Skip to main content
Log in

Conformal Invariance and Vector Operators in the O(N) Model

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

It is widely expected that, for a large class of models, scale invariance implies conformal invariance. A sufficient condition for this to happen is that there exists no integrated vector operator, invariant under all internal symmetries of the model, with scaling dimension \(-1\). In this article, we compute the scaling dimensions of vector operators with lowest dimensions in the O(N) model. We use three different approximation schemes: \(\epsilon \) expansion, large N limit and third order of the derivative expansion of Non-Perturbative Renormalization Group equations. We find that the scaling dimensions of all considered integrated vector operators are always much larger than \(-1\). This strongly supports the existence of conformal invariance in this model. For the Ising model, an argument based on correlation functions inequalities was derived, which yields a lower bound for the scaling dimension of the vector perturbations. We generalize this proof to the case of the O(N) model with \(N\in \left\{ 2,3,4 \right\} \).

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. This proof has been criticized in the mathematical community (see, for example, [28]). These observations apply also to the Polchinski’s work [29] and to our previous [30] and present work. In the present paper, as in most theoretical physics studies, we do not pretend to achieve an analysis with the level of rigor of the mathematical community.

  2. For \(N=1,\) the theory describes a single scalar with \({\mathbb {Z}}_2\) symmetry.

  3. At a fixed point, the running anomalous dimension \(\eta _k\) identifies with the anomalous dimension \(\eta \) which governs the decay of correlation functions at long distances.

  4. This property, called “decoupling” in [49], relies on the fact that correlation functions in non-exceptional configurations of momenta (or distances) have a finite limit when \(k\rightarrow 0\). In the case of turbulence, this property of decoupling is not fulfilled. This results in the fact that correlation functions are power-laws as in scale invariant theories, but the critical exponents for n-point correlation functions are not simply related one with another. This property is usually called multifractality (see, for example, [50]).

  5. The quantization of the eigenvalue spectrum is associated with the fact that we must only consider perturbations \(\tilde{\gamma }[\tilde{\phi }]\) which lead to correlation functions which are regular at long distances (small momenta) in the presence of the infrared regulator. Indeed, the initial condition of the flow involves correlation functions which can be Fourier transformed and are infinitely differentiable with respect to fields and wave vectors and this property is preserved by the flow.

  6. The regulator has been chosen to depend only on \(|x-y|\) and is therefore invariant under translations and rotations. These transformations therefore involve no variation of the regulator.

  7. Indeed, the Wilson–Fisher fixed point is rotational invariant but a non-zero value of \(a^i_{\mu }\) would break isotropy.

  8. Here and below we assume, as usual, that the matrix M is diagonalizable.

  9. As is well known [5], the integral \(J=1/(16\pi ^2)+{\mathcal {O}}(\epsilon ),\) independently of the particular choice of the regulator \(R_k\). This ensures the universality of the \(\beta \) functions given here.

  10. In fact, when considering the full large-N limit in the presence of couplings \(a_\mu ,\)\(b_\mu \) and \(c_\mu \) it is necessary to also rescale them (see Appendix C). However, this rescaling plays no role when considering the flows linearized in \(a_\mu ,\)\(b_\mu \) and \(c_\mu \) [see Eq. (37)]. For simplicity, we ignore this rescaling below.

  11. In the case of multicritical fixed points, the large N limit can be more subtle, see [56].

  12. The operator with 6 powers of the field and 3 derivatives has scaling dimension \(5+4\epsilon /3 +O(\epsilon ^2)\). The operators with 4 powers of the field and 5 derivatives have scaling dimension \(5 +O(\epsilon ^2),\)\(5 - 4\epsilon /9 +O(\epsilon ^2)\) and \(5 - 2 \epsilon /3 +O(\epsilon ^2)\).

  13. As it will be explained below it is important to choose centered finite differences in order to ensure that the vector operators are odd under parity.

  14. One can add further requirement, such as other cubic symmetries but we will not need them below.

  15. This implies that it is not possible to define their scaling dimension by looking at the power-law behavior of correlation functions at long distances. It is however possible to define a bona fide scaling dimension by a stability analysis of the RG flow around the fixed point. We stress that the constraint given in the introduction on the scaling dimension for the operator \({\varSigma }_\mu \) corresponds to this latter definition.

  16. It is often stated in the literature that redundant operators can be reabsorbed by a change of variables and are therefore not physically relevant. This however cannot be applied as such when testing whether a Ward identity is valid or not. Indeed, this would lead us to the absurd conclusion that a generic theory with two scalar fields can always be made O(2)-invariant by reabsorbing the redundant operator appearing in the right-hand-side of Eq. (55) through a field redefinition.

  17. We focus on the Ising case but the same discussion generalizes to other universality classes.

References

  1. Wilson, K.G., Kogut, J.B.: The Renormalization group and the epsilon expansion. Phys. Rep. 12, 75 (1974). https://doi.org/10.1016/0370-1573(74)90023-4

    Article  ADS  Google Scholar 

  2. Wetterich, C.: Exact evolution equation for the effective potential. Phys. Lett. B 301, 90 (1993). https://doi.org/10.1016/0370-2693(93)90726-X

    Article  ADS  Google Scholar 

  3. Ellwanger, U.: Collective fields and flow equations. Z. Phys. C 58, 619 (1993). https://doi.org/10.1007/BF01553022

    Article  ADS  Google Scholar 

  4. Morris, T.R.: The Exact renormalization group and approximate solutions. Int. J. Mod. Phys. A 9, 2411 (1994). https://doi.org/10.1142/S0217751X94000972

    Article  ADS  MathSciNet  MATH  Google Scholar 

  5. Berges, J., Tetradis, N., Wetterich, C.: Nonperturbative renormalization flow in quantum field theory and statistical physics. Phys. Rep. 363, 223 (2002). https://doi.org/10.1016/S0370-1573(01)00098-9

    Article  ADS  MathSciNet  MATH  Google Scholar 

  6. Delamotte, B.: An Introduction to the nonperturbative renormalization group. Lect. Notes Phys. 852, 49 (2012). https://doi.org/10.1007/978-3-642-27320-9_2

    Article  ADS  MathSciNet  MATH  Google Scholar 

  7. Zinn-Justin, J.: Quantum field theory and critical phenomena. Int. Ser. Monogr. Phys. 113, 1–1054 (2002)

    Google Scholar 

  8. Morris, T.R.: Derivative expansion of the exact renormalization group. Phys. Lett. B 329, 241 (1994). https://doi.org/10.1016/0370-2693(94)90767-6

    Article  ADS  MathSciNet  MATH  Google Scholar 

  9. Morris, T.R.: The Renormalization group and two-dimensional multicritical effective scalar field theory. Phys. Lett. B 345, 139 (1995). https://doi.org/10.1016/0370-2693(94)01603-A

    Article  ADS  Google Scholar 

  10. Nakayama, Y.: Scale invariance vs conformal invariance. Phys. Rep. 569, 1 (2015). https://doi.org/10.1016/j.physrep.2014.12.003

    Article  ADS  MathSciNet  Google Scholar 

  11. Polyakov, A.M.: Nonhamiltonian approach to conformal quantum field theory. Zh. Eksp. Teor. Fiz. 39, 23 (1974)

    Google Scholar 

  12. Ferrara, S., Grillo, A., Gatto, R.: Tensor representations of conformal algebra and conformally covariant operator product expansion. Ann. Phys. 76(1), 161 (1973)

    Article  ADS  MathSciNet  Google Scholar 

  13. Rattazzi, R., Rychkov, V.S., Tonni, E., Vichi, A.: Bounding scalar operator dimensions in 4D CFT. J. High Energy Phys. 2008(12), 031 (2008)

    Article  MathSciNet  Google Scholar 

  14. El-Showk, S., Paulos, M.F., Poland, D., Rychkov, S., Simmons-Duffin, D., Vichi, A.: Solving the 3D Ising model with the conformal bootstrap II. c-Minimization and precise critical exponents. J. Stat. Phys. 157, 869 (2014). https://doi.org/10.1007/s10955-014-1042-7

    Article  ADS  MathSciNet  MATH  Google Scholar 

  15. Simmons-Duffin, D.: A semidefinite program solver for the conformal bootstrap. J. High Energy Phys. 2015(6), 174 (2015)

    Article  MathSciNet  Google Scholar 

  16. Belavin, A.A., Polyakov, A.M., Zamolodchikov, A.B.: Infinite conformal symmetry in two-dimensional quantum field theory. Nucl. Phys. B 241, 333 (1984). https://doi.org/10.1016/0550-3213(84)90052-X

    Article  ADS  MathSciNet  MATH  Google Scholar 

  17. Belavin, A.A., Polyakov, A.M., Zamolodchikov, A.B.: Infinite conformal symmetry in two-dimensional quantum field theory. Nucl. Phys. B 241, 605 (1984)

    Article  MathSciNet  Google Scholar 

  18. El-Showk, S., Paulos, M.F., Poland, D., Rychkov, S., Simmons-Duffin, D., Vichi, A.: Solving the 3D Ising model with the conformal bootstrap. Phys. Rev. D 86, 025022 (2012). https://doi.org/10.1103/PhysRevD.86.025022

    Article  ADS  MATH  Google Scholar 

  19. Kos, F., Poland, D., Simmons-Duffin, D.: Bootstrapping mixed correlators in the 3D Ising model. JHEP 11, 109 (2014). https://doi.org/10.1007/JHEP11(2014)109

    Article  ADS  MATH  Google Scholar 

  20. Kos, F., Poland, D., Simmons-Duffin, D.: Bootstrapping the \(O(N)\) vector models. JHEP 06, 091 (2014). https://doi.org/10.1007/JHEP06(2014)091

    Article  ADS  MATH  Google Scholar 

  21. Kos, F., Poland, D., Simmons-Duffin, D., Vichi, A.: Bootstrapping the O(N) Archipelago. JHEP 11, 106 (2015). https://doi.org/10.1007/JHEP11(2015)106

    Article  ADS  MathSciNet  MATH  Google Scholar 

  22. Kos, F., Poland, D., Simmons-Duffin, D., Vichi, A.: Precision Islands in the Ising and \(O(N)\) models. JHEP 08, 036 (2016). https://doi.org/10.1007/JHEP08(2016)036

    Article  ADS  MathSciNet  MATH  Google Scholar 

  23. Polyakov, A.M.: Conformal symmetry of critical fluctuations. JETP Lett. 12, 381 (1970)

    ADS  Google Scholar 

  24. Polyakov, A.M.: Conformal symmetry of critical fluctuations. Pisma Zh. Eksp. Teor. Fiz. 12, 538 (1970)

    Google Scholar 

  25. Migdal, A.A.: On hadronic interactions at small distances. Phys. Lett. 37B, 98 (1971). https://doi.org/10.1016/0370-2693(71)90583-1

    Article  ADS  Google Scholar 

  26. Zamolodchikov, A.B.: Irreversibility of the flux of the renormalization group in a 2D field theory. JETP Lett. 43, 730 (1986)

    ADS  MathSciNet  Google Scholar 

  27. Zamolodchikov, A.B.: Irreversibility of the flux of the renormalization group in a 2D field theory. Pisma Zh. Eksp. Teor. Fiz. 43, 565 (1986)

    Google Scholar 

  28. Morinelli, V., Tanimoto, Y.: Commun. Math. Phys. (2019). https://doi.org/10.1007/s00220-019-03410-x

    Article  Google Scholar 

  29. Polchinski, J.: Scale and conformal invariance in quantum field theory. Nucl. Phys. B 303(2), 226 (1988)

    Article  ADS  MathSciNet  Google Scholar 

  30. Delamotte, B., Tissier, M., Wschebor, N.: Scale invariance implies conformal invariance for the three-dimensional Ising model. Phys. Rev. E 93(1), 012144 (2016)

    Article  ADS  MathSciNet  Google Scholar 

  31. Luty, M.A., Polchinski, J., Rattazzi, R.: The \(a\)-theorem and the asymptotics of 4D quantum field theory. JHEP 01, 152 (2013). https://doi.org/10.1007/JHEP01(2013)152

    Article  ADS  MathSciNet  MATH  Google Scholar 

  32. Dymarsky, A., Komargodski, Z., Schwimmer, A., Theisen, S.: On scale and conformal invariance in four dimensions. JHEP 10, 171 (2015). https://doi.org/10.1007/JHEP10(2015)171

    Article  ADS  MathSciNet  MATH  Google Scholar 

  33. Dymarsky, A., Farnsworth, K., Komargodski, Z., Luty, M.A., Prilepina, V.: Scale invariance, conformality, and generalized free fields. JHEP 02, 099 (2016). https://doi.org/10.1007/JHEP02(2016)099

    Article  ADS  MathSciNet  MATH  Google Scholar 

  34. Dymarsky, A., Zhiboedov, A.: Scale-invariant breaking of conformal symmetry. J. Phys. A 48(41), 41FT01 (2015). https://doi.org/10.1088/1751-8113/48/41/41FT01

    Article  MathSciNet  MATH  Google Scholar 

  35. Griffiths, R.B.: Correlations in Ising ferromagnets. I. J. Math. Phys. 8(3), 478 (1967)

    Article  ADS  Google Scholar 

  36. Kelly, D.G., Sherman, S.: General Griffiths’ inequalities on correlations in Ising ferromagnets. J. Math. Phys. 9(3), 466 (1968)

    Article  ADS  Google Scholar 

  37. Lebowitz, J.L.: GHS and other inequalities. Commun. Math. Phys. 35(2), 87 (1974)

    Article  ADS  MathSciNet  Google Scholar 

  38. Meneses, S., Penedones, J., Rychkov, S., Viana Parente Lopes, J.M., Yvernay, P.: A structural test for the conformal invariance of the critical 3D Ising model. JHEP 04, 115 (2019). https://doi.org/10.1007/JHEP04(2019)115

    Article  ADS  MathSciNet  MATH  Google Scholar 

  39. Delamotte, B., Tissier, M., Wschebor, N.: (2018). e-Print: arXiv:1802.07157 [hep-th]

  40. Gies, H., Wetterich, C.: Renormalization flow of bound states. Phys. Rev. D 65, 065001 (2002). https://doi.org/10.1103/PhysRevD.65.065001

    Article  ADS  MathSciNet  Google Scholar 

  41. Bricmont, J.: The Gaussian inequality for multicomponent rotators. J. Stat. Phys. 17(5), 289 (1977)

    Article  ADS  MathSciNet  Google Scholar 

  42. Dunlop, F.: Zeros of the partition function and Gaussian inequalities for the plane rotator model. J. Stat. Phys. 21(5), 561 (1979)

    Article  ADS  MathSciNet  Google Scholar 

  43. Dunlop, F.: Correlation inequalities for multicomponent rotators. Commun. Math. Phys. 49(3), 247 (1976)

    Article  ADS  MathSciNet  Google Scholar 

  44. Monroe, J.L., Pearce, P.A.: Correlation inequalities for vector spin models. J. Stat. Phys. 21(6), 615 (1979)

    Article  ADS  MathSciNet  Google Scholar 

  45. Kunz, H., Pfister, C.E., Vuillermot, P.A.: Correlation inequalities for some classical spin vector models. Phys. Lett. A 54(6), 428 (1975)

    Article  ADS  Google Scholar 

  46. Sokal, A.D.: Mean-field bounds and correlation inequalities. J. Stat. Phys. 28(3), 431 (1982)

    Article  ADS  MathSciNet  Google Scholar 

  47. de Gennes, P.G.: Exponents for the excluded volume problem as derived by the Wilson method. Phys. Lett. A 38, 339 (1972). https://doi.org/10.1016/0375-9601(72)90149-1

    Article  ADS  Google Scholar 

  48. Polchinski, J.: Renormalization and effective Lagrangians. Nucl. Phys. B231, 269 (1984). https://doi.org/10.1016/0550-3213(84)90287-6

    Article  ADS  Google Scholar 

  49. Canet, L., Delamotte, B., Wschebor, N.: Fully developed isotropic turbulence: nonperturbative renormalization group formalism and fixed point solution. Phys. Rev. E 93(6), 063101 (2016). https://doi.org/10.1103/PhysRevE.93.063101

    Article  ADS  Google Scholar 

  50. Frisch, U.: Turbulence: the legacy of AN Kolmogorov. Cambridge University Press, Cambridge (1995)

    Book  Google Scholar 

  51. Morris, T.R., Percacci, R.: Trace anomaly and infrared cutoffs. Phys. Rev. D 99(10), 105007 (2019). https://doi.org/10.1103/PhysRevD.99.105007

    Article  ADS  MathSciNet  Google Scholar 

  52. Rosten, O.J.: On functional representations of the conformal algebra. Eur. Phys. J. C 77(7), 477 (2017). https://doi.org/10.1140/epjc/s10052-017-5049-5

    Article  ADS  Google Scholar 

  53. Rosten, O.J.: A conformal fixed-point equation for the effective average action. Int. J. Mod. Phys. A 34(05), 1950027 (2019). https://doi.org/10.1142/S0217751X19500271

    Article  ADS  MathSciNet  MATH  Google Scholar 

  54. Fisher, M.E., Ma, S.K., Nickel, B.G.: Critical exponents for long-range interactions. Phys. Rev. Lett. 29, 917 (1972)

    Article  ADS  Google Scholar 

  55. Moshe, M., Zinn-Justin, J.: Quantum field theory in the large N limit: a review. Phys. Rep. 385, 69 (2003). https://doi.org/10.1016/S0370-1573(03)00263-1

    Article  ADS  MathSciNet  MATH  Google Scholar 

  56. Yabunaka, S., Delamotte, B.: Why might the standard large \(N\) analysis fail in the O(\(N\)) model: the role of cusps in the fixed point potentials. Phys. Rev. Lett. 121(23), 231601 (2018). https://doi.org/10.1103/PhysRevLett.121.231601

    Article  ADS  Google Scholar 

  57. Canet, L., Delamotte, B., Mouhanna, D., Vidal, J.: Nonperturbative renormalization group approach to the Ising model: a derivative expansion at order partial**4. Phys. Rev. B 68, 064421 (2003). https://doi.org/10.1103/PhysRevB.68.064421

    Article  ADS  Google Scholar 

  58. Canet, L., Delamotte, B., Mouhanna, D., Vidal, J.: Optimization of the derivative expansion in the nonperturbative renormalization group. Phys. Rev. D 67, 065004 (2003). https://doi.org/10.1103/PhysRevD.67.065004

    Article  ADS  Google Scholar 

  59. Balog, I., Chaté, H., Delamotte, B., Marohnić, M., Wschebor, N.: (2019) e-Print: arXiv:1907.01829 [cond-mat.stat-mech]

  60. Blaizot, J.P., Mendez Galain, R., Wschebor, N.: A new method to solve the non perturbative renormalization group equations. Phys. Lett. B 632, 571 (2006). https://doi.org/10.1016/j.physletb.2005.10.086

    Article  ADS  MathSciNet  MATH  Google Scholar 

  61. Benitez, F., Blaizot, J.P., Chate, H., Delamotte, B., Mendez-Galain, R., Wschebor, N.: Non-perturbative renormalization group preserving full-momentum dependence: implementation and quantitative evaluation. Phys. Rev. E 85, 026707 (2012). https://doi.org/10.1103/PhysRevE.85.026707

    Article  ADS  Google Scholar 

  62. Von Gersdorff, G., Wetterich, C.: Nonperturbative renormalization flow and essential scaling for the Kosterlitz–Thouless transition. Phys. Rev. B 64, 054513 (2001). https://doi.org/10.1103/PhysRevB.64.054513

    Article  ADS  Google Scholar 

  63. Codello, A., Defenu, N., D’Odorico, G.: Critical exponents of O(N) models in fractional dimensions. Phys. Rev. D 91(10), 105003 (2015). https://doi.org/10.1103/PhysRevD.91.105003

    Article  ADS  MathSciNet  Google Scholar 

  64. El-Showk, S., Paulos, M., Poland, D., Rychkov, S., Simmons-Duffin, D., Vichi, A.: Conformal field theories in fractional dimensions. Phys. Rev. Lett. 112, 141601 (2014). https://doi.org/10.1103/PhysRevLett.112.141601

    Article  ADS  MATH  Google Scholar 

  65. Nakayama, Y.: (2019) e-Print: arXiv:1906.07914 [hep-th]

  66. Sokal, A.D.: More inequalities for critical exponents. J. Stat. Phys. 25(1), 25 (1981)

    Article  ADS  MathSciNet  Google Scholar 

  67. Pohlmeyer, K.: The Jost–Schroer theorem for zero-mass fields. Commun. Math. Phys. 12, 204 (1969). https://doi.org/10.1007/BF01661574

    Article  ADS  MathSciNet  MATH  Google Scholar 

  68. Fernandez, R., Frohlich, J., Sokal, A.D.: Random Walks, Critical Phenomena, and Triviality in Quantum Field Theory. Springer, Berlin (1992)

    Book  Google Scholar 

  69. Griffiths, R.B.: Correlations in Ising ferromagnets. II. External magnetic fields. J. Math. Phys. 8(3), 484 (1967)

    Article  ADS  Google Scholar 

  70. Griffiths, R.B.: Correlations in Ising ferromagnets. III. Commun. Math. Phys. 6(2), 121 (1967)

    Article  ADS  Google Scholar 

  71. Krinsky, S., Emery, V.: Upper bound on correlation functions of Ising ferromagnet. Phys. Lett. A 50(3), 235 (1974)

    Article  ADS  Google Scholar 

  72. Griffiths, R.B.: Rigorous results for Ising ferromagnets of arbitrary spin. J. Math. Phys. 10(9), 1559 (1969)

    Article  ADS  MathSciNet  Google Scholar 

  73. Griffiths, R.B., Hurst, C.A., Sherman, S.: Concavity of magnetization of an Ising ferromagnet in a positive external field. J. Math. Phys. 11(3), 790 (1970)

    Article  ADS  MathSciNet  Google Scholar 

  74. Nakayama, Y.: Private communication

  75. Paulos, M.F., Rychkov, S., van Rees, B.C., Zan, B.: Conformal invariance in the long-range Ising model. Nucl. Phys. B 902, 246 (2016). https://doi.org/10.1016/j.nuclphysb.2015.10.018

    Article  ADS  MathSciNet  MATH  Google Scholar 

  76. Stevenson, P.M.: Optimized perturbation theory. Phys. Rev. D 23, 2916 (1981). https://doi.org/10.1103/PhysRevD.23.2916

    Article  ADS  Google Scholar 

Download references

Acknowledgements

The authors thank B. Delamotte, G. Tarjus, T. Morris and Y. Nakayama for fruitful discussions. The authors acknowledge financial support from the ECOS-Sud France-Uruguay Program U11E01. N. W. and D. P. thanks PEDECIBA (Programa de desarrollo de las Ciencias Básicas, Uruguay) and acknowledges funding through Grant from the Comisión Sectorial de Investigación Científica de la Universidad de la República, Project I+D 2016 (cod 412).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nicolás Wschebor.

Additional information

Communicated by Giulio Biroli.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Redundant Operators

This Appendix is devoted to a discussion of redundant vector operators, with an emphasis on conformal invariance. Redundant operators are of the form \((\delta S/\delta \phi (x)) {\mathcal {O}}(x)\). They are sometimes considered to be physically uninteresting. Indeed, they typically have short-range correlation functions.Footnote 15

1.1 Redundant Operators and Breaking of Conformal Invariance

The aim of this section is to explain why redundant operators can be ignored as possible breakings of conformal invariance. On general grounds, a redundant operator, with short-range correlations can be responsible for the breaking of Ward identities of a typical symmetry. The mere existence of such an operator would have strong physical consequences because correlation functions for other fields would not display the corresponding invariance. We illustrate this in a very simple situation. Consider a model with two scalar fields \(\phi _1\) and \(\phi _2\) whose dynamics is given by a general action S which needs not be O(2)-symmetric. If we perform, in the path integral of the partition function, a change of variable \(\phi _i\rightarrow \phi _i+\theta \epsilon _{ij}\phi _j\) (here \(\epsilon _{ij}\) is the bidimensional Levi–Civita tensor and \(\theta \) an infinitesimal angle) which corresponds to an infinitesimal rotation in internal space, we obtain:

$$\begin{aligned} \int d^dx\left( \epsilon _{ij}J_i\frac{\delta W}{\delta J_j}\right) =\int d^dx\left\langle \epsilon _{ij}\phi _i \frac{\delta S}{\delta \phi _j}\right\rangle . \end{aligned}$$
(55)

The brackets in the right-hand-side represent an average over the fields with the Boltzmann distribution in presence of sources \(J_i\) for the fields \(\phi _i\). Of course, if the action is O(2) symmetric, we recover the Ward identity for rotation in internal space. However, for a generic action S,  the right-hand-side does not vanish and the O(2) Ward identity is not satisfied. Now, what is of interest for us here is that the right-hand-side of the previous equation is the average of a redundant operator. The operator \(\epsilon _{ij}\phi _i \frac{\delta S}{\delta \phi _j}\) appearing in the right-hand-side of Eq. (55), which has only contact terms in its correlation functions, is physically important because it induces a breaking of O(2) invariance, at the level of Ward identities.

To make an analogy with the strategy followed in this article to study conformal invariance, suppose we want to prove that a model is invariant under O(2) by searching for putative operators that could appear in the right-hand-side of Eq. (55). Suppose that we can discard the existence of such operators which are not redundant but that we have no control on redundant ones. Then, the previous example shows that we have no way to conclude on the O(2) invariance of the theory. If, instead, we can discard both non-redundant and redundant operators, then we conclude that the theory is indeed invariant.Footnote 16

The situation is however different in the case of conformal invariance. Assume indeed that we find a redundant, integrated vector operator of dimension \(-1\). Such an operator would be of the formFootnote 17\(\int d^dx (\delta {\varGamma }_k)/(\delta \phi (x)){\mathcal {O}}_\mu (x)\) where the operator \({\mathcal {O}}_\mu \) depends on x only through the field argument (as explained in Sect. 3, an explicit x-dependence would be inconsistent with the translational invariance of the operator \({\varSigma }_\mu \)). This operator would yield a potential violation of the Ward identity:

$$\begin{aligned}&\int _{x} (K_{\mu }^{x}-2D_\star ^{\phi }x_{\mu }) \phi (x)\frac{\delta {\varGamma }_{k}}{\delta \phi (x)}- \frac{1}{2}\int _{x,y}\partial _{t}R_{k}(|x-y|)(x_{\mu }+y_{\mu })G_{k}(x;y) \nonumber \\&\quad =\int _x \frac{\delta {\varGamma }_k}{\delta \phi (x)}{\mathcal {O}}_\mu (x). \end{aligned}$$
(56)

The right-hand-side could actually be reabsorbed in a modification of the conformal transformation of the field \(\phi \) [74]. At odds with the case of internal symmetries, the modified conformal transformation \(\phi (x)\rightarrow \phi (x)+\epsilon _\mu [(K_{\mu }^{x}-2D_\star ^{\phi }x_{\mu })\phi (x)-{\mathcal {O}}_\mu (x)]\) is nontrivial because the x-independent term \({\mathcal {O}}_\mu \) cannot compensate the usual variations and the bracket is therefore non zero.

It can be shown that the modified conformal transformation, together with the usual translation, rotation, and scale transformation satisfy the conformal algebra.

To conclude, the existence of a putative redundant, integrated vector operator of dimension \(-1\) would not lead to a breaking of conformal invariance but, instead, to a modification of the transformation of the field.

1.2 Exact Scaling Dimensions of Some Redundant Operators

In this section, we show that some redundant operators have simple scaling dimensions. We work in the framework of NPRG described in Sect. 2. We can choose the Hamiltonian (or action) to be of the Ginzburg–Landau type:

$$\begin{aligned} S[\phi ]=\int _x \frac{1}{2}(\nabla \phi )^2+\frac{1}{2}r_{\varLambda }\phi ^2+\frac{u_{\varLambda }}{4!}\phi ^4, \end{aligned}$$
(57)

where \(\int _x=\int d^d x\). In order to determine the scaling dimension of an operator, we study the evolution of the corresponding coupling under the RG flow in the vicinity of the fixed point. To this end, we add to the action a part which couples to a vector operator:

$$\begin{aligned} S_{\text{ V }}[\phi ]=\int _x \frac{a^\mu _{\varLambda }}{3!}\phi ^3\partial _\mu {\varDelta }\phi . \end{aligned}$$
(58)

Up to integrations by parts, this operator is the same as the one considered in [38]. Moreover, it has been proved to be the most relevant integrated vector operator invariant under \({\mathbb {Z}}_2\) symmetry near \(d=4\) [30].

We perform an infinitesimal transformation of the integration variable: \(\phi \rightarrow \phi -a^\mu _{\varLambda }/u_{\varLambda }\partial _\mu {\varDelta }\phi \) in the path integral appearing in Eq. (3). It is readily found that the quadratic pieces in the action, including the regulating term \({\varDelta }S_k,\) are invariant under this transformation. The variation of the quartic part of the action is found to compensate exactly \(S_{\text{ V }}\). We thus find that

$$\begin{aligned} W_{k}\left[ J,a^\mu _{\varLambda }\right] =W_{k}\left[ J+\frac{a^\mu _{\varLambda }}{u_{\varLambda }}\partial _\mu {\varDelta }J,0\right] +{\mathcal {O}}\left( a^\mu _{\varLambda }a^\nu _{\varLambda }\right) . \end{aligned}$$
(59)

At the level of the effective average action, this relation implies

$$\begin{aligned} {\varGamma }_{k}\left[ \phi ,a^\mu _{\varLambda }\right] ={\varGamma }_k\left[ \phi +a^\mu _{\varLambda }\partial _\mu {\varDelta }\phi ,0\right] +{\mathcal {O}}\left( a^\mu _{\varLambda }a^\nu _{\varLambda }\right) . \end{aligned}$$
(60)

This last equation states that the evolution of the effective action with an infinitesimal \(a^\mu _{\varLambda }\) is related to the effective action at vanishing \(a^\mu _{\varLambda },\) up to a modification of the field. This can be used in the following way. Defining the running coupling constants \(u_k\) and \(a^\mu _k\) as the prefactors of, respectively, \(\int _x \frac{1}{4!}\phi ^4\) and \(\int _x \frac{1}{3!}\phi ^3\partial _\mu {\varDelta }\phi \) in \({\varGamma }_k,\) we obtain that \(a^\mu _k/u_k\) is constant along the flow. To obtain the scaling dimension of the vector operator, we introduce dimensionless, renormalized quantities (denoted with tilde) as

$$\begin{aligned} {\tilde{x}}&=k x, \end{aligned}$$
(61)
$$\begin{aligned} {\tilde{\phi }}({\tilde{x}})&=k^{-(d-2)/2}Z_k^{1/2}\phi (x), \end{aligned}$$
(62)

where \(Z_k\) scales as \( k^{-\eta }\) at the Wilson–Fisher fixed point with \(\eta \) the anomalous dimension. The renormalized coupling constants are thus:

$$\begin{aligned} {\tilde{u}}_k&=k^{d-4} Z_k^{-2} u_k, \end{aligned}$$
(63)
$$\begin{aligned} {\tilde{a}}_k^\mu&=k^{d-1} Z_k^{-2} a_k^\mu . \end{aligned}$$
(64)

At the critical point, \({\tilde{u}}\) flows to a fixed point value \(u_\star \). Consequently, when \(k\rightarrow 0,\)

$$\begin{aligned} {\tilde{a}}_k^\mu \sim a_{\varLambda }^\mu \frac{u_\star }{u_{\varLambda }}k^3 \end{aligned}$$
(65)

which shows that the scaling dimension of \(a^\mu \) is exactly 3.

The proof given above relies strongly on the particular microscopic action given in Eq. (57). This gives interesting non-universal information on the flow of the coupling \(a_k^\mu ,\) but confers a preeminent role to the peculiar form of the Hamiltonian. To overcome this issue, we now present an alternative proof of the same result, but which is based on the exact flow equation (6) for the effective average action, expressed in terms of dimensionless, renormalized, fields:

$$\begin{aligned} \begin{aligned} \partial _t {\varGamma }_k[{\tilde{\phi }}]=&\int _{{\tilde{x}}} \frac{\delta {\varGamma }_k}{\delta {\tilde{\phi }}({\tilde{x}})}\left( {\tilde{x}}^\rho \partial _{{\tilde{x}}^\rho }-\alpha \right) {\tilde{\phi }}({\tilde{x}})+\frac{1}{2}\int _{{\tilde{x}}{\tilde{y}}}\partial _t{\tilde{R}}({\tilde{x}}-{\tilde{y}}){\tilde{P}}_k({\tilde{x}},{\tilde{y}}), \end{aligned} \end{aligned}$$
(66)

where \( R_k( x)=Z_k k^{d+2}{\tilde{R}}(k x),\)\(\alpha =-(d-2+\eta )/2,\)\(t=\log (k/{\varLambda })\) and \({\tilde{P}}_k\) is the dimensionless, renormalized, propagator:

$$\begin{aligned} \int _{{\tilde{y}}} {\tilde{P}}_k({\tilde{x}},{\tilde{y}})\left[ \frac{\delta ^2{\varGamma }_k}{\delta {\tilde{\phi }}({\tilde{y}})\delta {\tilde{\phi }}({\tilde{z}})}+{\tilde{R}}({\tilde{y}}-{\tilde{z}})\right] =\delta ({\tilde{x}}-{\tilde{z}}). \end{aligned}$$
(67)

We now identify an exact eigenvector of the linearized flow. To this end, we add to the Wilson–Fisher fixed-point effective action \({\varGamma }_\star \) a small perturbation

$$\begin{aligned} {\varGamma }_k={\varGamma }_\star +{\tilde{r}}_\mu (t)\int _{ {\tilde{x}}} \frac{\delta {\varGamma }_\star }{\delta {\tilde{\phi }}({\tilde{x}})} \tilde{\partial }_\mu {\tilde{{\varDelta }}} {\tilde{\phi }}({\tilde{x}}) \end{aligned}$$
(68)

and we compute the flow of this functional at linear order in \({\tilde{r}}_\mu \).

$$\begin{aligned} \begin{aligned}&\partial _t{{\tilde{r}}}_\mu (t) \int _{{\tilde{x}}} \frac{\delta {\varGamma }_\star }{\delta {\tilde{\phi }}({\tilde{x}})} \tilde{\partial }_\mu \tilde{{\varDelta }} {\tilde{\phi }}({\tilde{x}}) \\&\quad = {\tilde{r}}_\mu (t)\int _{{\tilde{x}}{\tilde{y}}} \frac{\delta \ }{\delta {\tilde{\phi }}({\tilde{y}})}\left[ \frac{\delta {\varGamma }_\star }{\delta {\tilde{\phi }}({\tilde{x}})} {\tilde{\partial }}_\mu {\tilde{{\varDelta }}} {\tilde{\phi }}({\tilde{x}})\right] \left( {\tilde{y}}^\rho \partial _{{\tilde{y}}^\rho }-\alpha \right) {\tilde{\phi }}({\tilde{y}})\\&\qquad -\frac{1}{2} {\tilde{r}}_\mu (t)\int _{{\tilde{x}}{\tilde{y}} {\tilde{z}}{\tilde{t}}{\tilde{w}}}\partial _t{\tilde{R}} ({\tilde{x}}-{\tilde{y}}){\tilde{P}}_\star ({\tilde{y}}, {\tilde{z}}){\varGamma }_\star ^{(3)}({\tilde{z}},{\tilde{t}}, {\tilde{w}}){\tilde{P}}_\star ({\tilde{w}},{\tilde{x}}) {\tilde{\partial }}_\mu {\tilde{{\varDelta }}}{\tilde{\phi }}({\tilde{t}}). \end{aligned} \end{aligned}$$
(69)

On the other hand, if we derive the fixed point equation with respect to \({\tilde{\phi }}({\tilde{x}}),\) multiply by \( {\tilde{\partial }}_\mu {\tilde{{\varDelta }}} {\tilde{\phi }}({\tilde{x}})\) and integrate over \({\tilde{x}},\) we get

$$\begin{aligned} \begin{aligned} 0&=\int _{{\tilde{x}}{\tilde{y}}}{\tilde{\partial }}_\mu {\tilde{{\varDelta }}} {\tilde{\phi }}({\tilde{x}})\frac{\delta \ }{\delta {\tilde{\phi }}({\tilde{x}})}\left[ \frac{\delta {\varGamma }_\star }{\delta {\tilde{\phi }}({\tilde{y}})} \left( {\tilde{y}}^\rho \partial _{{\tilde{y}}^\rho }-\alpha \right) {\tilde{\phi }} ({\tilde{y}})\right] \\&\quad -\frac{1}{2} \int _{{\tilde{x}}{\tilde{y}}{\tilde{z}}{\tilde{t}} {\tilde{w}}}\partial _t{\tilde{R}}({\tilde{x}}-{\tilde{y}}) {\tilde{P}}_\star ({\tilde{y}},{\tilde{z}}){\varGamma }_\star ^{(3)} ({\tilde{z}},{\tilde{t}},{\tilde{w}}){\tilde{P}}_\star ({\tilde{w}}, {\tilde{x}}){\tilde{\partial }}_\mu {\tilde{{\varDelta }}}{\tilde{\phi }}({\tilde{t}}). \end{aligned} \end{aligned}$$
(70)

Combining the two equations, we obtain:

$$\begin{aligned} \begin{aligned} \partial _t {{\tilde{r}}}_\mu (t) \int _{{\tilde{x}}} \frac{\delta {\varGamma }_\star }{\delta {\tilde{\phi }}({\tilde{x}})} {\tilde{\partial }}_\mu {\tilde{{\varDelta }}}^n {\tilde{\phi }}({\tilde{x}})= {\tilde{r}}_\mu (t)\int _{{\tilde{x}}{\tilde{y}}} \frac{\delta {\varGamma }_\star }{\delta {\tilde{\phi }}({\tilde{x}})} \left[ {\tilde{\partial }}_\mu {\tilde{{\varDelta }}} ,{\tilde{x}}^\rho \partial _{{\tilde{x}}^\rho }\right] {\tilde{\phi }}({\tilde{x}}). \end{aligned} \end{aligned}$$
(71)

The commutator is easily evaluated to be equal to \(3{\tilde{\partial }}_\mu {\tilde{{\varDelta }}} \). From this we deduce that the small perturbation introduced in Eq. (68) is an exact eigenoperator of the flow around the fixed point, with eigenvalue 3. This is consistent with the result found in the one-loop calculation of [30].

We can generalize the previous result in different ways. First, we can change the power of the Laplacian in Eq. (68) from unity to a positive integer n. The main change appears at the level of Eq. (71), where the commutator is now \([{\tilde{\partial }}_\mu {\tilde{{\varDelta }}}^n ,{\tilde{x}}^\rho \partial _{{\tilde{x}}^\rho }]=(2n+1){\tilde{\partial }}_\mu {\tilde{{\varDelta }}}^n\). This implies that the associated eigenvector has dimension \(2n+1\). As a check of this result, we have considered the vector eigenoperators compatible with the \({\mathbb {Z}}_2\) symmetry whose scaling dimensions are 5 in \(d=4\) and we have computed their first correction in \(\epsilon =4-d\). There are 4 (independent) such operators: 1 (\(O_6^3\)) with 6 powers of the field and 3 derivatives and 3 (\(O_{4,i}^5\) with \(i\in \{1,2,3\}\)) with 4 powers of the field and 5 derivatives. A one-loop calculation shows that \(O_6^3\) has scaling dimension \(5-5\epsilon /3+{\mathcal {O}}(\epsilon ^2)\). The eigenvectors \(O_{4,i}^5\) have dimensions \(5+{\mathcal {O}}(\epsilon ^2),\)\(5-4\epsilon /9+{\mathcal {O}}(\epsilon ^2)\) and \(5-2\epsilon /3+{\mathcal {O}}(\epsilon ^2)\). The eigenoperator with scaling dimension \(5+{\mathcal {O}}(\epsilon ^2)\) is found to be \(\int _{{\tilde{x}}} {\tilde{\phi }}^3 {\tilde{\partial }}_\mu {\tilde{{\varDelta }}}^2{\tilde{\phi }},\) in agreement with the general result mentioned above. Other relations can be obtained if we consider in Eq. (68) an odd number of derivatives, with Lorentz indices not necessarily contracted together.

The present result also generalizes to the long-range Ising model, where the interaction between spins is not limited to nearest neighbors but decay as a power-law:

$$\begin{aligned} H=-\sum _{i,j}J(i-j)S_i S_j, \end{aligned}$$
(72)

where \(J(i-j)\sim |i-j|^{-d-\sigma }\) and \(\sigma \) is the exponent characterizing the decrease of the interactions. When \(0<\sigma <2-\eta ,\) the model still has an extensive free-energy but belongs to a different universality class than the local Ising model. The Ginzburg–Landau Hamiltonian is identical to the one given in Eq. (57) except that the quadratic part is now, in Fourier space,

$$\begin{aligned} \int \frac{d^dq}{(2\pi )^d}\phi (-q) q^{\sigma }\phi (q). \end{aligned}$$
(73)

It is easy to verify that all the present analysis still applies to this case. We have checked that the one-loop calculation around the upper critical dimension \(d_c=2\sigma \) gives that the most relevant integrated vector operator has scaling dimension \(3+{\mathcal {O}}(\epsilon ^2)\). This result is important because it justifies the use of the conformal bootstrap program in this model [75], at least near the upper critical dimension.

We can also generalize the result to other internal groups. For O(N) theories, an exact eigenoperator can be found by adding a common O(N) index on both the functional derivative and the field appearing in Eq. (68) and summing over this index. The associated eigenvalue is again 3 (or \(2n+1,\) if we change the power of the Laplacian). In [30] [see also Eq. (34)], we computed the scaling dimensions of the two vector operators of lowest dimension in an expansion in \(\epsilon \) and found \(3+{\mathcal {O}}(\epsilon ^2)\) and \(3-6\epsilon /(N+8)+{\mathcal {O}}(\epsilon ^2) \). This result is consistent with the nonrenormalization theorem proven here. Let us stress, however, that in the O(N) model the non renormalization theorem does not constraint the leading vector operator but the next-to-leading, as can be seen already at one-loop level [30].

For completeness, we give other exact eigenvectors which can be obtained following the same idea. These are, however, not invariant under the internal symmetries of the theory, involve even derivatives and are not relevant to our discussion on conformal invariance. A well-known example is the eigenvector associated with the external magnetic field: \(\int _{ {\tilde{x}}} \frac{\delta {\varGamma }_\star }{\delta {\tilde{\phi }}({\tilde{x}})}\) with eigenvalue \(\alpha =-(d-2+\eta )/2\). Another exact eigenvector, valid for the O(N) model is \(\int _{ {\tilde{x}}} \epsilon ^{ab}\frac{\delta {\varGamma }_\star }{\delta {\tilde{\phi }}^b({\tilde{x}})} {\tilde{{\varDelta }}}^n {\tilde{\phi }}^a({\tilde{x}})\) where \(\epsilon ^{ab}\) is antisymmetric, which is associated with the eigenvalue 2n.

It was shown in [38] that there exists other redundant operators in \(d=4-\epsilon \). In fact, all local operators of dimension up to 9 are redundant modulo total derivatives. Equivalently, all integrated vector operators of dimension up to 5 are redundant. For several of these, we were however not able to determine analytically their scaling dimensions.

\(\epsilon \)-Expansion

In this Appendix, we describe the 1-loop calculation of the two lowest scaling dimensions of vector operators. We present here the calculation in the framework of the NPRG, but of course, the calculation could be performed within more standard approaches, such as the Minimal Substraction scheme.

We start by observing that at tree-level (zero loop) the effective action takes its bare form:

$$\begin{aligned} {\varGamma }^{(tree)}_{k}=S=\int _x\Big \{r \rho +\frac{1}{3!}u\rho ^2+ \frac{1}{4}a_{\mu }\partial _{\mu }\rho \partial _{\nu }\phi _{i}\partial _{\nu }\phi _{i} +\frac{1}{2}b_{\mu }\partial _{\mu }\phi _{i}\partial _{\nu }\phi _{i}\partial _{\nu }\rho \Big \}. \end{aligned}$$
(74)

Differentiating successively with respect to \(\phi _{n_{i}}\left( x_{i} \right) \) and Fourier transforming, we obtain the form of the non-zero vertices:

$$\begin{aligned} {{\varGamma }_{i_{1}i_{2}}^{(2,tree)}}(p_1)&= \delta _{i_1i_2}(r+p_1^2) \end{aligned}$$
(75)
$$\begin{aligned} {{\varGamma }_{i_{1}i_{2}i_{3}i_{4}}^{(4,tree)}}(p_1,p_2,p_3)&= \frac{u}{3}\Big [\delta _{i_{1}i_{2}}\delta _{i_{3}i_{4}} +\delta _{i_{1}i_{3}}\delta _{i_{2}i_{4}}+\delta _{i_{1}i_{4}} \delta _{i_{2}i_{3}}\Big ] +i \frac{a_{\mu }-b_\mu }{2} \nonumber \\&\quad \times \Big [\left( p_{1}+p_{2}\right) ^{\mu }\big (p_{1}\cdot p_{2}- p_{3}\cdot p_{4}\big )\delta _{i_{1}i_{2}}\delta _{i_{3}i_{4}} \nonumber \\&\quad + \left( p_{1}+p_{3}\right) ^{\mu }\big (p_{1}\cdot p_{3}- p_{2} \cdot p_{4}\big )\delta _{i_{1}i_{3}}\delta _{i_{2}i_{4}}\nonumber \\&\quad + \left( p_{1}+p_{4}\right) ^{\mu }\big (p_{1}\cdot p_{4}- p_{2}\cdot p_{3}\big )\delta _{i_{1}i_{4}}\delta _{i_{2}i_{3}}\Big ]\nonumber \\&\quad -i\frac{b_{\mu }}{2} \left( p_{1}^{\mu }p_{1}^2+p_{2}^{\mu }p_{2}^2 +p_{3}^{\mu }p_{3}^2+p_{4}^{\mu }p_{4}^2\right) \nonumber \\&\quad \times \left( \delta _{i_{1} i_{2}}\delta _{i_{3}i_{4}} +\delta _{i_{1}i_{3}} \delta _{i_{2}i_{4}}+ \delta _{i_{1}n_{i}}\delta _{i_{2}i_{3}}\right) . \end{aligned}$$
(76)

In the previous equations, we have omitted for notation simplicity the index k on the coupling constants. The momentum \(p_4\) in \({\varGamma }^{(4)}\) is fixed by momentum conservation: \(\sum _{i=1}^{n} p_i=0\). The flow of \(a_{\mu }\) and \(b_{\mu }\) are deduced from the flow equation of \({\varGamma }^{(4)}\) at zero external field, which is obtained by differentiating four times the RG equation (6) and evaluating it at \(\phi =0\). We obtain:

$$\begin{aligned} \partial _{t}{\varGamma }^{(4)}_{i_{1}i_{2}i_{3}i_{4}}(p_1,p_2,p_3)&=\sum _{il}\int _{q}\partial _t R_k\left( q^2\right) G_k^2\left( q^2\right) \nonumber \\&\quad \bigg (-\frac{1}{2} {\varGamma }^{(6)}_{k,k,i_{1},i_{2},i_{3},i_{4}}(q,-q,p_1,p_2,p_3) \nonumber \\&\qquad + G_k\left( (q+p_{1}+p_{2})^2\right) {\varGamma }^{(4)}_{ki_{1}i_{2}l} (q,p_1,p_2) \nonumber \\&\quad {\varGamma }^{(4)}_{li_{3}i_{4}k}(q+p_1+p_2,p_3,p_4)+\text {2 perms.}\bigg ). \end{aligned}$$
(77)

At one loop, we can neglect the first term and replace the vertices in the right hand side of the flow equation by its tree-level form (76). Summing over i and l in the product of the \({\varGamma }_k^{(4,tree)},\) and keeping only terms which are at most linear in \(a_\mu \) and \(b_\mu ,\) we find that the result is independent of q:

$$\begin{aligned} \begin{aligned}&{{\varGamma }_{ki_{1}i_{2}l}^{(4,tree)}}(q,p_1,p_2) {{\varGamma }_{li_{3}i_{4}k}^{(4,tree)}}(-q,p_3,p_4)+2 \text {perms.} \\&\quad = \frac{u^{2}}{9}\left( N+8\right) \big (\delta _{i_{1}i_{2}} \delta _{i_{3}i_{4}}+\delta _{i_{1}i_{3}}\delta _{i_{2}i_{4}}+ \delta _{i_{1}i_{4}}\delta _{i_{2}i_{3}}\big )\\&\qquad +\frac{iua_{\mu }}{6}\Bigg \{\delta _{i_{1}i_{2}}\delta _{i_{3}i_{4}}\left[ (p_{1}+p_{2})^{\mu }(p_{1}\cdot p_{2}-p_{3}\cdot p_{4})\left( N+2\right) -2\sum _{k=1}^{4}(p_{k}^{\mu }p_{k}^{2})\right] \\&\qquad +2 \text { perms.}\Bigg \}\\&\qquad -\frac{iub_{\mu }}{6}\Bigg \{\delta _{i_{1}i_{2}}\delta _{i_{3}i_{4}} \left[ \left( p_{1}+p_{2}\right) ^{\mu }(p_{1}\cdot p_{2}-p_{3}\cdot p_{4})\left( N+2\right) +\sum _{k=1}^{4}(p_{k}^{\mu }p_{k}^{2})\left( N+6\right) \right] \\&\qquad +2 \text { perms.}\Bigg \}, \end{aligned} \end{aligned}$$
(78)

where cyclic permutations of the indices 2, 3 and 4 are not written explicitly.

The next step consists in identifying the prefactors of a given structure which involve both vector indices and momenta [see Eq. (76)] in the left hand side and right hand side of the flow equation (77). This implies that we must expand the r.h.s. in powers of the external momenta and extract terms of order zero and order three in momenta. However, Eq. (78) shows that the product of vertices already has a contribution with 0 and 3 powers of the external momenta. As a consequence, we can put, in the propagator, the external momenta to zero.

We can now extract the flows of \(a_\mu \) and \(b_\mu ,\) which read:

$$\begin{aligned} \partial _{t}u&=\frac{\left( N+8\right) }{3}u^{2}\int _{q}{{\dot{R}}\left( q\right) G^{3}\left( q\right) },\nonumber \\ \partial _{t}a_{\mu }&=\frac{u}{3}\left[ \left( N+4\right) a_{\mu }+4b_{\mu } \right] \int _{q}{{\dot{R}}\left( q\right) G^{3}\left( q\right) },\nonumber \\ \partial _{t}b_{\mu }&=\frac{u}{3}\left[ 2a_{\mu }+\left( N+6\right) b_{\mu } \right] \int _{q}{{\dot{R}}\left( q\right) G^{3}\left( q\right) }. \end{aligned}$$
(79)

The flow equations for the dimensionless variables are easily derived and correspond to those given in Eq. (29).

Large N Expansion

We discuss in this Appendix the large-N calculation of the two smallest scaling dimensions of the vector operators (which tend to 3 when \(d\rightarrow 4\)) as well as another one, (which tend to 5 when \(d\rightarrow 4\)). These can be deduced from the calculation of \({\varGamma }^{\left( 2\right) },\)\({\varGamma }^{\left( 4\right) }\) and \({\varGamma }^{\left( 6\right) }\) at vanishing external field in the large N limit. We recall that the large N limit is performed at fixed \({\hat{u}}=u N,\)\(\hat{a}_\mu =a_\mu N,\)\(\hat{b}_\mu =b_\mu N\) and \(\hat{c}_\mu =c_\mu N^2\). Moreover, we only need the flow of \(\hat{a}_\mu ,\)\(\hat{b}_\mu \) and \(\hat{c}_\mu \) at linear order in \(\hat{a}_\mu ,\)\(\hat{b}_\mu \) and \(\hat{c}_\mu \).

$$\begin{aligned}&\int _x \left\{ \frac{{\hat{u}}^{\varLambda }}{4!N}\phi _{i}^{2}\phi _{j}^{2} +\frac{\hat{a}^{\varLambda }_{\mu }}{4N}\phi _{i}\partial _{\mu }\phi _{i} \partial _{\nu }\phi _{j}\partial _{\nu }\phi _{j} +\frac{\hat{b}^{\varLambda }_{\mu }}{2N}\phi _{i}\partial _{\nu }\phi _{i} \partial _{\nu }\phi _{j}\partial _{\mu }\phi _{j} \right. \nonumber \\&\quad \left. +\frac{\hat{c}^{\varLambda }_{\mu }}{4N^2}\phi _{i}\partial _{\nu } \phi _{i}\phi _{j}\partial _{\nu }\phi _{j}\phi _{l}\partial _{\mu } \phi _{l}\right\} . \end{aligned}$$
(80)

The bare propagators and 4-point vertex are easily deduced from those given in Eq. (76). The 6-point vertex is given by:

$$\begin{aligned}&{{\varGamma }_{i_{1}i_{2}i_{3}i_{4}i_{5}i_{6}}^{(6,tree)}} (p_1,p_2,p_3,p_4,p_5) \nonumber \\&\quad = -i\frac{c_{\mu }^{{\varLambda }}}{2}\bigg \lbrace \delta _{i_{1}i_{2}}\delta _{i_{3}i_{4}}\delta _{i_{5}i_{6}} \Big [\left( p_{1}+p_{2}\right) ^{\mu }(p_{1}+ p_{2})^2+\left( p_{3}+p_{4}\right) ^{\mu }(p_{3}+ p_{4})^2 \nonumber \\&\qquad +\left( p_{5}+p_{6}\right) ^{\mu }(p_{5}+ p_{6})^2\Big ]+perm\bigg \rbrace , \end{aligned}$$
(81)

where perm represents the 14 permutations which lead to different combinations of Kronecker delta.

1.1 A Source of Simplification

A major simplification occurs in the calculation, which is a consequence of the following property: the contribution linear in \(a_\mu \) or \(b_\mu \) in the 4-point vertex which is proportional to \(\delta _{i_{1}i_{2}}\delta _{i_{3}i_{4}}\) vanishes in the exceptional configurations where the momenta are opposite by pairs in different delta’s (that is, if \(p_1+p_3=p_2+p_4=0\) or if \(p_1+p_4=p_2+p_3=0\)). As a consequence, in a diagram made of a chain of bubbles (see Fig. 2), if a \({\varGamma }^{(4)}\) connecting two bubbles is replaced by a perturbation \(a_\mu \) or \(b_\mu ,\) the diagram vanishes. Otherwise stated, in a chain of bubbles diagram, the perturbations \(a_\mu \) and \(b_\mu \) only occur when attached to an external leg (see Fig. 6).

A somewhat similar situation occurs when a \(c_\mu \) perturbation appears in a chain of bubble for the 4-point function. When this perturbation is inserted in a chain (whether in contact or not with external legs), given the structure of Eq. (81), the legs from the isolated loop cancels by itself and the remaining two pairs cancel with each other. This closely resembles the property mentioned previously that \(a_\mu \) and \(b_\mu \) don’t appear in an inner vertex of the chains. Moreover, the specific momentum structure appearing in the \(c_\mu \) vertex, implies that it can’t appear neither attached to an external leg. We conclude that \(c_\mu \) does not contribute to the flow of \(a_\mu \) and \(b_\mu \) at leading order in 1 / N.

The situation is even simpler in the calculation of \({\varGamma }^{(2)}\) because, by conservation of the momenta, the external legs have opposite momenta. In this situation, the perturbation 4-point vertex does not even contribute when attached to the external legs. As a consequence, the cactus diagrams for \({\varGamma }^{(2)}\) are independent of \(a_\mu \) and \(b_\mu \). This result is important because it implies that the inverse full propagator [which re-sums all cactus diagrams for \({\varGamma }^{(2)}\)] is independent of \(a_\mu ,\)\(b_\mu \) and \(c_\mu \).

1.2 Computation of \({\varGamma }^{(2)}\)

We first discuss the (standard) calculation of \({\varGamma }^{(2)}\) at leading order. As discussed above, we can remove \(a_\mu \) and \(b_\mu \) from this calculation. The sum of the cactus diagrams shown in Fig. 1 leads to

$$\begin{aligned} {\varGamma }_{ij}^{(2)}(p)=\delta _{ij}\Big \{p^2+r^{\varLambda }+\frac{\hat{u}^{\varLambda }}{6}\int _q\frac{1}{q^2+r^{\varLambda }+{\overline{{\varSigma }}}\left( r^{\varLambda }\right) }\Big \}=\delta _{ij}\Big \{p^2+r\Big \}, \end{aligned}$$
(82)

where \(r=r^{{\varLambda }}+{\overline{{\varSigma }}}\left( r^{\varLambda }\right) \) and \({\overline{{\varSigma }}}\left( r^{\varLambda }\right) \) satisfies the gap equation:

$$\begin{aligned} {\overline{{\varSigma }}}\left( r^{\varLambda }\right) =\frac{\hat{u}^{\varLambda }}{6}\int _q\frac{1}{q^2+r^{\varLambda }+{\overline{{\varSigma }}}\left( r^{\varLambda }\right) }. \end{aligned}$$
(83)

As is well known, the only effect of the cactus diagrams is to modify the mass.

1.3 Computation of \({\varGamma }^{\left( 4\right) }\)

In contrast to \({{\varGamma }}^{\left( 2\right) },\) the \({{\varGamma }}^{\left( 4\right) }\) vertex has corrections linear in \(a^{\varLambda }_{\mu }\) and \(b^{\varLambda }_{\mu }\) to leading order in the \(\frac{1}{N}\) expansion (i.e. to order \(\frac{1}{N}\)). It is convenient to decompose the 4-point vertex function as \({\varGamma }^{\left( 4\right) }={\varGamma }^{\left( 4\right) }_u+{\varGamma }^{\left( 4\right) }_{a_{\mu }}+{\varGamma }^{\left( 4\right) }_{b_{\mu }},\) where the first term is independent of \(a_\mu \) and \(b_\mu ,\) the second term is linear in \(a_\mu \) and the third is linear in \(b_\mu \). We omit all other terms which do not enter into the calculation of the scaling dimensions we are interested in.

The term \({\varGamma }^{\left( 4\right) }_u\) is the simplest one since it corresponds to the usual theory with \(a^{\varLambda }_{\mu }=b^{\varLambda }_{\mu }=0\). This gives the standard large-N result:

$$\begin{aligned}&{\varGamma }^{\left( 4\right) }_{u,i_1,i_2,i_3,i_4}(p_1,p_2,p_3) \nonumber \\&\quad =\frac{\hat{u}^{\varLambda }}{3N}\left[ \frac{\delta _{i_{1}i_{2}} \delta _{i_{3}i_{4}}}{1+\frac{\hat{u}^{\varLambda }}{6}{\varPi }\left( p_{1}+p_{2}\right) } \right. +\frac{\delta _{i_{1}i_{3}}\delta _{i_{2}i_{4}}}{1+\frac{\hat{u}^{\varLambda }}{6} {\varPi }\left( p_{1}+p_{3}\right) } +\left. \frac{\delta _{i_{1}i_{4}}\delta _{i_{2}i_{3}}}{1+\frac{\hat{u}^{\varLambda }}{6} {\varPi }\left( p_{1}+p_{4}\right) }\right] , \end{aligned}$$
(84)

where the function \({\varPi }\left( p\right) \) is defined as:

$$\begin{aligned} {\varPi }\left( p\right) =\int _q \frac{1}{q^2+r}\frac{1}{\left( q+p\right) ^2+r}. \end{aligned}$$
(85)

We now consider \({\varGamma }^{\left( 4\right) }_{a_{\mu }}\). For simplicity, we focus on the contribution proportional to \(\delta _{i_1i_2}\delta _{i_3i_4}\). The other contribution are obtained by permutations of the external legs. The set of diagrams which contribute is easy to characterize because the perturbation (\(a_\mu \) in this case) must be attached to the external legs (see Sect. C.1). The chain of bubbles diagrams for \({\varGamma }^{\left( 4\right) }_{a_{\mu }}\) are depicted in Fig. 6.

Fig. 6
figure 6

Diagrams contributing to \({\varGamma }^{\left( 4\right) }_{a_\mu }\)

The diagram with n couplings \(\hat{u}^{\varLambda }\) and one \(\hat{a}^{\varLambda }_{\mu }\) connected to \(p_{1}\) and \(p_{2}\) is equal to:

$$\begin{aligned}&\frac{i\hat{a}^{\varLambda }_{\mu }}{2N}\int _q\frac{\left( p_{1} +p_{2}\right) ^{\mu }\left[ p_{1}\cdot p_{2}+q\cdot \left( q+p_{1}+p_{2}\right) \right] }{\left[ q^{2}+r\right] \left[ \left( q+p_{1}+p_{2}\right) ^{2} +r\right] } \nonumber \\&\quad \times \left( \frac{-\hat{u}^{\varLambda }}{4!}\right) ^{n} \frac{4^{n}n!C^{n+1}_{n}}{\left( n+1\right) !} \left( {\varPi }\left( p_{1}+p_{2}\right) \right) ^{\left( n-1\right) }. \end{aligned}$$
(86)

If \(\hat{a}^{\varLambda }_{\mu }\) is connected to \(p_{3}\) and \(p_{4},\) we get:

$$\begin{aligned}&\frac{i\hat{a}^{\varLambda }_{\mu }}{2N}\int _q\frac{\left( p_{1}+p_{2} \right) ^{\mu }\left[ -q\cdot \left( q+p_{1}+p_{2}\right) -p_{3}\cdot p_{4}\right] }{\left[ q^{2}+r\right] \left[ \left( q+p_{1}+p_{2}\right) ^{2}+ r\right] } \nonumber \\&\quad \times \left( \frac{-\hat{u}^{\varLambda }}{4!}\right) ^{n} \frac{4^{n}n!C^{n+1}_{n}}{\left( n+1\right) !} \left( {\varPi }\left( p_{1}+p_{2}\right) \right) ^{\left( n-1\right) }. \end{aligned}$$
(87)

When adding both diagrams we get the result for n couplings \(\hat{u}^{\varLambda }\) and one \(\hat{a}^{\varLambda }_{\mu }\):

$$\begin{aligned} \frac{i\hat{a}^{\varLambda }_{\mu }}{2N}\left( -\frac{\hat{u}^{\varLambda }}{6}{\varPi }\left( p_{1}+p_{2}\right) \right) ^{n}\left( p_{1}+p_{2}\right) ^{\mu }\left[ p_{1}\cdot p_{2}-p_{3}\cdot p_{4}\right] . \end{aligned}$$
(88)

Note that the previous construction does not make sense for \(n=0\). However, it happens that Eq. (88) evaluated at \(n=0\) indeed represents the contribution of the first diagram of Fig. 6 with one \(a_\mu \) and no \({\hat{u}}\). It is straightforward to sum this general expression for all n to get:

$$\begin{aligned} {\varGamma }^{\left( 4\right) }_{a_{\mu }}= & {} \frac{i\hat{a}^{\varLambda }_{\mu }}{2N\left( 1+\frac{\hat{u}^{\varLambda }}{6}{\varPi }\left( p_{1}+p_{2}\right) \right) } \left( p_{1}+p_{2}\right) ^{\mu } \nonumber \\&\quad \times \left[ p_{1}\cdot p_{2}-p_{3}\cdot p_{4}\right] \delta _{i_1i_2}\delta _{i_3i_4}+\text {2 perms.} \end{aligned}$$
(89)

The calculation for \({\varGamma }_{b_\mu }^{(4)}\) proceeds in the same way. The contribution of diagrams with one \(b_\mu \) and n couplings \({\hat{u}}\) (again focusing on the contribution proportional to \(\delta _{i_1i_2}\delta _{i_3i_4}\)) is:

$$\begin{aligned}&\frac{-i\hat{b}^{\varLambda }_{\mu }}{2N} \left( -\frac{\hat{u}^{\varLambda }}{6} {\varPi }\left( p_{1}+p_{2}\right) \right) ^{n} \nonumber \\&\quad \times \left[ p_{1}^{\mu }p_{1}^{2}+p_{2}^{\mu }p_{2}^{2} +p_{3}^{\mu }p_{3}^{2}+p_{4}^{\mu }p_{4}^{2}+\left( p_{1} +p_{2}\right) ^{\mu }(p_{1}\cdot p_{2}-p_{3}\cdot p_{4})\right] . \end{aligned}$$
(90)

To sum up, the four-point vertex with at most one \(a_\mu \) or one \(b_\mu \) is

$$\begin{aligned} \begin{aligned} {\varGamma }^{\left( 4\right) }_{i_1i_2i_3i_4}(p_1,p_2,p_3)&=\frac{\delta _{i_{1}i_{2}}\delta _{i_{3}i_{4}}}{N\big (1+ \frac{\hat{u}^{\varLambda }}{6}{\varPi }\left( p_1+p_2\right) \big )} \\&\quad \times \bigg \lbrace \frac{\hat{u}^{\varLambda }}{3}+i \frac{\hat{a}^{\varLambda }_{\mu }-\hat{b}^{\varLambda }_{\mu }}{2} \left( p_{1}+p_{2}\right) ^{\mu }\left[ p_{1} \cdot p_{2} - p_{3} \cdot p_{4}\right] \\&\qquad -i \frac{\hat{b}^{\varLambda }_{\mu }}{2} \left( p_{1}^{\mu }p_{1}^2+p_{2}^{\mu }p_{2}^2+p_{3}^{\mu } p_{3}^2+p_{4}^{\mu }p_{4}^2\right) \bigg \rbrace +\text {2 perms,} \end{aligned} \end{aligned}$$
(91)

where, again, the permutations are obtained by a cyclic permutation of the external indices 2, 3 and 4.

1.4 Computation of \({\varGamma }^{\left( 6\right) }\)

The \({{\varGamma }}^{\left( 6\right) }\) vertex has corrections linear in \(c^{\varLambda }_{\mu }\) to leading order in the \(\frac{1}{N}\) expansion (i.e. to order \(\frac{1}{N^2}\)), but it may also have contributions coming from the types of diagrams shown in Fig. 7 where \(a_\mu \) or \(b_\mu \) is inserted at the core (i.e. the inner loop with three propagator) or, as before, attached to an external leg. However, these corrections do not contribute to the scaling dimensions of the operators under study because, as explained above, \(c_\mu \) does not contribute to the flows of \(a_\mu \) and \(b\mu \) at leading order in 1 / N. We therefore do not compute these corrections.

Fig. 7
figure 7

Left: a diagram contributing to \({\varGamma }^{\left( 6\right) },\) linear in \(a_\mu \) or \(b_\mu \). Right: a diagram contributing to \({\varGamma }^{\left( 6\right) }\) proportional to \(c_\mu \)

The diagrams to be computed are exceptionally simple since they have a \(c_\mu \) at the core with no loop and then just chain of bubbles with u perturbations, these are schematically shown in Fig. 7.

The diagram (proportional to \(\delta _{i_{1}i_{2}}\delta _{i_{3}i_{4}}\delta _{i_{5}i_{6}}\)) with a chain with \(n_1\) couplings \(\hat{u}^{\varLambda }\) attached to the external momentums \(p_1\) and \(p_2,\) a chain with \(n_2\) couplings \(\hat{u}^{\varLambda }\) attached to the external momentums \(p_3\) and \(p_4,\) a chain with \(n_3\) couplings \(\hat{u}^{\varLambda }\) attached to the external momentums \(p_5\) and \(p_6\) and one \(\hat{c}^{\varLambda }_{\mu }\) at the core is equal to:

$$\begin{aligned}&\frac{-i\hat{c}^{\varLambda }_{\mu }}{2N^2} \left( \frac{-\hat{u}^{\varLambda }}{4!}\right) ^{n_1}\left( \frac{-\hat{u}^{\varLambda }}{4!}\right) ^{n_2}\left( \frac{-\hat{u}^{\varLambda }}{4!}\right) ^{n_3}\frac{4^{n_1+n_2+n_3}(n_1+n_2+n_3)!C^{n_1+n_2+n_3+1}_{1}}{\left( n_1+n_2+n_3+1\right) !} \nonumber \\&\quad \times \left( {\varPi }\left( p_{1}+p_{2}\right) \right) ^{n_1}\left( {\varPi }\left( p_{3}+p_{4}\right) \right) ^{n_2}\left( {\varPi }\left( p_{5}+p_{6}\right) \right) ^{n_3} \nonumber \\&\quad \times \Big [\left( p_{1}+p_{2}\right) ^{\mu }(p_{1}+ p_{2})^2+\left( p_{3}+p_{3}\right) ^{\mu }(p_{3}+ p_{4})^2+\left( p_{5}+p_{6}\right) ^{\mu }(p_{5}+ p_{6})^2\Big ] \nonumber \\&\quad + \text {perms.} \end{aligned}$$
(92)

1.5 Running Couplings

A convenient way to deduce the scaling dimensions of the operators coupled to \(\hat{a}^{\varLambda }_{\mu }\) and \(\hat{b}^{\varLambda }_{\mu },\) is to introduce an infrared regulator in propagators:

$$\begin{aligned} \frac{1}{q^2+r^{\varLambda }}\rightarrow \frac{1}{q^2+r^{\varLambda }+R_k(q)} \end{aligned}$$
(93)

and study the running of the various couplings when varying the regulator. We thus define the renormalized couplings as:

$$\begin{aligned} {\varGamma }_{i_1i_2}^{(2)}(0)&=r_k \delta _{i_1i_2},\nonumber \\ {\varGamma }_{i_1i_2i_3i_4}^{(4)}(0,0,0)&=\frac{\hat{u}_k}{3N} \Big (\delta _{i_1i_2}\delta _{i_3i_4}+\delta _{i_1i_3} \delta _{i_2i_4}+\delta _{i_1i_3}\delta _{i_2i_4}\Big ),\nonumber \\ {\varGamma }^{\left( 4\right) , {\mathcal {O}}(p^3)}_{i_1i_2i_3i_4}(p_1,p_2,p_3)&=i\frac{\delta _{i_1i_2}\delta _{i_3i_4}}{2N} \bigg \lbrace (\hat{a}^k_{\mu }-\hat{b}^k_{\mu }) \left( p_{1}+p_{2}\right) ^{\mu }\left[ p_{1} \cdot p_{2} - p_{3} \cdot p_{4}\right] \nonumber \\&\quad -\hat{b}^k_{\mu }\left( p_{1}^{\mu }p_{1}^2+ p_{2}^{\mu }p_{2}^2+p_{3}^{\mu }p_{3}^2+p_{4}^{\mu }p_{4}^2\right) \bigg \rbrace \nonumber \\&\quad +\text {2 perms.}, \nonumber \\ {\varGamma }^{\left( 6\right) , \mathcal O(p^3)}_{i_1i_2i_3i_4i_5i_6}(p_1,p_2,p_3,p_4,p_5)&=-i\frac{\delta _{i_1i_2}\delta _{i_3i_4}\delta _{i_5i_6} \hat{c}^k_{\mu }}{2N^2} \nonumber \\&\quad \times \bigg \lbrace \left( p_{1}+p_{2}\right) ^{\mu }(p_{1}+ p_{2})^2 \nonumber \\&\quad +\left( p_{3}+p_{3}\right) ^{\mu }(p_{3}+ p_{4})^2+\left( p_{5}+p_{6}\right) ^{\mu }(p_{5}+ p_{6})^2 \bigg \rbrace \nonumber \\&\quad +\text {14 perms.} \end{aligned}$$
(94)

One can then conclude that the running couplings are:

$$\begin{aligned}&u_k =\frac{u}{1+\frac{u_{\varLambda }}{6}{\varPi }_k\left( 0\right) },\quad a^\mu _k=\frac{a^\mu _{\varLambda }}{1+\frac{u_{\varLambda }}{6}{\varPi }_k\left( 0\right) }, \quad b^\mu _k=\frac{b^\mu _{\varLambda }}{1+\frac{u_{\varLambda }}{6}{\varPi }_k\left( 0\right) } \nonumber \\&c^\mu _k =\frac{c^\mu _{\varLambda }}{\big (1+\frac{u_{\varLambda }}{6}{\varPi }_k\left( 0\right) \big )^3}+Y_a a^\mu _{\varLambda }+Y_b b^\mu _{\varLambda },\quad r_k=r^{{\varLambda }}+{\overline{{\varSigma }}}_{k}, \end{aligned}$$
(95)

where the functions \({\overline{{\varSigma }}}_{k}\) and \({\varPi }_{k}\) are calculated with the introduction of the infrared regulator:

$$\begin{aligned} {\overline{{\varSigma }}}_{k}\left( r_{\varLambda }\right)&=\frac{u_{\varLambda }}{6}\int _q\frac{1}{q^2+r_k+R_{k}\left( q^2\right) }, \end{aligned}$$
(96)
$$\begin{aligned} {\varPi }_{k}\left( 0\right)&=\int _q\frac{1}{\left( q^2+r_k+R_{k}\left( q^2\right) \right) ^{2}}. \end{aligned}$$
(97)

Taking this into account, we obtain the flow of the running couplings:

$$\begin{aligned} \partial _{t}r_k&=-\frac{\hat{u}_k}{6}\int _q \partial _{t}R_{k}(q) G_k^2(q),\nonumber \\ \partial _{t}\hat{u}_k&=-\frac{\hat{u}_k^3}{18}\int _q G_k^3(q)\int _{q'} \partial _{t}R_{k}(q') G_k^2(q') +\frac{\hat{u}_k^2}{3}\int _q \partial _{t}R_{k}(q) G_k^3(q),\nonumber \\ \partial _{t}\hat{a}_k^{\mu }&=-\frac{\hat{a}_k^\mu \hat{u}_k^2}{18} \int _q G_k^3(q)\int _{q'}\partial _{t}R_{k}(q') G_k^2(q') +\frac{\hat{a}_k^{\mu }\hat{u}_k}{3}\int _q \partial _{t} R_{k}(q) G_k^3(q),\nonumber \\ \partial _{t}\hat{b}_k^{\mu }&=-\frac{\hat{b}_k^\mu \hat{u}_k^2}{18} \int _q G_k^3(q)\int _{q'}\partial _{t}R_{k}(q') G_k^2(q') +\frac{\hat{b}_k^{\mu }\hat{u}_k}{3}\int _q \partial _{t}R_{k}(q) G_k^3(q), \nonumber \\ \partial _{t}\hat{c}_k^{\mu }&=-\frac{\hat{c}_k^\mu \hat{u}_k^2}{6}\int _q G_k^3(q)\int _{q'}\partial _{t}R_{k}(q') G_k^2(q') +\hat{c}_k^{\mu }\hat{u}_k\int _q \partial _{t}R_{k}(q) G_k^3(q) \nonumber \\&\quad +X_a \hat{a}_k^{\mu }+X_b \hat{b}_k^{\mu }, \end{aligned}$$
(98)

where in the previous equations,

$$\begin{aligned} G_k(q)= \frac{1}{q^2+r_k+R_{k}\left( q^2\right) }. \end{aligned}$$

Introducing dimensionless and renormalized variables as explained in the main text, we retrieve the flow equations (37).

Numerical Method

We describe in this section the details of the numerical method used to determine the scaling dimensions in the \({\mathcal {O}}(\partial ^3)\) approximation of NPRG.

We first determine the fixed point of the O(N) model at \({\mathcal {O}}(\partial ^2)\) in a \({\tilde{\rho }}\) grid with \(N_\rho +1\) sites in a box \({\tilde{\rho }} \in [0,L_\rho ]\) (this corresponds to a step in the \({\tilde{\rho }}\)-lattice \({\varDelta }\rho =L_\rho /N_\rho \)). The derivatives are approximated by centered finite differences with five points (with the exception of the edges, i.e. the first two and last two sites, where we use lateral finite differences).

The internal momentum integrals that appear are one dimensional (the angular part is just a constant) and are calculated by a Legendre–Gauss quadrature with \(N_q\) points in a box of size \(|L_q|\equiv \frac{q_{max}}{k}\).

The normalization condition is fixed as \(\tilde{Z}(\tilde{\rho }_i)|_{i=N_\rho /3}=1,\) where \(\tilde{Z}(\tilde{\rho })\) is the dimensionless version of \(Z_k(\rho )\) and \(\tilde{\rho }_i\) is the value of \(\tilde{\rho }\) at site i. On top of this, the size of the box \(L_\rho \) is adjusted every time a new set of parameter is considered so that the minimum of the potential at the fixed point falls in the site \(i=N_\rho /3\).

The parameters were chosen so as that the numerical error in the leading exponents (\(\eta \) and \(\nu \)) is below one per mille in \(d=3\). Then we varied the values of d for each N. In particular we used the exponential regulator \(R_{k}(q^2)={\mathcal {Z}}_{k}q^2 r(q^2/k^2)\) with:

$$\begin{aligned} r(y)=\frac{\alpha }{e^{y}-1}, \end{aligned}$$
(99)

where \({\mathcal {Z}}_k\) is the field renormalization which is related to the running anomalous dimension by \(\partial _t{\mathcal {Z}}_k=-\eta _k{\mathcal {Z}}_k\) (when approaching the fixed point \(\eta _k\) approaches the field anomalous dimension). The parameter \(\alpha \) was fixed in order to estimate the dependence of the results on the arbitrary regulating function \(R_k(q)\). In order to do so, for each N we employ the criterium of minimal sensitivity (PMS) [76]. That is, we choose the value of \(\alpha \) which minimizes the dependence on the regulator of some observable by varying the regulator in the family (99). This procedure has been tested before and shown to be very effective and predictive [58, 59]. In the present implementation we choose \(\alpha \) to be an extremum of the critical exponent \(\eta \) for \(d=3\) (no significative difference was observed when doing PMS on another quantity such as critical exponent \(\nu \)).

After obtaining the fixed point, we computed the eigenvectors of the \(3N_\rho \) linear system (\(N_\rho \) variables for each function \(\tilde{a}_{\mu },\)\(\tilde{b}_{\mu }\) and \(\tilde{c}_{\mu }\)) at the fixed points determined previously. To solve numerically the \({\mathcal {O}}(\partial ^3)\) of the DE for the O(N) model the following set of parameters was used:

$$\begin{aligned} N_\rho =60,\quad L_\rho =3\rho _0,\quad N_q=15,\quad L_q=4.2. \end{aligned}$$
(100)

Inequalities

In the present Appendix, we prove by induction the inequality (51) for \(T\ge T_c\). We use generalizations of Griffiths and Lebowitz inequalities which were proven for \(N=2,3,\) and 4 [41,42,43,44,45,46]. The inequalities obtained in those references are the following:

$$\begin{aligned}&\langle \phi _1(x^{(1)}_1)\ldots \phi _1(x^{(1)}_{n_1})\phi _2(x^{(2)}_1)\ldots \phi _2(x^{(2)}_{n_2})\ldots \phi _N(x^{(N)}_1)\ldots \phi _N(x^{(N)}_{n_N})\rangle \ge 0, \end{aligned}$$
(101)
$$\begin{aligned}&\langle \phi _\alpha (x_1)\ldots \phi _\alpha (x_n)\phi _\alpha (y_1)\ldots \phi _\alpha (y_m)\rangle \ge \langle \phi _\alpha (x_1)\ldots \phi _\alpha (x_n)\rangle \langle \phi _\alpha (y_1)\ldots \phi _\alpha (y_m)\rangle , \end{aligned}$$
(102)
$$\begin{aligned}&\langle \phi _\alpha (x_1)\ldots \phi _\alpha (x_n)\phi _\beta (y_1)\ldots \phi _\beta (y_m)\rangle \nonumber \\&\quad \le \langle \phi _\alpha (x_1)\ldots \phi _\alpha (x_n)\rangle \langle \phi _\beta (y_1)\ldots \phi _\beta (y_m)\rangle \,\mathrm {with}\,\alpha \ne \beta . \end{aligned}$$
(103)

Inequality (101) and (102) are very similar to the Griffiths inequalities I and II for a scalar field and (103) is very similar to the Lebowitz inequality.

First of all, it is clear that for \(T\ge T_c\) we need only to consider correlations with n and m with the same parity. In order to begin the induction, we observe that the inequality is trivially true for \(n+m=2\). We assume now the validity of (51) for all values of m and n such that \(m+n<N\). Under this hypothesis, we then prove its validity for all \(\{m,n\}\) such that \(m+n=N\).

We first consider the case where n and m are odd. This case is simpler because the second term of the l.h.s. of (51) is zero because we are considering \(T\ge T_{c}\) and (101) readily shows that \(\nonumber G_{m,n}^{odd}\ge 0\). By symmetry, the structure of the two point function must take the form:

$$\begin{aligned} G_{m,n}^{odd}\left( x,y\right)&\equiv \left\langle \varphi _{i_1}\left( x\right) \cdots \varphi _{i_m} \left( x\right) \right. \left. \varphi _{j_1}\left( y\right) \cdots \varphi _{j_n}\left( y\right) \right\rangle = \sum _{l=0}^{\frac{n-1}{2}}f_{l}\left( x,y\right) \nonumber \\&\quad \times \Big [\delta _{i_{1}j_{1}}\cdots \delta _{i_{2l+1}j_{2l+1}} \delta _{i_{2l+2}i_{2l+3}}\delta _{j_{2l+2}j_{2l+3}}\cdots \delta _{i_{m-1}i_{m}}\delta _{j_{n-1}j_{n}} \nonumber \\&\quad +\left( \frac{m!n!}{\left( 2l+1\right) ! \left( m-2l-1\right) !!\left( n-2l-1\right) !!}-1\right) \text {perms.}\Big ], \end{aligned}$$
(104)

where without loss of generality we focused on the case \(m\ge n\). In the previous sum, \(2l+1\) corresponds to the number of Kronecker delta which connect i’s with j’s. In the first step of the proof, we consider the following configuration of indices:

$$\begin{aligned}&i_{k}= {\left\{ \begin{array}{ll} 1 &{} \text {for}\quad k=1, \\ 2 &{} \text {for}\quad k=2,\ldots ,m, \end{array}\right. } \\&j_{k}= {\left\{ \begin{array}{ll} 2 &{} \text {for}\quad k=1,\ldots ,2s, \\ 1 &{} \text {for}\quad k=2s+1,\ldots ,n \end{array}\right. } \end{aligned}$$

with s ranging from 0 to \(\frac{n-1}{2}\). (For \(s=0,\) all the \(j_k\) are 1.) We can now apply inequality (103) for these index configurations to get:

$$\begin{aligned} {\tilde{G}}_{m,n}^{odd}\left( x,y\right)&\le \left\langle \varphi _{1}\left( x\right) \left( \varphi _{1}\left( y\right) \right) ^{n-2s}\right\rangle \left\langle \left( \varphi _{2}\left( x\right) \right) ^{m-1} \left( \varphi _{2}\left( y\right) \right) ^{2s}\right\rangle \nonumber \\&\le CG\left( x-y\right) , \end{aligned}$$
(105)

where the tilde on G indicates that it is taken in a particular configuration of indices and where we have used the recursion hypothesis (51) for \(m'+n'<m+n\) and the fact that \(\left\langle \left( \varphi _{2}\left( x\right) \right) ^{m-1}\left( \varphi _{2}\left( y\right) \right) ^{2s}\right\rangle \) is bounded by a constant to obtain the last line. Using, on the other hand the decomposition (104), we get:

$$\begin{aligned} \sum _{t=0}^{s} \frac{\left( 2s\right) !\left( n-2s\right) !!\left( m-1\right) !}{\left( 2t\right) !\left( 2s-2t\right) !!\left( m-1-2t\right) !!}f_{t}\left( x,y\right) \le CG\left( x-y\right) . \end{aligned}$$
(106)

By considering the different possible values of s,  we easily show that the \(f_t\) are bounded by \(G(x-y)\) (up to a multiplicative constant) which ensure that the inequality is valid for \(G_{m,n}^{odd}\).

The even case is a bit different since the second term of the l.h.s. of (51) is nonzero. We again choose \(m\ge n\) and look at the structure of the two point function:

$$\begin{aligned} G_{m,n}^{even}\left( x,y\right)&\equiv \left\langle \varphi _{i_1}\left( x\right) \cdots \varphi _{i_m} \left( x\right) \right. \left. \varphi _{j_1}\left( y\right) \cdots \varphi _{j_n}\left( y\right) \right\rangle = \sum _{l=0}^{\frac{n}{2}}g_{l}\left( x,y\right) \nonumber \\&\quad \times \left( \delta _{i_{1}j_{1}}\cdots \delta _{i_{2l}j_{2l}} \delta _{i_{2l+1}i_{2l+2}}\delta _{j_{2l+1}j_{2l+2}} \cdots \delta _{i_{m-1}i_{m}}\delta _{j_{n-1}j_{n}}\right. \nonumber \\&\quad +\left. \left( \frac{m!n!}{\left( 2l\right) !\left( m-2l\right) !! \left( n-2l\right) !!}-1\right) \text {perms.}\right) . \end{aligned}$$
(107)

Inequality (101) readily imposes that \(G_{m,n}^{even}\left( x,y\right) \ge 0\). We now proceed in the same way as for the odd case. We take a configuration of indices:

$$\begin{aligned}&i_{k}= {\left\{ \begin{array}{ll} 1 &{} \text {for}\quad k=1, \\ 2 &{} \text {for}\quad k=2,\ldots ,m, \end{array}\right. } \\&j_{k}= {\left\{ \begin{array}{ll} 2 &{}\text {for}\quad k=1,\ldots ,2s-1, \\ 1 &{}\text {for}\quad k=2s,\ldots ,n \end{array}\right. } \end{aligned}$$

with s ranging from 1 to \(\frac{n}{2}\). These configurations in combination with the inequality (103) impose an upper bound on a strict conical combinations of the \(g_{i}\) functions not involving the \(g_{0}\) function:

$$\begin{aligned} {\tilde{G}}_{m,n}^{even}\left( x,y\right)&=\sum _{t=1}^{t=s} \frac{\left( 2s-1\right) !\left( n-2s+1\right) !!\left( m-1\right) !}{\left( 2t-1\right) !\left( 2s-2t\right) !!\left( m-2t\right) !!}g_{t} \left( x,y\right) \nonumber \\&\le \left\langle \varphi _{1}\left( x\right) \left( \varphi _{1}\left( y\right) \right) ^{n-2s+1}\right\rangle \left\langle \left( \varphi _{2}\left( x\right) \right) ^{m-1} \left( \varphi _{2}\left( y\right) \right) ^{2s-1}\right\rangle \nonumber \\&\le CG^{2}\left( x-y\right) , \end{aligned}$$
(108)

where we have made use of the validity of recursion hypothesis (51) for all \(m'+n'<m+n\) and the tilde is again used to recall that a particular configuration of indices was used. We, again, find a lower and an upper bound on a strict conical combination of the functions \(g_{s},\) with \(s=1,\ldots ,\frac{n}{2}\). This implies that the absolute value of each of these \(g_{s}\) functions (with \(s\ne 0\)) is bounded by a constant times \(G^{2}\left( x-y\right) \) (by arguments identical to the odd case). To complete the argument we need to study the function \(g_{0}\). This one is clearly particular because it corresponds to the case where no i or j are connected through a Kronecker delta [see Eq. (104)]. To do this we consider the even simpler configuration where all indices \(i=1\) and all \(j=2,\) this yields for the two-point function:

$$\begin{aligned} 0 \le G_{m,n}^{even}\left( x,y\right) =g_{0}\left( x,y\right) \left( m-1\right) !!\left( n-1\right) !!\le \left\langle \varphi ^{m}_{1}\left( 0\right) \right\rangle \left\langle \varphi ^{n}_{1}\left( 0\right) \right\rangle \end{aligned}$$
(109)

from which immediately follows that:

$$\begin{aligned} g_{0}\left( x,y\right) -\frac{\left\langle \varphi ^{m}_{1}\left( 0\right) \right\rangle \left\langle \varphi ^{n}_{1}\left( 0\right) \right\rangle }{\left( m-1\right) !!\left( n-1\right) !!} \le 0. \end{aligned}$$
(110)

Let’s consider now all indices \(i,j=1\) to obtain a lower bound on a strict conical combination of all the g functions by making use of the inequality (102):

$$\begin{aligned} \sum _{l=0}^{\frac{n}{2}}\frac{m!n!}{\left( 2l\right) !\left( m-2l \right) !!\left( n-2l\right) !!}g_{l}\left( x,y\right) \ge \left\langle \varphi ^{m}_{1}\left( 0\right) \right\rangle \left\langle \varphi ^{n}_{1}\left( 0\right) \right\rangle \end{aligned}$$
(111)

and combining this with (110) we obtain a lower bound:

$$\begin{aligned} \begin{aligned}&g_{0}\left( x,y\right) -\frac{\left\langle \varphi ^{m}_{1}\left( 0\right) \right\rangle \left\langle \varphi ^{n}_{1}\left( 0\right) \right\rangle }{\left( m-1\right) !!\left( n-1\right) !!} \\&\quad \ge -\frac{1}{\left( m-1\right) !!\left( n-1\right) !!}\sum _{l=1}^{\frac{n}{2}} \frac{m!n!}{\left( 2l\right) ! \left( m-2l\right) !!\left( n-2l\right) !!}g_{l}\left( x,y\right) \\&\quad \ge C G^{2}\left( x-y\right) \end{aligned} \end{aligned}$$
(112)

with a constant \(C<0\).

So, we have bounded all the g functions with the exception of \(g_{0}\) for which the bound involves \(g_{0}\) minus a constant. It turns out that this constant has a simple interpretation: when inserted in Eq. (107), the multiplicative constants simplify and we are just left with \(\left\langle \varphi ^{m}_{1}\left( 0\right) \right\rangle \left\langle \varphi ^{n}_{1}\left( 0\right) \right\rangle \), which exactly compensates the disconnected part appearing in (51). This concludes the proof of (51).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

De Polsi, G., Tissier, M. & Wschebor, N. Conformal Invariance and Vector Operators in the O(N) Model. J Stat Phys 177, 1089–1130 (2019). https://doi.org/10.1007/s10955-019-02411-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-019-02411-3

Keywords

Navigation