cond-mat0005074/btex
1: %\documentstyle[preprint,aps,prl,floats,graphicx,amssymb]{revtex}
2: \documentstyle[aps,prl,floats,graphicx,amssymb]{revtex}
3: \begin{document}
4: \draft
5: \title{Hydrostatic theory of superfluid $^3$He-B
6: \thanks{Associated internet page: 
7: http://boojum.hut.fi/research/theory/btex.html}}
8: 
9: \author{E.V. Thuneberg} 
10: \address{Low Temperature Laboratory, Helsinki University of 
11: Technology, 02150 Espoo, Finland}
12: \date{\today}
13: \maketitle
14: \def\imagu{{\rm i}}
15: 
16: \begin{abstract}  The determination of the texture of the order
17: parameter is important for understanding many experiments in superfluid
18: $^3$He. In addition to reviewing the theory of textures in superfluid
19: $^3$He-B we give several new results, in particular on the surface
20: parameters in the Ginzburg-Landau region and bulk parameters at
21: arbitrary temperature.  Special attention is paid to separate the
22: results that are valid at all temperatures from those which are limited
23: to the Ginzburg-Landau region. We study the validity of a trivial
24: strong-coupling model, where the energy gap of the weak-coupling
25: theory is scaled by a temperature dependent factor. We compare
26: the theory with several experiments. For some
27: quantities the theory seems to work fine and we extract the
28: dipole-dipole interaction parameter from the measurements. 
29: \end{abstract} 
30: 
31: \section{Introduction} 
32: 
33: The superfluid phases of liquid $^3$He show complex behavior, which
34: still can be understood theoretically. Many phenomena have been studied
35: in a pure form in $^3$He, and the knowledge can then be applied
36: to other physical systems. For example, several structures of quantized
37: vorticity have been seen in both A and B phases of $^3$He
38: \cite{LounasT}. The effect of impurities has recently been studied in
39: many laboratories by aerogel immersed in liquid 
40: $^3$He \cite{aerogel}. Recent experiments on the Josephson effect
41: show unexpected behavior \cite{packard2}. Theoretical
42: understanding of all these phenomena requires good quantitative
43: understanding of the basic properties of superfluid $^3$He. This
44: provides the motivation for the present paper. 
45: 
46: The purpose of hydrodynamics is to determine the behavior of a fluid on
47: length and time scales that are long compared to some microscopic
48: lengths and times \cite{LLfluid}. Hydrostatics is a subfield of
49: hydrodynamics. It is restricted to study the equilibrium properties of
50: the fluid. In simple fluids hydrostatics reduces to statements about
51: the pressure variation in the fluid, which either rotates uniformly or
52: is exposed to some external field. The problem becomes more difficult,
53: if the fluid has some broken symmetry. Particular examples of these are
54: liquid crystals and superfluids $^4$He and $^3$He. In both superfluids,
55: the equilibrium mass current belongs to the scope of the hydrostatic
56: theory. The order parameter of $^3$He has also other degrees of
57: freedom. The structure of those, which is often called {\it texture},
58: also has to be incorporated. In $^3$He the hydrostatic theory is
59: limited to length scales that are large in comparison to the superfluid
60: coherence length $\xi_0\approx 10\,$nm. The theory is valid at all
61: temperatures.
62: The hydrostatic theory can still be applied when the motion of
63: the quasiparticles at low temperatures becomes ballistic rather than
64: diffusive.
65: The hydrostatic theory can be generalized to hydrodynamic theory by
66: using conservation laws and adding the transport coefficients. 
67: 
68: Our purpose is to make a systematic presentation of the
69: hydrostatic theory for the B phase of superfluid $^3$He. Part of the
70: reason is that presently the various results are scattered over a large
71: literature. Previous reviews treat hydrostatics only as a side topic
72: and cover only a small part of the subject
73: \cite{Leggett,BC,SRrev,Fetterrev,SVrev,Kharadze,VW}. The existing
74: papers are often unclear whether they treat general temperatures or are
75: restricted to the neighborhood of the superfluid transition
76: temperature $T_{\rm c}$. We note that the hydrostatics of the A phase is
77: better presented in the existing literature
78: \cite{BC,Fetterrev,SVrev,VW} than the B phase considered here. 
79: In addition to reviewing, we present several results that have
80: not been published before. 
81: 
82: We will begin with a general formulation of the hydrostatics of
83: $^3$He-B. Our approach is general enough to allow an external magnetic
84: field and uniform rotation, both in the leading order. We write down an
85: energy functional that consists of bulk terms and boundary conditions.
86: All structures on the length scale of $\xi_0$, such as surface layer or
87: vortex lines, have to be treated as boundary conditions. The theory is
88: found to split into two pieces: one for the superfluid velocity and the
89: other for the texture. The former is identical to the hydrostatics of
90: superfluid $^4$He, whereas the latter can be solved only after the
91: superfluid velocity is determined. 
92: 
93: The coefficients of the hydrostatic energy can either be
94: obtained experimentally or be calculated by some more
95: microscopic theory. The calculation is discussed in sections
96: \ref{s.qc} and \ref{s.gl}. The former considers
97: the quasiclassical theory of $^3$He \cite{SRrev}. The coefficients can
98: be calculated using the weak-coupling quasiclassical theory  at
99: arbitrary temperature. We also
100: discuss a ``trivial strong-coupling''  (TSC) model, where the
101: weak-coupling coefficients are improved by scaling the energy gap. A
102: different approach is studied in section 
103: \ref{s.gl}, where the hydrostatic coefficients are related to the
104: parameters of the phenomenological Ginzburg-Landau (GL) theory at
105: $T\approx T_{\rm c}$. 
106: 
107: Before any quantitative tests of the theory, we still have to
108: determine the parameters that the quasiclassical theory needs as input.
109: This is discussed in Section \ref{s.gd}, where we analyze several
110: experiments. We find that certain quantities are reasonably well fitted
111: using TSC model, but errors of 50\% may occur for other
112: quantities.  
113: 
114: \section{Hydrostatic free energy}\label{s.f}
115: 
116: The superfluidity in a Fermi system arises from formation of Cooper
117: pairs. A macroscopic number of pairs occupies the
118: same pair state in the superfluid state \cite{BCS}. The
119: relative orbital wave function of a pair has p-wave symmetry in $^3$He
120: and the spin state is a triplet
121: \cite{Leggett,WheatleyRev}.  The state of the pairs is thus described by
122: an order parameter, which is a complex
123: $3\times 3$ matrix $A_{\alpha i}$. It gives the projections of the
124: Cooper-pair wave function on the three p-wave orbitals ($p_x$, $p_y$,
125: and $p_z$, index
126: $i$) and on the three spin triplet states
127: ($\vert-\uparrow\uparrow+\downarrow\downarrow\rangle$,
128: ${\rm i}\vert\uparrow\uparrow+\downarrow\downarrow\rangle$, and
129: $\vert\uparrow\downarrow+\downarrow\uparrow\rangle$,
130: index $\alpha$). In
131: unperturbed B phase the order parameter has the form
132: \begin{equation} A_{\alpha j}=\Delta e^{\imagu\phi}R_{\alpha j}
133: \label{e.2.1}\end{equation} with real $\Delta$, $\phi$ and $R_{\alpha
134: j}$. Here the amplitude $\Delta$ has a fixed (temperature and pressure
135: dependent) value and
136: $R_{\alpha i}$ is constrained to be a rotation matrix, i.e.\
137: $R_{\alpha i}R_{\alpha j}=\delta_{ij}$. (Summation over repeated
138: indices is assumed.) The phase
139: $\phi$ and the more detailed form of the spin-orbit rotation matrix
140: $R_{\alpha i}$ are not fixed on the scale of the superfluid
141: condensation energy. These soft variables allow a dissipationless flow
142: of both mass and spin. The order parameter can be interpreted as the
143: wave-function of the center of mass of a pair. Using standard quantum
144: mechanics, we can then define a mass-flow velocity
145: \begin{equation} {\bf v}_{\rm s}={\hbar\over
146: 2m}{\mbox{\boldmath$\nabla$}}\phi,
147: \label{e.2.2}\end{equation}
148:  where $m=5.0097\cdot 10^{-27}$ kg is the mass of a
149: $^3$He atom. In a similar way one can also define a spin-flow velocity
150: \begin{equation} {\bf v}^{\rm spin}_{{\rm
151: s},\alpha}=-{\hbar\over4m}\epsilon_{\alpha\beta\gamma}R_{
152: \beta i}{\mbox{\boldmath$\nabla$}} R_{\gamma i}
153: \equiv {\hbar\over 4m}R_{\alpha i}\epsilon_{ijk}R_{\beta
154: j}{\mbox{\boldmath$\nabla$}} R_{\beta k},
155: \label{e.2.3}\end{equation} 
156:  where $\epsilon_{ijk}$ is the maximally
157: antisymmetric tensor. For example, if the axis of the spin-orbit
158: rotation is constant and parallel to $z$, the pairs with spin states 
159: $\vert\uparrow\uparrow\rangle$, 
160: $\vert\downarrow\downarrow\rangle$, and
161: $\vert\uparrow\downarrow+\downarrow\uparrow\rangle$
162: flow with velocities 
163: ${\bf v}_{\rm s}+ {\bf v}^{\rm spin}_{{\rm s},z}$,
164: ${\bf v}_{\rm s}- {\bf v}^{\rm spin}_{{\rm s},z}$, and 
165: ${\bf v}_{\rm s}$,
166: respectively. The
167: three-dimensional rotation matrices are conveniently parametrized by an
168: angle $\theta$ and an axis
169: $\hat{\bf n}$ of rotation as
170: \begin{equation} R_{ij}(\hat{\bf
171: n},\theta)=\cos\theta\delta_{ij}+(1-\cos\theta)\hat n_i\hat
172: n_j-\sin\theta\epsilon_{ijk}\hat n_k,
173: \label{e.2.4}\end{equation} where $\hat{\bf n}\cdot\hat{\bf n}=1$. We
174: note also the trivial identity
175: $\hat{\bf a}\cdot
176: \tensor{R}\cdot\hat{\bf a}=\cos\theta+(1-\cos\theta)(\hat{\bf
177: a}\cdot\hat{\bf n})^2$, where $\hat{\bf a}$ is an arbitrary unit vector.
178:  
179: The soft variables $\phi$ and $R_{\alpha i}$ are determined by the
180: interaction of the order parameter with various perturbations. The
181: perturbations can be divided into external fields and boundary
182: conditions. Experimentally, the most common field is the magnetic one
183: ${\bf H}$. It would also be straightforward to include the electric
184: field, but we will neglect it here because its coupling is very weak
185: \cite{SEP}. The motion of the $^3$He container can also be treated as
186: an external field. In equilibrium the normal fluid component (velocity
187: ${\bf v}_{\rm n}$) will follow the motion of the container, and the
188: only allowed motions are uniform translation and rotation. The former
189: is automatically taken into account because of Galilean invariance of
190: the theory. The rotation will appear as a field
191: $\nabla\times{\bf v}_{\rm n}$, which equals twice the angular velocity.
192: 
193: It is possible to construct a hydrostatic theory for any magnitude of
194: the external fields. Here we assume the fields are small enough that
195: the order parameter is not distorted strongly from the bulk form
196: (\ref{e.2.1}). Strong perturbations, such as a surface or a vortex
197: core, are treated as boundary conditions. These cause the order
198: parameter to deviate substantially from the bulk form within a length
199: scale of the coherence length
200: $\xi_0\approx 10$ nm, but on longer length scales also they act as weak
201: perturbations.
202:  
203: The degrees of freedom $\phi$ and
204: $R_{\alpha i}$ differ crucially in the following respect. The mass flow
205: can be written as
206: ${\bf j}_{{\rm s}}=\tensor{\rho}_{\rm s}\cdot{\bf v}_{{\rm s}}$, where
207: $\tensor{\rho}_{\rm s}$ is a phenomenological tensor. In the absence of
208: external fields, $\tensor{\rho}_{\rm s}$ must be a scalar because of
209: the isotropy of the unperturbed $A_{\alpha i}$ (\ref{e.2.1}). The mass
210: current has also to be conserved: $\nabla\cdot{\bf j}_{\rm s}=0$. Thus
211: we arrive at the Laplace equation
212: $\nabla^2\phi=0$. Adding the boundary conditions, this completely
213: determines
214: $\phi$. The external fields cause only a small correction to this.
215: Throughout the rest of this paper we neglect the small correction and
216: start out from the assumption that the Laplace equation for
217: $\phi$ is already solved, and thus ${\bf v}_{\rm s}$ is known.
218:  
219: The problem that remains is to determine the rotation matrix $R_{\alpha
220: i}$. What makes this different from $\phi$ is that there exists
221: interaction between the nuclear dipole moments of the $^3$He atoms. It
222: is of the form \cite{LeggettAnn}
223: \begin{equation} F_{\rm D}=\lambda_{\rm D}\int
224: d^3r(R_{ii}R_{jj}+R_{ij}R_{ji})=4\lambda_{\rm D}\int
225: d^3r\cos\theta(1+2\cos\theta).
226: \label{e.dipgen}\end{equation} Although this interaction is weak, it
227: partly removes the degeneracy with respect to the rotation matrix. This
228: means that the spin current is not conserved, but decays on a scale
229: $\xi_{\rm D}\approx 10\ \mu$m. On the same scale, the rotation angle
230: $\theta$ becomes fixed to $\arccos(-1/4)\approx 104^\circ$, which
231: corresponds to the minimum of
232: $F_{\rm D}$ (\ref{e.dipgen}), but the degeneracy with
233: respect to the rotation axis
234: $\hat{\bf n}$ remains. Because no conservation laws exist, the rotation
235: axis $\hat{\bf n}$ is more susceptible to all kinds of perturbations
236: than $\phi$. The subject of the rest of this paper is to study the
237: texture, i.e.\ $\hat{\bf n}({\bf r})$ on a length scale $\gg\xi_{\rm
238: D}$.
239:  
240: We write down the free energy functional that governs the texture. The
241: form of the energy terms is based on symmetry properties alone.  The
242: functional is valid in the limit of low fields and velocities, small
243: gradients of the order parameter and weak coupling between the spin and
244: orbital parts of the order parameter. The last condition is practically
245: always satisfied because the coupling is due to the dipole-dipole
246: interaction (\ref{e.dipgen}), which is small compared to the
247: superfluid condensation energy by factor
248: $10^{-6}(1-T/T_c)^{-1}$. We neglect all constant terms, i.e. terms that
249: do not depend on $\hat{\bf n}$. The leading terms in the expansion can
250: be written as follows
251: \begin{equation} F_{\rm DH}=-a\int d^3r(\hat{\bf n}\cdot{\bf H})^2
252: \label{e.2.5}\end{equation}
253: \begin{equation} F_{\rm DV}=-\lambda_{\rm DV}\int d^3r[\hat{\bf
254: n}\cdot({\bf v}_{\rm s} -{\bf v}_{\rm n})]^2
255: \label{e.2.6}\end{equation}
256: \begin{equation} F_{\rm HV}=-\lambda_{\rm HV}\int d^3r[{\bf
257: H}\cdot\tensor R\cdot ({\bf v}_{{\rm s}}-{\bf v}_{{\rm n}})]^2
258: \label{e.2.7}\end{equation}
259: \begin{equation} F_{\rm HV1}=-\lambda_{\rm HV1}\int d^3r{\bf H}\cdot
260: \tensor R\cdot\mbox{\boldmath$\nabla$}\times {\bf v}_{{\rm n}}
261: \label{e.2.8}\end{equation}
262: \begin{equation} F_{\rm G}=\int d^3r\left[\lambda_{\rm G1}{\partial
263: R_{\alpha i}\over\partial r_i}{\partial R_{\alpha j}\over\partial
264: r_j}+\lambda_{\rm G2} {\partial R_{\alpha j}\over\partial r_i}
265: {\partial R_{\alpha j}\over\partial r_i}\right].
266: \label{e.2.9}\end{equation}  
267:  The dipole-field term $F_{\rm DH}$ is discussed in Refs.\
268: \cite{EBA,SBE}, the gradient term $F_{\rm G}$ in Refs.\
269: \cite{BSOB,SBE}, the dipole-velocity
270: $F_{\rm DV}$ and field-velocity $F_{\rm HV}$ in Ref.\ \cite{BC},
271: and the first-order field-velocity term 
272: $F_{\rm HV1}$ in Ref.\ \cite{VolovikMineev}. Because of Galilean
273: invariance only the combination ${\bf v}_{\rm s}-{\bf v}_{\rm n}$
274: appears in Equations (\ref{e.2.6}) and (\ref{e.2.7}). The superfluid
275: velocity does not appear in the gyromagnetic term (\ref{e.2.8}) because
276: ${\bf v}_{\rm s}$ is curl free. Terms that are linear in
277: $\hat{\bf n}\cdot{\bf H}$,
278: $\hat{\bf n}\cdot({\bf v}_{\rm s}-{\bf v}_{\rm n})$ or ${\bf
279: H}\cdot\tensor R\cdot({\bf v}_{{\rm s}}-{\bf v}_{{\rm n}})$ are
280: prohibited by parity and time-reversal symmetry. Equations
281: (\ref{e.2.5})-(\ref{e.2.9}) serve as definitions of the parameters $a$,
282: $\lambda_{\rm DV}$,
283: $\lambda_{\rm HV}$, $\lambda_{\rm HV1}$,
284: $\lambda_{\rm G1}$ and $\lambda_{\rm G2}$, which depend on temperature
285: and pressure. The calculation of these parameters are discussed in
286: Sections
287: \ref{s.qc} and
288: \ref{s.gl}. For some of the parameters we use names given in Ref.\
289: \cite{SBE} instead of the more systematic names introduced here; for
290: example
291: $\lambda_{\rm DH}\equiv a$ and $\lambda_{\rm SH}\equiv d$.  
292:  
293: The derivation of the gradient energy (\ref{e.2.9}) deserves special
294: consideration. Originally, one starts from a general expression that is
295: quadratic in the spin velocity (\ref{e.2.3}). Making use of the
296: properties of the rotation matrices,
297: it is possible to simplify the energy to a form that
298: is bilinear in the rotation matrices $R_{\alpha i}$. 
299: In addition to the two terms in (\ref{e.2.9}), this form contains a
300: third  term of the form $\partial_iR_{\alpha j}\partial_jR_{\alpha
301: i}$.    The form (\ref{e.2.9}) can then be obtained by partial
302: integration which converts the third term into the form 
303: $\partial_iR_{\alpha i}\partial_jR_{\alpha j}$.
304: 
305: It should be noted that the partial integration of the gradient energy
306: (\ref{e.2.9})  produces a surface term that is similar to $F_{\rm SG}$
307: below. Thus the value of the surface coefficient
308: $\lambda_{\rm SG}$ is unique only if the form of the bulk gradient 
309: energy is properly defined. Here the uniqueness of $\lambda_{\rm SG}$ is
310: guaranteed by restricting the bulk gradient energy to the form
311: (\ref{e.2.9}).
312:   
313: The gradient term can also be expressed explicitly as a function
314: of
315: $\hat{\bf n}$ using the representation (\ref{e.2.4}) \cite{SBE}. The
316: needed identities are given in the Appendix. Our preference is to keep
317: the shorter form (\ref{e.2.9}) because a numerical algorithm can
318: directly be based on it.
319: 
320: The dipole length $\xi_{\rm D}$ is defined by $\xi_{\rm
321: D}=\sqrt{\lambda_{\rm G2}/\lambda_{\rm D}}$. It is conventional to
322: define dipole velocity
323: $v_{\rm D}$ and dipole field $H_{\rm D}$ by writing $\lambda_{\rm
324: HV}=2a/(5v_{\rm D}^2)$ and
325: $\lambda_{\rm DV}=aH_{\rm D}^2v_{\rm D}^{-2}$. We can also define a
326: magnetic coherence length
327: $\xi_{\rm H}=\sqrt{65\lambda_{\rm G2}/(8aH^2)}$, which is inversely
328: proportional to the field. The parameters defined here are temperature
329: dependent. Near $T_{\rm c}$ they reduce to constants that are commonly
330: used. For example, $\xi_{\rm H}\rightarrow R_{\rm c}H_{\rm B}/H$
331: defined in Ref. \cite{SBE}.
332:  
333: In addition to the bulk terms (\ref{e.2.5})-(\ref{e.2.9}), there are
334: boundary terms. These energy terms originate from regions where the
335: order parameter is strongly distorted from the form (\ref{e.2.1}). We
336: are here interested in two cases: surfaces and vortex cores. The
337: boundary terms below are valid in the limit that the length scale of
338: the distorted region ($\approx\xi_0$) is small compared to the dipole
339: length $\xi_{\rm D}$. In reality this is well the case. It guarantees
340: that the rotation angle $\theta$ is not affected by the boundary. The
341: form of the allowed boundary terms depends on the symmetry of the order
342: parameter at the boundary.
343: 
344: We assume that the surface structure has the maximal symmetry, i.e.\
345: time-inversion symmetry, rotation symmetry around the surface normal and
346: reflection symmetry in planes perpendicular to the surface. (We note
347: that also less symmetric states are possible \cite{T86}.)  We also
348: assume that the curvature of the surface is small. Such a surface gives
349: rise to the energy terms 
350: \begin{equation} F_{\rm SH}=-d\int_S d^2r({\bf H}\cdot\tensor R\cdot\hat
351: {\bf s})^2
352: \label{e.2.10}\end{equation}
353: \begin{equation} F_{\rm SHV1}=-\lambda_{\rm SHV1}
354: \int_Sd^2r{\bf H}\cdot\tensor R\cdot\hat {\bf s}\times({\bf v}_{{\rm
355: s}}-{\bf v}_{{\rm n}})
356: \label{e.2.11}\end{equation}
357: \begin{equation} F_{\rm SG}=\lambda_{\rm SG}\int_Sd^2r\hat s_jR_{\alpha
358: j}{\partial R_{\alpha i}\over\partial r_i}
359: \label{e.2.12}\end{equation}
360: \begin{equation} F_{\rm SD}=\int_Sd^2r[ b_4(\hat{\bf s}\cdot\hat{\bf
361: n})^4-b_2(\hat{\bf s}\cdot\hat{\bf n})^2].
362: \label{e.2.13}\end{equation}  Here
363: $\hat{\bf s}$ is a unit vector that is perpendicular to the surface and
364: points towards the superfluid. The surface-field term $F_{\rm SH}$ is
365: discussed in Ref.\
366: \cite{BSOB,SBE}, the surface-dipole term $F_{\rm SD}$ in Ref.\
367: \cite{BSOB,FominVuorio,SBE}, and the first-order surface-field-velocity
368: term
369: $F_{\rm SHV1}$ in Ref.\
370: \cite{VolovikMineev}. There are two contributions to the
371: surface-gradient coefficient, 
372: $\lambda_{\rm SG}=\lambda_{\rm SG}^{\rm a}+\lambda_{\rm SG}^{\rm
373: b}$. The former comes from the equilibrium spin current that flows
374: spontaneously along any surface\cite{ZKT}. In fact, the surface spin
375: current 
376: $J^{\rm ss}_{\alpha i}=\lambda_{\rm SG}^{\rm a}R_{\alpha
377: j}\epsilon_{ijk}\hat s_k$. (Note that $J^{\rm ss}$ contains the
378: factor 
379: $\hbar/2$ for each fermion and thus has the unit J/m.) The other
380: contribution
381: $\lambda_{\rm SG}^{\rm b}$ comes from the partial integration that
382: depends on the chosen form of
383: $F_{\rm G}$ \cite{SBE}. Note that there exists only one
384: surface-gradient term (\ref{e.2.12}) because
385: $R_{\alpha i}\nabla R_{\alpha j}$ is antisymmetric in $i$ and $j$. For
386: the same reason the term (\ref{e.2.12}) does not depend on the normal
387: derivative. We have constructed the definitions
388: (\ref{e.2.5})-(\ref{e.2.13}) so that all surface ($d$, $\lambda_{\rm
389: SHV1}$, $\lambda_{\rm SG}$,
390: $b_2$,
391: $b_4$)  and bulk coefficients  are non-negative, at least in the
392: Ginzburg-Landau region.
393: 
394:  
395: The order parameter is strongly distorted from the bulk form
396: (\ref{e.2.1}) in the cores of quantized vortex lines. Therefore the
397: cores must be treated as boundary regions. We describe a vortex line by
398: unit vector $\hat{\bf l}$ that is parallel to the line and points in
399: the direction of the circulation $\nabla\times{\bf v}_{\rm s}$. The
400: maximal point-symmetry operations of a vortex are  a) reflection in
401: plane perpendicular to $\hat{\bf l}$, b) rotation around $\hat{\bf l}$
402: (combined with a phase shift) and c) reflection in plane containing
403: $\hat{\bf l}$. The last one has to be combined with time inversion
404: because otherwise the circulation would change direction. Assuming that
405: the order parameter in the core has all these symmetries, we get the
406: phenomenological terms
407: \begin{equation} F_{\rm LH}=\lambda_{\rm LH}
408: \int_L d^3r({\bf H}\cdot\tensor R\cdot\hat {\bf l})^2
409: \label{e.2.14}\end{equation}
410: \begin{equation} F_{\rm LH1}=\lambda_{\rm LH1}\int_Ld^3r{\bf
411: H}\cdot\tensor R\cdot\hat {\bf l}
412: \label{e.2.15}\end{equation}
413: \begin{equation} F_{\rm LD}=\lambda_{\rm LD}\int_Ld^3r[(\hat{\bf
414: l}\cdot\hat{\bf n})^2+\hbox{corrections}].
415: \label{e.2.16}\end{equation}
416:  The line-field term $F_{\rm LH}$ is discussed in Ref.\
417: \cite{GGK}, the first-order line-field term $F_{\rm LH1}$ in Ref.\
418: \cite{HKSSBMV}, and line-dipole term
419: $F_{\rm LD}$ in Ref.\ \cite{T87}.  Here $L$ denotes the region where
420: vortices are present. In $F_{\rm LD}$ only the dominant term is written
421: explicitly. 
422: 
423: It is well known that the vortex cores do not have the maximal symmetry
424: \cite{T87}. In the A-phase-core vortex the symmetry (a) is broken.
425: Because this can take place in two different ways, we have to assign to
426: each vortex line a new variable
427: $q$ that equals either +1 (left-handed vortex) or -1 (right handed
428: vortex). This allows the line-gradient term \cite{SVrev}
429: \begin{equation} F_{\rm LG}=\lambda_{\rm LG}\int_Ld^3r\, \langle
430: q\rangle \hat l_jR_{\alpha j}{\partial R_{\alpha i}\over\partial r_i}, 
431: \label{e.2.17}\end{equation}
432: where $\langle \ldots\rangle$ denotes the average because $q$ may
433: change from one vortex to another. Similar to the surface term
434: $F_{\rm SG}$,
435: $F_{\rm LG}$ arises from spontaneous spin currents. For an
436: isolated vortex these currents form closed loops in the plane
437: perpendicular to
438: $\hat{\bf l}$. All vortices in $^3$He-B also have axial
439: spin currents but they do not
440: couple to external spin velocity in the lowest order because the net
441: current vanishes.
442: 
443: The
444: double-core vortex also allows the term (\ref{e.2.17}). Additionally,
445: the circular symmetry (b) is broken leaving only discrete symmetry in
446: rotations by
447: $\pi$. Thus an additional unit vector
448: $\hat{\bf b}$ perpendicular to $\hat{\bf l}$ is needed to describe the
449: vortex. This gives rise to line-anisotropy terms
450: \begin{equation} F_{\rm LAH}=\lambda_{\rm LAH}
451: \int_L d^3r\langle({\bf H}\cdot\tensor R\cdot\hat {\bf b})^2\rangle
452: \label{e.2.18}\end{equation}
453: \begin{equation} F_{\rm LAD}=\lambda_{\rm
454: LAD}\int_Ld^3r\langle(\hat{\bf b}\cdot\hat{\bf
455: n})^2+\hbox{corrections}\rangle.
456: \label{e.2.19}\end{equation}
457:  
458: We note that there is flexibility in the definitions of the different
459: terms. For example, the superflow around a vortex has to be counted
460: into term $F_{\rm LH}$ (\ref{e.2.14}) in the region where the order
461: parameter is strongly distorted but at larger distances it also can be
462: included as the bulk term $F_{\rm HV}$ (\ref{e.2.7}). 
463: 
464: \section{Connection to the quasiclassical theory}\label{s.qc}
465: 
466: The energy terms (\ref{e.dipgen})-(\ref{e.2.19}) contain a number of
467: phenomenological coefficients. They should either be determined
468: experimentally or calculated from a more microscopic theory than the
469: hydrodynamic one. Pursuing the latter, there exists the quasiclassical
470: theory \cite{SRrev}. This theory bypasses the difficult many-body
471: problem of strongly interacting $^3$He
472: atoms by concentrating in the low-energy range. It uses an
473: expansion, where the relevant expansion parameter for the
474: superfluid phases is the transition temperature divided by the Fermi
475: temperature, $T_{\rm c}/T_{\rm F}\sim 0.001$. The lowest nontrivial
476: order in this expansion is known as the {\it weak-coupling} theory. It
477: effectively contains the Bardeen-Cooper-Schrieffer theory as a special
478: case, but it also reduces to the Landau Fermi-liquid theory in the
479: normal state. This theory is adequate for some properties of superfluid
480: $^3$He, especially at low pressures, but it fails, for example, to
481: stabilize the A phase. For many purposes it is important to continue
482: the expansion to the next order in $T_{\rm c}/T_{\rm F}$.
483: We call this the {\it strong-coupling}
484: theory. (Serene and Rainer use the name ``weak-coupling plus'', but we
485: think this is too modest since there seems to be very little hope to
486: calculate further orders in the expansion.)
487: We will not go into the details of the quasiclassical
488: theory, which is extensively discussed by Serene and Rainer
489: \cite{SRrev}. 
490: 
491: It is important to realize that the 
492: quasiclassical theory is not microscopic in the sense that it would
493: depend only on fundamental constants. Instead, it needs several
494: parameters as input. This is especially a problem in the
495: strong coupling case, which needs as input the scattering amplitude of
496: quasiparticles (in the normal state) that is not accurately known.
497: Additionally, the needed calculations are rather complicated at general
498: temperature. There are two practical ways to proceed. The first is to
499: restrict to the temperature region close to
500: $T_{\rm c}$ and use the Ginzburg-Landau theory. This approach will
501: be described in Section \ref{s.gl}. The second way is to work at
502: arbitrary temperature but to use the weak-coupling approximation in the
503: quasiclassical theory. The latter approach is discussed in this section.
504: At the end of this section we discuss how to improve the weak-coupling
505: results by including a trivial strong-coupling correction.
506:  
507: In the weak-coupling theory, the properties of the normal state are
508: included via spin symmetric and antisymmetric Fermi-liquid
509: parameters, 
510: $F_l^{\rm s}$ and
511: $F_l^{\rm a}$ ($l=0$, 1, 2, $\ldots\infty$). We assume that the pairing
512: interaction is effective in the
513: p-wave channel only. The symmetry between particle and hole types of
514: quasiparticles is consistently assumed in the quasiclassical theory. It
515: turns out that all results presented below depend only on five
516: Fermi-liquid parameters: 
517: $F_1^{\rm s}$ and
518: $F_l^{\rm a}$ with $l=0$, 1, 2, and 3. (Infinite number of coefficients
519: is  needed in the hydrostatics of the A phase \cite{SRres}.) In
520: addition, the results depend on  the mass density $\rho$ of $^3$He
521: liquid, on the superfluid transition temperature $T_{\rm c}$ and on the
522: magnetic dipole-dipole interaction parameter $g_{\rm D}$. 
523: 
524: In the Bardeen-Cooper-Schrieffer model $g_{\rm D}$ has the expression 
525: \cite{LeggettAnn} (in SI units)
526: \begin{equation} g_{\rm D}={\mu_0\over 40}\bar{R^2}\left(\hbar\gamma
527: N(0)\pi k_{\rm B}T\sum_{\epsilon_n=-\epsilon_{\rm c}}^{\epsilon_{\rm c}}
528: {1\over\sqrt{\epsilon_n^2+\Delta^2}}\right)^2,
529: \label{e.gd}\end{equation} where $\bar{R^2}$ is a renormalization
530: constant and $\epsilon_{\rm c}$ a high energy cut-off. (Note that our
531: definition of $g_{\rm D}$ \cite{Fetterrev} is different from that in
532: Ref.\ \cite{LeggettAnn}.) The
533: Matsubara energies $\epsilon_n$ and the weak-coupling
534: energy gap  $\Delta(T)$ are defined in the Appendix, the
535: gyromagnetic ratio of the
536: $^3$He nucleus $\gamma=-2.04\cdot10^8\ ({\rm T\ s})^{-1}$, and the
537: density of states at the Fermi energy $2N(0)=(1+{1\over 3}F_1^{\rm
538: s})(3m^2\rho/\pi^4\hbar^6)^{1/3}$. It is very convenient that the
539: dependence of
540: $g_{\rm D}$ on temperature is so weak that we can safely ignore it. In
541: the weak-coupling approximation the constancy of $g_{\rm D}(T)$ would
542: be exact if the cut-off energy $\epsilon_{\rm c}$ in (\ref{e.gd}) were
543: the same in the gap equation (\ref{e.qcgap}). (For the standard choice
544: $\epsilon_{\rm c}\rightarrow\infty$ in the gap equation and
545: $\bar{R^2}=1$, the relative variation of
546: $g_{\rm D}(T)$ is less than
547: $10^{-5}$.) In the trivial strong-coupling model (see below) Eq.
548: (\ref{e.gd}) gives the maximum variation at the melting pressure, where
549: $g_{\rm D}$ decreases monotonically by 1.3\%  when $T$ decreases from
550: $T_{\rm c}$ to zero (assuming
551: $\bar{R^2}=1$). Because of uncertainties associated with $\epsilon_{\rm
552: c}$ and
553: $\bar{R^2}$ in Eq.\ (\ref{e.gd}), we prefer to extract $g_{\rm D}$ from
554: experiments, as will be discussed in section \ref{s.gd}. 
555: 
556: For completeness, we give the results for nuclear magnetic
557: susceptibility $\chi$
558: \cite{SRQ77}, superfluid density $\rho_{\rm s}$, and $\lambda_{\rm D}$
559: (\ref{e.dipgen})
560: \begin{equation} \chi=2\mu_0N(0)\left({\hbar\gamma\over
561: 2}\right)^2{{2\over 3}+({1\over 3}+{1\over 5}F^{\rm a}_2)Y\over
562: 1+F_0^{\rm a}({2\over 3}+{1\over 3}Y)+{1\over 5}F_2^{\rm a} ({1\over
563: 3}+({2\over 3}+F_0^{\rm a})Y)}
564: \label{e.suskis}\end{equation}
565: \begin{equation} \rho_{\rm s}=\rho{1-Y\over 1+{1\over 3}F_1^{\rm s}Y}
566: \label{e.rhos}\end{equation}
567: \begin{equation} \lambda_{\rm D}= g_{\rm D}\Delta^2.
568: \label{e.lambdad}\end{equation} All the following coefficients can be
569: understood as corrections to these. Here $Y(T)=1-Z_3(T)$ is the Yoshida
570: function, and the functions
571: $Z_j(T)$ are defined in the Appendix. 
572:  
573: The basic principle for calculating the hydrostatic parameters is
574: explained in Section VI of Ref.\
575: \cite{SRrev}.  For the coefficient of the dipole-field energy $F_{\rm
576: DH}$ the main part of the work, the calculation of the gap distortion,
577: is explained in detail in Ref.\
578: \cite{FS}. The result is
579: \begin{equation} a={5g_{\rm D}\over 2}\left[ {{1\over
580: 2}\hbar\gamma\mu_0(1+{1\over 5}F^{\rm a}_2)\over 1+F_0^{\rm a}({2\over
581: 3}+{1\over 3}Y)+{1\over 5}F_2^{\rm a} ({1\over 3}+({2\over 3}+F_0^{\rm
582: a})Y)}\right]^2 \left[5-{3Z_5\over Z_3} -{3F^{\rm a}_2Z_3\over
583: 5(1+{1\over 5}F^{\rm a}_2)}\right].
584: \label{e.3.1}\end{equation}   The coefficient of the dipole-velocity
585: energy (\ref{e.2.6}) can be calculated in a similar way and we obtain 
586: \begin{equation} \lambda_{\rm DV}=5g_{\rm D}\left({m^*v_{\rm F}\over
587: 1+{1\over 3}F_1^{\rm s}Y}\right)^2 \left(1-{3Z_5\over 2Z_3}\right).
588: \label{e.3.3}\end{equation} Here $m^*$ is the effective mass given by
589: $m^*/m=1+F^{\rm s}_1/3$. The Fermi velocity $v_{\rm F}$ is related to
590: basic parameters by $v_{\rm F}=\hbar(3\pi^2\rho/m)^{1/3}/m^*$. 
591: As far as we know, the expressions
592: (\ref{e.3.1}) and (\ref{e.3.3}) have not been published before. Rather
593: tedious calculation is needed for the coefficient in the field-velocity
594: energy (\ref{e.2.7}).  This is done in Ref.\ \cite{ghv}, and we quote
595: the result
596: \begin{eqnarray} \lambda_{\rm
597: HV}&=&{\rho\over\Delta^2}{m^*/m\over(1+{1\over 3}F_1^{\rm s}Y)^2}\left[
598: {{1\over 2}\hbar\gamma\mu_0(1+{1\over 5}F^{\rm a}_2)\over 1+F_0^{\rm
599: a}({2\over 3}+{1\over 3}Y)+{1\over 5}F_2^{\rm a} ({1\over 3}+({2\over
600: 3}+F_0^{\rm a})Y)}\right]^2\nonumber \\ &&\times\left[Z_3 -{9\over
601: 10}Z_5+{9\over 10}{Z_5^2\over Z_3}-{3\over 2}Z_7+{3F^{\rm a}_2Z_3\over
602: 50(1+{1\over 5}F^{\rm a}_2)}(3Z_5-2Z_3)\right].
603: \label{e.3.4}\end{eqnarray}   The gyromagnetic coefficient
604: $\lambda_{\rm HV1}=0$ because of particle-hole symmetry
605: \cite{Mineev}.   The gradient energy coefficients are calculated in
606: Ref.\ \cite{Dorfle}, and can also be found in Appendix F of Ref.\
607: \cite{SRrev}. They are
608: \begin{equation} \lambda_{\rm G2}={\hbar^2\rho\over 40mm^*} {(1+{1\over
609: 3}F_1^{\rm a})(1+{1\over 7}F_3^{\rm a})(1-Y)\over 1+{1\over 3}F_1^{\rm
610: a}({2\over 5}+{3\over 5}Y)+{1\over 7}F_3^{\rm a} ({3\over 5}+({2\over
611: 5}+{1\over 3}F_1^{\rm a})Y)}
612: \label{e.3.5}\end{equation}
613: \begin{equation} {\lambda_{\rm G1}\over\lambda_{\rm G2}}=2+ {({1\over
614: 3}F_1^{\rm a}-{1\over 7}F_3^{\rm a})(1-Y)\over (1+{1\over 7}F_3^{\rm
615: a})(1+{1\over 3}F_1^{\rm a}Y)}
616: \label{e.3.6}\end{equation}
617:  
618: The structure of the order parameter near surfaces has been studied for
619: a long time (for example in Refs.\
620: \cite{Buchholtz} and \cite{ZKT}), but the surface
621: terms have been evaluated only quite
622: recently. The gyromagnetic surface term (\ref{e.2.11}) vanishes
623: identically because of particle-hole symmetry. As discussed above, the
624: surface-gradient (\ref{e.2.12}) term has two contributions:
625: $\lambda_{\rm SG}=\lambda_{\rm SG}^{\rm a}+\lambda_{\rm SG}^{\rm b}$.
626: For the part that arises from the partial integration in the derivation
627: of (\ref{e.2.9}) we find 
628: $\lambda_{\rm SG}^{\rm b}=2\lambda_{\rm G2}$. The other part
629: $\lambda_{\rm SG}^{\rm a}$ coming from spontaneous spin currents has
630: recently been calculated in Ref.\
631: \cite{VT}. The same reference evaluates also the
632: surface dipole coefficients in (\ref{e.2.13}). The field term
633: (\ref{e.2.10}) has not been calculated. Until this
634: is done, we can use an extrapolation of the Ginzburg-Landau result 
635: $d={\mu_0\over 2}(\chi_{\rm n}-\chi)\xi_{\rm GL}d_0$, where
636: $d_0=d/g_{\rm H}\Delta^2\xi_{\rm GL}$ is plotted by solid lines in Fig.\
637: \ref{f.surfgl}. The Ginzburg-Landau coherence length
638: $\xi_{\rm GL}$ can be extrapolated to general temperature by
639: $\xi_{\rm GL}(T)=\hbar v_{\rm F}/\sqrt{10}\Delta(T)$. (Note that no
640: strong-coupling correction to the weak-coupling $\Delta$ is allowed in
641: this equation.)
642: $\chi_{\rm n}$ is the susceptibility in the normal
643: state [given by (\ref{e.suskis}) with $Y=1$].
644: 
645: For accurate calculation of the vortex terms, one needs a calculation
646: of the order parameter in the vortex core. This has been done at
647: general temperature by Fogelstr\"om and Kurkij\"arvi
648: \cite{FogelstromKurkijarvi}, but they do not give explicit values of
649: $\lambda_{\rm LH}$. However, at not too high rotation velocities, the
650: most of the contribution to
651: $\lambda_{\rm LH}$ comes from outside of the vortex core, and therefore
652: a reasonable estimate at all temperatures is 
653: \begin{equation} \lambda_{\rm LH}\approx {1\over 2}\lambda_{\rm
654: HV}\langle\vert{\bf v}_{\rm s}-{\bf v}_{\rm n}\vert^2\rangle_L\approx
655: {\hbar\over 2m}\Omega\lambda_{\rm HV}(\ln{R\over r_{\rm L}}-{3\over 4}),
656: \label{e.lambdaestimate}\end{equation} where $\Omega$ is the angular
657: velocity of rotation, $R=\sqrt{\hbar/2m\Omega}$ the unit cell radius of
658: a vortex, and $r_{\rm L}$ the radius of the vortex core. Because
659: $\lambda_{\rm LH}$ is rather insensitive to $r_{\rm L}$, we may use
660: $r_{\rm L}\approx\xi_{\rm GL}$ \cite{T87}, and use the same
661: extrapolation of
662: $\xi_{\rm GL}$ as for the surface term.
663: 
664: The weak coupling approximation used above is not expected to be
665: accurate at high pressures, where strong coupling corrections are
666: largest. As mentioned above, accurate strong-coupling calculations are
667: very cumbersome, and introduce the scattering amplitude, which is
668: poorly known. However, there is a simple procedure that is expected to
669: take into account a major part of the strong coupling effects. This
670: ``trivial strong coupling correction'' is to scale the energy gap
671: $\Delta$ by a temperature and pressure dependent factor. This factor is
672: tabulated by Serene and Rainer
673: \cite{SRrev} as a function of the temperature and the relative jump
674: $\Delta C/C_{\rm n}$ of the specific heat at
675: $T_{\rm c}$. The latter can be related to pressure according to the
676: measurements by Greywall
677: \cite{Greywall}. Such scaling of $\Delta$ affects all hydrostatic
678: coefficients (\ref{e.gd})-(\ref{e.lambdaestimate}) directly and/or via
679: modification of the functions $Z_j$ and $Y$.
680: 
681: 
682: \section{Connection to the Ginzburg-Landau theory}\label{s.gl}
683: 
684: The hydrostatic theory was based on expansion of the free energy in
685: small gradients and external fields.  The Ginzburg-Landau
686: (GL) theory is based on additional expansion in the amplitude of
687: the order parameter
688: \cite{GL}. The expansion can be limited to a small number of terms near
689: the superfluid transition temperature
690: $T_{\rm c}$, where the order parameter is small. In
691: most superconductors and in $^3$He, the GL theory gives
692: reliable results in the neighborhood of $T_{\rm c}$ because the
693: fluctuation range, where it becomes invalid, consists of a negligible
694: temperature range just at $T_{\rm c}$.
695: 
696: The order parameter in $^3$He is $3\times 3$ matrix $A_{\alpha
697: j}$. The GL theory consists of writing down the
698: terms in the free energy that are allowed by known symmetries. The
699: superfluid condensation energy must be invariant in separate rotations
700: in the spin and orbital spaces. This allows the leading terms 
701: \cite{MerminStare}
702: \begin{eqnarray}
703: F_{\rm cond}&=&\int d^3r\big[-\alpha A_{\mu i}^*A_{\mu i}
704: +\beta_1 A_{\mu i}^*A_{\mu i}^*A_{\nu j}  A_{\nu j}
705: +\beta_2 A_{\mu i}^*A_{\mu i}  A_{\nu j}^*A_{\nu j}\nonumber \\\mbox{}&&
706: +\beta_3 A_{\mu i}^*A_{\nu i}^*A_{\nu j}  A_{\mu j}
707: +\beta_4 A_{\mu i}^*A_{\nu i}  A_{\nu j}^*A_{\mu j}
708: +\beta_5 A_{\mu i}^*A_{\nu i}  A_{\nu j}  A_{\mu j}^*\big].
709: \label{e.gl}
710: \end{eqnarray}
711: The zero of the coefficient $\alpha=\alpha'(1-T/T_{\rm c})$ defines
712: the transition temperature $T_{\rm c}$. Other coefficients $\beta_i$
713: ($i=1...5$) as well as $\alpha'$ can be taken as constants in
714: the expansion of $F_{\rm cond}$ to order $(1-T/T_{\rm c})^2$. In
715: the presence of nonuniform order parameter one needs the gradient 
716: energy
717: \cite{AGR}
718: \begin{eqnarray}
719: F_{\rm G}&=&K\int d^3r\left[(\gamma-2\eta)(\partial_iA_{\mu
720: i})^*\partial_jA_{\mu j} +(\partial_iA_{\mu j})^*\partial_iA_{\mu
721: j}+(2\eta-1)(\partial_iA_{\mu j})^*\partial_jA_{\mu i}
722: \right]\nonumber\\
723: &=&K\int d^3r\left[(\gamma-1)({\partial_i}A_{\mu
724: i})^*\partial_jA_{\mu j} +(\partial_iA_{\mu j})^*\partial_iA_{\mu
725: j}-{\rm i}(2\eta-1)\epsilon_{kij}(\mbox{\boldmath$\nabla$}
726: \times{\bf v}_{\rm n})_kA_{\mu i}^*A_{\mu j} \right]\label{e.GLgradient}
727: \end{eqnarray}
728: with the Galilean-invariant derivative $\mbox{\boldmath$\partial$}=
729: \mbox{\boldmath$\nabla$}+2{\rm i}m{\bf v}_{\rm n}/\hbar$. 
730: In
731: addition there are the energy caused by the magnetic field ${\bf H}$
732: \cite{AmbeM,MerminStare}
733: \begin{equation}
734: F_{\rm H}=\int d^3r\left(-{\rm i}g_{\rm H1}
735: \epsilon_{\kappa\mu\nu}H_\kappa A_{\mu i}^*A_{\nu i}+g_{\rm H}H_\mu
736: A_{\mu i}^*A_{\nu i}H_\nu
737: +g_{\rm H}'H^2A_{\mu i}^*A_{\mu i}\right),
738: \label{e.GLfield}
739: \end{equation}
740: and the energy of the magnetic
741: dipole-dipole interaction \cite{LeggettAnn}  
742: \begin{equation}
743: F_{\rm D}=g_{\rm D}\int d^3r(A_{i i}^*A_{j j}+A_{i j}^*A_{j
744: i}-{\textstyle{2\over 3}}A_{\mu i}^*A_{\mu i}).
745: \label{e.GLdipole}
746: \end{equation}
747: We neglect all terms in the free energy that are independent of
748: $A_{\alpha j}$.  The 
749: gradient energy (\ref{e.GLgradient}) is parametrized using two
750: dimensionless parameters $\gamma$ and
751: $\eta$, which are related to parameters introduced by Serene and Rainer
752: \cite{SRres} as $\gamma=K_{\rm
753: L}/K_{\rm T}$ and $\eta=K_{\rm C}/K_{\rm T}$.
754: The two different forms (\ref{e.GLgradient}) are equivalent, as can
755: be verified by partial integration. In contrast to the hydrostatic
756: case (\ref{e.2.9}), the surface term in the partial integration
757: vanishes here because of the boundary condition $\hat s_iA_{\mu i}=0$. 
758: 
759: The parameters of the GL theory have been calculated using the
760: weak-coupling quasiclassical theory, and the results are well known (see
761: Refs.\ \cite{F75,T87}, for example). There are two alternatives to
762: incorporate the strong-coupling effects. One is to determine the
763: coefficients purely experimentally. The
764: $\beta_i$'s, or at least some combinations of them, have been determined
765: using experiments in the superfluid phases
766: \cite{THBG,gshift,HBBG}. The other alternative is to consider the
767: GL theory as a limiting case of the strong-coupling quasiclassical
768: theory near $T_{\rm c}$ \cite{RSbeta}. Here the
769: problem of the poorly-known scattering amplitude is encountered again,
770: but fortunately there exists model calculations for the most
771: important coefficients. We give here a short summary of the results.
772: 
773: There is a small correction to $\alpha'$ arising from finite lifetime
774: of quasiparticles \cite{RSspecific}. There are several suggestions for the
775: $\beta_i$'s that are based on different theoretical assumptions about
776: the scattering amplitude and measurements in the {\it normal state} of
777: $^3$He \cite{SS,Bedell,LevinValls}. Although the strong-coupling
778: corrections generally are small, they can be quite substantial in some
779: combinations of $\beta_i$'s. For example,
780: $\beta_{345}\equiv\beta_3+\beta_4+\beta_5$ may differ 50\% from its
781: weak-coupling value. The corrections to
782: $K$, $\gamma$, $\eta$,
783: $g_{\rm H}$, and $g_{\rm H}'$ are calculated by Serene and Rainer
784: in Ref.\ \cite{SRres}. They find that $\eta$ is unchanged from its
785: weak-coupling value 1, but $\gamma$  increases from its weak coupling
786: value 3 to
787: $\approx 3.1$ at the melting pressure.
788: $g_{\rm H}'$ is found to vanish even after strong-coupling corrections,
789: and therefore it is dropped in the following. 
790: The parameter
791: $g_{\rm H1}$ vanishes in the quasiclassical theory because of
792: particle-hole symmetry, but this term is still kept because it
793: is important in several situations. Its value is best extracted from
794: measurements of the splitting of the A transition
795: in magnetic field \cite{Sagan,Israelsson}.
796: We have assumed that the nontrivial corrections to the dipole energy
797: (\ref{e.GLdipole}) are small, and therefore use the same coefficient
798: $g_{\rm D}$ as already discussed in Sect.\ \ref{s.qc}.   
799: 
800: The calculations in the GL theory are considerably simpler than in the
801: general quasiclassical theory. Essentially all the hydrostatic
802: parameters appearing in equations (\ref{e.2.5})-(\ref{e.2.19}) have been
803: calculated. We list below the bulk hydrostatic coefficients as
804: functions of GL parameters.
805: \begin{equation} a={5g_{\rm D}g_{\rm H}\over 4\beta_{345}}
806: \label{e.4.1}\end{equation}
807: \begin{equation} \lambda_{\rm DV}={5m^2g_{\rm D}(\gamma-1)K
808: \over\hbar^2\beta_{345}}
809: \label{e.4.2}\end{equation}
810: \begin{equation} \lambda_{\rm HV}={2m^2g_{\rm H}(\gamma-1)K
811: \over\hbar^2\beta_{345}}
812: \label{e.4.3}\end{equation}
813: \begin{equation} \lambda_{\rm HV1}={mg_{\rm
814: H1}(\gamma-4\eta+1)K\over\hbar(\beta_4-3\beta_1-\beta_{35})}
815: \label{e.4.4}\end{equation}
816: \begin{equation} \lambda_{\rm G2}=K\Delta^2={\alpha K\over
817: 2(3\beta_{12}+\beta_{345})}
818: \label{e.4.5}\end{equation}
819: \begin{equation} {\lambda_{\rm G1}\over\lambda_{\rm G2}}=\gamma-1.
820: \label{e.4.6}\end{equation} 
821: The first equality in (\ref{e.4.5}) and (\ref{e.4.6}) can be obtained
822: trivially by substituting the B-phase order parameter (\ref{e.2.1})
823: into the gradient energy (\ref{e.GLgradient}). The amplitude $\Delta$ of
824: the order parameter is obtained by substitution into $F_{\rm cond}$
825: (\ref{e.gl}) and minimization with respect to $\Delta$. Equations
826: (\ref{e.4.1})-(\ref{e.4.3}) can be obtained by solving the GL equations
827: in simple cases of axially distorted B phase. For example, the
828: coefficient (\ref{e.4.1}) can be obtained by first calculating the
829: anisotropy of the gap due to a magnetic field and then evaluating the
830: dipole energy for this gap. The gyromagnetic coefficient
831: $\lambda_{\rm HV1}$ (\ref{e.4.4}) has been calculated by Mineev
832: \cite{Mineev}. Because of deviation of $\gamma$ from 3, it is
833: considerably larger than anticipated in Refs.\
834: \cite{VolovikMineev,Mineev}.
835:  
836: Accurate determination of the surface terms requires a self-consistent
837: solution of the order parameter near a wall. In the absence of fields,
838: the order parameter
839: $\tilde A_{\alpha i}$, which  is normalized to unit matrix in the
840: bulk, has  real components
841: $\tilde A_{xx}(x)$ and
842: $\tilde A_{yy}(x)=\tilde A_{zz}(x)$ near a surface located in the $y-z$
843: plane. The surface coefficients are then obtained by integration: 
844: \begin{equation} d=g_{\rm H}\Delta^2\xi_{\rm GL}\int_0^\infty
845: {dx\over\xi_{\rm GL}}(\tilde A_{yy}^2-\tilde A_{xx}^2)
846: \label{e.d0}\end{equation} 
847: \begin{equation}
848: \lambda_{\rm SG}=K\Delta^2\int_0^\infty dx2(\gamma-1)\tilde
849: A_{yy}{d\tilde A_{xx}\over dx}
850: \label{e.sg0}\end{equation} 
851: \begin{equation} b_2=g_{\rm D}\Delta^2\xi_{\rm GL}\int_0^\infty
852: {dx\over\xi_{\rm GL}} {5\over 4}(\tilde A_{xx}^2-6\tilde A_{xx}\tilde
853: A_{yy}+5\tilde A_{yy}^2)
854: \label{e.b2}\end{equation} 
855: \begin{equation} b_4=g_{\rm D}\Delta^2\xi_{\rm GL}\int_0^\infty
856: {dx\over\xi_{\rm GL}} {25\over 8}(\tilde A_{yy}-\tilde A_{xx})^2,
857: \label{e.b4}\end{equation} where $\xi_{\rm GL}=\sqrt{K/\alpha}$.  The
858: surface term $F_{\rm SHV1}$ can be found by calculating the order
859: parameter in the presence of phase gradient: $A_{\alpha j}(x,y)=\Delta
860: R_{\alpha i}\exp(\imagu ky)\tilde A_{ij}(x)$. The coefficient is then
861: given by
862: \begin{equation}
863: \lambda_{\rm SHV1}={2m\over \hbar}g_{\rm H1}\Delta^2\xi_{\rm
864: GL}^2\int_0^\infty {dx\over
865: \xi_{\rm GL}^2}\lim_{k\rightarrow 0}{1\over\imagu k}[\tilde A_{xj}\tilde
866: A_{yj}^*-\tilde A_{yj}\tilde A_{xj}^*].
867: \label{e.ks0}\end{equation}
868: 
869: 
870: The surface coefficients $d$, $b_2$, $b_4$ \cite{SBE}, and
871: $\lambda_{\rm SHV1}$
872: \cite{VolovikMineev} have been estimated before using  simple models
873: for the order parameter.  We calculate them here by solving the order
874: parameter numerically using the Sauls-Serene values
875: \cite{SS} for the coefficients $\beta_i$ \cite{T86}. For pressures
876: below 1.2 MPa we smoothly interpolate the parameters to the
877: weak-coupling values at zero pressure. We also assume the weak-coupling
878: value
879: $\gamma=3$. The calculations are done using boundary conditions
880: appropriate for both specular and diffuse scattering of quasiparticles
881: \cite{AGR}. (In the latter all the components of the order parameter
882: have to vanish at the surface.) The integrals in Equations
883: (\ref{e.d0})-(\ref{e.ks0}) (excluding the prefactors $g_{\rm H}\Delta^2\xi_{\rm GL}$ etc.) are dimensionless, and they are plotted
884: in Figures
885: \ref{f.surfgl} and \ref{f.surfgl2}. Because $b_2>2b_4>0$, $F_{\rm SD}$
886: (\ref{e.2.13}) is minimized by $\hat{\bf n}\parallel\hat{\bf s}$
887: \cite{FominVuorio}.
888: %23456789012345678901234567890123456789012345678901234567890123456789072
889: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
890: \begin{figure}[p]
891: %h=here,t=top,b=bottom,p=erilliselle kuvasivulle
892: \begin{center}\leavevmode
893: \includegraphics[width=0.4\linewidth]{GraphGLSurf.eps}
894: \caption{  The integrals in Equations (\protect\ref{e.d0}),
895: (\protect\ref{e.sg0}), and (\protect\ref{e.ks0}). The solid lines give 
896: $d/g_{\rm H}\Delta^2\xi_{\rm GL}$, dashed  $\lambda_{\rm SG}/K\Delta^2$
897: and dotted 
898: $\lambda_{\rm SHV1}\hbar/2mg_{\rm H1}\Delta^2\xi_{\rm GL}^2$. Out of
899: similar lines, the upper ones always correspond to specular surface
900: scattering and the lower to diffuse scattering.
901: }\label{f.surfgl}\end{center}\end{figure}
902: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
903: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
904: \begin{figure}[p]
905: %h=here,t=top,b=bottom,p=erilliselle kuvasivulle
906: \begin{center}\leavevmode
907: \includegraphics[width=0.4\linewidth]{GraphGLSurf2.eps}
908: \caption{  The integrals for surface-dipole terms in Equations
909: (\protect\ref{e.b2}) and (\protect\ref{e.b4}). The solid lines give 
910: $b_2/g_{\rm D}\Delta^2\xi_{\rm GL}$ and the dashed $b_4/g_{\rm
911: D}\Delta^2\xi_{\rm GL}$. For both quantities, the upper lines
912: correspond to specular surface scattering and the lower to diffuse
913: scattering.  }\label{f.surfgl2}\end{center}\end{figure}
914: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
915:  
916: All the vortex terms except $\lambda_{\rm LG}$ (\ref{e.2.17}) have been
917: calculated in Ref.\
918: \cite{T87}.
919: 
920: All the results of this and the previous section are, of course,
921: identical in the limit where both theories are valid: weak coupling
922: near $T_{\rm c}$.
923: 
924: \section{Determination of parameters}\label{s.gd}
925: 
926: In his section we analyze a few experiments in order to deduce the
927: values of parameters $F_0^{\rm a}$, $F_2^{\rm a}$,
928: and
929: $g_{\rm D}$. We apply trivial strong-coupling
930: corrections to the gap
931: $\Delta$, as explained at the end of Section \ref{s.qc}. For the molar
932: volume $v=mN_{\rm A}/\rho$ as a function of pressure we assume the fit
933: in Ref.\ \cite{Greywall}. 
934: 
935: The parameter $F_0^{\rm a}$ can be obtained from measurements in the
936: normal state: the specific heat and the nuclear susceptibility. For the
937: specific heat  we use the measurements by Greywall \cite{Greywall}. The
938: susceptibility has been measured by Ramm et al \cite{Ramm} and by
939: Hensley et al \cite{Hensley} with essentially identical results.
940: Unfortunately, it has been measured only below 2.9 MPa, and depending
941: on the extrapolation alone, the relative error in $1+F_0^{\rm a}$ may
942: be as large as 10\% at the melting pressure. Examples are the simple
943: fit $F_0^{\rm a}=-0.909+0.0055 v\ {\rm cm}^{-3}$ and the nonmonotonic
944: $F_0^{\rm a}(p)$ fit in Ref.\
945: \cite{HalperinVaroquaux}.
946: 
947: The reduced nuclear susceptibility in the superfluid state,
948: $\chi(T)/\chi_{\rm n}$ (\ref{e.suskis}), has been measured by
949: Corruccini and Osheroff \cite{CO}, by Ahonen et al \cite{AKP}, and by
950: Scholz et al \cite{Hoyt,Scholz}. (There has been a
951: discrepancy between the susceptibility measured by NMR and by a SQUID
952: magnetometer 
953: \cite{HBBG}, but that is probably caused by difficulties in
954: calibration \cite{Chris}.)
955: $\chi(T)/\chi_{\rm n}$ depends only on two parameters,
956: $F_0^{\rm a}$ and $F_2^{\rm a}$. (We treat the trivial strong-coupling
957: corrections as fixed, and use only the low-field limit of the Scholz
958: data.) If we assume
959: $F_0^{\rm a}$ given by Ramm and Hensley et al, only $F_2^{\rm a}$
960: remains to be fitted. We find that a nonzero {\it pressure-independent}
961: value of
962: $F_2^{\rm a}$ does not improve the fit essentially compared to
963: $F_2^{\rm a}\equiv 0$. Since we believe that $F_2^{\rm a}$ cannot have
964: strong pressure dependence, the simple choice $F_2^{\rm a}\equiv 0$
965: seems most attractive to us. Scholz finds $F_2^{\rm a}\approx -1$ with
966: a weak-coupling fit \cite{Scholz}, but this tendency is largely removed
967: by the inclusion of trivial strong-coupling corrections. Note that the
968: susceptibility data could be equally consistent with $F_2^{\rm a}\equiv
969: -0.7$, say, but that would imply a systematic reduction of $F_0^{\rm
970: a}$ by -0.025 from the results by Ramm and Hensley et al. Therefore we
971: take $F_2^{\rm a}\equiv 0$ in the following. For $F_0^{\rm a}$ we use
972: the simple fit given above because it is in better agreement with the
973: reduced susceptibility at the melting pressure \cite{CO} than the fit
974: by Halperin and Varoquaux. Note that at zero pressure our choice is not
975: far off from the relation between $F_0^{\rm a}$ and
976: $F_2^{\rm a}$ based on the "catastrophic relaxation" by Bunkov et al
977: \cite{bunkov}. There exists also other attempts to get $F_2^{\rm
978: a}$ \cite{FS88,Candela,MKL}.  
979: 
980: The dipole constant $g_{\rm D}$ has to be extracted from experiments
981: because its value cannot be calculated accurately in the quasiclassical
982: theory (Sect.\
983: \ref{s.qc}).   The most straightforward way to get $g_{\rm D}$ is to
984: measure the B phase longitudinal NMR frequency $\Omega_\parallel$. A
985: direct measurement of $\Omega_\parallel$ has been made by Bloyet et al
986: \cite{Bloyet} and  Candela et al \cite{Candela}. These experiments were
987: done at low temperatures in the collisionless regime. According to the
988: collisionless theory in a small magnetic field \cite{Fishman}
989: \begin{equation}
990: \Omega_\parallel^2={45\Delta^2g_{\rm D}\over\hbar^2N(0)}
991: \left({1\over\lambda}+{2\over 3}F_0^{\rm a}+{1\over 15}F_2^{\rm
992: a}\right)
993: \label{e.omegallcol}\end{equation} where the function $\lambda(T)$ is
994: defined in the Appendix.
995: An alternative is to extract
996: $\Omega_\parallel$ from transverse NMR frequency in surface-oriented
997: texture. These measurements have been done by Osheroff et al.\
998: \cite{OEBC}, by Ahonen, Krusius, and Paalanen
999: \cite{AKP} (results tabulated in Ref.\
1000: \cite{Wheatley}), by Spencer, Alexander and Ihas \cite{Spencer},
1001: and by Hakonen et al. \cite{HKSSSGVK}. Because the external magnetic
1002: field in these experiments reduces the frequency difference between
1003: normal and superfluid precession, these experiments have to be analyzed
1004: using hydrodynamic theory. There $\Omega_\parallel$ is related to
1005: $g_{\rm D}$ via \cite{LeggettAnn}
1006: \begin{equation}
1007: \Omega_\parallel^2={15\mu_0\gamma^2\Delta^2g_{\rm D}\over\chi}.
1008: \label{e.omegall}\end{equation}  Both relations (\ref{e.omegallcol})
1009: and (\ref{e.omegall}) imply that the temperature dependence of
1010: $\Omega_\parallel$ is fully determined by the energy gap
1011: $\Delta$ and the $\lambda$ function (\ref{e.lambda}) or the
1012: susceptibility $\chi$ (\ref{e.suskis}). 
1013: 
1014: A third way to get $g_{\rm D}$ is so-called g shift of the transverse
1015: NMR frequency $\omega$ from the Larmor frequency $\omega_0$. According
1016: to Ref.\ \cite{SBE}
1017: \begin{equation} {\omega-\omega_0\over\omega_0}={4\mu_0a\over 5\chi}.
1018: \label{e.gshift}\end{equation} The g shift was measured by Osheroff
1019: \cite{Osheroff} at the melting pressure and by Kycia et al
1020: \cite{gshift} below 2.17 MPa. [We note that these measurements are done
1021: in such a high field that the expressions given in this paper are no
1022: more reliable near $T_{\rm c}$. However, Kycia et al measure the
1023: g-shift as a function of magnetization, and this plot is found field
1024: independent, both experimentally and theoretically \cite{gshift}. The
1025: only consequence is that the temperature $T$ in Fig.\
1026: \ref{f.GdTemperature} is the temperature according to the weak-field
1027: susceptibility (\ref{e.suskis}), which differs from the true
1028: temperature near $T_{\rm c}$.]
1029: 
1030: We plot $g_{\rm D}$ obtained by all three methods in Fig.\
1031: \ref{f.GdTemperature}.  
1032: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1033: \begin{figure}[p]
1034: %h=here,t=top,b=bottom,p=erilliselle kuvasivulle
1035: \begin{center}\leavevmode
1036: \includegraphics[width=0.5\linewidth]{GdTAll.eps}
1037: \caption[GdTemp]{ Plotted as a function of temperature, $g_{\rm D}$
1038: interpreted from different experiments according to trivial
1039: strong-coupling model. In ideal case $g_{\rm D}$ should be independent
1040: of temperature. We have used the transverse NMR frequency measured by
1041: Osheroff et al.\
1042: \protect\cite{OEBC}  ($\blacktriangle$ 3.44 MPa), by Ahonen, Krusius,
1043: and Paalanen
1044: \protect\cite{AKP}  ($\lozenge$ 3.2 MPa, 
1045: $\blacktriangledown$ 2.9 MPa,
1046: $\triangledown$ 2.54 MPa, 
1047: $\blacksquare$ 2.11 MPa 
1048: $+$ 1.87 MPa,), and by Hakonen et al.\ \protect\cite{HKSSSGVK} 
1049: ($\bullet$ 2.5 MPa,
1050: $\circ$ 1.55 MPa, 
1051: $\blacklozenge$ 1.02 MPa,
1052: $\times$ 0.5 MPa, 
1053: $\vartriangle$ 0.05 MPa). We also used the longitudinal NMR frequency
1054: measured by Candela et al.\
1055: \protect\cite{Candela} ($\square$, from top to bottom 3.3, 2.1, 1.2,
1056: 0.61, and 0.03 MPa).  For comparison, there are curves that are
1057: determined from the g-shift measurements by Osheroff \cite{Osheroff}
1058: (curve with label $\times 5$, reduced by factor ${1/5}$) and by Kycia
1059: et al
1060: \cite{gshift} (curves, from top to bottom 2.17, 1.3, 0.7, 0.3, and 0.11
1061: MPa)   
1062:  }\label{f.GdTemperature}\end{center}\end{figure}
1063: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1064: Let us first ignore the g-shift data (lines). It can be seen that the
1065: data for
1066: $g_{\rm D}$ at each pressure is almost independent of temperature, as
1067: required by theory. We note that in order to reach this constancy it
1068: really is necessary that the longitudinal and transverse data of
1069: $\Omega_\parallel$ are analyzed with collisionless and hydrodynamic
1070: theories, respectively. Equally important is that we use trivial
1071: strong-coupling corrections. The value of $g_{\rm D}$ (but not the
1072: temperature dependence) also depends on
1073: $F_1^{\rm s}$ and $T_{\rm c}$, for which we use the measurements by
1074: Greywall \cite{Greywall}. 
1075: 
1076: The $g_{\rm D}$ data as a
1077: function of pressure is plotted in Fig.\ \ref{f.GdPressure}. It
1078: contains all the same data as Fig.\
1079: \ref{f.GdTemperature} and some additional data.  
1080: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1081: \begin{figure}[p]
1082: %h=here,t=top,b=bottom,p=erilliselle kuvasivulle
1083: \begin{center}\leavevmode
1084: \includegraphics[width=0.45\linewidth]{GdPAll.eps}
1085: \caption[GdPres]{ Plotted as a function of pressure, $g_{\rm D}$
1086: interpreted from different experiments according to the trivial
1087: strong-coupling model. The data from $\Omega_\parallel$ in transverse
1088: NMR ($+$), from longitudinal NMR by Candela et al ($\square$), and
1089: g shift (bars) are the same as in Fig.\
1090: \protect\ref{f.GdTemperature}. In addition there is data from zero-field
1091: longitudinal NMR measurements by Bloyet et al \protect\cite{Bloyet}
1092: ($\lozenge$) and transverse surface-oriented NMR measurements at
1093: saturated vapor pressure in the temperature range $T/T_{\rm c}=
1094: 0.83\ldots 0.86$ by Spencer et al \cite{Spencer} ($\times$). The solid
1095: and dashed lines approximate the data from longitudinal NMR and g shift,
1096: receptively. The dotted line is the model of Eq.\ (\ref{e.gdsimple})
1097: with 
1098: $\epsilon_c=0.45 k_{\rm B}T_{\rm F}$. It is argued in the text that 
1099: $g_{\rm D}$ given by $\Omega_\parallel$ is the correct one, and the
1100: difference between $g_{\rm D}$'s deduced from the g shift and
1101: $\Omega_\parallel$ is a measure of nontrivial strong-coupling 
1102: corrections.
1103:  }\label{f.GdPressure}\end{center}\end{figure}
1104: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1105: It can be seen that the longitudinal and transverse
1106: measurements of $\Omega_\parallel$ agree very well at intermediate
1107: pressures, but there is a difference at both high and low pressures. 
1108: Theoretically the ratio of collisionless and hydrodynamic
1109: $\Omega_\parallel$ (at the same pressure) depends only on the reduced
1110: temperatures
1111: $T_1/T_{\rm c}$ and $T_2/T_{\rm c}$ of the two measurements and on
1112: $F_0^{\rm a}$, $F_2^{\rm a}$, and $\Delta C/C_{\rm n}$. The difference
1113: in the effective
1114: $g_{\rm D}$ obtained by the two methods seems at low pressures much
1115: larger than the expected uncertainties of the parameters, and 
1116: remains unexplained. Both the positive slope of $g_{\rm D}(T)$ (Fig.\
1117: \ref{f.GdTemperature}) at high pressures and the difference between
1118: transverse and longitudinal data could be reduced by
1119: giving $F_2^{\rm a}$ a negative value
1120: ($\approx -0.7$ at high pressures), but only at the expense of impaired
1121: fit of the $\chi(T)/\chi_{\rm n}$.    
1122: 
1123: We believe that the nontrivial strong coupling
1124: corrections are small in the expressions for
1125: $\Omega_\parallel$ (\ref{e.omegallcol})-(\ref{e.omegall}), which are
1126: based on expectation values in an unperturbed order parameter. [There
1127: are nontrivial corrections to $\chi$
1128: \cite{SRQ77}, but as long as (\ref{e.suskis}) can reproduce (possibly
1129: with incorrect $F_2^{\rm a}$) the measured $\chi$,  the value obtained
1130: for $g_{\rm D}$ is unaffected.]  The collisionless $\Omega_\parallel$
1131: data is fitted by solid line in Fig.\ \ref{f.GdPressure}. The values
1132: obtained for 
1133: $g_{\rm D}$ depend on $F_0^{\rm a}$, $T_{\rm c}$, $N(0)$, $F_2^{\rm
1134: a}$, and
1135: $\Delta C/C_{\rm n}$, and must be revised if more accurate values of
1136: these become available, for example, via improved measurement of the
1137: temperature \cite{Solen}.
1138: 
1139: We can also estimate $g_{\rm D}$
1140: based on the simple model of Eq.\ (\ref{e.gd}). Assuming
1141: $\bar{R^2}=1$ and making the sum at
1142: $T=T_{\rm c}$ gives  
1143: \begin{equation} g_{\rm D}={\mu_0\over 40}\left(\hbar\gamma
1144: N(0)\ln{1.1339\epsilon_c\over k_{\rm B}T_{\rm c}}\right)^2.
1145: \label{e.gdsimple}\end{equation}
1146: We take the cut-off energy $\epsilon_c$ proportional to the Fermi
1147: temperature defined by $T_{\rm F}=3\rho/4N(0)k_{\rm B}m$. As shown by
1148: dotted line in Fig.\
1149: \ref{f.GdPressure}, the resulting expression fits nicely the
1150: experimental data for the constant of proportionality 
1151: $\epsilon_c/k_{\rm B}T_{\rm F}=0.45$. The
1152: agreement may be accidental, however, because there is no fundamental
1153: justification for the approximations made. 
1154: 
1155: Let us next study the $g_{\rm D}$ data based on the g-shift. It is also 
1156: rather independent of the temperature (lines in
1157: Fig.\ \ref{f.GdTemperature}). Fig.\ \ref{f.GdPressure} shows that the
1158: $g_{\rm D}$ data (bars) at different pressures are well consistent with
1159: each other: the low pressure data extrapolates well to the melting
1160: pressure data by Osheroff. However,
1161: $g_{\rm D}$ deduced from the g shift differs essentially from the
1162: determinations based on $\Omega_\parallel$, especially at high
1163: pressures. The reason for this is that the expression for $a$
1164: (\ref{e.3.1}) has substantial strong-coupling corrections that are not
1165: included in the scaling of the energy gap $\Delta$. This can be seen by
1166: comparing the
1167: $T\rightarrow T_{\rm c}$ limit of trivial strong-coupling $a$
1168: (\ref{e.3.1}), denoted by
1169: $a_{\rm TSC}$, with the Ginzburg-Landau limit $a_{\rm GL}$
1170: (\ref{e.4.1}). We find
1171: \begin{equation} {a_{\rm GL}\over a_{\rm TSC}}={a_{\rm GL}\over a_{\rm
1172: WC}}={g_{\rm H}\over g_{\rm H}^{\rm WC}} {\beta_{345}^{\rm WC}\over
1173: \beta_{345}}. 
1174: \label{e.nontrivial}\end{equation} where WC denotes weak-coupling. It
1175: is well known that $\beta_{345}$ differs substantially from its weak
1176: coupling value \cite{SS,THBG,gshift,HBBG}. This explains the difference
1177: in the apparent $g_{\rm D}$ deduced from g shift and $\Omega_\parallel$.
1178: The new thing in the present analysis compared to Ref.\
1179: \cite{gshift} is that the difference is not limited to
1180: the Ginzburg-Landau region, but because of the weak dependency of the
1181: apparent
1182: $g_{\rm D}$ on temperature (Fig.\ \ref{f.GdTemperature}), it persists
1183: almost unchanged at all temperatures. 
1184: 
1185: We conclude this section with a comparison
1186: of theoretical and experimental dipole velocity $v_{\rm D}$.
1187: Theoretically this quantity is related to coefficients $a$
1188: (\ref{e.3.1}) and
1189: $\lambda_{\rm HV}$ (\ref{e.3.4}) by the relation
1190: $v_{\rm D}^2=2a/(5\lambda_{\rm HV})$. It has been measured by Nummila
1191: et al \cite{nummila}. Originally they compared their result to a theory
1192: that turned out to be in error, see discussion in Refs.\
1193: \cite{ghv,GKV}. The comparison with the present theory is given in
1194: Fig.\ \ref{f.vd}. We have used trivial strong-coupling theory,
1195: parameters as described above and 
1196: $g_{\rm D}$ from solid line in Fig.\ \ref{f.GdPressure}.  
1197: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1198: \begin{figure}[p]
1199: %h=here,t=top,b=bottom,p=erilliselle kuvasivulle
1200: \begin{center}\leavevmode
1201: \includegraphics[width=0.45\linewidth]{vd.eps}
1202: \caption[vd]{ The dipolar velocity $v_{\rm D}$. The circles are
1203: experimental data from Ref.\ \protect\cite{nummila}. The solid lines are
1204: theory based on Equations (\ref{e.3.1}) and (\ref{e.3.4}), where all
1205: parameter values are fixed by other measurements as explained in the
1206: text. 
1207:  }\label{f.vd}\end{center}\end{figure}
1208: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1209: 
1210: Because both $a$ (\ref{e.4.1})
1211: and $\lambda_{\rm HV}$ (\ref{e.4.3}) are proportional to
1212: $\beta_{345}^{-1}$ in the Ginzburg-Landau region, the uncertainty
1213: discussed in connection with $a$ is expected to cancel out in $v_{\rm
1214: D}$. There also exists a direct measurement of $\lambda_{\rm HV}$
1215: \cite{ghv}. It also shows deviation from the trivial strong-coupling
1216: model, but the differences are not of similar type as for $a$, and are
1217: presently not understood.
1218: 
1219: Above we have discussed all input parameters of the
1220: trivial-strong-coupling hydrostatic theory except $F^{\rm a}_1$ and
1221: $F^{\rm a}_3$. Out of the bulk terms only the gradient coefficients
1222: (\ref{e.3.5}) and (\ref{e.3.6})
1223: depend on these. There is several independent evidence that $F^{\rm
1224: a}_1\approx -1$ at high pressures \cite{ORSB,Greywall83,VSexp} but
1225: $F^{\rm a}_3$ is not known. 
1226:  
1227: \section{Conclusions}\label{s.exp}
1228: 
1229: We have presented a  summary of the hydrostatic theory in superfluid
1230: $^3$He-B. Several new analytic and numerical results were included.
1231: Some experimental data was analyzed in order to extract the
1232: parameters of the theory. A particular goal was to understand how
1233: well the B phase is described by the trivial strong-coupling
1234: model. We found that some quantities ($\chi$,
1235: $g_{\rm D}$, $v_{\rm D}$) can successfully be calculated, but there are
1236: other quantities ($a$, $\lambda_{\rm HV}$) that may be wrong by
1237: 50\% in this model. The parameter $g_{\rm D}$ has direct relevance also
1238: for the A phase, where it has been used in comparison between theory and
1239: experiment (see Ref.\ \cite{KT}, for example).   
1240: 
1241: The first application of the results
1242: calculated here has been the comparison of the ratio
1243: $d/a$ [from equations (\ref{e.d0}) and (\ref{e.4.1})] to experiments in
1244: Ref.\ \cite{KGJKKMT}.  More recently, Kopu et al \cite{kopu} 
1245: applied the hydrostatic theory to a
1246: rotating cylindrical container. They calculated the NMR
1247: line shape and studied the optimal conditions for observing single
1248: vortex lines. Another application is the Josephson $\pi$ state 
1249: observed recently \cite{packard2}. This effect was found to depend
1250: crucially on the texture at the Josephson junction \cite{VTlett,VT}. 
1251: The texture is also essential in several experiments of superfluid 
1252: $^3$He in aerogel. For example, the identification of the B phase was
1253: based on its texture-dependent NMR spectrum \cite{AKWRNH}. For the
1254: present, the textural parameters in aerogel have been evaluated only in
1255: the Ginzburg-Landau region in the homogeneous scattering model
1256: \cite{verditz}. These developments demonstrate that there still are
1257: open problems in superfluid $^3$He and in many cases a proper
1258: understanding of the hydrostatic theory is a prerequisite for
1259: solving them.
1260: 
1261: \section*{Acknowledgments}
1262: 
1263: I thank the Academy of Finland for financial support.  
1264:  
1265: \section*{Appendix}
1266: 
1267: We give here some equations that complete the theory presented above.
1268: The weak-coupling energy gap $\Delta(T)$ is determined by the equation
1269: \begin{equation} 
1270: \ln{T\over T_{\rm c}}+\pi
1271: T\sum_{n=-\infty}^\infty\left[{1\over\vert\epsilon_n\vert}
1272: -{1\over\sqrt{\epsilon_n^2+\Delta^2}}\right]=0,
1273: \label{e.qcgap}\end{equation}
1274: where the
1275: Matsubara energies $\epsilon_n=\pi T(2n-1)$  with $n=0, \pm1,
1276: ...\pm\infty$. The $Z_j(T)$ functions are defined by 
1277: \begin{equation} Z_j=\pi k_{\rm B}T\Delta^{j-1}\sum_{n=-\infty}^\infty
1278: (\epsilon_n^2+\Delta^2)^{-j/2},
1279: \label{e.3.2}\end{equation} 
1280: $Y(T)=1-Z_3(T)$. The $\lambda(T)$ function \cite{Fishman}, which
1281: equals to
1282: $1-f(T)$ defined in Ref.\ \cite {LT}, can be written as 
1283: \begin{equation}
1284: \lambda=\pi k_{\rm B}T\sum_{n=-\infty}^\infty
1285: {\Delta\over\sqrt{\epsilon_n^2+\Delta^2}(\sqrt{\epsilon_n^2+\Delta^2}
1286: +\Delta)}.
1287: \label{e.lambda}\end{equation}
1288: The numerical calculation of the functions is discussed in Ref.\
1289: \cite{bcsgap}.
1290: 
1291: The gradient
1292: energies (\ref{e.2.9}) and (\ref{e.2.12}) can be written in different
1293: forms using the identities \cite{SBE}
1294: \begin{eqnarray} \partial_i R_{\alpha j}\partial_i R_{\alpha j}
1295: &=&4(1-\cos\theta)(\partial_i \hat n_j)^2
1296: =4(1-\cos\theta)\{(\nabla\times\hat{\bf n})^2+(\nabla\cdot\hat{\bf n})^2
1297:  +\nabla\cdot[(\hat{\bf n}\cdot\nabla)\hat{\bf n}
1298:  -\hat{\bf n}(\nabla\cdot\hat{\bf n})]\}\\
1299: \partial_i R_{\alpha i}\partial_j R_{\alpha j}
1300: &=&(1-\cos\theta)[2(\nabla\times\hat{\bf n})^2
1301:  +(1-\cos\theta)(\nabla\cdot\hat{\bf n})^2
1302:  -(1-\cos\theta)(\hat{\bf n}\cdot\nabla\times\hat{\bf n})^2
1303:  -2\sin\theta(\nabla\cdot\hat{\bf n})(\hat{\bf n}\cdot\nabla\times\hat{\bf n})]\\
1304: \partial_i R_{\alpha j}\partial_j R_{\alpha i}
1305: &=&(1-\cos\theta)\{2(\nabla\times\hat{\bf n})^2
1306:  +(1-\cos\theta)(\nabla\cdot\hat{\bf n})^2
1307:  -(1-\cos\theta)(\hat{\bf n}\cdot\nabla\times\hat{\bf n})^2\nonumber\\&&
1308:  -2\sin\theta(\nabla\cdot\hat{\bf n})(\hat{\bf n}\cdot\nabla\times\hat{\bf n})
1309:  +2\nabla\cdot[(\hat{\bf n}\cdot\nabla)\hat{\bf n}
1310:  -\hat{\bf n}(\nabla\cdot\hat{\bf n})]\}\\
1311: \hat s_i R_{\alpha i}\partial_j R_{\alpha j}
1312: &=&-(1-\cos\theta)\hat{\bf s}\cdot[(\hat{\bf n}\cdot\nabla)\hat{\bf n}
1313:  -\hat{\bf n}(\nabla\cdot\hat{\bf n})]\}.
1314: \label{e.drdr}\end{eqnarray} 
1315: Using these it can be seen that $F_{\rm G}$ (\ref{e.2.9}) has a pure
1316: divergence term $\propto\nabla\cdot[(\hat{\bf n}\cdot\nabla)\hat{\bf n}
1317:  -\hat{\bf n}(\nabla\cdot\hat{\bf n})]$. The prefactor of this term is
1318: half of the value of that by Smith, Brinkman and Engelsberg \cite{SBE}.
1319: With present definitions the other half is transferred to the surface
1320: term (\ref{e.2.12}).  
1321: 
1322: \begin{thebibliography}{99}
1323: \bibitem{LounasT} O.V. Lounasmaa and E.V. Thuneberg, Proc. Natl. Acad.
1324: Sci. USA {\bf 96}, 7760 (1999).
1325: 
1326: \bibitem{aerogel} A. Golov, J.V. Porto, D.A. Geller, N. Mulders, G.J.
1327: Lawes, and J.M. Parpia, Physica B {\bf 280}, 134 (2000).
1328: 
1329: \bibitem{packard2} A. Marchenkov, R.W. Simmonds, S. Backhaus, A.
1330: Loshak, J.C. Davis, and R.E. Packard, Phys. Rev. Lett. {\bf 83}, 3860
1331: (1999).
1332: 
1333: \bibitem{LLfluid} L.D. Landau and E.M. Lifshitz, {\it Fluid Mechanics}
1334: (Pergamon Press, Oxford, 1987).
1335: 
1336: \bibitem{Leggett} A.J. Leggett, Rev. Mod. Phys. {\bf 47}, 331
1337: (1975).
1338:  
1339: \bibitem{BC} W.F. Brinkman and M.C. Cross, in {\it Progress in Low
1340: Temperature Physics, Vol VIIA}, ed. D.F. Brewer (North Holland, 1978),
1341: p. 105.
1342:  
1343: \bibitem{SRrev} J.W. Serene and D. Rainer, Phys. Rep. {\bf 101}, 221
1344: (1983).
1345: 
1346: \bibitem{Fetterrev} A.L. Fetter, in {\it Progress in Low Temperature
1347: Physics, Vol. X}, ed. D.F. Brewer (Elsevier, 1986), p. 1.
1348: 
1349: \bibitem{SVrev} M.M. Salomaa and G.E. Volovik, Rev. Mod. Phys. {\bf
1350: 59}, 533 (1987), {\bf 60}, 573 (1988).
1351: 
1352: \bibitem{Kharadze} G.A. Kharadze, in {\it Helium Three}, ed. W.P.
1353: Halperin and L.P. Pitaevskii (Elsevier, Amsterdam 1990), p. 167. 
1354: 
1355: \bibitem{VW} D. Vollhardt and P. W\"olfle, {\it The superfluid phases
1356: of helium 3} (Taylor \& Francis, London 1990).
1357: 
1358: \bibitem{BCS} J. Bardeen, L.N. Cooper, and J.R. Schrieffer, Phys. Rev.
1359: {\bf 108}, 1175 (1957).
1360: 
1361: \bibitem{WheatleyRev} J.C. Wheatley, Rev. Mod.
1362: Phys. {\bf 47}, 415 (1975).
1363: 
1364: \bibitem{SEP} G.W. Swift, J.P. Eisenstein, and R.E. Packard, Phys. Rev.
1365: Lett. {\bf 45}, 1955 (1980). 
1366: 
1367: \bibitem{LeggettAnn} A.J. Leggett, Ann. Phys. (New York) {\bf 85}, 11
1368: (1974).
1369: 
1370: \bibitem{EBA} S. Engelsberg, W.F. Brinkman, and P.W. Anderson, Phys.
1371: Rev. A {\bf 9}, 2592 (1974).
1372: 
1373: \bibitem{SBE} H. Smith, W.F.
1374: Brinkman, and S. Engelsberg, Phys. Rev. B{\bf 15}, 199 (1977).
1375:  
1376: \bibitem{BSOB} W.F. Brinkman, H. Smith, D.D. Osheroff, and E.I. Blount,
1377: Phys. Rev. Lett {\bf 33}, 624 (1974).
1378: 
1379: \bibitem{VolovikMineev} G.E. Volovik and V.P. Mineev, Zh. Eksp. Teor.
1380: Fiz. {\bf 86}, 1667 (1984) [Sov. Phys. JETP {\bf 59}, 972 (1984)].
1381:  
1382: \bibitem{T86} E.V. Thuneberg, Phys. Rev. B {\bf 33}, 5124 (1986).
1383:  
1384: \bibitem{FominVuorio} I.A. Fomin and M. Vuorio, J. Low Temp. Phys. {\bf
1385: 21}, 271 (1975).
1386: 
1387: \bibitem{ZKT}W. Zhang, J. Kurkij\"arvi, and E.V. Thuneberg, Phys. Rev.
1388: B{\bf 36}, 1987 (1987).
1389:  
1390: \bibitem{GGK} A.D. Gongadze, G.E. Gurgenishvili, and G.A. Kharadze,
1391: Fiz. Nizk. Temp. {\bf 7}, 821 (1981) [Sov. J. Low Temp. Phys. {\bf 7},
1392: 397 (1981)].
1393:  
1394: \bibitem{HKSSBMV} P.J. Hakonen, M. Krusius, M.M. Salomaa, J.T. Simola,
1395: Yu.M. Bunkov, V.P. Mineev, and G.E. Volovik, Phys. Rev. Lett. {\bf 51},
1396: 1362 (1983).
1397: 
1398: \bibitem{T87} E.V. Thuneberg, Phys. Rev. B {\bf 36}, 3583 (1987).
1399:  
1400: \bibitem{SRres} J.W. Serene and D. Rainer, Phys. Rev. B {\bf 17}, 2901
1401: (1978).
1402: 
1403: \bibitem{SRQ77} J.W. Serene and D. Rainer, in {\it Quantum fluids and
1404: solids}, eds. S.B. Trickey, E.D. Adams, and J.W. Dufty (Plenum, New
1405: York, 1977), p. 111.
1406: 
1407: \bibitem{FS} R.S. Fishman and J.A. Sauls, Phys. Rev. B {\bf 33}, 6068
1408: (1986).
1409: 
1410: \bibitem{ghv} J.S. Korhonen, Yu.M. Bunkov, V.V. Dmitriev, Y. Kondo, M.
1411: Krusius, Yu.M. Mukharskiy, \"U. Parts, and E.V. Thuneberg, Phys. Rev. B
1412: {\bf 46}, 13983 (1992).
1413: 
1414: \bibitem{Mineev} V.P. Mineev, Zh. Eksp. Teor. Fiz. {\bf 90}, 1236
1415: (1986) [Sov. Phys. JETP {\bf 63}, 721 (1986)].
1416:  
1417: \bibitem{Dorfle} M. D\"orfle, Phys. Rev. B {\bf 23}, 3267 (1981); note
1418: incorrect powers of 2 in equations (5.3) and (5.4).
1419:  
1420: \bibitem{Buchholtz} L.J. Buchholtz, Phys. Rev. B {\bf 33}, 1579 (1986).
1421: 
1422: \bibitem{VT} J. Viljas and E.V. Thuneberg, to be published. 
1423:  
1424: \bibitem{FogelstromKurkijarvi} M. Fogelstr\"om and J. Kurkij\"arvi, J.
1425: Low Temp. Phys. {\bf 98}, 195 (1995); erratum {\bf 100}, 597 (1995).
1426: 
1427: \bibitem{Greywall} D.S. Greywall, Phys. Rev. B {\bf 33}, 7520 (1986).
1428: 
1429: \bibitem{GL} V.L. Ginzburg and L.D. Landau, Zh. Eksp. Teor. Fiz. {\bf
1430: 20}, 1064 (1950).
1431: 
1432: \bibitem{MerminStare}  N.D. Mermin and C. Stare, Phys. Rev.
1433: Lett. {\bf 30}, 1135 (1973).
1434: 
1435: \bibitem{AGR} V. Ambegaokar, P.G. deGennes, and D. Rainer, Phys. Rev. A
1436: {\bf 9}, 2676 (1974), {\bf 12}, 345 (1975).
1437:  
1438: \bibitem{AmbeM} V. Ambegaokar and N.D. Mermin, Phys. Rev. Lett. {\bf
1439: 30}, 81 (1973).
1440: 
1441: \bibitem{F75}  A.L. Fetter, in {\it Quantum Statistics and the
1442: Many-Body Problem}, eds. S.B. Trickey, W.P. Kirk, and J.W. Dufty
1443: (Plenum, New York, 1975), p. 127.
1444: 
1445: \bibitem{THBG}  Y.H. Tang, I. Hahn, H.M. Bozler, and C.M. Gould, Phys.
1446: Rev. Lett. {\bf 67}, 1775 (1991).
1447: 
1448: \bibitem{gshift} J.B. Kycia, T.M. Haard, M.R. Rand, H.H. Hensley, G.F.
1449: Moores, Y. Lee, P.J. Hamot, D.T. Sprague, W.P. Halperin, and E.V.
1450: Thuneberg, Phys. Rev. Lett. {\bf 72}, 864 (1994).
1451: 
1452: \bibitem{HBBG}  I. Hahn, S.T.P. Boyd, H.M. Bozler, and C.M. Gould,
1453: Phys. Rev. Lett. {\bf 81}, 618 (1998).
1454: 
1455: \bibitem{RSbeta} D. Rainer and J.W.
1456: Serene, Phys. Rev. B {\bf 13}, 4745 (1976).
1457: 
1458: \bibitem{RSspecific} D. Rainer and J.W. Serene, J. Low Temp. Phys. {\bf
1459: 38}, 601 (1980).
1460: 
1461: \bibitem{SS} J.A. Sauls and J.W. Serene, Phys. Rev. B {\bf 24},
1462: 183 (1981).
1463:  
1464: \bibitem{Bedell} K. Bedell, Phys. Rev. B {\bf 26}, 3747
1465: (1982).
1466: 
1467: \bibitem{LevinValls} K. Levin and O.T. Valls, Phys. Rep. {\bf
1468: 98}, 1 (1983).
1469: 
1470: \bibitem{Sagan} D.C. Sagan, P.G.N. deVegvar, E. Polturak, L. Friedman,
1471: S.-S. Yan, E.L. Ziercher and D.M. Lee, Phys. Rev. Lett. {\bf 53}, 1939
1472: (1984).
1473: 
1474: \bibitem{Israelsson}  U.E. Israelsson, B.C. Crooker, H.M. Bozler
1475: and C.M. Gould, Phys. Rev. Lett. {\bf 53}, 1943 (1984).
1476: 
1477: \bibitem{Ramm} H. Ramm, P. Pedroni, J.R. Thompson, and
1478: H. Meyer, J. Low Temp. Phys. {\bf 2}, 539 (1970).
1479: 
1480: \bibitem{Hensley}  H.H. Hensley, Y. Lee, P. Hamot, T. Mizusaki, and W.P.
1481: Halperin, J. Low Temp. Phys. {\bf 90}, 149 (1993).
1482: 
1483: \bibitem{HalperinVaroquaux} W.P. Halperin and E. Varoquaux, in {\it
1484: Helium Three}, ed. W.P. Halperin and L.P. Pitaevskii (Elsevier,
1485: Amsterdam 1990), p.\ 353.
1486: 
1487: \bibitem{CO} L.R. Corruccini and D.D. Osheroff, Phys. Rev. B {\bf 17},
1488: 126 (1978).
1489: 
1490: \bibitem{AKP} A.I. Ahonen, M. Krusius, and M.A. Paalanen, J. Low Temp.
1491: Phys. {\bf 25}, 421 (1976).
1492: 
1493: \bibitem{Hoyt} R.F. Hoyt, H.N. Scholz, and D.O. Edwards, Physica B {\bf
1494: 107}, 287 (1981).
1495: 
1496: \bibitem{Scholz} H.N. Scholz, Thesis (Ohio State University, 1981). 
1497: 
1498: \bibitem{Chris} C.M. Gould, private communication.
1499: 
1500: \bibitem{bunkov} Yu.M. Bunkov,
1501: S.N. Fisher, A.M. Gu\'enault, C.J. Kennedy, and G.R. Pickett, J. Low
1502: Temp. Phys. {\bf 89}, 27 (1992).
1503: 
1504: \bibitem{FS88} R.S. Fishman and J.A. Sauls, Phys. Rev. B {\bf 38}, 2526
1505: (1988).
1506: 
1507: \bibitem{Candela} D. Candela, D.O. Edwards, A.
1508: Heff, N. Masuhara, Y. Oda, and D.S. Sherrill, Phys. Rev. Lett. {\bf
1509: 61}, 420 (1988). 
1510: 
1511: \bibitem{MKL} R. Movshovich, N. Kim, and D.M. Lee, Phys. Rev.
1512: Lett. {\bf 64}, 431 (1990).
1513: 
1514: \bibitem{Bloyet} D. Bloyet, E. Varoquaux, C. Vibet, O. Avenel, P.M.
1515: Berglund, and R. Combescot, Phys. Rev. Lett. {\bf 42}, 1158 (1979); C.
1516: Vibet, thesis, L'Universit\'e Paris-Sud, Centre d'Orsay, 1979
1517: (unpublished).
1518: 
1519: \bibitem{Fishman} R.S. Fishman, Phys. Rev. B {\bf 36}, 79 (1987).
1520: 
1521: \bibitem{OEBC} D.D. Osheroff, S. Engelsberg, W.F. Brinkman, and L.R.
1522: Corruccini, Phys. Rev. Lett. {\bf 34}, 190 (1975).
1523: 
1524: \bibitem{Wheatley} J.C. Wheatley, in {\it Progress in Low Temperature
1525: Physics, Volume VIIA}, ed. D.F. Brewer (North-Holland, 1978), p. 1.
1526: 
1527: \bibitem{Spencer} G.F. Spencer, P.W. Alexander, and G.G. Ihas, Physica
1528: B {\bf 107}, 289 (1981).
1529: 
1530: \bibitem{HKSSSGVK} P.J. Hakonen, M. Krusius, M.M. Salomaa, R.H. 
1531: Salmelin, J.T. Simola, A.D. Gongadze, G.E. Vachnadze, and G.A. 
1532: Kharadze, J. Low Temp. Phys. {\bf 76}, 225 (1989). 
1533: 
1534: \bibitem{Osheroff} D.D. Osheroff, unpublished, g-shift
1535: $\approx (0.8373-1.1677t+1.8736t^2)\cdot 10^{-5}$ for $T/T_{\rm
1536: c}\equiv t$ in the range 0.39-0.74 at the melting pressure.  
1537: 
1538: \bibitem{Solen} R.J. Solen Jr. and W.E. Fogle, Physics Today {\bf 50},
1539: no. 8 Part 1, 36 (1997).
1540: 
1541: \bibitem{nummila} K.K. Nummila, P.J. Hakonen, and O.V. Magradze,
1542: Europhys. Lett. {\bf 9}, 355 (1989).
1543: 
1544: \bibitem{GKV} A. Gongadze, G. Kharadze, and G. Vachnadze, Fiz. Nizk.
1545: Temp. {\bf 23}, 546 (1997) [Low Temp. Phys. {\bf 23}, 405 (1997)].
1546: 
1547: \bibitem{ORSB} D.D. Osheroff, W. van Roosbroeck, H. Smith, and W.F.
1548: Brinkman, Phys. Rev. Lett. {\bf 38}, 134 (1977).
1549: 
1550: \bibitem{Greywall83}
1551: D.S. Greywall, Phys. Rev. B {\bf 27}, 2747 (1983).
1552: 
1553: \bibitem{VSexp} \"U. Parts, V.M.H. Ruutu, J.H. Koivuniemi, M. Krusius,
1554: E.V. Thuneberg, and G.E. Volovik, Physica B {\bf 210}, 311 (1995).
1555: 
1556: \bibitem{KT} J.M. Karim\"aki and E.V. Thuneberg, Phys. Rev. B {\bf 60},
1557: 15290 (1999).
1558: 
1559: \bibitem{KGJKKMT} J.S. Korhonen, A.D. Gongadze, Z. Jan\'u, Y. 
1560: Kondo, M. Krusius, Yu.M. Mukharsky, and E.V. Thuneberg, Phys.  Rev.
1561: Lett. {\bf 65}, 1211 (1990).
1562: 
1563: \bibitem{kopu} J. Kopu, R. Schanen, R. Blaauwgeers, V.B.
1564: Eltsov, M. Krusius, J.J. Ruohio, and E.V. Thuneberg, J. Low Temp. Phys.
1565: {\bf 120}, 213 (2000).
1566: 
1567: \bibitem{VTlett} J.K. Viljas and E.V. Thuneberg, Phys. Rev. Lett. {\bf
1568: 83}, 3868 (1999).
1569: 
1570: \bibitem{AKWRNH} H. Alles, J.J. Kaplinsky, P.S. Wootton, J.D. Reppy,
1571: J.H. Naish, and J.R. Hook, Phys. Rev. Lett. {\bf 83}, 1367
1572: (1999).
1573: 
1574: \bibitem{verditz} E.V. Thuneberg, in {\it Quasiclassical Methods in
1575: Superconductivity and Superfluidity, Verditz 96}, ed. D. Rainer and
1576: J.A. Sauls (1998) p. 53; cond-mat/9802044.
1577: 
1578: \bibitem{LT} A.J. Leggett and S. Takagi, Ann. Phys. (New
1579: York) {\bf 106}, 79 (1977).
1580: 
1581: \bibitem{bcsgap}http://boojum.hut.fi/research/theory/qc/bcsgap.html
1582: 
1583: \end{thebibliography}
1584: \end{document}
1585:  
1586:  
1587: