Skip to main content
Log in

Basic Introduction to Higher-Spin Theories

  • Review
  • Published:
International Journal of Theoretical Physics Aims and scope Submit manuscript

Abstract

This is a collection of my lecture notes on the higher-spin theory course given for students at the Institute for Theoretical and Mathematical Physics, Lomonosov Moscow State University. The goal of these lectures is to give an introduction to higher-spin theories accessible to master level students which would enable them to read the higher-spin literature. I start by introducing basic relevant notions of representation theory and the associated field-theoretic descriptions. Focusing on massless symmetric fields I review different approaches to interactions as well as the no-go results. I end the lectures by reviewing some of the currently available positive results on interactions of massless higher-spin fields, namely, holographic, Chern-Simons and chiral higher-spin theories.

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.

Similar content being viewed by others

Notes

  1. Strictly speaking, it is required that free fields realise projective representations, that is representations up to a phase. This extension is important if one needs to deal with fermions. Inclusion of fermions in the higher-spin context is relatively straightforward and, usually, does not add much except technical difficulties. Fermions will not be discussed here. For the discussion on fermions in the context of representation theory, see e.g. [7].

  2. See, e.g. [3] for a more precise statement and for the proof of this theorem.

  3. Note that this is only true for unitary representations. Indeed, if unitarity is not imposed, the best one can achieve when diagonalising even a single matrix is the Jordan canonical form.

  4. By the Poincare group we mean its component, which is continuously connected to the unity. For the whole Poincare group – with the time reversal and the parity transformation included – positive and negative modes belong to the same orbit.

  5. The above discussion on irreducibility applies to any signature. Though, keep in mind that tensorial representations for non-Euclidean signature are non-unitary.

  6. A systematic search of the minimal set of traceless symmetric fields that would be sufficient to write a Lagrangian for a free massive spin-s field was carried out in [10]. The Fronsdal action was obtained as the massless limit of [10]. The Fronsdal action for a massless spin-s field features traceless symmetric tensors of ranks s and \(s-2\), see (3.12), (3.13). The rank-s traceless field is, clearly, necessary as it features (3.1) explicitly. The rank-\((s-2)\) traceless field is auxiliary. It is needed because for \(\xi \) free of any differential constraints gauge transformation (3.2) violates tracelessness of \(\varphi \): even for \(\xi \) traceless, \(\varphi \) should have at least one non-trivial trace originating from the divergence of \(\xi \). Thus, at least one auxiliary symmetric and traceless rank-\((s-2)\) field is necessary to accommodate the trace of \(\varphi \). This proves that Fronsdal’s set of off-shell fields is, indeed, minimal provided we do not allow differential constraints on gauge parameters imposed off-shell.

  7. Relation between F and G is the higher-spin analogue of that between \(R_{\mu \nu }\) and \(G_{\mu \nu }\equiv R_{\mu \nu }-\frac{1}{2}Rg_{\mu \nu }\) in General Relativity.

  8. In [17] it was shown that the divergence-free constraint can be solved in terms of gauge parameters without differential constraints, which leads to a more complex pattern of reducible gauge transformations.

  9. Unlike most of the higher-spin literature, we define the cosmological constant with the same factors that are typically used in General Relativity. For example, in [8] the cosmological constant is defined as \(\Lambda \equiv -1/l^2\).

  10. Somewhat speculatively, the lowest-energy space for \(SO(d-1,2)\) representations can be compared with \(\varphi _{p,\sigma }\) space with \(p=(m,0,\dots ,0)\) for massive fields in flat space. In particular, for the latter states energy, indeed, acquires minimal value \(E_0=m\) and these also furnish a representation of \(SO(d-1)\), which is the Wigner little group in the massive case. Yet, there are, major differences between these two setups. Most importantly, transvections do not commute, so, only energy takes a definite value on \(|E_0,\mathbb {Y}_0\rangle \). Still, this analogy was used fruitfully to construct UIR’s of \(SO(d-1,2)\) in a form which is very intuitive from the flat space perspective [23].

  11. The fact that the shape \(\mathbb {Y}_0\) carries over to the field-theory description in such a straightforward manner is not entirely obvious. This can be shown, for example, by comparing the values of higher Casimir operators or, more directly, one can find the lowest-energy state on the field theory side and identify the associated \(E_0\) and \(\mathbb {Y}_0\). We will not do that, as our goal, anyway, is not to provide a mathematically strict proof of the equivalence of the two representations, but rather support this statement with some evidence and familiarise the reader with the frequently used tools.

  12. Due to the presence of the ambient metric we do not distinguish vectors and 1-forms. Moreover, the ambient metric allows us to unambiguously decompose ambient space vectors at \(X^2=-l^2\) into AdS tangent and AdS transverse parts, which are, moreover, orthogonal to each other. Namely, tangent vectors \(V^A\) are defined by the condition that these are annihilated by 1-form \(d(X^2+l^2)\), that is, \(V^AX_A=0\). Its orthogonal complement, spanned by vectors proportional to \(X^A\), can be identified as the AdS transverse vector space. By lowering indices, we get the analogous decomposition for 1-forms.

  13. For general embeddings, the ambient space Levi-Civita connection induces a Levi-Civita connection onto the embedded submanifold. The ambient and the induced geometric objects are connected by the Gauss-Codazzi equations.

  14. As usual, we ignore fermions in our analysis.

  15. This happens due to the isomorphism \(so(3,2)\sim sp(4,\mathbb {R})\). It may be instructive to consider a 4d symplectic manifold and Hamiltonians quadratic in the Darboux coordinates \(z=\{q^1,q^2,p^1,p^2 \}\). Via the Poisson bracket these generate vector fields \(\xi ^i(z)\partial _i\) that preserve the symplectic form and, moreover, being linear in the Darboux coordinates, \(\xi \propto z\), preserve the origin \(z=0\). Then, it is not hard to see that these vector fields induce linear transformations on the tangent bundle at the origin, which preserve the symplectic form at this point. Accordingly, these generate \( sp(4,\mathbb {R})\) and \(\partial _j \xi ^i(z)\) are the associated \(sp(4,\mathbb {R})\) matrices. Representation (6.12) gives a version of this construction, in which the Poisson bracket is replaced with the commutator.

  16. The fact that any operator on the Hilbert space of states generated from the vacuum by raising operators is expressible in terms of creation and annihilation operators is rather standard and can be found e.g. in [7].

  17. This analysis can be streamlined if one groups the oscillators into an \(sp(4,\mathbb {R})\) vector \(Y^A=\{a,a^\dagger ,b,b^\dagger \}\). In these terms it is not hard to see that the higher-spin algebra decomposes into symmetric tensors of \(sp(4,\mathbb {R})\).

  18. Strictly speaking, this argument shows that the representation is realised on the off-shell fields quotiented by pure gauge degrees of freedom. When quotienting out pure gauge degrees of freedom, off-shell gauge transformations and off-shell gauge parameters are used.

  19. Strictly speaking, there is a possibility that the first non-trivial vertex is quartic or higher order in fields, which we do not consider.

  20. We will derive this and analogous statements for general spins in the next section.

  21. Indeed, if

    $$\begin{aligned} f^{abc}f_{bd}{}^{e}=f^{abe}f_{bd}{}^{c} \end{aligned}$$
    (8.40)

    then

    $$\begin{aligned} f^{abc}f_{bd}{}^{e}=f^{abe}f_{bd}{}^{c}=-f^{eba}f_{bd}{}^{c}=-f^{ebc}f_{bd}{}^{a}=f^{cbe}f_{bd}{}^{a}= f^{cba}f_{bd}{}^{e}=-f^{abc}f_{bd}{}^{e}, \end{aligned}$$
    (8.41)

    which entails

    $$\begin{aligned} f^{abc}f_{bd}{}^{e}=0. \end{aligned}$$
    (8.42)

    Taking u and v any auxiliary vectors, we can construct

    $$\begin{aligned} F^a\equiv f^{abc}u_bv_c. \end{aligned}$$
    (8.36)

    Then (8.38) implies \(F^2=0\). Recalling that the internal space metric is \(\delta _{ab}\), this entails \(F=0\). Vanishing of F for any u and v implies that f equals zero.

  22. There is no summation over j here.

  23. This is related to the fact that amplitudes are obtained from the correlators which, in turn, are computed in the interaction picture, in which the external lines evolve with the free Hamiltonian.

  24. For a symmetric rank-s tensor \(\varphi ^{\mu (s)}\) its traceless part is given by an expression of the form

    $$\begin{aligned} \varphi _{0}^{\mu (s)}\equiv \varphi ^{\mu (s)}+\alpha _1 \eta ^{\mu \mu }\varphi _\nu {}^{\nu \mu (s-2)}+\alpha _2 \eta ^{\mu \mu }\eta ^{\mu \mu }\varphi _{\nu \rho }{}^{\nu \rho \mu (s-4)}+\dots , \end{aligned}$$
    (9.12)

    where coefficients \(\alpha _1\), \(\alpha _2\), \(\dots \) are defined from the requirement that \(\varphi _0\) is traceless. Clearly, the operation defined by (9.10) is a projection: for \(\varphi \) traceless one has \(\varphi _0=\varphi \). The traceless projection (9.10) can be alternatively implemented via contraction with the traceless projector \(\mathcal{P}_s^d\)

    $$\begin{aligned} \varphi _{0}^{\mu (s)}\equiv (\mathcal{P}_s^d)^{\mu (s)}{}_{\nu (s)}\varphi ^{\nu (s)} = (\delta ^\mu {}_\nu \dots \delta ^\mu {}_\nu + \alpha _1\eta ^{\mu \mu }\eta _{\nu \nu }\delta ^\mu {}_\nu \dots \delta ^\mu {}_\nu +\dots ) \varphi ^{\nu (s)} . \end{aligned}$$
    (9.13)

    Tensor \(\mathcal{P}_s^d\) is, clearly, symmetric and traceless on each group of indices. Employing polarisation vectors \(u_1\) and \(u_2\) we can turn it into a generating function that satisfies (9.12).

  25. To make this statement more precise, we should be more clear on what we mean by locality. For example, one can define that local vertices have finitely many derivatives. Then, the associated amplitudes are polynomial in the Mandelstam variables. One can consider a relaxed version of locality for which local amplitudes should not have singularities, that is these should be given by entire functions. These are two natural options for the analyticity properties required from local contact amplitudes.

  26. In the amplitude literature a somewhat different terminology is used. The property that the amplitude’s singularities match those contributed by exchanges is called unitarity, while locality refers to the property that amplitudes may only have single poles, no other singularities at tree level are allowed.

  27. We emphasise, that by on-shell trivial terms we mean those that vanish on the free equations of motion.

  28. Here \(\varphi ^\lambda \) is a vector potential against which the amplitude is supposed to be integrated, (9.4). We change notation from A to \(\varphi \) not to confuse it with the amplitude itself.

  29. Here ”\(\approx \)” denotes an approximate equality.

  30. Here, ”internal” refers to the fact that the internal algebra generators commute with the Poincare algebra. Equivalently, one can say that internal symmetries do not act neither on momenta nor on spin labels. In particular, according to this definition the U(1) associated with the Maxwell theory is regarded as internal symmetry.

  31. We remind the reader that we use ”internal” in the sense that the associated symmetries commute with the Poincare algebra. Abelian symmetries commute with all generators, hence, these are internal.

  32. In this section we will use the standard notation used in the AdS/CFT literature: D denotes the dimension of the AdS space, while \(d\equiv D-1\) is the dimension of its boundary.

  33. Strictly speaking, as we mentioned in the previous section, existence of a non-trivial Lie algebra with the spectrum consisting of flat space Killing tensors is not ruled out unless one makes additional assumptions. Still an important difference with the flat space is that in the AdS case the structure constants of the higher-spin algebra can, indeed, be induced from consistent cubic vertices, see e.g. [61]. In other words, the argument analogous to that of Section 10.3 does not rule out interactions in AdS.

  34. Mathematically, this means that the associated representation has the Gelfand-Kirillov dimension equal to \(D-1\). The Gelfand-Kirillov dimension is the standard way to characterise ”the size” of infinite-dimensional representations.

  35. Certainly, fields also become operators in quantum field theory. We will still use terminology of fields and operators depending on whether Klein-Gordon-like equations are involved.

  36. Lorentz invariance requires correlators to be functions of \(x^2_{ij}=(x_i-x_j)^2\). In the Euclidean case \(x_{ij}^2\ge 0\) and the only singularities of \((x^2_{ij})^\delta \) occur when \(x_i\rightarrow x_j\). Instead, in the Lorentzian signature \(x_{ij}^2\) changes the sign when \(x_i\) crosses the light cone with a vertex at \(x_j\), so at these points the correlator develops additional singularities. Depending on the way one analytically continues the correlator when crossing these singular light cones, one obtains correlators with different operator orderings.

  37. One often distributes powers of l between \(X^+\) and \(X^-\) differently: \((X^+,X^-) = z^{-1}(l^2,z^2+x^2)\). We will use (11.6) as it allows us to get rid of the l dependence in the coordinates for the AdS boundary.

  38. Note, however, a subtlety: the coordinate of an on-shell field cannot be well-defined as \(\delta (x-x_0)\) does not obey wave equations, so whether it is the AdS or the flat space case, an on-shell field is always a superposition of states with definite coordinates.

  39. In the higher-spin case we will be interested in \(\Delta =d-2\), which is smaller than \(d-\Delta =2\) for \(d<4\).

  40. This is the same difference as between the Green function and the kernel of a differential equation.

  41. In (11.38) all odd spin currents are vanishing. This can be avoided if we replace O(N) with U(N) and consider currents of the form \(J\sim \bar{\varphi }^a \varphi _a\).

  42. Here by plane waves we mean, essentially, \(e^{ipx}\), which in flat space contribute through the Fourier transform.

  43. It is often referred to as the Cartan-Weyl gravity.

  44. We would like to remind the reader that we use notation \(\mathbb {Y}\) to highlight the fact that in addition to possessing the Young symmetry a tensor is also traceless. Tensors possessing only Young symmetries and not satisfying any trace constraints are denoted \(\textbf{Y}\).

  45. This trace is not unique, which follows from the fact that \(so(2,2)\sim sl(2,\mathbb {R})_L\oplus sl(2,\mathbb {R})_L\), see below.

  46. There are some exotic theories in which kinematic generators receive corrections as well [7].

  47. These equations have more solutions when understood in the sense of distributions. We will not discuss distributional solutions here.

  48. As explained in Section 7, once a vertex is trivial on-the-free-shell, it can be eliminated by field redefinitions. Vertices of the form \((\partial _i^- + \partial _i\bar{\partial }_i /\partial _i^+)(\dots )\) clearly vanish on the free shell, therefore, field redefinitions allow one to trade \(\partial _i^-\) for \(- \partial _i\bar{\partial }_i /\partial _i^+\) and eliminate \(\partial ^-\) entirely.

  49. Kinematics of massless 3-point scattering is singular and requires a separate discussion.

  50. Here \(|p\rangle _m\) is what was denoted \(\varphi _{p,\sigma }\) in Section 2.

  51. See Section 2 for notations.

References

  1. Bekaert, X., Boulanger, N., Campoleoni, A., Chiodaroli, M., Francia, D., Grigoriev, M., et al. (2022) Snowmass White paper: higher spin gravity and higher Spin symmetry. arXiv:2205.01567

  2. Sorokin, D.: Introduction to the classical theory of higher spins. AIP Conf. Proc. 767, 172 (2005). https://doi.org/10.1063/1.1923335. arXiv:hep-th/0405069

    Article  ADS  MathSciNet  MATH  Google Scholar 

  3. Bekaert, X. and Boulanger, N. (2006) The Unitary representations of the Poincare group in any spacetime dimension. In: 2nd Modave Summer School in Theoretical Physics. vol. 11. arXiv:hep-th/0611263

  4. Vasiliev, M. (2014) Introduction into higher-spin gauge theory. Lectures given at Utrecht University

  5. Rahman, R. and Taronna, M. (2015) From Higher Spins to Strings: A Primer. arXiv:1512.07932

  6. Kessel, P. (2017) The Very Basics of Higher-Spin Theory. PoS, Modave2016:001. https://doi.org/10.22323/1.296.0001. arXiv:1702.03694

  7. Weinberg, S. (1995) The Quantum Theory of Fields. vol. 1,2,3, Cambridge University Press

  8. Didenko, V. and Skvortsov, E. (2014) Elements of Vasiliev theory. arXiv:1401.2975

  9. Fronsdal, C.: Massless fields with integer spin. Phys. Rev. D. 18, 3624 (1978). https://doi.org/10.1103/PhysRevD.18.3624

    Article  ADS  Google Scholar 

  10. Singh, L. and Hagen, C. (1974) Lagrangian formulation for arbitrary spin. 1. The boson case. Phys. Rev. D., 9:898. https://doi.org/10.1103/PhysRevD.9.898

  11. Francia, D., Sagnotti, A.: Free geometric equations for higher spins. Phys. Lett. B 543, 303 (2002). https://doi.org/10.1016/S0370-2693(02)02449-8. arXiv:hep-th/0207002

    Article  ADS  MathSciNet  MATH  Google Scholar 

  12. Francia, D., Sagnotti, A.: On the geometry of higher spin gauge fields. Class. Quant. Grav. 20, S473 (2003). https://doi.org/10.1088/0264-9381/20/12/313. arXiv:hep-th/0212185

    Article  ADS  MathSciNet  MATH  Google Scholar 

  13. Segal, A.Y. (2001) A Generating formulation for free higher spin massless fields. arXiv:hep-th/0103028

  14. Sagnotti, A., Tsulaia, M.: On higher spins and the tensionless limit of string theory. Nucl. Phys. B 682, 83 (2004). https://doi.org/10.1016/j.nuclphysb.2004.01.024. arXiv:hep-th/0311257

    Article  ADS  MathSciNet  MATH  Google Scholar 

  15. Skvortsov, E.D., Vasiliev, M.A.: Transverse Invariant Higher Spin Fields. Phys. Lett. B 664, 301 (2008). https://doi.org/10.1016/j.physletb.2008.05.043. arXiv:hep-th/0701278

    Article  ADS  MathSciNet  MATH  Google Scholar 

  16. Campoleoni, A., Francia, D.: Maxwell-like Lagrangians for higher spins. JHEP 03, 168 (2013). https://doi.org/10.1007/JHEP03(2013)168. arXiv:1206.5877

    Article  Google Scholar 

  17. Francia, D., Lyakhovich, S.L., Sharapov, A.A.: On the gauge symmetries of Maxwell-like higher-spin Lagrangians. Nucl. Phys. B 881, 248 (2014). https://doi.org/10.1016/j.nuclphysb.2014.02.001. arXiv:1310.8589

    Article  ADS  MathSciNet  MATH  Google Scholar 

  18. Fang, J., Fronsdal, C.: Massless Fields with Half Integral Spin. Phys. Rev. D 18, 3630 (1978). https://doi.org/10.1103/PhysRevD.18.3630

    Article  ADS  Google Scholar 

  19. Labastida, J.M.F.: Massless Particles in Arbitrary Representations of the Lorentz Group. Nucl. Phys. B 322, 185 (1989). https://doi.org/10.1016/0550-3213(89)90490-2

    Article  ADS  MathSciNet  Google Scholar 

  20. Zinoviev, Y.M. (2001) On massive high spin particles in AdS. arXiv:hep-th/0108192

  21. Schuster, P., Toro, N.: Continuous-spin particle field theory with helicity correspondence. Phys. Rev. D 91, 025023 (2015). https://doi.org/10.1103/PhysRevD.91.025023. arXiv:1404.0675

    Article  ADS  Google Scholar 

  22. Bekaert, X. and Skvortsov, E.D. (2017) Elementary particles with continuous spin. Int. J. Mod. Phys. A, 32:1730019. https://doi.org/10.1142/S0217751X17300198. arXiv:1708.01030

  23. Fronsdal, C. (1974) Elementary particles in a curved space. ii. Phys. Rev. D, 10:589. https://doi.org/10.1103/PhysRevD.10.589

  24. Vasiliev, M.: Higher spin superalgebras in any dimension and their representations. JHEP 12, 046 (2004). https://doi.org/10.1088/1126-6708/2004/12/046. arXiv:hep-th/0404124

    Article  ADS  MathSciNet  Google Scholar 

  25. Metsaev, R.R.: Massless mixed symmetry bosonic free fields in d-dimensional anti-de Sitter space-time. Phys. Lett. B 354, 78 (1995). https://doi.org/10.1016/0370-2693(95)00563-Z

    Article  ADS  MathSciNet  Google Scholar 

  26. Minwalla, S.: Restrictions imposed by superconformal invariance on quantum field theories. Adv. Theor. Math. Phys. 2, 783 (1998). https://doi.org/10.4310/ATMP.1998.v2.n4.a4. arXiv:hep-th/9712074

    Article  MathSciNet  MATH  Google Scholar 

  27. Evans, N.T.: Discrete series for the universal covering group of the 3 \(+\) 2 de sitter group. Journal of Mathematical Physics 8, 170 (1967). https://doi.org/10.1063/1.1705183

    Article  ADS  MathSciNet  MATH  Google Scholar 

  28. Mack, G.: All unitary ray representations of the conformal group SU(2,2) with positive energy. Commun. Math. Phys. 55, 1 (1977). https://doi.org/10.1007/BF01613145

    Article  ADS  MathSciNet  MATH  Google Scholar 

  29. Siegel, W.: All free conformal representations in all dimensions. Int. J. Mod. Phys. A 4, 2015 (1989). https://doi.org/10.1142/S0217751X89000819

    Article  ADS  MathSciNet  Google Scholar 

  30. Metsaev, R.R.: Arbitrary spin massless bosonic fields in d-dimensional anti-de Sitter space. Lect. Notes Phys. 524, 331 (1999). https://doi.org/10.1007/BFb0104614. arXiv:hep-th/9810231

    Article  ADS  MathSciNet  Google Scholar 

  31. Metsaev, R.R.: Fermionic fields in the d-dimensional anti-de Sitter space-time. Phys. Lett. B 419, 49 (1998). https://doi.org/10.1016/S0370-2693(97)01446-9. arXiv:hep-th/9802097

    Article  ADS  MathSciNet  Google Scholar 

  32. Enright, T., Howe, R. and Wallach, N. (1983) A classification of unitary highest weight modules. In: Representation Theory of reductive groups: proceedings of the university of utah conference 1982, Trombi, P.C. ed., (Boston, MA), pp. 97–143, Birkhäuser Boston, https://doi.org/10.1007/978-1-4684-6730-7_7DOI

  33. Ferrara, S. and Fronsdal, C. (2000) Conformal fields in higher dimensions. In: 9th Marcel Grossmann Meeting on Recent Developments in Theoretical and Experimental General Relativity, Gravitation and Relativistic Field Theories (MG 9). pp. 508–527, 6, arXiv:hep-th/0006009

  34. Bourget, A. and Troost, J. (2018) The conformal characters. JHEP, 04:055. https://doi.org/10.1007/JHEP04(2018)055. arXiv:1712.05415

  35. Hirai, T.: On irreducible representations of the lorentz group of n-th order. Proceedings of the Japan Academy 38, 258 (1962). https://doi.org/10.3792/pja/1195523378

    Article  MathSciNet  MATH  Google Scholar 

  36. Schwarz, F.: Unitary irreducible representations of the groups so0(n, 1). Journal of Mathematical Physics 12, 131 (1971). https://doi.org/10.1063/1.1665471

    Article  ADS  MathSciNet  MATH  Google Scholar 

  37. Dobrev, V., Mack, G., Petkova, V., Petrova, S., Todorov, I.: Harmonic analysis: on the n-dimensional lorentz group and its application to conformal quantum field theory. Lecture Notes in Physics, Springer, Berlin Heidelberg (1977)

    MATH  Google Scholar 

  38. Basile, T., Bekaert X. and Boulanger, N. (2017) Mixed-symmetry fields in de Sitter space: a group theoretical glance. JHEP, 05:081. https://doi.org/10.1007/JHEP05(2017)081. arXiv:1612.08166

  39. Mikhailov, A. (2002) Notes on higher spin symmetries. arXiv:hep-th/0201019

  40. Fronsdal, C. (1979) Singletons and Massless, Integral Spin Fields on de Sitter Space (Elementary Particles in a Curved Space. 7.. Phys. Rev. D, 20:848. https://doi.org/10.1103/PhysRevD.20.848

  41. Bekaert, X., Meunier, E.: Higher spin interactions with scalar matter on constant curvature spacetimes: conserved current and cubic coupling generating functions. JHEP 11, 116 (2010). https://doi.org/10.1007/JHEP11(2010)116. arXiv:1007.4384

    Article  ADS  MathSciNet  MATH  Google Scholar 

  42. Sleight, C. (2017) Interactions in Higher-Spin Gravity: a Holographic Perspective. J. Phys. A, 50:383001. https://doi.org/10.1088/1751-8121/aa820c. arXiv:1610.01318

  43. Fang, J., Fronsdal, C.: Massless, Half Integer Spin Fields in De Sitter Space. Phys. Rev. D 22, 1361 (1980). https://doi.org/10.1103/PhysRevD.22.1361

    Article  ADS  MathSciNet  Google Scholar 

  44. Buchbinder, I.L., Pashnev, A., Tsulaia, M.: Lagrangian formulation of the massless higher integer spin fields in the AdS background. Phys. Lett. B 523, 338 (2001). https://doi.org/10.1016/S0370-2693(01)01268-0. arXiv:hep-th/0109067

    Article  ADS  MathSciNet  MATH  Google Scholar 

  45. Buchbinder, I.L., Krykhtin, V.A., Reshetnyak, A.A.: BRST approach to Lagrangian construction for fermionic higher spin fields in (A)dS space. Nucl. Phys. B 787, 211 (2007). https://doi.org/10.1016/j.nuclphysb.2007.06.006. arXiv:hep-th/0703049

    Article  ADS  MATH  Google Scholar 

  46. Brink, L., Metsaev, R.R., Vasiliev, M.A.: How massless are massless fields in AdS(d). Nucl. Phys. B 586, 183 (2000). https://doi.org/10.1016/S0550-3213(00)00402-8. arXiv:hep-th/0005136

    Article  ADS  MathSciNet  MATH  Google Scholar 

  47. Deser, S., Waldron, A.: Partial masslessness of higher spins in (A)dS. Nucl. Phys. B 607, 577 (2001). https://doi.org/10.1016/S0550-3213(01)00212-7. arXiv:hep-th/0103198

    Article  ADS  MathSciNet  MATH  Google Scholar 

  48. Dirac, P.A.M.: A Remarkable representation of the 3 + 2 de Sitter group. J. Math. Phys. 4, 901 (1963). https://doi.org/10.1063/1.1704016

    Article  ADS  MathSciNet  MATH  Google Scholar 

  49. Flato, M. and Fronsdal, C. (1978) One Massless Particle Equals Two Dirac Singletons: Elementary Particles in a Curved Space. 6.. Lett. Math. Phys., 2:421. https://doi.org/10.1007/BF00400170

  50. Bae, J.-B., Joung, E. and Lal, S. (2016) One-loop test of free SU(N ) adjoint model holography. JHEP, 04:061. https://doi.org/10.1007/JHEP04(2016)061. arXiv:1603.05387

  51. Berezin, F.A., Shubin, M.A.: The Schroedinger Equation. Springer, Dordrecht (1991)

    Book  Google Scholar 

  52. Basile, T., Bekaert, X., Boulanger, N.: Flato-Fronsdal theorem for higher-order singletons. JHEP 11, 131 (2014). https://doi.org/10.1007/JHEP11(2014)131. arXiv:1410.7668

    Article  Google Scholar 

  53. Fradkin, E.S., Vasiliev, M.A.: Candidate to the Role of Higher Spin Symmetry. Annals Phys. 177, 63 (1987). https://doi.org/10.1016/S0003-4916(87)80025-8

    Article  ADS  MathSciNet  Google Scholar 

  54. Eastwood, M.G.: Higher symmetries of the Laplacian. Annals Math. 161, 1645 (2005). https://doi.org/10.4007/annals.2005.161.1645. arXiv:hep-th/0206233

    Article  MathSciNet  MATH  Google Scholar 

  55. Vasiliev, M.A.: Nonlinear equations for symmetric massless higher spin fields in (A)dS(d). Phys. Lett. B 567, 139 (2003). https://doi.org/10.1016/S0370-2693(03)00872-4. arXiv:hep-th/0304049

    Article  ADS  MathSciNet  MATH  Google Scholar 

  56. Iazeolla, C., Sundell, P.: A Fiber Approach to Harmonic Analysis of Unfolded Higher-Spin Field Equations. JHEP 10, 022 (2008). https://doi.org/10.1088/1126-6708/2008/10/022. arXiv:0806.1942

    Article  ADS  MathSciNet  MATH  Google Scholar 

  57. Govil, K., Gunaydin, M.: Deformed Twistors and Higher Spin Conformal (Super-)Algebras in Four Dimensions. JHEP 03, 026 (2015). https://doi.org/10.1007/JHEP03(2015)026. arXiv: 1312.2907

    Article  ADS  MathSciNet  MATH  Google Scholar 

  58. Joung, E., Mkrtchyan, K.: Notes on higher-spin algebras: minimal representations and structure constants. JHEP 05, 103 (2014). https://doi.org/10.1007/JHEP05(2014)103. arXiv: 1401.7977

    Article  ADS  MathSciNet  MATH  Google Scholar 

  59. Basile, T., Bekaert, X. and Joung, E. (2018) Twisted Flato-Fronsdal Theorem for Higher-Spin Algebras. JHEP, 07:009. https://doi.org/10.1007/JHEP07(2018)009. arXiv: 1802.03232

  60. Berends, F.A., Burgers, G., van Dam, H.: On the Theoretical Problems in Constructing Interactions Involving Higher Spin Massless Particles. Nucl. Phys. B 260, 295 (1985). https://doi.org/10.1016/0550-3213(85)90074-4

    Article  ADS  MathSciNet  Google Scholar 

  61. Joung, E., Taronna, M.: Cubic-interaction-induced deformations of higher-spin symmetries. JHEP 03, 103 (2014). https://doi.org/10.1007/JHEP03(2014)103. arXiv: 1311.0242

    Article  Google Scholar 

  62. Boulanger, N., Leclercq, S.: Consistent couplings between spin-2 and spin-3 massless fields. JHEP 11, 034 (2006). https://doi.org/10.1088/1126-6708/2006/11/034. arXiv:hep-th/0609221

    Article  ADS  MathSciNet  Google Scholar 

  63. Bekaert, X., Boulanger, N., Leclercq, S.: Strong obstruction of the Berends-Burgers-van Dam spin-3 vertex. J. Phys. A 43, 185401 (2010). https://doi.org/10.1088/1751-8113/43/18/185401. arXiv:1002.0289

    Article  ADS  MathSciNet  MATH  Google Scholar 

  64. Joung, E., Taronna, M.: Cubic interactions of massless higher spins in (A)dS: metric-like approach. Nucl. Phys. B 861, 145 (2012). https://doi.org/10.1016/j.nuclphysb.2012.03.013. arXiv:1110.5918

    Article  ADS  MathSciNet  MATH  Google Scholar 

  65. Aragone, C., Deser, S.: Consistency Problems of Hypergravity. Phys. Lett. B 86, 161 (1979). https://doi.org/10.1016/0370-2693(79)90808-6

    Article  ADS  MathSciNet  Google Scholar 

  66. Barnich, G., Henneaux, M.: Consistent couplings between fields with a gauge freedom and deformations of the master equation. Phys. Lett. B 311, 123 (1993). https://doi.org/10.1016/0370-2693(93)90544-R. arXiv:hep-th/9304057

    Article  ADS  MathSciNet  Google Scholar 

  67. Boulanger, N., Leclercq, S., Cnockaert, S.: Parity violating vertices for spin-3 gauge fields. Phys. Rev. D 73, 065019 (2006). https://doi.org/10.1103/PhysRevD.73.065019. arXiv:hep-th/0509118

    Article  ADS  MathSciNet  Google Scholar 

  68. Boulanger, N., Ponomarev, D., Skvortsov, E.D., Taronna, M.: On the uniqueness of higher-spin symmetries in AdS and CFT. Int. J. Mod. Phys. A 28, 1350162 (2013). https://doi.org/10.1142/S0217751X13501625. arXiv:1305.5180

    Article  ADS  MathSciNet  MATH  Google Scholar 

  69. Conde, E., Joung, E. and Mkrtchyan, K. (2016) Spinor-Helicity Three-Point Amplitudes from Local Cubic Interactions. JHEP, 08:040. https://doi.org/10.1007/JHEP08(2016)040. arXiv: 1605.07402

  70. Kessel, P. and Mkrtchyan, K. (2018) Cubic interactions of massless bosonic fields in three dimensions II: Parity-odd and Chern-Simons vertices. Phys. Rev. D, 97:106021. https://doi.org/10.1103/PhysRevD.97.106021. arXiv:1803.02737

  71. Woodard, R.P. (2015) Ostrogradsky’s theorem on Hamiltonian instability. Scholarpedia 10:32243. https://doi.org/10.4249/scholarpedia.32243. arXiv:1506.02210

  72. Kaparulin, D.S., Lyakhovich, S.L., Sharapov, A.A.: Classical and quantum stability of higher-derivative dynamics. Eur. Phys. J. C 74, 3072 (2014). https://doi.org/10.1140/epjc/s10052-014-3072-3. arXiv:1407.8481

    Article  ADS  Google Scholar 

  73. Barnich, G., Brandt, F. and Henneaux, M. (1995) Local BRST cohomology in the antifield formalism. 1. General theorems. Commun. Math. Phys. 174:57. https://doi.org/10.1007/BF02099464. arXiv:hep-th/9405109

  74. Barnich, G., Brandt, F. and Henneaux, M. (1995) Local BRST cohomology in the antifield formalism. II. Application to Yang-Mills theory. Commun. Math. Phys., 174:93. https://doi.org/10.1007/BF02099465. arXiv:hep-th/9405194

  75. Lucena Gómez, G. (2015) The Elegance of Cohomological Methods. arXiv:1508.07226

  76. Manvelyan, R., Mkrtchyan, K., Ruhl, W.: General trilinear interaction for arbitrary even higher spin gauge fields. Nucl. Phys. B 836, 204 (2010). https://doi.org/10.1016/j.nuclphysb.2010.04.019. arXiv:1003.2877

    Article  ADS  MathSciNet  MATH  Google Scholar 

  77. Sagnotti, A., Taronna, M.: String Lessons for Higher-Spin Interactions. Nucl. Phys. B 842, 299 (2011). https://doi.org/10.1016/j.nuclphysb.2010.08.019. arXiv:1006.5242

    Article  ADS  MathSciNet  MATH  Google Scholar 

  78. Buchbinder, I.L., Fotopoulos, A., Petkou, A.C., Tsulaia, M.: Constructing the cubic interaction vertex of higher spin gauge fields. Phys. Rev. D 74, 105018 (2006). https://doi.org/10.1103/PhysRevD.74.105018. arXiv:hep-th/0609082

    Article  ADS  MathSciNet  Google Scholar 

  79. Fotopoulos, A., Tsulaia, M.: On the Tensionless Limit of String theory, Off - Shell Higher Spin Interaction Vertices and BCFW Recursion Relations. JHEP 11, 086 (2010). https://doi.org/10.1007/JHEP11(2010)086. arXiv:1009.0727

    Article  ADS  MathSciNet  MATH  Google Scholar 

  80. Francia, D., Monaco, G.L. and Mkrtchyan, K. (2017) Cubic interactions of Maxwell-like higher spins. JHEP, 04:068. https://doi.org/10.1007/JHEP04(2017)068. arXiv:1611.00292

  81. Boulanger, N., Skvortsov, E.D. and Zinoviev, Y.M. Gravitational cubic interactions for a simple mixed-symmetry gauge field in AdS and flat backgrounds. https://doi.org/10.1088/1751-8113/44/41/415403 J. Phys. A, 44:415403 arXiv:1107.1872

  82. Bekaert, X., Boulanger, N., Henneaux, M.: Consistent deformations of dual formulations of linearized gravity: A No go result. Phys. Rev. D 67, 044010 (2003). https://doi.org/10.1103/PhysRevD.67.044010. arXiv:hep-th/0210278

    Article  ADS  MathSciNet  Google Scholar 

  83. Bekaert, X., Boulanger, N., Cnockaert, S.: No self-interaction for two-column massless fields. J. Math. Phys. 46, 012303 (2005). https://doi.org/10.1063/1.1823032. arXiv:hep-th/0407102

    Article  ADS  MathSciNet  MATH  Google Scholar 

  84. Boulanger, N., Damour, T., Gualtieri, L., Henneaux, M.: Inconsistency of interacting, multigraviton theories. Nucl. Phys. B 597, 127 (2001). https://doi.org/10.1016/S0550-3213(00)00718-5. arXiv:hep-th/0007220

    Article  ADS  MATH  Google Scholar 

  85. Peskin, M.E., Schroeder, D.V.: An introduction to quantum field theory. Westview, Boulder CO (1995)

    Google Scholar 

  86. Schwartz, M.: Quantum Field Theory and the Standard Model. Cambridge University Press, Quantum Field Theory and the Standard Model (2014)

    Google Scholar 

  87. ’t Hooft, G., Veltman, M.J.G.: Diagrammar. NATO Sci. Ser. B 4, 177 (1974). https://doi.org/10.1007/978-1-4684-2826-1_5

    Article  Google Scholar 

  88. Ponomarev, D. and Tseytlin, A.A. (2016) On quantum corrections in higher-spin theory in flat space. JHEP, 05:184. https://doi.org/10.1007/JHEP05(2016)184. arXiv: 1603.06273

  89. Francia, D., Mourad, J., Sagnotti, A.: Current Exchanges and Unconstrained Higher Spins. Nucl. Phys. B 773, 203 (2007). https://doi.org/10.1016/j.nuclphysb.2007.03.021. arXiv:hep-th/0701163

    Article  ADS  MathSciNet  MATH  Google Scholar 

  90. Bekaert, X., Joung, E., Mourad, J.: On higher spin interactions with matter. JHEP 05, 126 (2009). https://doi.org/10.1088/1126-6708/2009/05/126. arXiv:0903.3338

    Article  ADS  Google Scholar 

  91. Elvang, H. and Huang, Y.-t. (2013) Scattering Amplitudes. arXiv:1308.1697

  92. Taronna, M. (2017) On the non-local obstruction to interacting higher spins in flat space. JHEP, 05:026. https://doi.org/10.1007/JHEP05(2017)026. arXiv: 1701.05772

  93. Roiban, R. and Tseytlin, A.A. (2017) On four-point interactions in massless higher spin theory in flat space. JHEP, 04:139. https://doi.org/10.1007/JHEP04(2017)139. arXiv:1701.05773

  94. Benincasa, P. and Cachazo, F. (2007) Consistency Conditions on the S-Matrix of Massless Particles. arXiv:0705.4305

  95. Benincasa, P. and Conde, E. (2012) Exploring the S-Matrix of Massless Particles. Phys. Rev. D, 86025007. https://doi.org/10.1103/PhysRevD.86.025007, arXiv:1108.3078

  96. Bekaert, X., Boulanger, N., Sundell, P.: How higher-spin gravity surpasses the spin two barrier: no-go theorems versus yes-go examples. Rev. Mod. Phys. 84, 987 (2012). https://doi.org/10.1103/RevModPhys.84.987. arXiv:1007.0435

    Article  ADS  Google Scholar 

  97. Weinberg, S.: Photons and gravitons in \(S\)-Matrix theory: derivation of charge conservation and equality of gravitational and inertial mass. Phys. Rev. 135, B1049 (1964). https://doi.org/10.1103/PhysRev.135.B1049

    Article  ADS  MathSciNet  MATH  Google Scholar 

  98. Campoleoni, A. and Pekar, S. (2022) Carrollian and Galilean conformal higher-spin algebras in any dimensions. JHEP, 02150. https://doi.org/10.1007/JHEP02(2022)150. arXiv:2110.07794

  99. Flato, M., Fronsdal, C.: On DIS and Racs. Phys. Lett. B 97, 236 (1980). https://doi.org/10.1016/0370-2693(80)90591-2

    Article  ADS  Google Scholar 

  100. Sezgin, E., Sundell, P.: Massless higher spins and holography. Nucl. Phys. B 644, 303 (2002). https://doi.org/10.1016/S0550-3213(02)00739-3. arXiv:hep-th/0205131

    Article  ADS  MathSciNet  MATH  Google Scholar 

  101. Klebanov, I.R., Polyakov, A.M.: AdS dual of the critical O(N) vector model. Phys. Lett. B 550, 213 (2002). https://doi.org/10.1016/S0370-2693(02)02980-5. arXiv:hep-th/0210114

    Article  ADS  MathSciNet  MATH  Google Scholar 

  102. Fradkin, E.S., Vasiliev, M.A.: Cubic Interaction in Extended Theories of Massless Higher Spin Fields. Nucl. Phys. B 291, 141 (1987). https://doi.org/10.1016/0550-3213(87)90469-X

    Article  ADS  MathSciNet  Google Scholar 

  103. Petersen, J.L.: Introduction to the Maldacena conjecture on AdS / CFT. Int. J. Mod. Phys. A 14, 3597 (1999). https://doi.org/10.1142/S0217751X99001676. arXiv:hep-th/9902131

    Article  ADS  MathSciNet  MATH  Google Scholar 

  104. Aharony, O., Gubser, S.S., Maldacena, J.M., Ooguri, H., Oz, Y.: Large N field theories, string theory and gravity. Phys. Rept. 323, 183 (2000). https://doi.org/10.1016/S0370-1573(99)00083-6. arXiv:hep-th/9905111

    Article  ADS  MathSciNet  MATH  Google Scholar 

  105. D’Hoker, E. and Freedman, D.Z. (2002) Supersymmetric gauge theories and the AdS / CFT correspondence, In: Theoretical Advanced Study Institute in Elementary Particle Physics (TASI 2001): Strings, Branes and EXTRA Dimensions. pp. 3–158, 1. arXiv:hep-th/0201253

  106. Nastase, H. (2007) Introduction to AdS-CFT. arXiv:0712.0689

  107. Osterwalder, K., Schrader, R.: Axioms for euclidean green’s functions. Commun. Math. Phys. 31, 83 (1973). https://doi.org/10.1007/BF01645738

    Article  ADS  MathSciNet  MATH  Google Scholar 

  108. Osterwalder, K. and Schrader, R. (1975) Axioms for Euclidean Green’s Functions. 2.. Commun. Math. Phys., 42:281. https://doi.org/10.1007/BF01608978

  109. Mack, G.: Osterwalder-schrader positivity in conformal invariant quantum field theory. Lect. Notes Phys. 37, 66 (1975). https://doi.org/10.1007/3-540-07160-1_3

    Article  ADS  Google Scholar 

  110. Hartman, T., Jain, S. and Kundu, S. (2016) Causality Constraints in Conformal Field Theory. JHEP, 05:099. https://doi.org/10.1007/JHEP05(2016)099. arXiv:1509.00014

  111. Rychkov, S. (2017) EPFL Lectures on Conformal Field Theory in D\(>\)= 3 Dimensions. SpringerBriefs in Physics, Springer, Cham. https://doi.org/10.1007/978-3-319-43626-510.1007/978-3-319-43626-5. arXiv:1601.05000

  112. Simmons-Duffin, D. (2017) The Conformal Bootstrap. In: Theoretical advanced study institute in elementary particle physics: new frontiers in fields and strings, pp. 1–74. https://doi.org/10.1142/9789813149441_0001. arXiv:1602.07982

  113. Craigie, N.S., Dobrev, V.K., Todorov, I.T.: Conformally covariant composite operators in quantum chromodynamics. Annals Phys. 159, 411 (1985). https://doi.org/10.1016/0003-4916(85)90118-6

    Article  ADS  Google Scholar 

  114. Anselmi, D.: Higher spin current multiplets in operator product expansions. Class. Quant. Grav. 17, 1383 (2000). https://doi.org/10.1088/0264-9381/17/6/305. arXiv:hep-th/9906167

    Article  ADS  MathSciNet  MATH  Google Scholar 

  115. Bekaert, X., Erdmenger, J., Ponomarev, D., Sleight, C.: Towards holographic higher-spin interactions: Four-point functions and higher-spin exchange. JHEP 03, 170 (2015). https://doi.org/10.1007/JHEP03(2015)170. arXiv:1412.0016

    Article  MathSciNet  MATH  Google Scholar 

  116. Sleight, C. and Taronna, M. (2016) Higher spin interactions from conformal field theory: the complete cubic couplings. Phys. Rev. Lett., 116:181602. https://doi.org/10.1103/PhysRevLett.116.181602. arXiv:1603.00022

  117. Bekaert, X., Erdmenger, J., Ponomarev, D. and Sleight, C. (2015) Quartic AdS interactions in higher-spin gravity from conformal field theory. JHEP, 11:149. https://doi.org/10.1007/JHEP11(2015)149. arXiv:1508.04292

  118. Taronna, M. (2017) Pseudo-local theories: a functional class proposal. In: International workshop on higher spin gauge theories. pp. 59–84. https://doi.org/10.1142/9789813144101_0006. arXiv:1602.08566

  119. Bekaert, X., Erdmenger, J., Ponomarev, D. and Sleight, C. (2017) Bulk quartic vertices from boundary four-point correlators. In: International workshop on higher spin gauge theories. pp. 291–303. https://doi.org/10.1142/9789813144101_0015. arXiv:1602.08570

  120. Sleight, C. and Taronna, M. (2018) Higher-spin gauge theories and bulk locality. Phys. Rev. Lett., 121:171604. https://doi.org/10.1103/PhysRevLett.121.171604. arXiv:1704.07859

  121. Ponomarev, D. (2018) A Note on (Non)-locality in holographic higher spin theories. Universe, 4:2. https://doi.org/10.3390/universe4010002. arXiv:1710.00403

  122. Maldacena, J., Zhiboedov, A.: Constraining Conformal Field Theories with A Higher Spin Symmetry. J. Phys. A 46, 214011 (2013). https://doi.org/10.1088/1751-8113/46/21/214011. arXiv:1112.1016

    Article  ADS  MathSciNet  MATH  Google Scholar 

  123. Alba, V. and Diab, K. (2013) Constraining conformal field theories with a higher spin symmetry in d=4. arXiv:1307.8092

  124. Maldacena, J., Zhiboedov, A.: Constraining conformal field theories with a slightly broken higher spin symmetry. Class. Quant. Grav. 30, 104003 (2013). https://doi.org/10.1088/0264-9381/30/10/104003. arXiv:1204.3882

    Article  ADS  MathSciNet  MATH  Google Scholar 

  125. de Mello Koch, R., Jevicki, A., Jin, K., Rodrigues, J.P.: \(AdS_4/CFT_3\) Construction from collective fields. Phys. Rev. D 83, 025006 (2011). https://doi.org/10.1103/PhysRevD.83.025006. arXiv:1008.0633

    Article  ADS  Google Scholar 

  126. Ortin, T. (2015) Gravity and Strings. Cambridge Monographs on Mathematical Physics, Cambridge University Press, 2nd ed. ed.. https://doi.org/10.1017/CBO9781139019750

  127. Vasiliev, M.A. (1980) ’Gauge’ form of description of massless fields with arbitrary spin. (in Russian). Yad. Fiz. 32:855

  128. Vasiliev, M.A.: Free massless fields of arbitrary spin in the de sitter space and initial data for a higher spin superalgebra. Fortsch. Phys. 35, 741 (1987)

    Article  ADS  MathSciNet  Google Scholar 

  129. Lopatin, V.E., Vasiliev, M.A.: Free Massless Bosonic Fields of Arbitrary Spin in \(d\)-dimensional De Sitter Space. Mod. Phys. Lett. A 3, 257 (1988). https://doi.org/10.1142/S0217732388000313

    Article  ADS  MathSciNet  Google Scholar 

  130. Campoleoni, A.: higher spins in D = 2 + 1. Subnucl. Ser. 49, 385 (2013). https://doi.org/10.1142/9789814522519_0020. arXiv:1110.5841

    Article  MATH  Google Scholar 

  131. Campoleoni, A., Fredenhagen, S., Pfenninger, S., Theisen, S.: Towards metric-like higher-spin gauge theories in three dimensions. J. Phys. A 46, 14017 (2013). https://doi.org/10.1088/1751-8113/46/21/214017. arXiv:1208.1851

    Article  ADS  MathSciNet  MATH  Google Scholar 

  132. MacDowell, S.W., Mansouri, F.: Unified geometric theory of gravity and supergravity. Phys. Rev. Lett. 38, 739 (1977). https://doi.org/10.1103/PhysRevLett.38.739

    Article  ADS  MathSciNet  Google Scholar 

  133. Stelle, K.S., West, P.C.: Spontaneously broken de sitter symmetry and the gravitational holonomy group. Phys. Rev. D 21, 1466 (1980). https://doi.org/10.1103/PhysRevD.21.1466

    Article  ADS  MathSciNet  Google Scholar 

  134. Skvortsov, E.D., Vasiliev, M.A.: Geometric formulation for partially massless fields. Nucl. Phys. B 756, 117 (2006). https://doi.org/10.1016/j.nuclphysb.2006.06.019. arXiv:hep-th/0601095

    Article  ADS  MathSciNet  MATH  Google Scholar 

  135. Zinoviev, Y.M.: Frame-like gauge invariant formulation for massive high spin particles. Nucl. Phys. B 808, 185 (2009). https://doi.org/10.1016/j.nuclphysb.2008.09.020. arXiv:0808.1778

    Article  ADS  MathSciNet  MATH  Google Scholar 

  136. Ponomarev, D.S., Vasiliev, M.A.: Frame-like action and unfolded formulation for massive higher-spin fields. Nucl. Phys. B 839, 466 (2010). https://doi.org/10.1016/j.nuclphysb.2010.06.007. arXiv:1001.0062

    Article  ADS  MathSciNet  MATH  Google Scholar 

  137. Khabarov, M.V. and Zinoviev, Y.M. (2018) Infinite (continuous) spin fields in the frame-like formalism. Nucl. Phys. B, 928:182. https://doi.org/10.1016/j.nuclphysb.2018.01.016. arXiv:1711.08223

  138. Skvortsov, E.D.: Frame-like Actions for Massless Mixed-Symmetry Fields in Minkowski space. Nucl. Phys. B 808, 569 (2009). https://doi.org/10.1016/j.nuclphysb.2008.09.007. arXiv:0807.0903

    Article  ADS  MathSciNet  MATH  Google Scholar 

  139. Zinoviev, Y.M.: Toward frame-like gauge invariant formulation for massive mixed symmetry bosonic fields. Nucl. Phys. B 812, 46 (2009). https://doi.org/10.1016/j.nuclphysb.2008.12.003. arXiv:0809.3287

    Article  ADS  MathSciNet  MATH  Google Scholar 

  140. Alkalaev, K.B., Shaynkman, O.V., Vasiliev, M.A.: On the frame - like formulation of mixed symmetry massless fields in (A)dS(d). Nucl. Phys. B 692, 363 (2004). https://doi.org/10.1016/j.nuclphysb.2004.05.031. arXiv:hep-th/0311164

    Article  ADS  MathSciNet  MATH  Google Scholar 

  141. Boulanger, N., Iazeolla, C. and Sundell, P. (2009) Unfolding mixed-symmetry fields in AdS and the BMV conjecture: i. general formalism. JHEP, 07:013. https://doi.org/10.1088/1126-6708/2009/07/013. arXiv:0812.3615

  142. Skvortsov, E.D.: Gauge fields in (A)dS(d) and Connections of its symmetry algebra. J. Phys. A 42, 385401 (2009). https://doi.org/10.1088/1751-8113/42/38/385401. arXiv:0904.2919

    Article  ADS  MathSciNet  MATH  Google Scholar 

  143. Alkalaev, K.: FV-type action for \(AdS_5\) mixed-symmetry fields. JHEP 03, 031 (2011). https://doi.org/10.1007/JHEP03(2011)031. arXiv:1011.6109

    Article  ADS  MathSciNet  MATH  Google Scholar 

  144. Vasiliev, M.A.: Cubic vertices for symmetric higher-spin gauge fields in \((A)dS_d\). Nucl. Phys. B 862, 341 (2012). https://doi.org/10.1016/j.nuclphysb.2012.04.012. arXiv:1108.5921

    Article  ADS  MATH  Google Scholar 

  145. Boulanger, N., Skvortsov, E.D.: Higher-spin algebras and cubic interactions for simple mixed-symmetry fields in AdS spacetime. JHEP 09, 063 (2011). https://doi.org/10.1007/JHEP09(2011)063. arXiv:1107.5028

    Article  ADS  MathSciNet  MATH  Google Scholar 

  146. Boulanger, N., Ponomarev, D., Skvortsov, E.D.: Non-abelian cubic vertices for higher-spin fields in anti-de Sitter space. JHEP 05, 008 (2013). https://doi.org/10.1007/JHEP05(2013)008. arXiv:1211.6979

    Article  ADS  MATH  Google Scholar 

  147. Zinoviev, Y.M. (2014) Massive spin-2 in the Fradkin–Vasiliev formalism. I. Partially massless case. Nucl. Phys. B, 886:712 https://doi.org/10.1016/j.nuclphysb.2014.07.013.arXiv:1405.4065

  148. Blencowe, M.P.: A Consistent Interacting Massless Higher Spin Field Theory in \(D\) = (2+1). Class. Quant. Grav. 6, 443 (1989). https://doi.org/10.1088/0264-9381/6/4/005

    Article  ADS  MathSciNet  Google Scholar 

  149. Henneaux, M., Rey, S.-J.: Nonlinear \(W_{infinity}\) as asymptotic symmetry of three-dimensional higher spin anti-de sitter gravity. JHEP 12, 007 (2010). https://doi.org/10.1007/JHEP12(2010)007. arXiv:1008.4579

    Article  ADS  MATH  Google Scholar 

  150. Campoleoni, A., Fredenhagen, S., Pfenninger, S., Theisen, S.: Asymptotic symmetries of three-dimensional gravity coupled to higher-spin fields. JHEP 11, 007 (2010). https://doi.org/10.1007/JHEP11(2010)007. arXiv:1008.4744

    Article  ADS  MathSciNet  MATH  Google Scholar 

  151. Prokushkin, S.F., Vasiliev, M.A.: spin gauge interactions for massive matter fields in 3-D AdS space-time. Nucl. Phys. B 545, 385 (1999). https://doi.org/10.1016/S0550-3213(98)00839-6. arXiv:hep-th/9806236

    Article  ADS  MathSciNet  MATH  Google Scholar 

  152. Gaberdiel, M.R., Gopakumar, R.: An AdS\(_{3}\) dual for minimal model CFTs. Phys. Rev. D 83, 066007 (2011). https://doi.org/10.1103/PhysRevD.83.066007. arXiv:1011.2986

    Article  ADS  Google Scholar 

  153. Gaberdiel, M.R., Gopakumar, R.: Minimal model holography. J. Phys. A 46, 214002 (2013). https://doi.org/10.1088/1751-8113/46/21/214002. arXiv:1207.6697

    Article  ADS  MathSciNet  MATH  Google Scholar 

  154. Ponomarev, D. (2016) Off-shell spinor-helicity amplitudes from light-cone deformation procedure. JHEP, 12:117. https://doi.org/10.1007/JHEP12(2016)117. arXiv:1611.00361

  155. Zwiebach, B.: A first course in string theory, 1st edn. Cambridge Univ. Press, Cambridge (2014)

    MATH  Google Scholar 

  156. Perry, R.J. (1994) Hamiltonian light front field theory and quantum chromodynamics. In: Hadrons 94 Workshop. 7. arXiv:hep-th/9407056

  157. Burkardt, M.: Light front quantization. Adv. Nucl. Phys. 23, 1 (1996). https://doi.org/10.1007/0-306-47067-5_1. arXiv:hep-ph/9505259

    Article  Google Scholar 

  158. Ligterink, N. (1996) Light-front hamiltonian field theory. Covariance and renormalization., Ph.D. thesis, VU, Amsterdam

  159. Harindranath, A. (1996) An introduction to light front dynamics for pedestrians. In: International school on light-front quantization and non-perturbative QCD (To be followed by the Workshop 3-14 Jun 1996). vol. 5. arXiv:hep-ph/9612244

  160. Heinzl, T.: Light cone quantization: foundations and applications. Lect. Notes Phys. 572, 55 (2001). https://doi.org/10.1007/3-540-45114-5_2. arXiv:hep-th/0008096

    Article  ADS  MATH  Google Scholar 

  161. Mannheim, P.D., Lowdon, P. and Brodsky, S.J. (2021) Comparing light-front quantization with instant-time quantization. Phys. Rept., 891:1 https://doi.org/10.1016/j.physrep.2020.09.001. arXiv:2005.00109

  162. Dirac, P.A.M.: Forms of relativistic dynamics. Rev. Mod. Phys. 21, 392 (1949). https://doi.org/10.1103/RevModPhys.21.392

    Article  ADS  MathSciNet  MATH  Google Scholar 

  163. Ponomarev, D. and Skvortsov, E.D. (2017) Light-front higher-spin theories in flat space. J. Phys. A, 50:095401. https://doi.org/10.1088/1751-8121/aa56e7. arXiv:1609.04655

  164. Chalmers, G., Siegel, W.: The Selfdual sector of QCD amplitudes. Phys. Rev. D 54, 7628 (1996). https://doi.org/10.1103/PhysRevD.54.7628. arXiv:hep-th/9606061

    Article  ADS  Google Scholar 

  165. Ponomarev, D. (2017) Chiral Higher Spin Theories and Self-Duality. JHEP, 12:141. https://doi.org/10.1007/JHEP12(2017)141. arXiv:1710.00270

  166. Skvortsov, E.D., Tran, T. and Tsulaia, M. (2018) Quantum chiral higher spin gravity. Phys. Rev. Lett., 121:031601. https://doi.org/10.1103/PhysRevLett.121.031601. arXiv:1805.00048

  167. Krasnov, K., Skvortsov, E. and Tran, T. (2021) Actions for self-dual higher spin gravities. JHEP 08:076. https://doi.org/10.1007/JHEP08(2021)076. arXiv:2105.12782

  168. Bengtsson, A.K.H., Bengtsson, I., Brink, L.: Cubic interaction terms for arbitrary spin. Nucl. Phys. B 227, 31 (1983). https://doi.org/10.1016/0550-3213(83)90140-2

    Article  ADS  Google Scholar 

  169. Bengtsson, A.K.H., Bengtsson, I., Linden, N.: Interacting higher spin gauge fields on the light front. Class. Quant. Grav. 4, 1333 (1987). https://doi.org/10.1088/0264-9381/4/5/028

    Article  ADS  MathSciNet  Google Scholar 

  170. Metsaev, R.R.: Poincare invariant dynamics of massless higher spins: Fourth order analysis on mass shell. Mod. Phys. Lett. A 6, 359 (1991). https://doi.org/10.1142/S0217732391000348

    Article  ADS  MathSciNet  MATH  Google Scholar 

  171. Metsaev, R.R. (1991) S matrix approach to massless higher spins theory. 2: the case of internal symmetry. Mod. Phys. Lett. A, 6:2411. https://doi.org/10.1142/S0217732391002839

  172. Metsaev, R.R.: Cubic interaction vertices of massive and massless higher spin fields. Nucl. Phys. B 759, 147 (2006). https://doi.org/10.1016/j.nuclphysb.2006.10.002. arXiv:hep-th/0512342

    Article  ADS  MathSciNet  MATH  Google Scholar 

  173. Metsaev, R.R.: Cubic interaction vertices for fermionic and bosonic arbitrary spin fields. Nucl. Phys. B 859, 13 (2012). https://doi.org/10.1016/j.nuclphysb.2012.01.022. arXiv:0712.3526

    Article  ADS  MathSciNet  MATH  Google Scholar 

  174. Metsaev, R.R. (2018) Light-cone gauge cubic interaction vertices for massless fields in AdS(4). Nucl. Phys. B, 936:320. https://doi.org/10.1016/j.nuclphysb.2018.09.021arXiv:1807.07542

  175. Skvortsov, E. (2019) Light-front bootstrap for chern-simons matter theories, 06:058 https://doi.org/10.1007/JHEP06(2019)058 JHEP, arXiv:1811.12333

  176. Bengtsson, A.K.H. (2014) A Riccati type PDE for light-front higher helicity vertices. https://doi.org/10.1007/JHEP09(2014)105 JHEP, 09:105. arXiv:1403.7345

  177. Sleight, C. and Taronna, M. (2017) Higher-spin algebras, holography and flat space. JHEP, 02:095. https://doi.org/10.1007/JHEP02(2017)095. arXiv:1609.00991

  178. Ananth, S.: Spinor helicity structures in higher spin theories. JHEP 11, 089 (2012). https://doi.org/10.1007/JHEP11(2012)089. arXiv:1209.4960

  179. Vasiliev, M.A. (1999) Higher spin gauge theories: Star product and AdS space. arXiv:hep-th/9910096

  180. Bekaert, X., Cnockaert, S., Iazeolla, C. and Vasiliev, M.A. (2004) Nonlinear higher spin theories in various dimensions, in 1st solvay workshop on higher spin gauge theories. pp. 132–197. arXiv:hep-th/0503128

  181. Giombi, S., Yin, X.: Higher spin gauge theory and holography: the three-point functions. JHEP 09, 115 (2010). https://doi.org/10.1007/JHEP09(2010)115. arXiv:0912.3462

    Article  ADS  MathSciNet  MATH  Google Scholar 

  182. Giombi, S., Yin, X.: Higher Spins in AdS and twistorial holography. JHEP 04, 086 (2011). https://doi.org/10.1007/JHEP04(2011)086. arXiv:1004.3736

    Article  ADS  MathSciNet  MATH  Google Scholar 

  183. Giombi, S., Yin, X.: The higher spin/vector model duality. J. Phys. A 46, 214003 (2013). https://doi.org/10.1088/1751-8113/46/21/214003. arXiv:1208.4036

    Article  ADS  MathSciNet  MATH  Google Scholar 

  184. Boulanger, N., Kessel, P., Skvortsov, E.D. and Taronna, M. (2016) Higher spin interactions in four-dimensions: vasiliev versus fronsdal. J. Phys. A, 49:095402. https://doi.org/10.1088/1751-8113/49/9/095402. arXiv:1508.04139

  185. Skvortsov, E.D. and Taronna, M. (2015) On locality, holography and unfolding. JHEP, 11:044. https://doi.org/10.1007/JHEP11(2015)044. arXiv:1508.04764

  186. Didenko, V.E. and Vasiliev, M.A. (2017) Test of the local form of higher-spin equations via AdS / CFT. Phys. Lett. B, 775:352. https://doi.org/10.1016/j.physletb.2017.09.091. arXiv:1705.03440

  187. Didenko, V.E., Gelfond, O.A., Korybut, A.V. and Vasiliev, M.A. (2018) Homotopy properties and lower-order vertices in higher-spin equations. J. Phys. A, 51:465202. https://doi.org/10.1088/1751-8121/aae5e1. arXiv:1807.00001

  188. Joung, E., Nakach, S. and Tseytlin, A.A. (2016) Scalar scattering via conformal higher spin exchange. JHEP, 02125. https://doi.org/10.1007/JHEP02(2016)125. arXiv:1512.08896

  189. Beccaria, M., Nakach, S. and Tseytlin, A.A. (2016) On triviality of S-matrix in conformal higher spin theory. JHEP, 09:034. https://doi.org/10.1007/JHEP09(2016)034. arXiv:1607.06379

  190. Segal, A.Y. (2003) Conformal higher spin theory. https://doi.org/10.1016/S0550-3213(03)00368-7 Nucl. Phys. B 664:59 arXiv:hep-th/0207212

  191. Hähnel, P. and McLoughlin, T. (2017) Conformal higher spin theory and twistor space actions. J. Phys. A, 50:485401. https://doi.org/10.1088/1751-8121/aa9108. arXiv:1604.08209

  192. Adamo, T., Hähnel P. and McLoughlin, T. (2017) Conformal higher spin scattering amplitudes from twistor space. JHEP, 04:021. https://doi.org/10.1007/JHEP04(2017)021. arXiv:1611.06200

  193. Sperling, M. and Steinacker, H.C. (2017) Covariant 4-dimensional fuzzy spheres, matrix models and higher spin. J. Phys. A 50:375202. https://doi.org/10.1088/1751-8121/aa8295. arXiv:1911.03162

  194. Steinacker, H.C. (2020) On the quantum structure of space-time, gravity, and higher spin in matrix models. Class. Quant. Grav., 37:113001. https://doi.org/10.1088/1361-6382/ab857f. arXiv:1911.03162

  195. Coleman, S.R., Mandula, J.: All Possible Symmetries of the S Matrix. Phys. Rev. 159, 1251 (1967). https://doi.org/10.1103/PhysRev.159.1251

    Article  ADS  MATH  Google Scholar 

Download references

Acknowledgements

I would like to thank the students participating in the course for their feedback, which helped me improving these lectures. I would also like to thank K. Alkalaev, V. Didenko, M. Grigoriev, S. Pekar and, especially, E. Skvortsov, as well as all participants of the higher-spin course at UMONS – A. Bedhotiya, S. Dhasmana, J. O’Connor, M. Serrani, R. Van Dongen, – for their valuable comments on the preliminary versions of the text.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dmitry Ponomarev.

Additional information

Publisher's Note

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

Appendices

A Conventions

We use the mostly plus convention \(\eta = \textrm{diag}(-,+,\dots ,+)\).

To deal with symmetric tensors we use the following notation

$$\begin{aligned} \varphi ^{a(s)}=\varphi ^{a_1\dots a_s}. \end{aligned}$$
(A.1)

Moreover, we will use identical indices to indicate that indices have to be symmetrized. Symmetrization is defined with the normalization that makes it a projector. For example,

$$\begin{aligned} \partial ^a\varphi ^{a(s)}=\frac{1}{s+1}\left( \partial ^{a_1}\varphi ^{a_2\dots a_{s+1}}+\dots + \partial ^{a_{s+1}}\varphi ^{a_1\dots a_{s}}\right) , \end{aligned}$$
(A.2)

where the r.h.s. contains \(s+1\) non-trivial permutations of indices in the expression on the l.h.s. Alternatively, one can sum over all \((s+1)!\) permutations of indices and then the overall factor in front of the bracket would be \([(s+1)!]^{-1}\).

Levi-Civita Tensor

We will introduce the Levi-Civita tensor with lower indices so that

$$\begin{aligned} \epsilon _{1 \dots d}= 1. \end{aligned}$$
(A.3)

By raising indices we find the Levi-Civita tensor with upper indices

$$\begin{aligned} \epsilon ^{a_1 \dots a_d} = \eta ^{a_1b_1}\dots \eta ^{a_d b_d} \epsilon _{b_1\dots b_d}. \end{aligned}$$
(A.4)

It should be remarked that

$$\begin{aligned} \epsilon ^{1 \dots d} =\sigma , \qquad \sigma \equiv \textrm{det}[\eta ], \end{aligned}$$
(A.5)

so for the Minkowski metric \(\sigma =-1\).

There are a couple of useful formulas that involve the Levi-Civita tensor

$$\begin{aligned} A_{a_1}{}^{b_1}\dots A_{a_d}{}^{b_d} \epsilon _{b_1\dots b_d}=\textrm{det}[A]\epsilon _{a_1\dots a_d} \end{aligned}$$
(A.6)

and

$$\begin{aligned} \epsilon ^{k[n]l[d-n]}\epsilon _{k[n]m[d-n]}=\sigma n!(d-n)!\delta ^{[l_1}_{m_1}\dots \delta ^{l_{d-n}]}_{m_{d-n}}, \end{aligned}$$
(A.7)

where, as usual, antisymmetrization is understood as a projection.

B The Coleman-Mandual Theorem

In this appendix we sketch the proof of the Coleman-Mandula theorem stated in Section 10.2. For a more rigorous discussion we refer the reader to the original paper [195] and to [7] for shortcuts and clarifications. Here we mostly follow [7].

B.1 Step 1: Generators that Commute with Translations

The proof goes in several steps. We will start by considering only those symmetry transformations \(B_\alpha \) that commute with the translation generator

$$\begin{aligned}{}[B_\alpha ,P_\mu ]=0. \end{aligned}$$
(B.1)

It is convenient to choose the basis for the states in the theory so that they have definite momenta. Then, the action of B on single-particle states is of the formFootnote 50

$$\begin{aligned} B_{\alpha }|p\rangle ^m = (b_\alpha (p))^m{}_{m'} |p\rangle ^{m'}, \end{aligned}$$
(B.2)

where \(m, m'\) label the states with given momentum p, which also includes spin labels. Note that according to our assumptions – more precisely, the assumption of having finitely many mass shells with the mass below any fixed value – the space of states with fixed p is finite-dimensional.

The argument then works differently for \(B_{\alpha }\) that map into the pure trace \(b_\alpha \) – that is those proportional to the unit matrix \(\delta ^m{}_{m'}\) – and for those that map into traceless \(b_{\alpha }\). We will first consider the pure trace part of b.

By one of our assumptions, B act on two-particle states by acting on each single-particle state separately. Together with (B.2), this gives

$$\begin{aligned} (b_\alpha (p,q))^m{}_{m',}{}^{n}{}_{n'}= (b_\alpha (p))^m{}_{m'} \delta ^n{}_{n'}+(b_\alpha (q))^n{}_{n'} \delta ^m{}_{m'}. \end{aligned}$$
(B.3)

Next, we consider a two-particle scattering process of particles with momenta p and q into \(p'\) and \(q'\),

$$\begin{aligned} S(p,m;q,n \rightarrow p',m';q',n')\equiv \delta ^d(p'+q'-p-q) S(p',q';p,q)^{m'n'}{}_{mn}, \end{aligned}$$
(B.4)

where \(p^2 = (p')^2\) and \(q^2 = (q')^2\). Invariance of the S-matrix with respect to B implies

$$\begin{aligned} b_\alpha (p',q')S(p',q';p,q)=S(p',q';p,q)b_{\alpha }(p,q). \end{aligned}$$
(B.5)

Assuming that S is non-vanishing, we multiply both sides of (B.5) with \(S^{-1}\) and take the trace. Due to the cyclicity of the trace, we find

$$\begin{aligned} \textrm{tr}\, b_\alpha (p',q') = \textrm{tr}\, b_\alpha (p,q). \end{aligned}$$
(B.6)

The trace of the tensor product of two matrices is the product of their traces. Together with (B.3) and (B.6) this leads to

$$\begin{aligned} N(q^2)\textrm{tr}\, b_{\alpha }(p')+N(p^2)\textrm{tr}\, b_{\alpha }(q')=N(q^2)\textrm{tr}\, b_{\alpha }(p)+N(p^2)\textrm{tr}\, b_{\alpha }(q), \end{aligned}$$
(B.7)

where \(N(q^2)\) and \(N(p^2)\) result from taking the traces of the Kronecker delta’s and count the numbers of states on each mass shell. Equivalently, we have

$$\begin{aligned} \frac{\textrm{tr}\, b_{\alpha }(p')}{N(p^2)}+\frac{\textrm{tr}\, b_{\alpha }(q')}{N(q^2)}=\frac{\textrm{tr}\, b_{\alpha }(p)}{N(p^2)}+\frac{\textrm{tr}\, b_{\alpha }(q)}{N(q^2)}. \end{aligned}$$
(B.8)

Its general solution reads

$$\begin{aligned} \frac{\textrm{tr}\, b_\alpha (p)}{N(p^2)} = c^\mu _{\alpha } p_\mu +d_\alpha , \end{aligned}$$
(B.9)

where both c and d are p-independent. Indeed, the d-part on both sides of (B.8) cancels out identically, while the c-part cancels out due to momentum conservation. We, therefore, find that the pure trace part of \(b_\alpha \) is either proportional to momenta or is an internal symmetry.

We now proceed with \(B_{\alpha }\) that map into traceless \(b_\alpha \), which we will denote

$$\begin{aligned} (b^{\sharp }_\alpha )^{n'}{}_n\equiv (b_\alpha )^{n'}{}_n - \frac{\textrm{tr}\, b_\alpha (p)}{N(p^2)}\delta ^{n'}{}_n. \end{aligned}$$
(B.10)

From (B.9) it follows that the generators represented by \(b^{\sharp }_\alpha \) are

$$\begin{aligned} B^{\sharp }_{\alpha } \equiv B_\alpha - c^\mu _{\alpha } P_\mu -d_\alpha . \end{aligned}$$
(B.11)

On the two-particle states these act by

$$\begin{aligned} (b^{\sharp }_\alpha (p,q))^m{}_{m',}{}^{n}{}_{n'}= (b^{\sharp }_\alpha (p))^m{}_{m'} \delta ^n{}_{n'}+(b^{\sharp }_\alpha (q))^n{}_{n'} \delta ^m{}_{m'}. \end{aligned}$$
(B.12)

Our next goal is to show that \(B_\alpha ^{\sharp } \rightarrow b^{\sharp }_\alpha (p,q)\) is an isomorphism of Lie algebras. To this end, we need to show that

$$\begin{aligned} l^\alpha b^{\sharp }_\alpha (p,q)=0 \quad \Rightarrow \quad l^\alpha B_\alpha ^{\sharp }=0, \end{aligned}$$
(B.13)

where \(l^\alpha \) are some coefficients. Equation (B.13) implies that \(B_\alpha ^{\sharp } \rightarrow b^{\sharp }_\alpha (p,q)\) is invertible, so it is an isomorphism.

Since \(B^{\sharp }\) are symmetries of the S-matrix, we have

$$\begin{aligned} b^{\sharp }_\alpha (p',q')S(p',q';p,q)=S(p',q';p,q)b^{\sharp }_{\alpha }(p,q). \end{aligned}$$
(B.14)

Again, we assume that the S-matrix is non-vanishing. Then, (B.14) implies that \(b^{\sharp }_\alpha (p',q')\) and \(b^{\sharp }_\alpha (p,q)\) are related by the similarity transformation. This, in turn, implies that

$$\begin{aligned} l^\alpha b^{\sharp }_{\alpha }(p,q) = 0 \qquad \Rightarrow \qquad l^\alpha b^{\sharp }_{\alpha }(p',q') = 0. \end{aligned}$$
(B.15)

Considering (B.12) and that \(b^{\sharp }_\alpha (p)\) are traceless, one finds

$$\begin{aligned} l^\alpha b^{\sharp }_{\alpha }(p',q') = 0 \qquad \Rightarrow \qquad l^\alpha b^{\sharp }_{\alpha }(p') = 0. \end{aligned}$$
(B.16)

We, thus, have shown that \( l^\alpha b^{\sharp }_{\alpha }(p,q) = 0\) implies \( l^\alpha b^{\sharp }_{\alpha }(p') = 0\), where \(p'\) is constrained to be on the same mass shell as p and, moreover, there should exist \(q'\) on the same mass shell with q, so that \(p+q=p'+q'\). With some extra work, this limitation on \(p'\) can be lifted, that is one can prove that

$$\begin{aligned} l^\alpha b^{\sharp }_{\alpha }(p,q) = 0 \qquad \Rightarrow \qquad l^\alpha b^{\sharp }_{\alpha }(k) = 0, \end{aligned}$$
(B.17)

where k is an arbitrary on-shell momentum. The fact that \( l^\alpha b^{\sharp }_{\alpha }(k)\) vanishes for any k means that \( l^\alpha B^{\sharp }_{\alpha }=0\). Thus, we managed to show (B.13), which implies that the Lie algebra generated by \(B^{\sharp }_\alpha \) is isomorphic to its representation on two-particle states with the trace part removed, \( b^{\sharp }_{\alpha }(p,q)\).

Next, we apply the standard theorem, see e.g. [7], which tells us that a Lie algebra of finite-dimensional Hermitian matrices – like \( b^{\sharp }_{\alpha }(p,q)\) for fixed p and q – is at most the direct sum of a semi-simple Lie algebra and some number of U(1) Lie algebras. We will now explore the consequence of this theorem focusing on the semi-simple part. The associated symmetry generators will be denoted \(B^{\flat }_\alpha \).

The Lorentz group acts on these generators in the standard wayFootnote 51

$$\begin{aligned} G(\Lambda ,0): \qquad B^{\flat }_{\alpha } \quad \rightarrow \quad B^{\flat }_{\alpha }(\Lambda )\equiv U(G(\Lambda ,0))B^{\flat }_{\alpha }U^{-1}(G(\Lambda ,0)). \end{aligned}$$
(B.18)

Since \(B^{\flat }_{\alpha }\) commute with \(P_\mu \) it follows that \(B^{\flat }_{\alpha }(\Lambda )\) commute with \(\Lambda _\mu {}^\nu P_\nu \). Considering that \(\Lambda _\mu {}^\nu P_\nu \) is just a linear combination of translations, we conclude that \(B^{\flat }_{\alpha }(\Lambda )\) commutes with \(P_\mu \). This, in turn, entails that \(B^{\flat }_{\alpha }(\Lambda )\) is some linear combination of \(B^{\flat }_{\alpha }\)

$$\begin{aligned} U(G(\Lambda ,0))B^{\flat }_{\alpha }U^{-1}(G(\Lambda ,0)) = D^{\beta }{}_{\alpha }(\Lambda )B^{\flat }_{\beta }. \end{aligned}$$
(B.19)

Therefore, \(B^{\flat }_{\alpha }\) realise a representation of the Lorentz group. We would like to show that it is unitary.

To do that one notices that \(B^{\flat }_{\alpha }(\Lambda )\) commute the same way as \(B^{\flat }_{\alpha }\)U and \(U^{-1}\) factors cancel out. This means that the structure constants of the algebra generated by \(B^{\flat }_{\alpha }\) are invariant under Lorentz transformations, that is

$$\begin{aligned} f^{\gamma }{}_{\alpha \beta } = D^{\alpha '}{}_\alpha (\Lambda )D^{\beta '}{}_\beta (\Lambda )D^{\gamma }{}_{\gamma '}(\Lambda )f^{\gamma '}{}_{\alpha '\beta '}. \end{aligned}$$
(B.20)

As a consequence, the Lie algebra metric

$$\begin{aligned} g_{\beta \delta }\equiv f^{\gamma }{}_{\alpha \beta }f^{\alpha }{}_{\gamma \delta } \end{aligned}$$
(B.21)

is also Lorentz invariant. Moreover, since the Lie algebra generated by \(B^{\flat }_{\alpha }\) is semi-simple, metric (B.21) is positive-definite. Altogether, this implies that \(B^{\flat }_{\alpha }\) realise a unitary finite-dimensional representation of the Lorentz group. As we mentioned in Section 2, for finite-dimensional representations, this is only possible if the representation carried by \(B^{\flat }_{\alpha }\) is trivial. Thus, \(B^{\flat }_{\alpha }\) generate internal symmetries.

With some extra arguments, one can show that the U(1) part of \(B^{\sharp }_{\alpha }\) also commutes with the Lorentz algebra.

Summarising the results of the first part of the proof, we found that symmetries of the S-matrix that commute with momenta are either momenta – the first term on the right hand side of (B.9) – or internal symmetries – the second term in (B.9) and all generators \(B^{\sharp }_{\alpha }\).

B.2 Step 2: Locality in Momentum Space

On the next step, we take up the possibility of symmetry generators that do not commute with translations. In general, the symmetry generator in the momentum basis reads

$$\begin{aligned} A_\alpha |p\rangle ^n =\int d^dp' \mathcal{A}_\alpha (p',p)^n{}_{n'}|p'\rangle ^{n'}. \end{aligned}$$
(B.22)

Since the kernel \(\mathcal{A}\) maps physical states to physical states, it should vanish unless both p and \(p'\) are on the mass shell. Our goal is to show that \(\mathcal{A}\) vanishes for any \(p\ne p'\).

To achieve this, one considers a generator

$$\begin{aligned} A_\alpha ^f = \int d^dx e^{iPx}A_\alpha e^{-iPx}f(x), \end{aligned}$$
(B.23)

where f is an arbitrary function. It is a symmetry generator as it is defined via a composition of symmetry generators P and \(A_\alpha \). It is straightforward to show that \(A^f\) acts on the single-particle states as

$$\begin{aligned} A^f_\alpha |p\rangle ^n =\int d^dp' \tilde{f}(p'-p)\mathcal{A}_\alpha (p',p)^n{}_{n'}|p'\rangle ^{n'}, \end{aligned}$$
(B.24)

where \(\tilde{f}\) is the Fourier transform of f

$$\begin{aligned} \tilde{f}(p)\equiv \int d^d xe^{ixp}f(x). \end{aligned}$$
(B.25)

Next, we return to the analysis of the 2-to-2 scattering, \(p+q=p'+q'\). Let \(\Delta \) be such that \(p+\Delta \) is still on-shell, while all \(q+\Delta \), \(p'+\Delta \) and \(q'+\Delta \) are off-shell. Then, picking \(\tilde{f}\) in (B.24) with the support in the vicinity of \(\Delta \), we find that

$$\begin{aligned} A_\alpha ^f|q\rangle ^m=0 ,\qquad A_\alpha ^f|p'\rangle ^{n'}=0,\qquad A_\alpha ^f|q'\rangle ^{m'}=0. \end{aligned}$$
(B.26)

Indeed, outside the support of \(\tilde{f}\), the \(\tilde{f}\) factor vanishes in (B.24), while inside the support of \(\tilde{f}\), the \(\mathcal{A}\) factor vanishes, as \(\mathcal{A}\) only relates physical states. The condition of invariance of the S-matrix with respect to \(A^f\) reads

$$\begin{aligned} \langle p',q'| S |A^f p,q\rangle + \langle p',q'| S | p,A^f q\rangle = \langle A^f p',q'| S | p,q\rangle + \langle p',A^f q'| S | p, q\rangle , \end{aligned}$$
(B.27)

where we again used our assumption about the action of symmetries on multi-particle states. Due to (B.26) this reduces to

$$\begin{aligned} \langle p',q'| S |A^f p,q\rangle =0. \end{aligned}$$
(B.28)

This can happen for two reasons: either S or \(A^f\) is vanishing. By invoking some geometric considerations, it is not hard to see, that one can change on-shell p, q, \(p'\) and \(q'\), so that momentum conservation is still satisfied, moreover, \(p+\Delta \) remains on-shell, while \(q+\Delta \), \(p'+\Delta \) and \(q'+\Delta \) remain off-shell. In other words, the argument presented above holds in a certain continuous range of the Mandelstam variables. Keeping in mind our assumption that the S-matrix can only have isolated zeros, we conclude that (B.28) entails \(A^f=0\). This, in turn, implies that

$$\begin{aligned} A(p',p)=0,\qquad \text {for} \qquad p-p'=\Delta . \end{aligned}$$
(B.29)

The same argument can be applied to other \(\Delta \) that shift an on-shell momentum p to an on-shell momentum. Typically, one can choose the remaining three momenta so that the momentum conservation is still satisfied, while after a shift by \(\Delta \) they all go off-shell. The value of \(\Delta \) which is excluded by these arguments is \(\Delta =0\) as, clearly, all on-shell states remain on-shell. Accordingly, we find that A is supported only on \(p=p'\) or

$$\begin{aligned} A(p',p)=0, \qquad \forall \quad p'\ne p. \end{aligned}$$
(B.30)

B.3 Step 3: Constraining Derivatives in Momenta

With some technical assumptions the fact that the integral kernel \(A(p,p')\) is only supported on \(p=p'\) implies that it is given by \(\delta ^d(p-p')\) and its derivatives of finite order. In other words, the action of A on one particle states is given by

$$\begin{aligned} A_\alpha |p\rangle ^n = \sum _{i=0}^k A_\alpha ^{(i)|\mu _1\dots \mu _i |n}{}_{n'}(p)\frac{\partial }{\partial p^{\mu _1}}\dots \frac{\partial }{\partial p^{\mu _i}}|p\rangle ^{n'}. \end{aligned}$$
(B.31)

The last step of the proof is to reduce the analysis of symmetries of this type to those that commute with momenta, discussed in the first part of the proof.

To achieve this one considers a k-fold commutator of (B.31) with momenta, which is also a symmetry of the S-matrix

$$\begin{aligned}{}[P^{\mu _1},[P^{\mu _2}, \dots [P^{\mu _k},A_\alpha ]\dots ]]= (-1)^k A_\alpha ^{(k)|\mu _1\dots \mu _k}(p), \end{aligned}$$
(B.32)

where A is symmetric in \(\mu \) indices. It no longer contains derivatives of momenta, hence, it commutes with the translation generators. Therefore, the results of the first step of the proof can be applied and we have

$$\begin{aligned} A_\alpha ^{(k)|\mu _1\dots \mu _k|n}{}_{n'}(p) = b_\alpha ^{\mu _1\dots \mu _k|n}{}_{n'}+c_{\alpha }^{\mu | \mu _1\dots \mu _k}p_\mu \delta ^n{}_{n'}. \end{aligned}$$
(B.33)

Here c is just the pure trace c term from (B.9), while b combines p-independent internal traceless and pure-trace symmetries. Note that we dropped \(N(p^2)\) from (B.9) for the pure trace part. This can be done for the following reason. First, by our assumption, there are finitely many mass shells in the system, so \(p^2\) takes discrete eigenvalues. Moreover, as we showed, A acts locally in momentum space. Altogether, this implies that A acts within a single mass shell, so \(N(p^2)\) can be replaced with a number \(N(-m^2)\).

We will first focus on the case with \(m^2\ne 0\) and take into account that A may only act within a single mass shell. In general, invariance of the mass shell \(p^2+m^2=0\) with respect to transformation O implies

$$\begin{aligned} (P^2+m^2)O = O'(P^2+m^2), \end{aligned}$$
(B.34)

where \(O'\) is an arbitrary operator. Condition (B.34) can be rewritten as

$$\begin{aligned}{}[P^2,O] = (O'-O)(P^2+m^2). \end{aligned}$$
(B.35)

We would like to apply this conclusion to O defined by

$$\begin{aligned} O=[P^{\mu _2},[P^{\mu _3}, \dots [P^{\mu _k},A_\alpha ]\dots ]], \qquad k\ge 1. \end{aligned}$$
(B.36)

By evaluating \([P^2,O]\), we find

$$\begin{aligned}{}[P^2,O]= (-1)^k 2p_{\mu _1}(b_\alpha ^{\mu _1\dots \mu _k}+c_{\alpha }^{\mu | \mu _1\dots \mu _k}p_\mu ). \end{aligned}$$
(B.37)

This should be compared with the admissible form for \([P^2,O]\) on the right-hand side of (B.35). We find that this requires

$$\begin{aligned} (-1)^k 2p_{\mu _1}(b_\alpha ^{\mu _1\dots \mu _k}+c_{\alpha }^{\mu | \mu _1\dots \mu _k}p_\mu )=0. \end{aligned}$$
(B.38)

This constraint is applicable for \(k\ge 1\), because otherwise O does not exist, see (B.36).

Equation (B.38) should be satisfied for any p on the mass shell with \(m^2>0\). This leads to

$$\begin{aligned} b_\alpha ^{\mu _1\dots \mu _k}=0, \qquad c_{\alpha }^{\mu | \mu _1\dots \mu _k}=-c_{\alpha }^{\mu _1| \mu \dots \mu _k}. \end{aligned}$$
(B.39)

The symmetry condition on c can be solved non-trivially only for \(k=1\), see exercise 5 for \(k=2\) case. We, thus, find the only non-trivial solution to be

$$\begin{aligned} c^{\mu |\mu _1}=-c^{\mu _1|\mu }, \end{aligned}$$
(B.40)

which generates Lorentz transformations.

In summary, we are left with the following possibilities for symmetries of the S-matrix in massive theories: for \(k=1\) these may only contain Lorentz transformations, while for \(k=0\) A’s commute with momenta and, as was shown on previous steps of the proof, may be either internal symmetries or momenta themselves. This finishes the proof of the Coleman-Mandula theorem for massive particles.

This argument can be naturally extended to include massless particles. For massless particles the right-hand side of (B.35) does not have the mass term, so we find

$$\begin{aligned} (-1)^k 2p_{\mu _1}(b_\alpha ^{\mu _1\dots \mu _k}+c_{\alpha }^{\mu | \mu _1\dots \mu _k}p_\mu ) = (O'-O)p^2. \end{aligned}$$
(B.41)

In addition to the solutions that we have already discussed in the massive case, (B.41) can be solved as

$$\begin{aligned} c_{\alpha }^{\mu | \mu _1\dots \mu _k} = \eta ^{\mu ( \mu _1} d_\alpha ^{\mu _2 \dots \mu _k)} \end{aligned}$$
(B.42)

for some d. Solutions to (B.42) with \(k=1\) correspond to conformal symmetries. Solutions with \(k\ge 2\) can be argued away, e.g. by noticing that the associated A under commutator generate an unbounded number of derivatives in p, which contradicts (B.31). Summarising, we find that for massless particles, the symmetry of the S-matrix may consist of a direct product of the conformal algebra and the algebra of internal symmetries.

C Helicity

For the 4d Poincare group it is conventional to introduce the Pauli-Lubanski pseudovector

$$\begin{aligned} W_{\mu }\equiv \frac{1}{2}\varepsilon _{\mu \nu \rho \sigma }J^{\nu \rho }P^\sigma . \end{aligned}$$
(C.1)

It is straightforward to compute that \([P,W]=0\), so one can pick a basis in the space of states so that both P and W take definite values. We will use this basis in the following.

Focusing on massless fields, we take the standard momentum as in (2.23). Then, the only non-vanishing components of P are \(p_3=p_0\) and the only component of J, that is non-trivially realised is \(J^{12}\). It then straightforward to see that the only non-vanishing components of W are given by

$$\begin{aligned} W_0 = J^{12}p_3, \qquad W_3 = J^{12}p_0. \end{aligned}$$
(C.2)

This implies that for the chosen basis, for the states with the standard momentum, \(J^{12}\) also takes a definite value

$$\begin{aligned} J^{12}|p,\lambda \rangle = \lambda |p,\lambda \rangle . \end{aligned}$$
(C.3)

Equation (C.3) holds for the frame, in which momentum takes the standard form. To write it in the Lorentz-covariant form we note that (C.2) and (C.3) together entail

$$\begin{aligned} W_\mu |p,\lambda \rangle =\lambda P_\mu |p,\lambda \rangle . \end{aligned}$$
(C.4)

Since both \(W_\mu \) and \(P_\mu \) transform as vectors under parity-preserving Lorentz transformations, by keeping \(\lambda \) Lorentz invariant, (C.4) takes a manifestly covariant form. Together with the Wigner approach of induced representations, this implies that \(\lambda \) defined as the proportionality coefficient between W and P is the same for all the states in the representation and, hence, can be used to label different massless representations in the same way as spin does.

To understand the connection between helicity \(\lambda \) and spin, let us return to representations of the Wigner little group. In a given case it is SO(2), which for the standard momentum is generated by \(J^{12}\). In the main body of the text irreducible representations of SO(2) were given as traceless symmetric tensors of SO(2) with spin being the rank of a tensor. By counting the number of independent components of such a tensor, it is not hard to see that it is two for \(s>0\) and one for \(s=0\). This may seem to be in contradiction with (C.3), which suggests that a representation space of the Wigner little group is generated by a single vector \(|p,\lambda \rangle \), so it is one-dimensional.

To clarify what actually happens, we consider a simple example of the SO(2) vector representation. In this case \(J^{12}\) acts via a matrix

$$\begin{aligned} J_{12}= \left( \begin{array}{cc} 0 &{} i\\ -i &{} 0 \end{array} \right) . \end{aligned}$$
(C.5)

It is straightforward to find that its eigenvalues are \(\lambda =\pm 1\) and the associated eigenvectors are

$$\begin{aligned} v_{\pm }=\left( \begin{array}{c} 1\\ \pm i \end{array}\right) . \end{aligned}$$
(C.6)

Obviously, these are not real, so there is no contradiction with irreducibility of the real vector representation of SO(2).

Still, when we are dealing with a real vector representation, it may be convenient to use basis (C.6) in which \(J^{12}\) acts diagonally. At the same time, the coordinates of a vector in this basis should satisfy certain reality conditions to ensure that the associated vector is, indeed, real. The same refers to representations of the Poincare algebra obtained from these by the Wigner induced representation technique. A similar result holds for fields of any spin \(s>0\): a symmetric traceless tensor of rank s has two eigenvectors with respect to \(J^{12}\) with eigenvalues being \(\pm s\) and both eigenvectors corresponding to complex tensors.

C.1 Helicity in the Light-Cone Gauge

With the necessary background reviewed, let us demonstrate that the representation given in (13.5)-(13.8), indeed, has helicity \(\lambda \). To this end, we evaluate \(W_0\) for the standard momentum, see (C.2). With some simple algebra one finds that only the spin part of J contributes, moreover,

$$\begin{aligned} S^{12} = -i S^{x\bar{x}}. \end{aligned}$$
(C.7)

Taking into account, in addition, the relative factors due to the definition of generators (12.40), we find that (13.8) leads to

$$\begin{aligned} W_0 = \lambda p^3 = \lambda p_0. \end{aligned}$$
(C.8)

Therefore, the proportionality coefficient between W and P is, indeed, given by \(\lambda \).

Finally, let us mention the intuitive meaning of helicity. Remembering that only the spin part of J contributes to W, we have

$$\begin{aligned} W_0 = \frac{1}{2}\varepsilon _{0\nu \rho \sigma }S^{\nu \rho }P^\sigma = \lambda P_0. \end{aligned}$$
(C.9)

For the expression to be non-vanishing, indices \(\nu \), \(\rho \) and \(\sigma \) may take only spatial values. Moreover, the Levi-Civita tensor reduces to the spatial one

$$\begin{aligned} \frac{1}{2}\varepsilon _{ij k}S^{ij}P^k = \lambda P_0. \end{aligned}$$
(C.10)

Then, helicity becomes

$$\begin{aligned} \lambda = \frac{S_k p^k}{p_0}, \qquad S_{k} \equiv \frac{1}{2}\varepsilon _{ij k}S^{ij}. \end{aligned}$$
(C.11)

Since \(p_0=\pm \sqrt{p^k p_k}\), (C.11) implies that helicity, up to a sign, is a projection of spin on the spatial part of momentum.

D Fourier Transform for the Light-Cone Approach

In the light-cone deformation procedure it is convenient to make the Fourier transform with respect to spatial coordinates (13.48), which is then followed by the change of variables \(p=iq\). For readers convenience, we present here some of the useful formulas in the Fourier transformed form.

In these terms the canonical commutator reads

$$\begin{aligned}{}[\Phi ^{\lambda _1}(q^\perp _1,x^+),\Phi ^{\lambda _2}(q^\perp _2,x^+)] = \frac{\delta ^{\lambda _1+\lambda _2,0}\delta ^3(q^\perp _1+q^\perp _2)}{\beta _1-\beta _2} \end{aligned}$$
(D.1)

and the Noether charges are

$$\begin{aligned} \nonumber P_2^i&= \sum _{\lambda }\int d^3q^\perp _1 d^3q^\perp _2 \delta ^3(q^\perp _1+q^\perp _2)\beta _1 \Phi ^{-\lambda }(q^\perp _1,x^+) p^i_2(q_2,\partial _2) \Phi ^\lambda (q^\perp _2,x^+),\\ J_2^{ij}&= \sum _{\lambda }\int d^3q^\perp _1 d^3q^\perp _2 \delta ^3(q^\perp _1+q^\perp _2)\beta _1 \Phi ^{-\lambda }(q^\perp _1,x^+) j^{ij}_2(q_2,\partial _2) \Phi ^\lambda (q^\perp _2,x^+), \end{aligned}$$
(D.2)

where

$$\begin{aligned} {\begin{matrix} p_2^+ &{}=q^+, \qquad p_2^- = -\frac{q \bar{q}}{\beta }, \qquad p_2 = q, \qquad \bar{p}_2= \bar{q},\\ j_2^{+-} &{}= \frac{\partial }{\partial \beta }\beta , \qquad j_2^{x\bar{x}} = N_q- N_{\bar{q}} -\lambda ,\\ j_2^{x+}&{} = -\beta \frac{\partial }{\partial \bar{q}}, \qquad j_2^{x-} = \frac{\partial }{\partial \bar{q}}\frac{q \bar{q}}{\beta } + q \frac{\partial }{\partial \beta }+\lambda \frac{q}{\beta },\\ j_2^{\bar{x}+}&{} = -\beta \frac{\partial }{\partial q},\qquad j_2^{\bar{x}-} = \frac{\partial }{\partial q}\frac{q \bar{q}}{\beta } + \bar{q} \frac{\partial }{\partial \beta } -\lambda \frac{\bar{q}}{\beta } \end{matrix}} \end{aligned}$$
(D.3)

and

$$\begin{aligned} N_q \equiv q\frac{\partial }{\partial q}, \qquad N_{\bar{q}} \equiv \bar{q}\frac{\partial }{\partial \bar{q}}. \end{aligned}$$
(D.4)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ponomarev, D. Basic Introduction to Higher-Spin Theories. Int J Theor Phys 62, 146 (2023). https://doi.org/10.1007/s10773-023-05399-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10773-023-05399-5

Keywords

Navigation