% In: "50 years of BLTP", JINR, Dubna, 2006, p.21-30 \documentclass[12pt]{article} \usepackage{amsmath,amssymb} \textwidth 16cm \textheight 24cm \oddsidemargin -.5cm \topmargin -1.5cm \begin{document} \title{\bf Multiloop Calculations at BLTP:\\[0.4cm] 30 Years of Progress } \author{D.\,I.\,Kazakov and A.\,A.\,Vladimirov} \date{} \maketitle \vspace{-0.8cm} \begin{center} {\it Bogoliubov Laboratory of Theoretical Physics, Joint Institute for Nuclear Research, Dubna, Russia } \end{center} \begin{abstract} We briefly review the history of multiloop calculations in modern quantum field theory. They started, with our participation, in the Bogoliubov Laboratory of Theoretical Physics, JINR, more than 30 years ago, produced a number of remarkable results, and now developed into a worldwide computational industry. \end{abstract} A considerable progress achieved in multiloop calculations in 1970s-80s in the Laboratory of theoretical physics at JINR has been stimulated by the renaissance of perturbative quantum field theory (QFT) which followed the discovery of the asymptotic freedom in non-Abelian gauge theories. After several years of despair caused by a seemingly inevitable ghost-pole menace, the perturbation theory calculations were again given their chance to contribute to our knowledge of the properties of elementary particles, including strongly interacting ones -- hadrons. The first period of our activity in this field (1973-78), looking now only preparatory, was nevertheless very important. At that time, our main motivation was a problem of ultraviolet fixed point for the ``running" coupling. In quantum field theory at high energies, as a rule, we are interested in the dependence of observable quantities on momenta $p$ of interacting particles. By means of the renormalization group, this $p$-dependence can be effectively accumulated in one function $g(p)$ which is called effective charge, or running coupling. The name is due to the fact that a charge (coupling constant) $g$, being originally a true constant, now becomes a function of momenta $p$, and has to vary, or ``run", when $p$ goes to infinity. There are very few typical patterns of asymptotical behavior of $g(p)$. In the first nontrivial order of perturbation theory (called ``one loop" order due to Feynman diagrams with only one closed loop being involved) we have only two possibilities: $g(p)$ increases fast and reaches infinity at some finite value of $p$ (this corresponds to a notorious ghost pole -- not a good thing at all from physical point of view), or $g(p)$ tends to zero, and this is called {\it asymptotic freedom}. Though the rate of decrease of the running coupling in this case is rather slow, $g(p)\sim1/\ln p$, nevertheless $g(p)$ asymptotically vanish, and the corresponding particles decouple at high momenta: in effect, they become free. The phenomenon of asymptotic freedom discovered in 1973 in gauge theories with non-commutative (non-Abelian) groups, such as SU(3), made it possible to build a first reliable quantum field model of strong interactions -- Quantum Chromodynamics (QCD). Logarithmically decreasing asymptotics of the QCD effective charge appears to be in a good agreement with the present experimental data at high energies. On the theory side, asymptotically free theories are favored because they develop (in a strongly interacting model!) a new small parameter of the improved perturbation theory expansion -- a running coupling $g(p)$. However, taking into account the next (two-loop) order may drastically change the whole picture: it can produce a third kind of high-energy asymptotics, a so-called (ultraviolet) fixed point. This is the case when $g(p)$ tends to some finite value $g^*\neq0$. In principle, this effect could occur in theories troubled by a ghost pole (thus curing them) as well as in asymptotically free ones, keeping them manageable. Needless to say that in the fixed-point situation our (dis)ability to compute, or at least estimate the higher-loop corrections becomes of crucial importance. In 1973, motivated and supervised by D.\,V.\,Shirkov, we started a long programme of such calculations, firstly at the two- and three-loop orders, in a wide range of QFT models, with and without asymptotic freedom. These were by no means the first multiloop calculations known in the literature; rather, we tried to adapt the existing techniques devised for evaluating radiative corrections to the case of high-energy asymptotics treated by the renormalization group approach. Only a few aspects of our computational practice were really new, such as essential use of dimensional regularization and easy conversion from one renormalization scheme to another (Shirkov and Vladimirov, 1979). Also, the great computational facilities of the algebraic manipulation program SCHOONSCHIP were used throughout. Thus, the problem of doing tedious algebra was more or less settled. Of course, individual complicated integrals still required evaluation ``by hand". As a result, we succeeded in finding several examples of ultraviolet fixed points in different models (Belokurov, Kazakov, Slavnov, Shirkov and Vladimirov, 1973, 1974). However, these were not the results of immediate physical importance. And not only because most of the models involved were unphysical, toy models. There was a fundamental defect in the whole project: if one does discover a fixed point in the $n$-loop approximation, would it persist when $n+1$ loops are taken into account? As a rule, the answer is: No! A general reason for this is the divergent nature of perturbative series in quantum field theory. Paradoxically, the highest-precision physical predictions of Quantum Electrodynamics (QED) are actually obtained by taking several first terms of a divergent series! This seems to be a real mystery, even when an expansion parameter (a charge) is small, as in QED or QCD. But in the situation with the fixed point, this parameter (the asymptotical value $g^*$ of a running coupling) does not need to be small at all. Therefore, we cannot be sure that the fixed points found perturbatively do actually exist. So, we may say that our activity in 1973-78 produced a certain methodical, rather than physical, progress. However, it provided a base for a near breakthrough. Our principal achievements in multiloop calculations, dated roughly to 1979-83, resulted from a noticeable progress in the technique of analytic computation of higher-order renormalization group parameters, and also in half-empirical methods of ``summation" of divergent series. Let us retrace the key stages of this progress, paying main attention to the results that have been obtained at JINR. In the framework of the renormalization group method, effective charge $g(p)$ is unambiguously determined by a certain function, $\beta(g)$, with $g$ being original (not running) coupling constant of a quantum theory under consideration. This $\beta$-function is actually an infinite power series in $g$. Of course, in reality we are able to find only several first (lowest) terms of this series, by calculating the Feynman diagrams with appropriate number of loops. Thus, the one-loop calculations produce the first term of $\beta(g)$ (which already indicates the presence, or absence, of asymptotic freedom), the two-loop diagrams account for the second term, and so on. The coefficients at the specific powers of $g$ in the power series $\beta(g)$ are mere numbers: they cannot depend on momenta or masses which enter into algebraic expressions for the Feynman diagrams involved. This property has been used in the method elaborated by our group (Vladimirov, 1978, 1980) and known now as the ``infrared rearrangement", which provides highly efficient tools for computing coefficients of $\beta(g)$ by putting all, except one, momenta and masses to zero, thus making fairly manageable analytic evaluation of all the needed integrals. One cannot nullify all momenta and masses due to some subtle effects of interplay between ultraviolet and infrared divergences. Nevertheless, this method, supplemented by some very useful tricks (Gegenbauer polynomial expansion and integration by parts in the momentum space), invented by the INR group in Moscow (K.\,G.\,Chetyrkin, F.\,V.\,Tkachov, et al.) opened a possibility to compute analytically the three- and four-loop contributions to $\beta(g)$ in any renormalizable model. The other direction of activity that contributed much to the progress in multiloop calculations proved to be the development of summation techniques for divergent power series for $\beta(g)$ and the other renormalization-group functions. As mentioned above, the problem is that in the QFT perturbation theory the corresponding power series in $g$ are generally divergent so that their radius of convergence is actually zero. More precisely, these series are called ``asymptotic" in a sense of Poincar\'e, because their numerical coefficients increase like $n!$, with $n$ being a number of loops. As a result, the more terms are taken into account, the smaller are those values of the coupling constant $g$ for which the expansion is still reasonable. To overcome this difficulty and obtain a reliable information about a function from its perturbative expansion, there exist several methods for ``summation" of such series. These methods, as a rule, rely on some additional information about the relevant function, like its analytical properties in the coupling constant plane, exact asymptotics of the perturbation theory coefficients in higher orders, etc. We were not inventors of these methods, but we modified them when needed and, which was the most important, successfully applied them to extract physical predictions from the results of multiloop calculations. The main ingredient of our summation procedure (Kazakov, Shirkov and Tarasov, 1979; Kazakov and Shirkov, 1980) was a due account of analytical properties of the desired function through the use of the Laplace, or Borel, transformation. All this is usually called the Borel summation technique. After the transformation, a Borel-transformed power series contains extra $n!$ in the denominator and thus becomes convergent inside some circle. It follows from the known behaviour of the perturbation theory coefficients in higher orders which was estimated earlier for several QFT models. In particular, as was first calculated by L.\,N.\,Lipatov, in the $g\varphi^4$ theory, the coefficients of the $\beta$-function $\beta(g)=\sum \beta_n g^n$ behave like $\beta_n \sim (-a)^n n!n^{7/2}$, where $a=1/16\pi^2$. This implies that the corresponding Borel-transformed series is convergent inside the circle of radius $1/a$. The next important step is the analytical continuation from this circle to the entire integration domain which is done with the help of a special conformal map. The resulting series in a new conformal variable are to be cut at the highest known perturbation-theory coefficient. Miraculously, these almost empirical recipes did really work. We calculated the four-loop contributions (the first four terms) for $\beta(g)$ and some related functions of the scalar $g\varphi^4$ model in the framework of so-called $\varepsilon$-expansion formalism, carried out the summation procedure described above, and then applied all this to analysis of phase transitions of the second kind (Kazakov, Tarasov and Vladimirov, 1979). This is reasonable because certain characteristics of condensed matter near critical temperature, such as critical exponents, etc., are known to allow description in terms of appropriate QFT models. The results were impressive: an agreement with experiment and with other theoretical approaches was very good. In fact, this was the most precise calculation of the critical exponents reached so far, which was cited in the Nobel lecture by Kenneth Wilson. This activity found its continuation after several years when a group of people from BLTP and INR (Moscow) (K.\,G.\,Chetyrkin, S.\,G.\,Gorishny, S.\,A.\,Latin and F.\,V.\,Tkachov) developed computer codes for multiloop calculations and, to check their efficiency, performed five-loop calculations in the $g\phi^4$ theory. They also suggested a new method for evaluating complicated integrals in addition to the infrared rearrangement technique developed earlier in our laboratory. This is the so-called $R^*$-operation which combines the conventional $R$-operation developed by Bogoliubov in 1950s for eliminating ultraviolet divergences, with a similar infrared operation. The $R^*$-operation proved to be an extremely powerful computational tool because it allows to simplify the integrals by sending all external momenta to zero. The above-mentioned subtleties are circumvented here by preventive subtraction of the spurious infrared divergences. And though rigorous proof and full development of the $R^*$-operation method took several years (K.\,G.\,Chetyrkin and V.\,A.\,Smirnov; F.\,V.\,Tkachov) , its first applications were very exciting. The 5-loop calculation in $g\phi^4$ model was the top of the art at that time (and still is), though some integrals were evaluated in the form of infinite convergent series. It was natural then to look for a closed analytical answer in the form of Euler gamma functions and further Riemann zeta functions. It so happened that this task was partially fulfilled with the help of a new technique called the ``method of uniqueness", first suggested by A.\,N.\,Vasiliev with collaborators in St.\,Petersburg and then elaborated at BLTP (Kazakov, 1983, 1984). This method uses conformal properties of massless theories together with the so-called star-triangle relation (also called the uniqueness relation) for evaluating Feynman diagrams by either transforming them to exactly solvable ones or expanding them into power series in $\varepsilon$ with known coefficients. (Recently an elegant interpretation of the uniqueness relation in terms of noncommutative algebra was proposed by A.\,P.\,Isaev.) With the help of this method it became possible for the first time to complete the 5-loop calculation (Kazakov, 1985) of $\beta$-function in the $g\phi^4$ theory in explicit analytical form. Later on, this result was used for the calculation of critical exponents by the method discussed above. These achievements stimulated further activity in BLTP focused on analytical evaluation of multiloop Feynman diagrams (Gorishny and Isaev, 1985\,; Kotikov, 1991)\,. Multiloop results described above were obtained in scalar models, like $g\phi^4$. Proceeding now to our calculations in gauge theories we firstly have to mention, as our major achievement of that time, an evaluation of the three-loop $\beta$-function in QCD (Tarasov, Vladimirov and Zharkov, 1980). For technical reasons, this task is even more difficult than to compute four loops in $g\varphi^4$. The final formula (with $n_f$ being the number of flavors, i.e., sorts of quarks), $$ \beta_{QCD}(g^2) = \frac{g^4}{(4\pi)^2}(-11 + \frac{2}{3} n_f) + \frac{g^6}{(4\pi)^4} (-102 + \frac{38}{3} n_f) + \frac{g^8}{(4\pi)^6} (-\frac{2857}{2} + \frac{5033}{18} n_f - \frac{325}{54} n_f^2)\,, $$ obtained in early 1980, is actively exploited up to now in high-energy calculations based on Quantum Chromodynamics. It appeared to be one of the most frequently cited results produced in the Laboratory of theoretical physics. The next, fourth, term of $\beta_{QCD}$ has been computed only in 1997 by S\,.A.\,Larin and J.\,Vermaseren. The other example of application of the developed technique to QCD was the calculation of the structure functions of the deep inelastic scattering process in three loops. It required modification of the method of uniqueness as applied to integrals with arbitrary powers of momenta (Kazakov and Kotikov, 1987). As a result the complete quartic ($\alpha^2_s$) correction to the deep inelastic longitudinal structure function $F_L$ in QCD earlier considered by other groups was calculated (Kazakov and Kotikov, 1990; Kazakov, Kotikov, Parente, Sampayo and Sanchez-Guillen, 1990), which was the highest precision at that time. An additional problem, intimately related to any multiloop calculations, but fortunately not so serious as the calculations themselves, is a possible renormalization-scheme dependence of their results. To cope with ultraviolet divergences inevitably present in the Feynman diagrams of quantum field theory, one makes use of (a certain variant, or scheme, of) the renormalization procedure. Do final results of the corresponding computations depend on the scheme we choose? Generally they don't, provided that all orders of the perturbation expansion are taken into account. However, in practical situations it seems to be unmanageable. So, the results of calculations performed at any finite order of the perturbation theory do depend on the renormalization scheme. In the above case of $\beta_{QCD}$, such dependence starts from the three-loop order. Specifically, a numerical coefficient at $g^8$ in this formula has been found in the MS (minimal subtraction) renormalization scheme. But another calculations, using some other scheme, could produce different three- and higher-loop results. The situation looks quite unsatisfactory, but actually it is not that bad. In several papers (our paper (Vladimirov, 1979) being one of the first of them) it has been shown that the whole problem of scheme dependence can be substantially reduced by a clever choice of an expansion variable in the final expressions for appropriate physical quantities, where the $\beta$-function is only one of the ingredients. Despite the principal problem being thus fixed, a certain activity in which we also took part (Kazakov and Shirkov, 1985; Vladimirov, 1989) around the scheme dependence was noticeable during 1980s. It produced a couple of new, mostly empirical, recipes which proved to be quite useful in treating multiloop perturbative results. Almost as a by-product of our three-loop QCD calculations, we were lucky to obtain another result of great importance (maybe, rather theoretical than practical). It was known for several years that two first terms in the $\beta$-function of a special supersymmetric model, so-called $N=4$ SUSY Yang-Mills theory, surprisingly vanish. Without very great effort, by adding a set of diagrams required by supersymmetry, we succeeded in three-loop calculation (Avdeev, Tarasov and Vladimirov, 1980) of $\beta_{N=4}$ with the result being exactly zero! Of course, this still was not a proof that $\beta$-function in this model vanishes to all orders, but it was a true incentive to think so. And indeed, our result, obtained in its final form on 21 April 1980, has triggered much activity in this field. The result itself (three-loop zero) has been immediately verified by another computation. Moreover, a few years later, the all-order proof of vanishing of $\beta_{N=4}$ was presented by several groups. Thus, our 1980 work gave birth to studying the so-called finite quantum field models. These models have a twofold interest. First of all, finite field theories, or theories without ultraviolet divergences (the usual illness of any quantum field theory) are very attractive by themselves. In a sense, they realize a dream of a self-consistent quantum field theory where all quantities and parameters are directly meaningful, where we do not encounter infinite self energies, infinite polarizations, etc. Remarkable successes of the renormalization program in QFT helped us to get used to infinities implicitly absorbed into redefinitions of fields, masses, coupling constants and other quantities. But this was always considered as a weak point of a theory, if we think of it as a prescription to get meaningful results, without any divergences. Not criticizing this point of view, we talk now about a very restricted set of supersymmetric field theories which, happily, are free from these problems. These models give us an example of field theories with special properties, perhaps selected by Nature. Mathematically, these models might allow exact solutions. Thus, the already mentioned N=4 supersymmetric Yang-Mills theory is believed to be an integrable model, the property found so far exclusively in two-dimensional world. After those thrilling years of a quest for N=4 finiteness it was realized that N=2 supersymmetry was enough to get a finite model. Unfortunately these N=2 models are not explicitly applicable to high energy physics since they contain the so-called mirror particles creating severe problems for phenomenology. We were the first to propose how one can construct models with the ordinary (or N=1) supersymmetry which are also finite (Ermushev, Kazakov and Tarasov, 1987). Using the method developed earlier we performed the three-loop calculations in N=1 supersymmetric gauge theories and demonstrated that the realization of our algorithms really yields finite models. Later on we generalized the method to the theories with softly broken supersymmetry (Avdeev, Kazakov and Kondrashuk, 1998; Kazakov, 1998). These are not exactly the models that generalize the Standard Model of fundamental interactions, but rather supersymmetric Grand Unified theories with very special properties. Whether these models are chosen by Nature is the matter of future experiments. The second attractive aspect of finite theories is related to string theories. Here the multiloop calculations also played their role in determining the consistent theory. Formulation of the string theory on the world sheet leads to a two dimensional sigma model with the world-sheet supersymmetry. To be reparametrization invariant, this theory should have conformal invariance at the quantum level. This invariance is usually violated by conformal anomalies which are proportional to the renormalization group $\beta$-function. Hence, in a conformally invariant quantum theory the $\beta$-function should vanish. As mentioned above, it is ordinarily evaluated within the perturbation theory. The leading (one-loop) calculation yields the $\beta$-function proportional to the Ricci tensor of the corresponding target space manifold. Thus, the requirement of conformal symmetry, and hence of vanishing $\beta$-function, distinguishes the Ricci-flat manifolds (those with zero Ricci tensor), usually called Calabi-Yau manifolds, as possible compact components of multidimensional space-time. Of course, this leading order result is to be confirmed in higher orders. The first nontrivial contribution appears only at the four-loop level and produces a certain correction to the $\beta$-function (Grisaru, Milewski and Zanon). In this puzzling situation we participated in the 5-loop calculation which used all the skills and techniques invented earlier (Grisaru, Kazakov and Zanon, 1987). This was also a top of the art calculation that demonstrated that higher orders do not break the general statement concerning Ricci-flat manifolds. When dealing with perturbative calculations in supersymmetric theories, we encounter a bothersome problem of invariant regularization. In quantum field theory, some kind of regularization procedure is required to give a sense to (generally divergent) Feynman integrals. Actually, the regularization modifies original theory to a certain extent. In particular, it can destroy, or rather deform, its symmetry properties. But very often it is these symmetry features that made the theory to be worth dealing with! Of course, all the miraculous cancellations of divergences in supersymmetric models are due to this very supersymmetry. And we would be greatly disappointed if this symmetry do not persist at the regularized (and renormalized) level. Unfortunately, it appears to be uneasy task to adapt dimensional regularization (which is absolutely indispensable tool for multiloop calculations) to supersymmetry. Moreover, there was a wide spread opinion that it is quite impossible, and the most popular approach in this direction, so-called ``regularization by dimensional reduction", is contradictory. In 1981-82 we elaborated a method (Avdeev, Chochia and Vladimirov, 1981; Avdeev and Vladimirov, 1983) which enabled one to perform the calculations in a manifestly supersymmetric (and consistent) manner through a certain number of loops. However, at some higher order of perturbation theory (fourth or fifth, as a rule) the supersymmetry of the results could be lost. We managed to roughly estimate these potentially dangerous orders from below for the models of interest, and found, for example, that our three-loop zero result was not under suspicion (Avdeev and Vladimirov, 1983). Furthermore, the above-mentioned general arguments in favor of all-order finiteness of extended supersymmetric Yang-Mills theories seem not to be liable to this trouble at all. Looking back, we see the 30-years-long history of multiloop calculations, which included the development of cunning computational techniques, the record calculations performed in various models, as well as many successful applications to the quantum field theory, particle physics and condensed matter phenomena. In 1970s-80s we were at the center of this activity. It should be mentioned that the later development added a set of dedicated computer packages which are able not only to draw all the relevant Feynman diagrams, but also to perform automatically all the necessary Lorentz algebra, and even to evaluate multiloop integrals. In fact, we were among the first who started this activity which now has developed into a worldwide computational industry, extensively exploited in the high energy (especially QCD) calculations. The other major contribution of a more recent time period is related to the problem of evaluating Feynman diagrams with masses, where some new methods appeared. The breakthrough was achieved with the help of a mass expansion, which combines the power of asymptotic expansions with the $R^*$-operation. The very last achievement here was the analytical evaluation of a new class of integrals, the so-called boxes, which appear in many QCD problems and also in some SUSY models. Today there exists a very active and ambitious computational community around the globe (but mostly of Russian origin) which, as in the early days of QED, can perform unique calculations (and really does), thus pushing the theoretical predictions to the limit. \newpage %\begin{thebibliography}{99} \section*{References} \begin{itemize} \item[-] L.\,V.\,Avdeev, O.\,V.\,Tarasov and A.\,A.\,Vladimirov, \, Vanishing of the three loop charge renormalization function in a supersymmetric gauge theory, \,Phys.\,Lett. {\bf B96} (1980) 94. \item[-] L.\,V.\,Avdeev, G.\,A.\,Chochia and A.\,A.\,Vladimirov, \, On the scope of supersymmetric dimensional regularization, \, Phys.\,Lett. {\bf B105} (1981) 272; \item[-] L.\,V.\,Avdeev and A.\,A.\,Vladimirov, \, Dimensional regularization and supersymmetry, \, Nucl.\,Phys. {\bf B219} (1983) 262. \item[-] L.\,V.\,Avdeev, D.\,I.\,Kazakov and I.\,N.\,Kondrashuk, \, Renormalizations in Softly Broken SUSY Gauge Theories, \, Nucl.\,Phys. {\bf B510} (1998) 289. \item[-] V.\,V.\,Belokurov, D.\,I.\,Kazakov, D.\,V.\,Shirkov, A.\,A.\,Slavnov and A.\,A.\,Vladimirov, \,Ultraviolet behaviour of spontaneously broken gauge theories, \,Phys.\,Lett.\,{\bf B47} (1973) 359. \item[-] V.\,V.\,Belokurov, D.\,I.\,Kazakov, D.\,V.\,Shirkov, A.\,A.\,Slavnov and A.\,A.\,Vladimirov, \,Ultraviolet asymptotics in the presence of nonabelian gauge fields, \,Theor.\,Math.\,Phys. {\bf 19} (1974) 149. \item[-] A.\,V.\,Ermushev, D.\,I.\,Kazakov and O.\,V.\,Tarasov, \,Finite $N\!=\!1$ supersymmetric grand unified theories, \,Nucl.\,Phys. {\bf B281} (1987) 72. \item[-] S.\,G.\,Gorishny and A.\,P.\,Isaev, \,On an approach to the calculation of multiloop massless Feynman integrals, \,Theor.\,Math.\,Phys. {\bf 62} (1985) 232. \item[-] M.\,T.\,Grisaru, D.\,I.\,Kazakov and D.\,Zanon,\, Five-loop divergences for the $N\!=\!2$ supersymmetric non-linear sigma-model, \,Nucl.\,Phys. {\bf B287} (1987) 189. \item[-] D.\,I.\,Kazakov, D.\,V.\,Shirkov and O.\,V.\,Tarasov, \,Analytical continuation of perturbative results of the $g\phi^4$ model into the region $g\geq 1$, \, Theor.\,Math.\,Phys. {\bf 38} (1979) 15. \item[-] D.\,I.\,Kazakov, O.\,V.\,Tarasov and A.\,A.\,Vladimirov, \,On the calculation of critical exponents by the methods of quantum field theory, \,JETP {\bf 77} (1979) 1035. \item[-] D.\,I.\,Kazakov and D.\,V.\,Shirkov, \,Asymptotic series in quantum field theory and their summation, \,Fortsch. der Phys. {\bf 28} (1980) 465. \item[-] D.\,I.\,Kazakov, \,Calculation of Feynman diagrams by the method of uniqueness, \, Theor.\,Math.\,Phys. {\bf 58} (1984) 357. \item[-] D.\,I.\,Kazakov, \,The method of uniqueness, a new powerful technique for multiloop calculations, \,Phys.\,Lett. {\bf B138} (1983) 406. \item[-] D.\,I.\,Kazakov, \,Multiloop calculations: the method of uniqueness and functional equations, \,Theor.\,Math.\,Phys. {\bf 62} (1985) 127. \item[-] D.\,I.\,Kazakov and D.\,V.\,Shirkov, \,The account of quark masses in scheme-invariant perturbation theory in QCD, \,Sov.\,J.\,Nucl.\,Phys. {\bf 42} (1985) 768. \item[-] D.\,I.\,Kazakov, \,Finite SUSY field theories and dimensional regularization, \, Phys.\,Lett. {\bf B179} (1986) 352. \item[-] D.\,I.\,Kazakov, \,Finite $N=1$ SUSY gauge field theories, Mod.\,Phys.\,Lett. {\bf A2} (1987) 663. \item[-] D.\,I.\,Kazakov and A.\,V.\,Kotikov, \,The method of uniqueness: multiloop calculations in QCD, \,Theor.\,Math.\,Phys. {\bf 73} (1987) 348. \item[-] D.\,I.\,Kazakov and A.\,V.\,Kotikov, \,Total $\alpha_s$-correction to the deep-inelastic cross-section ratio $R=\sigma_L/\sigma_T$ in QCD, \,Nucl.\,Phys. {\bf B307} (1988) 721, {\bf B345} (1990) 299 (E). \item[-] D.\,I.\,Kazakov, A.\,V.\,Kotikov, G.\,Parente, O.\,A.\,Sampayo and J.\,Sanchez Guillen, \, Complete quartic ($\alpha^2_s$) correction to the deep inelastic longitudinal structure function $F_L$ in QCD, \,Phys.\,Rev.\,Lett. {\bf 65} (1990) 1535, {\bf 65} (1990) 2921 (E). \item[-] D.\,I.\,Kazakov, \,Finiteness of Soft Terms in Finite $N\!=1\!$ SUSY Gauge Theories, \,Phys.\,Lett. {\bf B421} (1998) 211. \item[-] A.\,V.\,Kotikov, \, Differential equations method: new technique for massive Feynman diagrams calculation, \, Phys.\,Lett. {\bf B254} (1991) 158. \item[-] D.\,V.\,Shirkov and A.\,A.\,Vladimirov, The renormalization group and ultraviolet asymptotics, \, Sov.\,Phys.\,Usp. {\bf 22} (1979) 860. \item[-] O.\,V.\,Tarasov, A.\,A.\,Vladimirov and A.\,Yu.\,Zharkov, \, The Gell-Mann-Low function of QCD in the three loop approximation, \, Phys.\,Lett. {\bf B93} (1980) 429. \item[-] A.\,A.\,Vladimirov, \, Methods of multiloop calculations and the renormalization group analysis of $\varphi^4$ theory, \, Theor.\,Math.\,Phys. {\bf 36} (1978) 732. \item[-] A.\,A.\,Vladimirov, \, Method for computing renormalization group functions in dimensional renormalization scheme, \,Theor.\,Math.\,Phys. {\bf 43} (1980) 417. \item[-] A.\,A.\,Vladimirov, \, Unambiguity of renormalization group calculations in QCD, \, Sov.\,J.\,Nucl.\,Phys. {\bf 31} (1980) 558. \item[-] A.\,A.\,Vladimirov, \,Scheme dependence in the renormalization group, \, \emph{JINR Preprint} E2-89-2 (1989). \end{itemize} \end{document}