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!
(p(-q,t)p(q / ,0))] =
^S(q,cof)S(-q9co-CDf)
.
(10.38)
Here we have taken into account that the dynamic structure factor S(q,<x>) is an even function of q. The approximation reduces S2(q,q';co) to a product of two density fluctuations of wavevector q and —q which propagate in the liquid without any interaction. S2(q,q';co) involves a frequency convolution over two dynamic structure factors. To the extent that this approximation is valid, the Raman-scattering intensity gives us no new information about the dynamics of the liquid, over and above what can be obtained from neutron-scattering studies (at least in principle). In evaluating Id(oj) based on the convolution approximation (10.38), one can use the quasiparticle ansatz (see Sections 1.1 and 2.1) S(q,co) = Z(q)S(co-coq)
(10.39)
in the region near the roton minimum. This gives 2
- 2wq)
(10.40)
103 Raman scattering from superfluid AHe
10
20 Energy shift (K)
251
30
Fig. 10.9. The Raman-scattering intensity vs. energy transfer for three different temperatures, at a pressure of 5 bar. The dashed curves are fits based on a Lorentzian approximation to S(Q,co) in (10.38) [Source: Ohbayashi et al, 1990].
and hence Id(co)~t}(QR)Z2(QR)p02((D)
(10.41)
Here the two-particle spectral density p°(Q= ^' co ) describing the excitation of two rotons with zero net momentum is given by the first term in (10.16). At temperatures above 1 K, a simple extension to include the roton finite width (due to scattering from thermally excited rotons) involves replacing the delta function in (10.39) by a Lorentzian. The convolution frequency integral (10.38) over two normalized Lorentzians
252
Two-particle spectrum in Bose-condensed fluids 10
Landau-
0.1 1.0
2.0 3.0 Temperature (K)
4.0
Fig. 10.10. The temperature dependence of the half-width of the peak in S(Q,co) at Q = 2 A"1 as determined from the fits such as shown in Fig. 10.9 [Source: Ohbayashi et al, 1990; Ohbayashi, 1991]. with a half-width T gives a normalized Lorentzian of half-width IT. The resulting generalization of (10.41) for Id(co) can then be used to extract information about the temperature dependence of the roton half-width T(T). This procedure has been used by Greytak and Yan (1971) for temperatures up to 1.8 K and more recently by Ohbayashi et al. (1990) to Tx and above. As shown in Fig. 10.9, it appears to be capable of giving good fits to the Raman data as a function of the temperature. The results so obtained for the temperature dependence of the roton energy and width (see Figs. 10.10 and 10.11) are in good agreement with those obtained directly from neutron-scattering data on S(Q,co), such as given by Stirling and Glyde (1990). However two comments should
4
10.3 Raman scattering from superfluid 10
•
s
1
1
•••
253
He
1
I : error
\
•• • I
a
1 1.0
1
1
2.0 Temperature (K)
1
1
3.0
Fig. 10.11. The temperature dependence of the peak position in S(Q,co) at Q = 2 A" 1 , as determined from fits such as shown in Fig. 10.9 [Source: Ohbayashi et aU 1990; Ohbayashi, 1991].
be made. First of all, results such as those shown in Figs. 10.10 and 10.11 are based on (10.38) and thus include the effect of the roton-roton interaction on the width of the resonance in S(Q,a>) but not on the shape of the two-particle spectral density p2(Q = 0,co) (for the case 2 ^ 0 , see Figs. 10.1 and 10.2). In many-body language, propagator renormalization effects are included but not vertex corrections (the effect of interactions between the excitations, which may lead to bound states). The second comment concerns the interpretation of the results in Figs. 10.10 and 10.11 near and above Tx in terms of the "roton" width and energy. In Section 7.2, we argued that a careful analysis of the S(Q,Q)) line shape at Q ~ 2 A" 1 in the region near Tx (see Fig. 7.11) was consistent with the idea that the roton peak intensity was vanishing, being replaced by a broad particle-hole spectrum (associated with thermally excited rotons) which characterizes the normal phase. If one models the resulting changes in the S(Q9co) line shape in terms of a single Lorentzian with a rapidly increasing width near Tx, one will naturally be led to results of the kind shown in Figs. 10.10 and 10.11. One cannot, however, interpret such fits as giving information about rotons near and above Tx. As we discussed in connection with neutron-scattering data in Section 7.2, a meaningful extraction of information about the
254
Two-particle spectrum in Bose-condensed fluids
quasiparticle spectrum at temperatures above about 1.7 K requires a rather sophisticated theoretical input. Even at low temperatures, the simple convolution approximation (10.38) has several deficiencies: (a) It ignores what one might call the excluded-volume effect due to the 4 He atom hard core. In the coordinate space form given in (10.36), it is clear that there can be no contribution from the regions |r3 — r4|, 1*1 —r2| < a, where a is the hard-core diameter. The simple decoupling (10.38) does not handle this constraint properly (for further discussion and references, see Halley, 1989). (b) As we mentioned above, by approximating (10.37) in terms of two non-interacting rotons, all the effects discussed in Section 10.1 have been ignored. In particular, we cannot discuss how h(co) will show the presence of a two-roton bound state which may arise when there is an attractive roton-roton interaction g4. (c) In approximations such as (10.38), the key step lies in treating the density fluctuations, rather than the field operators, as the fundamental variables. All available calculations of S2(q,qr;o>) which include the effect of two-roton bound states are based on identifying the density fluctuations as the elementary excitations. That is, something like p(q) = y/Z(q) [a+ + a_q]
(10.42)
is used, where a+ is the creation operator of an excitation (see Section 9.1). In this type of approach, the correlation function S2(q9q';co) involving four density operators is effectively reduced to a two-particle Green's function, such as G2(Q = 0, co) defined in (10.1). To this extent, the description of two-roton bound states can be taken over from the analysis given in Section 10.1. This approach of relating S2(q,q';&>) directly to G2(Q = 0,co) involves the same sort of approximation as taking the density-response function Xnn to be directly proportional to the single-particle Green's function, as in (10.31). Halley and Korth (1991) have extended the analysis based on (10.42) to include backflow by calculating S2(q,qr;co) starting from (10.37) using the correlated-basisfunction approach of Manousakis and Pandharipande (1986) discussed in Section 9.1. What one would like to see is a calculation of S2(q,q''9co) based on treating it as a true four-particle Green's function (involving eight quantum field operators). In earlier chapters, we saw the importance of keeping the
103 Raman scattering from superfluid 4He
10
255
30 Energy shift (K)
Fig. 10.12. Raman intensity vs. energy shift &>, measured at 0.65 K and SVP. The instrumental FWHM is 0.75 K. Weak structure above the two-roton peak at 2A is clearly evident in the expanded part of the high-energy data [Source: Ohbayashi, 1991].
distinction clear between inn and the more fundamental single-particle Green's functions Gap. The analogous investigation of S2(q,qr;co) has not been carried out in the literature to date. Writing (10.37) in the form (10.43)
one sees that reducing this expresion to products of pairs of single-particle operators (i.e., single-particle Green's functions) results in additional terms which are not included in (10.38) even in a normal Bose fluid. Moreover, when there is a Bose condensate present, we have a whole
256
Two-particle spectrum in Bose-condensed fluids
new class of pairings involving the off-diagonal averages, (a£(t)a*k) and (ak(t)a-k). We thus conclude that S2(q,q';co) describes dynamical correlations in superfluid 4 He which are not expressed simply in terms of density correlation functions, as in (10.38). This is a complication, but it also suggests that the Raman-scattering intensity may yield unique information not available from neutron-scattering experiments. Further studies of S2(q,q';c0) based on the field-theoretic analysis appropriate to a Bosecondensed fluid are clearly needed, as are studies of the two-particle Green's functions discussed in Section 10.1. As we have reviewed in Section 10.1, there is good evidence from the low-temperature Raman data that, at low pressure, there is a two-roton bound state. This was first observed by Greytak and Yan (1969) and has been confirmed by increasingly high-precision studies. Combining the Raman data of Murray et al. (1975) with the roton energy given by neutron scattering, the two-roton binding energy is estimated to be 0.27 + 0.04 K (Woods et al., 1977). More recent work is summarized by Ohbayashi (1989, 1991), who has also found evidence for additional fine-scale structure at higher energies (see Fig. 10.12). The latter may be due to higher-order resonances associated with maxon-maxon states as well as to bound states involving combinations of three or more maxons and rotons (for further discussion, see Iwamoto, 1989).
11 Relation between excitations in liquid and solid 4 He
In the early 1970's, attention was drawn to the remarkable similarity between the excitation spectra exhibited by S(Q,co) in solid 4 He and superfluid 4 He at low temperatures (Werthamer, 1972; Horner, 1972a; Glyde, 1974), as shown very dramatically in the theoretical results of Figs. 11.1 and 11.2. While various suggestions have been made as to the origin of this similarity, it remains an unresolved and intriguing problem. In this brief chapter, we compare the theoretical description of excitations in a quantum solid with those of a Bose-condensed liquid. While we review the key ideas, we assume that the reader has some familiarity with an introductory account of quantum crystals. (The modern theory of excitations in quantum crystals was essentially completed in the early 1970's. For background and a more detailed discussion of solid 4 He than we give in this chapter, we recommend the review by Glyde, 1976.) In both condensed phases, it is important to distinguish clearly between the elementary excitations and the density fluctuations. We argue that the phonons in solid 4 He are the natural analogue of the single-particle excitations in liquid 4 He. In Section 11.1, defining the phonons as the poles of the displacement correlation function, we briefly review theories which start with the self-consistent harmonic (SCH) approximation or something similar. In Section 11.2, we discuss the relation between the displacement-displacement and the density-density correlation functions in solid 4 He. This relation is based on the well known Green's function analysis of anharmonic crystals initiated by Ambegaokar, Conway and Baym (1965). Finally, in Section 11.3, we compare the expressions for S(Q,(o) in solid and superfluid 4 He. We do not discuss the interesting possibility of finding a Bose condensate in a quantum solid. For references, see Meisel (1992). 257
258
Relation between excitations in liquid and solid 4He
co (meV)
Fig. 11.1. Theoretical results for S(Q,co) as a function of Q (along the (111) direction) and co, in bcc solid 4 He. The peak intensity is seen to follow a "phonon-maxon-roton" type dispersion curve (dark line), with the shifting of spectral weight to free-particle-like behaviour (dashed line) at high Q [Source: Horner, 1974a]. 11.1 Phonons as poles of the displacement correlation function
The usual Hamiltonian describing an anharmonic crystal is obtained by expanding the interatomic potential energy in powers of the atomic displacements from the (Bravais) equilibrium sites. The degrees of freedom are described by the displacement field and thus the elementary excitations (phonons) correspond to the poles of the displacementdisplacement correlation function (usually called the phonon propagator). The lowest-order harmonic approximation consists of neglecting the cubic and higher-order anharmonic force constants. Expressing the atomic displacement of the / -th atom in terms of the usual phonon creation and annihilation operators, we have
2mNa>°qX
(11.1)
where AqX = aq)i+a*x and e x is the polarization vector of the X phonon branch. The Fourier transform of the (retarded) one-phonon Green's
11.1 Phonons as poles of the displacement correlation function
259
10
2 (A"1) Fig. 11.2. Theoretical dispersion curves (co vs. Q) for one-phonon and multiphonon structure in S(Q,co) in bcc solid 4 He (in three different directions). The heavy lines and shaded region give the mean energy and width, respectively, of the multiphonon scattering. The dash-dot line gives the free-atom recoil frequency. The heavy dots represent the peaks in the S(Q,co) data (Kitchens et al, 1972), with a width given by the dashed lines [Source: Horner, 1974b].
260
Relation between excitations in liquid and solid 4He
function Dfat) = -i9(t)([Aq,(t)J^(0)])
(11.2)
is given in a (self-consistent) harmonic approximation by Dx(q9co)=(
—0
^-0-) .
(11.3a)
The corresponding one-phonon spectral density is Ai(q,co) = -2 Im D^(q,co + «y)
= 2TT [5(o> - G)JA) - S(co + co,0,)] .
(11.36)
The effect of including higher-order anharmonic terms in the Hamiltonian can be treated using the usual diagrammatic perturbation methods of many-body theory (see, for example, Kwok, 1967; Horner, 1967). In such discussions, the phonon Green's function corresponding to some sort of harmonic approximation is the "building block", analogous to the single-particle Green's function for a non-interacting gas in discussions of quantum liquids. The effect of the anharmonic interactions (cubic, quartic, ...) gives rise to a phonon self-energy which leads to frequency shifts and damping of the original harmonic phonons. Extensive calculations of this kind are available in the literature (see, for example, Maradudin and Fein, 1962; Cowley, 1968; Glyde, 1971). In trying to formulate an analogous theory for solid 4He, one is immediately faced with the problem that small displacements about the equilibrium sites are unstable and thus the standard harmonic approximation is not a good starting point. Since the associated phonon frequencies are imaginary, self-consistent phonon (SCP) theories have been developed, based on renormalized force constants. The latter incorporate the effects of large zero-point fluctuations and lead to well defined phonon frequencies. These methods are based either on field-theoretic resummations of higher-order effects or on variational approximations to the many-body wavefunctions. In the latter approach, the physics is clear, since a wavefunction which is Gaussian in the atomic displacements is equivalent to some effective harmonic Hamiltonian. (For a lucid summary of how this procedure is carried out, we refer to Section II of Gillis, Werthamer and Koehler, 1968.) The problem is that if the strong short-range correlations (SRC) are included by an additional Jastrow-Feenberg-type function, we do not necessarily find any simple equivalence to an SC harmonic Hamiltonian. In particular, many theories which include SRC no longer guarantee that the excitations still
11.1 Phonons as poles of the displacement correlation function
261
correspond to poles of the displacement correlation function. More or less satisfactory self-consistent phonon theories with short-range correlations were developed in the early 1970's. A Green's function approach is given by Horner (1974c). One such theory of phonon excitations can be elegantly formulated in variational terms using the general approach of correlated basis functions (Feenberg, 1969). As we review in Section 9.1, this kind of approach has been extensively developed as a description of liquid 4 He at T = 0. Thus it is useful to discuss the analogous approach for solid 4 He. Here we briefly sketch the work of Koehler and Werthamer (KW, 1971), who have applied the CBF philosophy to the determination of the excited states of solid Helium in a consistent manner. Earlier work by Koehler (1967, 1968) treated the ground state in terms of a variational many-body wavefunction |Oo) which was a product of a Jastrow function describing the short-range correlations (SRC) and a Gaussian function |
(H.4)
The optimized energies of the KW one-phonon states obtained in this manner can be proven to be identical to the SC frequencies of a harmonic crystal whose force constants are defined only through the coefficients of the Gaussian part |OG) of the ground-state wavefunction. Thus the lowlying excited states defined by KW allow for a simple but still consistent interpretation as phonons even when SRC are included (see also Horner, 1971). A key feature is that the minimization of the ground-state energy with respect to the Gaussian coefficients in \0>G) automatically leads to the diagonalization of the Hamiltonian in the one-phonon-state subspace. Several authors have emphasized that the poles of the displacement correlation function are not necessarily exhausted by the phonon fre-
262
Relation between excitations in liquid and solid 4He
quencies given by SCP theory. A useful comparison can be made between SCH theories of quantum crystals and Hartree-Fock theories of quantum liquids. The Hartree-Fock approximation gives the best renormalized single-particle states but completely ignores collective modes. The SCH approximation, in contrast, gives the best collective modes but ignores single-particle excitations. Most SCP theories implicitly assume from the beginning that the elementary excitations can be classified in terms of some equivalent harmonic lattice. This assumption is built into the form used for the excited-state wavefunctions. In this connection, we note that the precise relation between collective vs. single-particle theories of excitations in solid 4 He has never really been resolved (Fredkin and Werthamer, 1965; Gillis and Werthamer, 1968; Werthamer, 1969). Some aspects of this relation are touched on by the numerical work of Horner (1972b), whose results for the spectral density Ax(Q,co) show the clear transition from phonon-like excitations at low Q and co to more single-particle-like excitations at high Q and co. The same transition is shown by the results for S(Q, co) given in Figs. 11.1 and 11.2. One should also recall that when anharmonic corrections are included in such theories, the phonon propagator spectral density is not usually strongly peaked as in (11.3b). As Q increases, the spectral density spreads out over a larger frequency region and develops a high-energy tail, as shown by the example in Fig. 11.3. Such self-energy effects inevitably arise when one corrects the SCH spectral density to include anharmonic interactions. Thus, the peak positions in Fig. 11.3 are quite different from the line centres (denoted by the arrows) defined as the normalized first frequency moment of Ax(Q,co). One is faced with how to define an appropriate mean or average phonon energy for given Q, X (Horner, 1972b). Indeed, one is led to question the usefulness of any such simplified description in calculating the thermodynamic properties of a quantum (or highly anharmonic) crystal in place of the full displacement-field spectral density A^(Q,co). Finally, we note that for small Q , the phonons in a solid can be also usefully classified as either hydrodynamic or collisionless, following the terminology of Section 6.2. Typically phonons studied by ultrasonics or Brillouin light scattering are in the low-energy, hydrodynamic domain and are often referred to as first sound. Solid 4 He also exhibits second sound, which may be viewed as an oscillation in the local number density of phonons. As Kwok and Martin (1966) discuss, both first sound and second sound appear as poles of the displacement correlation function of an anharmonic crystal in the long-wavelength limit (see also Kwok,
11.2 Phonons vs. density fluctuations in solid 4He
0.2
0.4 0.6 Frequency (THz)
0.8
263
1.2
Fig. 11.3. The one-phonon spectral density A(Q,co) as a function of Q and co, for a longitudinal mode in bcc solid 4He along the (1,0,0) direction. The phonon self-energy due to a single "bubble" has been computed self-consistently. In comparing these results with Fig. 3 of Horner (1972b), see ref. 20 in McMahan and Beck (1973) [Source: McMahan and Beck, 1973]. 1967). In contrast, the low-Q phonons in solids which are excited by neutron scattering are in the high-energy collisionless region. For this reason, such excitations are often referred to as zero sound phonons, in analogy to the collective modes in quantum liquids.
11.2 Phonons vs. density fluctuations in solid 4He Since the middle 1960's, it has been realized that, in general, the dynamic structure factor S(Q,co) of anharmonic crystals can be a complicated function of the underlying phonon excitations. The key simplifying feature of a crystal is that the dynamics involve displacements of the atoms with respect to a Bravais lattice, r/ = R/ + u/. Substituting (2.5) into the structure factor (2.7) gives (11.5)
Relation between excitations in liquid and solid 4He
264
Expanding (11.5) in powers in the atomic displacements, we obtain S(Q,co) = SBmgg(Q)S{co) + Si(Q,o>) + Sint(Q,co) + Smp(Q,co) ,
(11.6)
where
^2YiQiRR'f)Q'(ul(t)u,(0))-Q
(11.7)
is the "one-phonon" contribution involving the displacement-displacement correlation function DiV{t) = (M,(*)K/'(0)). In (11.7), d2(Q) = exp[-2W(Q)] is the Debye-Waller factor. The additional terms in (11.6) describe twophonon (and higher) contributions plus interference terms. The higherorder phonon contributions which arise in (11.6) complicate the situation considerably in a highly anharmonic quantum crystal like solid 4 He. These give rise to both a broad multiphonon continuum in S(Q,(o), and interference terms which contribute within the one-phonon region and considerably modify the contribution of Si (Q, co) to the total dynamic structure factor at large Q. It is now recognized that in solid 4 He, there is an important difference between the density fluctuation spectrum (peak positions in S(Q, co)) and the underlying phonon excitations (peaks in Ax{Q9o)) or Sx(Q9(o)). The many-body theory of the density-response function XnniQ, co) in solid 4 He can be formulated in terms of Dx(Q, co) following the analysis of Ambegaokar, Conway and Baym (ACB, 1965). The final result can be written schematically in the form Xm(Q,
+ xmp(Q,co) .
(11.8)
The density fluctuation spectrum is thus seen to include one-phonon as well as multiphonon contributions (involving two or more phonons). In addition, there are interference contributions described by the vertex functions Rx(Q,oo) in (11.8), by which higher-order phonon processes modify the contribution of the single-phonon scattering. Thus the S\ and Sint terms in (11.6) can be combined into a single contribution Sp(q,co) which may be called the physical "one-phonon" contribution to S(Q,co). This consists of all contributions to S[nt which contain a single-phonon propagator as an intermediate state. The effect of the interference terms on the one-phonon peak was first clearly exhibited in alkali halide crystals (Cowley and Woods, 1969). To the extent that it can be limited to Si(Q,co) in (11.6), the density fluctuation spectrum is identical to that of the displacement fluctuations (given by SCP theory, as discussed in Section 11.1). However, in a
11.2 Phonons vs. density fluctuations in solid 4He 1
1
i
1 o
1
O (1.5,0,0)
—
150
§f
'I ^\
100
\
O 50 so 1—
A 0
265
\o \
— 0
/
i
^r
i
0
1.0
2.0 Energy (meV)
°°^£ 3.0
4.0
Fig. 11.4. Scaled neutron-scattering intensity vs. co from longitudinal modes L[100] at two values of Q differing by a reciprocal lattice vector of the bcc solid. The instrumental resolution is shown by the bar. See caption of Fig. 11.5 [Source: Osgood, Kitchens, Shirane and Minkiewicz, 1972].
quantum crystal like solid 4 He, one cannot restrict oneself to SI(Q,G>). This is shown dramatically in Fig. 11.4 by experimental data for S(Q,a>), suitably normalized, at two values of Q, Q' which differ only by a reciprocal lattice vector T of the Bravais bcc lattice. Since we normalize the results by dividing by Q2d2(Q), 5i(Q,co) should be identical to Si(Q',co), where Q ; = Q + T. This equivalence follows immediately from the translational invariance of the displacement correlation function i.e., Dx(Q9co) = DI(Q + T,CO). In contrast, the data in Fig. 11.4 show that S(Q,co)/Q2d2(Q) is quite different at Q and Q ; . Because of the features outlined above, it is difficult to extract the phonon frequencies (or, more generally, the phonon spectral density) directly from the measured values of S(Q, co) at larger values of Q. Horner (1972a, 1974a), Glyde and Goldman (1976) and others have carried out detailed calculations of S(Q,a>) using (11.8), limiting themselves to one- and two-phonon processes plus their coupling. As shown in
266
Relation between excitations in liquid and solid 4He
Fig. 11.5, with enough care, these calculations can indeed reproduce the observed structure in S(Q, a>) starting from an appropriate phonon propagator D^(Q,(o). This figure also shows the different components making up the dynamic structure factor <S(Q,co). Clearly, however, without such microscopic calculations, one cannot hope to make any detailed comparison between the phonon spectral density Ax(Q,co) and S(Q,co) in solid 4 He. The spectra associated with the displacement and density correlation functions are simply quite different at larger wavevectors. ACB have derived some useful /-sum rules for the different contributions to S(Q,co). These are da> a>Su(Q,a>) =
( Q
'g^
d\Q)
= f dm o)SPA(Q,(o) ,
(11.9)
J/ —oo —o
where eqx is the polarization vector of the phonon branch being studied. These exact results show that both Si(Q,co) and SP(Q,CD) take up only a fraction d2(Q) of the total first frequency moment. This means that the effect of the interference vertex functions Rx(Q,co) in (11.8) is only to rearrange the spectral weight associated with the single-phonon propagator, not to change its total contribution to the first frequency moment /-sum rule. 11.3 Relation between S(Q,co) in superfluid and solid 4 He Clearly one can draw an analogy between the ACB result in (11.8) for solid 4 He with (3.47) and the dielectric formalism expression (5.24) in the case of superfluid 4 He. This similarity has been discussed by Wong and Gould (1974) as well as Glyde (1984). The single-particle Green's function Gap(Q,co) in Bose-condensed liquid 4 He plays the same role as the one-phonon (or displacement-field) correlation function Dx(Q,co) in solid 4 He. Wong (1979) has pointed out that the Debye-Waller factor d2(Q) in (11.8) plays a role analogous to the Bose-condensate order parameter \{ip)\2 = no which enters into the Bose vertex function Aa in (5.24). Despite the formal similarity of the structure of (11.8) and (5.24), however, there is a fundamental difference between an anharmonic crystal and a Bose-condensed fluid since, in the former, the phonon self-energy does not have a part analogous to the term in (5.75). In crystals, one has a relation similar to (5.24) but no equivalent of the inverse relation
11.3 Relation between S(Q,co) in superfluid and solid 4He 1
i
1
267
!
7
-
6
Q = (1.5,O,O)7t/tf S(Q,a»
5 S{ (Q, co) S2 (Q, co) 4
n\\
3
2
-
-
s
2
1
0
1
^
}
'
3 4
3
2
/v-
-
j
/
Q = (0.5, 0, 0) n/a
1
\
1
0 i
0
1
^ '
i
i
i
2
3
4
5
co (meV)
Fig. 11.5. Calculated values of S(Q, co)/Q2d2(Q) vs. co for two equivalent points in Q-space (see also Fig. 11.4). S(Q,co) is decomposed into the bare one-phonon component Si, the two-phonon component S2 of the multiparticle contribution, and the interference component Smt. Note that the scaled Si component is identical for the two values of Q [Source: Horner, 1972a].
268
Relation between excitations in liquid and solid 4He
(5.31) which would give the phonon displacement correlation function Dx(Q9co) in terms of the density-response function xnn(Q,co). As a result, the poles of the phonon displacement correlation function and the density correlation function are not coupled as in a Bose-condensed fluid. As we have discussed at length in Chapters 5 and 7, the strongly hybridized nature of the single-particle and zero sound modes means that the two contributions in (5.24) or (7.6) can strongly interfere with each other. At low Q it is better to start from the expression in (5.76). In this case, the dominant low-energy pole of both G^ and Xnn is a zero sound mode in superfluid 4 He. In the high-g region, in contrast, where the dominant maxon-roton pole in Xnn is interpreted to be an SP excitation originating as a pole of Gajg, (5.24) is more appropriate. It is in this SP region that the analogy to (11.8) is relevant. The formal similarity between (3.47) or (5.24) and (11.8) in the high-g, high-co region is especially intriguing in view of the very similar density fluctuation spectrum exhibited by both solid and superfluid 4 He (compare Fig. 11.2 with Fig. 1.6). Originally this similarity to superfluid 4 He was puzzling (Werthamer, 1972; Horner, 1972a), since the maxon-roton spectrum was thought to be specifically related to the "superfluid" nature of liquid 4 He. Our present interpretation (Section 7.2) of the maxon-roton as essentially a renormalized atomic-like excitation associated with the normal phase allows for a natural explanation of the essential similarity of S(Q, co) at large Q and co in the solid and superfluid phases of 4 He. This common origin presumably lies in the vibrational dynamics of a 4 He atom in a small cage formed by- its nearest neighbours, as Horner (1972a, 1974a) has suggested. Further theoretical studies would be highly desirable. As we discussed in Section 11.1, the variational correlated-basisfunction (CBF) approach has been used to treat the excitations in both solid and superfluid 4 He at low temperatures. However we call attention to a very basic difference between these discussions. In the case of solid 4 He, the CBF approach is used (Koehler and Werthamer, 1971, 1972) to evaluate the displacement correlation function D,t(Q,a>), i.e. the displacement-field excitations. The connection between these phonon excitations and the density fluctuation spectrum described by S(Q,a>) is then given by a separate ACB-type analysis (as in Section 11.2). In superfluid 4 He, in contrast, the CBF method is used to compute S(Q,co) directly, bypassing completely any discussion of the single-particle dynamics described by Gap(Q9co). In summarizing this whole chapter, it would seem that the word "phonon" should never be used in quantum solids without an appropriate
11.3 Relation between S(Q, co) in superfluid and solid 4He
269
adjective stating what "kind" of phonon one is dealing with. Much of the earlier literature is confusing because the distinction between various kinds of phonons was either not realized or insufficiently emphasized. The same comment is equally valid for "phonons" in quantum liquids, as we have seen in Chapters 5-7. More generally, both in Bose-condensed liquids and in quantum (or highly anharmonic) crystals, there is a subtle relation between the elementary excitations and the density fluctuations (the latter being measured in neutron scattering).
12 The new picture: some unsolved problems
In this book, we have developed the theory of the excitation spectrum of superfluid 4 He in which the Bose condensate plays the central role. In Chapter 5, we showed how a Bose broken symmetry inevitably leads to a mixing of the single-particle and density fluctuations. Combining the general results of the dielectric formalism with recent high-resolution neutron-scattering data over a wide range of wavevectors, energies and temperatures, we were led in Chapter 7 to a new interpretation of the well known phonon-maxon-roton dispersion curve. In Section 12.1, we briefly recapitulate this new scenario and discuss how it developed from preceding theoretical work. We also review earlier studies which had independently suggested that rotons were in fact atomic-like singleparticle excitations, quite different from the long-wavelength phonons. In addition, we address the question of what Feynman's work says about the nature of rotons. The most important topic which has not been covered in this book is superfluid 3 He- 4 He mixtures. The appropriate dielectric formulation has been developed by Talbot and Griffin (1984c), as we briefly summarize in Section 12.2. Much work remains to be done in using these formal results in a detailed analysis of experimental 5(Q, co) data, even at the level of Section 7.2 in the case of pure 4 He. Finally, in Section 12.3, we list some specific topics where further theoretical and experimental work would be useful. This list, which brings together suggestions scattered throughout the book, also acts as a convenient summary of our major themes. 12.1 Comments on the development of the new picture We recall that in the presence of a Bose broken symmetry, the singleparticle excitations, the particle-hole excitations and the two-particle 270
12.1 Comments on the development of the new picture Xnm (Q> « )
Particle - hole or density excitations
Av Single-particle excitations
211
Neutron scattering
X G2(Q, co) Two-particle or pair excitations
Raman scattering
Fig. 12.1. A block diagram showing how the condensate couples the various kinds of excitations in a Bose-condensed fluid.
excitations are all coupled into each other (Fig. 12.1). This sharing of excitations is the key dynamical consequence of a Bose condensate, as was first emphasized by Gavoret and Nozieres (1964) and Hohenberg and Martin (1965). The formal relations between the single-particle Green's function Ga^(Q,co), the particle-hole density-response function Xnn(Q,co) and the two-particle Green's functions G2(Q, co) have been discussed at length in Chapters 5 and 10. The precise implications of these coupled equations of motion, however, require the solution of very complex equations for self-energies, vertex functions, polarization parts etc. This has only been successfully carried out in the limit of small Q and co at T = 0 (as reviewed in Section 6.3). Even in the absence of rigorous calculations, however, the general structure of the coupled correlation functions made evident using the dielectric formalism (Chapter 5) does allow one to develop various scenarios concerning the excitations in a Bose fluid like superfluid 4 He. A quote from p. 23 of Bogoliubov (1947) applies equally well here: "All we can require from a molecular theory of superfluidity, at least at the first stage of investigation, is to be able to account for the qualitative picture of this phemonenon being based on a certain simplified scheme." In the interpretation of phonons, maxons and rotons developed in Chapter 7, the phonon is a zero sound (ZS) mode while the maxonroton is a single-particle (SP) excitation which has weight in S(Q,co) only because of the Bose broken symmetry. Physically, these two modes are interpreted as quite different excitation branches which are hybridized through the condensate to produce the observed phonon-maxon-roton dispersion curve (Glyde and Griffin, 1990). In this section, we review how
272
The new picture: some unsolved problems
this picture grew out of earlier studies. We also comment on what this picture assumes and implies about the nature of phonons and rotons. The standard microscopic view of excitations in superfluid 4 He was formulated in the classic paper by Miller, Pines and Nozieres (MPN, 1962) as well as by Nozieres and Pines in their 1964 monograph (see Section 7.3 of NP, 1964, 1990). This early work is based on Hugenholtz and Pines (1959), who first pointed out that in the presence of a Bose condensate, S(Q, co) would have a term directly proportional to the Beliaev single-particle Green's function Ga^(Q,co). Working to all orders in perturbation theory, Gavoret and Nozieres (1964) proved that, at small Q and co and at T = 0, both of these correlation functions exhibit the same phonon pole. (This behaviour already shows up in a dilute, weakly interacting Bose gas.) While explicit calculations were limited to the low <2,co region at low T, it was assumed that in superfluid 4 He: (a) The density fluctuation spectrum described by S(Q,(o) is directly proportional to Ga^(Q, co) at all values of Q. (b) Both functions exhibit a single excitation branch at high Q as well as low Q. The only difference between a gas and a liquid is that in a Bose liquid, the dispersion relation has a roton minimum at
g~2
A-1.
(c) This common excitation branch exhibited by both Ga^(Q, co) and S(Q,co) is crucially dependent on the existence of a Bose condensate. (d) The situation above Tx is completely different. This set of assumptions has had wide currency and still has strong proponents in the field-theoretic literature (see, in particular, Nepomnyashchy, 1992). It was first challenged by Pines (1966), who argued that in superfluid 4 He the phonon part of the quasiparticle spectrum in S(Q, co) was best described as a zero sound mode which exists both below and above Tx without much change. On the other hand, Pines did not discuss what this implied about the corresponding single-particle Green's functions Szepfalusy and Kondor (1974) and Griffin and Cheung (1973) first used the dielectric formalism of Ma and Woo (1967) to discuss the inter-related structure of both Ga^(Q,co) and S(Q,co), for all Q,co and T. A key feature of the dielectric formalism is how it draws attention to the possibility that a single-particle excitation and the zero sound can be hybridized by the Bose condensate no(T), allowing both renormalized excitations to appear in S(Q, co) below Tx. Precisely how this hybridization is realized in
12.1 Comments on the development of the new picture
273
superfluid 4 He, however, is not at all obvious (see the review by Griffin, 1991). The single-particle (SP) scenario described in the preceding paragraph was extended to finite temperatures by Griffin (1979a) as well as Griffin and Talbot (1981), in an attempt to explain the finite-temperature neutron-scattering data in the region 0.8 < Q < 2 A" 1 (Woods and Svensson, 1978). However, subsequent studies (Griffin, 1987, 1989) for small Q emphasized that, in fact, the phonon region is better described by the zero sound (ZS) limit (as suggested by Pines, 1966). Glyde and Svensson (1987) and Svensson (1989) suggested that these contradictory results could be understood if S(Q, co) described quite different excitation branches in the low- and high-Q regions. This was a very revolutionary concept since, starting with Landau (1947), the guiding assumption has always been that the phonon-maxon-roton quasiparticle dispersion relation described a single excitation branch in which there was no fundamental distinction between a phonon and a roton. The completely different temperature dependence of the intensity of the sharp peak observed in S(Q,co) at low Q (Woods, 1965b) and high Q (Woods and Svensson, 1978) was interpreted by Stirling and Glyde (1990) in terms of the dielectric formalism, based on the idea that the low-Q phonon was a collective density fluctuation while the high-g maxon-roton was an atomic-like excitation. This physical distinction between phonons and rotons goes back to Feynman (1954), MPN (1962) and Chester (1963). None of this early work, however, considered the specific role of the condensate no(T) in hybridizing the two excitations. Glyde and Griffin (1990) first illustrated how the coupling induced by the condensate could explain the continuous phonon-maxon-roton dispersion relation which is observed. As discussed in Section 6.3, the low-energy phonon part of the spectrum is somewhat subtle in the superfluid phase. Above 7^, such phonon resonances in S(Q,co) are zero sound density fluctuations, which also occur in normal liquid 3 He as well as in classical liquids (see Section 7.1). Below Tx, this zero sound mode still exists, although it is now associated with the effective fields produced by the condensate and non-condensate atoms, i.e., the two components of xnn as given by (5.11b). If the frequency spectrum associated with Gaj? and Xnn occurs at relatively low frequencies, the theory naturally leads to a zero sound phonon resonance in both Xnn and Gap which has a velocity which is temperature-independent (see Sections 6.3 and 7.2). It is perhaps not surprising that a zero sound density fluctuation
274
The new picture: some unsolved problems
dominates the spectrum of Xnn at low Q , both below and above Tx. Because of the Bose broken symmetry which sets in below Tx, however, this phonon mode also has finite weight in the single-particle Green's function Ga^(Q, co) and in fact dominates the /ow-energy elementary excitation spectrum at low temperatures. In a liquid, Bose statistics do not play a crucial role in the existence of this zero sound mode, but a condensate is crucial to ensure that this mode has finite weight as an elementary excitation in the superfluid phase. Thus we can say that the unique feature of superfluid 4 He is not so much that phonons exist as that the condensate allows these modes to play the role of elementary excitations below Tx. In contrast, the zero sound phonons in normal liquid 4 He or liquid 3 He do not play the role of elementary excitations. In the high-energy, short-wavelength region, we have argued that the maxon-roton peak in S(Q,co) has its origin as a pole in the singleparticle spectrum described by Gap(Q,co). This excitation probably exists above Tx without much change in the dispersion relation. These singleparticle states have high energy (>8 K) and exist only at relatively large wavevectors (^ 0.8 A" 1 ), and thus it seems physically reasonable that they would not be modified very much when the liquid goes from the normal to the superfluid phase (or for that matter, to the solid phase discussed in Chapter 11). The idea that a roton is physically quite different from a phonon has a long history. We recall that Feynman (1954) originally argued that the roton corresponds to a high-energy excited state involving the motion of a single atom in the potential well of its neighbors. Miller, Pines and Nozieres (1962) also viewed the roton as an atomic-like excitation modified by backflow. Chester gave several suggestive arguments that indicated that phonons and rotons were physically quite distinct excitations, concluding that (to quote from p. 65 of Chester, 1963): From Q = 0 up to Q — 1 A"1 the motions in the fluid consist of longitudinal density waves .... On the other hand, for Q ~ 2 A"1 we find that there is essentially no local density fluctuation and we have a single particle propagating fast through the liquid. The increase in effective mass of this atom comes from a large scale hydrodynamic backflow - with little or no change in the local fluid density. The region between Q = 1 A"1 and 2 = 2 A"1 must be interpreted as a rather complicated region in which we pass from a high frequency phonon mode at 1 A"1 to a single particle propagation at 2 A"1. This is clearly a possible transition in a liquid with an open structure and with low potential barriers. Our present analysis gives a reasonable scenario which allows for realiza-
12.1 Comments on the development of the new picture
275
tion of this kind of picture and extends these ideas to higher temperatures, to Tx and above. One of the most interesting implications of this new picture is that it suggests one should be able to understand the properties of normal liquid 4 He just above Tx in terms of a gas of roton-like excitations. Indeed, models based on excitations with a roton energy gap were used with partial success in the early literature (see, for example, Henshaw and Woods, 1961). Within the Glyde-Griffin picture, the "roton liquid" theory developed by Bedell, Pines and Fomin (1982) might also be appropriate above Tx. We recall that this phenomenological theory attempts to describe the thermodynamic properties of liquid 4 He in the region just below Tx in terms of an interacting gas of rotons, using a self-consistent field approach modelled after that used in the Landau Fermi liquid theory description of quasiparticle interactions in normal liquid 3 He. If the roton excitations are indeed intrinsic poles of the single-particle Green's functions (with an energy largely unaffected by the appearance of a condensate), the roton liquid theory formulation of Bedell, Pines and Fomin might be as relevant in the region immediately above Tx as it is below. (Of course, singularities in the specific heat and other thermodynamic quantities in the vicinity of Tx, specifically associated with the critical behaviour at a second-order phase transition, are not described by such a theory.) The relation of our scenario to the well known analysis of the excited states in a Bose liquid by Feynman (1954) is not completely clear. We recall that Feynman's analysis showed how Bose statistics was the key in ensuring that there were no other low-energy, long-wavelength excitations, apart from a phonon-like density fluctuation which by itself is not dependent on the nature of the quantum statistics involved. This work, however, made no reference to any role of a Bose condensate. The following remarks may be useful in understanding Feynman's work. In Section 1.1, we briefly reviewed the ideas behind the variational calculation of Feynman (1954). While Feynman (1953b, 1954) was motivated by a detailed microscopic examination of the various possible excited states in a Bose liquid, the variational theory he was led to is now viewed as a generic "single-mode approximation" for the densityresponse function. That is, if one starts with the assumption that S(Q,co) is dominated at all Q by a single undamped excitation of frequency COQ, the dispersion relation is given by the Feynman formula (1.6). This result is thus equally valid for normal liquid 4 He and liquid 3 He (for further discussion, see p. 915 of Mahan, 1990). The "dip" that such a disper-
276
The new picture: some unsolved problems
Fig. 12.2. The pair distribution function g(r) as a function of distance at 2.0 K, SVP. The solid line represents high-precision neutron data (Svensson et al, 1980) and the circles are path-integral Monte Carlo results. At this temperature, the calculated condensate fraction is 8%. When Bose symmetrization is not done, the only change in g(r) is shown by the dashed line [Source: Ceperley and Pollock, 1986].
sion relation exhibits occurs at the wavevector Q ~ 2 A l where the static structure factor S(Q) has its first maximum. Such a "roton dip" is thus a consequence of the short-range spatial correlations between atoms and has little to do with the quantum statistics the atoms obey. This fact is further emphasized by the path integral Monte Carlo simulations of Ceperley and Pollock (1986). As shown in Fig. 12.2, the short-range behaviour of the pair distribution function g(r) in superfluid 4 He is essentially unaffected when one does not carry out the Bose symmetrization. However, we recall the arguments given in Section 7.2 (see Fig. 7.17) against the assumption that the peak in S(Q,co) in normal liquids is associated with a single continuous excitation branch valid at both small and large Q. On the other hand, Bose statistics do play a crucial (but indirect) role in understanding why there are no other low-lying states in liquid 4 He besides the one generated by (1.4) (or the more modern versions such as (9.10)), at least at small wavevectors. By way of contrast, in normal liquid 3 He, at small wavevectors there are, in addition to the zero sound phonon mode, many other low-energy modes corresponding to the singleparticle Fermi quasiparticle excitations. It is Bose symmetrization which forces the excitations involving the motion of a single atom to have finite energy in the Feynman analysis. In turn, it is the Bose condensate that
12.2 Dielectric formalism for superfluid 3He-4He mixtures
277
allows the density fluctuation spectrum (phonons) to have a significant overlap on the field fluctuation spectrum described by the single-particle Green's function G(Q,co), justifying the implicit assumption of Feynman and Landau. Needless to say, by itself, our interpretation of the roton as an intrinsic single-particle excitation of normal liquid 4 He does not shed any light on why it has the precise energy spectrum hypothesized in Fig. 7.23. In this connection, Feynman's explicit discussion of the motion of a single atom in liquid 4 He is of great interest and deserves careful study. Very recently, Stringari (1992) has suggested a promising way of studying this question in a quantitative manner (see remarks at end of Section 9.1). 12.2 Dielectric formalism for superfluid 3 He- 4 He mixtures In this book, the dynamics of superfluid 4 He has been discussed starting from the microscopic theory of Bose-condensed liquids. The various scenarios we have developed concerning the nature of the excitations have the common feature that the Bose condensate plays a pivotal role. In order to test these ideas, we have seen that neutron-scattering experiments done at varying temperatures and pressures have been of special importance. Indeed the recent picture put forward by Glyde and Griffin (1990) was developed in an attempt to understand the temperature-dependent line shapes as the liquid goes from the superfluid to normal phase (see Section 7.2). In developing these ideas, the study of superfluid 3 He- 4 He mixtures is particularly promising since the 3 He concentration gives a new parameter which can be varied, in addition to the pressure and temperature. In particular, recent high-momentum-transfer neutron-scattering data (see Chapter 4) indicate that the condensate fraction in a superfluid 3 He- 4 He solution of 12% is considerably increased (Wang, Sosnick and Sokol, 1992). This suggests that hybridization effects induced by the 4 He condensate may be much stronger in mixtures than in pure superfluid 4 He. Talbot (1983) and Talbot and Griffin (1984c) have worked out the formal extension of the dielectric formalism given in Section 5.1 to a fully interacting system of Fermions and Bosons, with due allowance for a Bose condensate. This formalism can be applied to superfluid 3 He- 4 He mixtures. Just as in the case of pure liquid 4 He, the Bose broken symmetry leads to a coupling of the various correlation functions which is made most manifest by working in terms of irreducible, proper diagrams. The formal analysis is more complicated than that given in
278
The new picture: some unsolved problems
Section 5.1 by the necessity to include a whole new class of response functions involving the 3 He atoms. There is already a considerable body of literature on the elementary excitations and collective modes in superfluid 3 He- 4 He mixtures. A useful introduction to the literature is given in the review by Glyde and Svensson (1987). Fak et al. (1990) give a detailed analysis of recent neutron scattering data over a wide range of temperatures (0.07-1.5 K), with extensive references. The phenomenological quasiparticle theory of dilute 3 He- 4 He mixtures is discussed in detail by Baym and Pethick (1978), Ruvalds (1978) as well as in Chapter 24 of Khalatnikov (1965). Much of the work on mixtures to date has been concerned with low 3 He concentrations at low temperatures, with an emphasis on using the system as a way of studying a dilute Fermi gas of quasiparticles. In contrast, our major interest is to see how the phonon-maxon-roton excitation associated with the 4 He atoms is modified by the 3 He atoms. With this end in mind, it would be useful to have more neutron-scattering data at high 3 He concentrations, over a wide range of temperature (including above the superfluid transition). Besides the dielectric formalism we review in this section, the variational and phenomenological approaches reviewed in Chapter 9 have all been extended to superfluid 3 He- 4 He mixtures at T = 0. In connection with neutron-scattering data, the polarization potential approach of Section 9.2 is discussed by Hsu, Pines and Aldrich (1985) while the memory function formalism has been used by Lucke and Szprynger (1982) and Szprynger and Lucke (1985). As in the case of pure superfluid 4 He, these approaches can be brought into contact with the dielectric formalism by suggesting useful parameterizations of the "regular" functions in the latter theory. The coherent (spin-independent) neutron-scattering differential crosssection is proportional to (see Section 2.1)
V
)
)
,
(12.1)
where b3 and b* are the coherent neutron-scattering lengths for 3 He and He atoms, respectively, and x3 is the molar concentration of 3 He (see Fak et al., 1990, for further details). The dynamic structure factors are related to various number density response functions in the usual way:
4
Sab(Q,co) =
^=[N(co)
+ 1] Im Xab(Q,co) ,
(12.2)
12.2 Dielectric formalism for superfluid 3He-4He mixtures
279
where XabiQ, co) can be related to the Bose Matsubara Fourier components T
(pa(Q,T)pb(-Q))
(12.3)
(see (3.41) and (5.4) for analogous definitions for pure 4 He). Here a, b can stand for 3 He or 4 He and pa(Q) is the Fourier component of the number density operator of 3 He (a = 3) and 4 He (a = 4), defined as in (2.5) and (3.12). In the interpretation of neutron-scattering data on dilute mixtures, the cross-term S34 in (12.1) is often neglected, in which case the resonances in the total scattering intensity can be identified with renormalized 4 He and 3 He excitations associated with S44 and 533, respectively. Since 3 He and 4 He atoms have the same electronic polarizability, the cross-section for Brillouin light scattering (Stephen, 1976) can be expressed in terms of the dynamic structure factor associated with the total number density n = n^ + «4, Sab(Q,a/) .
(12.4)
a,b
As we discuss in Section 6.2, Brillouin light scattering probes the hydrodynamic domain. In superfluid 3 He- 4 He mixtures, this region is especially interesting since the intensity of second sound in Snn(Q,co) is greatly enhanced by the presence of the 3 He concentration fluctuations (Khalatnikov, 1965). The hydrodynamic spectrum of superfluid 3 He- 4 He mixtures has been thoroughly mapped out by light scattering (see, for example, Stephen, 1976; Rockwell, Benjamin and Greytak, 1975). It would be interesting to use the two-fluid equations for mixtures (Khalatnikov, 1965) to derive the single-particle Green's functions for the 4 He atoms in the asymptotic region of low Q and co (the analogous calculation for pure 4 He is sketched in Section 6.2). We now briefly summarize the results of Talbot and Griffin (1984c). These formally rigorous expressions are valid for all Q, co and T and are very relevant to inelastic neutron-scattering studies. In terms of irreducible contributions xab, one finds 133=
L
(12.5)
280
The new picture: some unsolved problems
where we define e(Q,oj) = 1 - 7(6)1X44 + 2X34 + Z33] , 1 > A(Q,CO) = Z447(0X33 - B 4 7 ( 0 X 4 3 • J
(12.6)
We make use of the fact that the bare interatomic potentials between the Helium atoms are identical, Vab = 7, for a, b = 3,4. As expected, all four response functions Xab in (12.5) share the same poles, which are given by the zeros of the dielectric function e(Q,a>) (Bartley et al, 1973). We also note that Xab = Z T7^ , (12.7) ajb
V
-
V
* ™
where xm = ^Xab- The simplest MFA approximation (see Section 3.4) aj>
corresponds to using the free-gas response functions for xah, i.e., xAA = X44, X 33 = X33
and
X34 = X34 = 0.
The crucial step in the dielectric formalism for Bose-condensed systems is to split irreducible contributions into proper (regular) and improper (condensate) parts, as discussed in Section 5.1. We first define (a = 3 or 4) the mixed density-field correlation functions
A= J ^ ^ ( / U Q , T ) S p ,
(12.8)
the generalization of (5.4). One finds that the irreducible parts are given by (see (5.11a)) lav(Q, iojn) = AflAI(Q, icon)G$(Q,
Uon) ,
(12.9)
where G$ is the matrix 4 He single-particle Green's function which only contains proper self-energies £j$ and the summation convention is used for repeated Greek subscripts. Thus in mixtures, we have two Bose symmetry-breaking vertex functions A4/l and A3^, which play a key role in that they determine the hybridization of the poles of Gj$ and XabThe irreducible contribution to the response function separates into two parts, lab = tab + fab >
(12-10)
with (see (5.11b)) ^K
(12.11)
12.2 Dielectric formalism for superfluid 3He-4He mixtures
281
The improper 4 He self-energy is given by ^
^ v 4
+ Av3) ,
(12.12)
where (see (5.10)) eR = 1 - V(Q)tn •
(12.13)
Similarly, one can derive the equivalent of (3.47) or (5.24) for the response functions Xab in mixtures, lab = \aG($Kb
+ XRab,
(12.14)
where A ^ = A^a/^R and eR is defined in (12.13). This is the key result of the dielectric formalism for mixtures since it shows how the spectrum of the 4 He single-particle Green's function G$ is coupled into the numberdensity-response functions Xab- The zeros of e(Q,co) in (12.6) determine the poles of both Xab and Gjfj. In contrast, the spectrum of the 3 He single-particle Green's function G(3)(Q,co) is not determined by the zeros of e(Q, co). This difference can already be seen by the fact that G^ and Xab involve Bose Matsubara frequencies while G(3) involve Fermi Matsubara frequencies. Talbot and Griffin (1984c) have derived the equivalent of the Ward identities given in Section 5.1 and used them to obtain various exact results in the zero-frequency limit. As one example, one can show that the vertex function A3^(Q, a> = 0) = 0. At the present time, this formalism is mainly useful in showing how the Bose broken symmetry leads to a certain inevitable coupling between the poles of Xab and G$ in mixtures. The equivalent of the analysis given in Section 7.2 for pure superfluid 4 He requires further high-resolution neutron-scattering data as a function of temperature, especially in the region below and above the superfluid transition. This seems like one of the most promising areas of research in the near future. Most phenomenological studies of the zeros of e(Q,a>) defined in (12.6) approximate ^33 as the Lindhard function of a non-interacting gas of Fermi quasiparticles and take %44 in a single-mode approximation appropriate to low temperatures (see, for example, Bartley et al, 1973; Hsu, Pines and Aldrich, 1985; Szprynger and Lucke, 1985). One immediately sees that the resulting response functions Xab will exhibit complicated hybridization effects between the Fermi quasiparticle p-h spectrum and the Boson pole of £44. In the polarization approach, for example, £44 is modelled by an expression similar to (9.28). In neutron-scattering studies,
282
The new picture: some unsolved problems
we need to know the 3 He Lindhard function at relatively large values of Q (say 1-1.5 A" 1 ). The original Landau-Pomeranchuk (LP) spectrum for a single 3 He atom in bulk liquid 4 He is Q2/2m*3, with an effective mass m\ ~ 2.4ra3. With this LP quadratic spectrum, the quasiparticle particle-hole spectrum is a band described by Q2/2m\ + Qvp. For a dilute solution, this is a very narrow band centred on the LP 3 He quasiparticle branch (at X3 = 0.06, we have kF ~ 0.3 A" 1 ). At such low concentrations, there is no collective zero sound mode associated with the 3 He atoms (in contrast to pure liquid 3 He, shown in Figs. 7.5, 7.6 and 7.18). Thus if we use the LP single-particle spectrum, the only 3 He density fluctuation branch in dilute mixtures is a narrow band which would cross the maxon-roton 4 He excitation at around Q ~ 1.7 A" 1 and co ~ 10 K. This would be expected to produce strong hybridization effects between these two branches in this cross-over region and a strong-level repulsion of the p-h branch (Bartley et al, 1973). This cross-over region has been extensively discussed in the literature (Hsu et a/., 1985; Szprynger and Lucke, 1985) with the conclusion that it is crucial to include the lower density of 4 He atoms in mixtures and also that the LP single-particle spectrum must be modified at the large values of Q probed in the neutron-scattering experiments. In a formal sense, these hybridization effects are analogous to the coupling between the single-particle maxon-roton spectrum and the 4 He p-h branch which we considered in the case of pure 4 He (see Section 7.2). A natural extension of the model calculations given at the end of Section 7.2 (see (7.23)) would be to add in the contribution X33 to describe the p-h spectrum of the 3 He atoms. The neutron-scattering line shapes from superfluid mixtures are clearly a rich area for future studies, especially at higher temperatures and concentrations where the collective dynamics of both 3 He and 4 He are important. The dielectric formalism should give a rigorous basis for understanding and describing the various hybridization effects which occur. 12.3 Suggestions for future research In this concluding section, we pull together a few suggestions as to where further work would be especially useful. Many of these suggestions have already been made in the earlier chapters. Our selection of topics for further research is influenced by what we believe is the most interesting feature of superfluid 4 He and 3 He- 4 He mixtures, namely the unique
12.3 Suggestions for future research
283
dynamical structure of various correlation functions which arises because of the Bose condensate.
Theoretical (1) Develop parameterized forms for <S(Q, co), as a function of temperature, which are consistent with the general structure induced by the Bose condensate. The preliminary analysis given at the end of Section 7.2 must be extended to include the two-excitation spectra before one can successfully deal with the very interesting maxon region Q ~ 1 A" 1 . (2) Work out the finite-temperature version of the Zawadowski, Ruvalds and Solana (1972) analysis of the condensate-induced hybridization between the one-roton and two-roton branches (Section 10.2). (3) The polarization potential approach (see Section 9.2) seems to give a useful way of parameterizing the general structure of correlation functions as predicted by the dielectric formalism. The PP approach should be extended to finite temperatures (> 1 K) by including the thermally excited particle-hole excitations. The high-energy multiparticle excitations have been discussed (at T = 0) by Hess and Pines (1988). (4) Formulate the Glyde-Griffin scenario at T = 0 in terms of a variational many-body wavefunction (see Section 9.1). The recent work of Stringari (1992) seems promising in this connection. (5) Make use of the first frequency-moment sum rules specific to Bosecondensed fluids (see Section 8.1 and 8.3). Extension of these to third frequency-moment sum rules would be of interest. (6) Work out the energy of single-particle excitations in a Bose liquid from first principles (Feynman, 1953b; Stringari, 1992). (7) The interaction V(Q) which appears in the final formulas of the dielectric formalism is assumed to be renormalized to some appropriate t-matrix when one includes multiple scattering. It would be useful to carry out this procedure more explicitly, summing up all contributions to regular quantities which involve two isolated propagator lines (see Section 5.4). (8) The one-loop approximation to the dielectric formalism (see Section 5.3) should be worked out in detail for superfluid 3 He- 4 He mixtures discussed in Section 12.2.
284
The new picture: some unsolved problems Experimental
(1) Measure the temperature dependence of the S(Q,co) line shape at high energies in the cross-over region around Q ~ 2.4 A" 1 (see Figs. 7.19 and 7.20). This should be done at a series of temperatures above and below 7^. (2) Measure the temperature dependence and dispersion of the highenergy peak which shows up in the low-Q neutron data (see Figs. 7.1 and 10.4). Is this a two-roton bound state (as usually assumed) or is it a remnant of a single-particle excitation, as suggested in Figs. 7.22 and 7.23? (3) High-resolution experiments on the temperature dependence of the S(Q,co) line shape at a series of different pressures would be very useful, especially in the intermediate maxon region. As shown very dramatically by Fig. 7.8, at a pressure of 20 bar, the normal distribution is peaked at a much higher energy than the maxon peak. As a result, S(Q,co) under high pressure exhibits the Glyde-Griffin scenario in a clearer fashion than the SVP data (where the superfluid and normal distributions are peaked at very similar energies). (4) Measure the pressure and temperature dependence of the neutronscattering line shapes in 3 He- 4 He mixtures in the vicinity of the tricritical point (i.e., at high concentrations and temperatures). (5) Develop new experimental probes of the existence and properties of rotons above the superfluid transition temperature. In the GlydeGriffin picture, rotons (more generally, maxon-rotons) are intrinsic single-particle excitations of normal liquid 4 He. The rotons develop finite weight in 5(Q,co) only as a result of the Bose broken symmetry. Thus neutron scattering is not a useful probe of rotons above T^. Quantum evaporation studies would seem one way of studying singleparticle excitations above and below Tx (see Caroli et al, 1976; Maris, 1992).
References
Abrikosov, A.A., Gor'kov, L.P. and Dzyaloshinskii, I.E. (1963), Methods of Quantum Field Theory in Statistical Physics (Prentice-Hall, Englewood Cliffs, N.J.). Ahlers, G. (1978), in Quantum Liquids, ed. by J. Ruvalds and T. Regge (North-Holland, New York), p. 1. Aldrich III, C.H., Pethick, C.G. and Pines, D. (1976), Phys. Rev. Lett. 37, 845. Aldrich III, C.H. and Pines, D. (1976), Journ. Low Temp. Phys. 25, 673. Aldrich III, C.H. and Pines, D. (1978), Journ. Low Temp. Phys. 32, 689. Allen, J.F. and Misener, A.D. (1938), Nature 141, 75. Ambegaokar, V., Conway, J. and Baym, G. (1965), in Lattice Dynamics, ed. by R.F. Wallis (Pergamon, New York), p. 261. Andersen, K.H., Stirling, W.G., Scherm, R., Stunault, A., Fak, B., Dianoux, A.J. and Godfrin, H. (1991), to be published. Anderson, P.W. (1966), Rev. Mod. Phys. 38, 298. Anderson, P.W. (1984), Basic Notions of Condensed Matter Physics (Benjamin/Cummings, Menlo Park, Calif.). Bartley, D.L., Robinson, J.E. and Wong, V.K. (1973), Journ. Low Temp. Phys. 12, 71. Bartley, D.L. and Wong, V.K. (1975), Phys. Rev. B12, 3775. Baym, G. (1969), in Mathematical Methods in Solid State and Superfluid Theory, ed. by R.C. Clark and G.H. Derrick (Oliver and Boyd, Edinburgh), p. 121. Baym, G., and Kadanoff, L.P. (1961), Phys. Rev. 124, 287. Baym, G. and Pethick, C.J., (1978), in The Physics of Liquid and Solid Helium, ed. by K.H. Bennemann and J.B. Ketterson (Wiley, New York), Vol. II. Bedell, K., Pines, D. and Fomin, I. (1982), Journ. Low Temp. Phys. 48, 417. Bedell, K., Pines, D. and Zawadowski, A. (1984), Phys. Rev. B29, 102. Beliaev, S.T. (1958a), Sov. Phys. - JETP 7, 289. Beliaev, S.T. (1958b), Sov. Phys. - JETP 7, 299. Bendt, P.J., Cowan, R.D. and Yarnell, J.H. (1959), Phys. Rev. 113, 1386. Bijl, A. (1940), Physica 7, 869. Bogoliubov, N.N. (1947), J. Phys. U.S.S.R. 11, 23. Bogoliubov, N.N. (1963, 1970), Lectures on Quantum Statistics (Gordon and Breach, New York), Vol. 2. Boon, J.P. and Yip, S. (1980), Molecular Hydrodynamics (McGraw-Hill, New York). Brown, G.V. and Coopersmith, M.H. (1969), Phys. Rev. 178, 327. 285
286
References
Brueckner, K.A. and Sawada, K. (1957), Phys. Rev. 106, 1117. Buyers, W.J.L., Sears, V.F., Lonngi, P.A. and Lonngi, D.A. (1975), Phys. Rev. All, 697. Campbell, C.E. (1978), in Progress in Liquid Physics, ed. by C.A. Croxton (Wiley, New York), p. 213. Campbell, C.E. and Clements, B.E. (1989), in Elementary Excitations in Quantum Fluids, ed. by K. Ohbayashi and M. Watabe (Springer-Verlag, Berlin), p. 8. Caroli, C, Roulet, B. and Saint-James, D. (1976), Phys. Rev. B13, 3875. Carraro, C. and Koonin, S.E. (1990), Phys. Rev. Lett. 65, 2792. Carraro, C. and Rinat, A.S. (1992), Phys. Rev. B45, 2945. Ceperley, D.M. and Pollock, E.L. (1986), Phys. Rev. Lett. 56, 351. Ceperley, D.M. and Pollock, E.L. (1987), Can. J. Phys. 65, 1416. Chester, G.V. (1963), in Liquid Helium, ed. by G. Careri (Academic Press, New York), p. 51. Chester, G.V. (1969), in Lectures in Theoretical Physics (Vol. XI-B): Quantum Fluids and Nuclear Matter, ed. by K.T. Mahanthappa and W.E. Britten (Gordon and Breach, New York), p. 253. Chester, G.V. (1975), in The Helium Liquids, ed. by J.G.M. Armitage and I.E. Farquhar (Academic Press, London), p. 1. Cheung, T.H. (1971), Ph.D. Thesis, University of Toronto, unpublished. Cheung, T.H. and Griffin, A. (1970), Can. Journ. Phys. 48, 2135. Cheung, T.H. and Griffin, A. (1971a), Phys. Lett. A35, 141. Cheung, T.H. and Griffin, A. (1971b), Phys. Rev. A4, 237. Clark, J.W. and Ristig, M.L. (1989), in Momentum Distributions, ed. by R.N. Silver and PE. Sokol (Plenum, New York), p. 39. Cohen, M. and Feynman, R.P. (1957), Phys. Rev. 107, 13. Copley, J.R.D. and Lovesey, S.W. (1975), Rep. Prog. Phys. 38, 461. Cowley, R.A. (1968), Rep. Prog. Phys. 31, 123. Cowley, R.A. and Buyers, W.J.L. (1969), J. Phys. C: Solid State Phys. 2, 2262. Cowley, R.A. and Woods, A.D.B. (1968), Phys. Rev. Lett. 21, 787. Cowley, R.A. and Woods, A.D.B. (1971), Can. J. Phys. 49, 177. Cummings, F.W., Hyland, G.J. and Rowlands, G. (1970), Phys. Condensed Matter 12, 90. Dietrich, O.W, Graf, E.H., Huang, C.G. and Passell, L. (1972), Phys. Rev. A5, 1377. Donnelly, R.J. (1991), Quantized Vortices in Helium II (Cambridge University Press, New York). Dzugutov, M. and Dahlborg, U. (1989), Phys. Rev. 40A, 4103. Einstein, A. (1925), Sitz. Berlin Preuss. Akad. Wiss. , p. 3. Fak, B. and Andersen, K.H. (1991), Phys. Lett. A160, 468. Fak, B , Guckelsberger, K., Korfer, M., Scherm, R. and Dianoux, A.J. (1990), Phys. Rev. B41, 8732. Family, F. (1975), Phys. Rev. Lett. 34, 1374. Feenberg, E. (1969), Theory of Quantum Fluids (Academic Press, New York). Ferrell, R.A., Menyhard, N , Schmidt, H., Schwabl, F. and Szepfalusy, P. (1968), Ann. Phys. (New York) 47, 565. Fetter, A.L. (1970), Ann. Phys. (New York) 60, 464. Fetter, A.L. (1972), Journ. Low Temp. Phys. 6, 487. Fetter, A.L. and Walecka, J.D. (1971), Quantum Theory of Many Particle Systems (McGraw-Hill, New York). Feynman, R.P (1953a), Phys. Rev. 91, 1291.
References
287
Feynman, R.R (1953b), Phys. Rev. 91, 1301. Feynman, R.R (1954), Phys. Rev. 94, 262. Feynman, R.R and Cohen, M. (1956), Phys. Rev. 102, 1189. Forster, D. (1975), Hydrodynamic Fluctuations, Broken Symmetry and Correlation Functions (W.A. Benjamin, Reading, Mass.). Fredkin, D.R. and Werthamer, N.R. (1965), Phys. Rev. 138A, 1528. Fukushima, K. and Iseki, F. (1988), Phys. Rev. B38, 4448. Gavoret, J. (1963), These de Doctorat es Sciences Physiques, Paris, Annales de Physique 8, 441. Gavoret, J. and Nozieres, P. (1964), Ann. Phys. (New York) 28, 349. Gay, C. and Griffin, A. (1985), Journ. Low Temp. Phys. 58, 479. Gersch, H.A. and Rodriguez, LJ. (1973), Phys. Rev. A8, 905. Gillis, N.S. and Werthamer, N.R. (1968), Phys. Rev. 167, 607. Gillis, N.S., Werthamer, N.R. and Koehler, T.R. (1968), Phys. Rev. 165, 951. Ginzburg, V.L. (1943), J. Exp. Theor. Phys. (U.S.S.R.) 13, 243. Giorgini, S., Pitaevskii, L. and Stringari, S. (1992), Phys. Rev. B46, 6374. Giorgini, S. and Stringari, S. (1990), Physica B 165-166, 511. Girardeau, M. and Arnowitt, R. (1959), Phys. Rev. 113, 755. Glyde, H.R. (1971), Can. Journ. Phys. 49, 761. Glyde, H.R. (1974), Can. Journ. Phys. 52, 2281. Glyde, H.R. (1976), in Rare Gas Solids, ed. by M.L. Klein and J.A. Venables (Academic Press, New York), vol. 1, p. 382. Glyde, H.R. (1984), in Condensed Matter Research Using Neutrons, ed. by S.W. Lovesey and R. Scherm (Plenum, New York) p. 95. Glyde, H.R. (1992a), Phys. Rev. B45, 7321. Glyde, H.R. (1992b), to be published. Glyde, H.R. and Goldman, V.V. (1976), Journ. Low Temp. Phys. 25, 601. Glyde, H.R. and Griffin, A. (1990), Phys. Rev. Lett. 65, 1454. Glyde, H.R. and Svensson, E.C. (1987), in Methods of Experimental Physics, ed. by D.L. Price and K. Skold (Academic Press, New York), vol. 23, ch. 13 . Gotze, W. and Lucke, M. (1976), Phys. Rev. B13, 3825. Greytak, T.J. (1978), in Quantum Liquids, ed. by J. Ruvalds and T. Regge (North-Holland, Amsterdam), p. 121. Greytak, T.J. and Kleppner, D. (1984), in New Trends in Atomic Physics, ed. by G. Grynberg and R. Stora (North-Holland, New York), Vol. II, p. 1127. Greytak, T.J. and Yan, J. (1969), Phys. Rev. Lett. 22, 987. Greytak, T.J. and Yan, J. (1971), Proc. of the 12th International Conference on Low Temperature Physics, Kyoto, 1970, ed. by E. Kanda (Keigaku Publ. Comp., Tokyo), p. 89. Griffin, A. (1972), Phys. Rev. A6, 512. Griffin, A. (1979a), Phys. Rev. B19, 5946. Griffin, A. (1979b), Phys. Lett. 71A, 237. Griffin, A. (1980), Phys. Rev. B22, 5193. Griffin, A. (1981), Journ. Low Temp. Phys. 44, 441. Griffin, A. (1984), Phys. Rev. B30, 5057. Griffin, A. (1985), Phys. Rev. B32, 3289. Griffin, A. (1987), Can. Journ. Phys. 65, 1368. Griffin, A. (1988), Physica C156, 12. Griffin, A. (1989), in Elementary Excitations in Quantum Fluids, ed. by K. Ohbayashi and M. Watabe (Springer-Verlag, Berlin), p. 23.
288
References
Griffin, A. (1991), in Excitations in 2D and 3D Quantum Fluids, ed. by A.F.G. Wyatt and H J . Lauter (Plenum, New York), p. 15. Griffin, A. and Cheung, T.H. (1973), Phys. Rev. A7, 2086. Griffin, A. and Payne, S. (1986), Journ. Low Temp. Phys. 64, 155. Griffin, A. and Svensson, E.C. (1990), Physica B165-166, 487. Griffin, A. and Talbot, E. (1981), Phys. Rev. B24, 5075. Griffin, A., Wu, W.C. and N. Lambert (1992), to be published. Halley, J.W. (1969), Phys. Rev. 181, 338. Halley, J.W. (1989), in Elementary Excitations in Quantum Fluids, ed. by K. Ohbayashi and M. Watabe (Springer-Verlag, Berlin), p. 106. Halley, J.W. and Korth, M.S. (1991), in Excitations in 2D and 3D Quantum Fluids, ed. by A.F.G. Wyatt and H.J. Lauter (Plenum, New York), p. 91. Hansen, J.P. and McDonald, I.R. (1986), Theory of Simple Liquids, Second Ed. (Academic Press, London). Harling, O.K. (1970), Phys. Rev. Lett. 24, 1046. Harling, O.K. (1971), Phys. Rev. A3, 1073. Henshaw, D.G. and Woods, A.D.B. (1961), Phys. Rev. 121, 1266. Herwig, K.W, Sokol, P.E., Snow, WM. and Blasdell, R.C. (1991), Phys. Rev. B44, 308. Herwig, K.W, Sokol, RE., Sosnick, T.R., Snow, WM. and Blasdell, R.C. (1990), Phys. Rev. B41, 103. Hess, D.W and Pines, D. (1988), Journ. Low Temp. Phys. 72, 247. Hilton, P.A., Cowley, R.A., Scherm, R. and Stirling, WG. (1980), J. Phys. C: Solid State Phys. 13, L295. Hohenberg, PC. (1967), Phys. Rev. 158, 383. Hohenberg, PC. (1973), Journ. Low Temp. Phys. 11, 745. Hohenberg, PC. and Martin, PC. (1964), Phys. Rev. Lett. 12, 69. Hohenberg, PC. and Martin, PC. (1965), Ann. Phys. (New York) 34, 291. Hohenberg, PC. and Platzman, P.M. (1966), Phys. Rev. 152, 198. Horner, H. (1967), Z. Physik 205, 72. Horner, H. (1971), Z. Physik 242, 432. Horner, H. (1972a), Phys. Rev. Lett. 29, 556. Horner, H. (1972b), Journ. Low Temp. Phys. 8, 511. Horner, H. (1974a), in Proceedings of the 13th International Conference on Low Temperature Physics, ed. by K.D.T. Timmerhaus, WJ. O'Sullivan and E.F. Hammel (Plenum Press, New York), vol. 1, p. 3. Horner, H. (1974b), in Proceedings of the 13th International Conference on Low Temperature Physics, ed. by K.D.T. Timmerhaus, WJ. O'Sullivan and E.F. Hammel (Plenum Press, New York), vol. 1, p. 125. Horner, H. (1974c), in Dynamical Properties of Solids, ed. by G.K. Horton and A.A. Maradudin (North-Holland, Amsterdam), vol. I, p. 451. Hsu, W, Pines, D. and Aldrich III, C.H. (1985), Phys. Rev. B32, 7179. Huang, K. (1964), in Studies in Statistical Mechanics, ed. by J. De Boer and G.E. Uhlenbeck (North-Holland, Amsterdam), vol. II. Huang, K. (1987), Statistical Mechanics, Second Edition (John Wiley, New York). Huang, K. and Klein, A. (1964), Ann. Phys. (New York) 30, 203. Hugenholtz, N. and Pines, D. (1959), Phys. Rev. 116, 489. Inkson, J.C. (1984), Many-Body Theory of Solids (Plenum, New York). Iwamoto, F. (1970), Prog. Theoret. Phys. (Kyoto) 44, 1135.
References
289
Iwamoto, F. (1989), in Elementary Excitations in Quantum Fluids, ed. by K. Ohbayashi and M. Watabe (Springer-Verlag, Berlin), p. 117. Jackson, H.W. (1969), Phys. Rev. 185, 186. Jackson, H.W. (1973), Phys. Rev. A8, 1529. Jackson, H.W. (1974), Phys. Rev. A10, 278. Juge, K.J. and Griffin, A. (1993), to be published. Kadanoff, L.P. and Baym, G. (1962), Quantum Statistical Mechanics (Benjamin, New York). Kadanoff, L.P. and Martin, P.C. (1963), Ann. Phys. (New York) 24, 419. Kalos, M.H., Lee, M.A., Whitlock, PA. and Chester, G.V. (1981), Phys. Rev. B24, 115. Kapitza, P. (1938), Nature 141, 74. Khalatnikov, I.M. (1965), An Introduction to the Theory of Superfluidity (Benjamin, New York). Kirkpatrick, T.R. (1984), Phys. Rev. B30, 1266. Kirkpatrick, T.R. and Dorfman, J.R. (1985), Journ. Low Temp. Phys. 58, 399, 301; 59, 1. Kitchens, T.A., Shirane, G., Minkiewicz, VJ. and Osgood, E.B. (1972), Phys. Rev. Lett. 29, 552. Koehler, T.R. (1967), Phys. Rev. Lett. 18, 654. Koehler, T.R. (1968), Phys. Rev. 17, 942. Koehler, T.R. and Werthamer, N.R. (1971), Phys. Rev. A3, 2074. Koehler, T.R. and Werthamer, N.R. (1972), Phys. Rev. A5, 2230. Kondor, I. and Szepfalusy, P. (1968), Ada Phys. Hung. 24, 81. Kwok, P.C. (1967), in Solid State Physics, ed. by F. Seitz, D. Turnbull and H. Ehrenreich (Academic Press, New York), vol. 20, p. 306. Kwok, P.C. and Martin, P.C. (1966), Phys. Rev. 142, 495. Landau, L.D. (1941), J. Phys. U.S.S.R. 5, 71. Landau, L.D. (1947), J. Phys. U.S.S.R. 11, 91. Landau, L.D. (1949), Phys. Rev. 75, 884. Landau, L.D. and Khalatnikov, I.M. (1949), Zh. Eksp. Teor. Fiz. 19, 637. Landau, L.D. and Lifshitz, E.M. (1959), Statistical Physics (Pergamon, Oxford). Lee, T.D., Huang, K. and Yang, C.N. (1957), Phys. Rev. 106, 1135. Leggett, A.J. (1975), Rev. Mod. Phys. 47, 331. Lifshitz, E.M. and Pitaevskii, L.P. (1980), Statistical Physics, Part 2, (Pergamon, Oxford). London, F. (1938a), Nature 141, 643. London, F. (1938b), Phys. Rev. 54, 947. London, F. (1950), Superfluids, Vol. 1: Macroscopic Theory of Superconductivity (Wiley, New York). London, F. (1954), Superfluids, Vol. 2: Macroscopic Theory of Superfluid Helium (Wiley, New York). Lucke, M. (1980), in Modern Trends in the Theory of Condensed Matter, Proc. of the 16th Karpacz Winter School of Theoretical Physics, ed. by A. Pekalski and J. Przystawa (Springer-Verlag, Berlin), p. 74. Lucke, M. and Szprynger, A. (1982), Phys. Rev. B26, 1374. Ma, S.K. (1971), Journ. Math. Phys. 12, 2157. Ma, S.K., Gould, H. and Wong, V.K. (1971), Phys. Rev. A3, 1453. Ma, S.K. and Woo, C.W (1967), Phys. Rev. 159, 165. Mahan, G. (1990), Many-Particle Physics, Second Edition (Plenum, New York).
290
References
Manousakis, E. (1989), in Momentum Distributions, ed. by R.N. Silver and RE. Sokol (Plenum, New York) p. 81. Manousakis, E. and Pandharipande, V.R. (1984), Phys. Rev. B30, 5064. Manousakis, E. and Pandharipande, V.R. (1986), Phys. Rev. B33, 150. Manousakis, E., Pandharipande, V.R. and Usmani, Q.N. (1985), Phys. Rev. B33, 7022. Maradudin, A.A. and Fein, A.E. (1962), Phys. Rev. 128, 2589. Maris, H.J. (1977), Rev. Mod. Phys. 49, 341. Maris, H.J. (1992), Journ. Low Temp. Phys. 87, 773. Maris, H.J. and Massey, W.E. (1970), Phys. Rev. Lett. 25, 220. Marshall, W. and Lovesey, S.W. (1971), Theory of Thermal Neutron Scattering (Oxford University Press, New York). Martel, P., Svensson, E.C., Woods, A.D.B., Sears, V.F. and Cowley, R.A. (1976), Journ. Low Temp. Phys. 32, 285. Martin, PC. (1965), in Low Temperature Physics - LT9, ed. by J.G. Daunt, D.O. Edwards, F.J. Milford and M. Yaqub (Plenum, New York), p. 9. Martin, PC. (1968), in Many-Body Physics, ed. by C. De Witt and R. Balian (Gordon and Breach, New York), p. 37. McMahan, A.K. and Beck, H. (1973), Phys. Rev. A8, 3247. McMillan, WL. (1965), Phys. Rev. 138, A442. Meisel, M.W. (1992), Physica B178, 121. Mezei, F. and Stirling, W.G. (1983), in 75th Jubilee Conference on Helium-4, ed. by J.G.M. Armitage (World Scientific, Singapore), p. 111. Miller, A., Pines, D. and Nozieres, P. (1962), Phys. Rev. 127, 1452. Mineev, V.R (1980), Sov. Phys. - JETP Lett. 32, 489. Minkiewicz, V.J., Kitchens, T.A., Shirane, G. and Osgood, E.B. (1973), Phys. Rev. A8, 1513. Mook, H.A. (1974), Phys. Rev. Lett. 32, 1167. Mook, H.A. (1983), Phys. Rev. Lett. 51, 1454. Mook, H.A., Scherm, R. and Wilkinson, M.K. (1972), Phys. Rev. A6, 2268. Murray, R.L., Woerner, R.L. and Greytak, T.J. (1975), J. Phys. C: Sol. State Phys. C8, L90. Nakajima, S. (1971), Prog. Theoret. Phys. (Kyoto) 45, 353. Nelson, D.R. (1983), in Phase Transitions and Critical Phenomena, ed. by C. Domb and J.L. Lebowitz (Academic Press, London), Vol. 7, p. 1. Nepomnyashchii, YA. (1983), Sov. Phys. - JETP 58, 722. Nepomnyashchy, YA. (1992), Phys. Rev. B46, 6611. Nepomnyashchii, YA. and Nepomnyashchii, A.A. (1974), Sov. Phys. - JETP 38, 134. Nepomnyashchii, YA. and Nepomnyashchii, A.A. (1978), Sov. Phys. - JETP 48, 493. Nozieres, P. (1964), Theory of Interacting Fermi Systems (Benjamin, New York). Nozieres, P. (1966), in Quantum Fluids, ed. by D.F. Brewer (North-Holland, Amsterdam), p. 1. Nozieres, P. (1983), Quantum Liquids and Solids, lecture notes, College de France. Nozieres, P. and Pines, D. (1964, 1990), Theory of Quantum Liquids, Vol. II: Superfluid Bose Liquids (Addison-Wesley, Redwood City, Calif.). O'Connor, J.T., Palin, C.J. and Vinen, WF. (1975), J. Phys. C: Solid State Phys. 8, 101. Ohbayashi, K. (1989), in Elementary Excitations in Quantum Fluids, ed. by K. Ohbayashi and M. Watabe (Springer-Verlag, Berlin), p. 32.
References
291
Ohbayashi, K. (1991), in Excitations in 2D and 3D Quantum Fluids, ed. by A.F.G. Wyatt and H.J.L. Lauter (Plenum, New York), p. 77. Ohbayashi, K., Udagawa, K., Yamashita, H., Watabe, M. and Ogita, N. (1990), Physica B165-166, 485. Osgood, E.B., Kitchens, T.A., Shirane, G. and Minkiewicz, V.J. (1972), Phys. Rev. A5, 1537. Palevsky, H., Otnes, K., Larsson, K.E., Pauli, R. and Stedman, R. (1957), Phys. Rev. 108, 1346. Parry, W.E. and ter Haar, D. (1962), Ann. Phys. (New York) 19, 496. Payne, S.H. and Griffin, A. (1985), Phys. Rev. B32, 7199. Penrose, O. (1951), Phil. Mag. 42, 1373. Penrose, O. (1958), in Proceedings of the International Conference on Low Temperature Physics, ed. by J.R. Dillinger (University of Wisconsin Press, Madison), p. 117. Penrose, O. and Onsager, L. (1956), Phys. Rev. 104, 576. Pike, E.R. (1972), Journ. de Physique 33, Cl, 25. Pike, E.R., Vaughan, J.M. and Vinen, W.F. (1970), J. Phys. C: Sol. State Phys. 3, L40. Pines, D. (1963), in Liquid Helium, ed. by G. Careri (Academic Press, New York), p. 147. Pines, D. (1965), in Low Temperature Physics - LT9, ed. by J.G. Daunt, D.O. Edwards, F.J. Milford and M. Yaqub (Plenum, New York), p. 61. Pines, D. (1966), in Quantum Fluids, ed. by D.F. Brewer (North-Holland, Amsterdam), p. 257. Pines, D. (1985), in Highlights of Condensed Matter Theory, ed. by F. Bassani, F. Fumi and M.P Tosi (North-Holland, New York), p. 580. Pines, D. (1987), Can. Journ. Phys. 65, 1357. Pines, D. (1989), Physics Today 42, 61. Pines, D. and Nozieres, P. (1966), Theory of Quantum Liquids, Vol. I: Normal Fermi Liquids (Addison-Wesley, Redwood City, Calif.). Pitaevskii, L.P. (1959), Sov. Phys. - JETP 9, 830. Pitaevskii, L.P. (1987), Sov. Phys. - JETP Lett. 45, 185. Pitaevskii, L.P. and Fomin, LA. (1974), Sov. Phys. - JETP 38, 1257. Pitaevskii, L.P. and Stringari, S. (1991), Journ. Low Temp. Phys. 85, 377. Placzek, G. (1952), Phys. Rev. 86, 337. Platzman, P. and Tzoar, N. (1965), Phys. Rev. 139, A410. Pollock, E.L. and Ceperley, D.M. (1987), Phys. Rev. B36, 8343. Popov, V.N. (1965), Sov. Phys. - JETP 20, 1185. Popov, V.N. (1983), Functional Integrals in Quantum Field Theory and Statistical Physics (Reidel, Dordrecht). Popov, V.N. (1987), Functional Integrals and Collective Excitations (Cambridge University Press). Popov, V.N. and Serendniakov, A.V. (1979), Sov. Phys. - JETP 50, 193. Puff, R.D. (1965), Phys. Rev. 137, 17406. Puff, R.D. and Tenn, J.S. (1970), Phys. Rev. Al, 125. Rahman, A., Singwi, K.S. and Sjolander, A. (1962), Phys. Rev. 126, 986. Reatto, L. and Chester, G.V (1967), Phys. Rev. 155, 88. Rickayzen, G. (1980), Green's Functions and Condensed Matter (Academic Press, New York). Rinat, A.S. (1989), Phys. Rev. B41, 4247.
292
References
Rockwell, D.A., Benjamin, R.F. and Greytak, TJ. (1975), Journ. Low Temp. Phys. 18, 389. Ruvalds, J. (1978), in Quantum Liquids, ed. by J. Ruvalds and T. Regge (North-Holland, New York), p. 263. Ruvalds, J. and Zawadowski, A. (1970), Phys. Rev. Lett. 25, 233. Scherm, R., Dianoux, A.J., Fak, B., Guckelsberger, K., Korfer, M. and Stirling, W.G. (1989), Physica B156 & 157, 311. Scherm, R., Guckelsberger, K., Fak, B., Skold, K., Dianoux, A.J., Godfrin, H. and Stirling, W.G. (1987), Phys. Rev. Lett. 59, 217. Schrieffer, J.R. (1964), Theory of Superconductivity (W.A. Benjamin, Reading, Mass.). Sears, V.F. (1969), Phys. Rev. 185, 200. Sears, V.F. (1981), Can J. Phys. 59, 555. Sears, V.F. (1985), Can J. Phys. 63, 68. Sears, V.F., Svensson, E.C., Martel, P. and Woods, A.D.B. (1982), Phys. Rev. Lett. 49, 279. Senger, G., Ristig, M.L., Campbell, C.E. and Clark, J.W. (1992), Ann. Phys. (N.Y.) 218, 160. Silver, R.N. (1988), Phys. Rev. B37, 3794. Silver, R.N. (1989), Phys. Rev. B38, 2283. Silver, R.N. and Sokol, P.E., eds. (1989), Momentum Distributions (Plenum, New York). Skold, K. and Pelizzari, C.A. (1980), Phil. Trans, of the Roy. Soc. of London B290, 605. Skold, K., Pelizzari, C.A., Kleb, R. and Ostrowski, G.E. (1976), Phys. Rev. Letters 37, 842. Smith, A.J., Cowley, R.A., Woods, A.D.B., Stirling, W.G. and Martel, P. (1977), J. Phys. C: Solid State Phys. 10, 543. Snow, WM. (1990), Ph.D. thesis, Harvard University, unpublished. Snow, WM., Wang, Y. and Sokol, P.E. (1992), Europhys. Lett. 19, 403. Sokol, P.E. (1987), Can. J. Phys. 65, 1393. Sokol, P.E. and Snow, WM. (1991), in Excitations in 2D and 3D Quantum Fluids, ed. by A.F.G. Wyatt and H.J. Lauter (Plenum, New York), p. 47. Sokol, P.E., Sosnick, T.R. and Snow, WM. (1989), in Momentum Distributions, ed. by R.N. Silver and P.E. Sokol (Plenum, New York), p. 139. Sosnick, T.R., Snow, WM., Silver, R.N. and Sokol, P.E. (1991), Phys. Rev. B43, 216. Sosnick, T.R., Snow, WM. and Sokol, P.E. (1990), Phys. Rev. B41, 11185. Stephen, M.J. (1969), Phys. Rev. 187, 279. Stephen, M.J. (1976), in The Physics of Liquid and Solid Helium, ed. by K.H. Benneman and J.B. Ketterson (Wiley, New York) Vol. 1, p. 307. Stirling, W.G. (1983), in 75th Jubilee Conference on Helium-4, ed. J.G.M. Armitage (World Scientific, Singapore), p. 109. Stirling, W.G. (1985), in Proc. 2nd International Conference on Phonon Physics, ed. by J. Kollar, N. Kroo, N. Menyhard, and T. Siklos (World Scientific, Singapore), p. 829. Stirling, W.G. (1991), in Excitations in 2D and 3D Quantum Fluids, ed. by A.F.G. Wyatt and H.J. Lauter (Plenum, New York), p. 25. Stirling, W.G. and Glyde, H.R. (1990), Phys. Rev. B41, 4224. Stirling, W.G., Talbot, E.F., Tanatar, B. and Glyde, H.R. (1988), Journ. Low Temp. Phys. 73, 33.
References
293
Straley, J.R (1972), Phys. Rev. A5, 338. Stringari, S. (1991), Journ. Low Temp. Phys. 84, 279. Stringari, S. (1992), Phys. Rev. B46, 2974. Svensson, E.C. (1989), in Elementary Excitations in Quantum Fluids, ed. by K. Ohbayashi and M. Watabe (Springer-Verlag, Heidelberg), p. 59. Svensson, E.C. (1991), in Excitations in 2D and 3D Quantum Fluids, ed. by A.F.G. Wyatt and H.J. Lauter (Plenum, New York), p. 59. Svensson, E.C, Martel, P., Sears, V.F. and Woods, A.D.B. (1976), Can. J. Phys. 54, 2178. Svensson, E.C, Martel, P. and Woods, A.D.B. (1975), Phys. Lett. 55A, 151. Svensson, E.C, Sears, V.F., Woods, A.D.B. and Martel, P. (1980), Phys. Rev. B21, 3638. Svensson, E.C and Tennant, D.C (1987), Japan Journ. Applied Phys. 26, (Supp. 26-3), 31. Szepfalusy, P. (1965), Acta Phys. Hungarica 19, 109. Szepfalusy, P. and Kondor, I. (1974), Ann. Phys. (New York) 82, 1. Szprynger, A. and Lucke, M. (1985), Phys. Rev. B32, 4442. Talbot, E.F. (1983), Ph.D. Thesis, University of Toronto, unpublished. Talbot, E., Glyde, H.R., Stirling, WG. and Svensson, E.C. (1988), Phys. Rev. B38, 11229. Talbot, E. and Griffin, A. (1983), Ann. Phys. (New York) 151, 71. Talbot, E. and Griffin, A. (1984a), Phys. Rev. B29, 2531. Talbot, E. and Griffin, A. (1984b), Phys. Rev. B29, 3952. Talbot, E. and Griffin, A. (1984c), Journ. Low Temp. Phys. 56, 141. Tanatar, B , Talbot, E.F. and Glyde, H.R. (1987), Phys. Rev. B36, 8376. Tarvin, J.A, Vidal, F. and Greytak, T.J. (1977), Phys. Rev. B15, 4193. Tisza, L. (1938), Nature 141, 913. Tserkovnikov, Yu. A. (1965), Sov. Phys. - Doklady 9, 1095. van Hove, L. (1954), Phys. Rev. 95, 249. Vinen, W E (1969), in Superconductivity, ed. by R.D. Parks (M. Dekker, New York) Vol. II, p. 1167. Vinen, W.F. (1971), J. Phys. C: Solid State Physics 4, L287. Vollhardt, O. and Wolfle, P. (1990), The Superfluid Phases of Helium 3 (Taylor and Francis, London). Wagner, H. (1966), Z. Physik 195, 273. Wang, Y, Sosnick, T.R. and Sokol, RE. (1992), to be published. Weichman, RB. (1988), Phys. Rev. B38, 8739. Werthamer, N.R. (1969), Amer. Journ. Phys. 37, 373. Werthamer, N.R. (1972), Phys. Rev. Lett. 28, 1102. Werthamer, N.R. (1973), Phys. Rev. A7, 254. West, G.B. (1975), Phys. Rep. 18C, 263. Whitlock, RA. and Panoff, R.M. (1987), Can. J. Phys. 65, 1409. Wilks, J. (1967), The Properties of Liquid and Solid Helium (Clarendon, Oxford). Wong, VK. (1977), Phys. Lett. 61A, 455. Wong, V.K. (1979), Phys. Lett. 73A, 398. Wong, V.K. and Gould, H. (1974), Ann. Phys. (New York) 83, 252. Wong, V.K. and Gould, H. (1976), Phys. Rev. B14, 3961. Woo, C.W. (1976), in The Physics of Liquid and Solid Helium, ed. by J.B. Ketterson and K.H. Bennemann (Wiley, New York) Vol. I, p. 349. Woods, A.D.B. (1965a), in Inelastic Scattering of Neutrons (IAEA, Vienna), p. 191.
294
References
Woods, A.D.B. (1965b), Phys. Rev. Lett. 14, 355. Woods, A.D.B. and Cowley, R.A. (1973), Rep. Prog. Phys. 36, 1135. Woods, A.D.B., Hilton, P.A, Scherm, R. and Stirling, W.G. (1977), J. Phys. C: Solid State Physics 10, L45. Woods, A.D.B and Sears, V.F. (1977), Phys. Rev. Lett. 39, 415. Woods, A.D.B. and Svensson, E.C. (1978), Phys. Rev. Lett. 41, 974. Woods, A.D.B., Svensson, E.C. and Martel, P. (1972), in Neutron Inelastic Scattering (IAEA, Vienna), p. 359. Woods, A.D.B., Svensson, E.C. and Martel, P. (1975), in Low Temperature Physics - LT14, ed. by M. Krusius and M. Vuorio (North-Holland, Amsterdam), Vol. 1, p. 187. Woods, A.D.B., Svensson, E.C. and Martel, P. (1976), Phys. Lett. 57 A, 439. Woods, A.D.B., Svensson, E.C. and Martel, P. (1978), Can. J. Phys. 56, 302. Yang, C.N. (1962), Rev. Mod. Phys. 34, 694. Yarnell, J.L., Arnold, G.P., Bendt, P.J. and Kerr, E.C. (1959), Phys. Rev. 113, 1379. Yoshida, F. and Takeno, S. (1987), Prog. Theor. Phys. (Kyoto) 77, 864. Yoshida, F. and Takeno, S. (1989), Phys. Reports C173, 301. Zawadowski, A. (1978), in Quantum Liquids, ed. by J. Ruvalds and T. Regge (North-Holland, New York), p. 293. Zawadowski, A., Ruvalds, J. and Solana, J. (1972), Phys. Rev. A5, 399.
Author index
Abrikosov, A.A. 233, 241 Ahlers, G. 69 Aldrich III, C.H. 209, 218-24, 230, 278, 281, 282 Allen, J.F. 2 Ambegaokar, V. 24, 204, 257, 264 Andersen, K.H. 6, 12, 23, 38, 154, 157-60, 162-3, 166, 168-70, 173-6, 180, 185 Anderson, RW. 1, 17, 51, 53 Arnold, G.P. 5 Arnowitt, R. 60, 116
Conway, J. 24, 204, 257, 264 Coopersmith, M.H. 59 Copley, J.R.D. 162, 225 Cowan, R.D. 5 Cowley, R.A. 3, 7, 13, 33-5, 41, 45, 78, 134, 156-61, 175, 204, 216, 246-7, 260, 264 Cummings, F.W. 69
Bartley, D.L. 199, 280-2 Baym, G. 24, 60, 70, 128, 204, 235, 257, 264, 278 Beck, H. 263 Bedell, K. 180, 236-9, 242, 275 Beliaev, ST. 17, 22, 51, 53, 60, 116 Bendt, RJ. 5 Benjamin, R.F. 279 Bijl, A. 9, 209 Blasdell, R.C. 83, 85 Bogoliubov, N.N. 2, 8, 16-18, 49-53, 56, 72, 130, 134, 271 Boon, J.P. 228 Brown, G.V. 59 Brueckner, K.A. 59 Buyers, WJ.L. 162 Caroli, C. 284 Campbell, C.E. 10, 210, 214 Carraro, C. 45, 83 Ceperley, D.M. 15-18, 48, 69, 81-5, 88, 128, 276 Chester, G.V. 6, 17, 72, 210, 217, 2 7 3 ^ Cheung, T.H. 19, 60, 93, 100, 107, 116, 140, 233-4, 272 Clark, J.W. 74,210-11,218 Clements, B.E. 210, 214 Cohen, M. 5, 10, 116,208,211
Dahlborg, U. 162 Dianoux, AJ. 6, 12, 23, 38-9, 154, 157-62, 166-70, 173, 175, 185, 278 Dietrich, O.W. 168 Donnelly, R.J. 21 Dorfman, J.R. 133 Dzugutov, M. 162 Dzyaloshinskii, I.E. 234, 241 Einstein, A. 48 Fak, B. 6, 12, 23, 38-9, 154, 157-63, 166-70, 173-6, 180, 185, 278 Family, F. 199 Feenberg, E. 210, 212, 261 Fein, A.E. 260 Ferrell, R.A. 72 Fetter, A.L. 32, 47, 55, 58-9, 62, 73, 131, 196, 249 Feynman, R.P. 5-6, 8-11, 15, 20, 48, 50, 116,208-11,273-5,283 Fomin, I. 239, 275 Forster, D. 51, 134, 138 Fredkin, D.R. 262 Fukushima, K. 243 Gavoret, J. 18, 22, 51, 54, 64, 71-2, 92-3, 101, 105, 116-30, 141, 147-8, 186, 208, 233, 271-2 Gay, C. 109, 111, 133, 136 Gersch, H.A. 43, 45, 78, 83 Gillis, N.S. 260-2
295
296
Author index
Ginzburg, V.L. 135 Giorgini, S. 72, 148, 199 Girardeau, M. 60, 116 Glyde, H.R. 11, 20-3, 33-8, 69, 79, 85, 88, 150, 153-62, 164, 168, 171, 175-7, 180, 183-94, 200-1, 244, 252, 257, 260, 265-6, 270-5, 277-8 Godfrin, H. 6, 12, 23, 38, 154, 157-8, 161-3, 169-70, 173 Goldman, V.V. 265 Gor'kov, L.P. 234, 241 Gotze, W. 225-8, 247 Gould, H. 19, 23, 66, 93-105, 110-16, 125, 151, 195-9, 217, 266 Graf, E.H. 168 Greytak, T.J. 109, 127, 136-8, 140, 231, 252, 256, 279 Griffin, A. 19-23, 60-4, 69-72, 88-9, 93, 100-2, 104, 107-16, 129-33, 136-40, 144, 146, 149-50, 153, 178, 180-4, 196-7, 200-7, 224, 233-4, 242, 248, 270-81 Guckelsberger, K. 160-1, 175, 278
Kleppner, D. 109 Koehler, T.R. 260-1, 268 Kondor, I. 19, 73, 93, 103, 107, 111, 150, 272 Koonin, S.E. 45, 83 Korfer, M. 161, 175, 278 Korth, M.S. 254 Kwok, RC. 260-3
Halley, J.W. 249, 254 Hansen, J.R 26, 162, 225, 228 Hading, O.K. 78 Henshaw, D.G. 5, 275 Herwig, K.W. 83, 85 Hess, D.W. 218, 221, 283 Hilton, RA. 5, 161, 256 Hohenberg, RC. 14, 18, 45, 50-1, 59-60, 64, 68, 72, 100, 110, 125, 128-30, 135-7, 140, 144, 148-50, 181, 195, 271 Homer, H. 257-68 Hsu, W. 218, 230, 278-82 Huang, K. 17, 50, 130 Huang, C.G. 168 Hugenholtz, N. 18, 51-3, 57-9, 62, 105, 272 Hyland, G.J. 69
Ma, S.K. 19-21, 51, 66, 93, 97, 110, 125, 133, 151, 272 Mahan, G. 33, 40, 54, 59, 63, 70, 139, 210, 275 Manousakis, E. 10, 74-5, 81-2, 210-15, 247, 254 Maris, H.J. 35, 156, 284 Marshall, W. 25, 27 Martel, P. 27, 33-5, 39, 43, 45, 70, 75, 78-9, 86-8, 134, 157, 170, 172, 175, 200, 243-7, 276 Martin, RC. 18-19, 50-1, 59-60, 64, 72, 100, 110, 125, 128-30, 134-6, 140, 144, 148-50, 262, 271 Massey, W.E. 156 McDonald, I.R. 26, 162, 225, 228 McMahan, A.K. 263 McMillan, W.L. 15, 197 Meisel, M.W. 257 Menyhard, N. 72 Mezei, F. 36, 154, 157, 159 Miller, A. 6, 10, 33, 67, 115-16, 178, 181, 211,216,272-4 Mineev, V.R 202 Minkiewicz, V.J. 259, 265 Misener, A.D. 2 Mook, H.A. 78, 88 Murray, R.L. 256
Iseki, F. 243 Iwamoto, F. 231, 249, 256 Jackson, H.W. 78, 214, 246 Juge, K.J. 248 Kadanoff, L.P. 60, 134, 235 Kalos, M.H. 17 Kapitza, P. 2 Kerr, E.C. 5 Khalatnikov, I.M. 3-4, 8, 22, 35, 127, 132-3, 136, 278-9 Kirkpatrick, T.R. 133, 199 Kitchens, T.A. 259, 265 Kleb, R. 172 Klein, A. 130
Lambert, N. 224 Landau, L.D. 2-8, 16, 35, 273 Larsson, K.E. 5 Lee, M.A. 17 Lee, T.D. 17 Leggett, A.J. 1 Lifshitz, E.M. 3, 8, 21, 28, 30, 49, 55, 72, 117-18, 138,241 London, F. 1, 4-7, 14-17, 48 Lonngi, D.A. 162 Lonngi, P.A. 162 Lovesey, S.W. 25-7, 162, 225 Lucke, M. 225-30, 247, 278, 281-2
Nakajima, S. 249 Nelson, D.R. 132 Nepomnyaschii, A.A. 127, 143-4, 147-8, 233
Author index Nepomnyashchii, Y.A. 127, 143-4, 147-9, 152, 233, 272 Nozieres, P. 1, 6, 10, 14, 18-23, 28, 32-3, 51-4, 64, 68, 71-2, 92-3, 100-1, 105, 115-30, 133, 141-2, 147-52, 178, 181, 186, 206, 208, 211, 216, 233, 239, 271-4 O'Connor, J.T. 137 Ogita, N. 251-4 Ohbayashi, K. 231, 251-5 Onsager, L. 12, 49-50 Osgood, E.B. 259, 265 Ostrowski, G.E. 172 Otnes, K. 5 Palevsky, H. 5 Palin, C.J. 137 Pandharipande, V.R. 10, 74, 81-2, 210-15, 247, 254 Panoff, R.M. 12, 75, 81-4 Parry, W.E. 59 Passell, L. 168 Pauli, R. 5 Payne, S.H. 107-11, 130, 187 Pelizzari, C.A. 172 Penrose, O. 12, 16, 49-50, 181, 210 Pethick, C.G. 218, 278 Pike, E.R. 134-5 Pines, D. 1, 6, 9, 14, 18-23, 28, 32-3, 51-3, 57-9, 62, 66-7, 105, 110, 115-6, 128, 133, 148, 161, 169, 178-81, 206, 209, 211, 216, 218-24, 226, 230, 236, 238, 242, 272-5, 278-83 Pitaevskii, L.P. 8, 19, 21, 24, 28, 30, 49, 56, 72-3, 104, 117-18, 148, 207, 231-3, 239, 241 Placzek, G. 25, 30 Platzman, P. 14, 45, 68, 77, 181 Pollock, E.L. 15-18, 48, 69, 81-5, 88, 128, 276 Popov, V.N. 108, 127, 132-3, 139, 145 Puff, R.D. 30, 78, 226 Rahman, A. 30-1, 41 Reatto, L. 72, 210 Rickayzen, G. 54, 58 Rinat, A.S. 45, 83 Ristig, M.L. 74, 210-11,218 Robinson, J.E. 280-2 Rockwell, D.A. 279 Rodriguez, L.J. 43, 45, 78, 83 Roulet, B. 284 Rowlands, G. 69 Ruvalds, J. 19, 24, 104, 180, 214, 229, 231-48, 278, 283 Saint-James, D. 284
297
Sawada, K. 59 Scherm, R. 5-6, 12, 23, 38, 78, 154, 157-63, 166-75, 185, 256, 278 Schmidt, H. 72 Schrieffer, J.R. 53 Schwabl, F. 72 Sears, V.F. 27, 33, 35, 39, 42-5, 68, 70, 75, 78-9, 86-8, 162, 175, 245, 276 Serendniakov, A.V. 127 Shirane, G. 259, 265 Silver, R.N. 43-5, 69, 78, 83 Singwi, K.S. 30-1, 41 Sjolander, A. 30-1, 41 Skold, K. 160-1, 172 Smith, A.J. 246-7 Snow, W.M. 14, 44, 78-91 Sokol, RE. 14, 22, 42-4, 69, 78-89, 277 Solana, J. 232, 238-9, 242, 248, 283 Sosnick, T.R. 44, 78-89, 277 Stedman, R. 5 Stephen, M.J. 127, 134, 137, 231, 249, 279 Stirling, W.G. 5-6, 11-12, 23, 34-9, 154-66, 168-77, 185, 200-1, 244-7, 252, 256, 273 Straley, J.R 150 Stringari, S. 23, 72, 148, 199, 207, 217-18, 277, 283 Stunault, A. 6, 12, 23, 38, 154, 157-8, 162-3, 166-70, 173, 185 Svensson, E.C. 23, 27, 33-9, 42, 45, 69-70, 75, 78-9, 86-8, 134-6, 154-7, 160-4, 167-72, 175, 178, 182, 185, 192-4, 200-3, 243-6, 273, 276-8 Szepfalusy, P. 19, 72, 93, 103-4, 107, 111, 150, 272 Szprynger, A. 229, 278, 281-2 Takeno, S. 227-9 Talbot, E.F. 23, 34, 37-8, 62, 95, 100, 102-5, 112, 115, 129-31, 139, 146, 160, 164, 168, 171, 175, 177-8, 185, 197, 200-1, 205-7, 270, 273, 277-81 Tanatar, B. 175, 177 Tarvin, J.A. 137^0 Tenn, J.S. 78 Tennant, D.C. 192 ter Haar, D. 59 Tisza, L. 3-5 Tserkovnikov, Yu. A. 107, 110 Tzoar, N. 77 Udagawa, K. 251-3 Usmani, Q.N. 74, 81-2 van Hove, L. 9, 25-6 Vaughan, J.M. 134 Vidal, F. 137-40
298
Author index
Vinen, W.F. 53, 134, 137 Vollhardt, O. 53, 122, 133 Wagner, H. 23, 70, 195-6 Walecka, J.D. 32, 47, 55, 58-9, 73, 196 Wang, Y. 90, 277 Watabe, M. 251-3 Weichman, P.B. 59, 105, 146-7 Werthamer, N.R. 257, 260-2, 268 West, G.B. 43 Wilkinson, M.K. 78 Whitlock, P.A. 12, 17, 75, 81-4 Wilks, J. 4, 9 Woerner, R.L. 256 Wolfle, P. 53, 122, 133 Wong, V.K. 19, 23, 66, 93-105, 110-16, 125, 151, 195-200, 204, 207, 217, 266, 280-2
Woo, C.W. 10, 19-21, 51, 93, 97, 210, 272 Woods, A.D.B. 3-7, 11-13, 23, 27, 33-6, 39-45, 70, 75, 78-9, 86-8, 104, 134-6, 156-7, 159, 162, 167-72, 175, 178, 194, 200-4, 216, 243-7, 256, 264, 275-6 Yamashita, H. 251-3 Yan, J. 252, 256 Yang, C.N. 17, 49 Yarnell, J.L. 5 Yip, S. 228 Yoshida, F. 227-9 Zawadowski, A. 19, 24, 104, 172, 180, 214, 229, 231-48, 283
Subject index
angular momentum rotating liquid Helium 7, 128 transverse perturbation 128 two-roton bound state 236, 239, 249 anharmonic (quantum) crystal 258-66 ACB sum rule 24, 204 definition of a phonon 260-2 anomalous dispersion 5, 150, 156 anomalous propagator Gn 54-7, 76, 156, 234 atomic-like excitations maxon-roton in superfluid 4 He 9, 88, 179, 183, 273-7 normal liquid 4 He 152, 183, 275, 284 polarization potential theory 221, 224 solid 4 He 268 spurious free-particle peaks 192 atomic motions 9-10, 26, 46, 68, 76, 219, 268, 274, 277 atomic polarizability 134, 249, 279 atomic scattering cross-section 45 attractive interaction (two-roton) 237-9 backflow 10, 114, 211-12, 216, 219, 223, 254 Beliaev matrix single-particle Green's function 55, 233-4 Beliaev second-order approximation 60-1, 116 Bethe-Salpeter integral equation 120-2, 232-3, 237 Bijl-Feynman wavefunction 9, 20, 211-13, 217, 275-6 Bogoliubov approximation 22, 56-61, 114 as description of superfluid 4 He 59, 64 dynamic structure factor 47, 61-3 effect of non-condensate atoms 59, 62, 108, 110 quasiparticle energy 57-8 zero sound interpretation 66, 160
Bogoliubov-Hohenberg-Martin sum rule 72, 130, 148 Bogoliubov prescription (for condensate) 104, 146 Born approximation (neutron scattering) 26, 278 Bose broken symmetry 1, 11, 18-21, 49-51, 64, 69, 104 Bose condensate direct evidence 69, 86-90, 181 as mean field 64-6, 151-2, 184 as reservoir 50, 104, 107 transitions (into or out of) 52-3, 181 Bose distribution function 7, 70-1 identities 32, 63 Bose-Einstein condensation 1, 49 dynamics of condensate atoms 19, 32, 104, 146-7 experimental evidence in liquid 4 He 14-15, 77-90 gas vs. liquid 22, 49 in a solid 257 Monte Carlo results 15, 81-2 relation to superfluidity 4, 15, 50, 128-31 Bose gas (ideal) 16, 47-8 density response 32, 106 Bose gas (weakly interacting), see WIDBG Bose liquid comparison to Bose gas 8, 22, 59 definition of condensate 49 Feynman treatment 9, 275 Bose order parameter 14 phase and amplitude 22, 49-50, 69, 91, 138 fluctuations 72, 132, 138-9, 147, 152 symmetry breaking 50 Bose statistics 10, 178, 274-6 Bose symmetrization, effect on excitation spectrum 10
299
300
Subject index
Feynman-Cohen state 217 momentum distribution 82 pair distribution function 276 bound states (roton) 19, 236-9, 256 Brillouin light scattering 23, 132-5, 137, 279 broadening collisional 44-5, 77, 177 see also final state effects, instrumental resolution broken symmetry (non-Bose) 14, 50, 141 broken-symmetry vertex functions 96-8, 100-3, 118, 123, 143, 177-84 high-frequency limits 197-8 key role 167, 178-9, 183-4, 187-9, 193 low-frequency limits 129, 143, 281 3 He- 4 He mixtures 280-1 Brueckner-Sawada theory 59 bubble diagrams 66, 119, 122, 179 central frequency moments 31, 45, 85 Chalk River neutron data (4He) key papers 33, 78, 154 momentum distribution (analysis) 78-9 multiparticle contribution 36-7, 40, 200, 246-7 phonon-maxon-roton dispersion curve 7, 35, 136 chemical potential 27, 47 determination of condensate fraction 105 Hugenholtz-Pines sum rule 57, 130 classical liquids 161-2, 174, 225 momentum distribution 46, 68-9 coherence factors (Bogoliubov) 57, 63, 76, 181, 235, 242 collisional broadening 44-5, 77, 177 collision-dominated, see hydrodynamic region collisionless regime 132, 149 relation to hydrodynamic region 23, 152, 263 commutation relations 104 compressibility sum rule 31, 143-5, 206, 225-6 compressional sound velocity 61, 71, 111, 150 Compton profile 43-4, 83-7 condensate atoms 51-2, 104 dynamics 146 condensate current 131, 139 condensate fraction ho appearance in sum rules 129-30, 196-8 effect on momentum distribution 69-75, 87 extraction from neutron data 14, 46, 86-91
magnitude and temperature dependence 15, 67, 90-1 Monte Carlo calculations 15-17, 81-4, 86, 88, 90-1 relation to pair-distribution 69 at superfluid-solid transition 91 variational calculations 75, 81-2 in 3 He- 4 He mixtures 91, 277 condensate-induced changes momentum distribution 69-75, 88 weight of maxon-roton in S(Q, co) 23, 69, 178-81, 189, 193 see also dielectric formalism, Glyde-Griffin, hybridization condensate wavefunction 49 conservation laws 60, 99, 104, 125 conserving approximations 60, 107, 116 continuity equation 50, 99, 104, 114, 125, 128, 131 continuum particle-hole (density) 32, 170-3, 240 particle-particle (pair) 63, 180, 187, 240, 242-3 thermally excited rotons 149, 179, 251 two-phonon 63, 115 Cooper pairs 1, 53, 133 correlated-basis-functions (CBF) 209-15 Feynman-Cohen basis 10, 74, 211-12 physical meaning 212, 217 solid 4 He 261, 268 role of condensate 20, 74 correlation (response) functions 95 current-current 95, 98, 114, 125, 192, 206 current-single particle 95 density-current 95, 98 density-density 29, 61, 64, 93, 101, 123 density-single-particle 95, 102, 198 four-point 234 3 He- 4 He mixtures 279-81 order parameter 146 single-particle 28, 54, 139 superfluid velocity 139, 144 correlation length 138 critical mode see Goldstone mode, second sound critical point (gas-liquid), 2 critical region (near 7^), 21 Bose order parameter 14, 91 correlation functions 69, 137-8 critical velocity 3, 16 cross-over phenomena collective and single-particle 20, 193 in 3 He- 4 He mixtures 230, 282 first and zero sound 71, 134, 136 zero sound and incoherent p-h excitations 170-5
Subject index cross-section 4 He- 4 He scattering 45 neutron- 4 He scattering 25-6 neutron- 3 He scattering 278 Raman scattering 249 current density operator 94, 125 Debye-Waller factor 264 analogue of Bose order parameter 266 decay of quasiparticles (stability) 19, 22, 156-7 deep-inelastic scattering 43, 199 density fluctuations (3He) 160-2, 172, 175, 275 density fluctuations (superfluid 4 He) as eigenstates 10, 211 coupling to single-particle excitations 20, 51, 100, 123, 193, 271 coupling to two-particle excitations 194, 240, 271 density matrix ODLRO 49 single-particle 49, 211, 218 two-particle 218 density operator 9, 26 second quantized 27, 51 density response function (Bose) 29, 93, 94, 101 Bogoliubov approximation 62 differences above and below Tx 102, 148-52 ideal gas approximation 32 RPA 106 relation to S(Q,co) 29, 61, 124 sum rules 30-1, 196-7 depletion of condensate 50, 143 Bose gas 73 superfluid 4 He 110 detailed-balance factor 30-2, 170, 199 diagrammatic perturbation theory (Bosons) key papers 18-19, 51, 93 treatment of condensate 51-2, 104-5 see also dielectric formalism, Gavoret-Nozieres diagrams 93-103 bubble 62, 119, 122 irreducible 94 ladder 58-9 one-loop 114 proper 94, 97-8, 129 regular 94, 98, 118 self-energy 57-8, 97-8 vertex function 96-8 zero-loop 105-7 dielectric formalism 21-2, 64, 92-105 condensate component 102-3, 143
301
dielectric functions 93, 96, 100, 102, 109, 115, 124-6, 184-9 Feynman diagrams 93, 95 generalized Ward identities 99, 1 0 3 ^ , 281 3 He- 4 He mixtures 24, 277-82 key papers, 19, 93 normal (non-condensate) contribution 97-8, 102-3, 115, 143, 149, 151, 184-5 relation to mean field 64-6, 279-80 summaries 92, 103, 123-6 two-component expressions 64, 96-8, 101-3, 123, 144, 150-2, 177-83 displacement field, see phonon Green's function divergences Popov treatment 139, 145 see also infrared divergences Doppler-broadening 46, 67, 181 oscillations 175-7 dynamic structure factor in liquids (general) 26, 124, 278-9 eigenstate representation 27 general properties 29-31 impulse-approximation 44, 67 sum rules 30-1 symmetric and antisymmetric components 78 dynamic structure factor in liquid 4 He 13, 33-9, 154-77 correlated-basis-functions 212-13 basic features 63 Feynman approximation 10 multi-particle contribution 13, 24, 33, 179 quasiparticle term 34, 178 relation to liquid 3 He 161-2, 169, 173-5 relation to solid 4 He 24, 266-9 temperature dependence of line shape 11, 158, 163-4, 169-71, 191 two-component forms 33, 64, 123, 177-9 two-fluid region 135 see also impulse approximation (high Q) dynamic structure factor in solid 4 He ACB sum rule 204, 266 Ambegaokar-Conway-Baym (ACB) formula 264 comparison to superfluid 4 He 12, 24, 258-9, 266-8 interference contribution 13, 264, 267 neutron scattering data 263-5 two-phonon contribution 244, 267 Dyson-Beliaev equations of motion 55-7, 99, 103 effective fields, 64, 219-20
302
Subject index
effective mass 3 He- 4 He mixtures 282 liquid 3 He 161 polarization potential theory 220, 224 energy gap 9, 60, 275 energy resolution, see instrumental resolution expansion parameter (gas) 58 Fermi liquids 3 He- 4 He mixtures 281-2 microscopic 118, 142, 224 normal Fermi liquid (Landau) 8, 218 particle-hole continuum 28, 32, 162 quasiparticle vs. zero sound 53 zero sound 160-2 Fermi wavevector ( 3 He- 4 He) 282 Feynman-Bijl excited state 9-10, 211 Feynman-Cohen excited state backflow 10,211,216 momentum space representation 211 relation to two-phonon states 10, 212 Feynman wavefunction relation to single-particle excitation 217, 274 role of Bose condensate 275-7 as single-mode approximation to
S(Q,w)275
variational aspects 9-10, 211 field-theoretic approach for Bose fluids field operators 49 key papers 18-19, 51, 93 history 17-20, 93, 272 field-theoretic approach for quantum solids definition of phonons 260, 262, 268 two-component form for S(Q,co) 264 final state effects (FSE) 45, 68 lifetime broadening 77 neutron scattering 46, 78, 80-6 normal liquid 4 He 46, 68, 78, 83 non-Lorentzian behaviour 45, 78, 85 solid 4 He 46-8 theories 45, 78, 83 finite-size effects 15, 74 finite-temperature perturbation theory 51, 55 first sound 5, 8, 135, 140, 150 fluctuations of order parameter, see Bose order parameter four-point correlation function 249-50 backflow 254 non-interacting approximation 250, 253 frequency moments compressibility sum rule 31, 143-5, 206 /-sum rules 30, 115, 195-207 third-moment sum rules 30, 226, 283 functional differentiation
conserving approximations 60 generation of Gi from G\ 60, 116 gapless spectrum (phonon) 57, 60, 147, 151, 184-5 gauge broken-symmetry 50, 53, 146 associated phonon mode 184 Gaussian approximation impulse approximation for <S(Q, co) 42 momentum distribution 43, 46, 68, 73, 85-86 Gavoret-Nozieres analysis (T = 0) 18, 23, 61, 105, 116-22, 141-3,233 Bethe-Salpeter equation 120, 233 common phonon pole 18, 61, 143 relation to dielectric formalism 121-2, 125 relation to two-fluid model 144-5 two-component expressions 64, 101, 143 Gavoret-Nozieres-Pines (GNP) scenario role of "normal fluid" component 149-52 summaries 125-6, 148-9, 272 zero temperature region 141-3 glory oscillations 177 Glyde-Grifrin scenario 178-94 development 2 7 1 ^ experimental evidence 164, 167, 183 key ideas 20, 153, 178, 183, 189, 193, 273 parameterizations 183-94 relation to other approaches 209, 215, 223, 272-3 relation to Ruvalds-Zawadowski 189, 248 Goldstone mode (symmetry-restoring) 14, 17, 50, 108, 147, 151, 184 Green's function Monte Carlo (GFMC), see Monte Carlo ground state wavefunction 9, 50, 210, 261 Hamiltonian (Bose) 52, 58 symmetry-breaking 51-2, 64, 146 hard-sphere Bosons 50, 58, 71, 197 Hartree-Fock approximation (Bose) self-consistent 60, 116, 196 self-energy 57, 113, 196-7 normal fluids 58, 113, 262 Helium atoms, see atomic polarisability, interatomic potential Helium-I vs. Helium-II 2 3 He- 4 He superfluid mixtures 24, 277-82 high-energy region 23 long-wavelength single-particle excitation 9, 151-2, 184-6, 245, 284 maxon 162-7 zero sound 165-7 see also pair excitations
Subject index high-frequency expansions 30, 196-8 tail in S{Q,co) 33, 199-200, 245 high-momentum neutron scattering 67-90 Chalk River analysis 78-9 condensate peak 67-9 impulse approximation 67 Sokol-Sosnick-Snow analysis 80-4 Hugenholtz-Pines theorem 57, 60-1, 71, 130, 147, 151 hybridization (condensate-induced) 178-94, 233, 271 one- and two-particle spectra 19, 24, 178-81, 189, 234-48 single-particle and density fluctuations 20, 23, 64-6, 97, 100, 109, 123, 151, 167-8, 183, 189, 192-3, 229 hydrodynamic region (Bose) 132-41 behaviour of response functions 114, 140-1, 145, 279 sum rules 31, 71, 129-30 two-fluid region (Landau) 23, 135, 139-40 vs. collisionless 132, 152 imaginary-time formulation (Matsubara) 29, 51, 54 impulse approximation 39-46, 67, 75-7, 198 additive corrections 78-9 dynamic structure factor 46, 67, 181 final-state corrections 45, 79, 83 neutron scattering data 14, 44, 77-87 incoherent scattering cross-section 26-31, 41 infrared (low OJ>) divergences 127, 142-8 Institut Laue-Langevin (ILL) neutron data (4He) maxon region 6, 12, 37-9, 163-4 past the roton 176 pressure dependence 165-6, 244 phonon region 11, 154-60 roton region 38, 169-71 instrumental resolution broadening 68, 78-80, 83-5, 154, 181 interaction vertex function 60, 120, 227 interatomic potential in 4 He 70 bare 4 He- 4 He interaction 45 effective interaction in liquid 4 He 177, 220-3, 235, 248, 283 hard-core 58, 71, 197, 199-200 interference contributions anharmonic crystals 201, 264, 267 superfluid 4 He 53, 183, 203, 242 WIDBG 62-4, 115 intermediate scattering function 26, 41 irreducible diagrams 93-103 irreducible response functions
303
two-component expressions 97-8, 123, 183^ Jastrow(-Feenberg) wavefunction 72, 208-10, 260-1 kinetic energy of atoms 73, 108 kinetic equations 122, 133 Kosterlitz-Thouless scenario (2D) 132 ladder diagrams (t-matrix approximation) 58-9, 122, 214, 222, 283 lambda transition (TA) region 8, 14-16, 20-1, 48, 67-9, 92, 102, 107, 113, 124-8, 134, 140-1, 152, 166-7, 178-82, 245, 253, 272-5 Landau-Feynman scenario 12, 20, 61, 272 Landau-Khalatnikov theory 8, 22, 133 Landau-Placzek ratio 134-7, 140, 158 lattice vibrations 258-62 lifetimes (thermal) 132 Lindhard function Bosons 32, 142, 179, 229 Fermions 169, 224, 227, 281-2 linear response theory 64, 146, 219 liquid-gas transition 2, 135 liquid 3 He 53, 77, 142, 160-1, 218, 221 Longitudinal current response function (Bose) 95 continuity equation 99, 125, 128, 192 zero-frequency limit 105, 128-9, 144 longitudinal susceptibility (Bose) 147 long-range order (LRO) 132 long-range spatial correlations Bose-condensed fluid 49, 128-9 relation to phonons 72 variational wavefunctions 210 Lorentzian approximation quasiparticle peak 34, 77, 160 Raman scattering 251 low-density Bose gas finite temperature 108-11 zero temperature 57-61 see also WIDBG low-energy, long-wavelength behaviour (Bose) effect of infrared divergence 145-8 Gavoret-Nozieres 18, 61, 141-3, 148-9 Goldstone mode 50, 151-2 Hugenholtz-Pines theorem 61 hydrodynamic vs. collisionless 150 Popov analysis 139, 145 regular functions 103, 184 sum rules 31, 128-31, 199 two-fluid region 132-40 macroscopic region (Bose), see two-fluid description
304
Subject index
macroscopic wavefunction 49 phase coherence 16, 50 relation to superfluidity 13-16 see also Bose order parameter Matsubara formalism (imaginary frequencies) analytic continuation 55, 61 Bose 29, 55, 147, 281 frequency sums 63, 113, 235 maxon region (intermediate Q) hybridization effects 167, 182-3, 187-90, 193, 203 neutron data 5-7, 163-7 maxon-roton branch hybridization with pair excitations 189 hybridization with phonons 20, 23, 182-9, 193 intrinsic to normal liquid 4 He 20, 126, 178, 183, 193, 273-5, 284 intrinsic to superfluid 4 He 125, 272 weight in S{Q,co) 23, 69, 178, 189, 284 mean-fields 108, 165 due to condensate 64-6, 151, 224 in 3 He- 4 He mixtures 280 zero-loop approximation 107 memory function formalism for S(Q,co) 23, 225-30 in 3 He- 4 He mixtures 229, 278 mode-mode coupling 227, 230, 281-2 momentum current response, see current response momentum distribution function (liquid 4 He) condensate peak 67, 81 data from neutron scattering 14, 43, 46, 77-85 extraction from 5(Q,co), theory 69, 78-9 Gaussian 46, 68, 79, 84-5 high-momentum behaviour 85, 88, 99 low-momentum behaviour 69-75, 79-81, 84, 88-9, 130, 148 Monte Carlo calculations 74, 81-5 non-condensate (normal) part 87-9 normalization 67, 84, 87 normal liquid 4 He 69, 79, 83, 88 pressure dependence 91 relation to single-particle spectral density 70-4 rigorous results 71, 75, 89, 148 temperature dependence 73 two-Gaussian fit 86-90 variational calculations 74, 81-2, 210, 214 Monte Carlo calculations Green's function (GFMC) 74, 81, 84, 86, 90-1, 210
path integral (PIMC) 15-18, 48, 69, 74, 81-5, 90, 128, 276 multiparticle spectrum neutron scattering data 13, 33, 155, 163-4, 167, 171, 244-7 overlap with density fluctuation 221, 244 overlap with single-particle excitations 13, 19, 240, 248 multiphonon see bound states, multiparticle spectrum multiple scattering (t-matrix), see ladder diagrams negative-energy poles in spectral density 57, 63, 72-3, 242 neutron scattering (inelastic) 26 dilute 3 He- 4 He mixtures 278, 281-2 liquid 3 He 160-2 liquid 4 He 33-40, 154-77 role of Bose condensate 19-21, 92, 160, 181, 276 solid 4 He 259, 267 see also dynamic structure factor neutron scattering lengths 26, 278 non-condensate atoms 5, 52, 56, 59-61, 73, 87 density response 32, 62, 115, 123, 149-51, 273 effect on SP spectrum 108 "normal" component in dielectric formalism 100-2, 109, 116, 123, 149-52, 179, 189, 197, 223 normal fluid density as measure of quasiparticle density 3, 5, 8, 110, 150, 191,223 PIMC results 15, 18 quasiparticle formula 7, 15, 131 relation to zero-frequency response functions 15, 18, 110, 128 temperature dependence 4, 8, 18 normal liquid 4 He 150, 268, 275 density response 11, 37, 46, 64, 79, 134, 158, 163-74 number density 28, 50-1 ODLRO 49 one-loop approximation 112-16, 151 in 3 He- 4 He mixtures 283 zero-frequency limit (sum rules) 131 order parameter (Bose) 49, 69 c-number approximation 104, 146 dynamics 146 macroscopic wavefunction 15-16, 49 see also Bose condensation, phase fluctuations
Subject index order parameter (non-Bose) 14, 50, 53, 109, 146 pair-distribution function 27, 45, 210 effect of Bose statistics 276 relation to condensate 69 pair excitation spectrum (Q j=0) 19, 180, 234-9, 243-7, 271 Bogoliubov approximation 63 coupling into S(Q,co) 212-14, 239 coupling to single-particle spectrum 19, 23, 240-8 temperature dependence 179-80, 213-15, 245 two-particle Green's functions 116-22, 271 particle-hole excitations (non-interacting) in 3 He- 4 He mixtures 281-2 liquid 3 He 28, 224 particle-hole excitations (liquid 4 He) 28 Doppler-broadened peak 168, 174-7 incoherent 32, 169, 170-5 Lindhard function (gas) 32, 179 see also zero sound particle-hole propagator, see density response function path integral Monte Carlo (PIMC), see Monte Carlo phase coherence (Bose) 16, 50 phase diagram for 4 He 2 phase fluctuations in superfluid 4 He 50, 72, 139, 146-8 phase shift (s-wave) 58 phonons in Bose-condensed fluids (theory) 141-52 WIDBG (T = 0) 57-61 W I D B G ( T ^ O ) 108-11 phonon Green's function (crystal) 260 analogue of single-particle Green's function 257, 260, 268 relation to S{Q,co) 264, 268 phonons in quantum crystals 258-63 collective vs. single-particle 262, 268 effect of short-range correlations (SRC) 260-1 as pole of displacement correlation function 24, 261, 268 relation to S{Q,co) 263-6 phonon scattering processes (superfluid 4 He) 8, 156-7 phonons in superfluid 4 He (experiment) 7 vs. maxon-roton branch 20, 153, 167 vs. normal liquid 4 He 11, 134, 158, 160 region around T;t 157-9 relation to zero sound 160-2 temperature independence (above 1 K) 11, 150, 157, 159, 184
305
widths 22, 155-7 see also anomalous dispersion Pitaevskii plateau 19, 180, 241 polarization potential theory 23, 218-24, 281-2 pressure dependence (liquid 4 He) 137, 164-8, 247 condensate fraction 91 effective polarization potentials 221-2 proper (and improper) diagrams 94, 97-8, 129 quantum evaporation 284 quantum hydrodynamics 8, 22, 216 Popov derivation 139, 145 quantum kinetic equations 122, 133 quasiparticle dispersion relation in liquid 4 He 7, 35 impulse-approximation region 174-7 maxon region 162-7 past the roton region 174, 180 phonon region 11, 158-60 roton region 168-74 single branch vs. two branches 183-93, 273 variational calculations 212-14 quasiparticles 33 Bogoliubov model 57 dilute 3 He- 4 He mixtures 281-2 interpretation 8, 12, 20, 148-53, 193, 272-6 normal liquid 4 He 12, 20, 277 see also phonons in quantum crystals, single-particle Green's function Raman light scattering 249 data analysis near TA 251-3 evidence for two-roton bound state 256 non-interacting pair approximation 250 s- and d-wave components, 236-7, 249 random phase approximation (RPA) 106-7, 218-19 Rayleigh central peak 134 recoiling atom energy 31-2 impulse-approximation 41-2, 67-8, 78, 175 regular diagrams 94-8, 114, 118, 280 regular single-particle Green's function G^ 96, 103, 124, 178, 183, 223, 280 long-wavelength 130, 151-2, 184 multiparticle spectrum induced by condensate 152 sum rules 130, 196-8 regularity assumption 116, 143—4 renormalization group 21, 59
306
Subject index
restricted ensemble 49-50 see also Bose broken-symmetry rotating superfluid 4 He 4, 7, 128 roton-liquid theory 275 rotons (experiment) 5 condensate-induced weight in S(Q,co) 178-82, 240-1, 273-4 existence above Tx 168, 178, 275, 284 lineshape 168-71, 284 overlap with p-h excitations 169, 172-3, 179 overlap with two-roton spectrum 19, 179-80 Ruvalds-Zawadowski theory 19, 104, 180, 214, 232^8 screened response functions in 3 He- 4 He mixtures 280-2 mean-field analysis 65 polarization potential theory 219-23 second sound Bose gas 133, 136 in 3 He- 4 He mixtures 279 quantum solid 262 in superfluid 4 He 5, 8, 133-40 weight in single-particle Green's functions 139-41 weight in S{Q,co) 135-7 self-consistent fields, see mean-fields self-consistent harmonic (SCH) phonon theory 260-1 self-consistent ^-matrix 59 self-energies (Bose fluids) 56-60 irreducible 97-8, 242 one-loop 112-15 relation to chemical potential 61, 130, 196 zero-frequency limit 130, 142-3, 147 zero-loop 106, 110, 114-15 shielded potential approximation (SPA) 65-6, 105-11, 113 short-range spatial correlations (SRC) liquid 4 He 45, 210, 219 solid 4 He 260-1, 268 short-time behaviour 30, 229 single-particle density matrix 49, 74, 211, 218 single-particle excitations in quantum crystals neutron data at high Q 259 relation to phonons 262, 268 single-particle excitations (spectrum) coupling into density 51 coupling into S(Q,w) 18-22, 60, 64, 178-83, 246, 284 at finite T 108, 110, 148-52
high-energy modes at long wavelengths 151, 184 single-particle Green's functions (Bose) 53-6 Bogoliubov approximation 57-9 in conserving approximations 60 low-frequency behaviour 18, 61, 148, 184 physical interpretation 148, 272 relation to S(Q,co) 18, 22, 29, 148-50, 177, 242-8, 272 spectral representation 28, 139 single-particle (SP) operator 217-18 single-particle spectral density (Bose) 28 Bogoliubov approximation 72 hydrodynamic (two-fluid) region 71, 139^0 relation to momentum distribution 70-4 rigorous sum rules 70, 196, 217 singular contributions, see diagrams, Gavoret-Nozieres analysis sound velocity 5 first sound 8, 111, 150 second sound 111, 135, 262 soft modes 109-11 thermodynamic (compressional) 31, 61, 111, 150 zero sound 111, 150, 263 specific heat near T-A 15-16 spin-polarized atomic Hydrogen 109 static structure factor S {Q) 27, 41, 211, 214, 226-7 quasiparticle contribution Z(Q) 34, 41, 181-2 Stokes (anti-Stokes) component 30, 32 strong-coupling limit 109-11 sum rules for Bose fluids Bogoliubov-Hohenberg-Martin 71, 130 compressibility 31, 145, 206 density-single-particle 198-9, 217 /-sum rule 30, 99, 115, 151, 185-6, 191, 196 Hugenholtz-Pines 57, 60-1, 71, 130, 151 third order frequency-moments 30, 226, 283 Wagner 70, 196, 198 Wong-Gould 197 superconductivity (BCS) 1, 53 supercurrents, relation to condensate 131, 139 superfluid density appearance in sum rules 129-30, 206 Landau two-fluid model 3, 5 relation to condensate 4, 22, 50, 131 relation to current-response functions 15, 128-9 temperature dependence near Tx 91 superfluid Fermi liquids 53, 122, 133
Subject index superfluidity, relation to Bose condensate 48, 50, 129, 131 superfluid-solid phase transition 90 superfluid velocity field 2, 49, 139 susceptibility dynamic 29 static 31, 225 symmetry-breaking perturbation 51-4, 104, 146 symmetry relations broken-symmetry vertex functions 97-8 dynamic structure factor 29-30 single-particle matrix Green's functions 55, 74 two-particle Green's functions 95 Taylor expansions at small Q, co Gavoret-Nozieres results 142-3 Nepomnyaschii-Nepomnyaschii results 143-4, 147-8 regularity assumption 116, 143-^ temperature fluctuations, see Brillouin light scattering, Landau-Lifshitz ratio, second sound thermal scattering processes as background 149 dilute Bose gas 32, 63 superfluid 4 He 142, 150, 179 thermodynamic derivatives ground state energy 61, 105 self-energies 61, 142-4, 196 thermodynamic potential 104-5 thermodynamic properties 4-7, 135-7 three-body correlations 210 Mnatrix, see ladder diagrams transport coefficients 4, 7, 133 transverse response functions 15, 95, 98, 128-9 tricritical point 281, 284 two-component forms for S(Q,co) dielectric formalism 64, 101, 123-6, 177-8, 183, 207, 223, 279-81 Gavoret-Nozieres {T = 0) 64, 143, 148-50 Glyde-Griffin 126, 183-94, 273 hydrodynamic (two-fluid) 135, 140-1 Szepfalusy-Kondor scenario 109-11, 150, 272 Woods-Svensson ansatz 23, 200-3 two-dimensional Bose fluids 132 two-fluid description equations of motion 4, 133, 279 Landau formulation 3, 5 3 He- 4 He mixtures 279 microscopic basis 18, 50, 128-32 Popov derivation 139, 145
307
at T = 0 limit 144-5, 149 vs. collisionless 152 two-particle density matrix 118 two-particle Green's functions (G2) 60, 116-22,232-9 two-particle interaction g4 (pair spectrum) 235, 248 two-phonon terms quantum crystal 261, 264, 267 Bose gas (WIDBG) 63, 115 two-roton spectrum in S(Q, co) 239-48 ultrasonic studies liquid 4 He 134, 155 solid 4 He 262 variational wavefunctions, see wavefunctions (many-body) velocity field (superfluid) 3, 139 vertex corrections 214, 253 viscosities of normal fluid 3, 133—4 vorticity in superfluid 4 He 21, 134 Wagner sum rule 70, 196-8 Ward identities (generalized) continuity equations 99, 104 high-frequency limit 197-8 importance 100, 103, 122, 129 one-loop approximation 112 zero-frequency limit 129-31, 281 zero-loop approximation 105-7 wavefunctions (many-body) 82, 209-14 Bose phase coherence 216 Bose statistics 9, 217 Feynman 9-10, 20 Feynman-Cohen 74, 213 Gaussian 260-1 hard-sphere Bosons 50 spatial correlations 72-4, 260-1 see also correlated-basis-functions (CBF) weak-coupling limit 63, 108, 111 WIDBG (weakly interacting dilute Bose gas) 1-2, 5 16, 20-1, 47, 56, 59, 66, 73, 92, 103-16 Woods-Svensson ansatz 23, 194, 200-3, 206 Y-scaling 42-4 zero-frequency limits 128-31, 143, 147 zero-loop approximation 105-11 zero-momentum atoms 48-50, 104, 273 zero sound in classical fluids 161-2, 183 zero sound in 3 He 160-1, 165, 169
308
Subject index
zero sound (ZS) in Bose fluids coupling to single-particle excitations 124-6, 183-9 damping 159 due to condensate mean field 66, 110, 151, 186, 273 at large Q 165, 168, 178
liquid 4 He (normal) 134, 165-8, 178, 183-5 liquid 4 He (superfluid) 20-2, 158-9, 182-3 normal fluid component 110, 126, 149, 151, 166-7, 185-9 superfluid vs. normal 151-2, 273