., de Carlos, B. and Copeland, E. J., Phys. Rev. D57, 7354 (1998). 31. Binetruy, P., Gaillard, M. K. and Wu, Y.-Y., Phys. Lett. B412, 288 (1997); Nucl. Phys. B493, 27 (1997). 32. Barrow, J. D. Phys. Lett. B235, 40 (1990). 33. Hill, C. and Ross, G. G., Nucl. Phys. B 311, 253 (1988); Phys. Lett. B 203, 125 (1988). 34. Affleck, I. et al., Nuc. Phys. B 256, 557 (1985).
© Physica Scripta 2000
Physica Scripta.Vol. T85, 183-190, 2000
Ultra High Energy Cosmic Rays - an Enigma A. A. Watson* Department of Physics and Astronomy, University of Leeds, Leeds LS2 9JT, UK Received March 3, 1999; accepted August 2, 1999
PACS Ref: 96.40.De, 96.40.Pq, 98.70.Sa
Abstract The methods of detection of high energy cosmic rays are briefly described and it is shown that results on the energy spectrum and arrival distribution above 4 x 1 0 " eV confound theoretical expectation. Some of the ideas which have been put forward to explain the data are outlined. There is an urgent need for better statistics as only about 12 events above 1020 eV have so far been detected. Furthermore the limit to the energy which cosmic rays can have is not known. In addition to the Pierre Auger Observatory, which is briefly described, the Hi-Res fluorescence detector is about to take data and an ambitious satellite observatory (Airwatch/OWL) is under discussion.
1. Introduction The cosmic ray energy spectrum extends from 10 9 eV to at least 3 x 1020eV with no evidence that the highest energy recorded thus far is Nature's upper limit. Although cosmic rays have been known since 1912 it remains uncertain as to where they come from. We believe that the majority of the particles are the nuclei of the common elements and at low energies, were direct observations can be made, the distribution of the elements is strikingly similar to that found in ordinary material in the solar system with the exception of elements such as Li, Be and B. These are the products of fragmentation of heavier nuclei and are very over-abundant by comparison with their local galactic abundance. The energy spectrum is nearly featureless lacking the lines or dips which would characterise an electromagnetic spectrum covering so many decades. It is because the particles are charged and travel through the tangled magnetic fields in our galaxy and beyond that it so difficult to trace them back to their origins. Only at the very highest energies does it seem reasonable to contemplate cosmic ray astronomy but the studies that have been possible so far have shown only a very unexpected isotropy. This result, coupled with the surprisingly great energies of the rarest particles, presents an enigma. In this short review the methods of detection of the highest energy cosmic rays will be described and results on the energy spectrum, mass composition and arrival direction distribution outlined. The enigma presented by the data will be highlighted and some of the ideas which have been suggested to resolve the puzzle will be reviewed. New methods of detection are needed and some of those planned will be outlined. There have been a number of recent reviews of the subject [1-5]. 2. Detection methods for ultra high energy cosmic rays Ultra High Energy Cosmic Rays (UHECR) are detected through the cascades of particles which are produced when * e-mail: [email protected] © Physica Scripta 2000
a high energy photon, proton or heavier nucleus interacts in the earth's atmosphere. The cascades are know as extensive air showers, the phenomenon having been discovered by Pierre Augur in 1938. His work established that the showers he was able to detect were initiated by entities (he thought that they were photons) of energy 10 15 eV and this changed the upper energy limit of energetic particles by nearly 106. The showers provide a detection route because the atmosphere acts both as a particle amplifier and a particle spreader. A primary of 1019eV will have about 10 10 particles in the resulting cascade when it reaches ground level while Coulomb scattering of the shower particles (mainly electrons) and the transverse momentum associated with the hadronic interactions spread the particles over a wide area. A shower from a 1019eV primary has a footprint on the ground of over 10 km 2 . The generic method of detection of air showers is a development of the technique used by Auger and his colleagues in their discovery work. The procedure is to spread an number of particle detectors (scintillators and water Cerenkov detectors are the current devices of choice) in a regular array on the ground. To study showers produced by primaries of 1019eV and above the spacing between the detectors can be about 1 km. The direction of the events can be measured to better than 3° by determining the relative time of arrival of the shower disk at the detectors. The particle density pattern can be used to infer the number of particles which reach the ground. To determine the primary energy it is necessary to use the predictions of a calculation which makes assumptions about the development of the shower in the atmosphere using extrapolations of the properties of interactions studied at accelerators. It is, of course, uncertain as to what the characteristics of hadronic interactions are at energies well above accelerator energies. The need to rely on extrapolations of known particle physics to estimate the primary energy is a major drawback of the generic method and has raised doubts about the accuracy of the energy estimates which derive from such studies. However the shower particles also excite the 2 + band of nitrogen and the resulting fluorescence is in the 350 - 450 nm band. Although only about 4.5 photons per metre of electron track are radiated, the light is detectable from showers above about 3 x 10 7 eV with arrays of photomultipliers on clear, moonless nights. This allows a calorimetric measurement of the shower energy in a manner similar to a technique familiar from particle physics. The resulting cascade shape gives the particle number as a function of depth and the energy of the primary can be obtained rather directly from the area under the curve. The angular accuracy of the fluorescence technique is very similar to that achieved with arrays of particle detectors. The fluorescence technique also allows the Physica Scripta T85
184
A. A. Watson
depth of maximum of the shower, Xmax to be determined to within ±50 gcm~ 2 . The particle array technique has been realised on scales > 10 km 2 at 5 sites world wide, Volcano Ranch (USA), Haverah Park (UK), Narribri (Australia), Yakutsk (Russia) and AG AS A (Japan). For a variety of reasons there are problems with data from the Narribri and Yakutsk experiments at the very highest energies (see [1,6] for a discussion) and in what follows data from the other three arrays will be emphasised. The fluorescence technique has so far been implemented only in the Dugway desert by the Fly's Eye group from the University of Utah: this has provided a key set of data. 3. The energy spectrum of high energy cosmic rays The arrays at Volcano Ranch and Haverah Park were set up with the primary motivation of measuring the arrival direction distribution of cosmic rays above 10 18 eV. In the late 1950s it was expected that the distribution would be anisotropic, possibly showing features associated with the magnetic field structure of our galaxy. No such anisotropics have so far been detected but in 1966 a prediction, following upon the discovery of the 2.7 K background radiation, led to a change of focus and to intense interest in measuring the spectral shape above 10 19 eV. It was recognised by Greisen [7] and independently by Zatsepin and Kuzmin [8] that cosmic rays of high enough energy would suffer serious energy losses as they passed through the microwave background radiation field. The interaction process is the excitation of the A + resonance if the highest energy cosmic rays are protons: the energies of interest are above 4 x l 0 1 9 e V . The protons lose energy rapidly and at 4 x 1019 eV 50% of the particles which reach the earth will have travelled less that 150 Mpc: at 10 20 eV the 50% distance is only 19 Mpc [5]. This effect has become known as the GZK effect. Heavier nuclei will also lose energy rapidly by photo-disintegration (the diffuse infra-red background is also important) and photons interacting with radio photons lose energy rapidly through pair production. The results from Haverah Park, Yakutsk, AGASA and Fly's Eye show that the cosmic ray energy spectrum steepens at around 2 x 1018 eV and then flattens above about 1019 eV. These spectral details were first claimed by the Haverah Park group [9] and were confirmed by later experiments with superior statistics. A very widely held interpretation of this result (e.g. [60]) is that galactic cosmic ray accelerators are no longer efficient above about 2 x 1018eV and that an extragalactic component begins to dominate at the highest energies. The origin of the extragalactic component remains unknown. Evidence for trans-GZK particles has been accumulating for many years. The first detection of a particle with energy above 10 20 eV was at the pioneering Volcano Ranch array [10] before the GZK prediction was made. Subsequent re-assessment of the energy has shown the original estimate to have been remarkably accurate. At Haverah Park 4 events above 1020 eV were reported from the exposure between 1967 and 1987 [11]. The deduction of the primary energy from the density pattern observed with ground arrays relies on the Monte Carlo calculations describing the development of the showers. While ground parameters were Physica Scripta T85
measured that have low susceptibility to uncertainties about the underlying particle physics, the astrophysical significance of the claimed energies was such that there was significant debate about their accuracy. In particular the first Fly's Eye spectrum [12], measured with a single eye, suggested evidence of a cut-off. However in 1993 the Fly's Eye group, using binocular data, reported a spectrum that agreed well with that measured at Haverah Park and, in addition, the group described the detection of an event with energy 3 x 1020 eV [13]. The calorimetric nature of the fluorescence method and the agreement with earlier measurements between 3 x 1017 and 3 x 1019 eV, where statistics are more numerous, removed many of the doubts about earlier claims for particles beyond the cut-off. In 1994 the AGASA group in Japan reported an event of 2 x 10 20 eV [14] and this group has recently published the results from a 7 year exposure with their 100 km 2 array. Six events with energies above 10 20 eV, well above the trans-GZK cut-off, are described [15]. Their data are reproduced in Fig. 1 where they are also compared with the final Haverah Park and Fly's Eye measurements. The agreement is extremely impressive bearing in mind the different techniques used and demonstrates that the systematic errors are understood at the 15% level. In Table I the exposures at 10 20 eV for the various instruments and the number of events recorded above that energy are shown. It is clear that within the very low statistics
ti
-
®
AGASA E'0.9
I
•
Akenolkm2
—
A
Haverah Park
7
-a u
< h
—
.
I
17
I
'-
9K Stereo Flys Eye
I
I
I
I
I
17.5
I
I
I
I
18
I
I
I
I
I
I
18.5
'
I
'
I
19
I
'
I
I
I
I
19.5
V I
I
I
I
20
I
I
I
I
20.5
21
log(ENERGY in eV) Fig. 1. A comparison of the spectrum reported1 by the AGASA group [15] with final spectra measured by the Haverah Park [11] and Fly's Eye [13]. I am grateful to Professor M. Nagano for providing me with this figure. © Physica Scripta 2000
Ultra High Energy Cosmic Rays - an Enigma Table I. Exposures of UHECR km 2 sr year > 1020 eV Volcano Ranch (1962) Haverah Park (1987) Fly's Eye (stereo) Fly's Eye (mono) (1993) AGASA (1998)
detectors.
N(> 1020 eV) Expected
Prob
~60
1
0.4
0.27
270
4
1.8
0.07
167
0
1.1
0.33
-600
1
4.1
0.07
667
6
4.5
0.13
12 events > 1020 eV Rate «* 1 km"2 century_1 sr 1 .
the rates from the different experiments are quite consistent. In addition the Yakutsk group have reported an event of 1.2 x 1020 eV [16]. The exposure for this array is not known with certainty as there are doubts about its efficiency for detecting high energy events [6,17]. We may conclude that the shape of the spectrum is now reasonably well known to about 5 x 1019 eV (Fig. 1) and that about 12 events have been detected above 1020 eV (Table I). Even if the energies of these events have been over-estimated by 20% it appears rather certain that the GZK cut-off is not observed. The rate of events above 10 20 eV is about 1 per km 2 per sr per century but this figure is uncertain to ±30%. 4. The arrival direction distribution The existence of events beyond the GZK cut-off implies that the sources of the events must be nearby. If a source could accelerate a proton to 1021 eV then, because of the loss processes associated with photo-pion production, the energy will fall to 3 x 1020 eV at 30 Mpc from it. For an iron nucleus the energy after travelling the same path would be even lower and for a photon lower still. Little is know about the magnetic field over such distances but, following Kronberg [18], it is often supposed that it has a mean strength of 10~9 G and is structured in cells of ~ 1 Mpc within which the field orientation changes. Such a configuration would deflect a proton of 1020 eV about 3° from a straight line over 30 Mpc. The celestial co-ordinates of the UHECR can be determined to better than 3° and one might therefore expect correlations between the arrival direction of the most energetic events with possible sources such as active radio galaxies. However one should look only at those objects which appear exceptional as there are many catalogues and there is always the possibility of finding some correlations from excessive searching when none really exists. One of the most careful pieces of work in this regard is that of Elbert and Sommers [19] who made a systematic but unsuccessful search for possible objects which might be the source of the 3 x 1020 eV event recorded by Fly's Eye. They argue that the objects of interest should be strong radio galaxies in which there would be extensive magnetic fields along with intense fluxes of high energy particles. They used the catalogue of strong radio sources (radio power 1041 ergs s ') compiled by Burbidge and Crowne [20] Within z < 0.125 (100 Mpc for H0 = 75 km s"1 Mpc" 1 ) © Physica Scripta 2000
185
there are 24 candidate objects, including well-known powerful systems such as For A, M82, M84, M87 and Cen A, but none lies close to the direction of the highest energy event. A more general search, using only data from the ground arrays, has been reported recently by Hillas [5]. His result is shown in Fig. 2. The restriction to ground array material, for which the exposure in right ascension is essentially uniform, allows a radial scale to be chosen so that an area in any part of the plot has the same exposure as an equal area in any other part of the plot. Events with energies above 1020 eV are shown with larger symbols. No correlations are found with galactic or extra-galactic features: the distribution of events is isotropic. The comprehensive plot shown in Fig. 2 has been preceded by other studies with smaller data sets. Using a data set dominated by Haverah Park events, Stanev et al. [21] claimed that cosmic rays above 4 x 1019 eV showed a correlation with the direction of the Super Galactic Plane: the level of significance was 2.5 — 2.8 a. Later studies with AGASA data [22] and with Fly's Eye data [23] did not support this claim. Very recently the AGASA group [24] have released details of 581 events above 10 19 eV recorded by them. Of these 47 are above 4 x 1019 eV and 7 are above 1020 eV. Within this self-consistent data set they find some evidence of clustering on an angular scale of 2.5° there are three doublets and one triplet, the chance occurrence of which is calculated as less than 1%. If such clusters are established in very much larger data sets they will have profound implications for our ideas on cosmic ray origin. It is interesting to note that one of the doublets becomes a triplet when a Haverah Park event of 10 20 eV is added: both triplets lie close to the Super Galactic Plane.
Fig. 2. The arrival directions of events with energy > 4 x 1019 eV The events are from the ground arrays for which the exposure in right ascension is uniform. The radial scale has been chosen so that equal areas in different parts of the plot correspond to equal exposures. The larger circles mark the directions of the events with energy above 1020eV. The super galactic plane is represented by the heavy dashed line running nearly vertically down the diagram. The solid line is the galactic plane: 'A' marks the galactic anticentre direction. The heavily stippled areas indicate regions where there are concentrations of radio galaxies and quasars [5]. Physica Scripta T85
186
A. A. Watson
5. Mass composition of UHECR Interpretation of the data on UHECR is hampered by lack of knowledge of the mass of the incoming particles. There are data from several experiments (AGASA, Fly's Eye and Haverah Park) which might be interpreted as showing a change from a dominantly iron beam near 3 x 10 I7 eV to a dominantly proton beam at 10 19 eV but there is still debate (see [25] for a recent discussion). The data are too limited to allow any reliable statement at energies beyond this. It is unlikely that the majority of the 12 events so far detected above 10 20 eV have photon parents as the showers seem to have normal numbers of muons (the tracers of hadronic primaries) [14] and the cascade profile of the most energetic fluorescence event is inconsistent with a photon primary [26]. It is also unlikely that the majority of events are created by neutrinos as the distribution of zenith angles would be different from what is observed. Indeed events of 1020eV look like events of 10 19 eV but ten times larger. However whether the particles are protons, iron nuclei or photons does not alter the problem posed by the existence of events above 10 20 eV with arrival directions that are distributed isotropically: this is the enigma of ultra high energy cosmic rays.
6. Theoretical interpretations The UHECR enigma is attracting significant theoretical attention and papers describing possible ways round the difficulty have appeared at the rate of 1 per week over the last 15 months. Many of the methods propose some form of electromagnetic process while others invoke ideas which demand new physics. 6.1. Electromagnetic
processes
It is important to recognise that the energy in a source capable of accelerating particles to 1020 eV and beyond must be extremely large as very general arguments demonstrate. A simple analysis [27,28], in which the size of the acceleration region is assumed to be comparable to the Larmor radius of the particle and the magnetic field is sufficiently weak to constrain synchrotron losses, shows that the energy in the magnetic field of the source grows as y5 where y is the Lorentz factor of the particle. For 10 20 eV the energy in the magnetic field must be » 10 57 ergs with B <0.1 Gauss. Such sources are likely to be strong radio emitters with radio power S> ,1041 ergs s"1, unless hadrons are being accelerated and electrons are not. It is worth emphasising that the present upper limit of 3 x 1020 eV is very probably set by the exposure rather than by any natural limit: the magnetic energy in the source of such a particle must be » 2 x 1059 ergs. The above analysis does not specify what acceleration process might be involved. Currently it is widely believed that cosmic rays with energies up to about 1015 eV are energised by diffusive shock acceleration with supernovae explosions identified as the most likely sites. At higher energies it is argued by some (e.g. [29]) that the same process continues but with the particles being accelerated by interaction with multiple supernova remnants as they move through the interstellar medium. This extended acceleration is supposed Physica Scripta T85
to take particles to energies close to 10 18 eV with the higher charge nuclei reaching the higher energies and becoming more dominant in the primary beam. This idea is consistent with the limited evidence on mass composition and with the steepening of the energy spectrum near 1018 eV. However there is, as yet, no direct evidence of acceleration of protons by supernova remnants at any energy. The diffusive shock acceleration process has been extensively studied since its conception in the late 1970s. In an accessible review Drury [30] has shown that the maximum energy attainable is given by E — kZeBRfic, where B is the magnetic field in the region of the shock, R is the size of the shock region and A: is a constant less than 1. The same result has been obtained by several authors including Hillas [31] who has used a simple but elegant plot of B vs. R to show that very few objects satisfy the conditions needed to achieve the maximum energy. The limit has been reconfirmed by Cesar sky [32] and by Blandford [33]. However some authors (e.g. [34]) claim that the diffusive shock acceleration process can be modified to give much higher energies than suggested by Drury's result and that radio galaxy lobes in particular are the acceleration sites. It is difficult to see how an energy of 3 x 1020 eV can be accounted for if the size of the shock region is 10 kpc and the magnetic field is 10 uG as even the optimum estimate of the energy is lower by a factor of 3 than this observational upper limit. Additionally Biermann and colleagues have speculated [35] that an energy spectrum and arrival direction distribution consistent with the data can be obtained if the magnetic field in the local Supercluster is 10~7 G. Thus two pieces of astrophysical understanding would need to be revised to explain the events: the parameters in the lobes of radio galaxies and the strength of the magnetic field in the local Supercluster. It should be noted that Elbert and Sommers [19] also proposed an unexpectedly strong extra-galactic magnetic field so that objects such as Cen A might be considered as candidate sources for the most energetic Fly's Eye event. There have been proposals on both theoretical and observational grounds that the highest energy cosmic rays are created when pairs of galaxies interact. Cesarsky [32] and Cesarsky and Ptushkin [36] were the first to examine this idea. They propose that the converging flows contain shock fronts which can accelerate the particles and claim that energies of ~ 10 20 eV can be reached if the primary nuclei are iron. They identify the Antennae system (NGC 4038/39) as a possible UHECR source. Observationally Al-Dargazelli et al. [37] have noted the association of a group of cosmic rays with energies above 10 19 eV with the galaxy pair VV338 while Takeda et al. [24] have tentatively identified an interacting galaxy Mrk 40 at z = 0.02 with a triplet of high energy events, the lowest energy of which is 5.35 x 1019 eV. It somewhat discouraging to find two colliding systems so identified: it lacks conviction. Furthermore Jones [38] has discussed the possibility of such colliding systems being sources and concluded that 10 17 eV is the maximum energy which is reasonably attainable. Amongst the most violent shocks in the Universe are the accretion shocks formed during the gravitational collapses which may lead to the formation of clusters of galaxies. These have been studied as a UHECR source by Kang, © Physica Scripta 2000
Ultra High Energy Cosmic Rays - an Enigma Ryu and Jones [39] but the limiting energy for protons is claimed to be only 6 x 1019 eV even if the field is as strong as 1 uG. They also claim that cosmic rays of around 1019eV could be produced in the Virgo cluster in such shocks. A further source of UHECR which has been considered is GRBs [40]. Here shocks with high Lorentz factors may give rise to conditions conducive to acceleration to the energies observed and it is argued that the energetics are such that attenuation by the microwave background radiation is not a problem. However the observational constraints on what is possible in GRBs may tighten in a few years and limit the energies reached in a way that we cannot envisage. In summary it seems difficult to find places where electromagnetic acceleration processes can raise particles to the energies observed. It should be stressed, however, that if the primaries were heavy nuclei there is some alleviation of the problem but much greater attention then has to be given to the photon environment in the acceleration region and to the possibility of fragmentation while they are being transported to earth: heavy nuclei of this energy are very fragile against photodisintegration. 6.2. Non-electromagnetic processes A number of proposals have been made which dispense with the need for electromagnetic acceleration. In general attention has been focused on the very highest energy events (> 10 20 eV) but in my view any mechanism that explains these events must also account for the events above about 1018 eV, where the galactic component probably disappears: the spectrum above this energy is too smooth to admit of two or more radically different components. Also the proposals must produce particles at the top of the atmosphere that can generate showers of the type we see: the characteristics of events about 10 20 eV are strikingly similar to those produced by primaries one or two decades lower in energy. One approach has been to invent mechanisms to evade the energy losses in the 2.7 K radiation field. For example Kolb, Farrar and collaborators [41] have speculated that a stable, massive, supersymmetric hadron, the S° (a combination of uds quarks and a gluino), may be responsible for producing the largest air showers. Such a particle, and similar massive particles dubbed 'uhecrons', might have a lower cross-section for the production of resonance particles at the energies observed. One positive feature of this proposal is that the particles would produce air showers rather similar in type to what is observed. As the S° is neutral the events would be expected to come from identifiable objects and Farrar and Biermann [42] have claimed an association with radio-loud QSOs for 5 of the most energetic events. Their statistical analysis has recently been challenged by Hoffman [43] and the idea is now capable of an independent test with the precise directions of the new AGASA events mentioned above [24]. Such an identification would stretch even further our understanding of particle acceleration processes as uhecrons of 3 x 1020eV would presumably be produced in conventional hadronic interactions that would require parent particles of even greater energy. Accelerator limits to the production cross-section of the S° imply an extremely severe energetics problem at the source, bearing in mind the y5 factor in the estimate of the magnetic field energy. © Physica Scripta 2000
187
Thus I find this hypothesis rather improbable on the grounds of energetics alone. Similarly vortons [44,45] represent a solution which is limited to the very highest energies. A similar idea, which I believe has the same generic difficulty with energetics, is the suggestion that ultra high energy neutrinos (10 22 eV) created in very distant sources produce cascades of photons, protons and electrons when they interact with background neutrinos of mass a few eV [46]. In addition to the production problem the predicted proton spectrum does not fit the experimental data down to 10 18 eV so that a variety of sources would be required to explain the observations above this energy. Kephart and Weiler [47] have revived an old idea [48] and proposed that monopoles of mass < 10 10 GeV might be the source of the high energy events. The monopoles would be accelerated in galactic magnetic fields and are not so numerous as to violate the Parker bound. However it is hard to see why an anisotropy associated with the galactic plane would not be expected and such monopoles have too low a Lorentz factor to produce the kind of showers observed. The later difficulty has recently been addressed by Huguet and Peter [49] who suggest that the monopoles have an internal structure which allows them to generate air showers in the conventional manner. Such entities would also be produced in sources and would not be in conflict with the observation of clusters of events from the same point in the sky. There is also speculation that UHECR arise from the decay of superheavy relic particles. In this picture cold dark matter is supposed to contain a small admixture of long lived superheavy particles with a mass > 1012 GeV and a lifetime greater than the age of the Universe [50]. It is argued that such particles can be produced during reheating following inflation or through the decay of hybrid topological defects such as monopoles connected by strings. It is hard to judge how speculative these ideas are but the decay cascade from a particular candidate [51] has been studied in some detail [52]. It is possible to produce the observed spectrum of cosmic rays, at least in the region above about 4 x 1019 eV, from the decay cascade and in addition the isotropy of the most energetic particles could be understood - within the limited statistics currently available - in terms of a galactic halo distribution of superheavy relic particles ([53], see also [5]). A further feature of the decay cascade is that an accompanying flux of photons and neutrinos is predicted which may be detectable with a large enough installation. The anisotropy question has been examined in some detail by Benson et al. [54] Berezinsky and Mikhailov [55] and Medina-Tanco and Watson [56]. The latter authors have made specific predictions for the anisotropy which would be seen by a Southern Hemisphere air shower array. The observation of the predicted anisotropy, plus the identification of appropriate numbers of neutrinos and photons, would be suggestive of a super heavy relic origin. However the existence of multiplets would not be expected in such a picture. Other ideas have been proposed including the suggestion [57] that there may be a departure from strict Lorentz invariance at the energies in question. Any of the classes of explanation outlined above would require a major reappraisal of our understanding of astrophysics, particle physics or cosmology. However without additional data we cannot hope to solve the puzzle Physica Scripta T85
188
A. A. Watson
set by the existence of UHECR and it is to provide these additional data that new projects have been planned.
7. New projects 7.1. The Pierre Auger Observatory The Pierre Auger Observatory has been conceived to measure the properties of the highest energy cosmic rays with unprecedented statistical precision. It will eventually consist of two instruments, constructed in the Northern and Southern Hemispheres, each covering an area of 3000 km 2 . The project will commence in late 1999 with the construction of an engineering array at Pampa Amarilla in Mendoza Province, Argentina. Subsequently work will start on the northern instrument in Millard County, Utah, USA. The Observatory has been designed by a consortium drawn from 18 nations and makes extensive use of the experience with previous ground arrays and with the Fly's Eye fluorescence detector. As well as providing an unsurpassed number of events, the Auger Observatory is designed to measure the energy and arrival direction of each event very accurately and to separate neutrinos and photons from hadronic primaries. Separation of hadronic primaries will be harder but there is a reasonable expectation of obtaining the fraction of protons and the fraction of iron nuclei as a function of energy. The design for the Observatory, which is the product of a six month planning operation supported principally by the Fermi National Accelerator Laboratory and UNESCO, calls for a hybrid detector system with 1600 particle detector elements and three or four fluorescence detectors at each of the sites. The principles of operation of a surface array and of a fluorescence detector separately were outlined in Section 2. The particle detectors will be 10 m 2 x 1.2 m deep water-Cerenkov tanks arranged on a 1.5 km hexagonal grid. These detectors are similar to those used at the Haverah Park array and have been selected because the water acts as a very effective absorber of the multitude of low energy electrons and photons found at distances of about 1 km from the shower axis. Both Monte Carlo calculations and experiments have shown that the mean energy of photons and electrons is about lOMeV. By contrast the muons, which are much less numerous, carry, on average, about 1 GeV of energy. At the Southern site fluorescence detectors will be set up at four locations, one near the centre of the particle array with the others on small promontories at the array edge: the site is close to the town of Malargue. During clear moonless nights signals will be recorded in both the fluorescence detectors and the particle detectors while for roughly 90% of the time only particle detector data will be available. It is much too costly to connect detectors by cables or optical fibres over an area of 3000 km 2 . Instead each detector will operate in a "stand-alone" mode with trigger data being sent to the central station using a Wireless LAN radio link operating at 915 MHz. The power for the surface detectors will be obtained from solar cells while data at all detectors will be time-stamped using GPS receivers. The GPS timestamp is sufficiently precise to allow accurate reconstruction of the direction of each incoming event. A major advantage of the use of detectors of two types - the hybrid array is that independent measurements of some parameters will Physica Scripta T85
be made for about 10% of the showers. For example the energy calibration of a surface array depends on the details of the interaction model used to convert the ground level measurements to the energy input at the top of the atmosphere. By contrast the fluorescence detector has direct energy calibration although correction must be made for the component of the detected light which is scattered Cerenkov light and for the transparency of the atmosphere, which needs to be monitored on a very regular basis. This later feature means that the aperture of the fluorescence detector grows as the energy increases in contrast to that of the surface detector which (see Fig. 3) is constant above an energy dictated only by the adopted spacing of the detectors. Thus for 1.5 km spacing the surface detector aperture remains constant at energies above 10 19 eV. The fluorescence detector can be used to measure the depth of shower maximum directly from the shape of the cascade curve. This is an important parameter but its interpretation does depend on the shower model used to describe the data. The ground array can measure the risetime of the signal at each detector, the lateral distribution of the signal amplitudes and the curvature of the shower front, each of which is related to the depth of shower maximum. The combination of the ground array and the fluorescence detector will also allow the core position of the shower to be found to within about 50 m, more precisely than with either detector alone. It seems probable that this combination of shower parameters, very accurately measured in 10% of events, will help decide which model is appropriate to describe the data and hence allow the primary mass to be extracted. In addition the water-tanks are equipped with Flash ADCs which enable the temporal form of the signal to be measured. It has been shown, both with Monte Carlo calculations and with a prototype 10 m 2 tank operated at the AGASA array, that the time-spread is sufficient to allow the large signals from through-going muons to be detected against the more numerous but smaller signals produced
I
t
P
S El £ 3.5 H 3 E ' O, 3 t,
J___L_l_iJ_LLLL—_J—L-J-i-LlULJm f I 1 I 1
I 1 I 1
I I I I I I I I h-4--l-h-l-^P^ 1 I I I I I I ^ I *__, r - T - | - r r r |
I >, I ,
I I I I I I I 1 t 1 - -+ - h f + + •*--! 17,5 * I I I I I I I q S , _ 1 . _ r r T r r J ,5 S
Energy feV)
Fig. 3. The upper diagram shows the triggering efficiency (upper line) as a function of energy while the lower line shows the way in which the number of triggered detectors increases with energy. The upper and lower lines in the lower figure show, respectively, the change of energy resolution and angular resolution with energy. (From the Auger Project Design Document). © Physica Scripta 2000
Ultra High Energy Cosmic Rays - an Enigma by the lower energy electromagnetic components. This provides a powerful method of identifying the fraction of muons in the shower, an important number which will be used to aid the derivation of the mass composition of the primary beam. This augurs particularly well for the identification of photon- initiated primaries. Neutrino events are expected to be detected most readily at angles greater than 75° from the vertical. The showers will be similar to those produced by hadrons in the near vertical direction and will contrast strongly with the muon dominated events which will characterise the remnants of showers produced by hadrons at such large angles. Detailed simulations of the performance of the ground array for energy and direction measurement have been carried out. The results are shown in Fig. 3. At 4 x 1019eV the energy resolution, with the ground array of particle detectors alone, will be ~ 10% and the angular resolution will be ~ 1.5° : on average about 11 detectors will be struck. The energy resolution and angular accuracy improves as the energy increases. The first stage of the Auger project will produce unique data as the Southern sky has not been studied with cosmic rays of these energies. There are some reasons to suppose that the spectrum and arrival direction distributions may not be the same as measured from the Northern hemisphere. These include the different distribution of luminous matter, the diffuse infrared photon flux (which has an influence on propagation of heavy nuclei) and the distribution of matter in the galactic halo. The Southern Hemisphere Auger detector thus offers the opportunity of unique science which, as explained above, could have a major impact on cosmology and/or particle physics. The engineering array is expected to be completed in late 2000. Although construction of the 3000 km 2 detector will take a further 4 years, exciting new data will be available as soon as 100 km 2 is operational with one fluorescence detector. 7.2. Other detectors: AGASA, Hi-Res and Airwatch/OWL At present the only experiment taking data above 10 19 eV is the AGASA instrument in Japan which has an aperture of 170 km 2 sr. From mid-1999 a successor to the Fly's Eye instrument, known as Hi-Res [58] started to take data in the Northern Hemisphere with a time averaged aperture of 340 km 2 sr at 1019 eV and 1000 km 2 sr at 10 20 eV. This is a stereo system which is expected to be able to measure the depth of shower maximum to within 30gem - 2 on an event-by-event basis. This precision is smaller than the expected fluctuations in depth of maximum for proton or Fe initiated showers. By contrast a single Auger observatory will have an aperture of 7000 km 2 sr above 1019 eV and will thus be 20 times as powerful as Hi-Res at the lower energy and 7 times more effective at the higher energy. The rate of events with fluorescence and ground array information from Auger will be comparable to the rate from Hi-Res. Achieving an exposure greater than that promised by the Auger Observatory is a formidable challenge. A promising line is the development of an idea of Linsley [59]. The concept is to observe fluorescence light produced by showers from space with satellite borne equipment. It is proposed to monitor ~ 10 5 km 2 sr (after allowing for the estimated 8% on-time). Preliminary design studies have been carried out in Italy [61] and the USA [62] and the groups interested © Physica Scripta 2000
189
have combined to plan an project known as Airwatch/OWL. Two satellites might be used to observe the fluorescence emission. The project requires considerable technological development but may be the only way to push to energies beyond whatever limit is found with the Auger instruments.
Acknowledgements I am gratefUl to the organising committee of the Nobel Symposium for inviting me to the meeting and for their generous hospitality and support.
References 1. Watson, A. A., Nucl. Phys. B (Proc. Suppl.) 22B, 116 (1990). 2. Watson, A. A., Nucl. Phys. B (Proc. Suppl.) 28B, 3 (1992). 3. Sokolsky, P., Sommers, P. and Dawson, B. R., Phys. Rep. 217, 225 (1995). 4. Yoshida, S. and Dai, H„ J. Phys. G 24, 905 (1998). 5. Hillas, A. M., Nature 395, 15 (1998). 6. Watson, A. A., Proc. Int. Symp. on Extremely High Energy Cosmic Rays: Astrophysics and Future Observatories (editor: M. Nagano: ICRR, Tokyo), 1996, p. 362. 7. Griesen, K., Phys. Rev. Lett. 16, 748 (1966). 8. Zatsepin, G. T. and Kuzmin, V. A., Sov. Phys. JETP Lett. 4, 78 (1966). 9. Cunningham, G. et al, Astrophys. J. 236, L71 (1980). 10. Linsley, J., Phys. Rev. Lett. 10, 146 (1963). 11. Lawrence, M. A., Reid, R. J. O. and Watson, A. A., J. Phys. G 17, 733 (1991). 12. Baltrusaitis, R. M. et al, Phys. Rev. Lett. 54, 1875 (1985). 13. Bird, D. J. et al, Astrophys. J. 424 491 (1994). 14. Hayashida, N. et al, Phys. Rev. Lett. 73, 3491 (1994). 15. Takeda, M. et al, Phys. Rev. Lett. 81, 1163 (1998). 16. Efimov, N. N. et al, Proc. ICCR International Symposium (Kofu): Astrophysical Aspects of the Most Energetic Cosmic rays, (World Scientific, 1990) (editors: M. Nagano and F. Takahara), p. 20. 17. Bower, A. J. et al, J. Phys. G 9, L53 (1982). 18. Kronberg, P., Rep. Prog. Phys. 57, 325 (1994). 19. Elbert, J. W. and Sommers, P., Astrophys. J. 441, 151 (1995). 20. Burbidge, G. and Crowne, A. H., Astrophys. J. Suppl. 40, 583 (1979). 21. Stanev, T. et al, Phys. Rev. Lett. 75, 3056 (1995). 22. Hayashida, N. et al, Phys. Rev. Lett. 77, 1000 (1996). 23. Bird, D. J. et al, Astrophys. J. 511, 739 (1999). 24. Takeda, M. et al, Astrophys. J. 522, 225 (1999). 25. Dawson, B. R., Meyhandan, R. and Simpson, K. M., Astroparticle Phys. 9, 331 (1998). 26. Halzen, F. et al, Astroparticle Phys. 3, 151 (1995). 27. Griesen, K., Proc. 19th ICRC (London) 2, 609 (1965). 28. Cavallo, G., Astron. Astrophys. 65, 415 (1978). 29. Ip, W. H. and Axford, W. I., AIP Conference Proc. 264 (Particle Acceleration in Cosmic Plasmas, 1991), p. 400. 30. Drury, L. O ' C , Contemporary Phys. 35, 232 (1994). 31. Hillas, A. M., Ann. Rev. Astron. Astrophys. 22, 425 (1984). 32. Cesarsky, C. J., Nucl. Phys. 288, 51 (1992). 33. Blandford, R. D., These proceedings (1999). 34. Biermann, P., J. Phys. G 23, 1 (1997). 35. Biermann, P. et al, astro-ph/9806283 (1998). 36. Cesarsky, C. J. and Ptushkin, V., Proc. 23rd ICRC (Calgary) 2,341 (1993). 37. Al-Dargazelli, S. S. et al, Proc. 25th ICRC (Durban) 4, 465 (1997). 38. Jones, F. C , AIP Conference Proc. 433 (Workshop on Observing Giant Air Showers from > 1020 eV particles from Space), 1998, p. 37. 39. Kang, H., Ryu, D. and Jones, T. W., Astrophys. J. 456, 422 (1996). 40. Waxman, E„ Phys. Rev. Lett. 75, 386 (1995). 41. Kolb, E. W., Farrar, G. H. et al, 1997, 1998, astro-ph/9707036 and hep-ph/9805288. 42. Farrar, G. H. and Biermann, P., Phys. Rev. Lett. 81, 3579 (1998). 43. Hoffman, C , astro-ph/9901026 (1999). 44. Bonazzola, S. and Peters, P., Astroparticle Phys. 7, 161 (1997). 45. Masperi, L. and Silva, G., Astroparticle Phys. 8, 173 (1998). Physica Scripta T85
190 46. 47. 48. 49. 50. 51. 52. 53. 54. 55.
A. A. Watson
Yoshida, S., Sigl, G. and Lee, S., hep-ph/9808324 (1998). Kephart, T. W. and Weiler, T. J., Astroparticle Phys. 4, 271 (1996). Porter, N . A., Nuovo Cimento 16, 958 (1960). Huguet, E. and Peter, P., submitted to Astroparticle Phys. (1999). Berezinsky, V., Kachelriess, M. and Vilenkin, A., Phys. Rev. Lett. 79, 4302 (1997). Benkali, K., Ellis, J. and Nanopoulos, D. V., hep-ph/9803333, Phys. Rev. D (in press) (1998). Birkel, M. and Sarkar, S„ Astroparticle Phys. 9, 297 (1998). Dubovsky, S. L. and Tinyakov, P. G., J.E.T.P. Lett. 68, 107 (1998). Benson, A., Smialkowski, A. and Wolfendale, A. W., Astroparticle Phys. 10, 313 (1998). Berezinsky, V. and Mikhailov, A. A., astro-ph/9810277, Phys. Lett. (submitted) (1998).
Physica Scripta T85
56 Medina-Tanco, G. A. and Watson, A. A., Astroparticle Phys. 12, 25 (1999). 57. Coleman, S. and Glashow, S. L., he-ph/9808446 (1998). 58. Sokolsky, P., AIP Conference Proceedings 433 (Workshop on Observing Giant Air Showers from > 1020 eV particles from Space), 1998, p. 65. 59. Linsley, J., 1979, USA Astronomy Survey Committee (Field Committee) Documents and Proc. 19th Int. Cos. Ray. Conf. (La Jolla) 3, 438 (1985). 60. Axford, W. I., Astrophys. J. Suppl. 90, 937 (1994). 61. Demarzo, C. N., AIP Conference Proc. 433 (Workshop on Observing Giant Air Showers from > 1020 eV particles from Space), 1998, p. 87. 62. Streitmatter, R. E., AIP Conference Proc. 433 (Workshop on Observing Giant Air Showers from > 1020 eV particles from Space), 1998; p. 95.
© Physica Scripta 2000
Physica Scripta.Vol. T85, 191-194, 2000
Acceleration of Ultra High Energy Cosmic Rays R. D. Blandford 130-33 Caltech Pasadena CA 91125 USA Received March 3, 1999; accepted August 2, 1999
extending from the proton rest mass, ~ 1 GeV to the "knee" at ~ 100 TeV -10 PeV. (10 GeV cosmic rays are about 10 m Some general features of cosmic ray acceleration are summarized along with apart and have an energy density comparable with that some inferences that can be drawn concerning the origin of the UHE of the microwave background. The spectrum steepens above component. The UHE luminosity density is found to be similar to that derived for GeV cosmic rays and its slope suggests a distinct origin. Reports of the knee: C / ( £ ) ~ 4 x 10" 18 (£/10 P e V ) " u J m" 3 . It then clustering on small angular scale, if confirmed, would rule out most proposed dips and flattens around the "ankle" (~ 1 — 10 EeV). source models. More generally, it is argued that the highest energy particles UHE cosmic rays - the toenail clippings of the universe 20 can only be accelerated in sites that can sustain an EMF £ :> 3 x 10 V and an associated power Lmiii :> £2/Z ~ 1039 W, where Z is the characteristic, - are observed up to 300 EeV and, with a little imagination, U(E) ~ 1.5 x 10"21CE/10 EeV)" 0 5 J m" 3 , comparable with electrical impedance, typically <100fi Shock acceleration, unipolar induction and magnetic flares are the three most potent, observed, accelerthe estimated, integrated background from y-ray bursts. ation mechanisms and radio jet termination shocks, •y-ray blast waves, dor(Despite the large uncertainty, and the fact that the number mant black holes in galactic nuclei and magnetars are the least density has fallen by ~ 10 orders of magnitude, we do implausible, "conventional" manifestations of these mechanisms that have measure the EeV spectrum better than the MeV spectrum, been invoked to explain the UHE cosmic rays. Each of these models presents of which we are, quite decently, ignorant.) problems and deciding between these and "exotic" origins for UHE cosmic rays, including those involving new particles or defects will require improved The ~ 1 GeV-100 TeV cosmic rays are of Galactic origin. statistical information on the energies, arrival times and directions, as should The ratio of Li, Be, B secondaries to C, N, O primaries be provided by the AUGER project. measures their range to be 1(E) ~ 100(£/10 GeV) - 0 6 kg m~2 (eg [6]). The cosmic ray luminosity of the Galaxy is then estimated as ~ MdU(E)c/l(E) ~ 2 x 10 33 (£yi GeV)" 0 1 W 1. Introduction where Mi is the gas mass of the disk. Scaling from the local I have been asked to summarize "conventional" schemes for galaxy luminosity density (per In is and assuming the acceleration of UHE cosmic rays, though any physical h = 0.6), we derive an average, cosmological, luminosity process capable of endowing a subatomic particle with density (per l n £ ) , < C > (E) ~ 4 x 10" 37 (^/10 GeV) -0 ' 1 W the kinetic energy of a well-hit baseball/cricketball can m~ 3 , for 10 GeV < E < 100 TeV. (For comparison, the hardly be considered conventional. This means that I shall stellar luminosity density is ~ 10~33 W m~ 3 .) leave others to review mechanisms that attribute the origin The UHE particles are almost surely extragalactic. As of these particles to topological defects, strings, monopole with y-ray bursts, there is no good evidence for disk, halo, decay, supersymmetric hadrons, cosmic necklaces, cryptons cluster or supercluster anisotropy (despite some tantalising and so on. Indeed, I suspect that the "hidden agenda" is hints in the past) [7]. Furthermore, magnetic confinement for me to fail at my appointed task, and, like my colleagues by the Galaxy is impossible - the Larmor radius r^(E) of on the MACHO experiment, to make the world safe for a 300 EeV cosmic ray in a uG field is ~ 300 kpc. If we assume elementary particle theorists. I shall not disappoint. that UHE cosmic rays are protons, (and assuming that they Many of the issues that I will cover have been recognised are not, only makes matters worse), then they have a short for some time and have been well-discussed in several excel- lifetime to photo-pion production on the microwave backlent reviews including Hillas [1] and Cronin [2] and the many ground [8,9]. The characteristic lifetime of a ~ 60— 300 EeV relevant contributions to the recent conference on this sub- cosmic ray is, very roughly, T(E) ~ 0.1 (is/300 EeV) -2 Gyr. ject [3], including, especially, the lively summary by the late This implies that the luminosity density increases with David Schramm. The conference proceedings edited by energy < £ > ( £ ) - U(E)/T(E)~ 10" 37 (£/300 EeV) 15 Chupp & Benz [4] is also relevant. Wm~ 3 . At the highest measured energy, the estimated cosmological luminosity density is not significantly different from that of 10 GeV cosmic rays. The change in slope in 2. The cosmic ray spectrum the source spectrum, above ~ 60 EeV, is a strong indication In order to give this topic some context, consider the com- that these UHE cosmic rays comprise a quite distinct plete cosmic ray spectrum (eg [5]). This extends over nearly component from their lower energy counterparts. In order to investigate this further, it is necessary to take twelve decades of energy from the proton rest mass, ~ 1 GeV, where their energy density is that of the microwave account of the fluctuations in energy loss. Taking the 17 background, to at least 300 EeV (= 50 J = 3 x 10" 8 m P i). We events reported by the AGASA collaboration above can consider the cosmic ray spectral energy density inferred 60 EeV, it is possible to derive a maximum likelihood estiat the solar system, U(E) = (4n/c)(dI/d\nE) ~ 5xl0~ 1 4 mate of the unnormalized energy density, uncorrected for -07 3 (is/10 G e V ) J m " (correcting for solar modulation) biases in detection efficiency. I find that if U(E) oc E"; Abstract
© Physica Scripta 2000
Physica Scripta T85
192
R. D. Blandford
E > .Emin = 60 EeV, then a = —1.2 ± 0.5.1 then calculate the probability that a particle of energy EQ has energy > E after time t, P(E, t\ Eo), following Aharonian and Cronin [10] and Bahcall and Waxman [11]. If we assume a power law for the luminosity density C(E) oc E&; E > Em{n = 60 EeV, then the logarithm of the likelihood for obtaining the observed events is
a£ln[(l - fi)£<^ JjEoE^2 J dt
dP 3\nE
(1)
Maximizing this function with respect to variation of /?, gives the estimate /? = 0.3 ± 0.2, and may even be consistent with a single, "top down" source with energy well above 300 EeV. A more sophisticated computation that takes into account the detection probabilities of the different events should be performed. There have been reports that UHE cosmic rays are significantly clustered on the sky. Specifically, in a sample of 47 events observed with AGASA, there are three pairs and one triple above 40 EeV with separations <2.5°, comparable with the positional errors [7]. (There are two more coincidences with events drawn from other samples.) There is no clear pattern for the associated particles to be ordered in energy and, in particular, one double has a 106 EeV particle arriving over 3 yr. after a 44 EeV particle. If these associations are real and typical, then there are three important implications. Firstly, as particles are likely to be deflected by intergalactic magnetic field through an angle S6 ~ (D£B)y2/rL(E) then they will be delayed by ~ D2lB/rL(E)2c oc E~2, where iB is the field correlation length and D ~ c7\150EeV) ~ 30 Mpc is the supposed source distance (cf [12]). Even Aesop would be challenged to explain how ~ 40 EeV cosmic rays precede ~ 100 EeV cosmic rays if they started at the same time and we would conclude that the source persists for several years, at least. This would rule out all particle/defect and y-ray burst models. Secondly, the small deflection angles at low energy limit the intergalactic field strength to B<20 (lB/\Mpc)~l/2 fT, far smaller than generally supposed, though probably not excludable by direct observation. Thirdly, the presence of three ~ 40 EeV cosmic rays associated with high energy cosmic rays of much shorter range, implies that the background of low energy cosmic rays not associated with high energy events must be larger than the incidence of clustered events by roughly the ratio of their typical lifetimes ~ 30, which more than accounts for the remainder of the low energy sample. This, in turn, implies that the high energy cosmic rays must come from a very few sources which are, consequently, quite energetic: E ~ lO^r/Syr) J, where T is their lifetime. (If the low energy cosmic rays are scattered through ~ 2.5°, then x > 105 yr and E ~ 3 x 1048 J.) However, this clustering hypothesis, which is necessarily a posteriori when expressed in detail, is only supported with modest confidence. (A simple, Monte Carlo simulation distributing 47 points at random on half the sky and looking for similar patterns is quite instructive.) Particle/defect/ burst explanations of UHE cosmic rays need not yet be rejected on these grounds. Physica Scripta T85
3. Cosmic ray acceleration The standard model of bulk cosmic ray production is first order Fermi acceleration at strong, super-Alfvenic, shocks associated with supernova remnants and, possibly, winds from hot stars {eg [13]). A typical relativistic proton will cross a shock, travelling with speed u, 0(c/u) times, gaining energy AE/E = 0(u/c) each traversal through scattering by hydromagnetic waves moving slowly with respect to the converging fluid flows on either side of the front. The net mean relative energy gain is 0(1), but the process is statistical and a kinetic calculation shows that the transmitted spectrum will be a power law in momentum, f(p) oc p-3r/(r-i)^ w h e r e rt ( = 4 for a strong shock), is the compression ratio. This mechanism can account, broadly, for the power (eg [14]), the slope (eg [6]) and the composition (eg [15]) of GeV cosmic rays. Shock acceleration is also, arguably, observed directly in SN1006 [16], as well as in the solar system (eg [17]). The maximum energy to which a particle can be accelerated at a shock front is dictated by the scattering mean free path, 1(E) ~ (B/SB)2LrL(E), where SB is the amplitude of resonant hydromagnetic waves with wavelength matched to the particle Larmor radius. The diffusion scale-length of cosmic rays ahead of the shock is ~ Ic/u and, assuming that this is limited by the size of the shock ~ R we arrive at the unsurprising result that the maximum energy achievable in shock acceleration, assuming SB < B and the presence of a large scale magnetic field, is Emax = e£ ~ euBR ~ ed<$>/&t, the product of the charge and the motional potential difference across the whole shock. Equivalently, we conclude that in order to accelerate a proton by this mechanism to an energy £e, the rate of dissipation of energy exceeds £„,;„ ~ £2/Z, where Z = \iQu — \XQE/B is the effective impedance of the accelerator in SI units. Imposing this condition for a supernova remnant in the interstellar medium leads to an estimate .Emaxsnr ~ 30 TeV (eg [6]), close to the knee. An additional source is needed between the knee and the ankle, where the source is generally supposed to be metagalactic. Larger shocks, especially those at Galactic wind termination shocks (eg [18]) and associated with gas flows around groups and clusters of galaxies have been invoked (eg [19]). These shocks are likely to be relatively weak and therefore to transmit steeper spectra, as observed. The major uncertainty is the strength of the magnetic field. If B ~ 30 pT at a galactic shock and ~ 10 pT at a cluster shock, then £ m a x ~ 10, 100 PeV respectively. Neither site is likely to accelerate the highest energy particles. An alternative accelerator is the unipolar inductor (eg [20]). The archetypical example is a pulsar - a spinning, magnetised, neutron star. The surface field will be quite complex but a certain quantity of magnetic flux $ can be regarded as "open" and tracable to large distances from the star, (well beyond the light cylinder). As the star is an excellent conductor, an EMF will be electromagnetically induced across these open field lines £ ~ £2$, where $ is the total, open magnetic flux. This EMF will cause currents to flow along the field and as the inertia of the plasma is likely to be insignificant the only appreciable impedance in the circuit is related to the electromagnetic impedance of free space Z ~ 0.3|ioc ~ 100 Q. The maximum energy © Physica Scripta 2000
Acceleration of Ultra High Energy Cosmic Rays to which a particle can be accelerated is Emax ~ e£ and the total rate at which energy is extracted from the spin of the pulsar is Lmin ~ £2/Z. Taking the Crab pulsar as an example, Emax ~ 30 PeV for protons and Z^n ~ 1031 W. As the stellar surface may well comprise iron, even the Crab pulsar has the capacity to accelerate ~EeV cosmic rays. However, it is not obvious that all of this potential difference will actually be made available for particle acceleration. In particular, this is unlikely to happen in the pulsar magnetosphere as a large electric field parallel to the magnetic field will be shorted out by electron-positron pairs, which are very easy to produce, and radiative drag is likely to be severe. A more reasonable site is the electromagnetic pulsar wind and the surrounding nebula where particles can gain energy as they undergo gradient drift between the pole and the equator [21]. Pulsars may well contribute to the spectrum of intermediate energy cosmic rays. A third, protoypical accelerator is a flare, for example one occuring on the solar surface or the Earth's magnetotail. Here magnetic instability leads to a catastrophic rearrangement, which must be accompanied by a large inductive EMF. Unless the instabilities are explosive, the effective impedance is again ~ \i0u, where u is a characteristic speed. Non-relativistic flares generally convert most of the dissipated magnetic energy into heat and are notoriously inefficient in accelerating high energy particles. Other acceleration mechanisms have been proposed and may contribute to the acceleration of the bulk of Galactic cosmic rays and relativistic electrons in non-thermal sources. These include a variety of second order processes and steady, magnetic reconnection. Many of them can be observed to operate within the solar system. However, they are thought to be too slow to be relevant to the acceleration of the highest energy cosmic rays.
4. Zevatrons Having argued that the three most potent, observed acclerators are shocks, unipolar inductors and flares, let us see how they can be modified to account for ~ ZeV cosmic rays. Firstly, note that, as u ~ c, mildly relativistic shocks minimise the power that has to be invoked to attain high energy. Specifically, we need a power > 1039 W to account for 300 EeV cosmic rays and this exceeds the bolometric luminosity of a powerful quasar. One of the few sites where such a large potential difference can be achieved is the termination shock of a powerful radio jet like that associated with Cygnus A (eg [22]). Stretching the numbers a little, we combine a field strength ~ 10 nT, with a speed ~ c and a transverse scale ~ 3 kpc which gives Em^ ~ 300 EeV. The problem with this model is that observed UHE cosmic rays are not positionally identified with the few known radio sources within D{E) ~ 30 Mpc that might be powerful enough to account for them (c/[23]). A more elaborate shock accelerator is the y-ray burst blast wave (eg [24]). Here, the shocks (assumed to be spherical) are ultrarelativistic with Lorentz factor f. The maximum energy, measured in the frame of the explosion, to which a proton can be accelerated in a dynamical timescale from an ultrarelativistic shock of radius R is Emm ~ eB'Rc, where Bl is the comoving field strength. The explosion power, © Physica Scripta 2000
193
adopting the most elementary of assumptions, is then Lmin ~ 4nr2(Emax/e)2/\iQC. Observed bursts have typical explosion powers estimated to be L exp ~ 1045 W, which can be consistent with 300 EeV proton acceleration as long as r < 300, which is just compatible with existing models. A serious physical constraint is the avoidance of radiative loss in this environment. An observational concern with this model is the improbability of having enough active bursts close to supply the highest energy particles roughly isotropically (cf [25]). The most relevant variant on unipolar induction is magnetic energy extraction from spinning, black holes, where the magnetic field is supported by external current, and the horizon is an imperfect conductor with resistance ~ 100Q (eg [26,27]). This impedance is matched to the electromagnetic load so that roughly half of the available spin energy ends up in the irreducible mass of the hole, the remainder being made available for particle acceleration. The total electromagnetic power needed to account for ~ 300 EeV acceleration is, once more, ~ 1039 W. A rapidly spinning, ~ 109 M© hole endowed with a field strength > 3 T or a 105 M G hole threaded by a > 3 x 104 T field suffices to accelerate 300 EeV particles. The major concern with this model is that the radiation background must be extremely low in order that catastrophic loss due to pion and pair production be avoided. Specifically, it is necessary that the microwave luminosity in an acceleration zone, of size R, be < 1034(i?/1014m) W, far smaller than the electromagnetic power. The best generalization of flare acceleration involves "magnetars" which are young, spinning neutron stars endowed with a ~ 10 — 100GT surface magnetic field as first postulated by Thompson and Duncan [28]. Now, the observation of 5 — 7 s period pulsations from three "soft gamma repeaters" effectively confirms their identification as old magnetars that have been decelerated by electromagnetic torque and which are now powered by magnetic energy which is released in a series of giant flares [29]. The inductive EMFs associated with an electromagnetic flare from a magnetar can be as high as AV ~ 3 x 1019 V, making them candidate UHE accelerators because the surface composition is likely to be Fe. However, the available reservoir of magnetic energy is only ~ 1040 J and the magnetar birthrate is no more than ~ 10 - 3 yr _ 1 in the Galaxy. This rules them out as an extragalactic source. Only if UHE cosmic rays have a Galactic origin, (and the large scale isotropy observations suggest quite strongly that do not), can there be enough power in magnetars to account for the UHE energy density.
5. Discussion I have argued, tentatively, that UHE cosmic rays are created in a new population of extragalactic sources with an average luminosity density that approaches that of Galactic cosmic rays. I have also described problems with each of the candidate "conventional" mechanisms for accelerating protons to these high energies. Quite different, general inferences have been drawn here, from the same data, by Waxman, and elsewhere by others. All of this underscores the need for better statistics which should be met by the AUGER proPhysica Scripta T85
194
R. D. Blandford
ject. Perhaps the most pressing need is to understand if particles of very different energy have a common origin. If true, this must rule out essentially all primordial particle/topological defect, neutron star, y-ray burst explanations, leaving only massive black holes and radio source models among the possibilities discussed above. In this case, it will be possible to seek identifications, especially at the highest energies, where the positions will be most accurate and the delays due to magnetic scattering the smallest. If, alternatively, clustering and its implications are not substantiated, then the next best clues will probably come from composition studies and detailing the large scale distribution on the sky. The most exciting outcome of all of this is that we are dealing with a new particle or defect with energy well out of the range of terrestrial accelerators. (For example, if there is a particle of energy Ex which decays with half life xx into N protons, then the cosmological energy density of these particles must be Qx ~ 3 x lO^AT'CExVlAeV)-1 (HQTX)~1.) Whatever happens, in a subject where the dullest and most conventional theories involve massive, spinning, black holes, ultrarelativistic blast waves and 100 GT fields threading nuclear matter, the future is guaranteed to be interesting.
Acknowledgements I am indebted to John Bahcall, Jim Cronin, Michael Hillas, Martin Rees, Alan Watson and Eli Waxman for stimulating discussions and the editors for their forbearance. I also gratefuly acknowledge the hospitality of the Insitute for Advanced Study (through the Sloan Foundation) and the Institute of Astronomy (through the Beverly and Raymond Sackler Foundation) as well as NASA grant 5-2837.
Physica Scripta T85
References 1. Hillas, M., Ann. Rev. Astron. Astrophys. 22, 245 (1984). 2. Cronin, J. "Unsolved Problems in Astrophysics" (Eds. J. Bahcall and J. Ostriker) (Princeton: Princeton University Press) p. 325. 3. Krizmanic, J. F., Ormes, J. F. and Streitmatter, R. E. (Eds.) (New York: AIP, 1998). 4. Chupp, E. L. and Benz, A. O. (Eds.) Astrophys. J. Suppl. 90, 511 (1994). 5. Berezinski, V. S. et al., "Astrophysics of Cosmic Rays" (Amsterdam North Holland, 1990). 6. Axford, W. I., Astrophys. J. Suppl. 90, 937 (1994). 7. Takeda, M. et al., astro-ph/9902239 (1999). 8. Greisen, K., Phys. Rev. Lett. 16, 748 (1996). 9. Zatsepin, G. T. and Kuz'min, V. A., J. Exp. Theor. Phys. 4, L78 (1966). 10. Aharonian, F. and Cronin, J., Phys. Rev. D 50, 1892 (1994). 11. Bahcall, J. N. and Waxman, E., 1999 (preprint). 12. Miralda-Escude, J. and Waxman, E„ Astrophys. J. 462, L59 (1996). 13. Blandford, R. D. and Eichler, D., Phys. Rep. 154, 1 (1987). 14. Malkov, M. A., Astrophys. J. 511, L53 (1999). 15. Ellison, D., Drury, L. O'C. and Meyer, J.-P., Astrophys. J. 487, 197 (1997). 16. Koyama, K. et al., Nature 378, 255 (1995). 17. Erdbs, G. and Balogh, A., Astrophys. J. Suppl. 90, 553 (1994). 18. Jokipii, J. R. and Morrill, G., Astrophys. J. 312, 170 (1987). 19. Norman, C. A., Melrose, D. B. and Achterberg, A., Astrophys. J. 454, 60 (1995). 20. Goldreich, P. and Julian, W. H., Astrophys. J. 157, 869 (1967). 21. Bell, A. R., Mon. Notes R. Astron. Soc. 257, 493 (1992). 22. Cavallo, G., Astron. Astrophys. 65, 415 (1978). 23. Farrar, G. R. and Biermann, P. L., Phys. Rev. Lett. 81, 3579 (1998). 24. Waxman, E., Phys. Rev. Lett. 75, 386 (1995). 25. Waxman, E. and Miralda-Escude, J., Astrophys. J. 472, 489 (1996). 26. Thorne, K. S., Price, R. M. and MacDonals, D., 'Black Holes: The Membrane Paradigm" (New Haven: Yale University Press, 1986). 27. Boldt, E. and Ghosh, P., astro-ph/9902342 (1999). 28. Thompson, C. A. and Duncan, R. C , Mon. Notes R. Astron. Soc. 275, 255 (1995). 29. Kouveliotou, C. et al, Nature 393, 235 (1998).
© Physica Scripta 2000
Physica Scripta.Vol. T85, 195-209, 2000
Particle Astrophysics with High Energy Photons T. C. Weekes* Whipple Observatory, Harvard-Smithsonian CfA, P.O. Box 97, Amado, AZ 85645-0097, U.S.A. Received January 19, 1999; accepted August 2, 1999
PACS Ref: 95.55Ka; 98.70Rz; 98.58MJ; 98.54Cm
Abstract The development of the atmospheric Cherenkov imaging technique has led to significant advances in y-ray detection sensitivity in the energy range from 200 GeV to 50 TeV The Whipple Observatory 10m reflector has detected the first galactic and extragalactic sources in the Northern Hemisphere; the Crab Nebula has been established as the standard candle for ground-based y-ray astronomy. The highly variable Active Galactic Nuclei, Markarian 421 and Markarian 501, have proved to be particularly interesting. A new generation of telescopes with improved sensitivity has the promise of interesting measurements of fundamental phenomena in physics and astrophysics. VERITAS (the Very Energetic Radiation Imaging Telescope Array System) is one such next generation system; it is an array of seven large atmospheric Cherenkov telescopes planned for a site in southern Arizona.
1. The relativistic universe It has to be with a certain sense of apology that one introduces the subject of photon astronomy to a symposium devoted to particle physics. Photons are, by any definition, rather dull specimens of the cosmic particle zoo. However one can argue that their very dullness, their lack of charge and mass, their infinite lifetime, their appearance as a decay product in many processes, their predictability, all combine to make them a valuable probe of the behavior of more exotic particles in distant, and therefore difficult to study, regions of the universe. Certainly no one can argue that photon astronomy at low energies (optical, radio and X-ray) has not had a major influence in our perception of the physical universe! Our universe is dominated by objects emitting radiation via thermal processes. The blackbody spectrum dominates, be it from the microwave background, the sun or the accretion disks around neutron stars. This is the ordinary universe, in the sense that anything on an astronomical scale can be considered ordinary. It is tempting to think of the thermal universe as THE UNIVERSE and certainly it accounts for much of what we see. However to ignore the largely unseen, non-thermal, relativistic, universe is to miss a major component and one that is of particular interest to the physicist, particularly the particle physicist. The relativistic universe is pervasive but largely unnoticed and involves physical processes that are difficult to emulate in terrestrial laboratories. The most obvious local manifestation of this relativistic universe is the cosmic radiation, whose origin, 86 years after its discovery, is still largely a mystery (although it is generally accepted, but not proven, that much of it arises in shock waves from galactic supernova explosions). The existence of a steady rain of particles, whose power law spectrum attests to their non-thermal origin and whose highest ener* e-mail: [email protected] © Physica Scripta 2000
gies extend far beyond that achievable in man-made particle accelerators, attests to the strength and reach of the forces that power this strange relativistic radiation. If thermal processes dominate the "ordinary" universe, then truly relativistic processes illuminate the "extraordinary" universe and must be studied, not just for their contribution to the universe as a whole but as the denizens of unique cosmic laboratories where physics is demonstrated under conditions to which we can only extrapolate. The observation of the extraordinary universe is difficult, not least because it is masked by the dominant thermal foreground. In places, we can see it directly such as in the relativistic jets emerging from AGNs but, even there, we must subtract the foreground of thermal radiation from the host elliptical galaxy. Polarization leads us to identify the processes that emit the radio, optical and X-ray radiation as synchrotron emission from relativistic particles, probably electrons, but polarization is not unique to synchrotron radiation and the interpretation is not always unambiguous. The hard, power-law, spectrum of many of the non-thermal emission processes immediately suggests the use of the highest radiation detectors to probe such processes. Hence hard X-ray and y-ray astronomical techniques must be the observational disciplines of choice for the exploration of the relativistic universe. Because the earth's atmosphere has the equivalent thickness of a meter of lead for this radiation, its exploitation had to await the development of space platforms for X-ray and y-ray telescopes. Although the primary purpose of the astronomy of hard photons is the search for new sources, be they point-like, extended or diffuse, it opens the door to the investigation of more obscure phenomenon in high energy astrophysics and even in cosmology and particle physics. Astronomy at energies up to 10 GeV has made dramatic progress since the launch of the Compton Gamma Ray Observatory in 1991 and that work has been summarized [1]. Beyond 10 GeV it is difficult to efficiently study y-rays from space vehicles, both because of the sparse fluxes which necessitate large collection areas and the high energies which make containment a serious problem. The development of techniques whereby y-rays of energy 100 GeV and above can be studied from the ground, using indirect, but sensitive, techniques is relatively new and has opened up a new area of high energy photon astronomy with some exciting possibilities and some preliminary results. The latter include the detection of TeV photons from supernova remnants and from the relativistic jets in AGNs. Such observations seriously constrain the models for such sources and in many cases lead to the development of a new paradigm. There remains the possibility that the annihilation lines from neutralinos might be discovered in the GeV-TeV region, that the evaporation of Physica Scripta T85
196
T. C. Weekes
primordial black holes might be manifest by the emission of bursts of TeV photons, that the infrared density of intergalactic space might be probed by its absorbing effect on TeV photons from distant sources, and even (in some models) that the fundamental quantum gravity energy scale might be constrained by the observation of short-term TeV flares in extragalactic sources. 2. Detection technique The techniques of ground-based Very High Energy (VHE) y-ray astronomy are not new but only achieved credibility in the late eighties with the detection of the Crab Nebula. The most sensitive technique, the atmospheric Cherenkov imaging technique, is the one that has been most successful and is now in use at some eight observatories. Its history and present status has been reviewed elsewhere [2]. It is an optical "telescope" technique and thus suffers the usual limitations associated with optical astronomy: limited duty cycle, weather dependence, limited field of view. But it also has the advantage that it is relatively inexpensive because it uses the same detector technology (photomultipliers) as optical astronomy, the same optical reflectors that borrow from solar energy investigations, and the same pulse processing techniques that are routinely used in high energy particle physics. In addition the Cherenkov technique operates in an energy regime where the physics of particle
interactions is relatively well understood and where there exist advanced Monte Carlo programs for the simulation of particle cascades. In recent years, VHE y-ray astronomy has been dominated by two advances in technique: the development of the atmospheric Cherenkov imaging technique, which led to the efficient rejection of the hadronic background, and the use of arrays of atmospheric Cherenkov telescopes to measure the energy spectra of y-ray sources. The former is exemplified by the Whipple Observatory 10m telescope (Fig. 1) with more modern versions CAT, the French telescope in the Pyrenees, and the Japanese-Australian CANGAROO telescope in Woomera, Australia. The most significant examples of the latter are thefivetelescope array of imaging telescopes on La Palma in the Canary Islands which is run by the Armenian-German-Spanish collaboration, HEGRA, and the four, soon to be seven, Telescope Array in Utah which is operated by a group of Japanese institutions. These techniques are relatively mature and the results from observations with overlapping telescopes are in good agreement. Vigorous observing programs are now in progress at all of these facilities; the vital observing threshold has been achieved whereby both galactic and extragalactic sources have been reliably detected. Many exciting results are anticipated as more of the sky is observed with this generation of telescopes. 3. Galactic sources
It is a measure of the maturity of this new discipline that the existence and study of galactic sources of TeV radiation is now considered ordinary and relatively uncontroversial. This is a dramatic change from only a decade ago when the existence of any galactic sources at all was hotly contested. These sources were always variable and difficult to confirm or refute [4]; it was not until the observation of steady sources, in particular, the observation of the Crab Nebula (which has become the standard candle), that the relative sensitivity of the different techniques could be assessed and some standards of credibility set. The Crab Nebula has been observed by some eight independent groups and no evidence for variability has been detected. It has been seen at energies from 200 GeV to more than 50 TeV and accurate energy spectra have been determined [3]. Originally predicted by Gould [5] as a TeV energy source based on a Compton-synchrotron model, the complete y-ray spectrum can now befittedby an updated version of the same model [3]. The variable parameter in this model is the magnetic field which is set by the TeV observations at 16±1 nanotesla, somewhat smaller than the value estimated from the equipartition of energy (Fig. 2). In practice, recent optical observations reveal a complex structure at the center of the nebula (where the TeV photons are believed to originate) and more sophisticated models are certainly called for. VHE y-rays have also been detected from other galactic sources. All of these detections are of sources with negative declinations, best seen in the Southern Hemisphere where there are fewer VHE observatories and hence the detections have largely been by one group. The exception is the y-ray Fig. 1. The Whipple Observatory 10 m reflector which will be the prototype pulsar PSR1706-44 which was discovered by the CANGAROO group [6] and confirmed by the Durham for the telescopes in VERITAS. Physica Scripta T85
© Physica Scripta 2000
Particle Astrophysics with High Energy Photons
197
certainly have sufficient energy and their occurrence rate is about right; also there is a known mechanism associated with shock fronts to explain acceleration. Hence when EGRET detected a small number of y-ray sources at 10 GeV energies which appeared to coincide with known SNRs [11], it was widely believed that the cosmic ray origin problem had been solved. However Drury et al. [12] had shown that the y-ray spectrum of such sources should be rather flat power-laws that would extend to TeV energies. Extensive observations by the Whipple collaboration have failed to find any evidence for TeV emission [13]. The upper limits 10 are shown in Fig. 3 along with the EGRET points. More elaborate models have been constructed that can be made to fit the observations [14]. It is also possible that the o Method 1 1988-9 EGRET source/SNR identifications are in error since the o Method 1 1995-6 sources are not strong and the galactic y-ray plane is a con* Method 2 1988-9 fused region at MeV/GeV energies. Either way, it would • EGRET A CANGAR00 be reassuring for theories of cosmic ray origins to see definite 10 detections from some shell-type SNRs where the emission is 10 10 10 10 10 10 E (TeV) consistent with n production in the shell. The next generation of VHE detectors should provide these definitive Fig. 2. TeV spectra of the Crab and the predicted inverse-Compton spectrum observations. for three magnetic field strengths (from [3]). Shown are data from EGRET, CANGAROO, and three different epochs/data analysis methods for Whipple.
group [7]; both of these groups operate from Australia. The source is detected by EGRET at MeV-GeV energies as 100% pulsed. There is no evidence in the TeV signal for pulsations but there is weak evidence that the pulsar is in a plerion which may be the source of the TeV y-rays. The CANGAROO group also report the detection of an unpulsed TeV signal from a location close to the Vela pulsar [8]; the position coincides with the birthplace of the pulsar and hence the signal may originate in a weak plerion left after the ejection of the pulsar. Another interesting result is the detection of Cen X-3 by the Durham group [9]. Perhaps the most surprising (and controversial) result is the detection of a TeV source that is coincident with one part of the shell of the supernova remnant, SN1006 [10]. X-ray observations had shown that there is non-thermal emission from two parts of the shell that is consistent with synchrotron emission from electrons with energy up to 100 TeV; hence the TeV y-ray detection is not a surprise. The TeV emission is consistent with inverse Compton emission from electrons which have been shock accelerated in the shell. However it is not clear why it should be seen from only one region. Because this represents the first direct detection of SNR shell emission this result, when confirmed, has great significance. Not only can the magnetic field be estimated but also the acceleration time; these two parameters are very important for shock acceleration theory. More sensitive observations may reveal the detailed energy spectrum, whether or not the source is extended, and the relative strength of the TeV emission from each shell. Ideally, of course, one would like to see direct evidence from VHE y-ray astronomy of emission from hadron collisions in SNR shells. These SNRs are widely believed to be the source of the hadronic cosmic rays seen in the solar system (at least up to proton energies of 100 TeV) which fill the galaxy. However this canonical model mostly rests on circumstantial evidence and it is highly desirable to find the smoking gun that would clinch the issue. Supernovae © Physica Scripta 2000
4. Extragalactic sources 4.1. Relativistic Jets One of the most surprising results to come from VHE y-ray astronomy has been the discovery of TeV-emitting blazars. Unlike the observation of galactic supernovae such as the Crab Nebula, which are essentially standard candles, the light-curves of blazars are highly variable. In Fig. 4 the nightly averages of the TeV flux from Markarian 421 (Mkn 421) in 1995 are shown as observed at the Whipple Observatory [15]. Although AGN variability was a feature of the AGNs observed by EGRET on the Compton Gamma Ray Observatory at energies from 30 MeV to 10 GeV, the weaker signals (because of the finite collection area) do not allow such detailed monitoring, particularly on short time-scales. Active galactic nuclei (AGN) are the most energetic on-going phenomena that we see in extragalactic astronomy. The canonical model of these objects is that they contain massive black holes (often at the center of elliptical galaxies) surrounded by accretion disks and that relativistic jets emerge perpendicular to the disks; these jets are often the most prominent observational feature. Blazars are an important sub-class of AGNs because they seem to represent those AGNs which have one of their jets aligned in our direction. Observations of such objects are therefore unique. The VHE y-ray astronomer is thus in the position of the particle physicist who is offered the opportunity to observe the accelerator beam, either head-on or from the side. For the obvious reason that there is more energy transferred in the forward direction the particle physicist usually chooses to put his most important detectors directly in the direction of the beam (or close to it) and its high energy products. While such observations give the best insight into the energetic processes in the jet, they do not give the best pictorial representation. Hence just as it is difficult to visualize the working of a cannon by looking down its barrel, it is difficult to get a picture of the jet by looking at it Physica Scripta T85
198
T. C. Weekes i MINII
i inull!
i IIIIIIII
i i imil
i i iMil
! IIIIIIII
i IIIIIIII
I IHMIII
i i mig
W44
(fi
7
I ! I Hill—I l l l l l l l l
I llllllll
I lllltlil
I llllllij
I llllllll
y — Cygni
6
I Hilt
:
10 E -7 o 10 c
-o
o 10 "o -9 Q. 1 0 10
10 •11
5 10 f»10 CD
^
c
•13
0 CM^.; H.l.llllil
10
l..U.U.llll...J..l.illJJii....l..l.l.l.lilll
I;
1
10
4
-14
10~
l-i-l.U!ill. i„i.i.iJIit[::...l..J.':lUliJil....i.S'j.lJJi
2
-2
10 6
M ±..LIUilll
Llj.llllil
-2
I llllllll
I llllllll
I I I MM?
I lllliill
I llllilll
l..l.llJiu[....i„Ullilit..„l..i.imijl
„
_ 2
10~
1
10 10 Energy (GeV)
I I illllll
c -=
IT |
I llllllll
UlUlliLjjjkllCjJJIllll
4
10 10 Energy (GeV)
I I lilt
IC-44-3
:
CM C
.Lllllllll
10
LWliULJjaiJlJlLjLLillllLj^
1
.0
10 10 Energy (GeV)
Fig. 3. Whipple Observatory flux upper limits (indicated by W) for shell-type SNRs. The lower energy points are from EGRET. The solid curves are extrapolations from the EGRET integral fluxes for y-rays arising from if decay. Upper limits from air shower arrays are also shown.
head-on. Observations at right angles to the jet give us our best low energy view of the jet phenomenon and indeed provide us with the spectacular optical pictures of jets from nearby AGNs (such as M87). 4.2. Sources
Mkn 421 is the closest example of an AGN which is pointing in our direction. It is a BL Lac object, a sub-class of blazars, so-called because they resemble the AGN, BL Lacertae which is notorious because of the lack of emission lines in its optical spectrum. Because such objects are difficult, and somewhat uninteresting, for the optical astronomer they Physica Scripta T85
were largely ignored until they were found to be also strong and variable sources of X-rays and y-rays. Mkn 421 achieved some notoriety largely because it was the first extragalactic source to be identified as a TeV y-ray emitter [16]. At discovery, its average VHE flux was «s 30% of the VHE flux from the Crab Nebula. Markarian 501 (Mkn 501), which is similar to Mkn 421 in many ways, was detected as a VHE source by the Whipple group in May 1995 [17]. It was only 8% of the level of the Crab Nebula and was near the limit of detectibility of the technique at that time. The discovery was made as part of an organized campaign to observe objects that were similar to Mkn 421 and were at small redshifts. This same campaign © Physica Scripta 2000
Particle Astrophysics with High Energy Photons
hours. The very large collection areas (> 10,000 m2) associated with atmospheric Cherenkov Telescopes is ideally suited for the investigation of short term variability. The VHE emission from the two best observed sources, Mkn 421 and Mkn 501 (Fig. 5), varies by a factor of a hundred. Although many hundreds of hours have now been devoted to their study, the variations are so complex that it is still difficult to characterize their emissions. It has been suggested [15] that for Mkn 421 the emission is consistent with a series of short flares above a baseline that falls below the threshold of the Whipple telescope (Fig. 4); the average flare duration is one day or shorter. The most important observations of Mkn 421 were in May, 1996 when it was found to be unusually active. On May 7, a flare was observed with the largest flux ever
later yielded the detection of the BL Lac object, 1ES 2344+514 ([18] which is also close (z = 0.044). Recently the Durham group has announced the detection of the BL Lac object, PKS2155-304 [19] which is also at a small redshift (z = 0.116). A more controversial, but potentially more important detection, is that of 3C 66A reported by the Crimean group [20]. These sources are summarized in Table I. Whereas the first two sources have been seen by a number of groups, the last three are reported by only one group and require confirmation. 4.3. Variability Perhaps the most exciting aspect of these detections is the observation of variability on time-scales from minutes to
Table I. Properties of the
Object Mkn 421 Mkn 501 1ES 2344+514 PKS 2155-304 3C66A
0.031 0.034 0.044 0.116 0.444
VHE BL Lac objects.
EGRET flux" (E > 100 MeV) (lO-'cnrV)
Average flux (E > 300 GeV) 12 2 (10 cm- s-')
1.4±0.2 3.2±1.3 <0.7 3.2±0.8 2.0±0.3
40 >8.1 <8.2 42 30b
a
Radio, optical, and X-ray data from [21]. [23]. b 1 TeV flux value.
(5 GHz)
M'
^x a (2 keV) (vJy)
14.4 14.4 15.5 13.5 15.5
3.9 3.7 1.1 5.7 0.6
720 1370 220 310 806
(mjy)
EGRET data from D.J. Thompson (priv. comm.), [22],
and
" i — i — i — i — i — i — i — i — i — i — i — i — i — i — i — r
?^ 2.5
199
" i — I — i — r
-
o
a: >s
o 2 cr I o E E 1.5 o o
o
0.5
0
•0.5
1
J_
49720
49740
J
49760
I
I
I
I
I L_
49780
_L
49800
J
i
i
i
49820
i
i
i
i •
49840
i
i
i
i
49860
i i_
49880
Modified Julian Day Fig.
4. Daily VHE y-ray count rates for Mkn 421 during 1995 (from [15]).
© Physica Scripta 2000
Physica Scripta T85
200
c
T. C. Weekes
a) Monthly Averages
2
**
_Q D
o CD D
0 6
1996
1995 _)
,
,
,
1_
H
1
1998
1997 1
1
— '
H
1 —
(b) Daily Averages 5 ~
3
2
0
49800
50000
50200
50400
50600
50800
51000 MJD
Fig. 5. Average monthly (a) and daily (b) VHE y-ray flux (in units of VHE Crab flux) for Mkn 501 between 1995 and 1998 (from [26]).
recorded from a VHE source. The observations began when the flux was already several times that of the Crab Nebula and it continued to rise over the next two hours before levelling off (Fig. 6). Observations were terminated as the moon rose but the following night it was observed at its quiescent level. One week later (May 15) a smaller, but shorter, flare was detected; in this case the complete flare was observed and the doubling time in the rise and fall was ^ 1 5 minutes. This is the shortest time variation seen in any extragalactic y-ray source at energies >10MeV (apart from in a y-ray burst). Mkn 501 is also variable, but as at other wavelengths, the characteristic time seems longer. Its baseline emission has varied by a factor of 15 over four years [26] (Fig. 5). Hour-scale variability has also been detected but its most important time variation characteristic appears to be the slow variations seen over the five months in 1997. Physica Scripta T85
4.4. Spectrum The atmospheric Cherenkov signal is essentially calorimetric and hence it should be possible to derive the y-ray energy spectrum from the observed light pulse spectrum. In practice it is more difficult because, unless an array of detectors is used, the distance to the shower core (impact parameter) is unknown. Although the extraction of a spectrum from even a steady and relatively steady source as the Crab Nebula required considerable effort and the development of new techniques, it was relatively easy to measure the spectra of Mkn 421 and Mkn 501 in their high state because the signal was so strong. The general features of the spectra derived from the Whipple observations are in agreement with those derived at the HEGRA telescopes [28]. The May 7, 1996 flare of Mkn 421 provided an excellent data base for the extraction of a spectrum; the data can be fit by a simple power-law (dN/dE oc E~26). There is © Physica Scripta 2000
Particle Astrophysics with High Energy Photons
201
Mrk421
T—i—i—I—i—i—i—i
T—i—r
'E
^o
May 07 1996J
16
I
I
I—r
c
14
'E
o
o cz
QC
I o
O
D
I
E E
E E
May 15 1996^
^ 7 6
5 h
o
o
o
o
4 3
6
2
4 2
0 J
I
I
4
I
I
I
I
I
5
I
I
l
I
I
6
I
I
I
I
I
I
L
J
7 UTC (hrs)
}
#
0
i
i
i
i
4
i
i
i
i
i
i
i
5
11 J f .1 I Mr 6
7 UTC (hrs)
Fig. 6. Mkn 421 flares of 1996 May 7 (left) and May 15 (right) (adapted from [25]).
the highest energy end of the spectrum. Large zenith angle observations at Whipple [27] and observation by HEGRA [28] confirm the absence of a cutoff out to 10 TeV. The generally high state of Mkn 501 throughout 1997 give data from the Whipple telescope that can be best fit by a curved spectrum of the form: dN/dE and £-2.20-o.45/og10£ [29] (Fig. 7). Here the spectrum extends to at least 10 TeV. The curvature in the spectrum could be caused by the intrinsic emission mechanism or by absorption in the source. Since Mkn 421 and Mkn 501 are virtually at the same redshift it is unlikely that it could be due to intergalactic absorption since Mkn 421 does not show any curvature [30]. 4.5. Multiwavelength
observations
The astrophysics of the y-ray emission from the jets of AGNs are best explored using multiwavelength observations. These are difficult to organize and execute because of the different observing constraints on radio, 10 Photon Energy (TeV) optical, X-ray, space-based y-ray and ground-based y-ray observatories. Of necessity observations are often Fig. 7. VHE spectra of Mkn 421 (filled circles) and Mkn 501 (open stars) as incomplete and, when complete coverage is arranged, the measured with the Whipple Observatory telescope [30]. source does not always cooperate by behaving in an interesting way! The first multiwavelength campaign on Mkn 421 no evidence of a cutoff up to energies of 5 TeV [24] (Fig. 7). Because of the possibility of a high energy cutoff due to coincided with a TeV flare on May 14-15, 1994 and showed intergalactic absorption there is considerable interest in some evidence for correlation with the X-ray band; however © Physica Scripta 2000
Physica Scripta T85
T. C. Weekes
202 1
-
(a)
• •
f
+
+
•
0
-
\
• *
+
't 40
Whipple Observations (> 300 G e V y - r a y s )
\
1 5
I •
/
\
0
*
i
(c)
135SE (50-150 keV)
V
/ •
_
t'°
ASCASIS0+SIS1 ( 0 . 7 - 7 . 5 keV X - r a y s )
(b)
20
0.4
Mrk 501 Observations, April 1997
Whipple Observations (> 350 GeV Y-rays)
v» *
EUVE DS/S ( 5 8 - 1 7 4 AXUV)
A *.
•""•v..
.•*•
-S-300
I
RXTEPCA • (2-10 keV) •(15-25 keV)(xl5)
; 200
3
-••-•./••
0.2
++
0.5
E
i
100 (d)
i
1.1
•i
1
0.9
+
,:
, , I 49830 49832
Optical Observations (650 nm)
t *
, , i 49834 49836
i
t
1 1
1
+ "44
If +
3
2
E
i
1
l
i , , , I , , , i , , , 49838 49840 49842 49844 Modified Julian Day
|
RXTE ASM (2-10 keV)
-+++ ++
1
t! " ^ * '
T t ++ + + + +++ + +-
CfA 1.2m (U-band) •
.
0.9
x 2 0.8
E
50540 50542 50544 50546 50548 50550 50552 50554 50556 50558 MJD Fig. 8. Left: Multi-wavelength observations of Mrk 421 (from [15]): (a) VHE y-ray, (b) X-ray, (c) extreme UV, and (d) optical lightcurves taken during the period 1995 April-May (April 26 corresponds to MJD 49833). Right: Multi-wavelength observations of Mkn 501 (adapted from [33]): (a) y-ray, (b) hard X-ray, (c) soft X-ray, (d) U-band optical taken during the period 1997 April 2-20 (April 2 corresponds to MJD 50540). The dashed line in (d) indicates the opticalfluxin 1997 March.
no enhanced activity was seen in EGRET [31]. A year later, in a longer campaign, there was again correlation between the TeV flare and the soft X-ray and UV data but with an apparent time lag of the latter by one day [15] (Fig. 8). The variability amplitude is comparable in the X-ray and TeV emission («s 400%) but is smaller in the EUV («200%) and optical (%20%) bands. In April, 1998 there was again a correlation seen between an X-ray flare observed by SAX and Whipple; in this case the TeV flare was much shorter (a few hours) compared to the X-ray (a day) [32]. The first multiwavelength campaign on Mkn501 was undertaken when the TeV signal was seen to be at a high level. The surprising result was that the source was detected by the OSSE experiment on CGRO in the 50-150 kev band (Fig. 8). This was the highest flux ever recorded by OSSE from any blazar (it has not detected Mkn 421) but the amplitude of the X-ray variations (^200%) was less than those of the TeV y-rays («400%) [33].
of Compton-synchrotron models, e.g., the Crab Nebula. Whereas the synchrotron peak in Mkn 421 occurs near 1 keV, that of Mkn 501 occurs beyond 100 keV which is the highest seen from any AGN. In 1998 the synchrotron spectrum peak in Mkn 501 shifted back to 5 keV and the TeV flux fell below the X-ray flux. 4.7. Implications The sample of VHE emitting AGNs is still very small but it is possible to draw some conclusions from their properties (summarized in Table I). •
4.6. Multiwavelength power spectra Because of the strong variability in the TeV blazars it is dif- • ficult to represent their multiwavelength spectra. In Fig. 9 we show the fluxes plotted as power (v x Fv) from Mkn 421 and Mkn 501 during flaring as well as the average fluxes. Both sources display the two peak distribution characteristic Physica Scripta T85
The first three objects, all detected by the Whipple group, are the three closest BL Lacs in the northern sky. Some 20 other BL Lacs have been observed with z < 0.10 without detectable emission. This could be fortuitous, because they are standard candles and these are closest (but the distance differences are small), or because they suffer the least absorption (but there is no cutoff apparent in their spectra). All of the objects are BL Lacs; because such objects do not show emission lines and therefore probably do not have strong optical/infrared absorption close to the source, it is suggested that BL Lacs are preferentially VHE emitters. © Physica Scripta 2000
Particle Astrophysics with High Energy Photons "T
T"
-
II
II
1,
ll
II
Jl
II
:
*1995 - .1996 ' • 1997 April 9-15 " • 1997 April 16 gf -io _ • 1997 April 9 B io : o Archival J? : o 1997 Average " °1998
May 1996 April 1995 May 1994 1977 to 1995
" ''J'
'I
' 1 ' 1
' 1M
'I
203
' 1 ' 1 ':
+
•sTrk
HT
*^ D n
+ M •
1
°°o ?
1 :
o
(ft °o
1
0
!l )
: .1 :
~p CD
"
:
O
•
o
0
-13
10
i :
-o
t\ I . . .
r8
•
o
,1
1
,1
, 1
, 1
,1
,1
,1
,1
,1
,1
,i
, i
, i
, i
,
I
12
20
24
26
28
v(Hz)
log v (Hz)
Fig. 9. Left: The multi-wavelength power spectrum of Mkn 421 (adapted from [15]). The dashed line shows an SSC model fit to the data. Right: The multi-wavelength power spectrum of Mkn 501 (adapted from [33]).
•
•
•
•
Four of the five sources are classified as XBLs which indicates that they are strong in the X-ray region and that the synchrotron spectrum most likely peaks in that range (and that the Compton spectrum peaks in the VHE y-ray range). The fifth, 3C 66A, is an RBL, like many of the blazars detected by EGRET; it is believed that these blazars have synchrotron spectra that peak at lower energies and Compton spectra that peak in the HE y-ray region. Only three (Mkn 421, PKS 2155-304 and 3C 66A) are listed in the Third EGRET Catalog; there is a weak detection reported by EGRET for Mkn 501. If 3C 66A is confirmed (and to a lesser extent PKS 2155-305), then the intergalactic absorption is significantly less than had been suggested from galactic evolution models. There is evidence for variability in all of the sources. The rapid variability seen in Mkn 421 indicates that the emitting region is very small which might suggest it is close to the black hole. In that case the local absorption must be very low (low photon densities). It seems more likely that the region is well outside the dense core.
There are three basic classes of model considered to explain the high energy properties of BL Lac jets: Synchrotron Self Compton (SSC), Synchrotron External Compton (SEC) and Proton Cascade (PC) Models. In the first two the progenitor particles are electrons, in the third they are protons. VHE y-ray observations have constrained the types of models that are likely to produce the y-ray emission but still do not allow any of them to be eliminated. For instance, the correlation of the X-ray and the VHE flares is consistent with the first two models where the same population of electrons radiate the X-rays and y-rays. There is little evidence for the IR component in BL Lac objects which would be necessary in the SEC models as the targets for Compton-scattering, so this particular type of model may not be likely for these objects. The PC models which produce © Physica Scripta 2000
the y-ray emission through e + e~ cascades also have great difficulty explaining the rapid cooling observed in the TeV emission from Mkn 421. Also the high densities of unbeamed photons near the nucleus, such as the accretion disk or the broad line region, are required to initiate the cascades and these cause high pair opacities to TeV y-rays [34]. Significant information comes from the multiwavelength campaigns (although thus far these have been confined to Mkn 421 and Mkn 501). Simultaneous measurements constrain the magnetic field strength (B) and Doppler factor (<5) of the jet when the electron cooling is assumed to be via synchrotron losses. The correlation between the VHE y-rays and optical/UV photons observed in 1995 from Mkn 421 indicates both sets of photons are produced in the same region of the jet; S > 5 is required for the VHE photons to escape significant pair-production losses [15]. If the VHE y-rays are produced in the synchrotronself-Compton process, 5 = 15 - 40 and B = 0.03 - 0.9G for Mrk 421 [35], [36] and d < 15 and B = 0.08 - 0.2G for Mkn 501 [29], [36]. On the other hand by assuming protons produce the y-rays in Mkn 421, Mannheim [37] derives 5 = 16 and B = 90G. The Mkn 421 values of 5 and B are extreme for blazars, but they are still within allowable ranges and are consistent with the extreme variability of Mkn 421. 5. Intergalactic absorption Thus far it has not been possible to make a direct measurement of the infrared background radiation at wavelengths more than 3.5 microns and less than 140 microns. This is unfortunate since the background potentially contains valuable information for cosmology, galaxy formation and particle physics. The problem for direct measurement is the presence of foreground local and galactic sources. However the infrared background can make its presence felt by the absorption it produces on the spectra of VHE Physica Scripta T85
204
T. C. Weekes
y-ray sources when they are at great distances. The absorption is via the yy -» e + e~ process, the physics of which is well understood. The maximum absorption occurs when the product of the energy of the two photons (y-ray and infrared) is approximately equal to the product of the rest masses of the electron-pair. Hence a 1 TeV y-ray is most heavily absorbed by O.leV (1.2 micron) infrared photon in head-on collisions. The importance of this effect for VHE and UHE y-ray astronomy was first pointed out by Nikishov [38]; its potential for making an indirect measurement of the infrared background was pointed out by Gould and Schreder [39] and, more recently, in the aftermath of the EGRET detections of AGNs, by Stecker and de Jager [40]. At the redshift of the AGNs detected at VHE energies to date (0.03 to 0.5) if the infrared density has the value assumed in some models [40], the effect is appreciable and should be apparent in carefully measured energy spectra in the range 1 to 50 TeV. Ideally for such a measurement the intrinsic emission spectrum of the y-rays from the distant source should be known. In practice this is not the case although thus far all the AGNs detected in the GeV-TeV range appear to have very smooth power-law spectra. Biller et al. [41] have made a conservative derivation of upper limits on the infrared spectrum based on the measured y-ray spectrum from 0.5 to 10 TeV from Mkn 421 and Mkn 501 by the Whipple and HEGRA groups (Table II). These upper limits apply to infrared energies from 0.025 to 1.0 eV; they are the best upper limits over this range. At some wavelengths, these limits are as much as an order of magnitude below the upper limits set by the DIRBE/COBE satellite (see Fig. 10). For the derivation the following assumptions are made:
A (/i,m) 12.5
125 10°
1
s
a
10"'
> 0)
1
i
i i i-
D
:
+ -H-
+
co
I
1
:
CO
o
••
1.25 1
10"
•
+ X ' ++ + £
+
D+
"a
+
a
~ -
<.a
-
10"
10°
(eV) Fig. 10. Upper limits to the background IR density. The solid lines show the limits derived assuming the VHE spectra of Mkn 421 and Mkn 501 extend to 10 TeV. Other symbols represent upper limits from direct measurements (squares are from COBE/DIRBE measurements). Figure from ([41)]. Physica Scripta T85
•
•
Then infrared densities are calculated such that they do not cause the shape of the observed VHE spectrum to deviate from the bounds set from the VHE measurements. This approach has the effect of anchoring the lower energy TeV data to the appropriate infrared upper limits and then extending these bounds so that they are consistent with those based on the shape and extent of the AGN spectra at the higher energies. Thus the maximum energy density in each interval of infrared energy is determined; these limits are given in Table II where the maximum energy in the measured y-ray spectrum is taken to be 10 TeV. They are plotted in Fig. 10; also shown are the upper limits from other methods. These upper limits do not conflict with the predictions of the infrared background based on detailed models of galactic evolution [42]. They do however allow some more cosmological possibilities to be eliminated. In particular in one scenario, density fluctuations in the early universe (z ^ 1000) could have produced very massive objects which would collapse to black holes at later times and could explain the dark matter. However although undetectable now, they would have produced an amount of infrared radiation that would have exceeded the above limits [41]. These limits also place some constraints on radiative neutrino decay.
6. Gamma ray bursts The contribution of TeV observations to the physics of y-ray bursts is at once the most speculative and most important (potentially) of all the scientific topics considered here. As yet, there is no positive detection of TeV photons during or immediately after a classical y-ray burst (GRB) (although there is one tantalizing but unconfirmed observation [43]). However since there is no turn over seen in the spectra of GRBs detected by EGRET at energies > 30 GeV, there is the potential for interesting observations at VHE energies. The observed EGRET spectra are power
Table II. Derived upper limits for e2.n(e).
, : 1
10"
•
the observed spectra of Mrk421 and 501 over the energy are represented by power law with differential index between —2.3 and —2.8, the differential index of the inherent source spectrum is less than —2.0, the infrared background spectrum can be represented by a relatively smooth spectrum with no abrupt changes or strong lines, the infrared energy distribution can be considered in broad intervals (Table II) and the infrared density in these intervals is constrained to be lower than the upper limits set by other methods.
.
co 10"
•
Upper limit to £2.H(E)
Energy interval (eV)
(eV/cm3)
0.025-0.042 0.042-0.083 0.082-0.167 0.167-0.330 0.330-0.625 0.625-1.000
0.0064 0.0027 0.0027 0.0035 0.0052 0.0120
© Physica Scripta 2000
Particle Astrophysics with High Energy Photons
A feature of the EGRET GRB observations was that there was evidence for delayed emission (up to 1.5 hours) from the burst site [46]. This may indicate a different component at these energies. Some models [47] predict that this delayed emission could persist for days and could hence be easily observed with narrow field of view instruments. The detection of a TeV y-ray component in a GRB would be a serious parameter for the emission models, in particular the Lorentz bulk motion in the source would be constrained. It would also be an independent distance indicator since the source would have to show absorption if the redshift was > 0.1.
y-ray source at the Galactic Center [49] at energies above 300 MeV (Fig. 12). There is no significant evidence for variability and the spectrum is unlike other galactic sources; this suggests that the source is unusual and unlike other galactic sources. Preliminary observations (4.5 hours) by the Whipple group [50] did not give a significant signal at 1.5 TeV energies. In a recent paper, Bergstrom, Ullio and Buckley [51] have estimated the flux from the annihilation radiation of neutrinos in the Galactic Center using the most recent models of the galactic mass distribution. The predicted line has a relative width of 10~3. Neither space nor ground-based detectors have energy resolution of this quality (even in the next generation of detectors) but the intensity of the line is such that it might be detectable even with relatively crude energy resolution. A broad spectrum of secondary y-rays would also be detected from the other annihilation modes giving a spectrum of the shape shown in Fig. 11 The wide range of possible parameters in supersymmetric space give a range of predicted y-ray fluxes. These are shown in Fig. 13 where the sensitivity of various ground-based y-ray experiments are also shown. The line labelled Whipple is the current sensitivity of the Whipple experiment and is seen to be unable to constrain the predictions. GRANITE III represents the expected sensitivity of a current upgrade to the Whipple telescope which should be completed in 1999. VERITAS is a next generation array of telescopes (see below) planned for southern Arizona. The sensitivity for a similar array in the Southern Hemisphere is also shown; the increased sensitivity of the telescope in the Northern Hemisphere comes from the enhanced performance of atmospheric Cherenkov telescopes when operated at large zenith angles.
-4
10
1—i—r'fi'Tn
1
1—i—i—r' " 1 1
NFW mode!, A i i ^ 1 0 " 3 sr
\s V
\
O
10" 8
N.
in M
£ ii X 3 LL C
o o sz
~= :
-N.
ml i i
>
"V
\y
\
\ .
10"
X Nv
9
: -
\ \
\.
>v
\ S.
-a
10
\
Q.
-11
10
io" ,a
\
-2 7
Background, E„ — - 300 GeV neutralino added 50 GeV neutralino added
-13
io-
; "= ;
\ \
N.
-10
10
I 1
: , :
10"!
10
© Physica Scripta 2000
1
Neutralino continuum gamma ray flux towards galactic centre -
J-l
The best candidate for the cold dark matter component is the neutralino, the lightest super symmetric particle. These particles annihilate with the emission of a single y-ray line whose energy is equal to that of the neutralino mass; however other annihilation modes are possible and there may be a y-ray continuum (Fig. 11). There are limits on the possible masses from cosmology and from accelerator experiments but the range from 30 GeV to 3 TeV is allowed. The upper part of this range would be accessible for study by ground-based y-ray telescopes. The neutralinos, if they exist, would be expected to cluster to the center of galaxies and might be detectable by their y-ray emission, either as a line or a continuum. Detailed numerical simulations indicate that there may be a strong density enhancement towards the centers of galaxies such as our own. Hence the Galactic Center is a prime candidate for observations. This hypothesis is given some credence by the detection of a somewhat extended
1—i—i—n-TT
105
-7
7. Neutralinos
1
r
!
2
3
4 5 6
2
10
3
\\
\
\
\\
\
\
-J
:
^ \ .
4 5 6
^ N. \
mill
laws with differential indices 1.95±0.25 [44]. The sensitivity of current ACTs is such that sources with spectral indices «» 2 would be easily detectable even for fluences as low as 5 x 10" 8 ergs/cm 2 [45]. Although only four of the very bright BATSE bursts were seen by EGRET, these were the brightest to occur within the field of view of EGRET and there is nothing to suggest that all bursts might not have GeV-TeV components. In fact, EGRET was not a very sensitive detector for GRBs both because of its limited collection area and its deadtime. There are now several models that suggest that TeV emission may be a strong feature of GRBs [47], [48]. There are however several negative factors concerning the possible detection of GRBs by ACTs. The narrow field of view combined with the low duty-cycle (clear, dark nights) lessens the chance of the serendipitous detection of the TeV component of a GRB. If the GRBs are truly cosmological (as they appear to be), then intergalactic absorption by pair production on infra-red photons must come into play at some point, steepening the apparent spectra. However the next generation of ACTs will have reduced energy thresholds, better flux sensitivities and rapid slew capabilities; these features combined with the more accurate source locations anticipated with the launch of HETE-2 may provide TeV detections at the rate of a few per year. In addition, the water Cherenkov detector, MILAGRO, will have all sky coverage (although reduced sensitivity below 1 TeV) and will have guaranteed coverage of some bursts detected by satellites.
205
\;
-
2
100
3
X, ,
1 III
4 5 6
1000
Photon energy [GeV]
Fig. 11. Total y-ray spectrum from neutralino annihilation for two different hypothetical neutralino masses. The spectrum shows narrow lines from direct annihilation to y-rays as well as the broad continuum emission resulting from other annihilation channels (from [51]). Physica Scripta T85
206
T. C. Weekes
20
15
10
5 0 335 Galactic Uvngiiutk
350
345
S
340
g
M5
350
345
340
Galactic Longitude
Fig. 13. Neutralino annihilation rate from the Galactic center ([51]). Shown are the sensitivity curves for Whipple, Granite III, and VERITAS. Each point represents a particular choice of supersymmetry model parameters in the experimentally allowed range.
J^ •o
1U
v
I
'
l
l
8 a.
-*C-12
10
i
•-.•
• • . • • « ; . = :•;!
:
i
1
i
l
l
= _ ~ "
.*' .1 £
'.":•' "tfSSVv. •*?:;«:, s w : , . o ?:-:'^
S*
' : : • .
«;«:••.-, ?..
:«^-;-:-K-. S :
X
a
naFI
!
H';
g JS
'•
~: I?s
V en
- « ; - > : • • • ;
•;'?.•;•».:..
• • : ; . • . • .
-13
t:^!^ffv~.
S';X5.:' S ' ' ? % -^:: : \/6i : f'5 V•#:>:.>*--. ^.vS; ; 'V:;;VS;y; •~^vVii : --'••i'-\ V : ^ - ^;i/W"iteS? « K -;^A;^ :.•..-:
".••
• ' .
.
•Vi* V
-14
10
\
. • • . ; - .
/•
•:..V.-:;..
•...
.'••••J:.;.v
f»^
i "
: • • ; . . ' . > . : ; • • : '•,:
..-:--.-. ''--;-.--.."-.v. I f f •••.'• " ••,,,<<£ " i . " •'.'-• ' ' ~ .- • \ .
:
""~~™
—-^. W^l^Il . ' • : ; " • • •
'
"'""ISP
$££••:•
10
in
-15
A?~ £
E L EQGC
(3)
^ h t Souther^ Arroy"""^--..^
10
2 « O
the first order time dispersion is given by:
r".
-16
' ' i
10
;•--;'•
\ . ' . '
1
.!
:•
- - ' ' •
'.•
.-
l"
10
E 7 [GeV] Fig. 12. EGRET observations of the Galactic Center region [49]. (a) background-subtracted, smoothed Ey > lGeV count map, (b) 300 MeV 1 GeV map.
The next generation y-ray satellite experiment, GLAST will give good coverage of parameter space at energies less than 300 GeV. Hence together GLAST and VERITAS will allow a sensitive search over the full range of allowed neutralino mass. 8. Quantum gravity Some quantum gravity models predict the refractive index of light in vacuum to be dependent on the energy of the photon. This effect, originating from the polarization of space-time, causes an energy-dependance to the velocity of light. Effectively, the quantum fluctuations are on distance scales near the Planck length, (Z,p ~ 10 - 3 3 cm), (corresponding to time-scales of l / £ p , the Planck mass (~ 1019GeV)). Different models of quantum gravity give widely differing predictions for the amount of time dispersion. In one model Physica Scripta T85
where A? is the time delay relative to propagation at the velocity of light, c, I; is a model-dependent factor of order 1, E is the energy of the observed photons, EQQ is the quantum energy scale, and L is the distance from the source. In most models £QG % EP but, in recent work in the context of string theory, it can be as low as 1016GeV [52]. Recently it has been suggested that astrophysical observations of transient high energy emission from distant sources might be used to measure (or limit) the quantum gravity energy scale. Amelino-Camelia et al. [53] suggested that BATSE observations of GRBs would provide a powerful method of probing this fundamental constant if variations on time-scales of milliseconds could be measured in the MeV signal in a GRB which was measured to be at a cosmological distance. Such time-scales and distances have been measured in GRBs but so far not in the same GRB. The absence of time dispersion in flares of TeV y-rays from AGNs at known distances provides an even more sensitive measure. Biller et al. [54] have used the sub-structure observed in the 15 minute flare in Mkn 421 observed by the Whipple group on April 15, 1996 [25] to derive a lower limit on EQQ. On a time-scale of 280 seconds there is weak (2
Particle Astrophysics with High Energy Photons 9. Future prospects It is clear that to fully exploit the potential of ground-based y-ray astronomy the detection techniques must be improved. This will happen by extending the energy coverage of the technique and by increasing its flux sensitivity. Ideally one would like to do both but in practice there must be trade-offs. Reduced energy threshold can be achieved by the use of larger but cruder mirrors and this approach is currently being exploited using existing arrays of solar heliostats (STACEE [61] and CELESTE [59]). A German-Spanish project (MAGIC) to build a 17 m aperture telescope using state-of-the-art technology has also been proposed. These projects may achieve thresholds as low as 20-30 GeV where they will effectively close the current gap in the y-ray spectrum from 20-200 GeV. Ultimately this gap will be covered by GLAST, the next generation y-ray space telescope (which will use solid-state detectors) which is scheduled for launch in 2005 by an international collaboration. Extension to even higher energies can be achieved by the atmospheric Cherenkov telescopes working at large zenith angles and by particle arrays at very high mountain altitudes. An interesting telescope that will soon come on line and will complement these techniques is the MILAGRO water Cherenkov detector in New Mexico which will operate 24 hours a day with wide field of view and will have good sensitivity to y-ray bursts and transients. VERITAS, with seven 10 m telescopes arranged in a hexagonal pattern with 80 m spacing (Fig. 14), will aim for the middle ground, with its primary objective being high sensitivity in the 100 GeV to 10 TeV range. It will be located in southern Arizona and will be the logical development of the Whipple telescope. It is hoped to begin construction in 1999 and to complete the array by 2004. The German-French HESS (initially four, and eventually perhaps sixteen, 10 m class telescopes) will be built in
Namibia and the Japanese NEW CANGAROO array (with three to four telescopes in Australia) will have similar objectives. In each case the arrays will exploit the high sensitivity of the imaging ACT and the high selectivity of the array approach. The relative flux sensitivities of the present and next generation of VHE telescopes as a function of energy are shown in Fig. 15, where the sensitivities of the wide field detectors are for one year and for the ACT for 50 hours; in all cases a 5
S 499 pixel cameras
207
Q
-»
10
/—S
W
fa
.9 10
VERITAS Whipple STACI3L O;: MAGIC GLAST (1 year) EGRET (1 year) MILAGSO 0 (5a, 50 hours, >10 events) 10
1
10
10
10
10
Et (GeV) Fig. 15. Comparison of the point source sensitivity of VERITAS to Whipple Fig. 14. Layout of the seven 10 m telescopes in VERITAS. Each telescope will [56], MAGIC [57], CELESTE/STACEE [59]; [60]; [61]; HEGRA [58], GLAST [62], EGRET [63], and MILAGRO [64]. have a camera with 499 pixels. © Physica Scripta 2000
Physica Scripta T85
208
T. C. Weekes
(electrons should show cut-offs which correlate with lower energy spectra, protons would not show a simple correlation). In addition, the recent efforts [65] to unify the different classes of blazar into different manifestations of the same object type can be tested. In addition the infrared background will be probed by the detection of sources over a range of redshifts. SNRs The existing data clearly indicate that in order to resolve the contributions of the various y-ray emission mechanisms, one needs more accurate measurements over a more complete range of energies. The NGGRTs and GLAST will be a powerful combination to address these issues. The excellent angular resolution of the NGGRTs will allow detailed mapping of the emission in SNRs. The sensitivity and energy resolution, combined with observations at lower y-ray and X-ray energies help to elucidate the y-ray emission mechanism. This may lead to direct the confirmation or elimination of SNRs as the source of cosmic rays. Gamma-ray pulsars The detection of VHE y-rays would be decisive in favoring the outer gap model over the polar cap model. Six pulsars are detected at EGRET energies and their high energy emission is already seriously constrained by the VHE upper limits. The detection of a pulsed y-ray signal above 50 GeV would be a major breakthrough.
4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34.
Unidentified galactic EGRET sources The legacy of EGRET may be more than 70 unidentified sources, many of which are in the Galactic plane. The positional uncertainty of these sources make identifications with sources at longer wavelengths unlikely. In the galactic plane, probable sources are SNRs and pulsars, particularly in regions of high IR density (e.g., OB associations), but some may be new types of objects. The N G G R T should have the sensitivity and low energy threshold necessary to detect many of these objects. Detailed studies of these objects with the excellent source location capability of the NGGRTs could lead to many identifications with objects at longer wavelengths. The energy resolution and energy coverage could also go a long way to explaining the y-ray emission mechanisms in any new types of source identified. In addition, any variability in these objects would be easily identified and measured with the NGGRT.
35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50.
Acknowledgements Research in VHE y-ray astronomy at the Whipple Observatory is supported by a grant from the U.S.D.O.E. Helpful comments from Mike Catanese, Stephen Fegan and Vladimir Vassiliev are also acknowledged.
References 1. Proceedings of the Fourth Compton Symposium, Williamsburg, Virginia, USA, April 1997, (Eds. C. D. Dermer, M. S. Strickman and J. D. Kurfess) AIP 410. 2. Ong, R. A., Physics Reports 305, 93 (1998). 3. Hillas, A. M. et al, Astrophys. J. 503, 744 (1998).
Physica Scripta T85
51. 52. 53. 54. 55. 56. 57. 58. 59.
Weekes, T. C , Space Sci. Rev. 59, 315 (1991). Gould, R. J., Phys. Rev. Lett. 15, 577 (1965). Kifune, T. et al, Astrophys. J. 438, L91 (1995). Chadwick, P. M. et al, Proc. 25th Int. Cos. Ray Conf., Durban 3, 189 (1997). Yoshikoshi, T. et al., Astrophys. J. 487, L65 (1997). Chadwick, P. M. et al., Astrophys. J. 503, 391 (1998). Tanimori, T. et al., Astrophys. J. 497, L25 (1998). Esposito, J. A. et al, Astrophys. J. 461, 820 (1996). Drury, L. O ' C , Aharonian, F. A. and Volk, H. J. Astron. Astrophys. 287, 959 (1994). Buckley, J. H. et al, Astron. Astrophys. 329, 639 (1998). Gaisser, T. K., Protheroe, R. J. and Stanev, T., Astrophys. J. 492, 219 (1998). Buckley, J. H. et al, Astrophys. J. 472, L9 (1996). Punch, M. et al, 358, 477 (1992). Quinn, J. et al, Astrophys. J. Lett. 456, L83 (1996). Catanese, M. et al, Astrophys. J. 501, 616 (1998). Chadwick, P. M. et al, Astropart. Phys. 9, 131 (1998). Neshpor, Yu. I. et al, Astron. Lett. 24, 134 (1998). Perlman, E. S. et al, Astrophys. J. Suppl. 104, 251 (1996). Mukherjee, R. et al, 490, 116 (1997). Kataoka, J. et al, Astrophys. J. 514, 138 (1999). Zweerink, J. et al, Astrophys. J. Lett. 490, L141 (1997). Gaidos, J. A. et al, Nature 383, 319 (1996). Quinn, J. et al, Astrophys. J. 518, 693 (1999). Krennrich, F. et al, Astrophys. J. 481, 758 (1997). Lorenz, E., Proc. of Workshop on TeV Astrophysics of Extragalactic Sources, Cambridge, MA, 131 (1998). Samuelson, F. W. et al, Astrophys. J. Lett. 501, L17 (1998). Krennrich, F. et al, Astrophys. J. 511, 149 (1999). Macomb, D. J. et al, Astrophys. J. 449, L99 (1995). Maraschi, L. et al, Proc. of Workshop on TeV Astrophysics of Extragalactic Sources, Cambridge, MA, 189 (1998). Catanese, M. et al, Astrophys. J. 487, L143 (1997). Coppi, P. S., Kartje, J. F. and Konigl, A. 1993, in Proc. Compton Symposium, (Eds. M. Friedlander, N. Gehrels and D. J. Macomb) (New York: AIP), 559. Catanese, M., Proc. of Symposium on BL Lac Phenomenon, Turku, Finland, 243 (1998). Tavecchio, F., Maraschi, L. and Ghiselline, G. Astrophys. J. 509, 608 (1998). Mannheim, K., Astron. Astrophys. 269, 67 (1993). Nikishov, A. J., Soviet Physics J.E.T.P. 14, 393 (1962). Gould, R. P. and Schreder, G. P. Phys. Rev. 155, 1408 (1967). Stecker, F. W. and De Jager, O. C. Astrophys. J. 415, L71 (1993). Biller, S. D. et al, Phys. Rev. Lett. 80, 2992 (1998a). MacMinn, D. and Primack, J. R., Space Sc. Rev. 75, 413 (1996). Padilla, L. et al, Astron. Astrophys. 337, 43 (1998). Dingus, B. et al, Proc. 4th Huntsville GRB Symposium, (AIP 428), (Eds. C. A. Meegan, R. D. Preece, T. M. Koshut) 349 (1998). Connaughton, V. et al, Astrophys. J. 479, 859 (1997). Hurley, K. et al, Nature 372, 652 (1994). Totani, T., Astrophys. J. 502, L13 (1998). Dermer, C. D., Chiang, J., Bottcher, M., Astrophys. J. (in press) (1998). Mayer-Hasselwander, H. A. et al, Astron. Astrophys. 335, 161 (1998). Buckley, J. H. et al, Proc. 25th Int. Cos. Ray Conf., Durban 3, 237 (1997). Bergstrbm, L., Ullio, P. and Buckley, J. H., Astropart. Phys. 9, 137 (1998). Witten, E., Nucl. Phys. B 471, 135 (1996). Amelino-Camelia, G. et al, Nature 383, 319 (1998). Biller, S. D. et al, Phys. Rev. Lett. 83, 2108 (1999). Vassiliev, V. V., Proc. of Workshop on TeV Astrophysics of Extragalactic Sources, Cambridge, MA, (in press). Weekes, T. C. et al, Astrophys. J. 342, 370 (1989). Barrio, J. A. et al, "The Magic Telescope", design study (MPI-PhE/98-5) (1998). Daum, A. et al, Astropart. Phys. 8, 1 (1997). Quebert, J. et al, in "Towards a Major Atmospheric Cherenkov Detector IV", (Padova, Italy), (Ed. M. Cresti), 248 (1995).
© Physica Scripta 2000
Particle Astrophysics with High Energy Photons 60. Bhat, C. L., Proc. 25th Int. Cos. Ray Conf., Durban 8, 211 (1997). 61. Ong, R. A. and Covault, C. E., in "Towards a Major Atmospheric Cherenkov Detector IV", (Padova, Italy), (Ed. M. Cresti), 247 (1998). 62. Leonard, P. J. T., Nature, 383, 394 (1996).
© Physica Scripta 2000
209
63. Kurfess, J. D. et al., Proceedings of the Fourth Compton Symposium, Williamsburg, Virginia, USA, April 1997, (Eds. C. D. Dermer, M. S. Strickman and J. D. Kurfess), (AIP 410), 509. 64. Sinnis, G. et al, Nucl. Phys. B (Proc. Suppl.) 43, 141 (1995). 65. Ghisellini, G. et al., Mon. Notes R. Astron. 301, 451 (1998).
Physica Scripta T85
Physica Scripta.Vol. T85, 210-220, 2000
Dark Matter and Dark Energy in the Universe Michael S. Turner Departments of Astronomy & Astrophysics and of Physics, Enrico Fermi Institute, The University of Chicago, Chicago, IL 60637-1433, USA NASA/Fermilab Astrophysics Center, Fermi National Accelerator Laboratory, Batavia, IL 60510-0500, USA Received January 10, 1999; accepted August 2, 1999
PACS Ref: 95.35.+d
Abstract
when the expansion rate,
For the first time, we have a plausible and complete accounting of matter and energy in the Universe. Expressed a fraction of the critical density it goes like this: neutrinos, between 0.3% and 15%; stars, between 0.3% and 0.6%; baryons (total), 5% ± 0 . 5 % ; matter (total), 40% ± 10%; smooth, dark energy, 80% ± 2 0 % ; totaling to the critical density (within the errors). This accounting is consistent with the inflationary prediction of a flat Universe and measurements of the anisotropy of the CBR. It also defines three "dark problems": Where are the dark baryons? What is the nonbaryonic dark matter? What is the nature of the dark energy? The leading candidate for the (optically) dark baryons is diffuse hot gas; the leading candidates for the nonbaryonic dark matter are slowly moving elementary particles left over from the earliest moments (cold dark matter), such as axions or neutralinos; the leading candidates for the dark energy involve fundamental physics and include a cosmological constant (vacuum energy), a rolling scalar field (quintessence), and a network of light, frustrated topological defects.
/r2 = ^ J >
1. Introduction The quantity and composition of matter and energy in the Universe is of fundamental importance in cosmology. The fraction of the critical energy density contributed by all forms of matter and energy, Qo =
Plot
£«,,
(1)
determines the geometry of the Universe: R2 curv
H 2
Q0
o -
1
(2)
Here, subscript '0' denotes the value at the present epoch, p crit = 3H$/SnG ~ 1.88/r2 x 10" 29 g cm" 3 , O, is the fraction of critical density contributed by component i (e.g., baryons, photons, stars, etc) and Ho = 100/ikms" 1 M p c - 1 . The sign of i? 2 urv specifies the spatial geometry: positive for a 3-sphere, negative for a 3-saddle and 0 for flat space. The present rate of deceleration of the expansion depends upon Oo as well as the composition of matter and energy,
1
(4)
R2 xurv
vanishes and R
4nG
(5)
J2 P/[l + wt]
is less than zero. If there were matter alone, a positively curved universe (Q0 > 1) eventually recollapses and a negatively curved universe (Qo < 1) expands forever. However, exotic components complicate matters: a positively curved universe with positive vacuum energy can expand forever, and a negatively curved universe with negative vacuum energy can recollapse. The quantity and composition of matter and energy in the Universe is also crucial for understanding the past. It determines the relationship between age of the Universe and redshift, when the Universe ended its early radiation dominated era, and the growth of small inhomogeneities in the matter. Ultimately, the formation and evolution of large-scale structure and even individual galaxies depends upon the composition of the dark matter and energy. Measuring the quantity and composition of matter and energy in the Universe is a challenging task. Not just because the scale of inhomogeneity is so large, around 10 Mpc; but also, because there may be components that remain exactly or relatively smooth (e.g., vacuum energy or relativistic particles) and only reveal their presence by their influence on the evolution of the Universe itself.
2. A complete inventory of matter and energy
2.1. Preliminaries 2.1.1. Radiation Because the cosmic background radiation (CBR) is known to be black-body radiation to very high precision (better than 0.005%) and its temperature is known to four signifi1 _ 0 WR\ = ^ o + - £ > , • Wi. (3) cant figures, T0 = 2.7277 ± 0.002 K, its contribution is very qo=2" Hi precisely known, Qyh2 = 2.480 x 10~5. If neutrinos are _4 The pressure of component /, pt = w,-p,-; e.g., for baryons massless or very light, mv <£ 10 eV, their energy density wt = 0, for radiation H>, = 1/3, and for vacuum energy is equally well known because it is directly related to that of the photons, Qv = | ( 4 / l l ) 4 / 3 O y (per species) (there is a Wi = — 1 . The fate of the Universe - expansion forever or recollapse small 1% positive correction to this number; see [1]). It is possible that additional relativistic species contribute - is not directly determined by Qo and/or qo. It depends upon precise knowledge of the composition of all components of significantly to the energy density, though big-bang matter and energy, for all times in the future. Recollapse nucleosynthesis (BBN) severely constrains the amount (the occurs only if there is a future turning point, that is an epoch equivalent of less than 0.2 of a neutrino species; see e.g., [2]) Physica Scripta T85
i Physica Scripta 2000
Dark Matter and Energy in the Universe
211
MATTER / ENERGY in the UNIVERSE
unless they were produced by the decay of a massive particle after the epoch of BBN. In any case, we can be confident that the radiation component of the energy density today is small. The matter contribution (denoted by £2M), consisting of particles that have negligible pressure, is the easiest to determine because matter clumps and its gravitational effects are thereby enhanced (e.g., in rich clusters the matter density typically exceeds the mean density by a factor of 1000 or more). With this in mind, I will decompose the present matter/energy density into two components, matter and vacuum energy,
TOTAL
MATTER COMPOSITION
J I I _ _
CDH,
"*
OAR* ENERSV I 0.6-/-0.2 J. WAITER y
I 0.35-/-O1
1
T < 0.15
Oo =
QM
+
&A
.
(6)
ignoring the contribution of the CBR and ultrarelativistic neutrinos. I will use vacuum energy as a stand in for a smooth component of energy (more later). Vacuum energy and a cosmological constant are indistinguishable: a cosmological constant corresponds to a uniform energy density of magnitude p vac = A/inG.
f
0.1 8ARWMS
j|
0.05./-0.005
NEUTRINOS
0.01
2.2. Q0 = 1 ± 0 . 2 There is a growing consensus that the anisotropy of the CBR offers the best means of determining the curvature of the Universe and thereby Qo, cf., Eq. (2). This is because the method is intrinsically geometric - a standard ruler on the last-scattering surface - and involves straightforward physics at a simpler time (see e.g., [3]). At last scattering baryonic matter (ions and electrons) was still tightly coupled to photons; as the baryons fell into the dark-matter potential wells the pressure of photons acted as a restoring force, and gravity-driven acoustic oscillations resulted. These oscillations can be decomposed into their Fourier modes; Fourier modes with k ~ lH$/2 determine the multipole amplitudes a/m of CBR anisotropy. Last scattering occurs over a short time, making the CBR is a snapshot of the Universe at t\s ~ 300 000 yrs. Each mode is "seen" in a well defined phase of its oscillation. (For the density perturbations predicted by inflation, all modes the have same initial phase because all are growing-mode perturbations.) Modes caught at maximum compression or rarefaction lead to the largest temperature anisotropy; this results in a series of acoustic peaks beginning at / ~ 200 (see Fig. 2). The wavelength of the lowest frequency acoustic mode that has reached maximum compression, ^max ~ vst\s, is the standard ruler on the last-scattering surface. Both l m a x and the distance to the last-scattering surface depend upon Q0, and the position of the first peak / ~ 200/V^o- This relationship is insensitive to the composition of matter and energy in the Universe (see [4]). CBR anisotropy measurements, shown in Figs. 2 and 3, now cover three orders of magnitude in multipole number and involve more than twenty experiments. COBE is the most precise and covers multipoles / = 2 — 20; the other measurements come from balloon-borne, Antarctica-based and ground-based experiments using both low-frequency (/~<100GHz) HEMT receivers and high-frequency (f > 100 GHz) bolometers. Taken together, all the measurements are beginning to define the position of the first acoustic peak, at a value that is consistent with a flat Universe. Various analyses of the extant data have been carried out, indicating QQ ~ 1 ± 0.2 (see e.g., [5]). It is certainly too early to draw definite conclusions or put too much weigh © Physica Scripta 2000
STARS T 00O5-/-O.0O2
0.001 —
—
Q Fig. 1. Summary of matter/energy in the Universe. The right side refers to an overall accounting of matter and energy; the left refers to the composition of the matter component. The contribution of relativistic particles, CBR photons and neutrinos, QTe\h2 = 4.170 x 10~5, is not shown. The upper limit to mass density contributed by neutrinos is based upon the failure of the hot dark matter model of structure formation [26,64] and the lower limit follows from the evidence for neutrino oscillations [69]. Here Ho is taken to be 6 5 k m s - 1 Mpc - 1 .
in the error estimate. However, a strong case is developing for a flat Universe and more data is on the way (Python V, Viper, TOCO, Maxima, Boomerang, DASI, and others). Ultimately, the issue will be settled by NASA's MAP (launch late 2000) and ESA's Planck (launch 2007) satellites which will map the entire CBR sky with 30 times the resolution of COBE (around 0.1°) (see [6]). 2.3. Matter 2.3.1. Baryons. For more than twenty years big-bang nucleosynthesis has provided a key test of the hot big-bang cosmology as well as the most precise determination of the baryon density. In 1995 comparison of the primeval abundances of D, 3 He, 4 He and 7 Li with their big-bang predictions defined a concordance interval, Q^h2 = 0.007 - 0.024 (see e.g., [7]; for another view, see [8]). Of the four light elements produced in the big bang, deuterium is the most powerful "baryometer" - its primeval abundance depends strongly on the baryon density (oc 1/pg7) - and the evolution of its abundance since the big bang is simple - astrophysical processes only destroy deuterium. Until recently deuterium could not be exploited Physica Scripta T85
212
Michael S. Turner 120
100 « i
-
o
+
1
10
M Tegmark, April 1998
100
1000
1
as a baryometer because its abundance was only known locally, where roughly half of the material has been through stars with a similar amount of the primordial deuterium destroyed. In 1998, the situation changed dramatically, launching BBN into the precision era of cosmology. Over the past four years there have been claims of upper limits, lower limits, and determinations of the primeval deuterium abundance, ranging from ( D / H ) = 1 0 -55 to ( D / H ) = 3 x 10 4 . In short, the situation was confusing. In 1998 Buries and Tytler clarified matters and established a strong case for (D/H)P = (3.4 ± 0.3) x 10~5 [9]. That case is based upon the deuterium abundance measured in four high-redshift hydrogen clouds seen in absorption against distant QSOs, and the remeasurement and reanalysis of other putative deuterium systems. In this enterprise, the Keck I 10-meter telescope and its HiRes Echelle Spectrograph have played the crucial role. The Buries - Tytler measurement turns the previous factor of three concordance range for the baryon density into a 10% determination of the baryon density, p B = (3.8 ±0.4)x 10"31 g e m - 3 or QBh2 = 0.02 ± 0.002 (see Fig. 4), with about half the error in p B coming from the theoretical error in predicting the BBN yield of deuterium. [A very recent analy-
Fig. 2. Summary of all CBR anisotropy measurements, where the temperature variations across the sky have been expanded in spherical harmonics, 5T(6, (j>) = ]T\ atm YimiO, <j>) and Q = (|a/m|2). In plain language, this plot shows the size of the temperature variation between two points on the sky separated by angle 8 vs. multipole number / = 200°/6 (1 = 2 corresponds to 100°, / = 200 corresponds to 0 = 1°, and so on). The curves illustrate the predictions of CDM models with fio = 1 (curve with lower peak) and J2o = 0.3 (darker curve). Note: the preference of the data for aflatUniverse and the evidence for the first of a series of "acoustic peaks." The presence of these acoustic peaks is a key signature of the density perturbations of quantum origin predicted by inflation (Figure courtesy of M. Tegmark).
B o n d . Jaffe a n d Knox
0.005 -1
1
0.01 1
1—I—I—!—I
0.02
0.03
I I I I I i : i I : 111 i i i u | i i l l i u i i
1996
100 S? 80 a.
k_l_
10 -10
,1,
2xl0~10
5x10 -10
I
I
10"
V 4
Fig. 4. Predicted abundances of He (mass fraction), D, 3He, and 7Li (number relative to hydrogen) as a function of the baryon density; widths of the curves indicate "2a" theoretical uncertainty. The dark band highlights Fig. 3. The same data as in Fig. 2, but averaged and binned to reduce error the determination of the baryon density based upon the recent measurement 2 bars and visual confusion. The theoretical curve is for the ACDM model of the primordial abundance of deuterium [2,9] QBh = 0.019 ± 0.0024 (95% _1 -1 with Ho = 65kms Mpc and QM = 0.4; note the goodness of fit (Figure cl); the baryon 22density is related to the baryon-to-photon ratio by pB = 6.88f7 x 10" (from [2]). courtesy of L. Knox). Physica Scripta T85
© Physica Scripta 2000
Dark Matter and Energy in the Universe sis has reduced the theoretical error significantly, and improved the accuracy of the determination of the baryon density, QBh2 = 0.019 ± 0.0012 [2].] This precise determination of the baryon density, made at a simpler time and based upon the early Universe physics of BBN, is consistent with two other measures of the baryon density, based upon entirely different physics. By comparing measurements of the opacity of the Lyman-a forest toward high-redshift quasars with high-resolution, hydrodynamical simulations of structure formation, several groups [10-12] have inferred a lower limit to the baryon density, Q^h2 > 0.015. The second test involves the height of the first acoustic peak: it rises with the baryon density (the higher the baryon density, the stronger the gravitational force driving the acoustic oscillations). Current CBR measurements are consistent with the Buries - Tytler baryon density; the MAP and Planck satellites should ultimately provide a 5% or better determination of the baryon density, based upon the physics of gravity-driven acoustic oscillations when the Universe was 300,000 yrs old (see e.g., [4]). This will be an important cross check of the BBN determination. 2.3.2. Weighing the dark matter: Qu = 0.4 ± 0.1. Since the pioneering work of Fritz Zwicky and Vera Rubin, it has been known that there is far too little material in the form of stars (and related material) to hold galaxies and clusters together, and thus, that most of the matter in the Universe is dark (see e.g. [13]). Weighing the dark matter has been the challenge. At present, I believe that clusters provide the most reliable means of estimating the total matter density. Rich clusters are relatively rare objects - only about 1 in 10 galaxies is found in a rich cluster - which formed from density perturbations of (comoving) size around lOMpc. However, because they gather together material from such a large region of space, they can provide a "fair sample" of matter in the Universe. Using clusters as such, the precise BBN baryon density can be used to infer the total matter density [14]. (Baryons and dark matter need not be well mixed for this method to work provided that the baryonic and total mass are determined over a large enough portion of the cluster.) Most of the baryons in clusters reside in the hot, X-ray emitting intracluster gas and not in the galaxies themselves, and so the problem essentially reduces to determining the gas-to-total mass ratio. The gas mass can be determined by two methods: (1) measuring the x-ray flux from the intracluster gas and (2) mapping the Sunyaev - Zel'dovich CBR distortion caused by CBR photons scattering off hot electrons in the intracluster gas. The total cluster mass can be determined three independent ways: (1) using the motions of clusters galaxies and the virial theorem; (2) assuming that the gas is in hydrostatic equilibrium and using it to infer the underlying mass distribution; and (3) mapping the cluster mass directly by gravitational lensing [15]. Within their uncertainties, and where comparisons can be made, the three methods for determining the total mass © Physica Scripta 2000
213
agree (see e.g., [15]); likewise, the two methods for determining the gas mass are consistent. Mohr et al. [16] have compiled the gas to total mass ratios determined from X-ray measurements for a sample of 45 clusters; they find / g a s = (0.075 ± 0.002)/r 3/2 . Carlstrom [17], using his S-Z gas measurements and X-ray measurements for the total mass for 27 clusters, finds / g a s = (0.06 ± 0.006) h _ 1 . (The agreement of these two numbers means that clumping of the gas, which could lead to an overestimate of the gas fraction based upon the X-ray flux, is not a problem.) Invoking the "fair-sample assumption," the mean matter density in the Universe can be inferred: GM = i2B//gas = (0.3 ± 0.05) h" 1 ' 2 (Xray) = (0.25 ± 0.04) h" 1 (S - Z)
(7)
= 0.4 ± 0.1 (my summary). I believe this to be the most reliable and precise determination of the matter density. It involves few assumptions, and most of them have now been tested (clumping, hydrostatic equilibrium, variation of gas fraction with cluster mass). There is one implicit assumption: that all the cluster baryons are in stars or hot gas. For this reason some authors quote the value from this method as an upper limit to QM. 2.3.3. Supporting evidence for QM = 0.4 ± 0.1. This result is consistent with a variety of other methods that involve different physics. For example, based upon the evolution of the abundance of rich clusters with redshift, Henry [18] finds Qu = 0.45 ± 0.1 (also see, [19,20]). Dekel and Rees [21] place a low limit Qu > 0.3 (95% cl) derived from the outflow of material from voids (a void effectively acts as a negative mass proportional to the mean matter density). The analysis of the peculiar velocities of galaxies provides an important probe of the mass density averaged over very large scales (of order several hundred Mpc). By comparing measured peculiar velocities with those predicted from the distribution of matter revealed by redshift surveys such as the IRAS survey of infrared galaxies, one can infer the quantity /? = QM6/°i where b\ is the linear bias factor that relates the inhomogeneity in the distribution of IRAS galaxies to that in the distribution of matter (in general, the bias factor is expected to be in the range 0.7 to 1.5; IRAS galaxies are expected to be less biased, b\ « 1.). Recent work by Willick and Strauss [22] finds )S = 0.5 ± 0.05, while Sigad et al. [23] find /? = 0 . 9 ± 0 . 1 . The apparent inconsistency of these two results and the ambiguity introduced by bias preclude a definitive determination of Qu by this method. However, Dekel [24] quotes a 95% confidence lower bound, Qu > 0.3, and the work of Willick & Strauss seems to strongly indicate that Qu is much less than 1; both are consistent with Qu ~ 0.4. Finally, there is strong, but circumstantial, evidence from structure formation that Qu is around 0.4 and significantly greater than QB. It is based upon two different lines of reasoning. First, there is no viable model for structure formation without a significant amount of nonbaryonic dark matter. The underlying reason is simple: in a baryons-only model, density perturbations grow only from the time of decoupling, z ~ 1000, until the Universe becomes curvature Physica Scripta T85
214
Michael S. Turner
dominated, z ~ Q^1 ~ 20; this is not enough growth to produce all the structure seen today with the size of density perturbations inferred from CBR anisotropy. With nonbaryonic dark matter and QM 3> ®B, dark-matter perturbations begin growing at matter - radiation equality and continue to grow until the present epoch, or nearly so, leading to significantly more growth and making the observed large-scale structure consistent with the size of the density perturbations inferred from CBR anisotropy. Second, the transition from radiation domination at early times to matter domination at around 10,000 yrs leaves its mark on the shape of the present power spectrum of density perturbations, and the redshift of matter - radiation equality depends upon Qu- Measurements of the shape of the present power spectrum based upon redshift surveys indicate that the shape parameter r = Quh ~ 0.25 ± 0.05 (see e.g., [25]). For h ~ 2/3, this implies QM ~ 0.4. (If there are relativistic particles beyond the CBR photons and relic neutrinos, the formula for the shape parameter changes and Qu ^> 0.4 can be accommodated; see [26]).
end of star formation in the Universe: 80% of star formation took place at a redshift greater than unity; see Fig. 6.) While the value for Qu derived from the cluster baryon fraction also relies upon clusters, the underlying assumption is far weaker and much more justified, namely that clusters provide a fair sample of matter in the Universe. Even if mass-to-light ratios were measured in the red (they typically are not), where the light is dominated by low-mass
2.4. Mass-to-light ratios and Qu-amazingly, the glass is half full! The most mature approach to estimating the matter density involves the use of mass-to-light ratios, the measured luminosity density, and the deceptively simple equation, (pM) = (M/L)C,
(8) 8
where L = 2 Ah x 10 LBQ Mpc~ is the measured (B-band) luminosity density of the Universe. Once the average mass-to-light ratio for the Universe is determined, Qu follows by dividing it by the critical mass-to-light ratio, (M/L) crit = 1200 h (in solar units). Though it is tantalizingly simple this method does not provide an easy and reliable method of determining QM- While computing a massto-light ratio is easy, actually determining the average mass-to-light ratio is probably as hard as determining the average mass density. Based upon the mass-to-light ratios of the bright, inner regions of galaxies, (M/L)t ~ few, the fraction of critical density in stars (and closely related material) has been determined, fl* ~ (0.003 ± 0 . 0 0 1 ) / r ' (see e.g., [27]). Persic and Salucci [28] derive a similar value based upon the observed stellar-mass function. Luminous matter accounts for only a tiny fraction of the total mass density and only about a tenth of the baryons. CNOC [29,30] have done a very careful job of determining a mean cluster mass-to-light ratio, (M/L) cluster = 240 ± 50, which translates to an estimate of the mean matter density, ^cluster = 0.20 ± 0.04. Because clusters contain thousands of galaxies and cluster galaxies do not seem radically different from field galaxies, one is tempted to take this estimate of the mean matter density seriously. However, it is significantly smaller than the value I advocated earlier, Qu = 0.4 ± 0 . 1 . Which estimate is right? I believe the higher number, based upon the cluster baryon fraction, is more reliable and that we should be surprised that the CNOC number is so close, closer than we had any right to expect! After all, only a small fraction of galaxies are found in clusters and the luminosity density £ itself evolves strongly with redshift and corrections for this effect are large and uncertain. (We are on the tail Physica Scripta T85
4
3
6
8
10
Tx [keV] Fig. 5. Cluster gas fraction as a function of cluster gas temperature for a sample of 45 galaxy clusters [16]. While there is some indication that the gas fraction decreases with temperature for T < 5 keV, perhaps because these lower-mass clusters lose some of their hot gas, the data indicate that the gas fraction reaches a plateau at high temperatures, _/^as =0.212 ±0.006 for h = 0.5 (Figure courtesy of Joe Mohr).
13
B u o
.5
it, W
u (0
V)
u
.25
3
2
J
0 0
I
L
J
I
I
L
5 10 Time (Gyr)
Fig. 6. Star formation history as a function of cosmic time (adapted from [58]). The two curves bracket the estimated uncertainty and the ordinate is given by the star formation rate times cosmic time. Note that today we are on the tail end of star formation. © Physica Scripta 2000
Dark Matter and Energy in the Universe 215 stars and reflects the integrated history of star formation rather than the current star-formation rate as it does in the blue, one would still require the fraction of baryons converted into stars in clusters to be identical to that in the field to have agreement between the CNOC estimate and that based upon the cluster baryon fraction. Apparently, the fraction of baryons converted into stars in the field and in clusters is similar, but not identical. To put this in perspective and to emphasize the shortcomings of the mass-to-light technique, had one used the cluster mass-to-x-ray ratio and the x-ray luminosity density, one would have inferred fiM ~ 0-05. A factor of two discrepancy based upon optical mass-to-light ratios does not seem so bad. Enough said.
1
—
5
-
a
\
\
\
^ \
\
\ \
C
s \ -1/3\
-
0) O
w=-l
9
\ radiation\
matter
1
1
-3
2.5. Missing energy?
'9-
© Physica Scripta 2000
1/3
\
"••
CD
0
The results Oo = 1 ± 0.2 and QM = 0.4 ± 0.1 are in apparent contradiction, suggesting that one or both are wrong. However, prompted by a strong belief in a flat Universe, theorists have explored the remaining logical possibility: a dark, exotic form of energy that is smoothly distributed and contributes 60% of the critical density [31,32]. Being smoothly distributed its presence would not have been detected in measurements of the matter density. The properties of this missing energy are severely constrained by other cosmological facts, including structure formation, the age of the Universe, and CBR anisotropy. So much so, that a smoking-gun signature for the missing energy was predicted: accelerated expansion [32]. To begin, let me parameterize the bulk equation of state of this unknown component: w = px/px. This implies that its energy density evolves as px oc R~" where n = 3(1 + w). The development of the structure observed today from density perturbations of the size inferred from measurements of the anisotropy of the CBR requires that the Universe be matter dominated from the epoch of matter - radiation equality until very recently. Thus, to avoid interfering with structure formation, the dark-energy component must be less important in the past than it is today. This implies that n must be less than 3 or w < 0; the more negative w is, the faster this component gets out of the way (see Fig. 7). More careful consideration of the growth of structure implies that w must be less than about —5 [33]. Next, consider the constraint provided by the age of the Universe and the Hubble constant. Their product, Hoto, depends the equation of state of the Universe; in particular, Hoto increases with decreasing w (see Fig. 8). To be definite, I will take ?0 = 14±1.5Gyr and i/ 0 = 65 ± 5 km s~' M p c - 1 (see e.g., [34,35]); this implies that H0t0 = 0.93± 0.13. Fig. 8 shows that w < — \ is preferred by age/Hubble constant considerations. To summarize, consistency between QM ~ 0.4 and O0 ~ 1 along with other cosmological considerations implies the existence of a dark-energy component with bulk pressure more negative than about —px/2. The simplest example of such is vacuum energy (Einstein's cosmological constant), for which w = — 1. The smoking-gun signature of a smooth, dark-energy component is accelerated expansion since qo = 0.5 + \.5wQx - 0.5 + 0.9w < 0 for w < 5
1
1 0
-2 -1 log (scale factor)
Fig. 7. Evolution of the energy density in matter, radiation (heavy lines), and different possibilities for the dark-energy component (w = — 1, — 1, i) vs. scale factor. The matter-dominated era begins when the scale factor was ~ 10~4 of its present size (off the figure) and ends when the dark-energy component begins to dominate, which depends upon the value of w: the more negative w is, the longer the matter-dominated era in which density perturbations can go into the large-scale structure seen today. These considerations require w<-i[33]. 1
!
,
|
1 1
1
1
i
1 i
1
.1 Hot,, = 1.06
1 .9 X
—
.8
= 0.8
.7 .6
—
7 °«== 1
!
0.4 1 L .1.. .1 i
1 1 1 1
1Is
0
-.5
w Fig 8. Hoto vs. the equation of state for the dark-energy component. As can be seen, an added benefit of a component with negative pressure is an older Universe for a given Hubble constant. The broken horizontal lines denote the l
2.6. Missing energy found! In 1998 evidence for the smoking gun signature of dark was found. It came in the form of the magnitude - redshift (Hubble) diagram for fifty-some type l a supernovae (SNe la) with redshifts out to nearly 1. Two groups, the Supernova Cosmology Project [36,37] and the High-z Supernova Search Team [38,39], working independently and using different Physica Scripta T85
216
Michael S. Turner
methods of analysis, each found evidence for accelerated expansion. Perlmutter et al. [36] summarize their results as a constraint to a cosmological constant (see Fig. 9), 4„
1,1
(9)
For £ 2 M ~ 0 . 4 ± 0 . 1 , this implies QA = 0.85 ± 0.2, or just what is needed to account for the missing energy! (A simple explanation of the SN la results may be useful. If galactic distances and velocities were measured today they would obey a perfect Hubble law, o0 = Hod, because the expansion of the Universe is simply a rescaling of all distances. Because we see distant galaxies at an earlier time, their velocities should be larger than predicted by the Hubble law, provided the expansion is slowing due to the attractive force of gravity. Using SNe la as standard candles to determine the distances to faraway galaxies, the two groups in effect found the opposite: distant galaxies are moving slower than predicted by the Hubble law, implying the expansion is speeding up!) Recently, two other studies, one based upon the x-ray properties of rich clusters of galaxies [40] and the other based upon the properties of double-lobe radio galaxies [41], have reported evidence for a cosmological constant (or similar dark-energy component) that is consistent with the SN la results (i.e., QA ~ 0.7). There is another powerful test of an accelerating Universe whose results are more ambiguous. It is based upon the fact that the frequency of multiply lensed QSOs is expected to be significantly higher in an accelerating universe [42]. Kochanek [43] has used gravitational lensing of QSOs to place a 95% cl upper limit, QA < 0.66; and Waga and Miceli [44] have generalized it to a dark-energy component with negative pressure: Qx < 1.3 + 0.55w (95% cl), both results for a flat Universe. On the other hand, Chiba and Yoshii
[45] claim evidence for a cosmological constant, QA = 0.7+^, based upon the same data. From this I conclude: (1) lensing excludes QA larger than 0.8, and (2) when larger objective surveys of gravitational-lensed QSOs are carried out (e.g., the Sloan Digital Sky Survey), there is the possibility of uncovering another smoking-gun for accelerated expansion. By far, the strongest evidence for dark energy is the SN l a
data. The statistical errors reported by the two groups are smaller than possible systematic errors. Thus, the behevability of the results turns on the reliability of SNe la as one-parameter standard candles. SNe la are thought to be associated with the nuclear detonation of Chandrasekhar-mass white dwarfs. The one parameter is the rate of decline of the light curve: The brighter ones decline more slowly (the so-called Phillips relation; see [46]). The lack of a good theoretical understanding of this (e.g., what is the physical parameter?) is offset by strong empirical evidence for the relationship between peak brightness and rate of decline, based upon a sample of nearby SNe la. It is reassuring that in all respects studied, the distant sample of SNe la appear to be similar to the nearby sample. For example, distribution of decline rates and dispersion about the Phillips relationship. The local sample spans a range of metallicity, likely spanning that of the distant sample, and further, suggesting that metallicity is not an important second parameter. At this point, it is fair to say that if there is a problem with SNe la as standard candles, it must be subtle. Cosmologists are even more inclined to believe the SN la results because of the preexisting evidence for a "missing-energy component" that led to the prediction of accelerated expansion.
2.7. Cosmic concordance
c.u J
ill.
/
M M 111l_
/ SNela
1.5 1
— 1 1 1 1 1
2
1 111111III "1 Matter^/ /
-\
.5
/
—
(
i
With the SN la results we have for the first time a complete and self-consistent accounting of mass and energy in the Universe (see Fig. 1). The consistency of the matter/ energy accounting is illustrated in Fig. 9. Let me explain this exciting figure. The SN la results are sensitive to the acceleration (or deceleration) of the expansion and constrain the combination | O M - QA- (Note, qo = \QM - &A, 5&M - QA corresponds to the deceleration parameter at redshift z ~ 0.4, the median redshift of these samples). The (approximately) orthogonal combination, QQ — Qu+ &A is constrained by CBR anisotropy. Together, they define a concordance region around Qo ~ 1, QM ~ 1/3, and QA ~ 2/3. The constraint to the matter density alone, Qu — 0.4 ± 0 . 1 , provides a cross check, and it is consistent with these numbers. Cosmic concordance!
But there is more. We also have a consistent and well motivated picture for the formation of structure in the Universe, ACDM. The ACDM model, which is the cold dark n 7\ 11 N 1 1 1 1 1 1 INI 1 1 1 II 1 II 1 f matter model with QB ~ 0.05, QCDM ~ 0-35 and QA ~ 0.6, is 2.5 a very good fit to all cosmological constraints: large-scale 0 .5 1.5 structure, CBR anisotropy, age of the Universe, Hubble conM stant and the constraints to the matter density and Fig. 9. Two-
\CBR Anisotropy
-
n
Physica Scripta T85
© Physica Scripta 2000
Dark Matter and Energy in the Universe
217
can be turned around to infer £2gas at the time clusters formed, redshifts z ~ 0 — 1, Qgash2 =fgasQMh2
0.4
0.5 0.6 0.7 H0 / 100 k m s~! Mpc-1
0.8
Fig. 10. Constraints used to determine the best-fit CDM model: PS = large-scale structure + CBR anisotropy; AGE = age of the Universe; CBF = cluster-baryon fraction; and HQ = Hubble constant measurements. The best-fit model, indicated by the darkest region, has Ho ~ 60— 65kms _ 1 Mpc _ 1 and QA — 0.55 — 0.65. Evidence for its smoking-gun signature - accelerated expansion - was found in 1998 (adapted from [47]).
characteristic of the Gaussian, curvature perturbations predicted by inflation. Until 1998, /lCDM's only major flaw was the absence of evidence for accelerated expansion. Not now. 3. Three dark matter problems While stars are very interesting and pretty to look at - and without them, astronomy wouldn't be astronomy and we wouldn't exist - they represent a tiny fraction of the cosmic mass budget, only about 0.5% of the critical density. As we have known for several decades, the bulk of the matter and energy in the Universe is dark. The present accounting defines three dark matter/energy problems; none is yet fully addressed. 3.1. Dark baryons By a ten to one margin, the bulk of the baryons are dark and not in the form of stars. With the exception of clusters, where the "dark" baryons exist as hot, X-ray emitting intracluster gas, the nature of the dark baryons is not known. Clusters only account for around 10% or so of the baryons in the Universe [49], and the (optically) dark baryons elsewhere, which account for 90% or more of all the baryons, could take on a different form. The two most promising possibilities for the dark baryons are diffuse hot gas and "dark stars" (white dwarfs, neutron stars, black holes or objects of mass around or below the hydrogen-burning limit). I favor the former possibility for a number of reasons. First, that's where the dark baryons in clusters are. Second, the cluster baryon fraction argument © Physica Scripta 2000
= 0.023 (£2M/0.4)(/!/0.65)1/2 .
(10)
That is, at the time clusters formed, the mean gas density was essentially equal to the baryon density (unless Quh1^2 is very small), thereby accounting for the bulk of baryons in gaseous form. Third, numerical simulations suggest that most of the baryons should still be in gaseous form today [11,50]. There are two arguments for dark stars as the baryonic dark matter. First, the gaseous baryons not associated with clusters have not been detected. Second, the results of the microlensing surveys toward the LMC and SMC [51] are consistent with about one-third of our halo being in the form of half-solar mass white dwarfs. I find neither argument compelling; gas outside clusters will be cooler ( r ~ 1 0 5 - 1 0 6 K ) and difficult to detect, either in absorption or emission. There are equally attractive explanations for the Magellanic Cloud microlensing events (e.g., self lensing by the Magellanic Clouds, lensing by stars in the spheroid, or lensing due to disk material that, due to flaring and warping of the disk, falls along the line of sight to the LMC; see [52-56]). The white-dwarf interpretation for the halo has a host of troubles: Why haven't the white dwarfs been seen [57]? The star formation rate required to produce these white dwarfs - close to 100yr _ 1 Mpc~ 3 - far exceeds that measured at any time in the past or present (see [58]). Where are the lower-main-sequence stars associated with this stellar population and the gas, expected to be 6 to 10 times that of the white dwarfs, that didn't form into stars [59]? Finally, there is evidence that the lenses for both SMC events are stars within the SMC [60-62] and at least one of the LMC events is explained by an LMC lens. The SMC/LMC microlensing puzzle can be stated another way. The lenses have all the characteristics of ordinary, low-mass stars (e.g., mass and binary frequency). If this is so, they cannot be in the halo (they would have been seen); the puzzle is to figure out where they are located. 3.2. Cold dark matter The second dark-matter problem follows from the inequality Qu — 0.4 » QB — 0.05: There is much more matter than there are baryons, and thus, nonbaryonic dark matter is the dominant form of matter. The evidence for this very profound conclusion has been mounting for almost two decades. This past year, the Burles-Tytler deuterium measurement anchored the baryon density and allowed the cleanest determination of the matter density, through the cluster baryon fraction, making the case for nonbaryonic matter almost airtight. Particle physics provides an attractive solution to the nonbaryonic dark matter problem: relic elementary particles left over from the big bang (see [63]). Long-lived or stable particles with very weak interactions can remain from the earliest moments of particle democracy in sufficient numbers to account for a significant fraction of critical density (very weak interactions are needed so that their annihilations cease before their numbers are too small). Physica Scripta T85
218
Michael S. Turner
The three most promising candidates are a neutrino of mass 30 eV or so (or J2i mv,~ 30 eV), an axion of mass 10~ 5±1 eV, and a neutralino of mass between 50GeV and 500 GeV. All three are motivated by particle-physics theories that attempt to unify the forces and particles of Nature. The fact that such particles can also account for the nonbaryonic dark matter is either a big coincidence or a big hint. Further, the fact that these particles interact with each other and ordinary matter very weakly, provides a simple and natural explanation for dark matter being more diffusely distributed. At the moment, there is significant circumstantial evidence against neutrinos as the bulk of the dark matter. Because they behave as hot dark matter, structure forms from the top down, with superclusters fragmenting into clusters and galaxies [64], in stark contrast to the observational evidence that indicates structure formed from the bottom up. (Hot + cold dark matter is still an outside possibility, with Qv ~ 0.15 or less; see [26,65]) Second, the evidence for neutrino mass based upon the atmospheric [66] and solar-neutrino [67,68] data suggests a neutrino mass pattern with the tau neutrino at 0.1 eV, the muon neutrino at 0.001 eV to 0.01 eV and the electron neutrino with an even smaller mass. In particular, the factor-of-two deficit of atmospheric muons neutrinos with its dependence upon zenith angle is very strong evidence for a neutrino mass difference squared between two of the neutrinos of around 10~2eV2 [69]. This sets a lower bound to neutrino mass of about 0.1 eV, implying neutrinos contribute at least as much mass as bright stars. WOW! The question is no longer whether there is nonbaryonic dark matter, but how much nonbaryonic dark matter and in what forms. Both the axion and neutralino behave as cold dark matter; the success of the cold dark matter model of structure formation makes them the leading particle dark-matter candidates. Because they behave as cold dark matter, they are expected to be the dark matter in our own halo; in fact, there is nothing that can keep them out [70]. As discussed above, 2/3 of the dark matter in our halo - and probably all the halo dark matter - cannot be explained by baryons in any form. The local density of halo material is estimated to be 10~ 24 gcm~ 3 , with an uncertainty of slightly less than a factor of 2 [71]. This makes the halo of our galaxy an ideal place to look for cold dark matter particles! An experiment at Livermore National Laboratory with sufficient sensitivity to detect halo axions is currently taking data [72,73] and experiments at several laboratories around the world are beginning to search for halo neutralinos with sufficient sensitivity to detect them [74]. The particle dark-matter hypothesis is compelling, and more importantly, it is now being tested. Finally, while the axion and the neutralino are the most promising particle dark-matter candidates, neither one is a "sure thing." Moreover, any sufficiently heavy particle relic (mass greater than a GeV or so) will behave like cold dark matter. A host of more exotic possibilities have been suggested, from solar-mass primordial black holes produced at the quark/hadron transition (see e.g., [75,76]) that masquerade as MACHOs in our halo to supermassive (mass greater than 10 10 GeV) particles produced by nonthermal processes at the end of inflation (see e.g., [77]). Lest we Physica Scripta T85
become overconfident, we should remember that Nature has many options for the particle dark matter.
3.3. Dark energy I have often used the term exotic to refer to particle dark matter. That term will now have to be reserved for the dark energy that is causing the accelerated expansion of the Universe - by any standard, it is more exotic and more poorly understood. Here is what we do know: it contributes about 60% of the critical density; it has pressure more negative than about — p/2; and it does not clump (otherwise it would have contributed to estimates of the mass density). The simplest possibility is the energy associated with the virtual particles that populate the quantum vacuum; in this case p = —p and the dark energy is absolutely spatially and temporally uniform. This "simple" interpretation has its difficulties. Einstein "invented" the cosmological constant to make a static model of the Universe and then he discarded it; we now know that the concept is not optional. The cosmological constant corresponds to the energy associated with the quantum vacuum. However, there is no sensible calculation of that energy (see e.g., [78-80]), with estimates ranging from 10122 to 1055 times the critical density. Some particle physicists believe that when the problem is understood, the answer will be zero. Spurred in part by the possibility that cosmologists may have actually weighed the vacuum (!), particle theorists are taking a fresh look at the problem (see e.g., [81,82]). Sundrum's proposal, that the energy of the vacuum is close to the present critical density because the graviton is a composite particle with size of order 1 cm, is indicative of the profound consequences that a cosmological constant has for fundamental physics. Because of the theoretical problems mentioned above, as well as the checkered history of the cosmological constant, theorists have explored other possibilities for a smooth, component to the dark energy (see e.g., [33]). Wilczek and I pointed out that even if the energy of the true vacuum is zero, as the Universe as cooled and went through a series of phase transitions, it could have become hung up in a metastable vacuum with nonzero vacuum energy [83]. In the context of string theory, where there are a very large number of energy-equivalent vacua, this becomes a more interesting possibility: perhaps the degeneracy of vacuum states is broken by very small effects, so small that we were not steered into the lowest energy vacuum during the earliest moments. Vilenkin [84] has suggested a tangled network of very light cosmic strings (also see, [85]) produced at the electroweak phase transition; networks of other frustrated defects (e.g., walls) are also possible. In general, the bulk equationof-state of frustrated defects is characterized by w = —N/3 where N is the dimension of the defect (N = \ for strings, = 2 for walls, etc.). The SN la data almost exclude strings, but still allow walls. An alternative that has received a lot of attention is the idea of a "decaying cosmological constant", a termed coined by the Soviet cosmologist Matvei Petrovich Bronstein in 1933 [86]. (Bronstein was executed on Stalin's orders in 1938, presumably for reasons not directly related to the cosmological constant; see [87].) The term is, of course, © Physica Scripta 2000
Dark Matter and Energy in the Universe
219
While we have many urgent questions, we can see a flood of precision cosmological and laboratory data coming that will help to answer these questions: High-resolution maps of CBR anisotropy (MAP and Planck); large redshift surveys (SDSS and 2dF); more SN la data; experiments to directly detect halo axions and neutralinos; more microlensing data (MACHO, EROSII, OGLE, AGAPE, and super MACHO); accelerator experiments at Fermilab and CERN, searching for the neutralino and its supersymmetric friends and further evidence for neutrino and its equation of motion by (see e.g., [88]) mass, and at the KEK and SLAC B-factories, revealing more about the nature of CP violation; and nonaccelerator '<j> + 3HJ>+V(k
4. Concluding remarks 1998 was a very good year for cosmology. We now have a plausible and complete accounting of matter and energy in the Universe; in ACDM, a model for structure formation that is consistent with all the data at hand; and the first evidence for the key tenets of inflation (flat Universe and adiabatic density perturbations). One normally conservative cosmologist has gone out on a limb by stating that 1998 may be a turning point in cosmology as important as 1964, when the CBR was discovered [98]. We still have important questions to address: Where are the dark baryons? What is the dark matter? What is the nature of the dark energy? What is the explanation for the complicated pattern recipe for our universe: neutrinos (0.3%), baryons (5%), cold dark matter particles (35%) and dark energy (60%)? Especially puzzling is the ratio of dark energy to dark matter: because they evolve differently with time, the ratio of dark matter to dark energy was higher in the past and will be smaller in the future; only today are they comparable. WHY NOW? © Physica Scripta 2000
16. Mohr, J., Mathiesen, B. and Evrard, A. E. Astrophys. J. 517, 627 (1998). 17. Carlstrom, J., this volume (2000). 18. Henry, P., BAAS 31, 1301H (1999). 19. Bahcall, N. and Fan, X., Astrophys. J. 504, 1 (1998). 20. Bahcall, N., this volume (2000). 21. Dekel, A. and Rees, M., Astrophys. J. Lett. 422, LI (1994). 22. Willick, J. and Strauss, M., Astrophys. J. 507, 64 (1998). 23. Sigad, et al, Astrophys. J. 495, 516 (1998). 24. Dekel, A., Ann. Rev. Astron. Astrophys., 32, 371 (1994). 25. Peacock, J. and Dodds, S., Mon. Not. R. Astron. Soc. 267, 1020 (1994). 26. Dodelson, S., Gates, E. I. and Turner, M. S., Science 274, 69 (1996). 27. Faber, S. and Gallagher, J., Ann. Rev. Astron. Astrophys. 17, 135 (1979). 28. Persic, M. and Salucci, P., Mon. Not. R. Astron. Soc. 258, 14p (1992). 29. Carlberg, R. et al, Astrophys. J. 436, 32 (1996). 30. Carlberg, R. et al, Astrophys. J. 478, 462 (1997). 31. Turner, M. S., Steigman, G. and Krauss, L., Phys. Rev. Lett. 52, 2090 (1984); Peebles, P. J. E. Astrophys. J. 284, 439 (1984). 32. Turner, M. S., Physica Scripta T36, 167 (1991). 33. Turner, M. S. and White, M., Phys. Rev. D 56, R4439 (1997). 34. Chaboyer, B. et al, Astrophys. J. 494, 96 (1998). 35. Freedman, W., this volume (2000). 36. Perlmutter, S. et al, Astrophys. J. 517, 565 (1999). 37. Goobar, A., this volume (2000). 38. Riess, A. et al, Astron. J. 116, 1009 (1998). 39. Schmidt, B. et al, Astrophys. J. 507, 46 (1998). Physica Scripta T85
220
Michael S. Turner
40. Mohr, J. et al., in preparation (2000). 41. Guerra, E. J., Daly, R. A. and Wan, L., Astrophys. J. submitted (astro-ph/9807249) (1998). 42. Turner, E. L„ Astrophys. J. Lett. 365, L43 (1990). 43. Kochanek, C , Astrophys. J. 466, 638 (1996). 44. Waga, I. and Miceli, A. P. M. R., Phys. Rev. D 59, 103507 (1999). 45. Chiba, M. and Yoshii, Y. Astrophys. J. 510, 42 (1999). 46. Phillips, M. M , Astrophys. J. Lett. 413, L105 (1993). 47. Krauss, L. and Turner, M. S., Gen. Rel. Grav. 27, 1137 (1995). 48. Ostriker, J. P. and Steinhardt, P. J., Nature 377, 600 (1995). 49. Turner, M. S., in Critical Dialogues in Cosmology, ed. N. Turok (World Scientific, Singapore 1997), p. 555. 50. Ostriker, J. P., this volume (2000). 51. Spiro, M., this volume (2000). 52. Sahu, K. C , Nature 370, 265 (1994). 53. Evans, W., Gyuk, G., Turner, M. S. and Binney, J., Astrophys. J. Lett. 501, L45 (1998). 54. Gates, E. I., Gyuk, G., Holder, G. and Turner, M. S., Astrophys. J. Lett. 500, 145 (1998). 55. Zaritsky, D. and Lin, D. N. C , Astron. J. 114, 254 (1997). 56. Zhao, H.-S., Mon. Not. R. Astron. Soc. 294, 139 (1998). 57. Graff, D. S., Laughlin, G. and Freese, K., Astrophys. J. 499, 7 (1998). 58. Madau, P., this volume (2000). 59. Fields, B., Mathews, G. J. and Schramm, D. N., Astrophys. J. 483, 625 (1997). 60. Alcock, C. et al, Astrophys. J. 518, 44 (1999). 61. EROS Collaboration, Astron. Astrophys. 332, 1 (1998). 62. EROS Collaboration, Astron. Astrophys. 337, L17 (1998). 63. Ellis, J.E., this volume (2000). 64. White, S. D. M., Frenk, C. and Davis, M., Astrophys. J. Lett. 274, LI (1983). 65. Gawiser, E. and Silk, J., Science 280, 1405 (1998). 66. Totsuka, Y., this volume (2000). 67. Kirsten, T., this volume (2000). 68. Bahcall, J., this volume (2000).
Physica Scripta T85
69. Fukuda, Y. et al. (SuperKamiokande Collaboration), Phys. Rev. Lett. 81, 1562 (1998). 70. Gates, E. I. and Turner, M. S., Phys. Rev. Lett. 72, 2520 (1994). 71. Gates, E. I., Gyuk, G. and Turner, M. S., Astrophys. J. Lett. 449, 123 (1995). 72. van Bibber, K. et al, Phys. Rev. Lett. 80, 2043 (1998). 73. Rosenberg, L. and van Bibber, K., Phys. Rep., in press (2000). 74. Sadoulet, B., this volume (2000). 75. Jedamzik, K., Phys. Rept. 307, 155 (1998). 76. Jedamzik, K. and Niemeyer, J., Phys. Rev. Lett. 80, 5481 (1998). 77. Kolb, E.W., this volume (2000). 78. Zel'dovich, Ya.B., JETP 6, 316 (1967). 79. Bludman, S. and Ruderman, M. Phys. Rev. Lett. 38, 255 (1977). 80. Weinberg, S., Rev. Mod. Phys. 61, 1 (1989). 81. Harvey, J., Phys. Rev. D 59, 026002 (1999). 82. Sundrum, R., JHEP 9907, 001 (1999). 83. Turner, M. S. and Wilczek, F., Nature 298, 633 (1982). 84. Vilenkin, A., Phys. Rev. Lett. 53, 1016 (1984). 85. Spergel, D. N. and Pen, U.-L., Astrophys. J. Lett. 491, L67 (1997). 86. Bronstein, M.P., Phys. Zeit. der Sowjetunion, 3, 73 (1933). 87. Kragh, H., "Cosmology and Controversy", (Princeton Univ. Press, Princeton, NJ, 1996). 88. Turner, M. S., Phys. Rev. D 28, 1243 (1983). 89. Huterer, D. and Turner, M. S., Phys. Rev. D 60, 081301 (1999). 90. Freese, K. et al, Nucl. Phys. B, 287, 797 (1987). 91. Ozer, M. and Taha, M. O., Nucl. Phys. B 287, 776 (1987). 92. Ratra, B. and Peebles, P. J. E., Phys. Rev. D 37, 3406 (1988). 93. Frieman, J., Hill, C , Stebbins, A. and Waga, I., Phys. Rev. Lett. 75, 2077 (1995). 94. Coble, K., Dodelson, S. and Frieman, J. A., Phys. Rev. D 55, 1851 (1996). 95. Caldwell, R., Dave, R. and Steinhardt, P.J., Phys. Rev. Lett. 80, 1582 (1998) 96. Steinhardt, P. J., this volume (2000). 97. Carroll, S., Phys. Rev. Lett. 81, 3067 (1998). 98. Turner, M. S., Pub. Astron. Soc. P a c , 111, 264 (1999).
© Physica Scripta 2000
Physica Scripta.Vol. T85, 221-230, 2000
Particle Candidates for Dark Matter John Ellis Theoretical Physics Division, CERN CH-1211 Geneva 23, Switzerland Received December 11, 1998; accepted August 2, 1999
PACS Ref: 12.10.Dm, 12.60.Jv, 14.60.Pq, 14.80.Ly, 95.85.Ry, 96.40.Pq, 98.80.Cq, 98.80.Es
Abstract Some particle candidates for dark matter are reviewed in the light of recent experimental and theoretical developments. Models for massive neutrinos are discussed in the light of the recent atmospheric-neutrino data, and used to motivate comments on the plausibility of different solutions to the solar neutrino problem. Arguments are given that the lightest supersymmetric particle should be a neutralino % and accelerator and astrophysical constraints used to suggest that 50GeV <mx< 600 GeV. Minimizing the fine tuning of the gauge hierarchy favours Qxh2 ~ 0.1. The possibility of superheavy relic particles is mentioned, and candidates from string and M theory are reviewed. Finally, the possibility of non-zero vacuum energy is discussed: its calculation is a great opportunity for a quantum theory of gravity, and the possibility that it is time dependent should not be forgotten.
It is of interest for our subsequent discussion to review in more detail [2] the upper limit (3) on the mass of a cold dark matter particle. One may write Qx =
Px 2x
mxnx Wh2T2'
(4)
where 7b ~ 2.73K is the present effective temperature of the cosmic microwave background radiation. To a good approximation, the comoving number density has remained essentially constant since the freeze-out temperature 7} at which annihilation terminated: T2 mp
«£ nx(T{) nx(T{)((Jznn(XX)vx)=-~-±.. 3 T ~ a 1. Introduction There is a wide range of possible masses for a candidate dark matter particle [1]. If it was once in thermal equilibrium, its number density n is almost independent of its mass, as long as the latter is <£ 1 MeV. Thus, for neutrinos with masses in the range discussed in the next section, nv ~ constant and hence pv = mvnv oc mv, leading to
°.»-&y-nm-
a)
where h is the present Hubble expansion rate in units of 100 kms _ 1 Mpc _ 1 and pc is the critical density. There is a region of masses for neutrinos, or similar particles, between 0(100)eV and 0(3)GeV where Qv > 1. Above this range, the cosmological density px of a particle X may be sufficiently suppressed by mutual annihilation: m
Px = xnx
• nx oc
a{xx^...y
(2)
that Qx % 1. In the case of neutrinos or similar particles, annihilation is particularly efficient when mv ~ ntz/2, leading to a local minimum in the relic density with Qv «; 1. Above this mass, the relic density in general rises again, depending on the behaviour of the annihilation cross-section. Thus one has three regions where the relic density of a neutrino or similar particle may be of cosmological interest: Qv ~ 1: 30eV, ~ fewGeV, or ~ lOOGeV.
(3)
The first of these possibilities corresponds to hot dark matter, and is of active interest for neutrinos as discussed in Section 2, whilst the other two correspond to cold dark matter, and are of interest for supersymmetric particles, as discussed in Section 3. The middle region has been essentially excluded by LEP, as we discuss in more detail later, whilst the third may be in the realm of the LHC. © Physica Scripta 2000
(5)
where a is the cosmological scale factor and mp ~ 1.2 x 1019 GeV is the Planck mass. For relic particles of interest, one typically finds that mx/T( ~ 20 to 30 and hence Qxh1-
10" (GzUXX^x)
10" \T0mP)
<
n
(XX)vx)TeVl
(6)
The TeV scale emerges naturally as the geometric mean of T0~2J3K and m P ~ 1 . 2 x l 0 1 9 GeV! Using the dimensional estimate (Ozrm(XX)vx) ~ C • a2/mx, where a is a generic coupling strength and C is a model-dependent numerical coefficient, we find mx
(16aVC)A/^TeV,
(7)
which yields the expectation that mx ;S 1 TeV. We will discuss in Section 3 the extent to which this argument implies that the LHC is "guaranteed" to discover a cold dark matter particle. For the moment, we just point out that the above discussion assumed that the particle was at one time in thermal equilibrium, which is not necessarily the case. Counter-examples include the axion discussed here by Turner [3], and the superheavy relic particles discussed here by Kolb [4], and in Section 4 of this talk. 2. Neutrinos Particle physics experiments [5] tell us that neutrino masses must be much smaller than those of the corresponding charged leptons and quarks: mVe <3.5eV vs. me ~ 0.511 MeV mVfi < 160keV vs. wM ~ 105 MeV,
(8)
wVt < 18MeV vs. mT ~ 1.78 GeV. Physica Scripta T85
222
John Ellis
There is another difference between neutrinos and other particles, namely that only left-handed neutrinos are known to exist, produced by the familiar V — A charged current:
following approximate eigenstates and eigenvalues: v
whereas both quarks and charged leptons have both left- and right-handed states ^L,R, ^L,R- Thus "Dirac" masses mP coupling them are possible: gh{THAI=y2,AL=0fRfh =^mf = gni[(0\HAI=y2,AL=0\0),
(10)
where the quantum numbers of the Standard Model Higgs have been indicated explicitly. The following puzzles then arise. If right-handed neutrinos VR exist, why are the mv (10) so small? If vR do not exist, can neutrinos acquire masses: wv^0? Most particle theorists believe that particles should be massless only if there is an exact gauge symmetry to guarantee this, as is the case for the photon and gluon. However, there is no such candidate symmetry to guarantee mv = 0, so most of us expect mv ^ 0. The fact that the v have no exact gauge quantum numbers enables them to have Majorana masses mM, since / R in (12) is replaced byf[ =f^C, where C is an antisymmetric matrix:
0 ( ^ ) v
R
:
W L
= ^ M - K )
t
= ^ ) , (15)
(9)
Jn = eyM(i - y 5 K + mv(l - 75K + ^ O - TsK,
L + V
R + ° ^ 7 F ) V L : W R = M.
M
Thus there are naturally very light neutrinos if M 3> mw, as would be expected in a GUT with M = 0(MQUJ). If the Dirac masses of the different neutrino generations scale like the corresponding quark or charged-lepton masses, one would expect ml or m2, m,
(16)
M,
and hence mVe «.mv
(17)
«mVt,
if the heavy Majorana mass matrix is diagonal, and if its eigenvalues M, are approximately the same. As an example, putting mj5 ~ 100 GeV for the third generation, one finds mV3 - 0 . 1 eV if M 3 ~ 1014 GeV. Before the advent of the atmospheric neutrino data [7-9], one might also have expected small neutrino mixing angles, analogous to those for quarks, which originate from Dirac mass matrices. M, This and the prejudice (17) are not necessarily supported m™v< vL = m„M.,T > i C v L = mfv • v . (11) h L by the atmospheric neutrino data [7-9], which suggest a large This is not possible for quarks and charged leptons, because mixing angle: sin2 26^ > 0.8 and a mass-squared difference both <7L • #L and 1-L • i-h have Qem ^ 0, whilst q^ • q^ also Amltmo ~ (1°~ 2 t o 10" 3 )eV 2 . We also recall that the solar has non-zero colour. Such a Majorana neutrino mass (11) neutrino data [10,11] favour one of three possible solutions: would require lepton-number violation: AL = 2 and weak the large-angle MSW solution with isospin AI = 1, which does contradict any sacred theoretical principles, but is not provided by the Higgs fields in the Stan- Am, lO- 5 eV 2 ,sin 2 20>O.8, (18a) solar dard Model or in the minimal SU(5) GUT. A Majorana mass term (13) could in principle be provided by a suitable the small-angle MSW solution with "exotic" Higgs field: 5 2 2 2 (18b) Am,solar 10~ eV ,sin 29~ 10" ,
gHw#d/=i,/iL=2VL • VL =*• mv = gH„WHAI=iiAL=2\0},
(12) or vacuum oscillations with
as appears in some non-minimal GUT models such as SO(10) with a 126 Higgs representation. However, there are difficulties with sich a scenario, since one would have expected a AI = 1 Majoron particle which should have been detected via invisible Z° decays. Alternatively, one can obtain m™ from a non-renormalizable coupling to the Standard Model Higgs:
M
(HAI=l/2Vh) • (HM^I/2VL)
•
rrC =gs-
M
— (13)
The question then arises: what could be the origin of the large mass parameter M? The most natural possibility is the exchange of a massive singlet neutrino field, traditionally called a right-handed neutrino vR, though this nomenclature is somewhat anachronistic. Given a VR, a Dirac mass m® =gHvv(0|^/=i/2|0) is also possible, and one arrives at the famous see-saw mass matrix [6]: 0 (o
I - M (VL,VR)
m
D
(14)
\mmrf v M Mj\vRJOne would expect the M^ entries in (16) to be of order mq or mt < {Mw), so the matrix diagonalization yields the 1
Physica Scripta T85
Am:solar
lO-' u eV 2 ,sin z 20£O.6.
(18c)
Which of these solutions might be favoured by postsuper-Kamiokande models of neutrino masses, and how plausible is neutrino hot dark matter of cosmological and astrophysical significance, which would require w v > l eV for at least one neutrino species? A first comment is that large neutrino mixing is (in retrospect) not at all implausible [12]. For one thing, it is not at all necessary that m ° oc mq or mi. For example, in a specific flipped SU (5) model used earlier to discuss quark and lepton masses, we found [13]
m:
(0(n)
0(1)
o '
V o
o
oil),
cm o(i) o
(19)
where r\ is a small parameter, which would yield at least one large neutrino mixing angle. Moreover, there is no good reason why the heavy singlet Majorana mass matrix should be (approximately) diagonal in the same basis. For example, in models with a U(l) flavour symmetry one expects matrix elements My a £"<+"; M G U T
(20) © Physica Scripta 2000
Particle Candidates for Dark Matter
Fig. 1. Light-neutrino mixing in a simple 2 x 2 model (22), as a function of the ratios of mass-matrix elements [13].
-10
223
Fig. 3. Possible renormalization-group running of the light-neutrino mixing angle [13].
sin 023 without fine-tuning of the ratios b/d, c/d, for example if b/d ~ 0.5 and c/d ~ 1.5. Moreover, the light-neutrino mixing angle may be enhanced by renormalization-group effects between the GUT and electro weak scales [15]:
0.25
b/d=0.5\
U -20
d ,2.mi+m22 16nl j (sinz 023) = -2(sin^ 023)(cos'! 023) x (A| - XA2) m 33 - m22 (23)
Pi
-30
as seen in Fig. 3. The renormalization-group enhancement is particularly important if A2 — X\ is large, as may happen at large tan /? in supersymmetric models, and/or if the diagonal 0.5 entries w 33 and W22 are almost equal. We infer that the hier1.5 archy (21) is plausible if Am2oXM ~ 10~5 eV2 as in the largeFig. 2. Light-neutrino mass hierarchy in a simple 2 x 2 model (22), as a funcand small-angle MSW solutions (18a,b). However, it seems tion of the ratios of mass-matrix elements [13]. difficult to stretch the hierarchy (24) as far as would be required by the vacuum solution (18c) to the solar neutrino problem. Our present inclination is therefore to favour where e is an expansion parameter. If n, = — rij, one gets a large off-diagonal entry My = 0{\) x MGUT, which looks (24) (lo -1 to I0"15)ev, m2 ~ l r r V v at least as plausible as having n, = 0, which would be mv required to give a large diagonal entry M„ = C ( M G U T ) - Such and the remaining question is whether the large- or a large off-diagonal entry would be another source of large small-angle MSW solution (18a,b) is to be favoured. In neutrino mixing. our specific flipped SU(5) model [13], we found not only large With so many different features contributing to the light off-diagonal entries in m® (22), but also large off-diagonal neutrino mass matrix, there is no obvious symmetry or other entries (23) in the singlet-neutrino mass matrix: reason why near-degeneracy w, ~ m; » |w, — mj\ should occur. Therefore, we expect that there may be a hierarchy of masses: (25) -40
m3 ~ y/Amltmo
>m2~
y^m 2 0 iar > mx.
(21)
If this is indeed the case, even the heaviest light neutrino would weigh < 0.1 eV, and would not be of great astrophysical and cosmological interest [14]. A corollary question is the extent to which large mixing is compatible with a large neutrino mass hierarchy. To address this, we may consider a simple two-state model [13] m
*=(d
c)
(22)
and diagonalize it to obtain eigenvalues m2 and m3 and a mixing angle 023- We see in Figs. 1 and 2 that a hierarchy mj, = O(10)w2 is quite compatible with large mixing © Physica Scripta 2000
where all of the indicated non-zero entries might a priori be comparable in magnitude. We therefore find the large-angle MSW solution (18a) to be at least as plausible as the small-angle MSW solution (18b), perhaps even more so. A final comment concerns the magnitudes of the entries in (28): we estimate their natural order of magnitude to be O(10 1 3 ± 2 ) GeV, corresponding to wV3 = 10 0±2 eV, overlapping comfortably with the desired range (21). In my view, although the super-Kamiokande data [7] make a very strong case for atmospheric neutrino oscillations, particle physicists will not be completely convinced until they have verified the effect using a beam with controllable energy, spectrum and flavour content, as provided by an accelerator neutrino beam. One such project, Physica Scripta T85
224
John Ellis
the K2K experiment [16], is to start taking data in 1999, and another, the NUMI/MINOS project [17], has also been approved. A third project, neutrino Gran Sasso (NGS), has been studied by a joint CERN-INFN working group [18], and is ready for construction if funding is approved. The K2K project is at relatively low energy, insufficient to produce the x lepton. The NUMI/MINOS and NGS projects are at higher energies, and should be able to reach down to Am2 ~ 10~3 eV2, though without much margin. Nevertheless, in a few years we may expect to know whether accelerator experiments confirm the super-Kamiokande results. However, as mentioned earlier, it currently seems unlikely that neutrinos will turn out to be an important component of the dark matter [14].
3. The lightest supersymmetric particle The motivation for supersymmetry at an accessible energy is provided by the gauge hierarchy problem [19], namely that of understanding why m w <§: mp, the only candidate for a fundamental mass scale in physics. Alternatively and equivalently, one may ask why (?F ~ g V ^ w ^ ^ N = 1/wp, where Mp is the Planck mass, expected to be the fundamental gravitational mass scale. Or one may ask why the Coulomb potential inside an atom is so much larger than the Newton potential, which is equivalent to why e2 = 0(1) » WpWe/wp, where mp,e are the proton and electron masses. One might think it would be sufficient to choose the bare mass parameters: mw <£ mp. However, one must then contend with quantum corrections, which are quadratically divergent: <5<w = OQA2
(26)
which is much larger than m w , if the cutoff A representing the appearance of new physics is taken to be 0(mp). This means that one must fine-tune the bare mass parameter so that it is almost exactly cancelled by the quantum correction (26) in order to obtain a small physical value of mw. This seems unnatural, and the alternative is to introduce new physics at the TeV scale, so that the correction (26) is naturally small. At one stage, it was proposed that this new physics might correspond to the Higgs boson being composite [20]. However, calculable scenarios of this type are inconsistent with the precision electroweak data from LEP and elsewhere. The alternative is to postulate approximate supersymmetry [21], whose pairs of bosons and fermions produce naturally cancelling quantum corrections: 5m\l = oQ\ml-ml\
(27)
that are naturally small: Sm2^ < m\, if |m|-m||
(28)
There are many other possible motivations for supersymmetry, but this is the only one that gives reason to expect that it might be accessible to the current generation of accelerators and in the range (7) expected for a cold dark matter particle. Physica Scripta T85
The minimal supersymmetric extension of the Standard Model (MSSM) has the same gauge interactions as the Standard Model, and the Yukawa interactions are very similar: Xd QDCH + X\LECH + XUQ UCH + /iHH
(29)
where the capital letters denote supermultiplets with the same quantum numbers as the left-handed fermions of the Standard Model. The couplings ld,i,u give masses to down quarks, leptons and up quarks respectively, via distinct Higgs fields H and H, which are required in order to cancel triangle anomalies. The new parameter in (22) is the bilinear coupling /i between these Higgs fields, that plays a significant role in the description of the lightest supersymmetric particle, as we see below. The gauge quantum numbers do not forbid the appearance of additional couplings XLLEC + X'LQDC + X UCDCDC
(30)
but these violate lepton or baryon number, and we assume they are absent. One significant aspect of the MSSM is that the quartic scalar interactions are determined, leading to important constraints on the Higgs mass, as we also see below. Supersymmetry must be broken, since supersymmetric partner particles do not have identical masses, and this is usually parametrized by scalar mass parameters m2, |0,-|2, gaugino masses \ Ma Va • Va and trilinear scalar couplings •AijkXyk4>i
l = ml + Cm2l/2, Ma = (-^— )mi/2
(31)
\aGUT/
where the C, are calculable coefficients [22] and MSSM phenomenology is then parametrized by fi, mo, m\/z, A and tan/? (the ratio of Higgs v.e.v.'s). Precision electroweak data from LEP and elsewhere provide two qualitative indications in favour of supersymmetry. One is that the inferred magnitude of quantum corrections favour a relatively light Higgs boson [23] m/, = 66+™±10GeV
(32)
which is highly consistent with the value predicted in the MSSM: m/, < 150 GeV [24] as a result of the constrained quartic couplings. (On the other hand, composite Higgs models predicted an effective Higgs mass > 1 TeV and other unseen quantum corrections.) The other indication in favour of low-energy supersymmetry is provided by measurements of the gauge couplings at LEP, that correspond to sin2 0w — 0.231 in agreement with the predictions of supersymmetric GUTs with sparticles weighing about 1 TeV, but in disagreement with non-supersymmetric GUTs that predict sin2 0 W ~ 0.21 to 0.22 [25]. Neither of these arguments provides an accurate estimate of the sparticle mass scales, however, since they are both only logarithmically sensitive to mo and/or m\/2© Physica Scripta 2000
Particle Candidates for Dark Matter The lightest supersymmetric particle (LSP) is expected to be stable in the MSSM, and hence should be present in the Universe today as a cosmological relic from the Big Bang [26]. This is a consequence of a multiplicativelyconserved quantum number called R parity, which is related to baryon number, lepton number and spin: 3B+L+2S
It is easy to check that R = +1 for all Standard Model particles and R = — 1 for all their supersymmetric partners. The interactions (33) would violate R, but not a Majorana neutrino mass term or the other interactions in SU(5) or SO(10) GUTs. There are three important consequences of R conservation: (i) sparticles are always produced in pairs, e.g., pp -> qgX, e + e~ ->• Ji+Jr, (ii) heavier sparticles decay into lighter sparticles, e.g., q -*• qg, ji-*- /iy, and (iii) the LSP is stable because it has no legal decay mode. If such a supersymmetric relic particle had either electric charge or strong interactions, it would have condensed along with ordinary baryonic matter during the formation of astrophysical structures, and should be present in the Universe today in anomalous heavy isotopes. These have not been seen in studies of H, He, Be, Li, O, C, Na, B and F isotopes at levels ranging from 10~" to 10~29 [27], which are far below the calculated relic abundances from the Big Bang: > 1(T6 to 10
1000
in =100
(a)
h S\ V*. *-. A-
(33)
* = (-!>
"relic
225
•10
CN
Qh =0.3
***>..."
* *
:• fK \
\.
nh =0.1
1001— -1000
-100
(a)
1000
(34)
for relics with electromagnetic or strong interactions. Except possibly for very heavy relics, one would expect these primordial relic particles to condense into galaxies, stars and planets, along with ordinary bayonic material, and hence show up as an anaomalous heavy isotope of one or more of the elements studied. There would also be a "cosmic rain" of such relics [28], but this would presumably not be the dominant source of such particles on earth. The conflict with (37) is sufficiently acute that the lightest supersymmetric relic must presumably be electromagnetically neutral and weakly interacting [26]. In particular, I believe that the possibility of a stable gluino can be excluded. This leaves as scandidates for cold dark matter a sneutrino v with spin 0, some neutralino mixture of y/H°/Z with spin 1/2, and the gravitino G with spin 3/2. LEP searches for invisible Z° decays require mj ^ 4 3 GeV [29], and searches for the interactions of relic particles with nuclei then enforce wj > few TeV [30], so we exclude this possibility for the LSP. The possibility of a gravitino G LSP has attracted renewed interest recently with the revival of gauge-mediated models of supersymmetry breaking [31], and could constitute warm dark matter if MA ~ 1 keV. In this talk, however, I concentrate on the y/H /Z° neutralino combination x, which is the best supersymmetric candidate for cold dark matter. The neutralinos and charginos may be characterized at the tree level by three parameters: m\/2, n and tan/?. The lightest neutralino % simplifies in the limit m\/2 -*• 0 where it becomes essentially a pure photino y, or n ->• 0 where it becomes essentially a pure higgsino H. These possibilities are excluded, however, by LEP and the FNAL Tevatron collider [29]. From the point of view of astrophysics and © Physica Scripta 2000
1000
(b) Fig. 4. Regions of the (ji, Mi) plane in which the supersymmetric relic density may lie within the interesting range 0.1 < Qxh2 < 0.3 [32].
cosmology, it is encouraging that there are generic domains of the remaining parameter space where Qxh2 ~ 0.1 to 1, in particular in regions where % is approximately a U(l) gaugino B, as seen in Fig. 4 [32]. Purely experimental searches at LEP enforce w x J>30 GeV, as seen in Fig. 5 [33]. This bound can be strengthened by making various theoretical assumptions, such as the Physica Scripta T85
226
John Ellis 300
ALEPH PRELIMINARY
(a)
60 >
M
•
o ==
m +m
200 GeV/c 2
X
50
=182 •
m ± =91
X'
LEP 183 "
x
250
"S
p =0.95
m =90
Qrf=0 1 200
P^O.SI -
/
y
y /^'
.
m =8 h
150 80
90
70
ll Fig. 5. Experimental lower limit on the lightest neutralino mass, inferred from unsuccessful chargino and neutralino searches at LEP [33].
s \
a)
80
\Qh=0.1
\
300
u<0
"
\
\ UHM
20
"""
LEP
250 100 3
4
5 6 7 8 9
C
"
\
110
m.=91 n
"••-,
120
Fig. 7. The regions of the (ji, M2) plane where the lightest supersymmetric particle may still by a Higgsino, taking into account the indicated LEP constraints [32]. The Higgsino purity is indicated by p2.
I ^>o
\ \ cosmo + UHM 60
\
H
10
b)
^ 80
^
\
)M + UHM
40
\
m +m =182\
\ cos mo + UHM
. m,±=91
'•-. \ \
universality of scalar masses mo,., including in the Higgs sector, the cosmoiogical dark matter requirement that Qy,h2 < 0.3 and the astrophysical preference that Qxh2 > 0.1. Taken together as in Fig. 6, we see that they enforce
DM + UHM UHM ~~~~—
40
LEP
""•••...
20
w x >42GeV 1
2
3
4
5
6
7
8
9
10
tan(3 Fig. 6. Theoretical lower limits on the lightest neutralino mass, obtained by using the unsuccessful Higgs searches (H), the cosmoiogical upper limit on the relic density ( Q , the assumption that all input scalar masses are universal, including those of the Higgs multiplets (UHM), and combining this with the cosmoiogical upper (cosmo) and astrophysical lower (DM) limits on the cold dark matter density [29]. Physica Scripta T85
(35)
and LEP should eventually be able to establish or exclude mx up to about 50 GeV. As seen in Fig. 7, LEP has already explored almost all the parameter space available for a Higgsino-like LSP, and this possibility will also be thoroughly explored by LEP [33]. Should one be concerned that no sparticles have yet been seen by either LEP or the FNAL Tevatron collider? One way to quantify this is via the amount of fine-tuning of © Physica Scripta 2000
Particle Candidates for Dark Matter
227
a)tm(5=3. n < 0
m_< n*
Ac(Gev) in*
riV/M-i/a
1 I 1 1 ! 1 I 1 I | 1 1 1 I | 1 I 1 I | I I I 1;
10*
: 1 | 1 1 1 | 1 1
100
200
300
• W&4£yM^^m '
10 2
400 m.„
500
600
100
c) tan (5 = !0,n<0
':*Jj^' - '
r
10
100
SO
309
400
500
'
400
WO
BOO
100C
500
600
;,'d)tanp=10,|i>0
1200
J 1 | 1 I 1 1 | i I i I | I I I I | I . . , , ,
^
-^|ji^':,#
JlllllF
ID 5
%
•
100
^
^
tanf?=10 : 1000
2000
'
60
100
110
120
200
300
400 m
10
0
600
MsfGeV) 10 3
-1O0O
500
tan|S=10 :
m , (GeV)
-2000
300 400 111,,.
-_
tanJ3=1Q ! 0
200
!.'2
500
600
iOO
200
300
400
™V2
Fig. 10. The change in the domain of parameter space allowed by the requirements 0.1 < Cixh2 < 0.3 after (shaded region) and before (dashed lines) including T co-annihilation [38].
130
M„ (GeV)
H (GeV)
Fig. 5. The fine-tuning price A0 imposed by LEP for tan/J = 10, as a function of model parameters [35].
the input parameters required to obtain the physical value of m w [34]: a/ 3ww (36) mw da\ where a\ is a generic supergravity input parameter. As seen in Fig. 8, the LEP exclusions impose [35] A0 = Max;
I a) ^(si
10 2 7
•
£^JlMi
;Lrf
4 , £8.
J80 10
§PP*
ro°o
0
20C
400
600
M1/2 (GeV)
0
500
1000
mD(GeV)
1500
2000
100
110
120
Although fine-tuning is a matter of taste, this is perhaps not large enough to be alarming, and could in any case be reduced significantly if a suitable theoretical relation between some input parameters is postulated [35]. It is interesting to note that the amount of fine-tuning A0 is minimized when Qxh2 ~ 0.1 as preferred astrophysically, as seen in Fig. 9 [36]. This means that solving the gauge hierarchy problem naturally leads to a relic neutralino density in the range of interest to astrophysics and cosmology. I am unaware of any analogous argument for the neutrino or the axion. As mx increases, the LSP annihilation cross-section decreases and hence its relic number and mass density increase. How heavy could the LSP be? Until recently, the limit given was m x <300 GeV [37]. However, it has now been pointed out that there are regions of the MSSM parameter space where co-annihilations of the % with the stau slepton T could be important, as seen in Fig. 10 [38]. These co-annihilations would suppress Qx, allowing a heavier neutralino mass, and we now find that [38]
M„(GeV)
Fig. 9 The correlation between the fine-tuning price A0 and the relic density Qxh2, showing dependences on model parameters [36]. Physica Scripta 2000
(37)
mx;$ 600 GeV
(38)
is possible. In the past, it was thought that all the Physica Scripta T85
228
John Ellis
E<5 x 1019 GeV, because of the reaction P + 7CMBR -*• ^ + [42]. However, no such GreisenZatsepin-Kuzmin cut-off is seen in the data! [43] The sasooj ultra-high-energy cosmic rays must originate nearby, and !'L«It= lOOfb"' — ^ should point back to any point-like sources such as AGNs. A,, = 0 , » B p - 2 , | i < 0 -A \ However, no such sources have been seen. v " Could the ultra-high-energy cosmic rays be due to the ^\ % decays of superheavy relic particles? These should be \ clustered in galactic haloes (including our own), and hence "- . ^x \ gaooo) 21 OS i x^ give an anisotropic flux [44], but there would be no obvious \ point sources. There have been some reports of anisotropies -. 2!SS "^~rrr——_ "^ in high-energy cosmic rays, but it is not clear whether they could originate in superheavy relic decays. We have analyzed [45] recently possible superheavy relic ' . - -__.. ,—-:: ;•-•••; ." 8(1500) candidates in string [46] and/or M theory. One expects Kaluza-Klein states when six excess dimensions are compactified: 10 —>• 4 or 11 —> 5, which we call hexons. However, , / - ~ — ^ these are expected to weigh > 1016 GeV, which may be too heavy, and there is no particular reason to expect hexons WOOD) \ to be metastable. In M theory, one expects massive states associated with a further compactification: 5 -* 4 dimensions, which we call pentons. Their mass could be ~ 1013 GeV, which would be suitable, but there is again 8(5001 \.l no good reason to expect them to be metastable. We are left with bound states from the hidden sector of string IM theory, which we call cryptons [46]. These could also have masses ~ 1013 GeV, and might be metastable for much the same reason as the proton in a GUT, decaying via higher-dimensional multiparticle operators. For example, m0 (GeV) in a flipped SU(5) model we have a hidden-sector Fig. 11. The region of the (mo, mi/2) plane accessible to sparticle searches at SU(4) x SO(10) gauge group, and the former factor confines four-constituent states which we call tetrons. Initial the LHC [40]. studies [46,45] indicate that the lightest of these might well have a lifetime >10 17 y, which would be suitable for the cosmologically-preferred region of MSSM parameter space' decays of superheavy dark matter particles. Detailed could be explored by the LHC [40], as seen in Fig. 11, but it simulations have been made of the spectra of particles of their decay now seems possible that there may be a delicate region close produced by the fragmentation products [47,48], and the ultra-high-energy cosmic-ray data to the upper bound (38). This point requires further study. are consistent with the decays of superheavy relics weighing ~ 1012 GeV, as seen in Fig. 12 [48]. Issues to be resolved here include the roles of supersymmetric particles in the fragmentation cascades, and the relative fluxes of 7, v and 4. Superheavy Relic Particles p among the ultra-high-energy cosmic rays. The expectation (7), exemplified by the MSSM range (41), was based on the assumption that the cold dark matter particles were at one time in thermal equilibrium. As dis- 5. Vacuum energy cussed here by Kolb [4], much heavier relic particles are possible if one invokes non-thermal production mechanisms. Data on large-scale structure [49] and high-redshift Non-thermal decays of inflatons in conventional models of supernovae [50] have recently converged on the suggestion cosmological inflation could yield Qx ~ 1 for mx ~ 1013 that the energy of the vacuum may be non-zero. In my view, GeV. Preheating via the parametric resonance decay of this represents a wonderful opportunity for theoretical the inflaton could even yield Qx ~ 1 for mx ~ 1015 GeV. physics: a number to be calculated in the Theory of EveryOther possibilities include a first-order phase transition at thing including quantum gravity. The possibility that the the end of inflation, and gravitational relic production vacuum energy may be non-zero may even appear more induced by the rapid change in the scale factor in the early natural than a zero value, since there is no obvious symmetry Universe [41]. It is therefore of interest to look for possible or other reason known why it should vanish. In the above paragraph, I have used the term vacuum experimental signatures of superheavy dark matter. energy rather than cosmological constant, because it may One such possibility is offered by ultra-high-energy cosmic not actually be constant. This option has been termed rays. Those coming from distant parts of the Universe quintessence here by Steinhardt [51], who has discussed a (D>100Mpc) are expected to be cut off at an energy classical scalar-field model that is not strongly motivated 1 There has recently been progress in implementing the constraints from the ab- by the Standard Model, supersymmetry or GUTs, though something similar might emerge from string theory. I prefer sence of charge and colour-breaking minima [39]. m.„ (GeV)
1
•\f
^--™—-—"""""
.„-.-——_
" ~ ~
""^^^
'•-..
'
^~^-^
:
:
s
•
:
^
x.^
i
^i
.
.
Physica Scripta T85
:i
••
*. • i
'i
>.
I^I
.
..i
|
. i.
:
ii)(l
© Physica Scripta 2000
Particle Candidates for Dark Matter
2
21 I 16
i i |
OlvMO '1(3*V
22 -
I
> I j
.
•
J
.
Li
,
J
!
1&«
19
IBS
20
20.5
21
21.S
22
l«9iatBW
Fig. 12. The ultra-high energy cosmic ray flux compared with a model calculation based on the decays of superheavy relic particles [48].
to think that a varying vacuum energy might emerge from a quantum theory of gravity, as the vacuum relaxes towards an asymptotical value (zero?) in an infinitely large and old Universe. We have recently given [52] an example of one such possible effect which yields a contribution to the vacuum energy that decreases as l/t2. This is compatible with the high-redshift supernova data, and one may hope that these could eventually discriminate between such a possibility and a true cosmological constant.
References 1. Griest, K. Jungman, G. and Kamionkowski, M., Phys. Rept. 267, 195 (1996). 2. Dimopoulos, S. Phys. Lett. B246, 347 (1990). 3. Turner, M. S. talk at this meeting; see also astro-ph/9811364, astro-ph/9811366, astro-ph/9811447, astro-ph/9811454. 4. Kolb, E. W. talk at this meeting; see also hep-ph/9810362. 5. Particle Data Group, Particle Data Group, C. Caso et al, Eur. Phys. J. C3, 1 (1998). 6. Yanagida, T., Proc. Workshop on the Unified Theory and the Baryon Number in the Universe (KEK, Japan, 1979); Slansky, R., Talk at the Sanibel Symposium, Caltech preprint CALT-68-709 (1979). 7. Super-Kamiokande collaboration, Fukuda, Y., et al., Phys. Rev. Lett. 81, 1562 (1998). 8. Kafka, T., for the Soudan II collaboration, Nucl. Phys. Proc. Suppl. 70, 340 (1999). 9. MACRO collaboration, Ambrosio, M., et al, Phys. Lett. B434, 451 (1998). 10. Bahcall, J. N., talk at this meeting; see also astro-ph/9808162; Bahcall, J. N., Krastev, P. I. and Yu. Smirnov, A., Phys.Rev. D58, 096016 (1998). 11. E. Lisi, talk at this meeting; see also Fogli, G. L., Lisi, E., Marrone, A. and Scioscia, G., Phys. Rev. D59, 033001 (1999). 12. For some recent work, see: King, S. F., Phys. Lett. B439, 350 (1998); Elwood, J. K., Irges, N. and Ramond, P., Phys. Rev. Lett. B81, 5064 (1998); Barger, V., et al, Phys. Rev. D58, 093016 (1998); Osland, P. and Vigdel, G. Phys. Lett. B438, 129 (1998); Joshipura, A. and Smirnov, A., Phys. Lett. B439, 103 (1998); Tanimoto, M., Phys. Rev. D59, 017304 (1999); Lazarides, G. and Vlachos, N., Phys. Lett. B441, 46 (1998); Barger, V., Weiler, T. and Whisnant, K., Phys. Lett. B440, 1 (1998); Pati, J., Nucl. Phys. Rev. Supp. 77, 299 (1999); Gonzalez-Carcia, M., et al., Nucl. Phys. B543, 3 (1999); Altarelli, G. and Feruglio, F., Phys. Lett. B439, 112 (1998). 13. Ellis, J., Leontaris, G., Lola, S. andNanopoulos, D. V. Eur. Phys. J. C9, 389 (1999). © Physica Scripta 2000
229
14. However, combinations of structure-formation and cosmic microwave-background measurements may be sensitive to smaller neutrino masses: Tegmark, M., talk at this meeting; see also Hu, W., Eisenstein, D. J., Tegmark, M. and White, M., Phys. Rev. D59, 023512 (1999). 15. Tanimoto, M., Phys. Lett. B360, 41 (1995). 16. Oyama, Y., for the K2K collaboration, hep-ex/9803014. 17. MINOS collaboration, Abies, E., et al., Fermilab Proposal P-875 (1995). 18. Acquistapace, G., et al., CERN report 98-02. 19. Maiani, L., Proc. Summer School on Particle Physics, Gif-sur-Yvette, 1979 (IN2P3, Paris, 1980) p. 3; G 't Hooft, in: (G 't Hooft et al, eds.), "Recent Developments in Field Theories" (Plenum Press, New York, 1980); Witten, E., Nucl. Phys. B188, 513 (1981); Kaul, R. K., Phys. Lett. 109B, 19 (1982). 20. For a review, see: Farhi, E. and Susskind, L. Phys. Rep. 74C, 277 (1981); Dimopoulos, S. and Susskind, L., Nucl. Phys. B155 237 (1979); Eichten, E. and Lane, K. Phys. Lett. B90, 125 (1990); Peskin, M. E. and Takeuchi, T., Phys. Rev. D46, 381 (1992); Altarelli, G., Barbieri, R. and Jadach, S., Nucl. Phys. B369, 3 (1992). 21. Fayet, P. and Ferrara, S., Phys. Rep. 32, 251 (1997); Haber, H. E. and Kane, G. I., Phys. Rep. 117, 75 (1985). 22. See, e.g., Kounnas, C , Lahanas, A. B., Nanopoulos, D. V. and Quiros, M., Phys. Lett. 132B, 95 (1983). 23. Gruunewald, M. and Karlen, D., talks at International Conference on High-Energy Physics, Vancouver 1998, http://www.cern.ch/ LEPEWWG/misc. 24. Okada, Y., Yamaguchi, M. and Yanagida, T., Progr. Theor. Phys. 85,1 (1991); Ellis, J., Ridolfl, G. andZwirner, F., Phys. Lett. B257, 83, (1991) Phys. Lett. B262, (1991) 477; Haber, H. E. and Hempfling, R., Phys. Rev. Lett. 66, 1815 (1991); Barbieri, R., Frigeni, M. and Caravaglios, F. Phys. Lett. B258, 167 (1991); Okada, Y., Yamaguchi, M. and Yanagida, T., Phys. Lett. B262, 54 (1991). 25. Ellis, J., Kelley, S. and Nanopoulos, D. V., Phys. Lett. B249,441 (1990) and Phys. Lett. B260, 131 (1991); Amaldi, U., de Boer, W. and Furstenau, H., Phys. Lett. B260, 447 (1991); Langacker, P. and Luo, M., Phys. Rev. D44, 817 (1991). 26. Ellis, J., Hagelin, J. S., Nanopoulos, D. V., Olive, K. A. and Srednicki, M., Nucl. Phys. B238, 453 (1994). 27. Rich, J., Spiro, M. and Lloyd-Owen, J., Phys. Rep. 151, 239 (1987); Smith, P. F., Contemp. Phys. 29, 159 (1998); Hemmick, T. K., et al, Phys. Rev. D41, 2074 (1990). 28. Mohapatra, R. N. and Nussinov, S., Phys. Rev. D57, 1940 (1998). 29. Ellis, J., Falk, T., Olive, K. and Schmitt, M., Phys. Lett. B388,97 (1996) and Phys. Lett. B413, 355 (1997). 30. Klapdor-Kleingrothaus, H. V. and Ramachers, Y., Eur. Phys. J. A3, 85 (1998). 31. Giudice, G. and Rattazzi, R., hep-ph/9801271. 32. Ellis, J., Falk, T., Olive, K. and Schmitt, M. Phys. Lett. B413, 355 (1997). 33. LEP Experiments Committee meeting, Nov. 12th, 1998, http:www.cern.ch/Committees/LEPC/minutes/LEPC50.html. This source provides some preliminary updates of this and other experimental plots from papers contributed to the International Conference on High-Energy Physics, Vancouver 1998: http://ichep. triumf.ca / main. asp. 34. J. Ellis, K. Enqvist, D.V. Nanopoulos and F. Zwirner, Mod. Phys. Lett. Al, (1986) 57; Giudice, G. F. and Barbieri, R. Nucl. Phys. B306, (1988) 63. 35. Chankowski, P. H., Ellis, J. and Pokorski, S., Phys. Lett. B423, (1998) 327; Chankowski, P. H., Ellis, J., Olechowski, M. and Pokorski, S., Nucl. Phys. B544, 39 (1999). 36. Chankowski, P. H„ Ellis, J., Olive, K. A. and Pokorski, S., Phys. Lett. B452, 28 (1999). 37. Olive, K. A. and Srednicki, M. Phys. Lett. B230, (1989) 78 and Nucl. Phys. B355, 208 (1991); Griest, K., Kamionkowski, M., and Turner, M. S., Phys. Rev. D41, 3565 (1990). 38. Ellis, J., Falk, T. and Olive, K. A., Phys. Lett. B444, 367 (1998). 39. Abel, S. and Falk, T., Phys. Lett. B444, 427 (1998). 40. Abdullin, S. and Charles, F., Nucl. Phys. B547, 60 (1999). 41. Kolb, E. W., Chung, D. J. H. and Riotto, A., hep-ph/9810361 and references therein. 42. Greisen, K., Phys. Rev. Lett. 16 (1966) 748; Zatsepin, G. T. and Kuzmin, V. A. Pis'ma Zh. Eksp. Teor. Fiz. 4 114 (1966). Physica Scripta T85
230
John Ellis
43. Watson, A. A., talk at this meeting; Takeda, M., et al, Phys. Rev. Lett. 81, 1163 (1998) and references therein. 44. Berezinsky, V. and Mikhailov, A., Phys. Lett. B449,237 (1999); Medina Tanco, G. A. and Watson, A. A., "Dark Matter Halos and the Anisotropy of Ultra-High Energy Cosmic Rays," (November 1998). 45. Benakli, K., Ellis, J. and Nanopoulos, D. V., Phys. Rev. D59, 047301 (1999). 46. Ellis, J., Lopez, J. and Nanopoulos, D. V., Phys. Lett. B247,257 (1990); Ellis, J., Gelmini, G., Lopez, J., Nanopoulos, D. V. and Sarkar, S., Nucl. Phys. B373, 399 (1992).
Physica Scripta T85
47. Berezinsky, V., Kachelriess, M. and Vilenkin, A., Phys. Rev. Lett. 79, 4302 (1997); Berezinsky, V. and Kachelriess, M. Phys. Lett. B434, 61 (1998). 48. Birkel, M. and Sarkar, S. Astropart. Phys. 9, 297 (1998). 49. Bahcall, N., talk at this meeting; see also Bahcall, N. A. and Fan, X. H., astro-ph/9804082. 50. Riess, A. G. et al, Astron. J. 116, 1009 (1998); Perlmutter, S., et al, astro-ph/9812133. 51. Steinhardt, P. J., talk at this meeting; see also Zlatev, I., Wang, L. W. and Steinhardt, P. J., Phys. Rev. Lett. 82, 896 (1999). 52. Ellis, J., Mavromatos, N. E. and Nanopoulos, D. V., gr-qc/9810086.
© Physica Scripta 2000
Physica Scripta. Vol. T85, 231-245, 2000
Early-Universe Issues: Seeds of Perturbations and Birth of Dark Matter Edward W. Kolb NASA/Fermilab Astrophysics Center, Fermi National Accelerator Laboratory, Batavia, Illinois 60510-0500, USA and Department of Astronomy and Astrophysics, Enrico Fermi Institute, The University of Chicago, Chicago, Illinois 60637-1433, USA Received January 26, 1999; accepted August 2, 1999
PACS Ref: 98.80 Cg
Abstract The signatures of events that occurred in the early universe can be seen today on the sky in the form of large-scale structures and fluctuations in the temperature of the cosmic background radiation. If we can read the signatures, we can learn what occurred in the early universe and something about the physics involved.
1. Introduction The theme of this meeting, "Particles and the Universe" represents perhaps the most striking illustration of the true unity of science. At first glance it is not at all clear that "particles" have anything to do with "the Universe." The realm of "particle" physics is nature on the smallest scales, while cosmology is concerned with the universe on the largest scales. Although the two fields are separated by the sizes of the objects they study, they are joined by the fact that it is impossible to understand the origin and evolution of large-scale structures in the universe without understanding the "initial conditions" that led to their development. The initial data was set in the very early universe when the fundamental particles and forces acted to produce the perturbations in the cosmic density field. A complete understanding of the present structure of the universe will also be impossible without accounting for the dark component in the density field. The most likely possibility is that this ubiquitous dark component is an elementary particle relic from the early universe. The study of the structure of the present universe may reveal insights into events that occurred in the early universe, and hence, into the nature of the fundamental forces and particles at an energy scale far beyond the reach of terrestrial accelerators. Perhaps the early universe, the ultimate particle accelerator, will provide the first glimpse of physics at the scale of Grand Unified Theories (GUTs), or even the Planck scale. The fact that "particles" and "the universe" is joined, may well be best expressed by a quote from the great American naturalist and conservationist John Muir (1838-1914): When one tugs at a single thing in nature, he finds it attached to the rest of the universe. In this paper I will concentrate on two events which occurred in the early universe. The first is the generation of perturbations in the density field during an early period of rapid expansion known as cosmic inflation. The second is the genesis of dark matter. The record of these events is written in the arrangement of galaxies, galaxy clusters, © Physica Scripta 2000
and the imperfections in the isotropy of the cosmic microwave background radiation. If we really understood particle physics we could predict the nature of those patterns. If we really knew how to read the story in the structures, we would learn something about particle physics. The story is there on the sky, patiently waiting for our wits to become sharp enough to read it. 2. Inflation One of the striking features of the CBR temperature fluctuations is that they appear to be noncausal. The CBR fluctuations were largely imprinted at the time of last-scattering, about 300 000 years after the bang. However, there are fluctuations on length scales much larger than 300 000 light years! How could a causal process imprint fluctuations on scales larger than the light-travel distance since the time of the bang? The answer is inflation. Sometime during the early evolution of the universe the expansion must have been such that a > 0. From Einstein's equations, the "acceleration" is related to the energy density and the pressure as a oc —{p + 3p). Therefore, an acceleration (positive a) requires an unusual equation of state with p + 3/7 < 0. This is referred to as "accelerated expansion" or "inflation." Including the inflationary phase, our best guess for the different epochs in the history of the universe is given in Table I. There is basically nothing known about the stringy phase, if indeed there was one. The earliest phase we have information about is the inflationary phase. As we shall see, the information we have is from the quantum fluctuations during inflation, which were imprinted upon the metric, and can be observed as CBR fluctuations and the departures from homogeneity and isotropy in the matter distribution, e.g., the power spectrum. Inflation was wonderful, but all good things must end. A lot of effort has gone into studying the end of inflation (For Table I. Different epochs in the history of the universe and the associated tempos of the expansion rate. tempo
passage
age
P
P
P + 3p
prestissimo presto
string dominated vacuum dominated (inflation) matter dominated radiation dominated matter dominated
< 10"43s ~ 10"38s
?
9
9
Pv
~P\
—
~ 10" s < 104yr
HjA
0
> 10V
Pmatter
allegro andante largo
36
r
+ 7*73 + 0 +
Physica Scripta T85
232
Edward W. Kolb
a review, see Kofman et al. [1]) It was likely that there was a inflaton potential energy density, fluctuations in the inflaton brief period of matter domination before the universe field lead to fluctuations in the energy density. Because of became radiation dominated. Very little is known about this the rapid expansion of the universe during inflation, these period after inflation. Noteworthy events that might have fluctuations in the energy density are frozen into superoccurred during this phase include baryogenesis, phase Hubble-radius-size perturbations. Later, in the radiation transitions, and generation of dark matter. We do know that or matter-dominated era they will come within the Hubble the universe was radiation dominated for almost all of the radius as if they were noncausal perturbations. first 10 000 years. The best evidence of the radiationThe spectrum and amplitude of perturbations depend dominated era is primordial nucleosynthesis, which is a relic upon the nature of the inflaton potential. Mukhanov [2] of the radiation-dominated universe in the period 1 second has developed a very nice formalism for the calculation to 3 minutes. The earliest picture of the matter-dominated of density perturbations. One starts with the action for gravera is the CBR. ity (the Einstein-Hilbert action) plus a minimally-coupled One important issue is how does one imagine a universe scalar inflatonfield4>\ dominated by vacuum energy making a transition to a matter-dominated or radiation-dominated universe. A -n\j^R-[g"\'t>U+Vi4') (1) simple way to picture this is by the action of a scalar field • / * (j) with potential V(
;R
if
2.1. Quantum fluctuations In addition to the classical motion of the inflaton field, during inflation there are quantum fluctuations. Since the total energy density of the universe is dominated by the
(inflaton)
spectra and amplitudes for the density perturbations. Before proceeding, now is a useful time to remark that in addition to scalar density perturbations, there are also fluctuations in the transverse, traceless component of the spatial part of the metric. These fluctuations (known as
(inflaton)
Fig. 1. Schematic illustrations of the inflaton potential energy. The potential on the left is a "large-field" model (Type-la). The middle figure illustrates a "small-field" model (Type-lb). The figure on the right illustrates an example of hybrid inflation (Type-Ic). Physica Scripta T85
© Physica Scripta 2000
Early-Universe Issues: Seeds of Perturbations and Birth of Dark Matter tensor fluctuations) can be thought of as a background of gravitons. Although the scalar and tensor spectra depend upon V(
233
Of course a classification scheme is only reasonable if there are some observable quantities that can differentiate between different schemes. It turns out that the different Type-I models fill in different regions of the n-r plane, as shown in Fig. 2 (from [3]). For a given spectral index n, small-field models have a smaller value of r. Shown as an ellipse is a very conservative estimate of the uncertainties in n and r that are expected after the next round of satellite observations. Although we don't know where the error ellipse will fall on the graph, an error ellipse at the indicated size will restrict models. So we expect a true inflation phenomenology, where models of inflation are confronted by precision observations. 2.3. Inflation models in the era of precision cosmology It was once said that the words "precision" and "cosmology" could not both appear in a sentence containing an even number of negatives. However that statement is now out of date, or at the very least, very soon will be out of date. A number of new instruments will come on line in the next few years and revolutionize cosmology. There is now a world-wide campaign to pin down the microwave anisotropics. In the near future, long-duration balloon flights, as well as observations from the cold, dry observatory in Antarctica will completely change the situation. Finally, in the next decade two satellites, a NASA mission—the Microwave Anisotropy Probe (MAP)— and an ESA mission—PLANCK, will culminate in a determination of the spectrum with errors much smaller than the present errors. Of course we don't know what the shape of the spectrum will turn out to be, but we can anticipate errors as small as shown in the figure. Withe errors of this magnitude, fitting the spectrum will allow determination of n and r, useful for inflation, as well as determination of the cosmological parameters (Ho, Qo, QB, A, etc.) to a few percent.
2.2. Models of inflation A quick perusal of the literature will reveal many models of inflation. Some of the familiar names to be found include: old, new, middle-aged, chaotic, quixotic, ergodic, exotic, heterotic, autoerotic, natural, supernatural, au natural, power-law, powerless, power-mad, one-field, two-field, home-field, modulus, modulo, moduli, self-reproducing, self-promoting, hybrid, low-bred, white-bread, first-order, second-order, new-world order, pre-big-bang, no-big-bang, post-big-bang, D-term, F-term, winter-term, supersymmetric, superstring, superstitious, extended, hyperextended, overextended, D-brane, p-brane, No-brain, dilaton, dilettante, . . . Probably the first step in sorting out different models is a classification scheme. One proposed classification scheme has two main types. Type-I inflation models are models based on a single inflaton field, slowly rolling under the influence of an inflaton potential V{<j>). This may seem like a 2.4. Reconstruction restrictive class, but in fact many more complicated models In addition to restricting the class of inflation models (Type can be expressed in terms of an equivalent Type-I model. la, Type lb, etc.), it may be possible to use the data from For instance "extended" inflation, which is a Jordan- precision microwave experiments to reconstruct a fragment Brans-Dicke model with an inflaton field and a JBD scalar of the inflationary potential. field, can be recast as an effective Type-I model. Anything that is not Type I is denoted as a Type-II model. There are subclasses within Type I. Type-la models are "large-field" models, where the inflaton field starts large and evolves toward its minimum. Examples of large-field models are chaotic inflation and power-law inflation. Type-lb models are "small-field" models, where the inflaton field starts small and evolves to its minimum at larger values of the inflaton field. Examples of small-field models are new inflation and natural inflation. Finally, hybrid-inflation models are classified as Type-Ic models. In Type-la and Type-lb models the vacuum energy is approximately zero at the end of inflation. Hybrid models have a significant vacuum energy at the end of inflation (see Fig. 1). Hybrid inflation is usually terminated by a first-order phase transition or by the action of a second scalar field. A more accurate description of large-field and small-field potential Fig. 2. Large-field (Type 1-a) inflation models, small-field (Type-lb) inflation is the sign of the second derivative of the potential: models, and hybrid (Type-Ic) inflation models seem to populate different regions of the n-r plane. Observational determination of n and r could rule large-field models have V" > 0 while small-field models have out entire classes of models. This figure is from Dodelson, Kinney and Kolb V" < 0. [3]© Physica Scripta 2000
Physica Scripta T85
234
Edward W. Kolb
Reconstruction of the inflaton potential (see Lidsey, et al. [4] for a review) refers to the process of using observational data, especially microwave background anisotropies, to determine the inflaton potential capable of generating the perturbation spectra inferred from observations [5]. Of course there is no way to prove that the reconstructed inflaton potential was the agent responsible for generating the perturbations. What can be hoped for is that one can determine a unique (within observational errors) inflaton potential capable of producing the observed perturbation spectra. The reconstructed inflaton potential may well be the first concrete piece of information to be obtained about physics at scales close to the Planck scale. As is well known, inflation produces both scalar and tensor perturbations, and each generate microwave anisotropies (see Liddle and Lyth [6] for a review). If V{4>) is known, the perturbation spectra can be computed exactly in linear perturbation theory through integration of the relevant mode equations [7]. If the scalar field is rolling sufficiently slowly, the solutions to the mode equations may be approximated using something known as the slow-roll expansion [8-10]. The standard reconstruction program makes use of the slow-roll expansion, taking advantage of a calculation of the perturbation spectra by Stewart and Lyth [11], which gives the next-order correction to the usual lowest-order slow-roll results. Two crucial ingredients for reconstruction are the primordial scalar and tensor perturbation spectra As(k) and A-r(k) which are defined as in Lidsey et al. [4]. The scalar and tensor perturbations depend upon the behavior of the expansion rate during inflation, which in turn depends on the value of the scaler inflaton field >. In order to track the change in the expansion rate, we define slow-roll parameters e, rj, and t, as
k = aH during inflation, and where C = — 2 + In 2 + y ~ —0.73 is a numerical constant, y being the Euler constant. These equations are the basis of the reconstruction process. From the expressions for A\ and A\ one can express the scalar and tensor spectral indices, defined as
ns
d l n
1
1 , ( 0 SE
~ An \_H(
«*) = An V
(
H"{
A\{k)
25n \mV\) e - [ l - ( 2 C + l ) e + C f j f
4
/in
A&)--ziz[-zr\ [l-(C+i)e]z 257i ywpi,
(9) where «* = n(k*), to the observed spectrum in order to extract the coefficients, where stars indicate the value at k*. The scale k* is most wisely chosen to be at the (logarithmic) center of the data, about k* = 0.01 h M p c - 1 . For reconstruction, one takes a Hamilton-Jacobi approach where the expansion rate is considered fundamental, and the expansion rate is parameterized by a value of the inflaton field. The Friedmann equation may be expressed in terms of the potential V(
12TC
H1 =
32n2
-V&)
(10)
or equivalently in terms of the slow-roll parameter £, as
(5)
(6)
(V)
where H, e, and r\ are to be determined at the value of <£ when Physica Scripta T85
ln3A + ..
(4)
As long as the slow-roll parameters are small compared to unity, the scalar and tensor perturbation amplitudes As(k) and AT(k) are given by (see Stewart and Lyth [11] for the normalization) 1
(8)
2 1 d« , ik 1 d« In 2 —+ 2 d In k K 6d(lnA:)2
, k In k*
m,
where here the prime superscript implies d/d^>. If the slow-roll parameters are small, e, rj, and £ can be expressed in terms of derivatives of the inflaton potential: m2Pl V" ^^Jn-V-
dln^ din A:
\n Al(k) = In A2s(kt) + (n. -1)
'PI
1/2
nT(k) =
in terms of slow-roll parameters. Perturbative reconstruction requires that one fits an expansion, usually a Taylor series of the form
4
, - ^ ( 3 - . , .
(11)
Subsequent derivatives of V may be expressed in terms of additional slow-roll parameters:
V' = -
^
\JAn
H2B^2(3
-n),
V" =H2(3e + 3ri-ri2 + Z2) . (12)
and so on. So if one can determine the slow-roll parameters, one has information about the potential. The slow-roll parameters needed to reconstruct a given derivative of the potential is given in Table II. Of course, the problem is that the slow-roll parameters are not directly observable! But one can construct an iterative scheme to express the slow-roll parameters in terms of observables. The result is shown in Table III. © Physica Scripta 2000
Early-Universe Issues: Seeds of Perturbations and Birth of Dark Matter
235
However if a tensor mode can de determined, then one Table II. A summary of the slow-roll parameters, H, e, r\, and £,, needed to reconstruct a given derivative of the potential to a may follow the following reconstruction procedure. /?*, certain order. Note that the next-order result is exact. «*, n't, ... are to be determined from observations. Fortunately, parameter estimation from the microwave next-order (exact) lowest-order background has been explored in some detail [12,13]. We shall use error estimates for Planck assuming polarized detectors are available, following the analysis of Zaldarriaga W.).«(<M.M.) et al. [13]. Most analyses have assumed that R* and «» are the only parameters needed to describe the spectra. In a recent paper [14], Copeland, Grivell, and Liddle have generalized their treatment to allow the power spectrum to deviate from Table III. The inflation parameters may be expressed in scale-invariance. Including extra parameters leads to a terms of the observables, A\, A\, n, and d«/dln k. deterioration in the determination of all the parameters, as it introduces extra parameter degeneracies. Fortunately, lowest-order next-order parameter for most parameters the uncertainty is not much increased A1 A1 H A\ by including the first few derivatives of n [14], but the par^T'^*S e AJ, As A~r , Ac » ft ameter n itself has a greatly increased error bar. If a A\,A\,n A^,A\,n,dn/d\nk 1 power-law is assumed it can be determined to around A\, A\,n,dn/d\nk { — An — 0.004 [13,14], but including scale dependence increases this error bar by a factor of ten or more. Notice that unless Table IV. A summary of the observables needed to one assumes a perfect power-law behavior, this increase reconstruct a given derivative of the potential to a certain in uncertainty is applicable even if the deviation from power-law behavior cannot be detected within the order. uncertainty. lowest-order next-order next-to-next-order From Grivell and Liddle [7], an estimate of the relevant uncertainties is 1 A V R,A\ V V" ym
R,A\ R,Al,n R,A\,n,dn/d
R,A\,n R,A\,n,dn/d\ak
Rm,A\,n,dn/d\nk
A(R*)~ 0.004 ; A(dn/d\nk)
In it
2
~ 0.04,
(14)
2
An~ 0.15 ; zl[d n/d(ln)t) ] ~ 0.005.
Using the notation ri^ to indicate d n / d l n ^ and Rt = Aj(k*)/Al(k*), the reconstruction equations are
Once R*, A\, «*, and n'^ are determined, then it is possible to find F((/>J, V'(4>J, and K"(>J from Eq. (13). Then one can express V(4>) as a Taylor series about F ( 0 J : V{4>) = V<4>.) + V'{
™n V(<j>.) ^ § 4 ( * * ) * * { 1 + [0.21.R.]},
+••
(15)
A(j) is found from an exact expression connecting changes in (f> with changes in k [4], x { l -[0.85J?.-0.53(1-/!„)]}, mp 2 V"(J>J ~ — A2s(k.)R.{9R.
- 1.5(1 - «,)
(13) dj> dink
n^H' 1 An H e-\
_ i-^ ~ S 4nK
, *''
a + R
(16)
-[24.3RI+ 0.25(1 - n , ) 2 - 14.8^(1 - 0 - 1 . 6 < ] } .
Let me illustrate reconstruction by considering two sample potentials. The first potential, discussed by Lidsey et al. [4] is a power-law potential, V{4>) = VQ exp(—a.
Physica Scripta T85
236
Edward W. Kolb
? 3 >*
-0.1 0 0.1 (0-*o)/ m pi
(*-0o)/ m Pi
>
0.2
0.3
0.4
-3-4 -
-0.4
-0.2
0 */m pl
0.2
0.4
Fig. 3. Sample reconstructions of two inflationary potentials. The upper two figures are the reconstruction of a power-law potential of the form V(4>) = VQ exp(—a^/mpi). The light curves are from the true potential, while the heavy curves are ten reconstructions. The lower figures are the actual potential and ten reconstructions of a potential of the form V((f>) = A4[l — (2/ji)tan~'(5<£/mpi)].
Also shown in Fig. 3 in the upper right-hand panel is information about the reconstruction of the first derivative of the potential. The second example of reconstruction was considered in Copeland et al. [15] The potential, first considered by Wang et al. [16] is a potential of the form (17) which is shown in the lower left-hand side of Fig. 3. A fit to a Taylor expansion to the exact spectrum, using k* as above, yields the following results: ln,4|(Jfc,) = 3.84 dn dlnfc
tf* = 0.014 d2«
;
n, = 0.579
= -0.134-d(\n ky = -0.052 .
(18)
From the estimated observational uncertainties of Eq. (14), we see that all these coefficients should be successfully determined at high significance and a simple "chi-by-eye" demonstrates that the spectrum reconstructed from these data is an adequate fit to the observed spectrum. This stresses the point that the more unusual a potential is, the more information one is likely to be able to extract about it, though the uncertainties on the individual pieces of information may be greater. We reconstruct the potential in a region about $„ = —0.22 (the reconstruction program does not determine $_), the width of the region given via Eq. (16). The result for the reconstructed potential including the effect of observational Physica Scripta T85
errors and cosmic variance are shown in Fig. 3, where it can immediately be seen that the reconstruction has been very successful in reproducing the main features of the potential while perturbations on interesting scales are being developed. The uncertainty is dominated by that of Aj(kt); although the gravitational waves are detectable in this model, it is only about a three-sigma detection and the error bar is thus large. Since the overall magnitude of the potential is proportional to Aj, the visual impression is of a large uncertainty. Fortunately, information in combinations of the higher derivatives is more accurately determined. Figure 3 also shows the reconstruction of V'/Vi/2 with observational errors; this combination is chosen as it is independent of the tensors to lowest order. Not only is it reconstructed well at the central point, but both the gradient and curvature are well fit too, confirming useful information has been obtained about not just V" but V" as well, which is only possible because of the extra information contained in the scale-dependence of the power spectrum. So rather accurate information is being obtained about the potential. 3. Dark matter There are many reasons to believe the present mass density of the universe is dominated by a weakly interacting massive particle (WIMP), a fossil relic of the early universe. Theoretical ideas and experimental efforts have focused mostly on production and detection of thermal relics, with mass typically in the range a few GeV to a hundred GeV. Here, I will review scenarios for production of nonthermal dark © Physica Scripta 2000
WIMPS
It is usually assumed that the dark matter consists of a species of a new, yet undiscovered, massive particle, traditionally denoted by X. It is also often assumed that the dark matter is a thermal relic, i.e., it was in chemical equilibrium in the early universe. A thermal relic is assumed to be in local thermodynamic equilibrium (LTE) at early times. The equilibrium abundance of a particle, say relative to the entropy density, depends upon the ratio of the mass of the particle to the temperature. Define the variable Y = nx/s, where nx is the number density of WIMP X with mass Mx, and s ~ T 3 is the entropy density. The equilibrium value of Y, YEQ, is proportional to exp(—x) for x » 1, while F E Q ~ constant for x <SC 1, where x = Mx/T. A particle will track its equilibrium abundance as long as reactions which keep the particle in chemical equilibrium can proceed rapidly enough. Here, rapidly enough means on a timescale more rapid than the expansion rate of the universe, H. When the reaction rate becomes smaller than the expansion rate, then the particle can no longer track its equilibrium value, and thereafter Y is constant. When this occurs the particle is said to be "frozen out." A schematic illustration of this is given in Fig. 4. The more strongly interacting the particle, the longer it stays in LTE, and the smaller its eventual freeze-out abundance. Conversely, the more weakly interacting the particle, the larger its present abundance. The freeze-out value of Y is related to the mass of the particle and its annihilation cross section (here characterized by
1 Mxmpioo
(19)
Since the contribution to Q is proportional to Mxnx, which in turn is proportional to MXY, the present contribution to Q from a thermal relic roughly is independent of its mass, 1 and depends only upon the annihilation cross section. The cross section that results in Qxh1 ~ 1 is of order 10~37cm2, of the order the weak scale. This is one of the attractions of thermal relics. The scale of the annihilation cross section is related to a known mass scale. The simple assumption that dark matter is a thermal relic is surprisingly restrictive. The largest the annihilation cross section possible is roughly Mx2. This implies that large-mass wimps would have such a small annihilation cross section that their present abundance would be too large. Thus one expects a maximum mass for a thermal WIMP, which turns out to be a few hundred TeV [18]. The standard lore is that the hunt for dark matter should concentrate on particles with mass of the order of the weak scale and with interaction with ordinary matter on the scale of the weak force. This has been the driving force behind 1 To first approximation the relic dependence depends upon the mass only indirectly through the dependence of the annihilation cross section on the mass.
© Physica Scripta 2000
o
1 , , , , 1 , , . .
-5 -
x >--10 3.1. Thermal relics—Wimpy
•
. . . .
matter. Since the masses of the nonthermal WIMPS are in the range 1012 to 1016 GeV, much larger than the mass of thermal wimpy WIMPS, they may be referred to as WIMPZILLAS. In searches for dark matter it may be well to remember that "size does matter."
237 I I I '
Early-Universe Issues: Seeds of Perturbations and Birth of Dark Matter
freeze o u t -> V \
1
^ _ ,„-> n x h
Y
\
-15-
V Y EQ
•
-20
10 2
10'
103
Mx/T Fig. 4. A thermal relic starts in LTE at T » Afx.When the rates keeping the relic in chemical equilibrium become smaller than the expansion rate, the density of the relic relative to the entropy density freezes out.
the vast effort in dark matter direct detection described in this meeting by Sadoulet. In view of the unitarity argument, in order to consider thermal WIMPZILLAS, one must invoke, for example, late-time entropy production to dilute the abundance of these supermassive particles [19], rendering the scenario unattractive. 3.2. Nonthermal
relics—WIMPZILLAS
There are two necessary conditions for the WIMPZILLA scenario. First, the WIMPZILLA must be stable, or at least have a lifetime much greater than the age of the universe. This may result from, for instance, supersymmetric theories where the breaking of supersymmetry is communicated to ordinary sparticles via the usual gauge forces [20]. In particular, the secluded and the messenger sectors often have accidental symmetries analogous to baryon number. This means that the lightest particle in those sectors might be stable and very massive if supersymmetry is broken at a large scale [21]. Other natural candidates arise in theories with discrete gauge symmetries [22] and in string theory and M theory [23,24]. It is useful here to note that wimpzilla decay might be able to account for ultra-high energy cosmic rays above the Greisen-Zatzepin-Kuzmin cutoff [25,26]. A wimpy little thermal relic would be too light to do the job, a WIMPZILLA is needed. The second condition for a WIMPZILLA is that it must not have been in equilibrium when it froze out (i.e., it is not a thermal relic), otherwise Qxh2 would be much larger than one. A sufficient condition for nonequilibrium is that the annihilation rate (per particle) must be smaller than the expansion rate: nx^lv] < H, where
,
(20)
where 7o is the present temperature. Using the fact that pJTJ = H(T*)MPlT2, nx(T*)/H(T*) = (Qx/Qy)T0MmTJ Physica Scripta T85
238
Edward W. Kolb
Mx- One may safely take the limit
\
Mx
)
\MX)
(21)
<1
This implies that if a nonrelativistic particle with Mx k, 200 TeV was created at T* < Mx with a density low enough to result in Qx < 1, then its abundance must have been so small that it never attained equilibrium. Therefore, if there is some way to create WIMPZILLAS in the correct abundance to give Qx ~ 1, nonequilibrium is automatic. Examples of WIMPZILLA evolution and freezeout are shown in Fig. 5. Any WIMPZILLA production scenario must meet these two criteria. Before turning to several WIMPZILLA production scenarios, it is useful to estimate the fraction of the total energy density of the universe in WIMPZILLAS at the time of their production that will eventually result in Q ~ 1 today. The most likely time for WIMPZILLA production is just after inflation. The first step in estimating the fraction of the energy density in WIMPZILLAS is to estimate the total energy density when the universe is "reheated" after inflation. Consider the calculation of the reheat temperature, denoted as TRn. The reheat temperature is calculated by assuming an instantaneous conversion of the energy density in the inflaton field into radiation when the decay width
of the inflaton energy, T^,, is equal to H, the expansion rate of the universe. The reheat temperature is calculated quite easily [17]. After inflation the inflaton field executes coherent oscillations about the minimum of the potential. Averaged over several oscillations, the coherent oscillation energy density redshifts as matter: p^ oc a" 3 , where a is the RobertsonWalker scale factor. If pl and a\ denotes the total inflaton energy density and the scale factor at the initiation of coherent oscillations, then the Hubble expansion rate as a function of a is
(22)
V 3 m„, \al
Equating H(a) and r$ leads to an expression for a\/a. Now if all available coherent energy density is instantaneously converted into radiation at this value of a\/a, one can define the reheat temperature by setting the coherent energy density, p^ — p,(ai/a) 3 , equal to the radiation energy density, pR = (n2/30)gtTRB, whereg* is the effective number of relativistic degrees of freedom at temperature TRH- The result is
TRH
= (g^T~)
V^V"PT
= 0-2(
)
V^V^PI
•
(23)
The limit from gravitino overproduction is 7/RH < 109 to 1010 GeV. Now consider the WIMPZILLA density at reheating. Suppose the WIMPZILLA never attained LTE and was nonrelativistic at the time of production. The usual quantity Qxh2 associated with the dark matter density today can be related to the dark matter density when it was produced. First write Px(to) PR('U)
x = Mx/T Fig. 5. A nonthermal relic typically has abundance much less than LTE at T » Mx. Here are two examples of the evolution of a nonthermal relic. In the left panel Y < YEQ at freezeout, while in the right panel Y » YEQ at freezeout. Again, the solid curve is the LTE abundance while the dashed curve is the actual abundance. Physica Scripta T85
=
PXPRH) / ? R H \
PR(tRH)\T0
J'
(24)
where p R denotes the energy density in radiation, px denotes the energy density in the dark matter, 7RH is the reheat temperature, 7o is the temperature today, to denotes the time today, and JRH denotes the approximate time of reheating. 2 To obtain PX('RH)/PR(*RH), one must determine when X particles are produced with respect to the completion of reheating and the effective equation of state between X production and the completion of reheating. At the end of inflation the universe may have a brief period of matter domination resulting either from the coherent oscillations phase of the inflaton condensate or from the preheating phase [27]. If the X particles are produced at time t = tt when the de Sitter phase ends and the coherent oscillation period just begins, then both the X particle energy density and the inflaton energy density will redshift at approximately the same rate until reheating is completed and radiation domination begins. Hence, the ratio of energy densities preserved in this way until the time of radiation 2
More specifically, this is approximately the time at which the universe becomes radiation dominated after inflation. © Physica Scripta 2000
Early-Universe Issues: Seeds of Perturbations and Birth of Dark Matter domination is PA-ORH) _ 8n PRORH)
px(te) 2
3 m xH2{te)
(25)
where m?\ w 1019 GeV is the Planck mass and most of the energy density in the universe just before time ?RH is presumed to turn into radiation. Thus, using Eq. 24, one may obtain an expression for the quantity Qx = PA-OO)/PCOO), where pc(t0) = ^Hgirip^Sn and H0 = \00h km s _ 1 Mpc - 1 : nx(te) \ T0 ) 3 \m?\) \mnHHtey
(26)
Here Q^h2 « 4.31 x 10~5 is the fraction of critical energy density in radiation today and nx is the density of X particles at the time when they were produced. Note that because the reheating temperature must be much greater than the temperature today (TRH/To> 4.2 x 1014), in order to satisfy the cosmological bound &xh2<\, the fraction of total WIMPZILLA energy density at the time when they were produced must be extremely small. One sees from Eq. (26) that Qxh2 ~ 1017 (TRH/l09GeV) (ftx(tt)/p(te)). It is indeed a very small fraction of the total energy density extracted in WIMPZILLAS.
This means that if the WIMPZILLA is extremely massive, the challenge lies in creating very few of them. Gravitational production discussed in Section 3.3.1 naturally gives the needed suppression. Note that if reheating occurs abruptly at the end of inflation, then the matter domination phase may be negligibly short and the radiation domination phase may follow immediately after the end of inflation. However, this does not change Eq. (26). 3.3.
WIMPZILLA
3.3.1. Gravitational
production Production
First consider the possibility that WIMPZILLAS are produced in the transition between an inflationary and a matterdominated (or radiation-dominated) universe due to the "nonadiabatic" expansion of the background spacetime acting on the vacuum quantum fluctuations [28]. The distinguishing feature of this mechanism is the capability of generating particles with mass of the order of the inflaton mass (usually much larger than the reheating temperature) even when the particles only interact extremely weakly (or not at all) with other particles and do not couple to the inflaton. They may still be produced in sufficient abundance to achieve critical density today due to the classical gravitational effect on the vacuum state at the end of inflation. More specifically, if 0.04 <, Mx/H\ < 2, where i/i ~ m^, ~ 10 13 GeV is the Hubble constant at the end of inflation (m^, is the mass of the inflaton), WIMPZILLAS produced gravitationally can have a density today of the order of the critical density. This result is quite robust with respect to the "fine" details of the transition between the inflationary phase and the matter-dominated phase, and independent of the coupling of the WIMPZILLA to any other particle. Conceptually, gravitational WIMPZILLA production is similar to the inflationary generation of gravitational perturbations that seed the formation of large scale structures. In the usual scenarios, however, the quantum © Physica Scripta 2000
239
generation of energy density fluctuations from inflation is associated with the inflaton field that dominated the mass density of the universe, and not a generic, sub-dominant scalar field. Another difference is that the usual density fluctuations become larger than the Hubble radius, while most of the WIMPZILLA perturbations remain smaller than the Hubble radius. There are various inequivalent ways of calculating the particle production due to interaction of a classical gravitational field with the vacuum (see for example [29-31]). Here, I use the method of finding the Bogoliubov coefficient for the transformation between positive frequency modes defined at two different times. For Mx/Hi < 1 the results are quite insensitive to the differentiability or the fine details of the time dependence of the scale factor. For 0.04 < Mx/Hi < 2, all the dark matter needed for closure of the universe can be made gravitationally, quite independently of the details of the transition between the inflationary phase and the matter dominated phase. Start with the canonical quantization of the X field in an action of the form (with metric ds2 = At2 — a2(t)dx2 = a2(rj)[drj2 — dx 2 ] where n is conformal time)
7<"/ A:
l
x
(VA-)2
•M2XX2-
£RX2
(27)
where R is the Ricci scalar. After transforming to conformal time coordinate, use the mode expansion
X(X) =
/(2^^K We *'* +a>^~ikX\
(28)
where because the creation and annihilation operators obey the commutator [a^, a\ ] = <5(3)(£i — £2), the hkS obey a normalization condition hktik* — h'knt = l t o satisfy the canonical field commutators (henceforth, all primes on functions of r\ refer to derivatives with respect to n). The resulting mode equation is (29)
**(»/) +"£(»/)A*fo) = 0, where wl = k2 + Mxo2 + (6f -
\)d'/a
(30)
The parameter £, is 1/6 for conformal coupling and 0 for minimal coupling. From now on, £ = 1/6 for simplicity but without much loss of generality. By a change in variable n -> k/a, one can rewrite the differential equation such that it depends only on H{r\), H'(n)/k, k/a(n), and Mx. Hence, the parameters H\ and a\ correspond to the Hubble parameter and the scale factor evaluated at an arbitrary conformal time nI, which can be taken to be the approximate time at which Xs are produced (i.e., n1 is the conformal time at the end of inflation). One may then rewrite Eq. (29) as
h'm+
(*+%*)h(n) = 0 ,
(31)
where fj = r\a\H\, a. = a/a\, and k = k/{a\H\). For simplicity of notation, drop all the tildes. This differential equation can be solved once the boundary conditions are supplied. Physica Scripta T85
Edward W. Kolb
Pxini) = Mxnx(rii)
=
1
MxHl
U
"(>7i).
, M * =0. 1
io-« ! io-«
••
ri/\a
!.... 10- 3 Of
-*l0-4
2TT 2
^Iftl ,
1
A•
r ( / ~-
r
1
r
(32)
where one should note that the number operator is defined at r\x while the quantum state (approximated to be the vacuum state) defined at n0 does not change in time in the Heisenberg representation. As one can see from Eq. (31), the input parameter is Mx/H\. One must also specify the behavior of a(r\) near the end of inflation. In Fig. 6 (from [28]), I show the resulting values of Qxh2 as a function of Mx/H\ assuming the evolution of the scale factor smoothly interpolates between exponential expansion during inflation and either a matter-dominated universe or radiation-dominated universe. The peak at Mx/H\ ~ 1 is similar to the case presented in Ref. [32]. As expected, for large Mx/Hi, the number density falls off faster than any inverse power of
«
it J
1
10- 6 r 2
m
•m-6
•10
,
, .\
l
(•
t iMfiil '
The number density of the WIMPZILLAS is found by a Bogoliubov transformation from the vacuum mode solution with the boundary condition at n = tj0 (the initial time at which the vacuum of the universe is determined) into the one with the boundary condition at n = nx (any later time at which the particles are no longer being created). n0 will be taken to be —oo while rjl will be taken to be at +00. Defining the Bogoliubov transformation as tip (n) = afcAjJ0 (n) +Pkfl*k°(7l) ( t n e superscripts denote where the boundary condition is set), the energy density of produced particles is
F-F-^
1 111111I
240
. . . . > < !
,
v/v,
10
, . , ,
Fig. 7. The evolution of the Bogoliubov coeflScient with conformal time for several wavenumbers. r\ = rjl corresponds to the end of the inflationary era.
lived) will these particles give contribution of the order of critical density. 3.3.2. Production during reheating
„
Another attractive origin for WIMPZILLAS is during the defrosting phase after inflation. It is important to recall that it is not necessary to convert a significant fraction of the available energy into massive particles; in fact, it must be an infinitesimal amount. I now will discuss how particles Now most of the action occurs around the transition from of mass much greater than TRH may be created in the correct inflation to the matter-dominated or radiation-dominated amount after inflation in reheating [33]. In one extreme is the assumption that the vacuum energy universe. This is shown in Fig. 7. Also from Fig. 7 one of inflation is immediately converted to radiation resulting can see that most of the particles are created with in a reheat temperature 7 R H . In this case Qx can be calculwavenumber of order H\. ated by integrating the Boltzmann equation with initial conTo conclude, there is a significant mass range (0. \H\ to H\, 13 dition Nx = 0 at T = r R H . One expects the X density to be where H\ ~ 10 GeV) for which WIMPZILLAS will have suppressed by exp(—2M x /T R u); indeed, one finds Qx ~ 1 critical density today regardless of the fine details of the for M X / T R H ~ 25 + 0.5 ln(m^-(o-| v|», in agreement with pretransition out of inflation. Because this production mechan9 ism is inherent in the dynamics between the classical vious estimates [25] that for TRH10 ~ 10 GeV, the WIMPZILLA gravitational field and a quantum field, it needs no fine mass would be about 2.5 x 10 GeV. A second (and more plausible) scenario is that reheating is tuning of field couplings or any coupling to the inflaton field. not instantaneous, but is the result of the slow decay of the However, only if the particles are stable (or sufficiently long inflaton field. The simplest way to envision this process is if the comoving energy density in the zero mode of the inflaton decays into normal particles, which then scatter •"'""•"'"T « i\r — and thermalize to form a thermal background. It is usually "3. matter assumed that the decay width of this process is the same T= H/2ti f 10 s Ihermaliaation 0 as the decay width of a free inflaton field. ; region ' \ 108 There are two reasons to suspect that the inflaton decay X"" 3T 'V~, width might be small. The requisite flatness of the inflaton potential suggests a weak coupling of the inflaton field to r 0 other fields since the potential is renormalized by the inflaton 0 coupling to other fields [34]. However, this restriction may be 2 1 evaded in supersymmetric theories where the nonrenormalif \ fc zation theorem ensures a cancelation between fields and "1 io-> W their superpartners. A second reason to suspect weak ....Li a coupling is that in local supersymmetric theories gravitinos 1C- 3 1 101 io- 2 io-> / are produced during reheating. Unless reheating is delayed, Mx/H, gravitinos will be overproduced, leading to a large undesired Fig. 6. The contribution of gravitationally produced WIMPZILLAS to Qxh2 as entropy production when they decay after big-bang a function of Mx/H\. The shaded area is where thermalization may occur if nucleosynthesis [35]. the annihilation cross section is its maximum value. Also shown is the conIt is simple to calculate the WIMPZILLA abundance in the tribution assuming that the WIMPZILLA is present at the end of inflation slow reheating scenario. It will be important to keep in mind with a temperature T = H\/2n. " " • i
•
1
Physica Scripta T85
.
. . . — 1
© Physica Scripta 2000
Early-Universe Issues: Seeds of Perturbations and Birth of Dark Matter that what is commonly called the reheat temperature, 7RH, is not the maximum temperature obtained after inflation. The maximum temperature is, in fact, much larger than 7RHThe reheat temperature is best regarded as the temperature below which the universe expands as a radiation-dominated universe, with the scale factor decreasing as g~ 1/3 T~x. In this regard it has a limited meaning [17,36]. One implication of this is that it is incorrect to assume that the maximum abundance of a massive particle species produced after inflation is suppressed by a factor of exp(-M/7"RH). To estimate WIMPZILLA production in reheating, consider a model universe with three components: inflaton field energy, p^, radiation energy density, pR, and WIMPZILLA energy density, px. Assume that the decay rate of the inflaton field energy density is 7"^. Also assume the WIMPZILLA lifetime is longer than any timescale in the problem (in fact it must be longer than the present age of the universe). Finally, assume that the light degrees of freedom are in local thermodynamic equilibrium. With the above assumptions, the Boltzmann equations describing the redshift and interchange in the energy density among the different components is
•
^AM
Px +
(
r
lHPx
[P\-{PIQ)2
= 0,
(33)
2
A- (P?) ]=O.
M>
where dot denotes time derivative. As already mentioned, (c|v|) is the thermal average of the X annihilation cross section times the Moller flux factor. The equilibrium energy density for the X particles, pxQ, is determined by the radiation temperature, T = (30pR/n2gt)1/A. It is useful to introduce two dimensionless constants, a^ and ax, defined in terms of TA, and (c|v|> as oc^Mf,
(
(34)
For a reheat temperature much smaller than M^, r^ must be small. From Eq. (32), the reheat temperature in terms of ax and Mx is r R H ^ a^M+Mn. For M+ = 1013 GeV, a 0 must be smaller than of order 10~13. On the other hand, ax may be as large as of order unity, or it may be small also. It is also convenient to work with dimensionless quantities that can absorb the effect of expansion of the universe. This may be accomplished with the definitions * = p
a
;
R
= pRa
The constants c\, c2, and c^ are given by M^0LX
c2 = c\ M a
x
;
= Px^x
^ •
XEQ is the equilibrium value of X, given in terms of the temperature T as (assuming a single degree of freedom for the X species) M3 / 1 \ 3 / 2
/ T
1/4
T M~x
U«v
M* RV* Mx x
3 m
l\ Hi 3
1
SnMlM2
(40)
Xl
The numerical value of x\ is irrelevant. Before numerically solving the system of equations, it is useful to consider the early-time solution for R. Here, early times means H » r^, i.e., before a significant fraction of the comoving coherent energy density is converted to radiation. At early times 0 ~
R~\cx(xV2-x\l2y\12
(//»/».
-^
T M '<*>
(H
(^\1/\w(*i\"\(±y3*-(±x4
\K2g*)
l
\X\)
\Xl)
R' = cx
.
X' = -c3
,X~2 (X2 JcPx + R V
© Physica Scripta 2000
0 + c2
.X
(x2 - X2n), 2
EX Q
). /
(36)
\Xi)
1/4
(42)
» /».
Thus, T has a maximum value of
wM,=o.7,(J|)" , ,r(|)" , 1/4
9 yvm2lHl
1/4
2«W
1/4
(43)
Ml
which is obtained at x/x\ = (8/3) 2/5 = 1.48. It is also possible to express a^ in terms of 7RH and obtain
TRH
•s/
(41)
Now express T in terms of R to yield the early-time solution for T:
= 0.77a
<*>,
(39)
It is straightforward to solve the system of equations in Eq. (47) with initial conditions at x = x\ of R{x\) = X{x\) = 0 and $(xi) = $i. It is convenient to express p^{x = x\) in terms of the expansion rate at x\, which leads to
-M^rm 1/8
- c i
<38>
The temperature depends upon R and g*, the effective number of degrees of freedom in the radiation:
J- MAX <*>' =
\3/2
' " • - s f l s ) * f e ) -*-"*">•
(35)
It is also convenient to use the scale factor, rather than time, for the independent variable, so one may define a variable x = aMj,. With this choice the system of equations can be written as (prime denotes d/dx)
c3 =c2 M x (37)
*
i>4, + ^Hp^ + r^p^ = 0,
241
1/4
(44)
For an illustration, in the simplest model of chaotic inflation Hi^Mj, with M^,~10 1 3 GeV, which leads to TMAX/ 3 TRH ~ 10 (200/g,) 1/8 for r R H = 10 9 GeV. We can see from Eq. (41) that for x/xj > 1, in the early-time regime T scales as a~3/8, which implies that Physica Scripta T85
242
Edward W. Kolb
i
entropy is created in the early-time regime [36]. So if one is spectrum, may be produced in decays of superheavy producing a massive particle during reheating it is necessary long-living particles [24-26,37]. In order to produce cosmic to take into account the fact that the maximum temperature rays of energies larger than about 1013 GeV, the mass of is greater than 7RH, and that during the early-time evol- the A'-particles must be very large, Mx ^ 1013 GeV and their ution, T oc a~3/8. lifetime zx cannot be much smaller than the age of the An example of a numerical evaluation of the complete sys- Universe, T ^ ^ I O 1 0 yr. With the smallest value of the tem in Eq. (36) is shown in Fig. 8 (from [33]). The model lifetime, the observed flux of ultra-high energy cosmic rays parameters chosen were M^ = 1013 GeV, a^ = 2 x 10~13, will be reproduced with a rather low density of A'-particles, 12 M A - = 1.15 x 10 12 GeV, a x = 10" 2 , and g* = 200. The Qx ~ 10~ . It has been suggested that A'-particles can be expansion rate at the beginning of the coherent oscillation produced in the right amount by usual collisions and decay period was chosen to be Hi = M$. These parameters result processes taking place during the reheating stage after inflation if the reheat temperature never exceeded Mx [37]. in 7 R H = 109 GeV and Qxh2 = 0.3. Figure 8 serves to illustrate several aspects of the problem. Again, assuming naively that that the maximum number Just as expected, the comoving energy density of <j> (i.e., density of a massive particle species X produced after 3/2 a 3 p^) remains roughly constant until f^ ~ H, which for inflation is suppressed by a factor of ( M ^ / 7 R H ) exp(—Mx/T u) with respect to the photon number density, R the chosen model parameters occurs around a/a\ ~ 5 x 108. But of course, that does not mean that the tempera- one concludes that the reheat temperature TRH should be 11 15 ture is zero. Notice that the temperature peaks well before in the range 10 to 10 GeV [25]. This is a rather high value "reheating." The maximum temperature, TMAX = 1012 GeV, and leads to the gravitino problem in generic superis reached at a/ai slightly larger than unity (in fact at symmetric models. This is one reason alternative production a/a\ = 1.48 as expected), while the reheat temperature, mechanisms of these superheavy A'-particles have been pro7RH = 109 GeV, occurs much later, around aja\ ~ 108. Note posed [28,38,39]. However, our analysis show that the situation is much more promising. Making use of Eq. (45), that TMAX ~ 10 3 7RH in agreement with Eq. (44). From the figure it is clear that X «; AfEQ at the epoch of the right amount of A'-particles to explain the observed freeze out of the comoving X number density, which occurs ultra-high energy cosmic rays is produced for around a/a\ ~ 102. The rapid rise of the ratio after freeze 3/14 / Mx out is simply a reflection of the fact that X is constant while / J R H \ _ /^*_\ (46) 10 15 U0 GeV/ V200/ Vl0 GeV A"EQ decreases exponentially. A close examination of the behavior of T shows that after where it has been assumed that (er|v|) ~ My . Therefore, the sharp initial rise of the temperature, the temperature particles as massive as 1015 GeV may be generated during decreases as a~3/8 [as follows from Eq. (42)] until the reheating stage in abundances large enough to explain H ~ f^, and thereafter T oc crl as expected for the the ultra-high energy cosmic rays even if the reheat temperaradiation-dominated era. ture satisfies the gravitino bound. For the choices of M$, a.^, g*, and ax used for the model illustrated in Fig. 8, Qxh2 = 0.3 for Mx = 1.15 x 1012 GeV, in excellent agreement with the mass predicted by using 3.3.3. Production during preheating an analytic estimate for the result [33] Another way to produce WIMPZILLAS after inflation is in a preliminary stage of reheating called "preheating" [27], where nonlinear quantum effects may lead to an extremely effective dissipational dynamics and explosive particle production. Here again, the results have also important implications for Particles can be created in a broad parametric resonance the conjecture that ultra-high cosmic rays, above the with a fraction of the energy stored in the form of coherent Greisen-Zatsepin-Kuzmin cut-off of the cosmic ray inflaton oscillations at the end of inflation released after only a dozen oscillation periods. A crucial observation for our 13 ,s z discussion is that particles with mass up to 1015 GeV may M„=10 GeV a=2xl0M,= 1.15xlO GeV a y =10" be created during preheating [38,40,41], and that their dis- ' ""•", ' """'1 ' ''"'"1 ' ••"'.' ( a / a , ) 3 [ p t ( a ) / p , ( a , ) J tribution is nonthermal. If these particles are stable, they 1 ••... may be good candidates for WIMPZILLAS [42]. 2:ip.< ( a r b i t r a r y u n i t s ) /\ 10"' The main ingredient of the preheating scenario introduced / / in the early 1990s is the nonperturbative resonant transfer of / \ energy to particles induced by the coherently oscillating lO•• / inflaton fields. It was realized that this nonperturbative / \ i 3 2 mechanism can be much more efficient than the usual n x h = 0.3 io\ s perturbative mechanism for certain parameter ranges of T EH =10 GeV i '• i the theory [27]. The basic picture can be seen as follows. Sup/ io| •1 a4PP. / rats) / a r b i t r a r y u pose there is a scalar field X with a coupling g^-^X2 where <j> r is a homogeneous classical inflaton field. The mode equation 10 ""1 10' 103 10s 104 10s 106 107 108 109 10 !0 for X field then can be written in terms of a redefined varia/a, able Xk = -^/fefl3/2 a s
r
2
i
••
-
,'IMim!
•
1
"
\
]
'
!
i
Fig. 8. The evolution of energy densities and T/Mx as a function of the scale factor. Also shown is A'/A'EQ. Physica Scripta T85
Xk(t) + [A + 2qcos(2t)]xk(t) = 0
(47) Physica Scripta 2000
Early-Universe Issues: Seeds of Perturbations and Birth of Dark Matter where A depends on the energy of the particle and q depends on the inflaton field oscillation amplitude. When A and q are constants, this equation is usually referred to as the Mathieu equation which exhibits resonant mode instability for certain values of A and q. In an expanding universe, A and q will vary in time, but if they vary slowly compared to the frequency of oscillations, the effects of resonance will remain. If the mode occupation number for the X particles is large, the number density per mode of the X particles will be proportional to \%k\2. If A and q have the appropriate values for resonance, Xk w m grow exponentially in time, and hence the number density will attain an exponential enhancement above the usual perturbative decay. This period of enhanced rate of energy transfer has been called preheating primarily because the particles that are produced during this period have yet to achieve thermal equilibrium.
243
The procedure to calculate the WIMPZILLA density is the same as in Section 3.3.1. Now, in addition to the parameter Mx/Hi, there is another parameter gmp\/H\. Now in large-field models H\ ~ 1013 GeV, so mp\/Hi might be as large as 106. The choice of g = 10~3 would yield gmp\/ Hj = 103. Fig. 9 (from [42]) shows the dependence of the WIMPZILLA density upon Mx/H\ for the particular choice gM-p\/H\ = 106. This would correspond to g ~ 1 in large-field inflation models where mp\/H\ = 106, about the largest possible value. Note that Qx ~ 1 obtains for Mx/Hi « 103. The dashed and dotted curves are two analytic approximations discussed in [42], while the solid curve is the numerical result. The approximations are in very good agreement with the numerical results. Figure 10 (from [42]) shows the dependence of the WIMPZILLA density upon gm?\/H\. For this graph Mx/H\ was chosen to be unity. This figure illustrates the fact that the dependence of Qxh2 on gm?\/H\ is not mono tonic. For details, see [42].
This resonant amplification leads to an efficient transfer of energy from the inflaton to other particles which may have stronger coupling to other particles than the inflaton, thereby speeding up the reheating process and leading to a higher reheating temperature than in the usual scenario. Another interesting feature is that particles of mass larger 3.3.4. Production in bubble collisions than the inflaton mass can be produced through this coherWIMPZILLAS may also be produced [39] if inflation is coment resonant effect. This has been exploited to construct a pleted by a first-order phase transition [43], in which the baryogenesis scenario [40] in which the baryon number universe exits from a false-vacuum state by bubble violating bosons with masses larger than the inflaton mass are created through the resonance mechanism. A natural variation on this idea is to produce WIMPZILLAS by the same g M„,/H,= 10« resonance mechanism. Interestingly enough, what was found [42] is that in the context of a slow-roll inflation with the potential V(
i,800 ^700 ~600 ?500
I^ 400 300tfsoo* 100 X
u
2
(a k+k
2
2
2 2\„2 2
+ (M x + g ^ )a
again taking t; = 1/6. © Physica Scripta 2000
0
M*/H,
4
(48) Fig. 10. An illustration of the nonmonotonic behavior of the particle density produced with the variation of the coupling constant. The value of Mx/Hi is set to unity [42]. Physica Scripta T85
244
Edward W. Kolb
nucleation [44]. When bubbles of true vacuum form, the energy of the false vacuum is entirely transformed into potential energy in the bubble walls. As the bubbles expand, more and more of their energy becomes kinetic as the walls become highly relativistic. In bubble collisions the walls oscillate through each other [45] and their kinetic energy is dispersed into low-energy scalar waves [45,46]. We are interested in the potential energy of the walls, MP = 4nr\R2, where r\ is the energy per unit area of a bubble wall of radius R. The bubble walls can be visualized as a coherent state of infiaton particles, so the typical energy E of the products of their decays is simply the inverse thickness of the wall, E ~ A'1. If the bubble walls are highly relativistic when they collide, there is the possibility of quantum production of nonthermal particles with mass well above the mass of the infiaton field, up to energy A~l = yAfy, with y the relativistic Lorentz factor. Suppose for illustration that the WIMPZILLA is a fermion coupled to the infiaton field by a Yukawa coupling g(j>XX. One can treat > (the bubbles or walls) as a classical, external field and the WIMPZILLA as a quantum field in the presence of this source. The number of WIMPZILLAS created in the collisions from the wall potential energy is Nx ~ fxMp/Mx, where fx parametrizes the fraction of the primary decay products in WIMPZILLAS. The fraction fx will depend in general on the masses and the couplings of a particular theory in question. For the Yukawa coupling g, it i s / x ~ g2ln{yM,j,/2Mx) [46,47]. WIMPZILLAS may be produced in bubble collisions out of equilibrium and never attain chemical equilibrium. Even with 7RH as low as 100 GeV, the present WIMPZILLA abundance would be Qx ~ 1 if g~ 10" 5 a 1/2 . Here a"1 «; 1 is the fraction of the bubble energy at nucleation in the form of potential energy at the time of collision. This simple analysis indicates that the correct magnitude for the abundance of WIMPZILLAS may be naturally obtained in the process of reheating in theories where inflation is terminated by bubble nucleation. 3.4. Wimpzilla conclusions In this talk I have pointed out several ways to generate nonthermal dark matter. All of the methods can result in dark matter much more massive than the feeble little weak-scale mass thermal relics. The nonthermal dark matter may be as massive as the GUT scale, truly in the WIMPZILLA range. The mass scale of the WIMPZILLAS is determined by the mass scale of inflation, more exactly, the expansion rate of the universe at the end of inflation. For large-field inflation models, that mass scale is of order 10 13 GeV. For small-field inflation models, it may be less, perhaps much less. The mass scale of inflation may one day be measured! In addition to scalar density perturbations, tensor perturbations are produced in inflation. The tensor perturbations are directly proportional to the expansion rate during inflation, so determination of a tensor contribution to cosmic background radiation temperature fluctuations would give the value of the expansion rate of the universe during inflation and set the scale for the mass of the WIMPZILLA.
Physica Scripta T85
Undoubtedly, other methods for WIMPZILLA production will be developed. But perhaps even with the present scenarios one should start to investigate methods for WIMPZILLA detection. While wimpy wimps must be color singlets and electrically neutral, WIMPZILLAS may be endowed with color and electric charge. This should open new avenues for detection and exclusion of WIMPZILLAS. The lesson here is that WIMPZILLAS may surprise and be the dark matter, and we may learn that size does matter! Acknowledgements This work was supported by the DOE and NASA under Grant NAG5-7092.
References 1. Kofman, L., Linde, A. D. and Starobinsky, A. A., Phys. Rev. D 56, 3258 (1997). 2. For a review, see Mukhanov, V. F., Feldman, H. A. and Brandenberger, R. H., Phys. Rep. 215, 203 (1992). 3. Dodelson, S., Kinney, W. H. and Kolb, E. W., Phys. Rev. D 56, 3207 (1997). 4. Lidsey, J. E. et al, Rev. Mod. Phys. 69, 373 (1997). 5. Copeland, E. J., Kolb, E. W., Liddle, A. R. and Lidsey, J. E., Phys. Rev. Lett. 71, 219 (1993), Phys. Rev. D 48, 2529 (1993), 49, 1840 (1994); Turner, M. S., Phys. Rev. D 48, 5539 (1993). 6. Liddle, A. R. and Lyth, D. H., Phys. Rep. 231, 1 (1993). 7. Grivell, I. J. and Liddle, A. R., Phys. Rev. D 54, 7191 (1996). 8. Salopek, D. S. and Bond, J. R., Phys. Rev. D 42, 3936 (1990). 9. Liddle, A. R. and Lyth, D. H., Phys. Lett. B 291, 391 (1992). 10. Liddle, A. R., Parsons, P. and Barrow, J. D., Phys. Rev. D 50, 7222 (1994). 11. Stewart, E. D. and Lyth, D. H., Phys. Lett. B 302, 171 (1993). 12. Jungman, G., Kamionkowski, M., Kosowsky, A. and Spergel, D. N., Phys. Rev. D 54,1332 (1996); Bond, J. R., Efstathiou, G. and Tegmark, M., Mon. Not. R. Astron. Soc. 291, L33 (1997). 13. Zaldarriaga, M., Spergel, D. and Seljak, U., Astrophys. J. 488,1 (1997). 14. Copeland, E. J., Grivell, E. J. and Liddle, A. R., Sussex preprint astro-ph/9712028 (1997). 15. Copeland, E. J., Grivell, I. J., Kolb, E. W. and Liddle, A. R., Phys. Rev. D58 043002 (1998) . 16. Wang, L., Mukhanov, V. F. and Steinhardt, P. J., Phys. Lett. B 414,18 (1997). 17. Kolb, E. W. and Turner, M. S., "The Early Universe", (AddisonWesley, Menlo Park, Ca., 1990). 18. Griest, K. and Kamionkowski, M., Phys. Rev. Lett. 64, 615 (1990). 19. Ellis, J.,Lopez, J. L. and Nanopoulos, D. V., Phys. Lett. B 247, 257 (1990). 20. See, for instance, Giudice, G. F. and Rattazzi, R., hep-ph/9801271. 21. Raby, S., Phys. Rev. D 56, (1997). 22. Hamaguchi, K., Nomura, Y. and Yanagida, T., hep-ph/9805346. 23. Benakli, K., Ellis, J. and Nanopoulos, D. V., hep-ph/9803333. 24. Nanopoulos, D. V., "Proceedings of DARK98", (1998) (L. Baudis and H. Klapdor-Kleingrothaus, eds). 25. Kuzmin, V. A. and Rubakov, V. A., Phys. Atom. Nucl. 61,1028 (1998). 26. Birkel, M. and Sarkar, S., hep-ph/9804285. 27. Kofman, L. A., Linde, A. D. and Starobinsky, A. A., Phys. Rev. Lett. 73, 3195 (1994); Khlebnikov, S. Yu. and Tkachev, 1.1., Phys. Rev. Lett. 77, 219 (1996); Phys. Lett. B 390, 80 (1997); Phys. Rev. Lett. 79, 1607 (1997); Phys. Rev. D 56, 653 (1997); Anderson, G. W., Linde, A. and Riotto, A., Phys. Rev. Lett. 77, 3716 (1996); see Kofman, L., "The origin of matter in the Universe: reheating after inflation", astro-ph/9605155, UH-IFA-96-28 preprint, 16 pp., to appear in "Relativistic Astrophysics: A Conference in Honor of Igor Novikov's 60th Birthday", (eds. B. Jones and D. Markovic), for a more recent review and a collection of references; see also Kofman, L., Linde, A. D. and Starobinsky, A. A., Phys. Rev. D 56, 3258 (1997); Traschen, J. and Brandenberger, R., Phys. Rev. D 42, 2491 (1990); Shtanov, Y., Traschen, J. and Brandenberger, R., Phys. Rev. D 51, 5438 (1995). 28. Chung, D. J. H., Kolb, E. W. and Riotto, A., hep-ph/9802238. 29. Fulling, S., Gen. Rel. and Grav. 10, 807 (1979). © Physica Scripta 2000
Early-Universe Issues: Seeds of Perturbations and Birth of Dark Matter 30. Birrell, N. D. and Davies, P. C. W., "Quantum Fields in Curved Space" (Cambridge University Press, Cambridge, 1982). 31. Chitre, D. M. and Hartle, J. B., Phys. Rev. D 16, 251 (1977); Raine, D. J. and Winlove, C. P., Phys. Rev. D 12, 946 (1975); Schaefer, G. and Dehnen, H., Astron. Astrophys. 54, 823 (1977). 32. Birrell, N. D. and Davies, P. C. W., J. Phys. A; Math. Gen. 13, 2109 (1980). 33. Chung, D. J. H., Kolb, E. W. and Riotto, A., hep-ph/9809453. 34. Lyth, D. H. and Riotto, A., hep-ph/9807278. 35. Ellis, J., Kim, J. and Nanopoulos, D. V., Phys. Lett. B 145, 181 (1984); Krauss, L. M., Nucl. Phys. B 227, 556 (1983); Khlopov, M. Yu. and Linde, A. D., Phys. Lett. 138B, 265 (1984); Ellis, J., Nanopoulos, D. V. and Sarkar, S., Nucl. Phys. B 461, 597 (1996). 36. Scherrer, R. J. and Turner, M. S., Phys. Rev. D 31, 681 (1985).
© Physica Scripta 2000
245
37. Berezinsky, V., Kachelriess, M. and Vilenkin, A., Phys. Rev. Lett. 79, 4302 (1997). 38. Kuzmin, V. and Tkachev, I. I., hep-ph/9802304. 39. Chung, D. J. H., Kolb, E. W. and Riotto, A., hep-ph/9805473. 40. Kolb, E. W., Linde, A. D. and Riotto, A., Phys. Rev. Lett. 77, 4290 (1996). 41. Kolb, E. W., Riotto, A. and Tkachev, 1.1., Phys. Lett. B423,348 (1998). 42. Chung, D. J., hep-ph/9809489. 43. La, D. and Steinhardt, P. J., Phys. Rev. Lett. 62, 376 (1989). 44. Guth, A. H., Phys. Rev. D 23, 347 (1981). 45. Hawking, S. W., Moss, I. G. and Stewart, J. M„ Phys. Rev. D 26, 2681 (1982). 46. Watkins, R. and Widrow, L., Nucl. Phys. B 374, 446 (1992). 47. Masiero, A. and Riotto, A., Phys. Lett. B 289, 73 (1992).
Physica Scripta T85
Physica Scripta.Vol. T85, 246-253, 2000
String Cosmology and the Beginning-of-Time Myth G. Veneziano Theory Division, CERN, CH-1211 Geneva 23, Switzerland Received January 27, 1999; accepted August 2, 1999
PACS Ref: 98.80.Bp, 98.80.Cq
Abstract The myth according to which the Universe - and time itself- started with a big bang will be challenged. Classical and quantum gravitational instabilities, can, respectively, inflate and warm up a primordial pre-bangian Universe satisfying a superstring-motivated principle of "Asymptotic Past Triviality". Properties of the pre-bangian Universe can be probed through its observable relics, which include: (z) a (possibly observable) stochastic gravitational-wave background; (//) a (possible) new mechanism for seeding the galactic magnetic fields; (z'z'O a (possible) new source of large-scale structure and CMB anisotropy.
only known candidate theory of quantum gravity, superstring theory. Rather, as we shall argue, superstring theory gives strong hints in favour of the second (pre-big bang) possibility through two of its very basic properties, the first in relation to its short-distance behaviour, the second from its modifications of General Relativity even at large distance. Let us briefly comment on both.
2. (Super)String inspiration 1. Introduction
2.1. Short distance
I would like to begin this talk by asking a very simple question: Did the Universe start "small"? The naive answer is: Yes, of course! However, a serious answer can only be given after defining the two keywords in the question: What do we mean by "start"? and What is "small" relative to? In order to be on the safe side, let us take the "initial" time to be a bit larger than Planck's time, fp ~ 10~43 s. Then, in standard Friedmann-Robertson-Walker (FRW) cosmology, the initial size of the (presently observable) Universe was about 10 _2 cm. This is of course tiny w.r.t. its present size (~ 1028cm), yet huge w.r.t. the horizon at that time, i.e. w.r.t. /p = ctf ~ 10~33cm. In other words, a few Planck times after the big bang, our observable Universe consisted of (1030)3 = 1090 Planckian -size, causally disconnected regions. More precisely, soon after t = tp, the Universe was characterized by a huge hierarchy between its Hubble radius and inverse temperature on one side, and its spatial- curvature radius and homogeneity scale on the other. The relative factor of (at least) 1030 appears as an incredible amount of fine-tuning on the initial state of the Universe, corresponding to a huge asymmetry between time and space derivatives. Was this asymmetry really there? And if so, can it be explained in any more natural way? It is well known that a generic way to wash out inhomogeneities and spatial curvature consists in introducing, in the history of the Universe, a long period of accelerated expansion, called inflation [1]. This still leaves two alternative solutions: either the Universe was generic at the big bang and became flat and smooth because of a long post-bangian inflationary phase; or it was already flat and smooth at the big bang as the result of a long pre-bungian inflationary phase. Assuming, dogmatically, that the Universe (and time itself) started at the big bang, leaves only the first alternative. However, that solution has its own problems, in particular those of fine-tuned initial conditions and inflaton potentials. Besides, it is quite difficult to base standard inflation in the
Since the classical (Nambu-Goto) action of a string is proportional to the area A of the surface it sweeps, its quantization must introduce a quantum of length As through:
Physica Scripta T85
S/h = A/A2 .
(1)
This fundamental length, replacing Planck's constant in quantum string theory [2], plays the role of a minimal observable length, of an ultraviolet cut-off. Thus, in string theory, physical quantities are expected to be bound by appropriate powers of 2S, e.g. H2~R~Gp> kBT/n
2.;2,
< d;\
(2)
In other words, in quantum string theory (QST), relativistic quantum mechanics should solve the singularity problems in much the same way as non-relativistic quantum mechanics solved the singularity problem of the hydrogen atom by putting the electron and the proton a finite distance apart. By the same token, QST gives us a rationale for asking daring questions such as: What was there before the big bang? Certainly, in no other present theory, can such a question be meaningfully asked. 2.2. Large distance Even at large distance (low-energy, small curvatures), superstring theory does not automatically give Einstein's General Relativity. Rather, it leads to a scalar-tensor theory of the JBD variety. The new scalar particle/field >, the so-called dilaton, is unavoidable in string theory, and gets reinterpreted as the radius of a new dimension of space in so-called M-theory [3]. By supersymmetry, the dilaton is massless to all orders in perturbation theory, i.e. as long as supersymmetry remains unbroken. This raises the question: Is the dilaton a problem or an opportunity? My answer is that it is possibly both; and while we can try to © Physica Scripta 2000
String cosmology and the beginning-of-time myth
247
cosmology going towards a singularity in the future. Of course, the lowest order approximation breaks down before either singularity is reached. A (stringy) moment away from their respective singularities, these two branches can easily be joined smoothly to give a single non-singular cosmology, (3) at least mathematically. Leaving aside this issue for the ' p M s ~ agauge moment (see Section 5 for more discussion), let us go back showing the basic unification of all forces in string theory to DDI. Since such a phase is characterized by growing and the fact that, in our conventions, the weak-coupling coupling and curvature, it must itself have originated from region coincides with <>/
4. Inflation as a classical gravitational instability
Present <> | = Fig. 1. Dilaton evolution in string cosmology. © Physica Scripta 2000
The assumption of APT entitles us to treat the early history of the Universe through the classical field equations of the low-energy (because of the small curvature) tree-level (because of the weak coupling) effective action of string theory. For simplicity, we will illustrate here the simplest case of the gravi-dilaton system already compactified to four space-time dimensions. Other fields and extra dimensions will be mentioned below, when we discuss observable Physica Scripta T85
248
G. Veneziano
Space
AT data given here
Fig. 2. Penrose diagram illustrating pre-big bang cosmology.
consequences. The (string frame) effective action then reads:
r,eff ^2 J dV=ge-*(ft + 3^4) :
(5)
In this frame, the string-length parameter As is a constant and the same is true of the curvature scale at which we have to supplement Eq. (5) with corrections. Similarly, string masses, when measured with the string metric, are fixed, while test strings sweep geodesic surfaces with respect to that metric. For all these reasons, even if we will allow metric redefinitions in order to simplify our calculations, we shall eventually turn back to the string frame for the physical interpretation of the results. We stress, however, that, while our intuition is not frame independent, physically measurable quantities are. Even assuming APT, the problem of determining the properties of a generic solution to the field equations implied by (5) is a formidable one. Very luckily, however, we are able to map our problem into one that has been much investigated, both analytically and numerically, in the literature. This is done by going to the so-called "Einstein frame". For our purposes, it simply amounts to the field redefinition (6)
ft-=$>«*-* in terms of which (7) becomes:
reff = ^fd4x^=^(n^
- ^E)9,<Mtf>) .
(7)
where 0O (/p = Ase^/2) is the present value of the dilaton (of Planck's length). Physica Scripta T85
Our problem is thus reduced to that of studying a massless scalar field minimally coupled to gravity. Such a system has been considered by many authors, in particular by Christodoulou [10], precisely in the regime of interest to us. In line with the APT postulate, in the analogue gravitational collapse problem, one assumes very "weak" initial data with the aim of finding under which conditions gravitational collapse later occurs. Gravitational collapse means that the (Einstein) metric (and the volume of 3-space) shrinks to zero at a space-like singularity. However, typically, the dilaton blows up at that same singularity. Given the relation (6) between the Einstein and the (physical) string metric, we can easily imagine that the latter blows up near the singularity as implied by DDL How generically does this happen? In this connection it is crucial to recall the singularity theorems of Hawking and Penrose [11], which state that, under some general assumptions, singularities are inescapable in GR. One can look at the validity of those assumptions in the case at hand and finds that all but one are automatically satisfied. The only condition to be imposed is the existence of a closed trapped surface (a closed surface from where future light cones lie entirely in the region inside the surface). Rigorous results [10] show that this condition cannot be waived: sufficiently weak initial data do not lead to closed trapped surfaces, to collapse, or to singularities. Sufficiently strong initial data do. But where is the border-line? This is not known in general, but precise criteria do exist for particularly symmetric space-times, e.g. for those endowed with spherical symmetry. However, no matter what the general collapse/singularity criterion will eventually turn out to be, we do know that: • it cannot depend on an over-all additive constant in 4>; • it cannot depend on an over-all multiplicative factor in g^. This is a simple consequence of the invariance (up to an over-all factor) of the effective action (7) under shifts of the dilaton and rescaling of the metric (these properties depend crucially on the validity of the tree-level low-energy approximation and on the absence of a cosmological constant). We conclude that, generically, some regions of space will undergo gravitational collapse, will form horizons and singularities therein, but nothing, at the level of our approximations, will be able to fix either the size of the horizon or the value of (p at the onset of collapse. When this is translated into the string frame, one is describing, in the region of space-time within the horizon, a period of DDI in which both the initial value of the Hubble parameter and that of
String cosmology and the beginning-of-time myth chaotic inflation [13]. Given some generic (though APT) initial data, we should ask which is the distribution of sizes of the collapsing regions and of couplings therein. Then, only the "tails" of these distributions, i.e. those corresponding to sufficiently large, and sufficiently weakly coupled regions, will produce Universes like ours, the rest will not. The question of how likely a "good" big bang is to take place is not very well posed and can be greatly affected by anthropic considerations. In conclusion, we may summarize recent progress on the problem of initial conditions by saying that [8]: Dilatondriven inflation in string cosmology is as generic as gravitational collapse in General Relativity. At the same time, having a sufficiently long period of DDI amounts to setting upper limits on two arbitrary moduli of the classical solutions. Our scenario is illustrated in Figs. 3 and 4, both taken from Ref. [8]. In Fig. 3,1 show, for the spherically symmetric case, a Carter-Penrose diagram in which generic (but asymptotically trivial) dilatonic waves are given around time-like (/") and null ( J " ) past-infinity. In the shaded region near I~,l~, a weak-field solution holds. However, if a collapse criterion is met, an apparent horizon, inside which a cosmological (generally inhomogeneous) PBB-like solution takes over, forms at some later time. The future singularity of the PBB solution at t = 0 is identified with the space-like singularity of the black hole at r = 0 (remember that r is a time-like coordinate inside the horizon). Figure 4 gives a (2 + l)-dimensional sketch of a possible PBB Universe: an original "sea" of dilatonic and gravity waves leads to collapsing regions of different initial size, possibly to a scale-invariant distribution of them. Each one of these collapses is reinterpreted, in the string frame, as the process by which a baby Universe is born after a period of PBB inflationary "pregnancy", with the size of each baby Universe determined by the duration of its pregnancy, i.e. by the initial size of the corresponding collapsing region. Regions initially larger than 10~13cm can generate Universes like ours, smaller ones cannot.
/
'»
/
<
/
/ X
i )
I
I weak-field expansion
|
| cosmological regime
Fig. 3. Penrose diagram for collapse/inflation in string cosmology. © Physica Scripta 2000
249
Fig. 4. Sketch of pre-big bang formation of our Universe.
A basic difference between the large numbers needed in (non-inflationary) FRW cosmology and the large numbers needed in PBB cosmology should be stressed at this point. In the former, the ratio of two classical scales, e.g. of total curvature to its spatial component, which is expected to be 0(1), has to be taken as large as 1060. In the latter, the above ratio is initially 0(1) in the collapsing/inflating region, and ends up being very large in that region thanks to DDI. However, the common order of magnitude of these two classical quantities is a free parameter, and is taken to be much larger than a classically irrelevant quantum scale. We can visualize analogies and differences between standard and pre-big bang inflation by comparing Figs. 5a and 5b. In these, I sketch the evolution of the Hubble radius and of a fixed comoving scale (here the one corresponding to the part of the Universe presently observable to us) as a function of time in the two scenarios. The common feature is that the fixed comoving scale was "inside the horizon" for some time during inflation, and possibly very deeply inside at its onset. Also, in both cases, the Hubble radius at the beginning of inflation had to be large in Planck units and the scale of homogeneity had to be at least as large. The difference between the two scenarios is just in the behaviour of the Hubble radius during inflation: increasing in standard inflation (a), decreasing in string cosmology (b). This is what makes PBB's "wine glass" more elegant, and stable! Thus, while standard inflation is still facing the initial-singularity question and needs a non- adiabatic phenomenon to reheat the Universe (a kind of small bang), PBB cosmology faces the singularity problem later, combining it to the exit and heating problems (discussed in Sections 5 and 6.2, respectively). Physica Scripta T85
250
G. Veneziano
Here
Now
Post big bang expansion Space
Space
Pre big bang inflation
s f >
i!
M i l
Big bang? *
,
K_
Start of collapse/inflatic
A
Towards Triviality Fig. 5. Wine glass of standard inflation.
Fig. 6. Wine glass of pre-big bang scenario.
In the end, what saves PBB cosmology from fine-tuning is (not surprisingly!) super symmetry. This is what protects us from the appearance of a cosmological constant in the weak-coupling regime. Even a relatively small cosmological constant would invalidate our scale-invariance arguments and force DDI to be very short [5]. Thus, amusingly, while an effective cosmological constant is at the basis of standard (post-big bang) inflation, its absence in the weak coupling region is at the basis of PBB inflation. This may allow us to speculate that the absence (or extreme smallness) of the present cosmological constant may be related to a mysterious degeneracy between the perturbative and the non-perturbative vacuum of superstring theory.
It was recently pointed out [15] that the reverse order of events is also possible. The coupling may become large before the curvature. In this case, at least for some time, the low-energy limit of M-theory should be adequate: this limit is known [3] to give D = 11 supergravity and is therefore amenable to reliable study. It is likely, though not yet clear, that, also in this case, strong curvatures will have to be reached before the exit can be completed. In the following, we will assume that:
5. The exit problem/conjecture We have argued that, genetically, DDI, when studied at lowest order in derivatives and coupling, evolves towards a singularity of the big bang type. Similarly, at the same level of approximation, the non-inflationary solutions emerge from a singularity. Matching these two branches in a smooth, non-singular way has become known as the (graceful) exit problem in string cosmology [14]. It is, undoubtedly, the most important theoretical problem facing the whole PBB scenario. There has been quite some progress recently on the exit problem. However, for lack of space, I shall refer the reader to the literature [14] for details. Generically speaking, toy examples have shown that DDI can flow, thanks to higher-curvature corrections, into a de-Sitter-like phase, i.e. into a phase of constant H (curvature) and constant (j>. This phase is expected to last until loop corrections become important (see next section) and give rise to a transition to a radiation-dominated phase. If these toy models serve as an indication, the full exit problem can only be achieved at large coupling and curvature, a situation that should be described by the newly conjectured M-theory [3]. Physica Scripta T85
• the big bang singularity is avoided thanks to the softness of string theory; • full exit to radiation occurs at strong coupling and curvature, according to a criterion given in Section 6.2.
6. Observable relics and heating the pre-bang Universe 6.1. PBB relics Since there are already several review papers on this subject (e.g. [16]), I will limit myself to mentioning the most recent developments, after recalling the basic physical mechanism underlying particle production in cosmology [17]. A cosmological (i.e. time-dependent) background coupled to a given type of (small) inhomogeneous perturbation *F enters the effective low-energy action in the form: l
-jdr,d3xS(r,)[r2-(VY)2].
(10)
Here r\ is the conformal-time coordinate, and a prime denotes d/drj. The function S(rj) (sometimes called the "pump" field) is, for any given W, a given function of the scale factor a(r\), and of other scalar fields (four-dimensional dilaton (j>(r\), moduli &,•(>/), etc.), which may appear non-trivially in the background. While it is clear that a constant pump field S can be reabsorbed in a rescaling of W, and is thus ineffective, a time-dependent S couples non-trivially to the fluctuation © Physica Scripta 2000
String cosmology and the beginning-of-time myth
251
Thus, converting the primordial seeds into those that may have existed at the proto-galaxy formation epoch is by no means a trivial exercise. Work is in progress to try to adapt existing codes [25] to the evolution of our primordial seeds. Gravity waves, dilaton : S = a2e~^ • Finally, for Kalb-Ramond fields and axions, a and cj) work in the same direction and spectra can be large even at large Heterotic gauge bosons : S = e~^ (9) scales [26]. An interesting fact is that, unlike the GW Kalb - Ramond, axions : S = a~2e~^ . spectrum, that of axions is very sensitive to the cosmological behaviour of internal dimensions during A distinctive property of string cosmology is that the dilaton the DDI epoch. On one side, this makes the model less
and leads to the production of pairs of quanta (with equal and opposite momenta). One can easily determine the pump fields for each one of the most interesting perturbations. The result is:
QGW < lO" 14 .
(10)
Since the GW spectra of PBB cosmology are "blue", the bound by COBE is automatically satisfied, with no implication on the GW yield at interesting frequencies. Values of &GW m t n e range of 10" 6 -10~ 7 are possible in some regions of parameter space, which, according to some estimates of sensitivities [20], could be inside detection capabilities in the near future. • For gauge bosons there is no amplification of vacuum fluctuations in standard cosmology, since a conformally flat metric (of the type forced upon by inflation) decouples from the electromagnetic (EM) field precisely in D = 3 + 1 dimensions. As a very general remark, apart from pathological solutions, the only background field which, through its cosmological variation, can amplify EM (more generally gauge-field) quantum fluctuations is the effective gauge coupling itself [21]. By its very nature, in the pre-big bang scenario the effective gauge coupling inflates together with space during the PBB phase. It is thus automatic that any efficient PBB inflation brings together a huge variation of the effective gauge coupling and thus a very large amplification of the primordial EM fluctuations [22-24]. This can possibly provide the longsought origin for the primordial seeds of the observed galactic magnetic fields. Notice, however, that, unlike GW, EM perturbations interact quite considerably with the hot plasma of the early (post-big bang) Universe. 1
This, however, refers just to first-order tensor perturbations; the mechanism —described below— of seeding CMB anisotropy through axions would also give a tensor (and a vector) contribution whose relative magnitude is being computed. © Physica Scripta 2000
{AT/T(n)AT/T(ri))
= J ^ ( 2 / + l)C,Pi(cos d),
(12)
l
and the parameters Hmax,kmax, a are defined by the primordial axion energy spectrum in critical units as: QUk)=(j^)(k/kmaxT.
(13)
In string theory, as repeatedly mentioned, we expect # m a x / M P ~ M s / M p ~ l / 1 0 and r]0kmax ~ 1030, while the exponent a depends on the explicit PBB background with the above-mentioned HZ case corresponding to a = 0. The standard tilt parameter n = ns (s for scalar) is given by n = 1 + 2a and is found, by COBE, to lie between 0.9 and 1.5, corresponding to 0 < a < 0.25 (a negative a leads to some theoretical problems). With these inputs we can see that the correct normalization (C2 ~ 10~10) is reached for a ~ 0.2, which is just in the middle of the allowed range. In other words, unlike in standard inflation, we cannot predict the tilt, but when this is given, we can predict (again unlike in standard inflation) the normalization. Our model, being of the isocurvature type, bears some resemblance to the one recently advocated by Peebles [28] and, like his, is expected to contain some calculable amount of non-Gaussianity, which is being calculated and will be checked by the future satellite measurements (MAP, PLANCK). • Many other perturbations, which arise in generic compactifications of superstrings, have also been studied, and lead to interesting spectra. For lack of time, I will refer to the existing literature [23,24]. Physica Scripta T85
252
G. Veneziano
6.2. Heat and entropy as a quantum gravitational instability Before closing this section, I wish to recall how one sees the very origin of the hot big bang in this scenario. One can easily estimate the total energy stored in the quantum fluctuations, which were amplified by the pre-big bang backgrounds. The result is, roughly, " quantum ~ ^ e f f / / m a x ,
(14)
where JVefr is the effective number of species that are amplified and // m a x is the maximal curvature scale reached around t = 0. We have already argued that // m a x ~ M s = k~x, and we know that, in heterotic string theory, Nea is in the hundreds. Yet this rather huge energy density is very far from critical, as long as the dilaton is still in the weakcoupling region, justifying our neglect of back-reaction effects. It is very tempting to assume [23] that, precisely when the dilaton reaches a value such that p q u a n t u m is critical, the Universe will enter the radiation-dominated phase. This PBBB (PBB bootstrap) constraint gives, typically: e * - ~ 1/JVeir ,
(15)
i.e. a value for the dilaton close to its present value. The entropy in these quantum fluctuations can also be estimated following some general results [29]. The result for the density of entropy S is, as expected S~AA ef f// max .
(16)
It is easy to check that, at the assumed time of exit given by (19), this entropy saturates a recently proposed holography bound [30]. This also turns out to be a physically acceptable value for the entropy of the Universe just after the big bang: a large entropy on the one hand (about 1090); a small entropy for the total mass and size of the observable Universe on the other, as often pointed out by Penrose [31]. Thus, PBB cosmology neatly explains [32] why the Universe, at the big bang, looks so fine-tuned (without being so) and provides a natural arrow of time in the direction of higher entropy. 7. Conclusions • Pre-big bang (PBB) cosmology is a "top-down" rather than a "bottom-up" approach to cosmology. This should not be forgotten when testing its predictions. • It does not need to invent an inflaton, or to fine-tune its potential; inflation is "natural" thanks to the duality symmetries of string cosmology. • It makes use of a classical gravitational instability to inflate the Universe, and of a quantum instability to warm it up. • The problem of initial conditions "decouples" from the singularity problem; it is classical, scale-free, and unambiguously defined. Issues of fine tuning can be addressed and, I believe, answered. • The spectrum of large-scale perturbations has become more promising through the invisible axion of string theory, while the possibility of explaining the seeds of galactic magnetic fields remains a unique prediction of the model. • The main conceptual (technical?) problem remains that of providing a fully convincing mechanism for (and a detailed description of) the pre-to-post-big bang transition. It is Physica Scripta T85
very likely that such a mechanism will involve both high curvatures and large coupling and should therefore be discussed in the (yet to be fully constructed) M-theory [3]. New ideas borrowed from such theory and from D-branes [33,15] could help in this respect. • Once/if this problem will be solved, predictions will become more precise and robust, but, even now, with some mild assumptions, several tests are (or will soon become) possible, e.g. — the tensor contribution to AT/T should be very small (see, however, footnote Section 6); — some non-Gaussianity in AT/T correlations is expected, and calculable. — the axion-seed mechanism should lead to a characteristic acoustic-peak structure, which is being calculated; — it should be possible to convert the predicted seed magnetic fields into observables by using some reliable code for their late evolution; — a characteristic spectrum of stochastic gravitational waves is expected to surround us and could be large enough to be measurable within a decade or so.
References 1. Kolb, E. W. and Turner, M. S., "The Early Universe" (AddisonWesley, Redwood City, CA, 1990); A. D. Linde, "Particle Physics and Inflationary Cosmology" (Harwood, New York, 1990). 2. Veneziano, G., Europhys. Lett. 2, 133 (1986); "The Challenging Questions", Erice, 1989, (ed. A. Zichichi) (Plenum Press, New York, 1990), p. 199. 3. See, e.g., Witten, E., Nucl. Phys. B 443, 85 (1995); Horawa, P. and Witten, E., Nucl. Phys. B 460, 506 (1996). 4. Taylor, T. R. and Veneziano, G., Phys. Lett. B 213, 459 (1988). 5. Veneziano, G., Phys. Lett. B 265, 287 (1991). 6. Gasperini, M. and Veneziano, G., Astropart. Phys. 1, 317(1993), Mod. Phys. Lett. A 8, 3701 (1993), Phys. Rev. D 50 2519 (1994). 7. An updated collection of papers on the PBB scenario is available at http://www.to.infn.it/~gasperin/. 8. Buonanno, A. Damour, T. and Veneziano, G., Nucl. Phys. B 543, 275 (1999); see also, Veneziano, G., Phys. Lett. B 406, 297 (1997); Buonanno, A., Meissner, K. A., Ungarelli, C. and Veneziano, G., Phys. Rev. D 57, 2543 (1998), and references therein. 9. Penrose, R., "Structure of space-time", in "Battelle Rencontres", (ed. Dewitt, C. and Wheeler, J. A), (Benjamin, New York, 1968). 10. Christodoulou, D., Commun. Pure Appl. Math. 56, 1131 (1993), and references therein. 11. Penrose, R., Phys. Rev. Lett. 14, 57 (1965); Hawking, S. W. and Penrose, R., Proc. Roy. Soc. Lond. A 314 529 (1970). 12. Turner, M. and Weinberg, E., Phys. Rev. D 56, 4604 (1997); Kaloper, N. Linde, A. and Bousso, R., Phys. Rev. D 59 043508 (1999). 13. Linde, A., Phys. Lett. 129B, 177 (1983). 14. Brustein, R. and Veneziano, G., Phys. Lett. B 329,429 (1994); Kaloper, N., Madden, R. and Olive, K. A., Nucl. Phys. B 452, 677 (1995), Phys. Lett. B 371, 34 (1996); Easther, R., Maeda, K. and Wands, D., Phys. Rev. D 53, 4247 (1996); Gasperini, M., Maggiore, M. and Veneziano, G., Nucl. Phys. B 494, 315 (1997); Brustein, R. and Madden, R., Phys. Lett. B 410, 110 (1997), Phys. Rev. D 57, 712 (1998). 15. Maggiore, M. and Riotto, A., "D-branes and Cosmology", hep-th/ 9811089; see also Banks, T., Fishier, W. and Motl, L., "Duality versus Singularities", hep-th/9811194. 16. Veneziano, G., in 'String Gravity and Physics at the Planck Energy Scale", Erice, 1995, (eds. N . Sanchez and A. Zichichi), (Kluver Academic Publishers, Boston, 1996), p. 285; Gasperini, M., ibid., p. 305. 17. See, e.g., Mukhanov, V. F., Feldman, A. H. and Brandenberger, R. H., Phys. Rep. 215, 203 (1992). 18. Brustein, R., Gasperini, M., Giovannini, M. and Veneziano, G., Phys. Lett. B361,45 (1995); Brustein, R. et a!., Phys. Rev. D 51, 6744 (1995). 19. Smoot, G. F. et al., Astrophys. J. 396, LI (1992); Bennet, C. L. et al., Astrophys. J. 430, 423 (1994). © Physica Scripta 2000
String cosmology and the beginning-of-time myth 20. Astone, P. et al., Phys. Lett. B 385, 421 (1996); Allen, B. and Brustein, R., Phys. Rev. D 55, 970 (1997). 21. Ratra, B., Astrophys. J. Lett. 391, LI (1992). 22. Gasperini, M. Giovannini, M. and Veneziano, G., Phys. Rev. Lett. 75, 3796 (1995); Lemoine, D. andLemoine, M., Phys. Rev. D 52,1955 (1995). 23. Buonanno, A., Meissner, K. A., Ungarelli, C. and Veneziano, G., JHEP 1, 4 (1998). 24. Brustein, R. and Hadad, M., Phys. Rev. D 57, 725 (1998). 25. Kulsrud, R. M., Cen, R., Ostriker, J. P. and Ryu, D., Astrophys. J. 480, 481 (1997). 26. Copeland, E. J., Easther, R. and Wands, D., Phys. Rev. D 56, 874 (1997); Copeland, E. J., Lidsey, J. E. and Wands, D., Nucl. Phys. B 506, 407 (1997). 27. Durrer, R., Gasperini, M., Sakellariadou, M. and Veneziano, G., Phys. Lett. B 436, 66 (1998); Gasperini, M. and Veneziano, G., Phys. Rev. D 59, 043511 (1999); Phys. Rev. D 59, 043503 (1999). 28. Peebles, P. J. E., "An isocurvature CDM cosmogony. I and II", astro-ph/9805194, and astro-ph/9805212.
© Physica Scripta 2000
253
29. Gasperini, M. and Giovannini, M., Phys. Lett. B 301,334 (1993); Class. Quant. Grav. 10, L133 (1993); Brandenberger, R., Mukhanov, V. and Prokopec, T., Phys. Rev. Lett. 69, 3606 (1992); Phys. Rev. D 48, 2443 (1993). 30. Fischler, W. and Susskind, L., "Holography and Cosmology", hep-th/9806039; see also Bak, D. and Rey, S.-J. "Holographic principle and string cosmology", hep-th/9811008; Biswas, A. K., Maharana, J. and Pradhan, R. K., Phys. Lett. B 462, 243 (1999). 31. see, e.g., Penrose, R., "The Emperor's new mind", (Oxford University Press, New York, 1989), Chapter 7. 32. Veneziano, G., Phys. Lett. B 454, 22 (1999). 33. Lukas, A., Ovrut, B. A. and Waldram, D., Phys. Lett. B 393, 65 (1997); Nucl. Phys. B 495,365 (1997); Larsen, F. and Wilczek, F., Phys. Rev. D 55, 4591 (1997); Kaloper, N., Phys. Rev. D 55, 3394 (1997); Lu, H., Mukherji, S. and Pope, C. N., Phys. Rev. D 55, 7926 (1997); Poppe, R. and Schwager, S., Phys. Lett. B 393, 51 (1997); Lukas, A. and Ovrut, B. A., Phys. Lett. B 437, 291 (1998); Kaloper, N., Kogan, I. and Olive, K. A., Phys. Rev. D 57, 7340 (1998).
Physica Scripta T85
Physica Scripta.Vol. T85, 254-258, 2000
A Search for Galactic Dark Matter M. Spiro, Eric Aubourg and Nathalie Palanque-Delabrouille EROS Collaboration, DAPNIA/SPP, Centre d'Etudes de Saclay, F-91191 Gif sur Yvette, France Received November 30, 1998; accepted August 2, 1999
Abstract Ten years after Paczynski's proposal to use gravitational microlensing as a tool for discovering dark stars and four years after the identification of the first candidate events, searches for microlensing events have opened a new window in astronomy and started to yield quantitative information on the contribution of dark compact objects to the dark halo of the Galaxy. A review of the results obtained so far is presented, along with recent highlights in the direction of the Small Magellanic Cloud.
massive deflector located near the line of sight between the star and the observer. This forms two distorted images of the source, as illustrated in Fig. 2. In the particular configuration where all three objects are perfectly aligned, the two images merge into a ring, whose radius is called the Einstein radius RE RE = ^—2-DOSX(1-X)
1. Introduction For the past 70 years, evidence has been accumulated for a large amount of Dark Matter on all scales, from individual galaxies to the whole Universe. The most compelling piece of evidence comes from the observation of the velocities at which matter rotates about galaxies. The rotation curves of most spiral galaxies show the following trend: after an initial rise, they flatten and tend to remain more or less constant as one goes to larger and larger radii, instead of falling with the expected l/*/R dependence beyond the visible component. This implies a mass distribution extending far beyond the optical radius of the galaxy. To make up for this discrepancy, a dark halo component is added to the galactic structure. This situation is quite universal, and the Milky Way is no exception: its flat rotation curve out to about 20 kpc from the Galactic Center implies a mass at least three times as large as that derived from the study of its visible components, including stars and gas. There is a wide range of candidates to explain this dark matter, both baryonic — molecular Hydrogen clouds, stellar remnants such as neutron stars or black holes, brown dwarfs (< 0.08 M 0 ) , red dwarfs (~ 0.1 M Q ), white dwarfs (~ 1 M Q ) - and non-baryonic - WIMPS, axions or massive neutrinos. The comparison of the limits on the content of the Universe in baryonic dark matter derived from Big Bang Nucleosynthesis, with the amount of Dark Matter required from dynamical arguments imply that both categories should exist (see Fig. 1). However, it is remarkable that the predicted range for baryonic dark matter and the additional mass needed on the scale of galaxies coincide exactly, making it reasonable to fill the halos of galaxies with baryonic dark matter. This prompted interest in looking for dark compact objects wandering in the halo of the Milky Way, using a new detection technique first proposed by Paczyhski [1]: microlensing.
2. Gravitational microlensing According to the principles of general relativity, the light rays from a source star are deflected by the presence of a Physica Scripta T85
(1)
where x = Don/Dos is the ratio of the distance observer-deflector to the distance observer-source and M the mass of the deflector. When probing the dark matter content of the Galactic halo, the source star is typically 60 kpc away (located in one of the Magellanic Clouds) from the observer and the lens typically a solar-mass object or lighter, so the angular separation between the images (~ 2 / ? E / D O D ) is only of the order of the milliarcsecond. Given the limited resolution of optical telescopes with current technology, only the combined light intensity can therefore be recorded. The total magnification, however, is always greater than what the observer would receive from the source in the absence of the lens, which makes the latter detectable. As
-
1 i
-
' '
1 I I :1 1II j '
J ; 1 1 1 1 111 ^ , i i o o;
\
o
\
*-i '
O
\J~^~~^
'«
V
%
"•
'•
%
.001
iI
;
\
'i i Ii Ii m m ll
.01
\ ° •
_
X
!
.i ii i| •• m i | |l |
.1
V
j
\
K % \ \
i|
II
^ \ - '
', ^ ' \ fa \ • *
1 1 1 Mil
21
\ Ci \
' %
0 I
<J
\ S 2. i
1 1 1 1 1 111
Ho>35
.j
'
iI tIN1,1 .i m III;
1
i I i I 1I1 1I 1I i M i il
10
n Fig. 1. Ho - Q plot (assuming A = 0), indicating allowed (shaded) and excluded regions. The fraction of visible matter in the Universe, £2V;S is shown, along with the fraction of baryonic matter QB resulting from nucleosynthesis, the average value of the fraction of matter in halos of galaxies flhaio a n d t n e average value of the fraction of matter in clusters of galaxies f2ciusters- Two more limits are indicated: the lower bound Ho < 35 km s _ 1 M p c - 1 obtained from white dwarf stars and supernovae, and the lower bound on the age of the Universe f0 > 10 Gyr obtained essentially from the age of globular clusters, under the reasonable assumpton that the Universe can hardly be younger than its components (adapted from [2]). © Physica Scripta 2000
A Search for Galactic Dark Matter with EROS 2 light rays trajectory
Observer O „ |3| ^
• DeftectorD
.
e n
Source starS _ /
255
in plane of deflector
,
_ D-* SQJu
'.
Image 12
Kg. 2. Deflection of light by a massive body D located near the line of sight between the observer O and the source star S. The dotted circle is the Einstein ring.
the lens moves in the halo with respect to the line-of-sight, the typical time scale At of a microlensing event is given by At = — ~ 9 ( V M / M G v
(2)
t
where v, is the transverse velocity of the lens which to first order can be taken as the rotation velocity of the galactic halo, ie 220 km/s. The probability that a given star is amplified at a given time is very low, typically 5 • 10~7 Millions of stars thus have to be monitored for years in order to ever be able to detect such a rare event. More details on the general principles of microlensing can be found in [3]. The only targets far enough to probe a large fraction of the galactic halo but close enough to resolve millions of stars are the large and the small Magellanic Clouds (LMC and SMC respectively), observable from the southern hemisphere. Various experiments are involved in the survey of stars located in these two satellite galaxies: EROS 2 (French experiment observing in Chile), MACHO (American experiment observing in Australia) and OGLE 2 (Polish experiment observing in Chile). The following sections present the latest results obtained by these experiments on the contribution of dark compact objects to the mass of the Galactic halo. Because the mass range to probe extends from planetary objects (~ 10~7 M Q ) to stellar dark objects (a few solar masses), the event time-scales that the experiments have to be sensitive to vary from a few hours to a few months (see eq. 2). Dedicated experiments and analyses are therefore performed separately to search for either small or large mass deflectors.
o i 10' 8
i 10" 7
] 10" 6
i 10" 5
i 10" 4
i 10- 3
i i i 10" 2 10- 1 1 M a s s o f t h e d e f l e c t o r s (M 0 )
Fig. 3. Exclusion diagram showing the 95% CL limits as obtained by EROS (full curve), by MACHO (dotted curve) and by a combined analysis of EROS + MACHO data (dot-dashed curve). Also shown is the 95% CL region compatible with the 6 events detected by the MACHO two year analysis.
indicated as the abscissa of the graph. The vertical axis represents the maximum halo mass fraction of the halo that could be composed of objects of a given mass M. As explained above, EROS and MACHO have chosen very different analysis techniques, and there is little overlap in exposure for the two projects. Combining both sets of data after removal of this small overlap thus yields even stronger limits [6]. The new exclusion diagram obtained is also illustrated in Fig. 3. It can be seen that not more than ~20% of the halo can be composed of objects in the mass range [10 -7 - 0.02] M 0 , at the 95% confidence level. Because we are using 5-function mass distributions and since the limit is quite flat in the mass interval mentioned above, any mass function that peaks in this range is also excluded at the same confidence level.
4. Contribution of high mass objects 3. Limits on contribution of small mass objects
4.1. Present results toward the LMC To search for planetary mass dark matter in the galactic Both the EROS 1 experiment (using photographic plates) halo, the EROS 1 CCD experiment, on the one hand, moni- and the MACHO experiment (using CCD's) have searched tored 150 thousand stars in the LMC, with a high efficiency and found long time-scale microlensing candidates. A total of ~ 80%, thanks to a very good time sampling. On the other of 10 events have been detected, two by EROS [7] and eight hand, the MACHO experiment monitored 8.6 million stars by MACHO [8], although one of them is slightly asymmetric thanks to a large coverage of the LMC, but with only a and thus often disregarded as a microlensing candidate and ~ 1% efficiency to short time-scale events since their another is an LMC binary event where the lens is most probobservational strategy was optimized for long time-scale ably located in the LMC itself. events. Neither of the two experiments found any such event, The typical Einstein radius crossing time associated to the which allowed them to set quite stringent limits on the maxi- LMC events is of the order of 40 days, which implies a surmum contribution of small mass objects to the dark mass of prisingly high most probable mass for the lenses: the Galactic halo [4,5]. Figure 3 shows the limits obtained ~ 0.5 M Q . This mass is much larger than the upper limit by either project, under the hypothesis of a "standard" of 0.08 M Q for brown dwarfs, and the lenses could be interisotropic and isothermal spherical halo entirely filled of dark preted as, instead, white dwarfs or black holes. The optical compact objects all having the same mass M which is depth implied by the mean duration of the events is com© Physica Scripta 2000
Physica Scripta T85
M. Spiro, Eric Aubourg and Nathalie Palanque-Delabrouille
patible with about half that required to account for the dynamical mass of the dark halo. Such an interpretation, however, is not quite accepted among astronomers: by observing younger galaxies where we could detect the light due to a significant white dwarf component in the halo, a limit of 10% has been set on their contribution [9]; furthermore, if indeed half of the dark halos of galaxies consisted of such stellar remnants, we should observe an enrichment of the interstellar medium in helium, which we do not. There is thus no consensus yet as to the nature of the deflectors causing the observed events. The 95% CL region allowed by the MACHO experiment due to the detection of 6 events {i.e. disregarding the LMC binary and the asymmetric event) is illustrated in Fig. 3. 4.2. Future data The interpretation of the present data is ambiguous, and huge error bars remain on both the most probable mass of the deflectors and the halo fraction in compact objects. More statistics is thus required, and several experiments are accumulating data to answer the questions of the presence or not and the nature of dark compact objects in the halo of the Galaxy: •
•
•
The EROS 2 experiment is now taking data with a completely redesigned setup and a new strategy. Using a wide field CCD camera (data taken in two colors simultaneously, with in each color a 1 square degree mosaic consisting of eight 2048 x 2048 CCD's), EROS covered 66 deg2 on the LMC during the first year of observation (August 1996 - May 1997) and a total of 88 deg 2 the second year. The exposure times and time sampling are adapted to a search for long time-scale events. The MACHO experiment is presently analyzing four years of data on 15 deg 2 in the LMC, which means an increase of a factor of 2 in time scale and 1.4 in area. Preliminary results indicate 6 new events with time-scales ranging between 15 and 110 days with an average of about 50 days. This would confirm a high mass for halo deflectors if the lenses are indeed located in the halo of the Galaxy. The OGLE 2 experiment uses an upgraded setup and started taking data in summer 1997. They now also cover fields in the Large and the Small Magellanic Clouds (their previous strategy concerned only fields toward the Galactic Center, thus probing disk Dark Matter and not halo Dark Matter).
5. Highlights toward the SMC The Small Magellanic Cloud gives a new line-of-sight through the Milky Way halo and a new population of source stars. The use of various lines-of-sight is very important since a comparison of the event rates is a powerful tool for discriminating between various shapes for the dark halo [10,11]. In addition, this allows for discrimination between various theories for the populations responsible for the LMC lenses [12]. The EROS and MACHO experiments (and more recently the OGLE experiment) thus monitor stars in the SMC to Physica Scripta T85
o
2400
log(Flux
256
4.95 4.9 4.85 4.8 4.75 4.7 4.65 4.6
2500
. _
2600
M\
M .> # /*
" ,
•
*
»
/
•
•
X
2700
2800 time (in days)
* \ \
\
X
".%» >*** **-" * -
2400
*•• f
, . . . . ,
1
2500
2600
.
2700
.
.
i
.
v ,
,
.
2800 time (in days)
Fig. 4. Light curve of microlensing candidate SMC # 1 , with a microlensing fit including blending superimposed. Time is in days since Jan. 0, 1990 (Julian date 2,447,891.5). Red light curve on top, blue on the bottom.
search for microlensing events. EROS recently published the first analysis on SMC data, whose results are presented hereafter. 5.1. First analysis of SMC data The EROS 2 experiment covers the densest 10 square degrees of the Small Magellanic Cloud. On these 10 fields, a total of 5.3 millions light curves were built and subjected to a series of selection criteria and rejection cuts to isolate microlensing candidates [13]. Ten light curves passed all cuts and were inspected individually. Several correspond to physical processes other than microlensing (one of them, for instance, is the light curve of a nova that exploded in the SMC), and only one of the candidates passes this final visual inspection. Its light curve is shown in figure 4. Once corrected for blending - because of the high stellar density of the fields monitored in microlensing surveys, the flux of each reconstructed star generally results from the superposition of the fluxes of many source stars— the event light curve is well fitted by that of a microlensing event with an Einstein radius crossing time of 123 days, a maximum magnification of 2.6 occurring on January 11, 1997 and a X2/d.o.f. = 332/217 = 1.5. The best microlensing fit is for 70% of the monitored flux being amplified and 30% being the contribution of blending unamplified light. The source star being very bright, the value of the reduced X2 of the fit is surprisingly high. A search for periodicity was therefore performed on the light curve, and a modulation was detected, with an amplitude of 2.5% and a period of 5.2 days. Fitting again the candidate light curve for microlensing allowing for a periodic modulation yields much more satisfactory residuals than before: #2/d.o.f. = 199/214. This strongly supports the microlensing interpretation of the observed magnification. The modulation detected was later confirmed by the OGLE experiment, on their own data taken after the event occurred (first points in June 1997). They also confirm the value of the blending © Physica Scripta 2000
A Search for Galactic Dark Matter with EROS 2 coefficient of the source star since their new camera allows the separation of the two components of the blend and thus the individual measure of each of the two fluxes. 5.2. Estimate of Halo fraction in compact objects and lens mass The time-scale of the observed event allows one to estimate the fraction of the halo that can be composed of dark compact objects generating microlensing events, independently of their mass. Assuming that the deflector is in the halo of the Milky Way, and considering a standard halo model {i.e. an isotropic and isothermal spherical halo), the EROS experiment estimated that the detected event is compatible with about 50% of the mass of the halo in dark compact objects. This fraction can vary by as much as a factor of two when considering other halo models (flattened halos for instance, or thinner halos and thicker disks so as to reproduce the rotation curve of the Galaxy but have less mass in the dark halo component). Using a likelihood analysis also based on the time-scale of the detected event, the most probable mass of the deflector generating the event can be estimated. Under the assumption ofa standard halo composed ofdark compact objects having a single mass M (ie the mass function is supposed to be a Dirac-function), the most probable mass of the Halo deflector, given with \o error bars, is:
M = 2.6t\l
M
o-
(3)
The event has the highest time-scale observed so far, and consequently the highest most probable mass. Only a neutron star or a black hole could be that massive and yet be dark. It is even harder than for the LMC events to explain how the halo of the Galaxy could be filled (even partially) with such heavy dark objects. Other interpretations therefore have to be looked at seriously. 5.3. Interpretation as SMC self lensing The very long time-scale of the observed event suggests that it could show measurable distortions in its light curve due to the motion of the Earth around the Sun (the parallax effect: [14]), provided that the Einstein radius projected onto the plane of the Earth is not much larger than the Earth orbital radius. The first detection of parallax in a gravitational microlensing event was observed by [15]. No evidence for distortion due to parallax is detectable on the light curve, implying either a very massive deflector with a very large Einstein radius, or a deflector near the source. The absence of parallax detection implies the following relation between the mass of the deflector and its distance to the observer, at the 95% CL:
where, as previously, x = Z>OD/.DOSIf the deflector is in the halo of t h e Galaxy (assuming a standard halo, x < 0.66 at the 95% CL) this yields a lower limit on the mass of the deflector: M > 0.6 M©. If the deflector is located in the S M C itself, 1 - J C ~ 1/10 and the mass of the deflector is then M ~ 0.1M©, typical of a brown dwarf or faint star in the SMC. © Physica Scripta 2000
257
To validate a possible SMC self lensing interpretation of the first event detected toward this new line-of-sight, it is necessary to check whether the SMC stellar population could provide such an event, in terms of duration and optical depth (probability that at a given time, a given star be magnified). Various authors have suggested that the SMC is quite elongated along the line-of-sight, with a depth varying from a few kpc to as much as 20 kpc, depending on the region under study. We will approximate the SMC density profile by a prolate ellipsoid:
where z is along the line-of-sight and r is transverse to the line-of-sight. The depth h will be a free parameter, allowed to vary between 2.5 and 7.5 kpc. The values of the various parameters of the model are fit to the isophote levels of the cloud (which yields So — 400 M 0 pc~ 2 and r<j = 0.5 kpc) and considering a mass-to-light ratio of 3 M 0 /L© (typical of the values measured in the disk of the Milky Way). For h = 2.5, 5.0 or 7.5 kpc, the predicted optical depths are T = 1.0 • 10~7, 1.7 • 10~7 or 1.8 • 10~7 respectively, to be compared with the experimental optical depth of 3.3 • 10~7. Considering the very limited statistics we have, the model is consistent with the observations. Finally, considering a velocity dispersion of 30 km/s in the SMC (and the 123 days time-scale of the event), the mass M of the deflector can be estimated according to an assumed distance between the source star and the lens. If the deflector is 5 kpc (resp. 2.5 kpc) from the source, its mass is M ~ 0 . 1 M G (resp. 0.2 M Q ), compatible with the results obtained from the parallax analysis. An SMC self-lensing interpretation of the first microlensing event detected toward this new line-of-sight is thus quite plausible. 5.4. A Binary Lens towards the SMC After the detection of this first SMC event, the MACHO collaboration alerted the microlensing community of an ongoing microlensing event (IAU circular 6935), which was later identified as a binary source event. In that case, the variation of the amplification is no longer simple: the gravitational potential of the double lens gives rise to cautic lines. When the source star crosses such a caustic, the amplification becomes singular. It is thus possible to resolve the finite size of the star by measuring the duration of the caustic crossing. The measurements obtained on this microlensing event allowed to predict the date of the second caustic crossing, June 18,1998. All microlensing collaborations took data this night. Among them, the PLANET collaboration obtained well sampled data at the time of the maximum, and the EROS collaboration equivalent data at the end of the caustic crossing (see Fig. 5). Both data sets allowed to put constraints on the duration of the caustic crossing [16,17]. Combining this result with an estimate of the size of the source star, it was then possible to put limits on the proper motion of the lens. The most plausible interpretation for this event is that the lens lies in the SMC itself: only 7% of the halo population has a proper motion compatible with the one measured. Physica Scripta T85
258
M. Spiro, Eric Aubourg and Nathalie Palanque-Delabrouille EROS - DIFFERENTIAL PHOTOMETRY
C
8000 -
.
:
ti
Dotes from June 0 UT 1998
6000 -
T
4000 2000
I i
-
,,
.
.
I
18.2
I
I
I
18.25
I
I
I
18.3
,
,
,
,
I
13.35
.
,
I
,
I
18.4
I
,
,
,
16.45 time (days)
m
14000 r
g 12000
r
10000 j-
1 t+ f
The status of microlensing experiments and their implications on the galactic structure can be summarized in a few words. With about 100 microlensing events detected toward the Galactic Center, nearly 10 toward the LMC, 1 toward the SMC and none yet toward the Andromeda Galaxy (M31), there is strong evidence for the existence of a bar in the bulge and for lenses residing either in the halo of our Galaxy or in the LMC/SMC themselves. The main question now raised by these results is to determine where the lenses generating the detected events belong. Are they halo objects or intrinsic to either cloud? More statistics is still being accumulated. The MACHO experiment will run until 1999, EROS 2 plans to run until 2002 and OGLE 2 is just starting to take new data. The answer to this problem can then come from at least three possible studies:
*
8000 6000 4000 -
•
L •• %
•• 2000 4 0 '* » , w . % * # * . A . « . A . « . „ . * . . « . ¥ . t -•..•*.,— - , , , , I . . , . I , i i . I i , , i I , , , . 18.2 18.25 18.3 18.35 18.4 18.45
•
time (days)
Fig. 5. Differential photometry of EROS data taken on 18 June. R data on top, B data on bottom, in ADU.
• 6. Conclusions Two targets have been explored so far, in the search for dark matter in the halo of the Milky Way. They are the Large and the Small Magellanic Clouds. The LMC data collected by the MACHO and the EROS experiments have allowed them to exclude any major contribution to the dark mass of the halo from compact objects in the mass range 1 0 " 7 M o -O.O2M 0 . Eight events compatible with microlensing by halo lenses were detected, with an average time-scale of 40 days, which could be interpreted as about 50% of the halo dark matter in the form of ~ 0.5M© objects. A huge controversy remains as to the nature of these objects. The SMC data has yielded one event found during the analysis of the first year of data, and one binary event detected online by the MACHO group. The first event has the longest time-scale observed so far: 123 days. If the lens causing the event is assumed to be in the halo of the Galaxy then its most probable mass is 2.6M Q , with a lower limit of 0.6M o coming from parallax analysis. Such a high mass is very hard to explain. A more plausible explanation is to assume that the lens and the source star are both located in the SMC. For the second event, caustic crossing time indicates clearly that the lens is in the SMC. The mass of the deflectors would then be typical of that of a faint star in the SMC, and the experimental optical depth compatible with a "thick disk" model of the SMC.
Physica Scripta T85
The comparison of the time-scales of the events toward the LMC and the SMC. They are expected to be similar if the deflectors are in the Galactic halo but different (due to the different velocity dispersions) if they are intrinsic to each cloud. The analysis of the spatial distribution of LMC events. The events are expected to be distributed evenly over the entire cloud if the lenses belong to the Galactic halo, while they should follow the stellar density of the LMC if they are LMC stars themselves. Finally, because the disk of the Andromeda Galaxy is slanted with respect to the line-of-sight, different fractions of its halo will be probed according to which end of the disk is being monitored; this will yield a larger number of events on the far side than on the near side, for microlensing events produced by deflectors in the halo of M31. Several experiments are exploring this line of sight and the first results are expected soon.
References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19.
Paczynski, B., Astrophys. J. 304, 1 (1986). Copi, C. and Schramm, D. Comments Nucl. Part. Phys., 22, 1 (1996). Paczynski, B., Ann. Rev. Astron. Astrophys., 34 (1996). Renault, C. et al., (EROS coll.), Astron. Astrophys. 324, L69 (1997). Alcock, C. et al., (MACHO coll.), Astrophys. J. 471, 774 (1996). Afonso, C. et al., (EROS and MACHO coll.), Astrophys. J. 199, LI 2 (1998). Ansari, R. et al., (EROS coll.), Astron. Astrophys. 314, 94 (1996). Alcock, C. et al., (MACHO coll.), Astrophys. J. 490, 59 (1997). Chariot, S. and Silk, J., Astrophys. J. 445, 124 (1995). Sackett, P. and Gould, A., Astrophys. J, 419, 648 (1993). Frieman, J. and Scoccimarro, R., Astrophys. J. 431, L23 (1994). Zhao, H., submitted to Astrophys. J. (astro-ph/9703097) (1997). Palanque-Delabrouille, N. et al, (EROS coll.), Astron. Astrophys. 332, 1 (1998). Gould A., Astrophys. J. 392, 442 (1992). Alcock, C. et al., (MACHO coll.), Astrophys. J. 454, L125 (1995). Afonso, C. et al, (EROS coll.), Astron. Astrophys. 337, L17 (1998). Albrow, M. D. et al., (PLANET coll.), 1998, astro-ph/9807086. Palanque-Delabrouille, N., PhD thesis, University of Chicago and Universit'e de Paris 7 (1997). de Vaucouleurs, G. and Freeman, K. C , in "Galaxies," 163 (1970).
© Physica Scripta 2000
Physica Scripta.Vol. T85, 259-266, 2000
Dark Matter Tomography J. A. Tyson Bell Labs, Lucent Technologies, Murray Hill, NJ 07974, USA Received January 1, 1999; accepted August 2, 1999
PACS Ref: 98.62.Sb, 95.30.Sf
Abstract Gravitational lensing provides an opportunity to directly image dark matter concentrations. This article is a short review of recent progress in strong and weak gravitational lens tomographic imaging of dark matter. The lens-distorted images of background galaxies are inverted, yielding a 2-D map of the projected dark matter overdensity. The maps cover a range of scales and thus constrain the nature of dark matter. On 10 kpc scales, high resolution strong lens inversions reveal a soft core in the mass distribution, not seen in scale-free cold dark matter «-body simulations. On 10 Mpc scales, statistical weak lensing inversion maps of mass show the outer mass profile of overdensities in the linear regime.
1. Introduction The nature and distribution of dark matter plays a central role in cosmology and structure formation. The average matter density Qm has been estimated mainly through extrapolating indirect observations of luminous clusters of galaxies to the rest of the Universe. If we could explore mass structure and pin down Qm via direct observations, this would significantly clarify cosmology. It may be premature to conclude that we have accounted for all the dark matter, since until very recently we have relied on the notion that radiation traces mass. The cold dark matter (CDM) + inflation scenario is consistent with a large body of indirect observations: measurements of the anisotropy of the cosmic microwave background, redshift surveys of the distribution of luminous matter today, and supernova probes of acceleration. However, observations have only begun to discriminate between different inflationary models and versions of CDM. Direct sampling of the mass distribution would be a useful addition to the array of cosmological observations. Most of the matter in the Universe is of unknown form and dark: Stars (and related material) contribute a tiny fraction of the critical density, Qlum = 0.003 ± 0.001/r 1 ~ 0.004, while the amount of matter known to be present from its gravitational effects contributes around 100 as much. Several methods for estimating the amount of matter all seem to converge on £2matter ~ 1/3; they include measurements of the masses of clusters of galaxies and the peculiar motions of galaxies. Finally, the theory of big-bang nucleosynthesis and the recently measured primeval abundance of deuterium [1] tightly constrain the baryon density: ^baryon = (0.02 ± 0.002)/r 2 ~ 0.05. The factor of ten discrepancy between this number and dynamical and lensing measurements of the matter density is evidence for nonbaryonic dark matter. It is the prospect of directly measuring Qm and mass structure that is the subject of this paper. Gravitational lensing by individual stars or galaxies is beyond the scope of this review. © Physica Scripta 2000
A key problem with weighing the Universe with traditional techniques (velocity dispersion, X-ray gas luminosity) is that we have been literally looking under lamp posts. So it is necessary to take what we currently know about the mass content of the Universe with a degree of caution. With the new tool of gravitational lens tomography, there are several questions that may be addressed: Have we found it all? Perhaps there is dark matter outside luminous clusters and superclusters. We already know mass-light segregation occurs on some spatial scales. Does the distribution of mass inside clusters agree with the «-body theory predictions for scale-free CDM? Does the mass halo cut off sharply at large radius, as seen in n-body simulations? Is there dark matter elsewhere and what are its morphological properties? Fundamentally, we need to weigh space. The path of a photon from a distant source is bent as it passes by the foreground mass, making the source appear at an altered position. This light-bending is accompanied by another first-order effect, systematic image distortions. If the background source image is resolved then this image stretching is observable. Gravitational lens distortion of background galaxies enables calibrated measurements of the distribution of dark matter in the Universe. This lens phenomenon is most naturally divided into two broad classes: strong lensing and weak lensing. In weak lensing the gravitational deflection angles are very small and single sources produce single distorted images. In strong lensing, sources appear highly distorted and can form multiple images. Whether a given source is weakly or strongly lensed depends on the impact parameter of the ray: whether its image appears outside or inside the critical "Einstein" radius [2]. Strong lensing analysis constrains the mass distribution in the parts of the lens which exceed the critical density for image splitting. In cases where multiple images of a source are created by the lens, the details of the position and distortion of these sub-images are highly sensitive to the projected 2dimensional mass distribution within the lens. Weak gravitational lensing can give a direct measure of mass overdensity on scales of several kpc to tens of Mpc. In statistical weak lensing it is necessary to average over the apparent orientation of tens of sources for each mass resolution element: typically, weak lens statistical inversion uses tens of thousands of "arclets" (distorted background galaxies) over a wide field [3,4]. The largest source of noise in weak lens inversion of deep optical images is the ellipticity noise of the source population itself; galaxies are not round and they come in all orientations. A number of nonparametric algorithms have been developed for inverting the arclets, and regularized approaches have been developed to avoid edge effects and systematics near the strong lens regime. Physica Scripta T85
260
J. A. Tyson
Up to a constant sheet of mass, lensing inversion can be calibrated and has no adjustable parameters. Calibration of the mass scale can be done both through simulations and via observation of a mass standard. Mass calibration via strong lensing forms an independent check on weak lensing mass scale calibration. There are now several of these mass calibrator clusters over the sky; the redshift of the arc is known so that the mass enclosed within that radius is known. Dark matter tomographic imaging of rich mass structures in different environments may lead to some understanding of the process of structure formation. There are a variety of gravitational lens studies of projected mass on several scales which are underway or which appear feasible in the near future: •
•
•
Mass distribution within clusters of galaxies: Weak and strong lensing studies can map the mass of rich clusters of galaxies over a wide range of spatial scales, from kpc to Mpc. Of interest is the mass profile near the center (from strong lensing observations), the correlation with luminosity, and mass segregation between galaxies and the totally dark component. Outer halos of clusters: Wide-field weak lensing studies can now map the mass out towards the linear regime beyond 2 Mpc radius. N-body CDM simulations and some observations show a steeper than isothermal mass profile at large radii. Trends in mass-to-light ratio (bias) with radius and environment will be of particular interest. Normal clusters: Rich X-ray luminous clusters are rare. If there is any hope of using clusters to estimate Qm the distribution of mass in and around normal clusters must be understood. I will describe a weak lensing survey of X-ray selected normal clusters. Inter-cluster volume: Statistically significant probes of mass in so-called "blank fields" will become feasible soon. Already, there are several pilot projects under way to map shear over large angular scales in order to study the local mass distribution on tens of Mpc scales.
into the blue. In a redshift-magnitude plot of hundreds of galaxies the trend to redshift ~ 1 at 25th B magnitude is clear. A typical galaxy seen at z = 1 may be a 0.1 L* galaxy, so an imaging survey at 26th magnitude would cover a wide range in redshifts extending from 1 to 3. While the redshift of each background galaxy is not accurately known, most of them are relatively far behind the foreground lenses studied to date. For arclet inversion of z < 0.3 lenses, the lack of detailed redshift data for each of these sources produces less than a 10% mass scale error. One can do even better: if deep imaging is done at several wavelengths, there is enough low resolution spectral information to supply "photometric redshifts", accurate to Az = 0.1, so galaxies may be separated into broad foreground and multiple background redshift bins. Without luminosity evolution these distant background galaxies would have prohibitively faint surface brightness (SB ~ (1 + z)~ ). Even with their luminosity evolution, very faint levels of surface brightness must be achieved. At any surface brightness limit, there is redshift sample selection; there is a huge increase in the number of galaxies behind a redshift 0.5 lens by going from 26 to 29 mag arcsec - 2 surface brightness. Galaxies fainter than 26 B magnitude also cover a wide range of angular scales, with typical half-light diameters of 1 arcsecond [5]. Image quality is critical since galaxies must be resolved in order to measure their ellipticity. 3. Clusters of galaxies
On very large scales, the currently favored notion of hierarchical structure formation predicts that mass and light should be correlated. As overdense regions grow and their density rises into the non-linear regime their internal density • fluctuation spectrum is modified. However, it is hoped that at least the matter content of clusters is representative of the Universe. It may be too much to hope that their mass-to-light (M/L) ratio is also representative. Traditionally, clusters have been found via their optical or X-ray luminosity. Masses of galaxy clusters have played a central role in our understanding of dark matter problems. 2. The sources Cluster masses can be estimated by three different techNature has provided us with a high density of distant blue niques which, on scales of 1 Mpc, give consistent results. galaxies, one every few arcseconds on the sky; log N = 2.03 The first, which dates back to Zwicky (1933), uses the + 0.45(2?-27) arcmin - 2 mag - 1 . Since many of the faint measured velocities of cluster galaxies and the virial theorem galaxies are resolved and are distributed up to high redshifts to determine the total mass (i.e., KEgai ~ \PEgai\/2). they may be used in statistical gravitational lens studies of Assuming further that the M/L ratio of clusters is the same foreground mass distributions. If galaxies were a few times as the mean M/L ratio in the Universe, one can combine smaller or less luminous in the past, this "cosmic mirage" measurements of the dispersion of velocities of hundreds would be extremely difficult to use. The sources must have of galaxies in several clusters with a measurement of the mean redshifts large compared with the lens, a number density light density to arrive at an estimate [7] of the mean mass denon the sky sufficient to sample the lens shear field on relevant sity QM ^ 0-25. If clusters have more luminosity per mass scales, an intrinsic angular diameter larger than the ratio of than average, this technique would underestimate QM- Since seeing FWHM to the magnification of the lens, and other the bias relation between mass and light is probably environproperties (blue color and unique surface brightness) ment dependent through detailed mechanisms for star enabling efficient separation of the sources from formation, the ansatz of a universal M/L is in some doubt. superimposed galaxies within the lens and foreground. The second method uses the temperature of the hot X-ray The extreme blue colors and enhanced luminosity of the emitting intracluster gas and the virial theorem to arrive faint galaxies results from their large look-back times: we at the total mass. In this method, one must assume that see these young galaxies as they were billions of years the hot gas is pressure supported and in equilibrium. The ago just forming a generation of hot and UV-bright stars density and temperature profiles of the hot gas must also at redshifts of 1 to 3, so that their UV excess flux is redshifted be obtained. Physica Scripta T85
© Physica Scripta 2000
Dark Matter Tomography 261 The third and most direct method is gravitational lensing: light from distant galaxies scatters off the mass gravitationally. For small impact parameters, close to the cluster center, lensing is strong enough to produce multiple images. Farther out, lensing distorts the shape of distant galaxies, allowing the cluster surface mass density to be mapped directly. X-ray measurements can determine the amount of hot intracluster gas, and with the plausible assumption that clusters are fair samples of matter in the Universe, baryogenesis yields an indirect estimate of Qu- A compilation of the existing data on cluster baryonic mass [8] and cluster total mass inside the same radius gives 3/2 MB/MTOT = (0.07 ± 0.007) h~ ~ 0.15. Invoking the "fair sample" ansatz QB/®M = MS/MTOT, the accurate determination of QB can be used to infer QM = (0.3 ±0.05) h~l/2 ~ 0.4. [h = H0 in units of 100 km s _1 Mpc"1.] Note that this is about twice the estimate obtained via the universal M/L assumption. How is this dark matter distributed?
MASS CONTOURS
-125
-25C -250
250
4. Strong lensing in clusters Multiple lensed images of the same resolved source galaxy contain rich information related to the mass distribution in the lens. In cases where portions of the lens exceed critical density, producing multiple images of some sources, it is possible to obtain higher resolution in the mass map by combining weak lens inversion at large radius with the strong lensing constraints in the inner region. One approach to solving this inverse problem is to use a parametric lens model in which multiple mass components are parametrized by their centroid, mass profile, and ellipticity. A regularized iteration using the detailed multiple images as well as the weak lens inversion solution for the mass at large radius, can lead to a robust solution for the lens mass map at high resolution. This has the desirable property that the resulting mass map has high resolution in regions of high mass density. In practice we have found that a reliable way of converging to the solution is to demand that the strong lensed source images, when unlensed by the estimated mass distribution in the lens, reconstruct to an identical single image [9], and that this image of the source maps to the observed arcs (and only these arcs) in the image plane when lensed by this same mass. [It is necessary to use the lens mass distribution, rather than starting with a 2-dimensional lens potential, because non-circular lens potentials can be unphysical in their corresponding mass reconstructions.] With my colleagues Ian Dell'Antonio and Greg Kochanski, we have developed a code for this weak-strong mass reconstruction and applied it to complex strong lens simulations as well as deep HST imaging of two massive clusters, the cluster CL 0024+17 at z = 0.4 and A1689 at z = 0.2. As an example, Figure 1 shows a high resolution mass map of the cluster 0024+1654, based on parametric inversion [10]. This lens creates eight well resolved whole or partial sub-images of a background galaxy, seen in deep imaging with the Hubble Space Telescope. Of course we do not know what the source galaxy looks like, so we cannot start with that prior; the source galaxy optical image must be part of the model to be optimized along with the mass distribution in the lens. The parametrized source is then © Physica Scripta 2000
Fig. 1. A map of the projected mass density in the redshift 0.4 cluster CL0024+1654, obtained by fitting the observed distorted images of a z = 1.6 source galaxy. Dark matter halos of individual galaxies in the cluster are seen, but the major dark matter component is smoother than CDM simulations.
ray-traced through the lens plane, and the resulting image is compared pixel-by-pixel with the HST image. We parametrize the mass distribution as a cluster of mass concentrations ("mascons"). Each mascon is based on a power law model [11], for the mass density S(r) vs projected radius r, with an inner core radius and an outer cutoff radius. Mass components not associated with optically observed galaxies have their positions in the lens plane free. In the case of CL0024, the mass and source models were determined by 512 parameters. However, there are over 3800 significantly nonzero pixels in the arcs. The sparse sampling of the lens plane by the arcs yields the highest mass resolution only near the arcs; cluster mascons projected near an arc have the effect of their mass distribution highly magnified. For galaxies that are more than about 5 arcseconds from the arcs, only their total mass matters. Galaxies farther than about 20 arcseconds from the arcs are parameterized in groups. In this way 18 cluster galaxies have well determined masses via strong lensing, and this is used to calibrate the remaining cluster galaxies via the fundamental-plane relation. The optimizer uses simulated annealing and bootstrap resampling to avoid false minima. The developing model predicted the central image, based on the three major arcs, and then correctly predicted the multiple subimages near the outer arcs. One of these was buried under a cluster galaxy and was recovered as a result of this prediction by modeling and subtracting the foreground galaxy's light. The mass distribution found in CL0024 is remarkably relaxed. Excluding mass concentrations centered on visible galaxies, more than 98% of the remaining mass is represented by a smooth elliptical concentration of dark matter centered near the brightest cluster galaxies, with a 33 A-1 kpc soft core. The ellipticity is similar to the X-ray isophotes. Physica Scripta T85
262
J. A. Tyson
The non-galaxy dark matter distribution we observe in CL0024 shows no evidence of in-falling massive clumps, and is far smoother and symmetric than typical simulated clusters in either Q = 1 or Q = 0.3 CDM cosmologies. Integrated to 107/r1 kpc radius, the rest-frame mass to light ratio is M/Lv = 238 ± 16 h (M/LV)Q, rising with radius (see Figure 2). Because galaxies were brighter in the past, this translates to a M/L ratio of 400 h now. The soft core suggests either nongravitational interactions or gravitational coupling to a dissipative baryonic component of significant mass fraction. While HDM can produce soft cores, HDM is not consistent with the high density of dark matter that we find in the individual cluster galaxies. Because ofthe relatively low X-ray luminosity (0.7 L£), it is unlikely that this can be attributed to hot gas alone. High resolution mass maps of clusters of galaxies will be useful to compare with future n-body/ gas-dynamical simulations. None of the recent simulations show evidence of a soft core, in disagreement with these observations. Indeed, as the resolution of simulations increases from 100 to 30 kpc the central mass becomes more singular [12]. We have recently found a simiar soft mass core in the more massive cluster A1689. While internal cluster mass maps may constrain the nature of dark matter, there are two ways in which confining the study of mass to clusters may lead to a biased measure of the mass content of the Universe. First, the M/L ratio within the cluster may not be typical; i.e. a higher baryon fraction may lead to more light per mass. Second, being selected because of their luminosity, clusters may be a statistically biased sample of mass. If we examine other regions in equal detail, we will be able to examine these possibilities.
stretched single images of many independent background galaxies projected near the cluster [4]. This technique was applied first to massive clusters where there was already indirect evidence for a high mass contrast. If the source angle relative to the lens is larger than the Einstein radius, multiple images of a source will not be formed, but the single image of the source will be elongated. Since there are 50-100 independent resolved sources per square arcminute, this method can be applied anywhere on the sky, and gives good signal-to-noise inversions in regions of low shear which are much more representative of random regions. The gravitational lens distortion is quantified using the intensity-weighted second moment of the galaxy image orthogonal and along the radius relative to the lens center [4]. A dimensionless scalar alignment T, calculated from these principle axis transformed source ellipticities, is related to the projected mass density clumping and is defined at each point in the image plane via the (r,6) principal-axis transformed second moments of the background galaxy image: j v r ) _ *< ' "> ~ ' " • _
iee + hr
2
(1~K)y
(l-K)
2
2
2
+y
y(r)
m
1 - »c(r)'
where the convergence [13] K(r) = Z(r)/Ic and the shear y(r) = [I(r) - Z(r)]/Zc, and where Lc is the critical surface mass density, related to the distance ratio: Ec = c2/ (4nGD). The distance ratio for a foreground-background pair is [6] D =
{\ ~ q<> - d\d2)(dx - d2) (\-qo~d2){\-d2){\+zt%y
(2)
5. Statistical weak lensing
where dx = JI + q0zti and d2 = y i +q0zbg. The tangential second moments are formed from the intensity weighted moments, regularized by a galaxy light distribution prior:
Historically, the first solution to this inverse imaging problem was via statistical weak lens inversion, in which the mean shear field is estimated from the tangential alignment of
iee — M20 sin2 <> / + M02 cos2
NFW model
{)
where
where 8x = (x — < x >g) and similarly for 8y, the sky intensity near this background galaxy is given by /„> and optimal normalized Gaussian weights (adaptive kernel) W(dx,8y, I(x, y)) are calculated from the half luminosity radius. An unlensed population of galaxies randomly placed and oriented will give a net distortion T(r) of zero at every point in the image plane, while a population of lensed galaxies will give a positive value at the point corresponding to the lens center. At some point off the lens center T(r - r0) is 1 10 100 R ( h-' kpc ) proportional the smaller projected mass at that new point, as shown in Fig. 3. Fig. 2. The projected mass density of CL0024 and the projected light density In 3-D simulations of deep shear surveys, it is apparent are plotted vs radius. The diffuse dark matter component dominates, and has a soft core. Note also that the M/L ratio rises with radius. that in addition to the shear noise arising from intrinsic Physica Scripta T85
© Physica Scripta 2000
Dark Matter Tomography
263
by E(r) = ZcCB(r) /
Fig 3. The sum of the tangential component of the shear around the solid circle is simply related to the average mass density interior. This is also true for the offset (dashed) circle. This provides a method for mapping the projected mass density on the sky, producing images of the dark matter. The resolution is given by the size of the circle. In practice, a weighted average of an array of sizes is used in deriving the mass map.
source galaxy random ellipticities there are errors arising from occasional chance projections of one arclet very near or on top of another. We have found that a regularized iteration, in which the kernel W(Sx, Sy) is elliptical and maximum likelihood matched to each arclet, is even more effective in reducing the tail of the arclet ellipticity noise.
6. Weak lens mapping of clusters To construct an image of the gravitational lens projected mass distribution, the distortion statistic T(r) may be computed over a grid of positions as candidate lens centers. In the weak lensing limit it can be shown that the tangential alignment T is a measure of the mass contrast [3,13]: T(r) = 2[IU
(5)
(7)
T(r) d log r + I(r, r m a x ),
'/
where C is the seeing correction obtained via simulations. The function B is given by #(r) = (1 - r 2 / ^ ) - 1 , a n d 2(r, r max ) is the average density in the annulus between r and r max . For a sufficiently large field, I(r, r max ) is small compared with with peak density. This arclet inversion is called "aperture mass densitometry." Until recently, weak lens studies have been confined to rather massive clusters [4,15-17]. An example is the massive X-ray luminous cluster RXJ 1347 at z = 0.45, Lx = 2 x 1046 erg s _1 , whose mass profile appears isothermal out to 1.5 h~l Mpc [18]. Figure 4 shows this mass profile reconstruction, with 1st order strong lensing corrections included. Other massive clusters show isothermal profiles over a limited range of radius, becoming steeper at large radii, as predicted by n-body simulations. From the inversion of 6000 arclets surrounding the rich z = 0.18 cluster A1689, a steeper than isothermal profile was found beyond 300 / r 1 kpc radius [19]. The mass core radius is smaller than most observed X-ray core radii in nearby clusters, suggesting that the X-ray gas may be less relaxed dynamically than the dark matter. In rich compact clusters mass appears to trace the cluster red light, on scales of a few xlOO hrx kpc, with rest-frame V band M/L ratios of a few hundred h in solar units. Lensing by a group of clusters apparently undergoing merging has been detected [20]. Because faint blue galaxies become more common at redshifts above 0.5, most studies of weak lensing have been confined to lenses around redshift 0.2 to 0.4. However, clusters exist at higher redshift and two of them near z = 0.8 have been studied recently by weak lensing [21]. As multiband deep imaging becomes more common in lensing work, color-redshifts will enable weak lens studies of clusters at higher redshift as well as direct surveys for low mass structures. 6.1. Normal cluster survey The massive clusters are rare and may not have properties typical of the rest of the Universe, such as M/L ratio. Moreover, the more common less massive clusters are easier
At any point r in the image plane we can sum over the tangential alignment of all source images about that point, creating a continuous scalar distortion statistic T:
Projected R a d i u s ( a r c s e c ) 60
80
100
200
6000
t(r)=
f K{u)T(r - u)Au,
(6)
where the apodization kernel K(u) weights source images at large radius less, and is generally of the form K(u) = (u2 + ulY1. For t to be simply related to the mass [14] K{s) must asymptotically approach the power law s~2 at large s. The distortion t at any point is related to the mass density contrast, since any mass can be represented as a sum of cylindrically symmetric distributions, and light bending angles from different mass components add. The solution to this inverse problem for the contrast of the projected lens mass density E is then given by a simple integral of the shear over radius from the lens center. The average projected mass density interior to radius r is given © Physica Scripta 2000
4000
: IOOO
•J 800 i 600 400
"H.*
*"HH
0.1 "
iT
T
-I 0.05-1
iiii- J 100
500 Projected Radius ( h _ l kpc)
1000
1500
Fig. 4. The radial mass distribution of the cluster RXJ1347, with strong lensing corrections [18], as in Eq. (1). The cluster RXJ 1347.5-1145, at a redshift of z = 0.451, has an X-ray luminosity Lx = 2 x 1046 erg s -1 . K = Z/Eait- The dotted line is an isothermal distribution. Physica Scripta T85
264
J. A. Tyson
to characterize in rc-body simulations, due to their larger numbers. Using the BTC we have undertaken a rigorous weak lensing survey of clusters in a fixed X-ray luminosity range of 1 — 4 x 1044 near a common redshift, forming an ensemble which may be compared to similar clusters formed in simulations at afixedepoch. Deep imaging over a 5 Mpc field around eight of these "normal" clusters, in 3-4 photometric bands, has been carried out. The resulting colorredshifts permit efficient calibration of the mass maps. On these large spatial scales any non-local bias between mass and light will be apparent. Several of these clusters show subsidiary mass concentrations where there is little or no light; multiple mass concentrations with occasional mass bridges are found. The result of applying equation 7 to one of these normal clusters is shown in Fig. 5, where the seeing-corrected mass density is followed out to 3 h~x Mpc radius. In this annular average, systematic ellipticity error is less than 0.0014, after correction for seeing.
Fig. 6. A test of low level shear detection limit. In this simulation, two dwarf cluster mass concentrations with peak density 6% of Iz [left] cause a small shear perturbation in the background galaxies. Weak lens inversion using 11,000 arclets recovers the masses [right] at high significance with shear noise tfshear = 0.003.
recovery of the two mascons is S/N = 5. Dividing by the shear-to-mass efficiency, the calibrated rms mass noise is 0.013ZC, smaller than the rms in CDM theories. No false 6.2. Detection limits mascons with the density of these two test masses were found Cold dark matter models predict a filamentary structure of in the reconstructed mass density image or in repeated mass emerging by redshift 0.5, with peak densities of realizations. The input diffuse mascons were thus detected 105pc. But most of this dark matter is distributed in less spec- at high significance. Larger smoothingfiltersin the inversion tacular mass concentrations, with projected (2-D) densities yield higher significance for overdensities covering large less than Zc ~ 1 g cm~2 and rms fluctuations around 0.02 angles. Longer integrations yield over 100,000 arclets in a Zc. What are the chances of detecting these low level 45-arcminute field, producing lower mass noise at a given mass angular resolution. Of course mass profiles from mascons with gravitational lens tomographic searches? annular averages as in Fig. 5 achiveve even lower noise. Simulated low level shear in a CCD survey for dark matter What will ultimately limit tomographic mass detection? is shown in Fig. 6, where two dwarf cluster masses (ten times The random mass density error scales as N~^e2ts, and is limsmaller than those discussed above) were placed at redshift 0.3 in a 3-D distribution of galaxies extending out to redshift ited by integration time. There are two sources of noise 3. These mascons had 40 arcsec wide soft cores (0.0627c), 600 and one source of systematic error. The photon shot noise km sec -1 velocity dispersion, and an outer mass cutoff at 500 from the night sky diffuse surface brightness may be overcome with longer integration times. In the same way, higher arcsec. redshift (fainter) galaxies may be reliably photometered. This simulates a worst case: a small 2000 x 2000 pixel part This raises the efficiency, especially for higher redshift lenses. of thefieldof the Big Throughput Camera, in moderate seeBut the source galaxy random ellipticity orientation noise ing (0.9 arcsec FWHM) and including all blue galaxies to will remain about ermS = 0.3, so lower level mass density 26th magnitude. The angular resolution of the inversion detections will have to use lower angular resolution, is 2 arcminutes, and random noise is 0.003 27c. Depolarcapitalizing on the root N averaging of source ellipticity ization due to seeing reduces the efficiency of mass recnoise over larger resolution elements. There is also systemonstruction but the resulting signal to noise ratio for atic shear error due to the finite number density of psf calibration stars; rms shear variation of stellar images has been reduced to levels of 0.0007 averaged over a 45 arcminute field. Most of this systematic shear error is fixed, and is due to the optics. This may be calibrated to better than 0.001 per square arcminute, but there remain induced systematics from focus variations. In practice, these effects limit two arcminute resolution mass reconstructions to about 0.002 Zc in multi-color deep imaging to 26th magnitude. This is over ten times lower than the rms mass density fluctuations in blank fields predicted in any CDM model.
7. Large scale dark matter Fig. 5. The projected radial mass density of a normal low X-ray luminosity cluster from the BTC Normal Cluster survey in a deep 5 Mpc field [22]. The mass density I(r) follows an isothermal relation out to 1 h~l Mpc. The dashed line is an isothermal profile. Physica Scripta T85
By concentrating on the largest luminous structures known, clusters of galaxies, we are biased to radiation. Larger scale applications of this dark matter mapping may eventually find clumped dark matter unrelated luminosity. Mosaics of CCDs make such a large scale search for coherent align© Physica Scripta 2000
Dark Matter Tomography 265 ment in the distant faint galaxies particularly attractive, and dark matter on angular scales up to degrees can in principle be studied in this way. It will be important to produce images of the dark matter structures not directly associated with clusters of galaxies. If we wish to extract only information on the statistics of the foreground mass overdensities, rather than map them, there is a tensor statistic analogous to the scalar two-point correlation function [13,23-27]. Correlations between foreground galaxies in the lens-plane and their induced weak-lens shear can in principle yield constraints on non-linear and non-local biasing [28]. The observational challenge in low shear measurement is systematic errors due to variations across the detector of the point spread function, optics astigmatism, and variable field distortion due to atmospheric refraction. These systematics can be overcome by chopping and trailing techniques, and optics calibration in rich starfields.Preliminary measurements of the arclet orientation correlation function in random fields yield shear correlations which are below the SCDM prediction. The key in finding the rest of the dark matter, even in places where there are no clusters, is to measure shear over much larger areas. Shear systematics have to be controlled at the 0.1 percent level. In collaboration with Gary Bernstein at U. Michigan, we have built a large area blue-sensitive CCD mosaic camera for the 4-meter Blanco telescope at CTIO. With this instrument we have a chance of viewing mass in ordinary places, rather than in the inner parts of massive rare overdensities. The Big Throughput Camera is a mosaic of backilluminated CCDs, giving it high throughput over a large field of view, as well as good blue sensitivity [29]. [Note: the Sloan CCD mosaic just had its "first light"; this is a much bigger array of the same type of CCDs as the BTC. Other CCD mosaics are now being applied to weak lensing observations [30]]. The BTC format is a 2x2 array of thinned, broad-band antireflection-coated 2048x2048 SITe CCDs. Mounted at the prime focus of the 4-meter Blanco Telescope at CTIO, the camera covers a 35 arcmin field of view per exposure. Each CCD is separately adjustable in position to ~ 5 urn accuracy and cooled to 170 K. The BTC's largefieldof view is ideal for searches for rare objects or events, and statistical studies of large numbers of objects. In its first two years of operation, the BTC has been used for faint quasar surveys, high-redshift supernova surveys, faint intergalactic stars, detection of faint Kuiper belt objects, and a variety of weak gravitational lensing projects. A large contiguous field of view is required to map the dark matter structures. This is done with many "shiftand-stare" exposures, resulting in a contiguous field of 47 arcminutes or more. The unique problem of large focalplane mosaics is field distortion, since they cover most of the usable field of view of large telescopes. Multiple shifted exposures cannot be combined into a single deeper image until corrections for optics distortion and differential refraction are made. These effects are corrected using a parametric model of the entire optical system and observations of astrometric standard fields at several airmasses with each filter used in the observing run. For each set of images to be combined, we fit a distortion function to each exposure, forcing the coordinates of high signal-to-noise objects to agree among exposures. To correct for astigmatism, the image is con© Physica Scripta 2000
volved with a position-dependent flux-conserving kernel designed to circularize stellar images. The flow diagram for this pipeline software is shown in Fig. 7. 7.1. Surveying dark matter So far, lens search programs have concentrated on obvious clusters of galaxies where one knows there is a deep gravitational potential well coincident with the luminous baryons. Weak lensing analysis can now be applied to blank fields to explore mass generally, without resort to light. Furthermore, evidence is mounting from strong lensing observations that the input physics for the «-body work is incomplete. It is fair to say that the process of mass clustering is not yet well understood. The recent primordial deuterium measurement together with the observed high cluster baryon density ( the "baryon catastrophe"), indicate ^matter = 0.4; this is also suggested by large-scale peculiar velocities. This excess mass in dark structures, over what has been inferred from studies of clusters, will be clustered throughout a deep probe out to high redshift. Such a mass probe of random areas relies heavily on wide-field deep imaging in superb conditions and over a wide wavelength range. While deep lensing probes of many blankfieldswill be necessary in order to test this and measure the cosmic variance, a hint of what may be in store was found in the second blank field studied: a large mass structure with peak mass density of 0.02 Zc was found in a BTC field chosen for its absence of obvious clusters of galaxies. In some blank fields examined in this way with the BTC we have discovered clusters by first detecting the mass overdensity via weak lens inversion.
8. Cosmic complementarity Along with other probes, weak lensing can play a role in untangling the cosmological constant A (if non-zero) and Qm. A significantly non-zero QA generates larger shear at high source redshifts, but negligible effects at very low redshift. Combining the data from weak lensing surveys on different scales and depths will result in an independent estimate of QM and QA- TWO weak lens surveys are in progress: one covering a very widefieldbut low lens redshift
Fig. 7. The dataflowin the Big Throughput Camera data reduction pipeline. The observed shapes of stars in the field are used to correct for low-level residual astigmatism. Physica Scripta T85
266
J. A. Tyson
matter
cision to detect less than 10 percent of the statistical weak lensing shear predicted by current Cold Dark Matter theory and the mass fluctuation spectral index. A dedicated 6-meter "Dark Matter Telescope" optimized for wide field surveys would vastly enhance our understanding of dark matter and cosmology. Wide spectral coverage, together with low aberration requirements are a challenge for normal telescopes which have multi-element glass correctors. A design being seriously discussed is the Paul-Baker 3-mirror telescope. This corrects for aberrations and is intrinsically achromatic and non-dispersive. In terms of "throughput" (collecting area times imaging solid angle), it would be 40 times more efficient than any of the wide-field mosaics now being built for existing or new telescopes.
Fig. 8. The current situation in the cosmological constant - matter plane. Three key experiments are indicated: Supernova measurements of the accel- Acknowledgements eration (SN), cosmic microwave background anisotropy measurements (CMB), and direct measurements of mass (shading). All three will see My principal collaborators in this research are Gary Bernstein, Ian Dell' Antonio, Philippe Fischer, Michael Jarvis, David Kirkman, Raja decreased errors in the next few years. Guhathakurta, Greg Kochanski, Deano Smith, and David Wittman.
and the other probing deeply to higher redshift in several References 2-degree fields. The observational situation in the A — QM plane, 1. Buries, S. and Tytler, D., Astrophys. J. 499, 699 (1998). 2. Narayan, R. and Bartelmann, M. In "Formation of Structure in the late-1998, is reviewed in Fig. 8. Three complementary experUniverse", (A. Dekel and J. P. Ostriker, Eds) (Cambridge U. Press); iments are plotted. The preliminary results of two groups astro-ph/9606001. using Type la supernovae as standard candles are shown, 3. Tyson, J. A., Valdes, F. Jarvis, J. F. and Mills, A. P., Astrophys. J. Lett. including an estimate of systematic error (neglecting effects 281, 59 (1984). of sample selection, evolution, and grey dust.) The pattern 4. Tyson, J. A., Valdes, F. and Wenk, R. A., Astrophys. J. Lett. 349, LI (1990). of anisotropy in the CMB depends upon the total energy 5. Im, M., Casertano, S., Griffiths, R. E., Ratnatunga, K. U. and Tyson, density in the Universe. The first acoustic peak in the multipole power spectrum is /peak — 200/ +JQM + ®A- These
current likelihood contours are shown as dashes. The CMB anisotropy does not measure mass at z = 1000 directly; a model relating three mass-derived effects on the radiation must be fit to the CMB anisotropy data. Note that the likelihood contours for these first two experiments are roughly orthogonal. As the new generation of satellite CMB observations map out the fine scale structure of the microwave sky, the position our Universe occupies in this plane will be strongly constrained. Finally, the shaded region is the range of fimatter: the lower limit based on extrapolating cluster masses to the field assuming identical MIL ratios; the upper limit is less well determined. As direct lensing observations of dark matter explore more of the Universe, this range will become narrower and perhaps move to the right if future explorations for dark matter discover significant high M/L structures in the vast regions between clusters. Note that the recent determination of the cosmic abundance of deuterium suggests £2matter = 0.4. It will be interesting, and a test of our cosmological model, if the results of all three tests (SN, CMB, and Lensing) are consistent. It will be even more interesting if they are not. Such a dark matter and dark energy survey is feasible, using current technology. A definitive measure of the density spectrum of the dark matter on supercluster scales is technically within reach. Deep imaging of millions of distant galaxies at several wavelengths (to get statistical redshifts from their colors) repeated in many different directions, would have the prePhysica Scripta T85
J. A., Astrophys. J. 441, 494 (1994). 6. Blandford, R.D. and Narayan, R., Ann. Rev. Astron. Astrophys. 311 (1992). 7. Carlberg, R. G. et al, Astrophys. J. 462, 32 (1996). 8. Evrard, A. E. et al, Astrophys. J. 469, 474 (1996). 9. Colley, W. N., Tyson, J. A. and Turner, E. L„ Astrophys. J. Lett. 461, 83 (1996). 10. Tyson, J. A., Kochanski, G. and Dell'Antonio, I. P., Astrophys. J. Lett. 498, 107 (1998). 11. Schneider, P., Ehlers, J. and Falco, E. E., "Gravitational Lenses" (Springer, NY, 1993), 244. 12. Syer, D. and White, S. D. M., Mon. Notes R. Astron. Soc. 293, 337 (1998). 13. Miralda-Escude, J., Astrophys. J. 370, 1 (1991). 14. Kaiser, N. and Squires, G., Astrophys. J. Lett. 404, 441 (1993). 15. Fort, B., Mellier, Y., Dantel-Fort, M., Bonnet, H. and Kneib, J. P. 1996, Astron. Astrophys. 310, 705 (1996). 16. Seitz, C, Kneib, J. P., Schneider, P. and Seitz, S., Astron. Astrophys. 314, 707 (1996). 17. Smail, I. et al, Astrophys. J. 479, 70 (1997). 18. Fischer, P. and Tyson, J. A., Astron. J. 114, 14 (1997). 19. Tyson, J. A. and Fischer, P., Astrophys. J. Lett. 446, 55 (1995). 20. Squires, G. et al, Astrophys. J. 482, 648 (1997). 21. Clowe, D., Kaiser, N., Luppino, G., Henry, J. P. and Gioia, I. M. Astrophys. J. 497, 61 (1998). 22. Dell'Antonio, I. et al, Astrophys. J., in press. 23. Gunn, J. E., Astrophys. J. 147, 61 (1967). 24. Valdes, F„ Tyson, J. A. and Jarvis, J. F. Astrophys. J. 271, 431 (1983). 25. Blandford, R. D., Saust, A. B., Brainerd, T. and Villumsen, J. V., Mon. Notes R. Astron. Soc. 251, 600 (1991). 26. Kaiser, N., Astrophys. J. 498, 26 (1998). 27. Seljak, U., Astrophys. J. 506, 64 (1998). 28. Schneider, P., Astrophys. J. 498, 43 (1998). 29. Wittman, D., SPIE, 3355, 626 (1998). 30. Luppino, G., Tonry, J. L. and Stubbs, C, SPIE, 3355, 469 (1998).
© Physica Scripta 2000
Physica Scripta.Vol. T85, 267-273, 2000
Status of Models for Gamma Ray Bursts Martin J. Rees Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 OHA, U.K. Received January 15, 1999; accepted Augiust 2, 1999
It did not take long for the theorists to become enthusiastically engaged. At the "Texas conference on Relativistic The nature of gamma-ray bursts, an enigma for more than 25 years, is now Astrophysics", held in December 1974, Ruderman [2], gave coming into focus. They involve extraordinary power outputs, and highly a review of models and theories. He presented an exotic relativistic dynamics. The "trigger" is probably a binary merger or collapse involving stellar-mass compact objects. The most plausible progenitors, menu of alternatives that had already appeared in the ranging from NS-NS mergers to various hypernova-like scenarios, eventually literature, involving supernovae, neutron stars, flare stars, lead to the formation of a black hole with a debris torus around it, the antimatter effects, relativistic dust, white holes, and some 54 15 extractable energy being up to 10 ergs. Magnetic fields may exceed 10 even more bizarre options. He noted also the tendency (still G. Details of the afterglow may be easier to understand than the initial trigger. often apparent; not only among astrophysics!), for theorists Bursts at very high redshift can be astronomically-important as probes of the distant universe. to "strive strenuously tofitnew phenomena into their chosen specialities". During the 1970s and 1980s, data on GRBs accumulated, 1. Introduction thanks to a number of satellites. Particular mention should Astrophysics is an observation-led subject: theorists gener- be made of the impressive contributions by Mazets and ally play a subsidiary role - certainly a more modest one his colleagues in Leningrad. Also important were the than their counterparts in, for instance, particle physics. extended observations made by the Pioneer Venus Orbiter But in the case of gamma ray bursts the lag between gath- (PVO). The number of detected bursts rose faster than ering data and making sense of it has been specially the number of proposed models - indeed, some of the crazier embarrassing, even by astrophysical standards. Until less early conjectures were actually ruled out. than two years ago, there was absolutely no consensus on During that period, three classes of models were pursued: what, or even where, the bursts are. Owing primarily to those in which the bursts were respectively in the Galactic the impetus of the Italian/Dutch Beppo-SAX satellite, there Disc (at distances of a few hundred parsecs), in the halo is now general agreement that the bursts (or at least a sub- (at distances of tens of kiloparsecs), and at cosmological stantial subset of them) are at high redshifts. Theorists (several gigaparsec) distances. The most popular idea during had already suggested the outline of some models for the 1980s was that the bursts were relatively local, probably "cosmological" bursts. There is now a stronger motive to in our Galactic Disc, and due to magnetospheric phenomena develop these in fuller detail. In this paper I shall review or "glitches" on old neutron stars (defunct pulsars). the basic data, and then discuss the physics of the energy It was already clear that two kinds of statistical inforproduction and its conversion into an intense burst and a mation could in principle decide the location of GRBs as prolonged afterglow. soon as enough data had accumulated, and selection effects were understood. One was the number-versus-intensity of the events, which tells us whether they are uniformly dis2. History tributed in Euclidean space, or whether we are in some sense The story of gamma ray bursts (GRBs) started in the late probing out to the edge of the distribution. The other was the 1960s, when American scientists at Los Alamos had devel- degree of anisotropy. The counts of GRBs were already suspected to be natter oped the Vela satellites, whose purpose was to monitor clandestine nuclear tests in space by detecting the associated than the classic Euclidean slope (-1.5 on a log-log plot), since gamma-ray emission. Occasional flashes, lasting a few otherwise it was hard to understand why large-area detectors seconds, were indeed recorded. It took several years before flown in balloons didn't detect more faint bursts. Flat counts these were realised to be genuine; and to be natural, rather would not of course have been unexpected if the bursts came than sinister phenomena. In 1973 a paper was published from within a bounded system such as our Galactic Disc. by Klebesadal, Strong and Olson entitled Observations of However, one would then have expected an anisotropic disGamma-ray Bursts of Cosmic Origin [1]. This classic paper tribution of sources over the sky: an enhancement towards reported 16 short bursts of photons in the energy range the Galactic plane, and perhaps also towards the Galactic between 0.2 and 1.5 MeV, which had been observed during Centre. It was therefore a real surprise when the Compton a three-year period using widely separated spacecraft. Gamma Ray Observatory (GRO) satellite, whose Burst The burst durations ranged from less than 0.1 second up and Transient Source Experiment (BATSE) offered systemto about 30 seconds, but complicated time-structure was atic all-sky coverage, with good sensitivity over the photon observed within the longer bursts. It was apparent that energy range 30 kev - 1.9 Mev, revealed that the bursts the bursts came neither from the Earth nor from the Sun, are highly isotropic over the sky. More than 2500 have now been recorded, and there is still no statistical evidence but little else was clear at that time. Abstract
© Physica Scripta 2000
Physica Scripta T85
268
Martin J. Rees
for any dipole or quadrupole anisotropy, nor for any two-point correlation [3]. The lack of any enhancement either towards the plane of the Galaxy, or towards the Galactic Centre, was recognised as a very severe constraint on the hypothesis that bursts come from the Galaxy. (Note that the bursts cannot be so ultra-local that we were not even probing as far as a scale height in the Disc: this would yield isotropy, but is ruled out by the flatness of the number counts.) The "non-Euclidean" counts imply that the surveys are probing to distances where the sources are, for some reason, thinning out; the problem is to reconcile this with the isotropy. The BATSE experiment has produced a large body of data on the spectra and time structure of bursts. Despite the large variety, there is little doubt that gamma-ray bursts are a well-defined class of objects, distinguished spectrally from phenomena such as X-ray bursters, and also from the so-called 'soft gamma repeaters' which have substantially softer spectra and are associated with young highlymagnetised neutron stars within our Galaxy. There are some apparent correlations (though no specially strong ones) between various GRB properties. For instance, shorter bursts tend to be stronger and to have somewhat harder spectra; the histogram of burst durations exhibits two distinct-seeming peaks; and the counts deviate most from the Euclidean slope (i.e. are flatter) for bursts with harder spectra [4]. The isotropy evidence tilted the balance of opinion strongly towards a cosmological interpretation of GRBs. It still remained conceivable, however, that the bursts were in our Galactic Halo, but sufficiently far out that the Sun's 8-kpc offset from the Galactic Centre did not produce a dipole asymmetry. In April 1995 an interesting debate took place on the location of GRBs, in which the two main protagonists were Don Lamb and Bohdan Paczyhski (a written version of the argument appears in [5] and Paczynski [6]). It was held in the Washington Museum of Natural History, to commemorate the 75th anniversary of the famous debate that took place there between Shapley and Curtis on whether some of the so called "nebulae" were stellar systems (i.e. other galaxies) beyond our Milky Way. I had the privilege of moderating this debate, perhaps because I was one of the few people who had explored both options (cf [7]). There was an agreement among all participants on the kind of new evidence that could settle the issue. Most valuable of all would be firm identification with objects detectable in other wavebands. The stumbling block here is the poor positional accuracy of most gamma-ray detectors. BATSE itself has error circles of 1 or 2 degrees for the brightest bursts, and more than 5 degrees for the fainter ones. However, there are some bursts whose locations have been pinned down with a precision of minutes of arc or better by triangulation experiments involving deep space probes; this technique utilises the rapid time structure, which, when recorded and timed by detectors separated by 10 light minutes or more, allows accurate positioning. It generally took several days to recover and correlate the required data, and to calculate positions by this technique: nothing was found by looking in any of the resulting error boxes. The controversies in the Shapley-Curtis debate were settled within a few years; our knowledge of extragalactic Physica Scripta T85
astronomy thereby made a forward leap, and astronomers moved on to address more detailed issues. The GRB distances were actually settled even more quickly and decisively: the crucial step was the detection of graduallyfading afterglows within some of the arc-minute-scale error boxes that the BeppoSax satellite was able to supply within a few hours of the burst. The first such detection occurred in February 1997 [8,9]; still more crucial was an event, GRB 970528, detected in May of that year, whose optical afterglow [10] displayed strong absorption features with z = 0.835, indicating that it probably lay in a galaxy with that redshift. Subsequently, further interest was aroused by the report of an afterglow for the burst GRB971214 at a redshift z = 3.4, whose energy output in y-rays would amount to 1053,5) erg if the emission were isotropic [11]. A recent tentative claim [12] that one of the strongest bursts may have a redshift of order 5, would signal a still higher energy release.
3. What is the trigger? The photon luminosity, for the few-second duration of a typical burst, is of course colossal: it exceeds by many thousands the most extreme output from any active galactic nucleus (thought to involve supermassive black holes), and is 14 orders of magnitude above the Eddington limit for a stellar-mass object. The total energy, however, is not out of line with some other phenomena encountered in astrophysics - indeed it is reminiscent of the energy released in the core of a supernova, the big difference being that the primary sudden event (with a timescale of seconds) is not smothered by a stellar envelope, as in a supernova, but manifests itself in hard radiation that escapes more promptly. Unless they are beamed into less than one percent of the solid angle, the triggers for GRBs are thousands of times rarer than supernovae. The most widely favoured and "conventional" possibility is coalescence of binary neutron stars (see, for example, [13]). Systems such as the famous binary pulsar will eventually coalesce, when gravitational radiation drives them together. When a neutron star (NS) binary coalesces, the rapidly-spinning merged system would be too massive (for most presumed equations of state) to form a single NS; on the other hand, the total angular momentum is probably too large to be swallowed immediately by a black hole. The expected outcome, after a few milliseconds, would therefore be a spinning black hole (BH), orbited by a torus of neutron-density matter. Other types of progenitor have been suggested - e.g. a NS-BH merger, where the neutron star is tidally disrupted before being swallowed by the hole; the merger of a white dwarf with a black hole [WD-BH] ; or a category labeled as hypernovae or collapsars, where the collapsing core is too massive to become a neutron star, but has too much angular momentum to collapse quietly into a black hole (as in a so called "failed supernova"). The simple point that I wish to stress, however, is that a BH plus debris torus is a common ingredient of all these models; moreover the overall energetics of these various progenitors differ by at most an order of magnitude, the spread reflecting the differing spin energy in the hole and the different masses left behind in an orbiting torus. (There has been some confusion on this point in recent literature, through failure to appreci© Physica Scripta 2000
Status of Models for Gamma Ray Bursts ate that the dominant energy from a NS-NS event comes after a black hole forms, rather than during the precursor stage that Narayan et al. [13] discussed.) How might such a system generate relativistic outflow or a release of electromagnetic energy? 4. Energy from a black hole and debris torus? Two large reservoirs of energy are in principle available: the binding energy of the orbiting debris, and the spin energy of the black hole. The first can provide up to 42% of the rest mass energy of the torus, for a maximally rotating black hole: the second can provide up to 29% (for a maximal spin rate) of the mass of the black hole itself. How can the energy be transformed into outflowing relativistic plasma after such a coalescence event? There seem to be two options. The first is that some of the energy released as thermal neutrinos is reconverted, via collisions outside the dense core, into electron-positron pairs or photons. The rate of this process depends on the square of the neutrino luminosity. The second option is that strong magnetic fields anchored in the dense matter convert the rotational energy of the system into a Poynting-dominated outflow, rather as in pulsars. Let us consider these two options in turn. (i) Neutrinos could give rise to a relativistic pair-dominate wind if they converted into pairs in a region of low baryon density (eg along the rotation axis, away from the equatorial plane of the torus). The vv -> e+e~ process can tap the thermal energy of the torus produced by viscous dissipation. For this mechanism to be efficient, the neutrinos must escape before being advected into the hole; on the other hand, the efficiency of conversion into pairs (which scales with the square of the neutrino density) is low if the neutrino production is too gradual. Typical estimates suggest a limit of <1051 erg [14-17], except perhaps in the "collapsar" or failed SN lb case where Popham et al. [17] estimate 1Q52.3 e r g S for 0 p t j m u m parameters. If the pair-dominated plasma were collimated into a solid angle Qj then of course the apparent "isotropized" energy would be larger by a factor {An/Qj), but unless Qj is < 10~2 — 10 -3 this may fail to satisfy the apparent isotropized energy of 10535 ergs implied by a redshift z = 3.4 for GRB 971214. (ii) An alternative way to tap the torus energy is via magnetic fields threading the torus [13,18-20]. Even before the BH forms, a NS-NS merging system might lead to winding up of the fields and dissipation in the last stages before the merger [21,22]. The above mechanisms tap the rotational energy available in the debris torus. However, a hole formed from a coalescing compact binary is guaranteed to be rapidly spinning, and, being more massive, could contain a larger reservoir of energy than the torus; this energy, extractable in principle through MHD coupling to the rotation of the hole by the Blandford and Znajek [23] (B-Z) effect, could be even larger than that contained in the orbiting debris [19,24]. Collectively, any such MHD outflows have been referred to as Poynting jets. Simple scaling from the familiar results of pulsar theory tell us that fields of order 1015 G, are needed needed to carry away the rotational or gravitational energy in the time scales of tens of seconds [25,26]. If the magneticfieldsdo not thread the BH, then a Poynting outflow can at most carry the © Physica Scripta 2000
269
gravitational binding energy of the torus. For a maximally rotating and for a non-rotating BH this is 0.42 and 0.06 of the torus rest mass, respectively. The torus mass in a NS-NS merger is Mt ~ 0.1 M 0 [16], and for an NS-BH or WD-BH merger it may be M, ~ 1M© [24,27]. The extractable energy could amount to several times 1053e(Mt/MG) ergs, where e is the efficiency in converting gravitational into MHD jet energy. Tori masses even higher than ~ 1M© may occur in scenarios involving massive supernovae. Conditions for the efficient escape of a high-r jet may, however, be less propitious if the "engine" is surrounded by an extensive envelope. If magnetic fields of comparable strength thread the BH, its rotational energy offers an extra (and even larger) source of energy that can in principle be extracted via the B-Z mechanism [19]. For a maximally rotating BH, this is 0.29Mt,hC2 ergs, multiplied, of course, by some efficiency factor. A near-maximally rotating black hole is guaranteed in a NS-NS merger. The central BH will have a mass of about 2.5MG; the NS-BH merger and hypernova models may not produce quite such rapidly-spinning holes, but the hole masses are larger, so the expected rotational energy should be comparable. Spinning holes can thus power a jet of up to ~ 1.5 x 1054 ergs. Even allowing for low total efficiency (say 30%), a system powered by the torus binding energy would only require a modest beaming of the y-rays by a factor {An/Qj) ~ 20, or no beaming if the jet is powered by the B-Z mechanism, to produce the equivalent of an isotropic energy of 1053-5 ergs. 5. The gamma-ray emission mechanism Well-known arguments connected with opacity, variability timescales and so forth require highly relativistic outflow. Best-guess numbers are Lorentz factors r in the range 102 to 103, allowing rapidly-variable emission to occur at radii in the range 1014 to 1016 cms. The entrained baryonic mass would need to be below 10_4M© to allow these high relativistic expansion speeds. Because the emitting region must be several powers of ten larger than the compact object that acts as atrigger®, there is a further physical requirement: the original energy outflowing in a magnetised wind would, after expansion, be transformed into bulk kinetic energy (with associated internal cooling). This energy cannot be efficiently radiated as gamma rays unless it is re-randomised. This requires relativistic shocks. Impact on an external medium would randomise half of the initial energy merely by reducing the expansion Lorentz factor by a factor of 2. Alternatively, there may be internal shocks within the outflow: for instance, if the Lorentz factor in an outflowing wind varied by a factor more than 2, then the shocks that developed when fast material overtakes slower material would be internally relativistic [28]. In an unsteady outflow, if T were to vary by a factor of 2 on a timescale St, internal shocks would develop at a distance r2cSt, and randomise most of the energy. For instance, if r ranged between 500 and 2000, on a timescale of 1 second, efficient dissipation would occur at 3 x 1016 cms. There is a general consensus that the longer complex bursts must involve internal shocks, though simple sharp pulses could arise from an external shock interaction (the Physica Scripta T85
270
Martin J. Rees
latter would in effect be the precursor of the afterglow). An external shock moving into a smooth medium would obviously give a burst with a simple time-profile. A blobby external medium could give features, but only if the covering factor of blobs is low, implying modest efficiency. Even if the bursts were caused by a completely standardised set of objects, their appearance would be likely to depend drastically on orientation relative to the line of sight. Along any given line of sight, the time-structure would be determined partly by the advance of jet material into the external medium, but probably even more by internal shocks within the jet, which themselves depend on the evolution of the torus, from its formation to its eventual swallowing or dispersal. The radiation processes for the gamma rays are probably no more than synchrotron radiation. This would imply the presence of magnetic fields where the shocks occur. If the outflow from the central trigger is Poynting-dominated, then a field of 1015 G at (say) 107 cm would imply a comoving field of 10 7 (r/100) _ 1 G out at 1013 cm - strong enough to ensure rapid cooling of shocked relativistic electrons. (Note, conversely, that even if magnetic fields were not important near the central trigger, they must be present, with about the same amount of flux that Poynting-dominated models require, at the location of the actual gamma-ray emission.) We are a long way from modelling what triggers gamma ray bursts. If we had a precise description of the dynamics, along with the baryon content, magnetic field, and Lorentz factor of the outflow, we could maybe predict the gross time-structure. But we could not predict the intensity or spectrum of the emitted radiation - still less answer key questions about the emission in other wavebands - without also having an adequate theory for particle acceleration in relativistic shocks. We need the answers from plasma physicists to the following poorly-understood questions: (i) Do ultra-relativistic shocks yield power laws? The answer probably depends on the ion/positron ratio, and on the relative orientation of the shock front and the magnetic field (e.g. [29]). (ii) In ion-electron plasmas, what fraction of the energy goes into the electrons? (Hi) Even if the shocked particles establish a power law, there must be a low-energy break in the spectrum at an energy that is in itself relativistic. But will this energy, for the electrons, be TmpC2, or (or even, if the positive charges are heavy ions like Fe, TrnpeC2)? (iv) Can ions be accelerated up to the theoretical maximum where the gyroradius becomes the scale of the system? If so, the burst events could be the origin of the highest energy cosmic rays, (v) Do magnetic fields get amplified in shocks? This is relevant to the magnetic field in the swept-up external matter outside, the contact discontinuity, and determines how sharp the external shock actually is.
6. Intrinsic time scales A question which has remained largely unanswered so far is what determines the characteristic duration of bursts, which can extend to tens, or even hundreds, of seconds. This is of course very long in comparison with the dynamical or orbital time scale for the "triggers", which is measured in milliseconds. While bursts lasting hundreds of seconds can easily be derived from a very short, impulsive energy input, this is generally unable to account for a large fraction Physica Scripta T85
of bursts which show complicated light curves. This hints at the desirability for a "central engine" lasting much longer than a typical dynamical time scale. Observationally [30] the short ( < 2 s) and long ( > 2 s) bursts appear to represent two distinct subclasses, and one early proposal to explain this was that accretion induced collapse (AIC) of a white dwarf (WD) into a NS plus debris might be a candidate for the long bursts, while NS-NS mergers could provide the short ones [31]. As indicated by Ruffert et al. [15], vv annihilation will generally tend to produce short bursts < 1 s in NS-NS systems, requiring collimation by 10 _1 — 10 - 2 , while Popham, Woosley and Fryer [27] argued that in collapsars and WD/He-BH systems longer vv bursts may be possible. An acceptable model requires that the surrounding torus should not completely drain into the hole, or be otherwise dispersed, on too short a time scale. There have been some discussions in the literature of possible 'runaway instabilities' in relativistic tori [32-34]: these are analogous to the runaway Roche lobe overflow predicted, under some conditions, in binary systems. These instabilities can be virulent in a torus where the specific angular momentum is uniform throughout, but are inhibited by a spread in angular momentum. In a torus that was massive and/ or thin enough to be self-gravitating, bar-mode gravitational instabilities could lead to further redistribution of angular momentum and/or to energy loss by gravitational radiation within only a few orbits. Whether a torus of given mass is dynamically unstable depends on its thickness and stratification, which in turn depends on internal viscous dissipation and neutrino cooling. The disruption of a neutron star (or any analogous process) is almost certain to lead to a situation where violent instabilities redistribute mass and angular momentum within a few dynamical time scales (i.e. in much less than a second). A key issue for gamma ray burst models is the nature of the surviving debris after these violent processes are over: what is the maximum mass of a remnant disc/torus which is immune to very violent instabilities, and which can therefore in principle survive for long enough to power the bursts? It is the mass of this residual torus - i.e what is left after violent instabilities on a dynamical timescale have done their work - that is the relevant M t in the above expressions (in Section 4) for the extractable energy of the torus. If the trigger is to liberate its energy over a period 10-100 sec via Poynting flux - either through a relativistic wind "spun off" the torus or via the B-Z mechanism - the required field is a few times 1015 G. A weaker field would extract inadequate power; on the other hand, if the large-scale field were even stronger, then the energy would be dumped too fast to account for the longer complex bursts. It is not obvious why the fields cannot become even higher. Note that the virial limit is Bs ~ 1017 G. Kluzniak and Ruderman [35] note that, starting with 1012 G, it only takes of order a second for simple winding to amplify the field to 1015 G; amplification in a newly-formed torus could well occur more rapidly, for instance via convective instabilities, as in a newly formed neutron star (cf [26,36]). Kluzniak and Ruderman suggest, however, that the amplification may be self-limiting because magnetic stresses would then be strong enough for flares to break out. © Physica Scripta 2000
Status of Models for Gamma Ray Bursts A magnetic field configuration capable of powering the bursts is likely to have a large scale structure. Flares and instabilities occurring on the characteristic (millisecond) dynamical time scale would cause substantial irregularity or intermittency in the overall outflow that would manifest itself in internal shocks. There is thus no problem in principle in accounting for sporadic large-amplitude variability, on all time scales down to a millisecond, even in the most long-lived bursts. Note also that it only takes a residual torus (or even a cold disk) of 10~3MQ to confine a field of 1015 G, which can extract energy from the black hole via the B-Z mechanism. Even if the evolution time scale for the bulk of the debris torus were no more than a second, enough may remain to catalyse the extraction of energy from the hole at rate adequate to power a long-lived burst.
271
comoving field strength, even out at 1013 cm, is of order 106 G. Trapped filaments of iron-rich thermal, with density up to 1019cm~3 and with kT of order a keV, could be confined by such fields. Such filaments must of course have a small volume-filling factor: otherwise they would load down the jet too much. However, in these strong fields the gyroradii would be so small that filaments could survive against thermal conduction and other diffusion processes even if their dimensions (transverse to the field) were less than 100 cm, such thin filaments can provide a large covering factor even whilefillinga tiny fraction of the volume. If they were moving relativistically outwards, they could contribute ultra-blueshift spectral features - for instance, K-edges of Fe could be shifted up to hundreds of keV. 8. Brief comments on the afterglows
7. Isotropic or beamed outflows?
The discovery of afterglows has not only has extended observations to longer time scales and other wavebands, making the identification of counterparts possible, but also provided confirmation for much of the earlier work on the fireball shock model of GRB, in which the y-ray emission arises at radii of 1013 - 1015 cm [28,40-44]. In particular, this model led to the prediction of the quantitative nature of the signatures of afterglows, in substantial agreement with subsequent observations [8,22,45-49]. Astrophysicists understand supernova remnants reasonably well, despite continuing uncertainty about the initiating explosion; likewise, we may hope to understand the afterglows of gamma ray bursts, despite the uncertainties about the "trigger" that I have already emphasised. The simplest hypothesis is that the afterglow is due to a relativistic expanding blast wave. The complex time-structure of some bursts suggests that the central trigger may continue for up to 100 seconds. However, at much later times all memory of the initial time-structure would be lost: essentially all that matters is how much energy and momentum has been injected, its distribution in angle, and the mass fractions in shells with different Lorentz factors. The simplest spherical afterglow model - where a relativistic blast wave decelerates as it runs into ambient matter, leading to a radiative output with a calculable spectrum, and a characteristic power law decay - has been remarkably successful at explaining the gross features of the GRB 970228, GRB970508 and other afterglows (e.g. [49]). The gamma-rays we receive come only from material whose motion is directed within one degree of our line of sight. They therefore provide no information about the ejecta in other directions: the outflow could be isotropic, or concentrated in a cone of any angle substantially larger Two further effects render the computational task of than one degree (provided that the line of sight lay inside simulating jets even more challenging. The first stems from the cone). At observer times of more than a week, the blast the likelihood that any entrained matter would be a mixture wave would however be decelerated to a moderate Lorentz of protons and neutrons (neutrons, being unconstrained factor, irrespective of the initial value. The beaming and by magneticfields,could also drift into a jet from the denser aberration effects are thereafter less extreme, so we observe walls at its boundary). If a streaming velocity builds up afterglow emission not just from material moving almost between ions and neutrons (i.e. if they have different Lorentz directly towards us, but from a wider range of angles. factors in the outflow) then interactions can lead to dissiThe afterglow is thus a probe for the geometry of the pation even in a steady jet where there are no shocks [37]. ejecta - at late stages, if the outflow is beamed, we expect A second possibility [38,39] is that entrained ions in a a spherically-symmetric assumption to be inadequate; the relativistic jet could become concentrated in dense filaments deviations from the predictions of such a model would then confined by the magnetic field. As already mentioned, the tell us about the ejection in directions away from our line
Computer simulations of compact object mergers and black hole formation can address the fate of the bulk of the matter, but there are some key questions that they cannot yet tackle. In particular, high resolution of the outer layers is needed because even a tiny mass fraction of baryons loading down the outflow severely limits the attainable Lorentz factor for instance a Poynting flux of 1053 ergs could not accelerate an outflow to r > 100 if it had to drag more than ~ 10~4 solar masses of baryons with it. Further 2D numerical simulations of the merger and collapse scenarios are under way largely using Newtonian dynamics, and the numerical difficulties are daunting. There may well be a broad spread of Lorentz factors in the outflow - close to the rotation axis r may be very high; at larger angles away from the axis, there may be an increasing degree of entrainment, with a corresponding decrease in T. Even if the outflow is not narrowly collimated, some beaming is expected because energy would be channeled preferentially along the rotation axis. Moreover, we would expect baryon contamination to be lowest near the axis, because angular momentum flings material away from the axis, and any gravitationally-bound material with low angular momentum falls into the hole. In hypernovae, the envelope is rotating only slowly and thus would not initially have a marked centrifugal funnel; even 1053 ergs would not suffice to blow out more than a narrow cone of the original envelope with a Lorentz factor or more than 100. So in these models the gamma rays would be restricted to a narrow beam, even though outflow with a more moderate Lorentz factor (relevant to the afterglow) could be spread over a wider range of angles. A wide variety of burst phenomenology could be attributable to a standard type of event being viewed from different orientations.
© Physica Scripta 2000
Physica Scripta T85
272
Martin J. Rees
of sight. It is quite possible, for instance, that there is relativistic outflow with lower r (heavier loading of baryons) in other directions (e.g. [49]); this slower matter could even carry most of the energy [24]. An argument for a broad beaming angle in the energy outflow is that, if the energy were channeled into a solid angle Qj then, [50], one expects a faster decay of r after it drops below QJ112. A simple calculation using the usual scaling laws leads then to a steepening of the flux power law in time. The lack of such an observed downturn in the light curve has been interpreted as further supporting the sphericity of the entire fireball. There are several important caveats, however. The first one is that the above argument assumes a simple, impulsive energy input (lasting < than the observed y-ray pulse duration), characterized by a single energy and bulk Lorentz factor value. Estimates for the time needed to reach the non-relativistic regime, or T < QJ / < few, could then be under a month [51,52], especially if an initial radiative regime with r oc r~3 prevails. (It is however unclear whether, even when electron radiative time scales are shorter then the expansion time, such a regime applies, as it would require strong electron-proton coupling [53]. Waxman, et al. [54] have also argued on observational grounds that the longer lasting r oc r~ 3/2 (adiabatic regime) is more appropriate.) Furthermore, even the simplest reasonable departures from this ideal model (e.g. a substantial amount of energy and momentum ejected behind the main shell with a lower Lorentz factor) would drastically extend the afterglow lifetime in the relativistic regime, by providing a late "energy refreshment" to the blast wave on time scales comparable to the afterglow time scale [55]. Anisotropy in the burst outflow and emission affects the light curve at the time when the inverse of the bulk Lorentz factor equals the opening angle of the outflow. If the critical Lorentz factor is less than 3 or so (i.e. the opening angle exceeds 20°) such a transition might be masked by the transition from ultrarelativistic to mildly relativistic flow, so quite generically it would difficult to limit the late-time afterglow opening angle in this way if it exceeds 20°. Since some afterglows are unbroken power laws for over 100 days (e.g. GRB 970228), if the energy input were indeed just a a simple impulsive shell the opening angle of the late-time afterglow at long wavelengths is probably greater than 1/3, i.e. £2 opt >0.4. However, even this still means that the energy estimates from the afterglow assuming isotropy could be 30 times too high. The beaming angle for the gamma ray emission could be far smaller, and is much harder to constrain directly. The ratio of Qy/Qx has been considered by Grindlay [66] using data from Ariel V and HEAO-A1/A2 surveys, who did not find evidence for a significant difference between the deduced gamma-ray and X-ray rates, and concluded that higher sensitivity surveys would be needed to provide significant constraints. More promising for the immediate future, the ratio Qy/Qopt can also be investigated observationally (see also [50]). The rate of GRB with peak fluxes above 1 p h c m ~ 2 s - 1 as determined by BATSE is about 300/yr, i.e. 0.01 /sq. deg/yr. According to Wijers et al. [56] this flux corresponds to a redshift of 3. If the gamma rays were much more narrowly beamed than the optical afterglow there should be many "homeless" afterglows, i.e. ones without a GRB preceding them. The transient sky at faint Physica Scripta T85
magnitudes is poorly known, but there are two major efforts under way to find supernovae down to about R — 23 [57,58]. These searches have by now covered a few tens of "square degree years" of exposure and would be sensitive to afterglows of the brightness levels thus far observed. It therefore appears that the afterglow rate is not more than a few times 0.1/sq. deg/yr. Since the magnitude limit of these searches allows detection of optical counterparts of GRB brighter than 1 ph c m - 2 s _ 1 it is fair to conclude that the ratio of homeless afterglows to GRB is unlikely to exceed ~ 20. It then follows that Qy > 0.05£2opt, which combined with our limit to Qopt yields Qy > 0.02. The true rate of events that give rise to GRB is therefore at most 600 times the observed GRB rate, and the opening angle of the ultrarelativistic, gamma-ray emitting material is no less than 5°. Combined with the most energetic bursts, this begins to pose a problem for the neutrino annihilation type of GRB energy source. Obviously, the above calculation is only sketchy and should be taken as an order of magnitude estimate at present. However, it should improve as more afterglows are detected and the modelling gets more precise. 9. Conclusions and prospects Simple blast-wave models seem able to accommodate the present data on afterglows. However we can at present only infer the energy per solid angle; there are reasons to suspect that the afterglow is not too narrowly beamed; on the other hand the constraints on the angle-integrated y-ray energy are not strong - these could be concentrated in a high-Lorentz-factor beam only a few degrees across. As regards the trigger, there are two key questions. First, does it involve a black hole orbited by a dense torus (which I've advocated as a "best buy")? Second, if so, can we decide between the various alternative ways of forming it: NS-NS, NS-BH or hypernova? The locations should help to settle the second question. This is because a hypernova would be expected to lie in a region of recent star formation; on the other hand, a neutron star binary could take hundreds of millions of years to spiral together, and could by then (epecially if given a kick velocity on formation) have moved many kiloparsecs from its point of origin [59]. There is also already tentative evidence that some detected afterglows arise in relatively dense gaseous environments - e.g. by evidence for dust in GRB970508 [60] and the absence of an optical afterglow and strong soft X-ray absorption in GRB 970828 [61,62]. On the other hand, fits to the observational data on GRB 970508 and GRB 971214 suggest external densities in the range of 0.04-0.4 cm - 1 , which would be more typical of a tenuous interstellar medium [65]. We must also remain aware of other possibilities. For instance, we may be wrong in supposing that the central object becomes dormant after the gamma-ray burst itself. It could be that the accretion-induced collapse of a white dwarf, or (for some equations of state) the merger of two neutron stars, could give rise to a rapidly-spinning pulsar, temporarily stabilised by rapid rotation. The afterglow could then, at least in part, be due to a pulsar's continuing power output (cf [25]). It could also be that mergers of unequal mass neutron stars, or neutron stars with other © Physica Scripta 2000
Status of Models for Gamma Ray Bursts compact companions, lead to the delayed formation of a black hole. Such events might also lead to repeating episodes of accretion and orbit separation, or to the eventual explosion of a neutron star which has dropped below the critical mass, all of which would provide a longer time scale, episodic energy output. And there could be more subclasses of classical GRB than just short ones and long ones. For instance, GRB with no high energy pulses (NHE) appear to have a different (but still isotropic) spatial distribution from those with high energy (HE) pulses [4]. Some caution is needed in interpreting this, since selection effects could lead to a bias against detecting HE emission in dim bursts [63]. Then, there is the apparent coincidence of GRB 980425 with the SN Ib/Ic 1998bw [64]. Much progress has been made in understanding how gamma-rays can arise in fireballs produced by brief events depositing a large amount of energy in a small volume, and in deriving the generic properties of the long wavelength afterglows that follow from this. There still remain a number of mysteries, especially concerning the identity of their progenitors, the nature of the triggering mechanism, the transport of the energy and the time scales involved. Gamma-ray bursts, even if we do not understand them, may still be useful as powerful beacons for probing the high redshift (z > 5) universe. Even if their total energy is reduced by beaming to a "modest" ~ 1052 — 10525 ergs in photons, they are the most extreme phenomena that we know about in high energy astrophysics. The modeling of the burst itself - the trigger, the formation of the ultrarelativistic outflow, and the radiation processes - is a formidable challenge to theorists and to computational techniques. It is, also, a formidable challenge for observers, in their quest for detecting minute details in extremely faint and distant sources. And if the class of models that we have advocated here turns out to be irrelevant, the explanation of gamma-ray bursts will surely turn out to be even more remarkable and fascinating. Acknowledgements I am especially grateful to Peter Meszaros and Ralph Wijers for extended collaboration, and to Josh Bloom and Stan Woosley for discussions. This research has been supported by the Royal Society.
References 1. Klebesadal, R. W., Strong, I. B. and Olsen, R. A., Astrophys. J. 182, L85 (1973). 2. Ruderman, M., Ann. N.Y. Acad. Sci. 262, 164 (1975). 3. Briggs, M. S. et al, Astrophys. J. 469, 40 (1997). 4. Pendleton, G. et al, Astrophys. J. 464, 606 (1996). 5. Lamb, D., PASP 107, 1152 (1995). 6. Paczynski, B., PASP, 107, 1167 (1995). 7. Podsiadlowski, P., Rees, M. J. and Ruderman, M., Mon. Notes R. Astron. Soc. 273, 755 (1995). 8. Costa, E. et al, Nature 387, 783 (1997) 9. van Paradijs, J. et al, Nature 386, 686 (1997). 10. Metzger, (1997). 11. Kulkarni, S. et al, Nature 393, 35 (1998). 12. Fruchter, A., submitted to Astrophys. J. Lett. (1999). 13. Narayan, R., Paczynski, B. and Piran, T. Astrophys. J. Lett. 395, L83 (1992). 14. Ruffert, M., Astron. Astrophys. 317, 793 (1997).
© Physica Scripta 2000
273
15. Ruffert, M., Janka, H.-T., Takahashi, K., Schaefer, G., Astron. Astrophys. 319, 122 (1997). 16. Ruffert, M. and Janka, H.-T. Aston. Astrophys. in press (1998). 17. Popham, R., Woosley, S. and Fryer, C , Astrophys. J. subm (astroph/98070 28) (1998). 18. Paczynski, B., Acta. Astron. 41, 257 (1991). 19. Meszaros, P. and Rees, M. J., Astrophys. J. 482, L29 (1997). 20. Katz, J. and Piran, T., Astrophys. J. 490, 772 (1997). 21. Meszaros, P. and Rees, M. J., Astrophys. J. 397, 570 (1992). 22. Vietri, M., Astrophys. J. 478, L9 (1997a). 23. Blandford, R. D. and Znajek, R. L., Mon. Notes R. Astron. Soc. 179, 433 (1997). 24. Paczynski, B., Astrophys. J. 494, L45 (1998). 25. Usov, V. V., Mon. Notes R. Astron. Soc. 267, 1035 (1994). 26. Thompson, C , Mon. Notes R. Astron. Soc. 270, 480 (1994). 27. Fryer, C. and Woosley, S., submitted to Astrophys. J. Lett. (1998) (astro-ph/9804167). 28. Rees, M. J. and Meszaros, P., Astrophys. J. 430, L93 (1994). 29. Gallant, Y. A., Hoshino, M., Langdon, A. B., Arons, J. and Max, C. E., Astrophys. J. 391, 73 (1992). 30. Kouveliotou, C. et al, Astrophys. J. 413, L101 (1993). 31. Katz, J. and Canel, L. M., Astrophys. J. 471, 915 (1996). 32. Nishida, S., Lanza, A., Eriguchi, Y. and Abramawicz, M. A., Mon. Notes R. Astron. Soc. 278, L41 (1996). 33. Abramowicz, M. A., Karas, V. and Lanza, A., Astron. Astrophys. (in press) (1998) (astro-ph/9712245). 34. Daigne, F. and Mochkovitch, R., Mon. Notes R. Astron. Soc. 285, LI 5 (1997). 35. Kluzniak, W. and Ruderman, M., Astrophys. J. 505, L113 (1998). 36. Thompson, C. and Duncan, R. C , Astrophys. J. 408, 194 (1998). 37. Derishev, E. V., Kocharovsky, V. V. and Kocharovsky, Vl.V preprint (1999). 38. Meszaros, P. and Rees, M. J., Astrophys. J. Lett. 502, L105 (1998). 39. Meszaros, P. and Rees, M. J., Mon. Notes R. Astron. Soc. 299, L10 (1998). 40. Rees, M. J. and Meszaros, P., Mon. Notes R. Astron. Soc. 258, 41P (1992). 41. Meszaros, P. and Rees, M. J., Astrophys. J. 405, 278 (1993). 42. Paczynski, B. and Xu, G., Astrophys. J. 427, 708 (1994). 43. Katz, J., Astrophys. J., 422, 248 (1994). 44. Sari, R. and Piran, T., Astrophys. J. 455, L143 (1995). 45. Meszaros, P. and Rees, M. J., Astrophys. J. 476, 232 (1997). 46. Tavani, M., Astrophys. J. 483, L87 (1997). 47. Waxman, E. Astrophys. J. 489, L33 (1997). 48. Reichart, D., Astrophys. J. 485, L57 (1997). 49. Wijers, R. A. M. J., Rees, M. J. and Meszaros, P., Mon. Notes R. Astron. Soc. 288, L51 (1997). 50. Rhoads, J. E., Astrophys. J. 487, LI (1997). 51. Vietri, M., Astrophys. J. 488, L105 (1997). 52. Huang, Y. F., Dai, Z. G. and Lu, T., Astron. Astrophys. in press (1998) (astro-ph/9807061) 53. Meszaros, P., Rees, M. J. and Wijers, R. A. M. J., Astrophys. J. 499, 301 (1998). 54. Waxman, E., Kulkarni, S. and Frail, D., Astrophys. J. 497, 288 (1998). 55. Meszaros, P. and Rees, M. J., Astrophys. J. Lett. 496, LI (1998). 56. Wijers, R. A. M. J., Bloom, J. S., Bagla, J. S. and Natarajan, P., Mon. Notes R. Astron. Soc. 294, L13 (1998). 57. Garnavich, P. et al, Astrophys. J. 493, L53 (1998). 58. Perlmutter, S. et al, Nature 391, 51 (1998). 59. Bloom, J., Sigurdsson, S. and Pols, O., Mon. Notes R. Astron. Soc. in press (1998) (astro-ph/9805222). 60. Reichart, D., Astrophys. J. 495, L99 (1998). 61. Groot, P. etal, in "Gamma-Ray Bursts," (Meegan, C , Preece, R. and Koshut, T., eds.), (AIP: New York 1997), p. 557. 62. Murakami, T. et al, in "Gamma-Ray Bursts," (ed Meegan, C. et al), (AIP, New York 1997). 63. Norris, J, private commun. (1998). 64. Galama, T. et al, Nature, 395, 670 (1998) (astro-ph/9806175). 65. Wijers, R. A. M. J. and Galama, T., Astrophys. J. subm (astroph/9805341) (1998). 66. Grindlay, J., preprint (1998)
Physica Scripta T85
NOBEL SYMPOSIA published by the Royal Swedish Academy in the periodical PHYSICA SCRIPTA NS 27
NS 46
NS 50
NS 59
NS 67
NS 73
NS 79
Super-Heavy Elements - Theoretical Predictions and Experimental Generation (1974) 3 p Edited by Sven Gosta Nilsson and Nils Robert Nilsson Physica Scripta vol 10A (1974)
186 pages. Price per copy GBP 25
Many-Body Theory of Atomic Systems (1979) Edited by Ingvar Lindgren and Stig Lundqvist Physica Scripta vol 21: 3-4 (1980)
370 pages. Price per copy GBP 35
Nuclei at very High Spin (1980) Edited by Georg Leander and Hans Ryde Physica Scripta vol 24: 1-2 (1981)
297 pages. Price per copy GBP 28
The Physics of Chaos and Related Problems (1984) Edited by Stig Lundqvist Physica Scripta vol T9 (1984)
224 pages. Price per copy GBP 25
Unification of Fundamental Interactions (1986) Edited by Lars Brink, Robert Marnelius, Jan S. Nilsson, Per Salonionson and Bo-Sture Skagerstam Physica Scripta vol T15 (1987)
212 pages. Price per copy GBP 25
Physics of Low-dimensional Systems (1988) Edited by Stig Lundqvist and Nils Robert Nilsson Physica Scripta vol T27 (1989)
168 pages. Price per copy GBP 25
The Birth and Early Evolution of Our Universe (1990) Edited by Jan S. Nilsson, Bengt Gustafsson and Bo-Sture Physica Scripta vol T36 (1991)
Skagerstam 304 pages. Price per copy GBP 31
Nobel Jubilee Symposium 1991
NS 85
NS 91
NS 99
Low Dimensional Properties of Solids (1991) Edited by Mats Jonson and Tord Claeson Physica Scripta vol T42 (1992)
220 pages. Price per copy GBP 25
Heavy-Ion Spectroscopy and QED Effects in Atomic Systems (1992) Edited by Ingvar Lindgren, Indrek Martinson and Reinhold Schuch Physica Scripta vol T46 (1993)
272 pages. Price per copy GBP 28
Trapped Charged Particles and Related Fundamental Physics (1994) Edited by Ingmar Bergstrom, Conny Carlberg and Reinhold Schuch Physica Scripta vol T59 (1995)
440 pages. Price per copy GBP 44
Heterostructures in Semiconductors (1996) Edited by Hermann G. Grimmeiss Physica Scripta vol T68 (1996)
122 pages. Price per copy GBP 25
NS 104 Modem Physics of Basic Quantum Concepts and Phenomena (1997) Edited by Erik B. Karlsson and Erkki Brandas Physica Scripta vol T76 (1998)
234 pages. Price per copy GBP 25
NS 109 Particle Physics and the Universe (1998) Edited by Lars Bergstrom, Per Carlson and Claes Fransson Physica Scripta vol T85 (2000)
274 pages, Price per copy GBP 82
All Nobel Symposia published in Physica Scripta can be ordered from Physica Scripta The Royal Swedish Academy of Sciences Box 50005 SE-104 05 STOCKHOLM Sweden Physica Scripta T85
© Physica Scripta 2000