1. Introduction
Continuous phase transitions controlled by an order parameter play an essential role in the description of condensed systems. This applies equally to the well-known phase transitions, which have been included in textbooks for hundreds of years, and to the exotic ones, obtained relatively recently in powerful fields or at superhigh pressures. A wide variety of critical points are known: these points represent a set of thermodynamic parameters near which the behavior of the system is non-analytical. The anomalies demonstrate “universality”, expressed, for example, in the existence of universal critical indices. The latter is determined by rather general properties of the system under consideration, for example, by the tensor structure of the order parameter and space dimension.
Despite a rich history, the theory of critical behavior began to develop only in the 20th century, when L.D. Landau proposed the simplest description of continuous phase transitions based on the analysis of symmetry and dimension [
1,
2]. However, Landau’s theory was based on an incorrect account of fluctuations in the considered systems and therefore gave incorrect predictions. Significant progress in the theory of critical phenomena was associated with the emergence of the so-called Wilson renormalization group (RG) [
3,
4,
5] and the
-expansion. Here, for the first time, it was possible to take into account the renormalization of critical exponents due to fluctuation effects. Similar techniques have yielded both the static properties of the critical behavior and the dynamic properties of equilibrium fluctuations, which describe the growth of the relaxation times in the vicinity of the critical point. Unfortunately, the consistent application of the Wilsonian RG turned out to be associated with considerable computational difficulties. Therefore, in this computational formalism, it was only possible to obtain significant results within the scope of the first-order perturbation theory. Of course, such accuracy did not correspond to experimental expectations and did not allow making practically important predictions, which the theory of condensed matter so badly needed. Fortunately, it turned out that the equivalent and more advanced technique of quantum field RG [
6,
7,
8] solves the problem much better. On its basis, the development of theoretical studies of critical behavior continued.
In the theory of critical behavior, there are quite a few different renormalization group approaches. Among them, in addition to the Wilsonian RG mentioned above, we will name the RG in the real space dimension [
9,
10,
11] as well as the functional RG [
12]. Our article is mainly devoted to the standard quantum field RG approach. It allows the calculation of multiloop diagrams most efficiently with proper use of the arbitrariness of renormalization scheme. The result for critical exponents is constructed as a series in
parameter, which is the deviation of the space dimension from the logarithmic dimension, for which the Lagrangian coupling constant has specific scaling properties. In presenting the principles of multiloop calculations, we assume that the reader is familiar with quantum field renormalization group methods in statistical physics as well as with the theory of ultraviolet renormalization. For our part, we recommend here the classic monographs [
6,
7,
8]. In the process of time-consuming multiloop calculations, it turned out that progress is influenced not only by technical methods but also by the regularization and renormalization schemes used. For dimensional regularization, the minimal subtraction scheme in the massless theory is the most efficient (see, for example, [
8] and references there). With this approach, the best results of multiloop calculations in critical dynamics are achieved.
In this article, we will refer to the results obtained in the study of “critical dynamics”. Here, we have in mind the dynamics of fluctuations near the equilibrium state in the vicinity of the critical point. It should be noted that the presented methods for calculating diagrams and for resummation of perturbation series are also successfully applied to models studied by the renormalization group in the real space dimension and to microscopic quantum models. Furthermore, they are also successfully used in various stochastic models that do not tend to the Gibbs equilibrium in the limit of long relaxation time. However, these models still need to sufficiently implement multiloop calculations.
The article is organized as follows.
Section 1.1 outlines the main diagram calculation techniques developed for static field theories. The basic models of critical dynamics are described in
Section 2.
Section 3 presents the currently developed methods for calculating dynamics diagrams and a simple illustration of some calculations. The fluctuation–dissipation theorem leads to the connections between the static and corresponding dynamics models. This leads to significant simplifications in the multiloop dynamics calculations. This is a basis of diagram reduction for the critical dynamics models described in
Section 4, where the result of five-loop calculations for the
symmetric model
A is presented.
Section 5 is devoted to methods of resuming divergent series in critical dynamics models. The resummed values of dynamic critical exponents
z for the model
A are presented.
1.1. Basic Diagram Calculation Methods
The standard
model, illustrating multiloop calculation methods, is considered as the theory of the scalar field
over the Euclidean space of dimension
. The system is described by the action
As is customary in the literature, in similar expressions below, we will omit integration over spatial coordinates and time. Although the investigated functional has the meaning of a Hamiltonian, in the quantum field approach, it is denoted by
S, inheriting the notation from the relativistic QFT in the Minkowski space. The model under consideration is multiplicatively renormalizable. The renormalization is performed by stretching the fields
and the parameters
,
, where
is the renormalization mass. In the minimal subtraction scheme, the renormalization constants have purely pole contributions generated by divergent diagrams
The method of studying critical behavior is based on the RG equation (see, for example, [
8] and references therein)
where
is the scale parameter and
is the
n-point correlation function. The so-called renormalization group functions
and
(here
are the canonical dimensions of
i,
correspond to the anomalous dimensions) are expressed in a known way in terms of the renormalization constants [
8]. Small
s corresponds to the large-scale—or infra-red (IR)—behavior of the system’s correlators. The IR region is reached at
, for example, when
p is the only argument of the correlation function in the momentum representation. Thus, the standard problem in this approach is to calculate the contributions of all diagrams to the renormalization constants for a given order of the perturbation expansion. Further, a fairly routine recalculation of them into the
-expansion coefficients of critical exponents is required. Let us list the main techniques for calculating diagrams that have been developed in statics field theories [
8,
13].
Alpha-representation. It consists of the integral representation of a power factor in the denominator of the diagram in the form
and the subsequent replacement of the orders of integration in the diagram. As a result, the
d-dimensional integrals over the momenta are calculated first, and the parametric integrals are processed at the end.
Let us consider an example of applying the mentioned techniques to calculate the pole singularity of the following diagram in the dimensional regularization
The calculation of the last expression is simplified if we recall the independence of the calculated pole contribution in
from the external momentum
p. Setting
and using the
-representation, we obtain
The pole contribution in in this expression does not depend on , which is the basis for using the “massless” calculation scheme with . In more complex diagrams, after calculating the d-dimensional integrals over the arguments of the propagators, a nontrivial problem is the evaluation of the integrals over the remaining set of parameters. Let us designate this set of parameters and and list the main techniques employed to compute them.
Sector decomposition method. Within this machinery [
14] the parametric multiple integral is divided into sectors
,
, etc., and then the appropriate scaling of variables
is performed in each sector. As a result, all integrals again take the form
. The integral over the large parameter
z is easy to calculate. It produces poles in
. The sector decomposition procedure can be repeated for diagrams with divergent subgraphs to obtain a convergent integral.
Kompaniets–Panzer method. The method [
15] is based on the fact that integrals over parameters
of alpha-representations in the diagrams without divergent subgraphs or in the subgraphs can be calculated with a certain order of integration using analytical formulas for polylogarithms.
Alternating transitions to the coordinate and momentum representations. A significant simplification in analytical calculations was given by the method presented in [
8], using formulas for the Fourier transform
of arbitrary power functions in an arbitrary space dimension
and formulas for the inverse Fourier transform
Here, we introduce magnitudes
For example, calculating a two-loop diagram using this technique looks like this
This expression explicitly contains the required
pole contributions. For calculating more complex diagrams, the convolution of lines formula is also useful:
Further development of this method was based on the use of the vertex uniqueness method [
16] and the
-operation [
17,
18,
19,
20]. The described methods made it possible to evaluate the critical exponents in the
static model up to six-loop diagrams [
15] and present the critical indices in the form of the
-expansion up to the corresponding order. As a rule, the
-series obtained in the static models turns out to be asymptotic, so the resulting segment of the series required a resummation procedure.
2. Models of Critical Dynamics
The study of static properties of correlated equilibrium systems in the vicinity of the critical point or close to the continuous phase transitions is a major area of statistical field theory. However, time-dependent phenomena, e.g., relaxation dynamics, are more diverse and no less interesting because they capture inter-mode interaction effects, which die out in equilibrium. A key experimental manifestation of dynamic effects is critical slowing down in the relaxation time, which increases abnormally. The most investigated and known models of critical dynamics are listed in [
21] and in more detail discussed in [
8,
22,
23].
Critical dynamics is described by a set of physically meaningful fields
, which obey the system of stochastic differential equations. For a wide class of systems, these equations can be written in the general Langevin form
where
is a symmetric operator
,
is a skew-symmetric one
and satisfies the condition
, where summation over the index
a is assumed [
8]. It may be verified that this condition is sufficient to ensure relaxation towards an equilibrium state with the Gibbs distribution functional
. The additive noise
is the Gaussian random field and is fully characterized by the first two moments
The explicit form of the static action S, which encodes the types of underlined symmetry and strengths of the assorted interactions, depends on the phase transitions or critical points we explore.
Moving on, let us review the paradigmatic models which are at the heart of a tremendous number of research papers on critical dynamical behavior. The simplest one is
the model A that encompasses the critical dynamic of a single nonconserved order parameter. The corresponding action is given by
The model B is determined by the same static action, Equation (10), and describes relaxation of the conserved field, resulting in the differential operator with a positive kinetic coefficient .
The rich collective dynamics of a nonconserved field
and conserved scalar one
m are captured by the effective action of
the model C
with a positive set of coefficients
and two independent coupling constants
.
The study of fully conserved dynamic of two fields
and
m with the equilibrium action, Equation (11), gives rise to
the model D, where the operator
has the form
Until now, the above definitions of the stochastic models have not incorporated all the possibilities of the entire equation, Equation (8). One can specify the operator responsible for the inter-mode interaction during the relaxation process. On the one hand, additional couplings lead to a more diverse physical picture. On the other hand, this sufficiently complicates the theoretical analysis, which is reflected in the absence of multi-loop renormalization group calculations, and the available results that some authors led are currently controversial.
The static action of
the model F of the complex valued field
and the real scalar one
m is
while
where
are additional real coupling constant. We also can arrive at
the model E by setting
and
.
The model G describes relaxation of the nonconserved
and conserved
three-component (
) real fields with the action
and
For notational clarity, one can concatenate these fields and introduce the six-component “doublet”
, such that
and
. Then, the
matrix
is determined by its elements
with a coupling constant
v.
Model H is based on the action
where
is a three-component divergence-free velocity field and
with a real coupling constant
v, and the superscript
T denotes the transpose of an operator.
The variety of non-equilibrium physical phenomena and complex systems and, as a consequence, a set of the respective stochastic field models are beyond the scope of this classification. Among them, one can refer to the most notable and key topics: fully developed turbulence [
24], advection–diffusion problems [
25], share flows [
26], chemical reactions [
27,
28], self-organized criticality [
29], random walks [
30], and Kardar–Parisi–Zhang dynamics. Overall, the diagrammatic formulation of stochastic field theory is a bit more unwieldy than that of static models. In addition to the emergence of the integration over internal frequencies, there appear new fields, various types of interaction vertices, and propagators, which bare structure may be more complicated. Thus, the implementation of the renormalization group program is a tricky business. For a long time, it was believed that the calculations in critical dynamics lag behind the gains made in static theories by at least two loops of the perturbation expansion.
One of the first who employed diagrammatic methods in the stochastic field theory was Wyld [
31]. To analyze Navier–Stokes turbulence, he put forward a systematic perturbation approach analogous to Feynman diagrams. After a while, through the work of Martin, Siggia, and Rose [
32], Janssen [
33], and De Dominicis [
34] it became clear that the study of a certain class of stochastic equations, in particular Equation (8) is tantamount to that of the effective field models with an action
. Thus, dynamical correlation functions were put in one-to-one correspondence to the Green functions determined by the generating functional
with the action
Eventually, this field theory can be renormalized, similar to the usual quantum field theory. Moreover, the renormalization constants of fields and parameters presented in both the static S and the dynamic actions are the same. This is the first simplification of calculations in dynamic models.
For the description of the critical dynamics near the superfluid phase transition, in the vicinity of the
-point, the phenomenological
F and
E models have been proposed [
21], but the characteristics of the critical dynamics of this phase transition have not been determined. Two-loop calculations did not lead to an unambiguous choice of a stable fixed point of the corresponding RG equation. Recently, the authors of [
35,
36] made theoretical progress by approaching this problem in an ab initio manner. The system has been investigated within the formalism of time-dependent Green functions at the finite temperature on the basis of the microscopic action
for boson particles with the local repulsive interaction of the strength
. Here
,
are complex conjugated fields,
denotes the particle mass and
is the chemical potential in the unit system with
,
stands for the Laplace operator. The evolution parameter
t belongs to the Schwinger–Keldysh contour
C,
Figure 1, in the complex
t-plane [
37,
38]. The reference points of the contour usually were sent to infinity
and
to enable the use of the Fourier transforms with respect to time.
To construct the perturbation expansion, the fields of the action, Equation (22), were replicated on different parts of the contour by the introduction of new fields
,
,
, where the subscripts
R,
A, and
T refer to the corresponding parts of the contour with retarded, advanced, and temperature propagators, respectively. Then, the diagram technique in the model is similar to that in the stochastic dynamics. Three-loop calculation [
39] in this model leads to the conclusion that the dynamics near
point is described by the simple two-component A model. As was shown by the analysis of stochastic equations, it is because E and F models for compressible fluids are unstable.
3. Calculation of Dynamic Diagrams
Let us consider for definiteness some basic techniques of computing the analytical expressions corresponding to ultra-violet divergent parts of dynamic diagrams at the example of ones in the simplest model
A. Using the the explicit form of the static action, Equation (10), together with the general formulae, Equation (21), one obtains the respective dynamic action
The free propagators of the theory are readily established in the
representation and take the form
The one-loop diagram shown in
Figure 2 gives rise to a renormalization of the coupling constant
g at the leading order and corresponds (here, for simplicity, we fix the incoming momentum and frequency to be zero) to the following integral
which involves divergences and thus must be suitably regularized, e.g., by applying dimensional regularization
. As a first step, we need to calculate the frequency integral using the residue theorem. The remaining momentum integral
is then amenable to further analytical evaluation based on the above mentioned methods in
Section 1. An alternative way of computing loop diagrams, which may be more appropriate in the multi-loop case, is based on the
representation, where the bare propagators cast in the form
In
the temporal variable
is the argument of the response field
. Now, the loop integral of the same diagram, Equation (25), reads as
and can again be straightforwardly evaluated in a standard manner.
However, the maximal progress in the three-loops pure analytic calculation in the dynamical A model was achieved in [
40]. There, an approach with Fourier and inverse Fourier transformations between frequency—momentum and time—coordinate representations was used.
3.1. Method of Calculation
Let us describe the calculation method [
40] in more detail. The origin of the approach lies in the possibility of performing in a closed form the Fourier transform of the function
with respect to either momentum
p or temporal variable
t to arrive at the function of
or frequency
, respectively. Comparison of these results leads to the following explicit expression for the Fourier transform from the
representation to the
one. In case
, the result for arbitrary values of
a is
where
d is the dimension of the (coordinate) space.
The main advantage of the approach considered is that the loop diagrams with lines in the form of r.h.s. Equation (29) can be simply calculated, and we need not calculate corespondent integrals in momenta and frequency. After the inverse Fourier transform of all lines into the representation, expressions in the form of l.h.s. Equation (29) must be multiplied.
3.2. Convolution of Lines
The loop-like diagrams can be quite easily calculated by the approach considered, but the convolution of subgraphs similar to Equation (7) is a rather tricky problem. Nevertheless, it should be exploited if the result leads to some new loop diagram or subgraph. Using the profitable properties, Equation (29), the propagator lines are cast into the
representation in which the convolution of two arbitrary lines is turned into a simple product
or
depending on the temporal arguments of the
-function in the propagators. The Feynman identity, Equation (4),
produces
In the second expression, the integral over
can be split into a sum of two. The first one is over
, and the second is over
. These parts contribute to different
-functions in the
representation. It is convenient to rescale the
variables to obtain integrals in new variables
in the limits
. After the inverse Fourier transform, the convolution integrals over
may be expressed as
The derived expressions are fairly voluminous, but, despite this, they can be evaluated numerically. Note that integrals in some z type variables can be calculated analytically.
3.3. Example
Here, we demonstrate the calculation method by evaluating the two loops diagram,
Figure 3a, in the massless model
A. The first step is to represent all propagators, Equation (26), in the form Equation (28). Note the arrow lines already have the necessary form. While the plain line representing the bare propagator
can be decomposed into a retarded and an advanced component
The next step is to transform both components appropriately. Since they have a common structure, we will describe transformation only for the first one, while the second can be treated analogously
Here, we got rid of the denominator by introducing an integration over an auxiliary variable s. Now, being Gaussian in the momentum p, the propagator can be brought to the representation by employing the momentum Fourier transform. Under the s-integral we recognize the function as that of the form Equation (28) with the parameters and . Performing the same steps as before, one can reduce all propagators in the diagram to a set of propagators of the form Equation (28) in the representation.
The diagram,
Figure 3a, produces a logarithmic ultra-violet divergence. To calculate this contribution to renormalization constants, we put the momentum and frequency of the external line in the upper right corner equal to zero. Then we have to calculate the amputated graph shown in
Figure 3b. To do this, we multiply the lines in the
presentation in the right loop labeled by 2 and 3. Then using Equation (30) we calculate its convolution with the upper
line marked by 1. The result of the convolution is a sum of new lines obtained. Then every one of these lines is multiplied by the lower line
marked by 4. After the transform to the
representation with the use of Equation (29), we obtain the result in the form of integrals over two variables similar to that over
s in Equation (32) and over
z in Equation (30). The
function, due to the Fourier transform, contains the leading pole in
. The second pole in
can be simply extracted from the integral over
z. The remaining integrals are readily calculated analytically or numerically. In some cases, this approach can be used to obtain closed-form expressions for the diagrams.
Note that the four loop results in the model
A have been achieved using another scheme [
41]. Unlike other approaches, a massive theory was considered. This scheme was targeted to pure numerical computations of diagrams at zero momenta and frequencies after the so-called
-operation [
8]. Nevertheless, the five loops results were achieved in [
42] using the method of sector decomposition [
14] in the massless theory. To illustrate this approach, let us overwrite the calculation of the same diagram,
Figure 3b. In the
representation, one obtains two denominators after integration over temporal variables
t of the vertexes between lines numbered
and
. These denominators are quadratic in momenta. The other two denominators one has due to lines 2 and 3. Using Equations (4) and (3), one can evaluate all integrals in momenta and obtain the integrals over four Feynman parameters
. Then we split the integration domain into sectors:
,
, etc. In each sector, the smaller variables are scaled, e.g.,
,
, etc., to restore the unit upper limits of integration. Then the integral in the largest
can be calculated analytically, and the procedure is repeated until obtaining the convergent integral in the remaining parameters. Numerical calculations are needed for the remaining integrals obtained.
4. Simplification for Critical Dynamic Models: Diagram Reduction
As one can see, the calculation of the dynamics models faces a significant amount of problems in comparison to static models. However, several approaches can help to reduce the complexity of the considered dynamics model. First, the renormalization constants of the static fields and parameters in the chosen renormalization scheme are combined with those calculated in the static theory, where the calculations are much simpler [
8]. However, the main interest of the dynamics model is in renormalization constants that correspond to the dynamics field and parameters.
The complexity of the Feynman diagrams and integrands corresponding to them increases the computational time drastically compared to the same order static models. First of all, the existence of different kinds of fields increases the number of possible propagators and, as a result, the number of diagrams in each order. For example, the statical
model in the fifth order contains only 11 two-point diagrams. However, the simplest dynamical model A, which has only one more extra field, already has 95 dynamics diagrams. Moreover, another difference from the static case, as we showed in
Section 3, is additional integration by time (or frequency). For using the same calculation technique as in a static model (in particular, the sector decomposition method), it is necessary to take the time variable integration in a diagram that generates the so-called time versions. The time version has a nontrivial structure in comparison to the static version of the diagram, which requires a slight modification of the sector decomposition. To go back to the model A, the 95 dynamics diagrams mentioned lead to 1025 corresponding time versions. This fact shows the necessity of advanced techniques for reducing the dynamics diagrams.
The problem with a nontrivial denominator was solved by the adaption of the sector decomposition method to the diagrams of critical dynamics. And the problem with the growing number of time versions was largely overcome due to the developed diagram reduction scheme. The proposed approach allows one to automate the calculation process as much as possible and use most of the tools from the static case. Below, we illustrate these procedures based on the model A.
Time versions. As an example, we consider three-loop
-diagram. Due to the translational invariance of the propagators (see Equation (26)) over time, the time at one of the vertices (for example,
for (33) can be considered to be fixed. The area of integration over the remaining times
and
is limited by the presence of the function
in the propagator
. This area can be divided into the following regions (time versions):
,
, and
. In accordance with these inequalities, the relative position of the vertices in the diagram (in decreasing order from left to right) was fixed. We represent this diagram as a sum of three contributions:
Integration over relative time in each of the time intervals is easily performed. The result can be represented in the form of diagrams containing only integration over momentum:
The lines without any hatch with momentum
k correspond to the “static propagator”
, where
. The lines with a hatch correspond to 1. Dashed lines correspond to a factor in the denominator, equal to the sum of the “energies”
lines that cross dashed lines. In particular, for the third time version, we have
Diagram Reduction Scheme The integration over time significantly decreases the number of considering integrands. However, the diagram reduction scheme that was proposed in Refs. [
41,
42], can solve this issue. The idea is based on grouping the time versions according to certain rules, which leads to the construction of the modified diagrams, the number of which is much smaller than the original.
Using diagram reduction, it is easy to show that the combination of the dynamics diagram is equal to some of one static diagrams with the same topology. It corresponds to the correlation between dynamic and static renormalization constants. To illustrate it, let us consider the sum of two-time versions for one dynamic diagram; see Equation (34). The time version here implies that the times in the vertexes in the picture are ordered due to the vertical cross-sections. The coefficients in front of them correspond to the symmetric factor.
Achieved progress. Since the 1970s, several works [
43,
44,
45,
46,
47] used a grouping of the diagrams to simplify the answer. However, none of this procedure was generalized. Thus, the automatization of the procedure was not even remotely possible. For the dynamical model A, in recent years, the generic rules of diagram grouping—the diagram reduction scheme—were proposed in [
41,
42]. That can help reduce the number of considered diagrams and simplify corresponding integrands. As a result, it leads to a significant reduction in computational time. For model A, the amount of reduced diagrams in the fifth order becomes 201, which is five times less than without reduction. The generalization of the diagram–reduction method allowed calculating the critical index z until th fifth order:
The achieved progress was expended on more complicated models. For example, in the E-model, the leading order turned out to be insufficient for determining the IR of a stable fixed point; therefore, in the papers [
45,
48] (due to disagreements between the proposed results), the second-order
-expansion was calculated. Later [
49] recalculate the second order. However, it also turned out to need to be more sufficient to answer the question about the type of critical dynamic behavior. The third order of perturbation theory can change the IR stability of the fixed points. In work [
50], the recipe for a reduction diagram was proposed, which can be easily implemented for higher orders too.
6. Conclusions
This review was devoted to modern methods for calculating critical dynamic diagrams and the results obtained with their help. For a long period of time, no noticeable progress has been observed in this area, since the calculations in dynamics are rather cumbersome compared to the static model. However, the developed modern methods and the automation of their processes, which we have described in this article, help to study significant advances in this area.
Since the calculation of the dynamic model is based on the statistical model, we started with well-known approaches for static models, such as the Alpha representation, Feynman representation, sector decomposition method, Kompaniets–Panzer method, and alternating transitions to coordinate and momentum representations. Then, we extended them to a dynamic case and showed modern techniques for simplifying calculations of dynamic diagrams based on their connection with static. We also explain the Borel resummation procedure for divergent perturbation series in dynamical models and described the necessary basis for resummation.
The advantage of these approaches is that they can be automated and easily implemented. Furthermore, these methods have already been successfully used to promote higher orders of perturbation theory. We hope that the given overview will be useful for further research on the critical dynamics and behavior of stochastic scaling.