1: \section{Weakly Interacting Bose Gas at Zero Temperature}
2: \label{t0}
3: In this section, we discuss in some detail the weakly interacting Bose gas
4: at $T=0$. We begin with a description using effective field theory
5: and formulate a perturbative framework that can be used
6: for practical calculations.
7: We then calculate the leading corrections in the
8: low-density expansion to the energy density, depletion, and
9: long-wavelength excitations. Finally, we
10: discuss nonuniversal effects.
11:
12:
13: \subsection{Effective Field Theory}
14: \label{eftt}
15: Effective field theory (EFT) is a general approach that can be used to analyze
16: the low-energy behavior of a physical system in a systematic
17: way~\cite{eft,lep2,kap}.
18: EFT takes advantage of the separation of scales in a system to make
19: model-independent predictions at low energy. The effective Lagrangian
20: ${\cal L}_{\rm eff}$ that describes the low-energy physics is written
21: in terms of only the long-wavelength degrees of freedom and the
22: operators that appear are determined by these degrees of freedom and the
23: symmetries present at low energy.
24: The effective Lagrangian generally
25: includes an infinite tower of nonrenormalizable interactions, but they
26: can be organized according to their importance at low energy; to a certain
27: order in a low-energy expansion, only a finite number of operators contribute
28: to a physical quantity and one can carry out renormalization in the standard
29: way order by order in this expansion.~\footnote{\label{elg}
30: The counterterms that
31: are used to cancel the divergences in the calculations of
32: one physical quantity are the same as those required in the
33: calculations of another.
34: Thus, having determined the counterterms once and for all, the
35: effective theory can be used to make predictions about other physical
36: quantities.}
37: Since the coefficients of these operators
38: encode the short-wavelength physics, we do not need to make any detailed
39: assumptions about the high-energy dynamics to make
40: predictions at low energy.
41:
42: In some cases, one can determine the coefficients
43: of the low-energy theory as functions of the coupling constants in the
44: underlying theory
45: by a perturbative matching procedure. One calculates physical
46: observables at low energies perturbatively
47: and demand they be the same in the
48: full and in the effective theory. The coefficients of the effective
49: theory then encode the short-distance physics. If one cannot determine
50: these short-distance coefficients by matching,
51: they can be
52: taken as phenomenological parameters that are determined by experiment.
53:
54: A classic example of an effective field theory is chiral perturbation theory
55: which is a low-energy field theory for pions~\cite{chiral}. Pions are
56: interacting particles whose fundamental description is provided
57: by QCD. However, QCD is strongly interacting and confining at low energies
58: and so perturbative calculations using the QCD Lagrangian are hopeless.
59: So instead of using the quark and gluon degrees of freedom in QCD, one
60: writes down the most general Lagrangian for the pions, which are the relevant
61: low-energy degrees of freedom. The terms that appear in the chiral Lagrangian
62: are determined by the global symmetries of QCD. The coefficients of the
63: chiral Lagrangian cannot be determined as functions of the couplings
64: and masses in QCD using perturbative methods.
65: They can in principle be determined using a nonperturbative method
66: such as lattice gauge theory, but in practice they are normally
67: determined by experiment.
68:
69: Nonrelativistic
70: QED~\cite{nrqed} (NRQED)
71: is an example of an effective field theory whose coefficients
72: are tuned so they reproduce a set of low-energy scattering amplitudes
73: of full QED. NRQED is
74: tailored to perform low-energy (bound state) calculations, where
75: one takes advantage of the nonrelativistic nature of the
76: bound states by
77: isolating the contributions from the relativistic
78: momentum scales. These are encoded in the coefficients of the
79: various local operators in the effective Lagrangian.
80: Traditional approaches involving the Bethe-Salpeter equation do not
81: take advantage of the separation of scales in bound state problems and are
82: therefore much more difficult to solve.
83:
84: Landau's quasiparticle model for $^4$He can also be viewed as an effective
85: theory. In order to explain that the specific heat varies as $T^3$
86: for temperatures much smaller than $T_c$, he suggested that the
87: low-lying excitations are phonons with a linear dispersion relation.
88: More generally, he proposed a spectrum that is linear
89: for small wavevectors and has a local minimum around $p=p_0$.
90: This part of the spectrum behaves like
91: \bqa
92: \epsilon(p)&=&\Delta +{(p-p_0)^2\over2m_0}\;,
93: \eqa
94: where $\Delta$, $p_0$, and $m_0$ are phenomenological parameters.
95: Excitations near $p_0$ are referred to as rotons.
96: Assuming that the elementary excitations are noninteracting, one can
97: use spectrum to calculate the specific heat.
98: Landau determined the parameters by
99: fitting the calculated specific heat to experimental data~\cite{landau22}.
100:
101: The weakly-interacting Bose gas is a system where effective field theory
102: methods can be applied successfully.
103: %This approach was first
104: %formulated and applied in~\cite{eric}.
105: The starting point is the action:
106: \bqa\nonumber
107: S[\psi^*,\psi]&=&\int dt\;
108: \Bigg\{\int d^dx\;
109: \psi^*({\bf x},t)
110: \left[
111: i{\partial\over\partial t}
112: +%{\hbar^2\over2m}
113: \nabla^2+\mu
114: \right]
115: \\ &&\nonumber
116: \times\psi({\bf x},t)
117: -{1\over2}\int d^dx\int d^dx^{\prime}\;
118: \psi^*({\bf x},t)\psi^*({\bf x}^{\prime},t)
119: \\ &&
120: \times
121: V_0({\bf x}-{\bf x}^{\prime})
122: \psi({\bf x},t)\psi({\bf x}^{\prime},t)+...
123: \Bigg\}\;.
124: \label{ac}
125: \eqa
126: Here, $\psi^*({\bf x},t)$ is a complex field operator that creates a
127: boson at the
128: position ${\bf x}$ at time $t$,
129: $\mu$ is the chemical potential, and $V_0({\bf x})$
130: is the two-body potential.
131: The ellipses indicate terms that describe possible
132: interactions between
133: three or more bosons.
134: The chemical potential
135: $\mu$ must be adjusted to get the correct number density $n$.
136:
137: The interatomic potential can be divided into a central
138: part $V_0^c(x)$ and a remainder.
139: The central of the potential only depends on the separation $x$ of the atoms
140: and their electronic spins. It conserves separately the total orbital
141: angular momentum and the total electronic spin of the atoms.
142: The noncentral part of the interaction conserves the total angular momentum,
143: but not separately the orbital
144: angular momentum and the total electronic spin of the atoms.
145: An example of a term in the noncentral part of the interaction is the
146: magnetic dipole-dipole interaction.
147:
148: The central part of the potential consists of a short-range
149: part with range $x_0$ and a long-range van der Waals tail. The latter goes as
150: $1/x^6$ as $x\rightarrow\infty$.
151: A typical two-body potential is
152: shown in Fig.~\ref{pot1}.
153: A model potential of this kind is
154: the sum of a hard-core potential with range $x_0$
155: and a van der Waals potential:
156: \bqa
157: V_0^c(x)&=&
158: \Bigg\{\begin{array}{cc}
159: +\infty\;,&x<x_0 \\
160: -{C_6\over x^6}\;,&x>x_0
161: \end{array}
162: \eqa
163: where $C_6$ and $x_0$ are constants.
164: Another example is the hard-core square-well
165: potential:
166: \bqa
167: V_0^c(x)&=&
168: \Bigg\{\begin{array}{cc}
169: +\infty\;,&x<x_c \\
170: -V_0\;,&x_0<x<x_c\\
171: 0\;,&x<x_c
172: \end{array}
173: \eqa
174: where $x_0$ and $x_c$ are constants. This potential sustains a number
175: of two-body bound states depending on the values of $x_0$ and $x_c$.
176: Many real potentials used in experiments sustain bound states and the
177: ground state of the system is no longer a homogeneoues gas, but rather
178: a state of clusters of atoms. However, if the scattering length is positive,
179: the homogeneous Bose gas can exist as a long-lived metastable state.
180:
181: In Fig.~\ref{pot2}, we have shown the Fourier transform $V({\bf k})$
182: of a typical
183: short-range two-body potential with range $x_0$. Since a true interatomic
184: potential vanishes for large momenta $k$, one will never face
185: ultraviolet divergences using it in actual (perturbative) calculations.
186:
187: In this paper we restrict our calculations to the spinless Bose gas.
188: In experiments with trapped alkali atoms
189: (e.g. $^7$Li, $^{23}$Na $^{85}$Rb, $^{87}$Rb, and $^{133}$Cs),
190: the situation is more complicated. In all these atoms, there is a single
191: $s$-electron outside closed shells. Consequently, these atoms have
192: electron spin $S=1/2$. The total spin of a colliding pair of alkali atoms
193: is therefore either $S=0$ or $S=1$. The central part of the potential
194: $V_0^c(x)$ depends on the total spin and one refers to these potentials
195: as the {\it singlet} and {\it triplet} potentials, respectively.
196: The singlet potential is generally much deeper and sustains many more
197: bound states than does the triplet potential.
198: For example, the singlet potential of $^{87}$Rb is deeper than the
199: triplet potential by more than one order of magnitude.
200: The scattering lengths are denoted by $a_s$ and $a_t$.
201: In the case of atomic hydrogen, they
202: have been calculated from first
203: principles; $a_s=0.3a_0$ and $a_t=1.3a_0$, where $a_0$
204: is the Bohr radius. Similarly, the coefficient of the van der Waals
205: tail has also been determined; $C_6=6.499a_0$~\cite{atom2}.
206:
207: One can associate a natural length scale $l$ with any atomic potential
208: $V_0^c(x)$. For a short-range potential, this length is the range $x_0$ itself.
209: For a long-range potential, it is a little more complicated.
210: The length scale associated with the van der Waals tail is
211: $l_{\rm vdW}\sim C_6^{1/4}$. In this case, the length scale $l$ is
212: either the range or $l_{\rm vdW}$, whichever is larger.
213: For a generic potential, the low-energy observables such as the scattering
214: length and the effective range are of the order $l$.
215: There is nothing that forbids the these quantities to be much larger
216: than $l$, but it is {\it unnatural} and typically it requires fine-tuning
217: of one or more parameters in the potential.
218: In the case of the alkali atoms, $l_{\rm vdW}$ is much larger than the
219: range of the short-range part of the potential and is thus the
220: natural length scale for these atoms. The spin-singlet scattering length
221: for $^{85}$Rb is $a_s=+2800a_0$ and is more than one order of magnitude
222: larger than $l=l_{\rm vdW}=164a_0$. This can be viewed as a fine-tuning of
223: the mass of the atom. This can be seen from the fact that the mass of
224: $^{87}$Rb is only 2.3$\%$ larger and the
225: spin-singlet scattering length has the more natural value $a_s=+90.4a_0$.
226: One can also obtain unnaturally large scattering lengths by tuning an
227: external parameter in experiments. One way of doing this, is to
228: tune an external magnetic field to a Feshbach resonance~\cite{fesh}.
229: This is currently receiving a lot of attention both
230: theoretically\cite{stw,ties} and experimentally~\cite{ino}.
231:
232:
233: In the remainder of this work, we do not specify the
234: magntiude of the scattering length. We only demand the {\it diluteness
235: condition} be satisfied, namely that
236: \bqa
237: na^3\ll1\;.
238: \eqa
239: We now return to the action~(\ref{ac}), which
240: is invariant under a global phase transformation
241: \bqa
242: \psi({\bf x},t)\rightarrow e^{i\alpha}\psi({\bf x},t)\;.
243: \label{gpha}
244: \eqa
245: The global $U(1)$ symmetry reflects the conservation of atoms. It also
246: ensures that the number density $n$ and current
247: density ${\bf j}$ satisfy the continuity equation
248: \bqa
249: \dot{n}+\nabla\cdot{\bf j}=0\;,
250: \eqa
251: where the dot denotes differentiation with respect to time.
252:
253:
254: %\begin{wrapfigure}[]%{l}[0pt]{8cm}
255: \begin{figure}[htb]
256: \epsfysize=6.6cm
257: \epsffile{pot.eps}
258: %\vspace{2mm}
259: \caption[a]{Typical behavior of a two-body potential $V_0^c(x)$.}
260: \label{pot1}
261: \end{figure}
262:
263: %\begin{wrapfigure}[]%{l}[0pt]{8cm}
264: \begin{figure}[htb]
265: \epsfysize=6.6cm
266: \epsffile{potp.eps}
267: %\vspace{2mm}
268: \caption[a]{Solid line:
269: Typical behavior of the potential in momentum space $V({\bf k})$.
270: Dotted line: Fourier transform of a contact potential.}
271: \label{pot2}
272: \end{figure}
273:
274:
275:
276: The nonlocal evolution equation that follows from the action~(\ref{ac}),
277: cannot always be used for practical calculations, but can be replaced
278: by a local one.
279: Suppose we are interested in the properties of the system at
280: momenta $k$
281: such that the de Broglie wavelength $1/k$ is
282: much longer than the
283: range of the interatomic potential $V_0({\bf x})$.
284: The interactions therefore appear pointlike
285: on the scale of the de Broglie wavelength,
286: and they can
287: be mimicked by local interactions
288: The parameters of these local
289: interactions must be tuned so that they reproduce low-energy observables
290: to sufficient accuracy.
291: However, if the potential is long range,
292: the scattering amplitude depends in a nonanalytic way on the wave vector
293: ${\bf k}$ characterizing the incoming atoms in the CM frame.
294: Such behavior cannot be reproduced by local operators.
295: If $V_0^c(x)$ falls of like $1/x^6$, this nonanalytic behavior
296: enters first at order $k^4$~\cite{e+h}.
297:
298: The effective Lagrangian for the bosons can be constructed using the
299: methods of effective field theory. Once the symmetries have been
300: identified, one writes down the most general local
301: effective Lagrangian consistent
302: with these symmetries.
303: At zero temperature, the symmetries are Galilean invariance, time-reversal
304: symmetry, and the global phase symmetry~(\ref{gpha}).
305: These symmetries severely restrict the possible terms in the
306: effective action. One finds~\cite{eric}:
307: \bqa\nonumber
308: &&S[\psi^*,\psi]=\int dt\;
309: \int d^dx\;\Bigg\{
310: \psi^*
311: \left[
312: i%\hbar
313: {\partial\over\partial t}
314: +\nabla^2+\mu
315: \right]\psi
316: \\ && \nonumber
317: -{1\over2}
318: g\left(\psi^*\psi\right)^2
319: -{1\over2}h\left[\nabla\left(\psi^*\psi\right)\right]^2
320: %\\ &&
321: -{g_3\over36}\left(\psi^*\psi\right)^3+...\Bigg\}
322: \;,
323: \\ &&
324: \label{act}
325: \eqa
326: where $g$, $h$, and $g_3$ are coupling constants that can be determined
327: by a matching procedure.
328: The dots denote operators that are higher order in the field $\psi$ or its
329: derivative $\nabla\psi$ and respect the symmetries.
330: One demands that the effective field theory
331: represented by the action~(\ref{act}) reproduces a set of low-energy
332: observables to some desired accuracy.
333: An example is the coefficients of the expansions in $\sqrt{na^3}$ of
334: the ground state energy density. Another example coefficients in the
335: low-energy expansions for the scattering amplitudes in the $n$-body sector.
336: When the coefficients in the action~(\ref{act}) have been determined to some
337: accuracy in a low-energy expansion, effective field theory guarantees that
338: all other observables can be determined with the same accuracy.
339: %We return to this issue below.
340:
341: The quantum field theory defined by the action~(\ref{act}) has ultraviolet
342: divergences that must be removed by renormalization of the
343: parameters $\mu, g, h, g_3...$ .
344: They arise because we are treating the interactions
345: between the atoms as pointlike down to arbitrarily short distances.
346: For instance the operator $g(\psi^{\dagger}\psi)^2$ can be thought of as
347: a contact potential with strength $g$;
348: $V_0({{\bf x}-{\bf x}^{\prime}})=g\delta({{\bf x}-{\bf x}^{\prime}})$.
349: The Fourier transform is then
350: a constant in momentum space; $V_0({\bf k})=g$.
351: This is illustrated is Fig.~\ref{pot2},
352: where the dotted line shows $V({\bf k})$. Thus $V({\bf k})$
353: does not vanish for large momenta and this is the reason why one encounters
354: ultraviolet divergences in the calculation of Feynman diagrams.
355: In order to make the theory well defined, we must introduce an ultraviolet
356: cutoff.
357: This is indicated in Fig.~\ref{pot2}, where we exclude wave numbers
358: $k>M$ in momentum integrals.
359:
360: If we use a simple momentum cutoff $M$ to cut off the ultraviolet
361: divergences, there will be terms that are proportional to
362: $M^n$, where $n$ is a positive integer.
363: There are also terms that are proportional to $\log(M)$.
364: The coefficients of the power divergences depend on the method
365: we use to regulate the integrals, while the coefficients
366: of $\log(M)$ do not. Thus the power divergence are artifacts of the
367: regulator, while the logarithmic divergences represent real physics.
368: In this paper, we will be using dimensional regularization~\cite{jerry}
369: to regulate
370: infrared as well as ultraviolet divergences in the loop integrals.
371: In dimensional regularization, one
372: calculates the loop integrals in $d=3-2\epsilon$ dimensions for values of
373: $\epsilon$ for which the integrals converge. One finally analytically continues
374: back to $d=3$ dimensions. In dimensional regularization, an arbitrary
375: momentum scale $M$ is introduced to ensure that loop integrals have their
376: canonical dimensions also away from three dimensions. This scale can be
377: identified
378: with the simple momentum cutoff mentioned above. Advantages of dimensional
379: regularization
380: are that it respects symmetries such as rotational symmetry and gauge
381: invariance.
382: Dimensional regularization sets power divergences to zero
383: and logarithmic divergences show up as poles in $\epsilon$.
384: With dimensional regularization, the only ultraviolet divergences
385: that require explicit renormalization are therefore logarithmic
386: divergences. This simplifies calculations significantly as
387: we shall see. In fact all the divergences encountered
388: in the one-loop calculations that we present here, can be
389: removed by the renormalization of $\mu$, $g$, and the vacuum
390: energy ${\cal E}$
391: and in three dimensions, these are power divergences.
392: Thus no explicit renormalization is required.
393:
394:
395:
396: We now return to the determination of the parameters $g, h, g_3...\;.$
397: We follow~\cite{e2} and determine them
398: by demanding that the
399: effective field theory~(\ref{act})
400: reproduces the low-momentum expansions for the
401: scattering amplitudes in the vacuum
402: for $2\rightarrow2$ scattering, $3\rightarrow3$ scattering,
403: etc. to some desired accuracy.
404:
405: To calculate the coupling constant $g$,
406: consider the scattering of two atoms in the vacuum
407: with initial wave numbers ${\bf k}_1$
408: and ${\bf k}_2$, and final wave numbers ${\bf k}_1^{\prime}$
409: and ${\bf k}_2^{\prime}$. The probability amplitude for the
410: $2\rightarrow2$ scattering is given by the T-matrix element
411: ${\cal T}$. The tree-level contribution to ${\cal T}$ comes from
412: the leftmost diagram in Fig.~\ref{loopfig} and is given by
413: \bqa
414: {\cal T}_0&=&-2g\;,
415: \eqa
416: where the subscript indicates the number of loops. The quantum corrections
417: to the tree-level result come from the loop diagrams in Fig.~\ref{loopfig}.
418: The leading quantum correction comes from the one-loop diagram and reads
419: \bqa\nonumber
420: {\cal T}_1(q)&=&-2ig^2\int{d\omega\over2\pi}\int{d^dk\over(2\pi)^d}
421: {1\over\omega-k^2+i\varepsilon}
422: \\ &&
423: \times
424: {1\over(\omega-k_1^2-k^2_2)+({\bf k}-{\bf k}_1-{\bf k}_2)^2+i\varepsilon}
425: \;,
426: \eqa
427: where $q={1\over2}|{\bf k}_1-{\bf k}_2|$, and we have used that the
428: free propagator $\Delta_0(\omega,{\bf k})$
429: in vacuum corresponding to the action~(\ref{act}) is
430: \bqa
431: \Delta_0(\omega,{\bf k})&=&{i\over \omega-k^2+i\epsilon}\;.
432: \eqa
433: The integral over $\omega$ is performed using contour integration.
434: After changing variables
435: ${\bf k}\rightarrow {\bf k}+|{\bf k}_1+{\bf k}_2|/2$, we obtain
436: \bqa
437: {\cal T}_1(q)&=&g^2\int{d^dk\over(2\pi)^d}
438: {1\over k^2-q^2-i\varepsilon}\;.
439: \label{cvar}
440: \eqa
441: Note that the integral over $k$ is linearly divergent in the ultraviolet.
442: This divergence is set to zero in dimensional regularization.
443: Using dimensional regularization,
444: the result of the integration of $k$ can be written as
445: \bqa
446: {\cal T}_1(q)&=&g^2M^{2\epsilon}{\Gamma(1-{d\over2})\over(4\pi)^{d/2}}\left[
447: -q^2-i\varepsilon
448: \right]^{d-2\over2}
449: \label{m11}
450: \;,
451: \eqa
452: where $\Gamma(x)$ is the gamma function.
453: The limit $d\rightarrow3$ is regular and Eq.~(\ref{m11}) reduces to
454: \bqa
455: {\cal T}_1(q)&=&i{g^2q\over16\pi}\;.
456: \eqa
457: This expression is simply Fermi's golden rule.
458: The quantum corrections from higher orders are given by the
459: diagrams like the two-loop graph in Fig.~\ref{loopfig}.
460: They form a geometric series that can be summed up exactly and the
461: exact $2\rightarrow2$ amplitude becomes
462: \bqa
463: {\cal T}(q)&=&-{2g^2\over g+{\cal T}_1(q)}\;.
464: \label{match}
465: \eqa
466: The scattering amplitude for $2\rightarrow2$ scattering in the
467: underlying theory described by the action~(\ref{ac})
468: can be calculated by solving the two-body scattering problem in the
469: potential $V_0({\bf x}-{\bf x}^{\prime})$.
470: The contribution from $s$-wave scattering is~\cite{landau}
471: \bqa
472: {\cal T}&=&{16\pi\over q}e^{i\delta_0(q)}\sin[\delta_0(q)]\;,
473: \label{m1}
474: \eqa
475: where $\delta_0(q)$ is the $s$-wave phase shift.
476: We next write the low-momentum expansion as follows
477: \bqa
478: q\cot[\delta_0(q)]&=&
479: \left[-{1\over a}+{1\over2}r_sq^2+..\right]\;.
480: \label{lmexp}
481: \eqa
482: This expansion defines
483: the scattering length $a$ and the effective range $r_s$.
484: Using the identity $e^{ix}\sin(x)=1/(\cot(x)-i)$,
485: we can expand the $T$-matrix in powers of momentum $q$.
486: Matching the expressions~(\ref{match}) and~(\ref{m1})
487: through first order in the external momentum $q$ using~(\ref{lmexp}),
488: we obtain
489: \bqa
490: g=8\pi a\;.
491: \label{ga}
492: \eqa
493: The parameter $h$ can be determined by going to the next order in the
494: low-momentum expansion. We will do this in subsec.~\ref{noneff}.
495: Similarly, the parameter $g_3$ can be determined by solving the
496: three-body scattering problem in the potential
497: $V_0({\bf x}-{\bf x}^{\prime})$ and demand that the $3\rightarrow3$
498: scattering amplitudes in the full and in the effective theory
499: be the same at low momentum.
500:
501:
502: %\begin{wrapfigure}[]%{l}[0pt]{8cm}
503: \begin{figure}[htb]
504: \epsfysize=1.4cm
505: \vspace{0.1cm}
506: \epsffile{loop.eps}
507: %\vspace{2mm}
508: \caption[a]{Diagrams contributing to the $2\rightarrow2$
509: scattering amplitude.}
510: \label{loopfig}
511: \end{figure}
512: Traditionally, the starting point has been the interatomic potential
513: $V_{0}({\bf x}-{\bf x}^{\prime})$. At low densities, it can be shown
514: that the ladder diagrams are of equal importance and must be summed.
515: The summation of these diagrams can be expressed in terms of an effective
516: interaction $\Gamma$
517: that satisfies an integral equation which also involves
518: $V_{0}({\bf x}-{\bf x}^{\prime})$~\cite{beli,fetter}.
519: The interatomic potential appearing in $\Gamma$ can be eliminated
520: in favor of the scattering amplitude for two-particle scattering.
521: In the low-momentum limit, the effective interaction reduces to the
522: $8\pi a$. The mean-field self-energies and therefore the first order
523: propagator and spectrum
524: can be expressed in terms of the effective interaction.
525: Thus the spectrum reduces in the low-momentum limit to the Bogoliubov
526: spectrum.
527:
528:
529:
530: \subsection{Perturbative Framework}
531: \label{pf}
532: We next discuss the perturbative framework set up by Braaten and
533: Nieto~\cite{eric}
534: that can be used to systematically
535: calculate the low-energy properties of a weakly interacting Bose gas.
536:
537: We first parameterize the quantum field $\psi$
538: in terms of a time-independent
539: condensate $v$ and a quantum fluctuating field $\tilde{\psi}$:
540: \bqa
541: \psi=v+\tilde{\psi}\;.
542: \label{shift}
543: \eqa
544: The fluctuating field $\tilde{\psi}$ can
545: be conveniently written in terms of two real fields:
546: \bqa
547: \tilde{\psi}={1\over\sqrt{2}}\left(\psi_1+i\psi_2\right)\;.
548: \label{split}
549: \eqa
550: Substituting Eqs.~(\ref{shift}) and~(\ref{split})
551: into Eq.~(\ref{act}), the action can
552: be decomposed into three terms
553: \bqa
554: \label{terms}
555: S[v,\psi_1,\psi_2]=S[v]+S_{\rm free}[v,\psi_1,\psi_2]
556: +S_{\rm int}[v,\psi_1,\psi_2]\;,
557: \eqa
558: where we have indicated that the action depends on $v$ as well, and
559: switched to the variables $\psi_1$ and $\psi_2$ instead of
560: $\psi^*$ and $\psi$.
561: $S[v]$ is the classical action
562: \bqa
563: S[v]=\int dt\int d^dx
564: \left[\mu v^2-{1\over2}gv^4
565: \right]\;,
566: \label{cact}
567: \eqa
568: while the free part of the action is
569: \bqa\nonumber
570: S_{\rm free}[v,\psi_1,\psi_2]&=&\int dt\int d^dx
571: \left[
572: {1\over2}\left(\dot{\psi}_1\psi_2-\psi_1\dot{\psi_2}\right)
573: \right.\\
574: &&\left.
575: \hspace{-2cm}+{1\over2}
576: \psi_1\left(\nabla^2+X\right)\psi_1
577: +{1\over2}
578: \psi_2\left(
579: \nabla^2+Y\right)\psi_2
580: \right]\;.
581: \label{free}
582: \eqa
583: where
584: \bqa
585: X&=&\mu-3gv^2\;,
586: \label{x}
587: \\
588: Y&=&\mu-gv^2\;.
589: \label{y}
590: \eqa
591: The terms $3gv^2$ and $gv^2$ in $X$ and $Y$ are often referred to as
592: mean-field self-energies.
593: The interaction part of the action is
594: \bqa\nonumber
595: S_{\rm int}[v,\psi_1,\psi_2]&=&
596: \int dt\int d^dx\left[
597: \sqrt{2}J\psi_1
598: \right. \\ &&\left.
599: \hspace{-1cm}
600: +{1\over\sqrt{2}}Z\psi_1\left(\psi_1^2+\psi_2^2\right)
601: -{1\over8}g\left(\psi_1^2+\psi_2^2\right)^2
602: \right]\;,
603: \label{inter}
604: \eqa
605: The sources in Eq.~(\ref{inter}) are
606: \bqa
607: J&=&\left[\mu-gv^2\right]v
608: \label{t}
609: \;,\\
610: Z&=&-gv\;.
611: \eqa
612: The propagator that corresponds to the free action
613: $S_{\rm free}[v,\psi_1,\psi_2]$
614: in Eq.~(\ref{free}) is
615: \bqa
616: \label{prop}
617: D(\omega,p)&=&\frac{i}{\omega^2
618: -\epsilon^2(p)+i\varepsilon}
619: \left(\begin{array}{cc}
620: p^2-Y&-i\omega \\
621: i\omega&p^2-X
622: \end{array}\right)\;.
623: \eqa
624: Here, $p=|{\bf p}|$, where ${\bf p}$ is the wavevector,
625: $\omega$
626: is the frequency, and
627: $\epsilon(p)$ is the dispersion relation:
628: \bqa
629: \label{disp}
630: \epsilon(p)=\sqrt{(p^2-X)(p^2-Y)}\;.
631: \eqa
632: Note that one can diagonalize the matrix~(\ref{prop})
633: by a field redefinition, which is equivalent to a Bogoliubov transformation
634: in the operator approach. Such a field redefinition makes, however,
635: the interaction terms much more complicated and
636: increases the number of diagrams that one needs to evaluate.
637: For practical purposes, we therefore stick to the above propagator.
638:
639:
640:
641: The partition function ${\cal Z}$ can be expressed as a path integral
642: over the quantum fields $\psi_1$ and $\psi_2$~\cite{nege}:
643: \bqa
644: {\cal Z}=\int{\cal D}\psi_1{\cal D}\psi_2\,e^{iS[v,\psi_1,\psi_2]}\;.
645: \eqa
646: All the thermodynamic observables can be derived from
647: the partition function
648: ${\cal Z}$.
649: For instance, the free energy density ${\cal F}$ is given by
650: \bqa
651: {\cal F}(\mu)=i{\log{\cal Z}\over{\cal V}}\;,
652: \eqa
653: where ${\cal V}$
654: is the spacetime volume of the system. The pressure ${\cal P}(\mu)$
655: is
656: \bqa
657: {\cal P}(\mu)&=&-{\cal F}(\mu)\;.
658: \eqa
659: The number density $n$ is given by the expectation value
660: $\langle\psi^*\psi\rangle$ in the ground state:
661: \bqa
662: n(\mu)&=&\int{\cal D}\psi_1{\cal D}\psi_2\,(\psi^*\psi)\;e^{iS[v,\psi_1,\psi_2]}\;.
663: \eqa
664: It can therefore be expressed
665: as
666: \bqa
667: \label{rhodef}
668: n(\mu)=-{\partial {\cal F}(\mu)\over \partial \mu}\;.
669: \eqa
670: The energy density ${\cal E}$ is given by the Legendre transform of
671: the free energy density ${\cal F}$:
672: \bqa
673: {\cal E}(n)={\cal F}(\mu)+n\mu\;.
674: \label{rela}
675: \eqa
676:
677:
678:
679:
680:
681:
682:
683: The free energy ${\cal F}$ is given by all
684: {\it connected vacuum diagrams} which are
685: Feynman diagrams with no external legs.
686: The sum of the vacuum graphs is independent of the condensate $v$.
687: At this point it is convenient to
688: introduce the thermodynamic potential $\Omega(\mu,v)$.
689: The thermodynamic potential is given by all
690: {\it one-particle irreducible vacuum diagrams}
691: and can be expanded in number of loops:
692: \bqa
693: \Omega(\mu,v)&=&
694: \Omega_0(\mu,v)+\Omega_1(\mu,v)+\Omega_2(\mu,v)+...\;,
695: \label{loopp}
696: \eqa
697: where the subscript $n$ denotes the contribution from the $n$th order in the
698: loop expansion.
699: If $\Omega$ is evaluated at a value of the condensate that satisfies the
700: condition
701: \bqa
702: \bar{v}&=&\langle\psi\rangle\;,
703: \label{condit}
704: \eqa
705: all {\it one-particle reducible diagrams} (those that are disconnected
706: by cutting a single propagator line) vanish.
707: Thus evaluating the thermodynamic potential at the value
708: of the condensate that satisfies~(\ref{condit}), one obtains the free
709: energy:
710: \bqa
711: {\cal F}(\mu)&=&
712: \Omega_0(\mu,\bar{v})+\Omega_1(\mu,\bar{v})+\Omega_2(\mu,\bar{v})+...\;.
713: \label{loop}
714: \eqa
715: Using Eq.~(\ref{split}), the
716: condition~(\ref{condit}) reduces to
717: $\langle\psi_2\rangle=0$ and $\langle\psi_1\rangle=0$.
718: The first condition can be automatically satisfied by a suitable
719: choice of the phase of $\psi$.
720: The second condition is then equivalent to
721: \bqa
722: {\partial\Omega(\mu,v)\over\partial v}&=&0\;.
723: \label{tad1}
724: \eqa
725: The value of the condensate that satisfies~(\ref{tad1}) is denoted
726: by $\bar{v}$.
727: The free energy can also be expanded in
728: powers of quantum corrections around the
729: mean-field result ${\cal F}_0(\mu)$:
730: \bqa
731: {\cal F}(\mu)&=&{\cal F}_0(\mu)+{\cal F}_1(\mu)+{\cal F}_2(\mu)+...
732: \label{fh}
733: \eqa
734: The loop expansion~(\ref{loop}) of ${\cal F}(\mu)$
735: does not coincide with the expansion~(\ref{fh})
736: of ${\cal F}(\mu)$ in powers of quantum corrections because of its dependence
737: on $\bar{v}$.
738: To obtain the expansion of ${\cal F}$ in powers of
739: quantum corrections, we must
740: expand the condensate $\bar{v}$ in powers of
741: quantum corrections:
742: \bqa
743: \bar{v}=v_0+v_1+v_2+...\;,
744: \label{vexp}
745: \eqa
746: where $v_0$ is the classical minimum, which satisfies
747: \bqa
748: {\partial\Omega_0(\mu,v)\over\partial v}&=&0\;.
749: \eqa
750: %It follows from~(\ref{cact}) that $v_0^2=\mu/g$.
751: By expanding Eq.~(\ref{tad1})
752: about $v_0$, one obtains the quantum corrections $v_1, v_2,...$
753: to the condensate. For instance,
754: the first quantum correction $v_1$ to the classical minimum $v_0$ is
755: \bqa
756: v_1=-{\partial\Omega_1(\mu,v)\over\partial v}\Bigg|_{v=v_0}\Bigg/
757: {\partial^2\Omega_0(\mu,v)\over\partial v^2}\Bigg|_{v=v_0}
758: \;.
759: \label{v11}
760: \eqa
761: The mean-field free energy density is
762: \bqa
763: {\cal F}_0(\mu)&=&\Omega_0(\mu,v_0)
764: \eqa
765: Inserting~(\ref{vexp}) into~(\ref{loop}) and expanding in powers
766: of $v_1, v_2...$, we obtain the quantum expansion of the free
767: energy density.
768: The first quantum correction to the free energy density is
769: \bqa
770: {\cal F}_1(\mu)=\Omega_1(\mu,v_0)\;,
771: \eqa
772: and the second quantum correction to the free energy density
773: is
774: \bqa\nonumber
775: {\cal F}_2(\mu)&=&\Omega_2(\mu,v_0)+
776: v_1{\partial \Omega_1(\mu,v)\over\partial v}\Bigg|_{v=v_0}\\
777: &&+
778: {1\over2}v^2_1{\partial^2\Omega_0(\mu,v)\over\partial v^2}\Bigg|_{v=v_0}\;.
779: \eqa
780: Expressions for higher order corrections to the free energy can be
781: derived in the same way.
782:
783: The value of the condensate $v$ that minimizes the classical
784: action~(\ref{cact}) is given by $v_0=\sqrt{\mu/g}$.
785: The linear term
786: in Eq.~(\ref{inter}) then vanishes since $J=0$.
787: At the minimum of the classical action,
788: both the propagator~(\ref{prop})
789: and the dispersion
790: relation~(\ref{disp}) simplify significantly since we also have $Y=0$.
791: At the minimum,
792: these equations reduce to
793: \bqa
794: D(\omega,p)&=&\frac{i}{\omega^2
795: -\epsilon^2(p)+i\varepsilon}
796: \left(\begin{array}{cc}
797: p^2&-i\omega \\
798: i\omega&\epsilon^2(p)/p^2
799: \end{array}\right)\;,
800: \label{prop2}
801: \\
802: \epsilon(p)&=&p\sqrt{p^2+2\mu}\;.
803: \label{disps}
804: \eqa
805: The spectrum~(\ref{disps}) was first derived by Bogoliubov in 1947~\cite{bogo}.
806: The dispersion relation is gapless and linear for small wave vectors.
807: This reflects the spontaneous breakdown of the $U(1)$-symmetry.
808: The dispersion relation changes from being linear to being quadratic
809: in the vicinity of $p=\sqrt{2\mu}$, which is called the
810: inverse coherence length.
811: For very large wave vectors, the
812: dispersion relation is approximately $\epsilon(p)=p^2+\mu$, where
813: the second term represents the mean-field energy due to the
814: interaction with the condensed particles.
815:
816:
817: We next comment on how traditional approaches fit into the more
818: general perturbative framework discussed in this section.
819: The starting point is the second quantized grand canonical Hamiltonian
820: that includes a two-body potential $V_0({\bf x})$:
821: \bqa\nonumber
822: {\cal H}&=&
823: \int d^dx\;
824: \psi^*({\bf x})
825: \left[
826: -\nabla^2-\mu
827: \right]\psi({\bf x})
828: \\ &&\hspace{-0.5cm}\nonumber
829: +{1\over2}\int d^dx\int d^dx^{\prime}\;
830: \psi^*({\bf x})\psi^*({\bf x}^{\prime})
831: V_0({\bf x}-{\bf x}^{\prime})
832: \psi({\bf x})\psi({\bf x}^{\prime})\;.
833: \\ &&
834: \label{ham1}
835: \eqa
836: This Hamiltonian is nonlocal and is often replaced by a local one. This is
837: done by approximating the true two-body potential by a local
838: two-body interaction, whose strength $g$ is tuned to reproduce the
839: scattering length of $V_{0}({\bf x}-{\bf x}^{\prime})$.
840: This yields
841: \bqa
842: {\cal H}&=&
843: \int d^dx\;\left\{
844: \psi^*
845: \left[
846: -\nabla^2-\mu
847: \right]\psi
848: +{1\over2}g\left(\psi^*\psi\right)^2
849: \right\}\;.
850: \label{ham2}
851: \eqa
852: The grand canonical Hamiltonian
853: is expressed in terms of creation and annihilation operators that satisfy
854: the standard equal-time commutation relations. Bogoliubov's idea was
855: to treat the ${\bf p}=0$ momentum separately~\cite{bogo}.
856: Since this state is macroscopically occupied, the
857: creation and annihilation operators for bosons with
858: ${\bf p}=0$ commute to very good approximation.
859: Thus they can be treated classically and be replaced by a constant which
860: is the condensate density. This step is equivalent to splitting the
861: quantum field $\psi$ into a condensate $v$ and a fluctuating field
862: $\tilde{\psi}$, Eq.~(\ref{shift}).
863:
864: In the Bogoliubov approximation~\cite{bogo},
865: one makes a quadratic approximation to the Hamiltonian by neglecting
866: terms with three and four operators.
867: Since the resulting Hamiltonian contains products of two annihilation
868: and products of two creation operators, it must be
869: diagonalized by a canonical transformation.
870: The resulting quasi-particle spectrum is then given by Eq.~(\ref{disps}).
871:
872: In the Beliaev approximation~\cite{beli}, one goes one step
873: further by calculating the leading quantum corrections to the
874: quasi-particle spectrum~(\ref{disps}).
875: This is done by including all one-loop diagrams, that is
876: all diagrams up to second order in the interaction,
877: in the self-energies and calculating the poles of the propagator.
878: The
879: correction to the Bogoliubov spectrum~(\ref{disps}) was
880: first calculated by
881: Beliaev~\cite{beli} and coincides with a leading-order calculation in the
882: approach outlined here. We shall return to that calculation in
883: Sec.~\ref{belsec}.
884:
885: The perturbative framework has been formulated in terms of two real fields
886: $\psi_1$ and $\psi_2$.
887: As noted in the introduction, one has traditionally
888: presented the theory in terms of normal and anomalous
889: Green's functions $G_{11}(\omega,p)$ and $G_{12}(\omega,p)$
890: and the self-energies
891: $\Sigma_{11}(\omega,p)$ and $\Sigma_{12}(\omega,p)$.
892: This formulation corresponds to using the fields $\psi^{*}$
893: and $\psi$ instead of $\psi_1$ and $\psi_2$.
894: In either formulation, one ends up with a $2\times2$ matrix for the
895: propagator and self-energies and the amount of work to calculate most
896: quantities is comparable.
897: For readers who want to translate intermediate results to
898: a more familiar language, we note that
899: \bqa
900: \Sigma_{11}(\omega,p)&=&
901: %2gv^2+
902: {1\over2}\left[\Pi_{11}(\omega,p)+\Pi_{22}(\omega,p)
903: \right]\;,
904: \label{r1}
905: \\
906: \Sigma_{12}(\omega,p)&=&%gv^2+
907: {1\over2}\left[\Pi_{11}(\omega,p)-\Pi_{22}(\omega,p)\right]
908: \label{r2}
909: \;,
910: \eqa
911: where $\Pi_{ij}(\omega,p)$
912: are the components of the $2\times2$ self-energy matrix.
913:
914: We next comment on the Hugenholz-Pines theorem~\cite{hug}.
915: The Hugenholz-Pines theorem ensures that the spectrum does not exhibit a
916: gap and it has been proven to all orders in perturbation theory.
917: It is simply the Goldstone theorem for a dilute Bose gas with
918: a spontaneously broken continuous symmetry.
919: The Hugenholz-Pines theorem is normally given in terms of
920: normal and anomalous self-energies:
921: \bqa\nonumber
922: \mu&=&
923: \Sigma_{11}(0,0)-\Sigma_{12}(0,0)\;.
924: \eqa
925: In terms of the self-energies $\Pi_{ij}(\omega,p)$, the theorem
926: takes a particular
927: simple form:
928: \bqa
929: \mu&=&\Pi_{22}(0,0)\;.
930: \eqa
931:
932:
933:
934:
935:
936:
937: \subsection{Ground State Energy Density and Condensate Depletion}
938: \label{subcon}
939: In this subsection, we calculate the leading quantum correction to the
940: ground state energy and the depletion of the condensate. While these
941: results are standard textbook material~\cite{fetter}, it is instructive
942: to see how they are derived within the present framework.
943:
944:
945: The mean-field thermodynamic potential $\Omega_0(\mu,v)$ is given
946: by the terms in the classical action~(\ref{cact}):
947: \bqa
948: \Omega_0(\mu,v)=-\mu v^2+{1\over2}gv^4\;.
949: \label{o0}
950: \eqa
951: The minimum of the mean-field thermodynamic potential is given
952: by $v_0=\sqrt{\mu/g}$ and
953: the mean-field free energy ${\cal F}_0$ is obtained by
954: evaluating Eq.~(\ref{o0}) at the minimum:
955: \bqa
956: {\cal F}_0(\mu)&=&-{\mu^2\over 2g}\;.
957: \label{f0}
958: \eqa
959: The mean-field number density follows from differentiating~(\ref{f0})
960: with respect to $\mu$. The chemical potential in the mean-field
961: approximation is then obtained by
962: inversion:
963: \bqa
964: \mu_0&=&gn\;.
965: \eqa
966: The mean-field energy ${\cal E}_0$ is easily from~(\ref{rela}) found to be
967: \bqa\nonumber
968: {\cal E}_0(n)&=&{1\over2}gn^2\\
969: &=&4\pi an^2\;.
970: \label{e0}
971: \eqa
972: The mean-field result for the energy density was first obtained by
973: Bogoliubov~\cite{bogo}.
974:
975: The one-loop contribution to the thermodynamic potential $\Omega(\mu,v)$ is
976: given by \bqa
977: \Omega_1(\mu,v)&=&i{\log{\cal Z}_0\over{\cal V}}\;,
978: \eqa
979: where ${\cal Z}_0$ is the path integral involving
980: the quadratic quantum fluctuations around the mean field.
981: \bqa\nonumber
982: {\cal Z}_0&=&\int{\cal D}\psi_1{\cal D}\psi_2\;
983: e^{iS_{\rm free}[v,\psi_1,\psi_2]}\\
984: &=&e^{i\int dt\int d^3x\det D^{-1}(\omega,p)}
985: \eqa
986: where $D^{-1}(\omega,p)$
987: is the inverse of the propagator~(\ref{prop})
988: and $\det$ denotes a determinant in the functional sense.
989: Using the fact that ${\rm Tr}\log A=\log\det A$ for any matrix $A$,
990: we obtain
991: \bqa\nonumber
992: \Omega_1(\mu,v)&=&-{1\over2}i\int{d\omega\over2\pi}
993: \int{d^dk\over(2\pi)^d}\log\det D^{-1}(\omega,k)
994: \\ &&
995: +\Delta_1\Omega\;.
996: \label{1om}
997: \eqa
998: Here we have added
999: $\Delta_1\Omega$, which is the one-loop counterterm.
1000: The counterterm is added to cancel the ultraviolet divergences
1001: that one encounters when evaluating the integral in~(\ref{1om}).
1002: After integrating over $\omega$ using~(\ref{wi1}), we obtain
1003: \bqa\nonumber
1004: \Omega_1(\mu,v)&=&{1\over2}\int{d^dk\over(2\pi)^d}\sqrt{(k^2-X)(k^2-Y)}
1005: \\&&
1006: +\Delta_1\Omega(\mu,v)\;,
1007: \label{2om}
1008: \eqa
1009: The one-loop contribution ${\cal F}_1$ to the free energy is obtained by
1010: evaluating Eq.~(\ref{2om}) at the classical minimum,
1011: where $Y=0$ and $X=-2\mu$.
1012: The one-loop free energy then reduces to
1013: \bqa\nonumber
1014: {\cal F}_{0+1}(\mu)&=&
1015: -{\mu^2\over 2g}+{1\over2}\int{d^dk\over(2\pi)^d}k\sqrt{k^2+2\mu}
1016: +\Delta_1{\cal F}(\mu)\\
1017: &=&-{\mu^2\over 2g}+{1\over2}I_{0,-1}(2\mu)+\Delta_1{\cal F}(\mu)\;,
1018: \label{f01}
1019: \eqa
1020: where the integral $I_{m,n}(\Lambda)$ is defined in the appendix and
1021: $\Delta_1{\cal F}$ is the one-loop counterterm.
1022: The integral $I_{0,-1}(\Lambda)$ has quintic, cubic, and linear ultraviolet
1023: divergences that are
1024: set to zero in dimensional
1025: regularization.
1026: The counterterm $\Delta_1{\cal F}$ is therefore zero and
1027: the limit $d\rightarrow3$ is
1028: regular. We obtain
1029: \bqa
1030: {\cal F}_{0+1}(\mu)&=&-{\mu^2\over2g}
1031: \left[1-{4\sqrt{2\mu g^2}\over15\pi^2}
1032: \right]\;.
1033: \label{fr0+1}
1034: \eqa
1035: It might
1036: be useful to see how the renormalization procedure works with a simple
1037: ultraviolet cutoff. In that case, the one-loop contribution to the
1038: free energy ${\cal F}_{1}$
1039: can be written as
1040: \bqa
1041: {\cal F}_{1}(\mu)&=&{1\over2}\int^M{d^3k\over(2\pi)^3}
1042: k\sqrt{k^2+2\mu}+\Delta_1{\cal F}
1043: \;,
1044: \eqa
1045: where the integral is calculated in $d=3$ dimensions and the superscript
1046: $M$ indicates that $|{\bf k}|<M$ has been imposed.
1047: We can now rewrite this as
1048: \bqa\nonumber
1049: {\cal F}_{1}(\mu)&=&{1\over2}\int^M{d^3k\over(2\pi)^3}
1050: \left[k\sqrt{k^2+2\mu}
1051: -k^2-\mu+{\mu^2\over2k^2}
1052: \right]\\ &&
1053: +\int^M{d^3k\over(2\pi)^3}
1054: \left[
1055: k^2+\mu-{\mu^2\over2k^2}
1056: \right]
1057: +\Delta_1{\cal F}
1058: \eqa
1059: The first integral is now convergent in the limit $M\rightarrow\infty$
1060: and the divergences have been isolated in the second integral.
1061: This first term goes like $M^5$
1062: and is independent of $\mu$ and $g$. It
1063: can therefore be removed by a vacuum energy counterterm $\Delta_1{\cal E}$.
1064: The form of $\Delta_1{\cal F}$ can be found by substituting
1065: $g\rightarrow g+\Delta_1g$ and $\mu\rightarrow \mu+\Delta_1{\mu}$
1066: in ${\cal F}_{0}(\mu)$
1067: and expanding to first order in $\Delta_1g$ and $\Delta_1\mu$.
1068: Including the vacuum counterterm, we obtain
1069: \bqa
1070: \Delta{\cal F}_1(\mu)&=&-{\mu\over g}\Delta_1\mu
1071: +{\mu^2\over2g^2}\Delta_1g+\Delta_1{\cal E}\;.
1072: \eqa
1073: The counterterms needed to cancel the quintic, cubic and linear
1074: divergences can then be found by inspection. One obtains
1075: \bqa
1076: \Delta_1{\cal E}&=&
1077: {1\over2}\int^M{d^3k\over(2\pi)^3}{k^2}\;,
1078: \\
1079: \label{dgcut}
1080: \Delta_1g&=&{1\over2}g^2\int^M{d^3k\over(2\pi)^3}{1\over k^2}\;,\\
1081: \Delta_1\mu&=&{1\over2}g\int^M{d^3k\over(2\pi)^3}\;.
1082: \eqa
1083: Note that the counterterm $\Delta_1g$ in Eq.~(\ref{dgcut}) is
1084: precisely what is needed to cancel the divergence appearing
1085: in the one-loop correction to the scattering length, Eq.~(\ref{cvar}).
1086: This is in accord with our comment on counterterms in footnote~\ref{elg}.
1087: The renormalized one-loop
1088: free energy can then be written as
1089: \bqa\nonumber
1090: {\cal F}_{0+1}(\mu)&=&-{\mu^2\over2g}+
1091: {1\over2}\int^M{d^3k\over(2\pi)^3}\left[\epsilon(k)-
1092: k^2
1093: \right.\\ &&
1094: \left.
1095: -\mu+{\mu^2\over2k^2}
1096: \right]\;.
1097: \eqa
1098: We can now take the limit $M\rightarrow\infty$.
1099: Evaluating the integral,
1100: we recover Eq.~(\ref{fr0+1}).
1101:
1102: Finally, we mention that the {\it pseudo-potential method}~\cite{huangyang}
1103: is an alternative way of treating the
1104: ultraviolet divergences. One replaces the contact potential
1105: $g\delta^3({\bf x})/2$ by the pseudo potential
1106: $g\delta^3({\bf x})\left(\partial/\partial x\right)x/2$ and one can
1107: avoid the ultraviolet divergences by evaluating the partial derivative
1108: at the right stage of the calculation. We shall not discuss this method
1109: any further.
1110:
1111:
1112:
1113:
1114: Using Eqs.~(\ref{rhodef}),~(\ref{f01}) and~(\ref{alge1}),
1115: we obtain the number
1116: density in the
1117: one-loop approximation
1118: \bqa\nonumber
1119: n_{0+1}(\mu)&=&{\mu\over g}-{1\over2}I_{1,1}(2\mu)\\
1120: &=&{\mu\over g}\left[1-{\sqrt{2\mu g^2}\over3\pi^2}
1121: \right]
1122: \;.
1123: \label{1lden}
1124: \eqa
1125: Inverting Eq.~(\ref{1lden}) to obtain the chemical potential as a function
1126: of the number density, one finds
1127: \bqa\nonumber
1128: \mu_{0+1}(n)&=&gn+{1\over2}gI_{1,1}(2gn)\\
1129: &=&8\pi an\left[
1130: 1+{32\over3}\sqrt{{na^3\over\pi}}
1131: \right]\;.
1132: \label{muuu}
1133: \eqa
1134: Note that here and in the following, we are replacing the argument
1135: $\mu$ by $gn$ in the loop integrals $I_{m,n}$, since corrections are
1136: of higher order in $\sqrt{na^3}$.
1137: Using Eqs.~(\ref{rela}),~(\ref{f01}) and~(\ref{1lden}),
1138: we obtain the energy density
1139: in the one-loop approximation:
1140: \bqa\nonumber
1141: {\cal E}_{0+1}(n)&=&{1\over2}gn^2+{1\over2}I_{0,-1}(2gn)\\
1142: &=&4\pi an^2\left[
1143: 1+{128\over15}\sqrt{{na^3\over\pi}}
1144: \right]
1145: \;.
1146: \label{ed}
1147: \eqa
1148: The leading quantum correction to the mean-field result~(\ref{e0})
1149: was first derived
1150: by Lee, Huang and Yang~\cite{leeyang,leehu} for a hard-sphere potential.
1151: Later it has been shown that it is {\it universal} in the sense that
1152: it applies to all short-range potential with scattering length
1153: $a$~\cite{bs,beli,lieb}.
1154: Note that the result~(\ref{ed}) is
1155: nonanalytic
1156: in the scattering length $a$. This shows that the result is nonperturbative
1157: from the point of view of ``naive'' perturbation theory, where one uses
1158: free particle propagators.
1159: It corresponds to the summation of an infinite set of one-loop diagrams,
1160: with repeated insertions of the operator ${3\over2}gv^2\psi_1^2$.
1161: The structure of these diagrams is the same as the ring
1162: diagrams first discussed by Gell-Mann and Brueckner
1163: for the nonrelativistic electron gas~\cite{gm},
1164: and are summed in the same manner.
1165: It is interesting to note that each diagram in the series
1166: is increasingly divergent in the infrared, but the sum is infrared
1167: convergent.
1168:
1169: Using~(\ref{shift}) and the parametrization~(\ref{split}), the
1170: expression for the number density becomes
1171: \bqa
1172: n&=&
1173: v^2+{1\over2}\langle\psi^2_1+\psi_2^2\rangle\;.
1174: \label{ndef}
1175: \eqa
1176: The condensate density $n_0$ is given by the expectation value
1177: $v^2=|\langle\psi\rangle|^2$
1178: At the mean-field level,
1179: one neglects the fluctuations of $\psi_1$ and $\psi_2$ and
1180: replaces $\langle\psi^*\psi\rangle$ by $|\langle\psi\rangle|^2$.
1181: The total number density
1182: is then
1183: equal to the condensate density. Taking quantum fluctuations into account,
1184: this is no longer the case. Due to interactions, some of the particles
1185: are kicked out of the condensate and are not in the ${\bf k}=0$
1186: momentum state. The difference $n-n_0$
1187: is called the depletion of the condensate and is proportional to the
1188: diluteness parameter $\sqrt{na^3}$.
1189: The one-loop diagrams that contribute to the
1190: expectation values $\langle\psi_1^2\rangle$ and $\langle\psi_2^2\rangle$
1191: are shown in Fig.~\ref{depl}.
1192: The solid line denotes the diagonal propagator
1193: for $\psi_1$ and the dashed line the diagonal propagator for $\psi_2$.
1194: The blobs denote an insertion of the operator
1195: $\psi_1^2$ or $\psi_2^2$.
1196: Taking these one-loop effects into account,
1197: Eq.~(\ref{ndef}) reduces to
1198: \bqa\nonumber
1199: n_{0+1}(n_0)&=&n_0+{1\over2}i
1200: %\left({e^{\gamma}M^2\over4\pi}\right)^{\epsilon}
1201: \int{d\omega\over2\pi}\int{d^dk\over(2\pi)^d}
1202: %\\ && \nonumber\times
1203: \left[{k^2+\epsilon^2(k)/k^2\over\omega^2-\epsilon^2(k)+i\varepsilon}\right]\\
1204: &=&n_0+{1\over4}\left[I_{1,1}(2gn_0)+I_{-1,-1}(2gn_0)\right]
1205: \;.
1206: \label{t0n}
1207: \eqa
1208: Using the expressions for the integrals in the appendix, we obtain
1209: \bqa
1210: n_{0+1}(n_0)&=&n_0\left[1+{8\over3}\sqrt{n_0a^3\over\pi}\right]
1211: \label{deplbog}
1212: \;.
1213: \eqa
1214: The result~(\ref{deplbog}) for the depletion
1215: was first obtained by Bogoliubov in 1947~\cite{bogo}.
1216: In recent experiments, Cornish {\it et al.}~\cite{corny} were able to vary
1217: the $s$-wave scattering length
1218: $a$ for $^{85}$Rb atoms
1219: over a large range by applying a strong external magnetic field
1220: and exploiting the existence of a Feshbach resonance at $B\sim 155\;G$.
1221: Values for $\sqrt{na^3}$ up to approximately
1222: 0.1 were obtained and should be sufficiently large to see deviations
1223: from the mean field in experiments. In order to observe this quantum
1224: phenomenon, it is essential that experiments are carried out at sufficiently
1225: low temperature so that the thermal depletion of the
1226: condensate is negligible.
1227:
1228:
1229: %\begin{wrapfigure}[]%{l}[0pt]{8cm}
1230: \begin{figure}[htb]
1231: \epsfysize=1.6cm
1232: \epsffile{depl.eps}
1233: %\vspace{2mm}
1234: \caption[a]{One-loop diagrams contributing to
1235: expectation values $\langle\psi_1^2\rangle$
1236: and $\langle\psi_2^2\rangle$.
1237: }
1238: \label{depl}
1239: \end{figure}
1240:
1241:
1242:
1243: \subsection{Collective Excitations}
1244: \label{belsec}
1245: The Bogoliubov spectrum is given by Eq.~(\ref{disps})
1246: and was derived from the microscopic theory represented by the
1247: action~(\ref{act}).
1248: It is linear for
1249: small momentum $p$ with slope $\sqrt{2\mu}$.
1250: The spectrum $\omega(p)$ of collective excitations is given by
1251: the poles of the propagator. The poles are the solutions to the
1252: equation
1253: \bqa
1254: \det\left[D^{-1}_0(\omega,p)-\Pi(\omega,p)\right]&=&0\;,
1255: \label{colll}
1256: \eqa
1257: where $D_0(\omega,p)$ is the real-time version of the free
1258: propagator~(\ref{propf})
1259: and $\Pi(\omega,p)$ is the $2\times2$ self-energy matrix.
1260: The dispersion relation $\omega(p)$ is generally complex and can
1261: therefore be written as
1262: \bqa
1263: \omega(p)&=&{\rm Re}\,\omega(p)-i\gamma(p)\;.
1264: \eqa
1265: The real part ${\rm Re}\,\omega(p)$
1266: gives the energies of the excitations, while
1267: the imaginary part $\gamma(p)$
1268: represents the damping of the excitations.
1269: The Bogoliubov spectrum is purely real and in this approximation,
1270: the excitations therefore have an infinite lifetime.
1271:
1272:
1273: In the following, we calculate the leading quantum correction to the
1274: real part of the spectrum in the long-wavelength limit
1275: and thus reproduce Beliaev's classic result~\cite{beli}.
1276: The one-loop Feynman diagrams that contribute to the self-energies
1277: are shown in
1278: Figs.~\ref{selfff}--\ref{s12}\footnote{Note that the sum of all
1279: {\it one-particle reducible} diagrams that contribute to the self-energies
1280: vanishes. This sum is proportional to the derivative of the effective
1281: potential and is zero when evaluated at the minimum of the effective
1282: potential.}.
1283: The solid line denotes the diagonal propagator
1284: for $\psi_1$ and the dashed line the diagonal propagator for $\psi_2$.
1285: The off-diagonal propagators for $\psi_1$ and $\psi_2$ are represented
1286: by lines that are half solid and half dashed and vice versa.
1287: One-loop contributions to self-energies are down by a factor of $\sqrt{na^3}$
1288: compared to the mean-field terms in the inverse propagator.
1289: It is therefore consistent to evaluate the self-energies
1290: using the mean-field dispersion relation $\epsilon(p)$ since
1291: corrections would be suppressed by at least a factor of $na^3$.\\
1292:
1293: %\begin{wrapfigure}[]%{l}[0pt]{8cm}
1294: \begin{figure}[htb]
1295: \epsfysize=3.0cm
1296: \epsffile{pi.eps}
1297: %\vspace{2mm}
1298: \caption[a]{One-loop diagrams contributing to the self-energy
1299: $\Pi_{11}(\omega,p)$.}
1300: \label{selfff}
1301: \end{figure}
1302: \vspace{1cm}
1303:
1304: %\begin{wrapfigure}[]%{l}[0pt]{8cm}
1305: \begin{figure}[htb]
1306: \epsfysize=3.0cm
1307: \epsffile{p2.eps}
1308: %\vspace{2mm}\
1309: \caption[a]{One-loop diagrams contributing to the self-energy
1310: $\Pi_{22}(\omega,p)$.}
1311: \label{s222}
1312: \end{figure}
1313: \vspace{1cm}
1314:
1315: %\begin{wrapfigure}[]%{l}[0pt]{8cm}
1316: \begin{figure}[htb]
1317: \epsfysize=0.6cm
1318: \epsffile{p1.eps}
1319: %\vspace{2mm}
1320: \caption[a]{One-loop diagrams contributing to the self-energy
1321: $\Pi_{12}(\omega,p)$.}
1322: \label{s12}
1323: \end{figure}
1324:
1325: In the following, we calculate the off-diagonal self-energy
1326: $\Pi_{12}(\omega,p)$ explicitly. The expression is
1327: \bqa\nonumber
1328: \Pi_{12}^{}(\epsilon(p),p)&=&g^2v^2
1329: \int{d\omega\over2\pi}
1330: \int{d^dk\over(2\pi)^d}
1331: \\ &&\nonumber
1332: \hspace{-2.4cm}
1333: \times\Bigg\{
1334: {3\left[\omega+\epsilon(p)\right]k^2
1335: \over[(\omega+\epsilon(p))^2-\epsilon^2(|{\bf p}+{\bf k}|)+i\varepsilon]
1336: [\omega^2-\epsilon^2(k)+i\varepsilon]}
1337: \\ &&
1338: \hspace{-2.5cm}
1339: -
1340: {\left[\omega+\epsilon(p)\right]\epsilon^2(k)/k^2
1341: \over\left[(\omega+\epsilon(p))^2-\epsilon^2(|{\bf p}+{\bf k}|)
1342: +i\varepsilon\right]
1343: \left[\omega^2-\epsilon^2(k)+i\varepsilon\right]}
1344: \Bigg\}\;,
1345: \eqa
1346: %where the superscript indicates the number of loops in the quantum loop
1347: %expansion.
1348: After integrating over $\omega$, the imaginary
1349: part of $\Pi_{12}(\epsilon(p),p)$
1350: becomes
1351: \bqa\nonumber
1352: {\rm Im}\,\Pi_{12}^{}(\epsilon(p),p)&=&-g^2v^2
1353: %\left({e^{\gamma}M^2\over4\pi}\right)^{\epsilon}
1354: \int{d^dk\over(2\pi)^d}
1355: \\ && \nonumber
1356: \times\Bigg\{
1357: {[\epsilon(p)+\epsilon(k)][3k^2/\epsilon(k)-\epsilon(k)/k^2]
1358: \over\left[\epsilon^2(|{\bf p}+{\bf k}|)-\epsilon^2(p)-\epsilon^2(k)\right]}
1359: \\ &&\nonumber
1360: %\hspace{-1cm}
1361: +{3k^2-\epsilon^2(k)/k^2
1362: \over\left[(\epsilon(k)+\epsilon(p))^2
1363: -\epsilon^2(|{\bf p}+{\bf k}|)\right]}
1364: \Bigg\}\;.
1365: \\ &&
1366: \label{pi12t}
1367: \eqa
1368: Finally, we expand Eq.~(\ref{pi12t})
1369: in powers of the external momentum $p$. After
1370: simplifying using Eqs.~(\ref{alge1})--(\ref{alge3}), the self-energy
1371: reduces to
1372: \bqa\nonumber
1373: {\rm Im}\,\Pi_{12}^{}(\epsilon(p),p)&=&{1\over8}g(2gv^2)^{3/2}
1374: \left[3I_{1,3}-I_{-1,1}
1375: \right]p
1376: \\ &&
1377: +{\cal O}\left(p^3/g^{3/2}v^3\right)\;.
1378: \label{p1}
1379: \eqa
1380: Notice in particular that $\Pi_{12}^{}(0,0)$ vanishes. This
1381: property holds to all orders in perturbation theory and follows from
1382: time-reversal invariance. Note also that
1383: $\Pi_{21}^{}(\omega,p)=-\Pi_{12}^{}(\omega,p)$, which also holds
1384: to all orders in loop expansion.
1385: The real part of the self-energies $\Pi_{11}(\epsilon(p),p)$ and
1386: $\Pi_{22}(\epsilon(p),p)$ are expanded
1387: about zero external momentum in the same way.
1388: Including the mean-field self-energies, one finds:
1389: \bqa
1390: \nonumber
1391: {\rm Re}\,\Pi_{11}(\epsilon(p),p)&=&3gv^2+
1392: {1\over4}g\Big[
1393: 3I_{1,1}+I_{-1,-1}-gv^2\left(9I_{2,3}
1394: \right.\\ && \left.
1395: -6I_{0,1}+I_{-2,-1}
1396: \right)
1397: \Big]
1398: +{\cal O}\left(p^2/gv^2\right)
1399: \label{p2}
1400: \;,\\ \nonumber
1401: {\rm Re}\,\Pi_{22}(\epsilon(p),p)&=&gv^2+
1402: {1\over12}g\Big[9I_{1,1}+3I_{-1,-1}
1403: \\ &&%\hspace{-1.2cm}
1404: \nonumber
1405: -p^2gv^2\left(6I_{3,5}-11I_{1,3}+3I_{-1,1}
1406: \right)\Big]
1407: \\ &&
1408: +{\cal O}\left(p^4/(g^2v^4\right)
1409: \;.
1410: \label{p3}
1411: \eqa
1412: The expressions for the
1413: self-energies $\Pi_{11}(\epsilon(p),p)$, $\Pi_{22}(\epsilon(p),p)$
1414: and $\Pi_{12}(\epsilon(p),p)$
1415: are infrared
1416: divergent. The integrals $I_{-2,-1}(2gn_0)$ and $I_{-1,1}(2gn_0)$
1417: both have
1418: a logarithmic divergence as the loop momentum $k$ goes to zero.
1419: These divergences show up as poles in $\epsilon$ in
1420: Eqs.~(\ref{ir1})--(\ref{ir2}).
1421: However, it is important to point out that they
1422: cancel in the final results for physical quantities, as we shall see
1423: below.
1424:
1425: The real part of equation~(\ref{colll}) can now be written as
1426: \bqa\nonumber
1427: [\omega-{\rm Im}\,\Pi_{12}(\epsilon(p),p)]^2&=&
1428: \left[p^2-\mu+{\rm Re}\,\Pi_{11}(\epsilon(p),p)\right]
1429: \\ &&\hspace{-1cm}
1430: \times
1431: \left[p^2-\mu+{\rm Re}\,\Pi_{22}(\epsilon(p),p)\right]\;.
1432: \label{cnew}
1433: \eqa
1434: The next step is to eliminate the chemical potential from Eq.~(\ref{cnew})
1435: by minimizing the one-loop thermodynamic potential
1436: $\Omega_{0+1}(\mu,v)$, This is
1437: found by differentiating the sum of Eqs.~(\ref{o0}) and~(\ref{2om})
1438: with respect to the condensate $v$
1439: and setting it to zero. This yields
1440: \bqa\nonumber
1441: 0&=&-\mu+gv^2
1442: +{1\over4}g
1443: %\left({e^{\gamma}M^2\over4\pi}\right)^{\epsilon}
1444: \int{d^dp\over(2\pi)^d}{3(p^2-Y)+(p^2-X)
1445: \over\sqrt{(p^2-X)(p^2-Y)}}
1446: \;.
1447: \\ &&
1448: \label{gol}
1449: \eqa
1450: It is consistent to the order in quantum corrections at which we are
1451: calculating, to evaluate the one-loop contribution to~(\ref{gol})
1452: at the classical minimum $v=v_0$.
1453: Eq.~(\ref{gol}) then reduces to
1454: \bqa\nonumber
1455: 0&=&-\mu+gv^2+{1\over4}g\left[3I_{1,1}+I_{-1,-1}\right]\\
1456: &=&-\mu+\Pi_{22}(0,0)\;.
1457: \label{eqs}
1458: \eqa
1459: Eq.~(\ref{eqs}) shows that the Goldstone theorem is satisfied
1460: at the stationary point of the thermodynamic potential.
1461: Using (\ref{t0n}), we see that the chemical potential obtained
1462: from~(\ref{muuu}) agrees with~(\ref{eqs}).
1463: Solving for the chemical potential and substituting the result as well
1464: as the self-energies given by Eqs.~(\ref{p1})--(\ref{p3})
1465: into Eq.~(\ref{cnew}), we obtain
1466: \bqa\nonumber
1467: &&
1468: \left[\omega+{1\over8}g(2gv^2)^{3/2}
1469: \left(3I_{1,3}-I_{-1,1}
1470: \right)p\right]^2=
1471: \\ && \nonumber
1472: p^2\left[p^2+2gv^2\left(1-{1\over8}g
1473: (9I_{2,3}+I_{-2,-1}-6I_{0,1})\right)\right] \\
1474: &&
1475: \times
1476: \left[1-{1\over12}g^2v^2\left(6I_{3,5}-11I_{1,3}+3I_{-1,1}
1477: \right)\right]\;.
1478: \label{om}
1479: \eqa
1480: We can now solve for $\omega$ in Eq.~(\ref{om}).
1481: In the long-wavelength limit, $p\ll\sqrt{2gv^2}$, we obtain
1482: \bqa\nonumber
1483: {\rm Re}\,\omega(p)&=&p\sqrt{2gv^2}\left[1-{g^2v^2\over24}
1484: \left(6I_{3,5}+7I_{1,3}-3I_{-1,1}
1485: \right)
1486: \right]\\ && \nonumber
1487: \times
1488: \left[1-{1\over16}g\left(
1489: 9I_{2,3}-6I_{0,1}+I_{-2,-1}
1490: \right)\right]\\
1491: &=&p\sqrt{2gn_0}\left[
1492: 1+{28\over3}\sqrt{n_0a^3\over\pi}
1493: \right]\;.
1494: \label{pre}
1495: \eqa
1496: The infrared divergent terms that appear in the
1497: expressions for the self-energies
1498: cancel algebraically after having used
1499: the relations~(\ref{alge1})--(\ref{alge3}).
1500: The ultraviolet divergences associated with the integrals in
1501: Eq.~(\ref{pre}) are again power divergences. Thus one immediately
1502: obtains a finite result and this is another example of the
1503: convenience of employing dimensional regularization.
1504: The result~(\ref{pre}) was first obtained by Beliaev~\cite{beli}.
1505: One can easily check that the slope of $\omega$ in Eq.~(\ref{pre}) is the same
1506: as the macroscopic speed of sound, $c$, that one obtains from differentiating
1507: the pressure with respect to the number density~\cite{fetter}.
1508: This equality has been proven
1509: to all orders in perturbation theory by Gavoret and Nozieres~\cite{gav}, and by
1510: Hohenberg and Martin~\cite{hohen}.
1511:
1512: The effective field theory approach presented in this section
1513: is based on a cartesian parametrization of the quantum field
1514: $\tilde{\psi}$ in Eq.~(\ref{split}).
1515: A similar effective field theory approach
1516: was developed by Popov~\cite{popov2} and later extended by Liu~\cite{w1,w2}.
1517: Instead of using a cartesian parametrization of the field $\psi$,
1518: it was parametrized using the density $n$
1519: and the phase $\chi$:
1520: \bqa
1521: \psi({\bf x},t)&=&\sqrt{n({\bf x},t)}\;e^{i\chi({\bf x},t)}\;.
1522: \label{np}
1523: \eqa
1524: The density field is shifted in analogy with the shift in Eq.~(\ref{shift}):
1525: \bqa
1526: n({\bf x},t)&=&v^2+\sigma({\bf x},t)\;.
1527: \label{shift2}
1528: \eqa
1529: We recall that the condition~(\ref{condit}) simplifies calculations,
1530: since it makes all one-particle reducible diagrams vanish.
1531: This condition is replaced by
1532: \bqa
1533: \langle\sigma\rangle&=&0\;.
1534: \eqa
1535: Using the parametrization~(\ref{np}) together
1536: with~(\ref{shift2}), the action~(\ref{act}) takes the form
1537: \bqa\nonumber
1538: S&=&S[v]+\int dt\int d^dx\left[{1\over v}J\sigma
1539: +{1\over2}\left(\chi\dot{\sigma}-\dot{\chi}\sigma
1540: \right)
1541: -{1\over2}g\sigma^2
1542: \right.\\ \nonumber
1543: &&\left. \hspace{2.3cm}
1544: -v^2(\nabla\chi)^2-{1\over4v^2}(\nabla\sigma)^2
1545: -\sigma(\nabla\chi)^2
1546: \right.\\ &&\left.\hspace{2.3cm}
1547: +{1\over4v^4}\sigma\left(\nabla\sigma\right)^2+...
1548: \right]\;,
1549: \label{s22}
1550: \eqa
1551: where the dots indicate an infinite series of higher order
1552: operators.
1553: The classical action $S[v]$ and the source $J$ are given by Eqs.~(\ref{cact})
1554: and~(\ref{t}), respectively.
1555: After a rescaling of the fields $\sigma$ and $\chi$, the free propagator
1556: corresponding to the action~(\ref{s22}) is identical to Eq.~(\ref{prop2}).
1557: However, the interaction vertices
1558: are different. In particular, the vertex
1559: corresponding to the operator
1560: $\sigma(\nabla\chi)^2$ is momentum
1561: dependent. One feature of the perturbative expansion that follows from
1562: the action~(\ref{s22}) is the absence of infrared divergences in
1563: individual
1564: diagrams~\cite{popov,ericag}. The momentum dependence of the trilinear
1565: interaction $\sigma(\nabla\chi)^2$ compensates the singular behavior
1566: of the propagator at low momenta.
1567: This difference should not be viewed as
1568: being fundamental since the infrared divergences always cancel in physical
1569: quantities. We have already seen one example of this, when we
1570: considered the quantum correction to the Bogoliubov dispersion relation.
1571: On the other hand, individual diagrams are more severely ultraviolet
1572: divergent, but these cancel when the diagrams are added.
1573: The above features merely represent a different
1574: way of organizing the perturbative calculations. The equality of
1575: calculations of physical quantities order by order in perturbation theory
1576: simply reflects the reparametrization invariance
1577: of the functional integral.
1578:
1579:
1580:
1581: As mentioned before, the imaginary part of the dispersion relation
1582: represents the damping of the collective excitations.
1583: In the case of a dilute Bose gas,
1584: this term was first calculated by Beliaev~\cite{beli}.
1585: In the long-wavelength limit, his
1586: calculations showed that the damping rate is proportional to $p^5$:
1587: \bqa
1588: \gamma(p)&=&{3p^5\over320\pi n_0}\;.
1589: \label{imw}
1590: \eqa
1591: The imaginary part $\gamma(p)$
1592: is connected with one phonon decaying into two phonons
1593: with lower energy. It is often referred to as Beliaev damping.
1594: The action~(\ref{s22}) has later been used by several
1595: authors~\cite{popov,w1,w2} to rederive
1596: the result~(\ref{imw}).
1597: The calculations represent a significant
1598: simplification compared to the original derivation.
1599:
1600:
1601: \subsection{Nonuniversal Effects}
1602: \label{noneff}
1603: In the previous subsection, it was shown that the dominant effects of the
1604: interaction between the atoms could be subsumed in a single
1605: coupling constant called the $s$-wave scattering length.
1606: Thus all interatomic potentials with the same $s$-wave scattering length
1607: will have the same properties to leading order in the
1608: low-density expansion, and this property is called universality.
1609: However, at higher orders in the low-density expansion, physical
1610: quantities will depend on the
1611: details of the interatomic potential such as the effective range $r_s$.
1612: These are called nonuniversal effects. A detailed analysis of nonuniversal
1613: effects can be found in the paper by Braaten, Hammer and Hermans~\cite{e2}.
1614: We discuss these next.
1615:
1616: Including the operator
1617: $\left[\nabla(\psi^*\psi)\right]^2$ in Eq.~(\ref{act}), we can
1618: again calculate exactly the scattering amplitude for $s$-wave scattering.
1619: Summing the contributions from the diagrams in Fig.~\ref{loopfig},
1620: we obtain~\cite{e2}
1621: \bqa
1622: {\cal T}(q)&=&-\left[{1\over2g+4hq^2}+i{q\over16\pi}\right]^{-1}\;.
1623: \label{m2}
1624: \eqa
1625: The coupling constant $h$ is then related to the effective range
1626: $r_s$ of the true potential~\cite{e2} and it is determined by
1627: matching Eqs.~(\ref{m1}) and~(\ref{m2}) through third order in $q$.
1628: This yields
1629: \bqa
1630: h&=&%8\pi a{ar_s\over4}\;.
1631: 2\pi a^2r_s\;.
1632: \label{hdef}
1633: \eqa
1634: After performing the shift~(\ref{shift}), the operator
1635: $\left[\nabla(\psi^*\psi)\right]^2$ also
1636: contributes to the free part of the action.
1637: Including the effects of this operator, one obtains a modified
1638: propagator and a modified dispersion relation:
1639: \bqa
1640: D(\omega,p)&=&\frac{i}{\omega^2
1641: -\epsilon^2(p)+i\varepsilon}
1642: \left(\begin{array}{cc}
1643: p^2&-i\omega \\
1644: i\omega&\epsilon^2(p)/p^2
1645: \end{array}\right)\;,
1646: \label{propexp}
1647: \eqa
1648: where the new dispersion relation is
1649: \bqa
1650: \epsilon(p)&=&p\sqrt{(1+2hv^2)p^2+2gv^2}
1651: \label{newspec}
1652: \;.
1653: \eqa
1654: The spectrum~(\ref{newspec}) has the Bogoliubov form with modified
1655: coefficients. It is therefore straightforward to recalculate the
1656: ground state energy density and the depletion of the condensate
1657: by rescaling the momentum ${\bf p}\rightarrow {\bf p}\sqrt{1+2hv^2}$.
1658: For instance, the expression for the number density is in analogy with
1659: equation~(\ref{t0n}) given by
1660: \bqa\nonumber
1661: n_{0+1}(n_0)&=&n_0
1662: +{1\over4}
1663: %\left({e^{\gamma}M^2\over4\pi}\right)^{\epsilon}
1664: \left[\int{d^dp\over(2\pi)^d}
1665: {p^2\over\epsilon(p)}+{\epsilon^2(p)\over p^2}
1666: \right]\\ \nonumber
1667: &=&n_0+{1\over4}\left[1+2hv^2\right]^{-{d+1\over2}}I_{1,1}
1668: \\ &&
1669: +{1\over4}\left[1+2hv^2\right]^{-{d-1\over2}}I_{-1,-1}\;.
1670: \eqa
1671: The limit $d\rightarrow3$ is regular and we obtain
1672: \bqa
1673: n_{0+1}(n_0)&=&n_0\left[1+
1674: {1\over24\pi^2}{\sqrt{8n_0g^3}\over(1+2hv^2)^2}\left(1-2hv^2\right)
1675: \right]\;.
1676: \eqa
1677: Using Eqs.~(\ref{ga}) and~(\ref{hdef}), and
1678: expanding to first order in the effective range
1679: $r_s$, we find
1680: \bqa\nonumber
1681: n_{0+1}(n_0)&=&n_0\left[1+
1682: {8\over3}\sqrt{n_0a^3\over\pi}
1683: \right.\\ &&\left.
1684: -{32\pi^2r_s\over a}\left(
1685: {n_0a^3\over\pi}\right)^{3/2}\right]
1686: \;.
1687: \eqa
1688: Similarly, we can calculate the ground state energy density.
1689: Expanding to first order in the effective range
1690: $r_s$, one finds~\cite{e2}:
1691: \bqa\nonumber
1692: {\cal E}_{0+1}(n)&=&
1693: 4\pi an^2\left[
1694: 1+{128\over15}\sqrt{{na^3\over\pi}}
1695: \right.
1696: \\ &&
1697: -{1024\pi^2r_s\over15a}\left({na^3\over\pi}\right)^{3/2}
1698: \Bigg]
1699: \label{expl}
1700: \;.
1701: \eqa
1702: Effective field theory can be used to determine at which order in the
1703: low-density expansion a given operator starts to contribute to
1704: various physical quantities. As an example, we consider the
1705: energy density ${\cal E}$. Each power of $\psi^*\psi$ contributes
1706: a factor of $n$. Each power of $\nabla$ contributes a factor of $\sqrt{na}$.
1707: Each loop order in the quantum loop expansion contribitutes a factor of
1708: $\sqrt{na^3}$. The derivative interaction $[\nabla(\psi^*\psi)]^2$
1709: does not contribute to the energy density at the mean-field level for
1710: a homogeneous Bose gas since $\nabla v$ obviously vanishes in this case.
1711: It first contributes at the one-loop level and this gives one factor
1712: of $\sqrt{na^3}$
1713: There are two powers of $\psi^*\psi$ and two powers of $\nabla$, which give
1714: two factors of $n$ and $\sqrt{na}$, respectively. This yields a contribution
1715: proportional to $r_sn^2({na^3})^{3/2}$ in accordance with the explicit
1716: calculation~(\ref{expl}).
1717:
1718:
1719: We next comment on the nonuniversal effects that arise in
1720: higher-order calculations
1721: of the energy density. For simplicity, we ignore
1722: all other coupling constants than $g$.
1723: The leading term is the
1724: mean-field contribution given in~(\ref{e0}).
1725: A one-loop calculation gives rise to a
1726: universal correction proportional to $\sqrt{na^3}$ shown in~(\ref{ed}).
1727: At the two-loop level, one
1728: encounters two terms. The first is a universal logarithmic
1729: corrections proportional to $na^3\log na^3$. This
1730: term was first calculated by
1731: Wu~\cite{wu}. The second term is a nonuniversal term proportional
1732: to $na^3$.
1733: This term was first calculated
1734: by Braaten and Nieto~\cite{eric}.
1735: It comes about as follows. At the two-loop level, there is a
1736: logarithmic ultraviolet divergence
1737: that cannot be cancelled by a local
1738: two-body counterterm of the form $\delta g(\psi^*\psi)^2$.
1739: The only way to cancel it, is to add to the Lagrangian
1740: a local counterterm of the
1741: form $\delta g_3(\psi^*\psi)^3$ and absorb the divergence in the
1742: coefficient of this operator. One can see the necessity of such
1743: an operator by considering
1744: $3\rightarrow3$ scattering. At the two-loop level, or fourth order in $a$,
1745: there are Feynman diagrams that depend logarithmically on the
1746: ultraviolet cutoff. They give an additional momentum-independent
1747: contribution to the $3\rightarrow3$
1748: scattering amplitude.
1749: In order to reproduce the low-energy scattering
1750: of three atoms, the local momentum-independent
1751: operator $\delta g_3(\psi^*\psi)^3$
1752: must be included in the effective Lagrangian. The counterterm of this
1753: operator that removes the logarithimic divergences from the
1754: $3\rightarrow3$ scattering amplitude is then exactly the same
1755: counterterm needed to remove the logarithmic divergence
1756: in the energy density.
1757:
1758:
1759: The coefficient $g_3$ generally depends
1760: on the properties of the two-body and
1761: three-body potentials.
1762: The operator $(\psi^*\psi)^3$
1763: in Eq.~(\ref{act}) takes into account not only the contribution from
1764: $3\rightarrow3$ scattering
1765: from a possible three-body potential, but also the contribution from
1766: the successive $2\rightarrow2$ scattering via the potential $V_0({\bf x})$.
1767: One way to determine the coefficient $g_3$ would be to solve the
1768: $3\rightarrow3$ scattering problem for the potential $V_0({\bf x})$.
1769: Alternatively, one can determine it from
1770: calculating the ground state energy density of
1771: bosons interacting through $V_0({\bf x})$.
1772: Such a strategy was recently used by Braaten, Hammer and
1773: Hermans~\cite{e2} using the Monte Carlo calculations of the
1774: condensate fraction and energy density for four different model potentials
1775: by Giorgini, Boronat, and Casulleras~\cite{georg}.
1776: The four potential were a hard-sphere potential with radius $a$,
1777: two soft-sphere potentials with height $V_0$ and radii $R=5a$ and $R=10a$, and
1778: a hard-sphere square-well potential with depth $V_0$ with inner and outer radii
1779: of $R=a/50$ and $R=a/10$, respectively.
1780: These four potentials
1781: all have the same $s$-wave scattering length $a$,
1782: but different effective range $r_s$.
1783: By calculating the energy density for the homogeneous Bose in the low-density
1784: expansion and matching it onto the Monte Carlo results,
1785: Braaten, Hammer and Hermans were able
1786: to estimate the coefficient $g_3$. Due to the large statistical errors,
1787: they could not find any deviation from universality in the three-body
1788: contact parameter. In order to determine
1789: the coefficient more accurately, one needs
1790: data with higher statistics at various densities.
1791: