This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Self-dual pp-wave solutions in chiral higher-spin gravity

Tung Tran
Abstract

We show that chiral higher-spin gravity with a vanishing cosmological constant admits a class of exact self-dual pp-wave solutions derived from harmonic scalar functions and two principal spinors. These solutions satisfy both the linear and non-linear equations of motion, as they annihilate all higher-order vertices, leading to the equations of motion for free fields on a self-dual background sourced by a positive-helicity spin-2 field. Our method employs a simple light-cone ansatz for positive-helicity chiral higher-spin fields, along with a modified Kerr-Schild ansatz adapted for the self-dual gravity framework.

1 Introduction

By extending the usual GR with massless higher-spin fields, one can enlarge the symmetries of gravity by those associated with higher-spin fields, see e.g. [1] for an overview. This led to various higher-spin gravities (HSGRA)s, which serve as simple toy models toward a theory of quantum gravity, one that is presumably UV-finite and free of black hole-like singularities.

Intriguingly, to construct theories of interacting massless higher-spin fields, one often needs to relax some of the usual assumptions underlying field theories such as unitarity or parity invariance, which, to some extent, is not a surprise fact (see discussion in [2, 3, 4, 5, 6, 7, 8]).

At the moment, QFT-compatible HSGRAs with propagating degrees of freedom are either higher-spin extension of (self-dual) Weyl gravity [9, 10, 11, 12, 13, 14], or chiral higher-spin theories, see e.g. [15, 16, 17, 18, 19, 20, 21, 22]. There are also in 3d3d HSGRA [23, 24, 25, 26, 27, 28, 29, 30, 31] and a higher-spin extension of Jackiw-Teitelboim gravity in 2d2d [32]. Another interesting example is the higher-spin gauge theory induced by the IKKT matrix model on an FLRW cosmological background with a truncated higher-spin spectrum [33, 34].111See e.g. [35, 36, 37] and references therein for the recent development of this model.

In this work, we focus on chiral higher-spin gravity and explore the space of its exact solutions to examine how much they differ from GR.222See [38] for the lower-spin BPST solution in chiral HSGRA.

One of the key features of chiral HSGRA is that its gravitational background is self-dual [39], i.e. background with Euclidean or split signature where half of the component of the Weyl tensor (the anti self-dual part) is zero. Of course, one can choose to work with a complexified setting, which provides more flexibility in doing computation. In fact, this philosophy, which has deep roots in twistor theory [40], is what we want to employ in this paper in finding exact solutions of chiral HSGRA using its equations of motion [41, 42].

The rest of the paper is organized as follows:

  • -

    In Section 2, we present crucial information, including the spectrum and explicit forms of local vertices, to study chiral higher-spin gravity (HSGRA) through its equations of motion derived from certain minimal model with local LL_{\infty}-structures of a certain Q-manifold [43, 44].

  • -

    Section 3 studies the exact solutions of chiral HSGRA sourced by a set of harmonic scalar functions {𝙷s}\{\mathtt{H}_{s}\} and the principal spinors (λ,λ~)(\lambda,\tilde{\lambda}) à la Newman-Penrose. Since all local vertices of chiral HSGRA are smooth in the cosmological constant Λ\Lambda, as shown in [42], it seems reasonable to first explore the case Λ=0\Lambda=0. This is exactly what we do in this paper.

    The spin-2 sector of the theory needs some special attention, since it generates the proposed self-dual pp-wave background away from a flat background. In particular, we will adapt the Cartan structure equations to the context of self-dual GR as discussed in [45], which admits a smooth deformation from flat space to (A)dS. By considering an appropriate harmonic ansatz for fluctuation, akin to those in [46, 47], we show that the spin-2 sector of chiral HSGRA indeed admit a class of exact solutions, which solve the full non-linear EoMs, for any harmonic profile function 𝙷2\mathtt{H}_{2}. From here, the construction for exact higher-spin solutions is completely analogous. Note that our solution has a simple form, and it makes most of the structure maps of the underlying LL_{\infty}-algebra (for Λ=0\Lambda=0) vanish, see earlier discussion in [41]. Moreover, it is reasonable to speculate that once the space of solutions is found, it should be ‘portable’ to other nearby local higher-spin theories [48, 49] as well.

    In the concluding part of Section 3, we also derive the effective action corresponding to our exact solutions and show that this action indeed reduces to the standard kinetic term for free spinning fields propagating on the proposed self-dual pp-wave background. That is, despite having non-trivial interactions, the theory appears to be free when the scattering data are those obtained exact solutions.

  • -

    We end with a discussion in Section 4. There are also appendices which the reader can refer to in due time.

Notation.

We use lowercase Roman letters a,b=1,2,3,4a,b=1,2,3,4 to denote tangent indices, while uppercase letters A,B=1,2,3,4A,B=1,2,3,4 will be used for twistor or 𝔰𝔭(4)\mathfrak{sp}(4)-indices. The Greek letters α,β=0,1\alpha,\beta=0,1 and α˙,β˙=0˙,1˙\dot{\alpha},\dot{\beta}=\dot{0},\dot{1} are used to for spinorial indices. Symmetrized indices are denoted by the same letters, e.g. AaBa=12(Aa1Ba2+Aa2Ba1)A_{a}B_{a}=\frac{1}{2}(A_{a_{1}}B_{a_{2}}+A_{a_{2}}B_{a_{1}}) while fully symmetric rank-s tensor will be denoted by Ta(s)=Ta1asT_{a(s)}=T_{a_{1}\ldots a_{s}}.

2 Review

This section provides the mathematical background used to construct explicit representations of chiral HSGRA with vanishing cosmological constant. We shall be brief and refer the reader to [41, 42] for the details and to [50, 51] for a recent development.

2.1 Free differential algebra

Let \mathcal{M} be a 4d4d spacetime manifold and C(T[1])Ω()C^{\infty}(T[1]\mathcal{M})\equiv\Omega^{\bullet}(\mathcal{M}) be its associated graded commutative algebra of differential forms endowed with the exterior differential dd, which squares to zero. Then, the pair (T[1],Qd)(T[1]\mathcal{M},Q_{\mathcal{M}}\equiv d) present the simplest example of a QQ-manifold.

Now, consider a map ΦA(x,dx):T[1]𝒩\Phi^{A}(x,dx):T[1]\mathcal{M}\rightarrow\mathcal{N} from the source T[1]T[1]\mathcal{M} to another target QQ-manifold (𝒩,Q𝒩)(\mathcal{N},Q_{\mathcal{N}}) with coordinates ΦA\Phi^{A} preserving the differential. Then, around a stationary point 𝚙𝒩\mathtt{p}\in\mathcal{N}, the associated set of pde. for ΦA\Phi^{A} takes the form

dΦA\displaystyle d\Phi^{A} =Q𝒩A(Φ).\displaystyle=Q_{\mathcal{N}}^{A}(\Phi)\,. (2.1)

This is also known as free differential algebra (FDA) [43, 52, 53, 54]. The nilpotency of Q𝒩QQ_{\mathcal{N}}\equiv Q is equivalent to the statement:

QBΦBQA=0±QBQAM1Mn=N1BNk0.\displaystyle Q^{B}\frac{\partial}{\partial\Phi^{B}}Q^{A}=0\qquad\Leftrightarrow\qquad\sum\pm Q^{B}{}_{M_{1}\ldots M_{n}}Q^{A}{}_{N_{1}\ldots B\ldots N_{k}}=0\,. (2.2)

This gives us an LL_{\infty}-algebra [44, 55] whose LL_{\infty}-structure maps are the Taylor’s coefficients QM1NnBQ^{B}_{M_{1}\ldots N_{n}}.

In the simplest setting, which works for generic gauge theories and gravity, we can choose Φ(x)=(ω(x),𝖢(x))\Phi(x)=(\omega(x),\mathsf{C}(x)) to be the graded coordinates of 𝒩\mathcal{N}. Here, ω\omega is a gauge field of degree one taking values in some Lie algebra 𝔤\mathfrak{g}, and 𝖢\mathsf{C} is a ‘matter’ field of degree zero taking values in some representation ρ(𝔤)\rho(\mathfrak{g}) of 𝔤\mathfrak{g}. Then, the equations of motion resulting from (2.1) with the above field content, in general, are

dω\displaystyle d\omega =𝒱(ω,ω)+𝒱(ω,ω,𝖢)+𝒱(ω,ω,𝖢,𝖢)+,\displaystyle=\mathcal{V}(\omega,\omega)+\mathcal{V}(\omega,\omega,\mathsf{C})+\mathcal{V}(\omega,\omega,\mathsf{C},\mathsf{C})+\ldots\,, (2.3a)
d𝖢\displaystyle d\mathsf{C} =𝒰(ω,𝖢)+𝒰(ω,𝖢,𝖢)+𝒰(ω,𝖢,𝖢,𝖢)+.\displaystyle=\mathcal{U}(\omega,\mathsf{C})+\mathcal{U}(\omega,\mathsf{C},\mathsf{C})+\mathcal{U}(\omega,\mathsf{C},\mathsf{C},\mathsf{C})+\ldots\,. (2.3b)

where the notations (𝒱,𝒰)(\mathcal{V},\mathcal{U}) are obvious. To sum things up, we can define a theory whenever we can specify {(ω,𝖢)}\{(\omega,\mathsf{C})\} and {𝒱n(ω,ω,𝖢,),𝒰n(ω,𝖢,)}\{\mathcal{V}_{n}(\omega,\omega,\mathsf{C},\ldots)\,,\mathcal{U}_{n}(\omega,\mathsf{C},\ldots)\}.333The subscripts of 𝒱\mathcal{V} and 𝒰\mathcal{U} denote the number of arguments that enter the corresponding vertices.

Note that (2.3) can be referred to as a deformation of the following differential graded Lie algebra (dgLa)

dω=12[ω,ω],d𝖢=ρ(ω)𝖢.\displaystyle d\omega=\frac{1}{2}[\omega,\omega]\,,\qquad d\mathsf{C}=\rho(\omega)\mathsf{C}\,. (2.4)

The first equation is simply the definition of a flat connection associated to the one-form ω\omega, while the second expresses the covariant constancy of the 0-form 𝖢\mathsf{C} wrt. the flat connection.

2.2 Chiral HSGRA

2.2.1 Chiral FDA data

The local isomorphism SO(4,)SL(2,)×SL(2,)SO(4,\mathbb{C})\simeq SL(2,\mathbb{C})\times SL(2,\mathbb{C}) in 4d4d complexified spacetime allows us to write any tensorial fields in terms of spin-tensor fields Tα(m)α˙(n)T_{\alpha(m)\,\dot{\alpha}(n)}.

In the case of chiral HSGRA, (ω,𝖢)(\omega,\mathsf{C}) can be written in terms of generating functions as

1-form :\displaystyle: ω\displaystyle\omega =s1(2s)!ωA(2s)YA(2s)=m,n01m!n!ωα(m)α˙(n)yα(m)y~α˙(n),m+n2,\displaystyle=\sum_{s}\frac{1}{(2s)!}\omega_{A(2s)}Y^{A(2s)}=\sum_{m,n\geq 0}\frac{1}{m!n!}\omega_{\alpha(m)\,\dot{\alpha}(n)}\,y^{\alpha(m)}\tilde{y}^{\dot{\alpha}(n)}\,,\qquad m+n\in 2\mathbb{N}\,, (2.5a)
0-form :\displaystyle: 𝖢\displaystyle\mathsf{C} =s1(2s)!𝖢A(2s)YA(2s)=m,n01m!n!𝖢α(m)α˙(n)yα(m)y~α˙(n),m+n2.\displaystyle=\sum_{s}\frac{1}{(2s)!}\mathsf{C}_{A(2s)}Y^{A(2s)}=\sum_{m,n\geq 0}\frac{1}{m!n!}\mathsf{C}_{\alpha(m)\,\dot{\alpha}(n)}\,y^{\alpha(m)}\tilde{y}^{\dot{\alpha}(n)}\,,\qquad m+n\in 2\mathbb{N}\,. (2.5b)

where YA=(yα,y~α˙)Y^{A}=(y^{\alpha},\tilde{y}^{\dot{\alpha}}). Here, yα,y~α˙y^{\alpha},\tilde{y}^{\dot{\alpha}} are commutative spinorial variables, and yα(n)=yα1yαny^{\alpha(n)}=y^{\alpha_{1}}\ldots y^{\alpha_{n}} etc. is our standard abbreviation. The dynamical bosonic fields, which give us free action, are singled out as

helicity +s+s :\displaystyle:\qquad 𝒜:\displaystyle\mathcal{A}: =s1(2s)!𝒜α(2s)yα(2s)=ω(y,y~=0),\displaystyle=\sum_{s\in\mathbb{N}}\frac{1}{(2s)!}\mathcal{A}_{\alpha(2s)}y^{\alpha(2s)}=\omega(y,\tilde{y}=0)\,, (2.6a)
helicity s-s :\displaystyle:\qquad :\displaystyle\mathcal{B}: =s1(2s)!Bα(2s)yα(2s)=𝖢(y,y~=0),\displaystyle=\sum_{s\in\mathbb{N}}\frac{1}{(2s)!}B_{\alpha(2s)}y^{\alpha(2s)}=\mathsf{C}(y,\tilde{y}=0)\,, (2.6b)
helicity    0 :\displaystyle:\qquad φ:\displaystyle\varphi: =𝖢(y=0,y~=0).\displaystyle=\mathsf{C}(y=0,\tilde{y}=0)\,. (2.6c)

Using Bα(2s)B^{\alpha(2s)} and 𝒜α(2s2)\mathcal{A}_{\alpha(2s-2)} one can write down the free action for chiral spinning fields as [39]

S=Bα(2s)Hαα𝒜α(2s2),Hαα:=hαγ˙hα.γ˙\displaystyle S=\int B^{\alpha(2s)}\wedge H_{\alpha\alpha}\nabla\mathcal{A}_{\alpha(2s-2)}\,,\qquad H_{\alpha\alpha}:=h_{\alpha\dot{\gamma}}\wedge h_{\alpha}{}^{\dot{\gamma}}\,. (2.7)

In the above, hαα˙h_{\alpha\dot{\alpha}} is the background vierbein of any SD spacetime where the ASD component Cα(4)C_{\alpha(4)} of the Weyl tensor vanish, and \nabla is the associated self-dual connection. This action is invariant under δ𝒜α(2s2)=ξα(2s2)+hαηα(2s3)β˙β˙\delta\mathcal{A}_{\alpha(2s-2)}=\nabla\xi_{\alpha(2s-2)}+h_{\alpha}{}^{\dot{\beta}}\eta_{\alpha(2s-3)\,\dot{\beta}}, which can be checked immediately by noting that 2ξα(2s2)=Hαξα(2s3)ββ\nabla^{2}\xi_{\alpha(2s-2)}=H_{\alpha}{}^{\beta}\xi_{\alpha(2s-3)\beta} and Hααhα=β˙0H_{\alpha\alpha}\wedge h_{\alpha}{}^{\dot{\beta}}=0.

As explained in [39], the physical one-form 𝒜α(2s2)ωα(2s2)\mathcal{A}_{\alpha(2s-2)}\equiv\omega_{\alpha(2s-2)} can be written as

ωα(2s2)=hββ˙Aα(2s2)ββ˙+hαϑα(2s3)α˙α˙.\displaystyle\omega_{\alpha(2s-2)}=h^{\beta\dot{\beta}}A_{\alpha(2s-2)\beta\,\dot{\beta}}+h_{\alpha}{}^{\dot{\alpha}}\vartheta_{\alpha(2s-3)\,\dot{\alpha}}\,. (2.8)

Since we can use the gauge transformation δηωα(2s2)=hαηα(2s3)α˙α˙\delta_{\eta}\omega_{\alpha(2s-2)}=h_{\alpha}{}^{\dot{\alpha}}\eta_{\alpha(2s-3)\,\dot{\alpha}} to gauge away the second component ϑ\vartheta in ωα(2s2)\omega^{\alpha(2s-2)}, we will simply consider for most of the time

ωα(2s2)=hββ˙Aα(2s2)ββ˙.\displaystyle\omega_{\alpha(2s-2)}=h^{\beta\dot{\beta}}A_{\alpha(2s-2)\beta\,\dot{\beta}}. (2.9)

Now, at the free level the FDA for chiral HSGRA is simple

ωα(2s2k)α˙(k)\displaystyle\nabla\omega_{\alpha(2s-2-k)\,\dot{\alpha}(k)} =hαωα(2s3k)α˙(k)β˙β˙,0k2s2,\displaystyle=h_{\alpha}{}^{\dot{\beta}}\omega_{\alpha(2s-3-k)\,\dot{\alpha}(k)\dot{\beta}}\,,\qquad\qquad\quad 0\leq k\leq 2s-2\,, (2.10a)
ωα˙(2s2)\displaystyle\nabla\omega_{\dot{\alpha}(2s-2)} =H~β˙β˙Cα˙(2s2)β˙β˙,H~α˙α˙:=hγα˙hγα˙,\displaystyle=\widetilde{H}^{\dot{\beta}\dot{\beta}}C_{\dot{\alpha}(2s-2)\dot{\beta}\dot{\beta}}\,,\qquad\qquad\qquad\widetilde{H}^{\dot{\alpha}\dot{\alpha}}:=h_{\gamma}{}^{\dot{\alpha}}\wedge h^{\gamma\dot{\alpha}}\,, (2.10b)
Cα(k)α˙(2s+k)\displaystyle\nabla C_{\alpha(k)\,\dot{\alpha}(2s+k)} =hββ˙Cα(k)βα˙(2s+k)β˙,\displaystyle=h^{\beta\dot{\beta}}C_{\alpha(k)\beta\,\dot{\alpha}(2s+k)\dot{\beta}}\,, (2.10c)
Bα(2s+k)α˙(k)\displaystyle\nabla B^{\alpha(2s+k)\,\dot{\alpha}(k)} =hββ˙Bα(2s+k)βα˙(k)β˙,\displaystyle=h_{\beta\dot{\beta}}B^{\alpha(2s+k)\beta\,\dot{\alpha}(k)\dot{\beta}}\,, (2.10d)

where Cα(k)α˙(2s+k)C_{\alpha(k)\,\dot{\alpha}(2s+k)} are the \nabla-generated auxiliary fields of Cα˙(2s)C_{\dot{\alpha}(2s)}, and Bα(2s+k)α˙(k)B^{\alpha(2s+k)\,\dot{\alpha}(k)} are the \nabla-generated auxiliary fields of Bα(2s)B^{\alpha(2s)}, cf. (2.6).

Since the next (auxiliary) fields can be generated by the previous (auxiliary) fields by acting \nabla on the physical fields again and again, the whole spectrum of fields can be obtained starting from the chiral fields 𝒜α(2s2)\mathcal{A}_{\alpha(2s-2)} and Bα(2s)B^{\alpha(2s)} and the scalar field φ\varphi (see Fig. 1) .

Refer to caption
Figure 1: The horizontal/vertices axes represent the number of dotted and un-dotted spinorial indices that a tensorial field has. Here, [red] are fields with positive helicity, and [blue] are fields with negative helicity. The arrows indicate the directions in which auxiliary fields will be generated by acting \nabla on the previous ones. All fields generated this way are referred to as chiral FDA data.

2.2.2 Vertices

The next ingredients we need are the structure maps/vertices (𝒱,𝒰)(\mathcal{V},\mathcal{U}) in (2.3). Conveniently, we can write all vertices of chiral HSGRA as poly-differential operators:

𝖥(Φ1,,Φn):=𝖥(y,y~,iα,~iα˙)Φ1(y1,y~1)Φn(yn,y~n)|yi,y~i=0,Φ=(ω,𝖢).\displaystyle\mathsf{F}(\Phi_{1},\ldots,\Phi_{n}):=\mathsf{F}(y,\tilde{y},\partial_{i\,\alpha},\tilde{\partial}_{i\,\dot{\alpha}})\Phi_{1}(y_{1},\tilde{y}_{1})\ldots\Phi_{n}(y_{n},\tilde{y}_{n})\Big{|}_{y_{i},\tilde{y}_{i}=0}\,,\quad\Phi=(\omega,\mathsf{C})\,. (2.11)

For conciseness, we will suppress the notation |yi=0,y~i=0\big{|}_{y_{i}=0,\,\tilde{y}_{i}=0} at the end of each expression. Here, 𝖢=(C,B)\mathsf{C}=(C,B) is the master 0-form field, which contains both negative-helicity BB’s and positive-helicity CC’s fields and the scalar field φ\varphi.

Below, we present some of the key features of these structure maps. The technical details in deriving these maps can be found in [41, 42] (see also Appendix B for a recap).

\star-product

The first structure map in the minimal model of interest is the binary structure map 𝖥2\mathsf{F}_{2}. It is given by a graded associative star-product:

Φ1Φ2\displaystyle\Phi_{1}\star\Phi_{2} =exp(YAY1A+YAY2AIABY1AY2B)Φ1(Y1)Φ2(Y2)\displaystyle=\exp\Big{(}Y^{A}\frac{\partial}{\partial Y_{1}^{A}}+Y^{A}\frac{\partial}{\partial Y_{2}^{A}}-I^{AB}\frac{\partial}{\partial Y_{1}^{A}}\frac{\partial}{\partial Y_{2}^{B}}\Big{)}\Phi_{1}(Y_{1})\Phi_{2}(Y_{2})
=exp(y1+y2+Λ12+[y1]+[y2]+[12])Φ1Φ2,\displaystyle=\exp\Big{(}\langle y\,\partial_{1}\rangle+\langle y\,\partial_{2}\rangle+\Lambda\langle\partial_{1}\,\partial_{2}\rangle+[y\,\partial_{1}]+[y\,\partial_{2}]+[\partial_{1}\,\partial_{2}]\Big{)}\Phi_{1}\Phi_{2}\,, (2.12)

where Λ\Lambda is the cosmological constant, IABI^{AB} is the infinity twistor [40] obeying

IAB=(Λϵαβ00ϵα˙β˙),IAB=(ϵαβ00Λϵα˙β˙),IACIBC=ΛδA.B\displaystyle I^{AB}=\begin{pmatrix}\Lambda\epsilon^{\alpha\beta}&0\\ 0&\epsilon^{\dot{\alpha}\dot{\beta}}\end{pmatrix}\,,\qquad I_{AB}=\begin{pmatrix}\epsilon_{\alpha\beta}&0\\ 0&\Lambda\epsilon_{\dot{\alpha}\dot{\beta}}\end{pmatrix}\,,\qquad I_{AC}I^{BC}=\Lambda\delta_{A}{}^{B}\,. (2.13)

Note that

yi=yαiα,ij=iαjα;[yi]=y~α˙~iα˙,[ij]=~iα˙~jα˙,\displaystyle\langle y\,\partial_{i}\rangle=y^{\alpha}\partial_{i\alpha}\,,\quad\langle\partial_{i}\,\partial_{j}\rangle=\partial_{i}^{\alpha}\partial_{j\alpha}\,;\qquad[y\,\partial_{i}]=\tilde{y}^{\dot{\alpha}}\tilde{\partial}_{i\dot{\alpha}}\,,\quad[\partial_{i}\,\partial_{j}]=\tilde{\partial}_{i}^{\dot{\alpha}}\tilde{\partial}_{j\dot{\alpha}}\,, (2.14)

where iα=yiα\partial_{i\alpha}=\frac{\partial}{\partial y_{i}^{\alpha}}, ~iα˙=y~iα˙\tilde{\partial}_{i\dot{\alpha}}=\frac{\partial}{\partial\tilde{y}_{i}^{\dot{\alpha}}}. Here, eyie^{\langle y\,\partial_{i}\rangle} may be referred to as a shift operator with the property eyiΦ(yi)=Φ(yi+y)e^{\langle y\,\partial_{i}\rangle}\Phi(y_{i})=\Phi(y_{i}+y). We will use the following convention to raise and lower spinorial indices

vα=vβϵαβ,vα=vβϵβα,vα˙=vβ˙ϵα˙β˙,vα˙=vβ˙ϵβ˙α˙.\displaystyle v^{\alpha}=v_{\beta}\epsilon^{\alpha\beta}\,,\quad v_{\alpha}=v^{\beta}\epsilon_{\beta\alpha}\,,\quad v^{\dot{\alpha}}=v_{\dot{\beta}}\epsilon^{\dot{\alpha}\dot{\beta}}\,,\quad v_{\dot{\alpha}}=v^{\dot{\beta}}\epsilon_{\dot{\beta}\dot{\alpha}}\,. (2.15)

The Weyl algebra 𝖠2\mathsf{A}_{2} of polynomial functions in YA=(yα,y~α˙)Y^{A}=(y^{\alpha},\tilde{y}^{\dot{\alpha}}) defined in terms of the \star-product cf. (2.2.2) admits the decomposition 𝖠2=𝖠1(y)𝖠~1(y~)\mathsf{A}_{2}=\mathsf{A}_{1}(y)\otimes\widetilde{\mathsf{A}}_{1}(\tilde{y}) where the subscript denotes the number of canonical pairs. We then define

𝔥𝔰Λ:=𝖠1(𝖠~1𝙼𝚊𝚝(n,𝕂)),\displaystyle\mathfrak{hs}_{\Lambda}:=\mathsf{A}_{1}\otimes\big{(}\widetilde{\mathsf{A}}_{1}\otimes\mathtt{Mat}(n,\mathbb{K})\big{)}\,, (2.16)

to be the underlying higher-spin algebra of the system. Note that when Λ=0\Lambda=0, the Poisson structure defined by the infinity twistor is degenerate, and 𝖠1\mathsf{A}_{1} becomes a commutative algebra while 𝖠~1𝙼𝚊𝚝(n,𝕂)\widetilde{\mathsf{A}}_{1}\otimes\mathtt{Mat}(n,\mathbb{K}) remains to be an non-commutative (NC) algebra. This is the setting that we consider in this work.

Higher-order vertices.

The details of deriving higher-order vertices for chiral FDA, cf. (2.3), can be found in e.g. [51]. Here, we shall emphasize some of the main points, which are relevant for our analysis in Section 3. Henceforth, we focus on the case where Λ=0\Lambda=0.

\diamond 𝒱n3(ω,ω,𝖢,,𝖢)\mathcal{V}_{n\geq 3}(\omega,\omega,\mathsf{C},\ldots,\mathsf{C}) vertices. The 𝒱n(ω,ω,𝖢,)\mathcal{V}_{n}(\omega,\omega,\mathsf{C},\ldots) maps consist of (n2)=n(n1)/2\binom{n}{2}=n(n-1)/2 sub-vertices:

𝒱n(ω,ω,𝖢,,𝖢)=Vω,ω,𝖢,,𝖢(n)+Vω,𝖢,ω,𝖢,,𝖢(n)+\displaystyle\mathcal{V}_{n}(\omega,\omega,\mathsf{C},\ldots,\mathsf{C})=V^{(n)}_{\omega,\omega,\mathsf{C},\ldots,\mathsf{C}}+V^{(n)}_{\omega,\mathsf{C},\omega,\mathsf{C},\ldots,\mathsf{C}}+\ldots (2.17)

We shall denote ii and jj to be the positions of ω\omega’s in the sub-vertices VV’s (assuming i<ji<j) since they can be useful references for describing the sub-vertices VV’s. It is worth noting that 𝒱n\mathcal{V}_{n} are LL_{\infty} structure maps obtained by suitable graded symmetrization of a certain AA_{\infty} structure maps/sub-vertices V(n)V^{(n)} (see e.g. Appendix B).

For instance, a vertex of the form V𝖢,,ωi,𝖢,,ωj,,𝖢(n)V^{(n)}_{\mathsf{C},\ldots,\omega_{i},\mathsf{C},\ldots,\omega_{j},\ldots,\mathsf{C}} has the following expressions [50]:

V𝖢,,ωi,𝖢,,ωj,,𝖢(n)=()i+j1σGijnijn2Δn2exptr(𝖯Gij𝖰T),\displaystyle V^{(n)}_{\mathsf{C},\ldots,\omega_{i},\mathsf{C},\ldots,\omega_{j},\ldots,\mathsf{C}}=(-)^{i+j-1}\sigma_{G_{ij}}\,\mho_{n}\langle\partial_{i}\,\partial_{j}\rangle^{n-2}\int_{\Delta_{n-2}}\exp\mathrm{tr}\,(\mathsf{P}_{G_{ij}}\mathsf{Q}^{T})\,, (2.18)

where

𝖰\displaystyle\mathsf{Q} =(u1u2un21uiv1v2vn21vi),\displaystyle=\begin{pmatrix}u_{1}&u_{2}&\ldots&u_{n-2}&1-\sum u_{i}\\ v_{1}&v_{2}&\ldots&v_{n-2}&1-\sum v_{i}\end{pmatrix}\,, (2.19a)
𝖯Gij\displaystyle\mathsf{P}_{G_{ij}} =(i1ij0in2yij10jijn2yj).\displaystyle=\begin{pmatrix}\langle\partial_{i}\,\partial_{1}\rangle&\ldots&-\langle\partial_{i}\,\partial_{j}\rangle&\ldots&0&\ldots&\langle\partial_{i}\,\partial_{n-2}\rangle&-\langle y\,\partial_{i}\rangle\\ \langle\partial_{j}\,\partial_{1}\rangle&\ldots&0&\ldots&-\langle\partial_{j}\,\partial_{i}\rangle&\ldots&\langle\partial_{j}\,\partial_{n-2}\rangle&-\langle y\,\partial_{j}\rangle\end{pmatrix}\,. (2.19b)

The integration domain Δn2n\Delta_{n}\in\mathbb{R}^{2n} in (2.18) is a compact convex nn-gon and the integrating variables uku_{k} and vkv_{k} obey

u1v1u2v2unvn1U1V,U=ui,V=vi.\displaystyle\frac{u_{1}}{v_{1}}\leq\frac{u_{2}}{v_{2}}\leq\ldots\leq\frac{u_{n}}{v_{n}}\leq\frac{1-U}{1-V}\,,\qquad U=\sum u_{i}\,,\quad V=\sum v_{i}\,.

Furthermore, GijG_{ij} denotes the corresponding tree/graph in homological perturbation theory from which VV’s come from. Additionally, σGij\sigma_{G_{ij}} represents the permutation that arranges fields into ‘natural’ order of the associated graph. Lastly,

n:=exp(i=1n[yi]+i<jn[ij]),\displaystyle\mho_{n}:=\exp\Big{(}\sum_{i=1}^{n}[y\,\partial_{i}]+\sum_{i<j\leq n}[\partial_{i}\,\partial_{j}]\Big{)}\,, (2.20)

is a poly-differential operator acting solely on the y~\tilde{y}-sector by virtue of the \star-product (2.2.2).

Canonically, we can take G12=(ω1,ω2,𝖢3,,𝖢n2)G_{12}=(\omega_{1},\omega_{2},\mathsf{C}_{3},\ldots,\mathsf{C}_{n-2}) to be the reference graph where all 𝖢\mathsf{C}’s are on the right of ω\omega’s. For concreteness, the explicit form of the sub-vertex Vω,ω,𝖢,,𝖢(n)V^{(n)}_{\omega,\omega,\mathsf{C},\ldots,\mathsf{C}} associated with G12G_{12} has the following form

Vω1,ω2,𝖢,,𝖢(n)(y,y~;i,~i)=n12n2×\displaystyle V^{(n)}_{\omega_{1},\omega_{2},\mathsf{C},\ldots,\mathsf{C}}(y,\tilde{y};\partial_{i},\tilde{\partial}_{i})=\mho_{n}\langle\partial_{1}\,\partial_{2}\rangle^{n-2}\times
×Δn2exp[(1k=1n2uk)y1+(1k=1n2vk)y2+m1,2um1m+m1,2vm2m].\displaystyle\quad\times\int_{\Delta_{n-2}}\exp\Bigg{[}\Big{(}1-\sum_{k=1}^{n-2}u_{k}\Big{)}\langle y\,\partial_{1}\rangle+\Big{(}1-\sum_{k=1}^{n-2}v_{k}\Big{)}\langle y\,\partial_{2}\rangle+\sum_{m\neq 1,2}u_{m}\langle\partial_{1}\,\partial_{m}\rangle+\sum_{m\neq 1,2}v_{m}\langle\partial_{2}\,\partial_{m}\rangle\Bigg{]}\,. (2.21)

Then, any other graph can be obtained from G12G_{12} by moving some number of 𝖢\mathsf{C}’s in between ω\omega’s and cyclicing all arguments to a specific position e.g. V𝖢,,ω,𝖢,,ω,𝖢,𝖢V𝖢,𝖢,𝖢,,ω,𝖢,,ωV_{\mathsf{C},\ldots,\omega,\mathsf{C},\ldots,\omega,\mathsf{C},\mathsf{C}}\rightarrow V_{\mathsf{C},\mathsf{C},\mathsf{C},\ldots,\omega,\mathsf{C},\ldots,\omega}. This allows us to constructs the remaining vertices/graphs from the reference ones.

\diamond 𝒰(ω,𝖢,,𝖢)\mathcal{U}(\omega,\mathsf{C},\ldots,\mathsf{C}) vertices. It turns out that (𝒱,𝒰)(\mathcal{V},\mathcal{U}) vertices can be related to each others through a \mathbb{C}-linear pairing inherited from a degree 2 cyclic AA_{\infty}-algebra:

|:𝒩𝒩[1][2],wherea|b=()(|a|+1)(|b|+1)b|a.\langle-|-\rangle:\mathcal{N}\otimes\mathcal{N}^{\vee}[-1]\rightarrow\mathbb{C}[-2]\,,\quad\text{where}\quad\langle a|b\rangle=(-)^{(|a|+1)(|b|+1)}\langle b|a\rangle\,.

The natural pairings in the yy-sector used to construct all vertices [41, 42, 51] are as follows:

ω|𝖢=𝖢|ω=exp[12]ω(y1)𝖢(y2)\langle\omega|\mathsf{C}\rangle=-\langle\mathsf{C}|\omega\rangle=\exp[\langle\partial_{1}\,\partial_{2}\rangle]\omega(y_{1})\mathsf{C}(y_{2})
V(ω,ω,𝖢,,𝖢)|C=+ω|U(ω,𝖢,,𝖢)V(1,,n|y)+U(y,1,,n1|n)\!\begin{aligned} \big{\langle}V(\omega,\omega,\mathsf{C},\ldots,\mathsf{C})|C\big{\rangle}&=+\big{\langle}\omega|U(\omega,\mathsf{C},\ldots,\mathsf{C})\big{\rangle}\\ V(\partial_{1},\ldots,\partial_{n}|y)&\mapsto+U(y,\partial_{1},\ldots,\partial_{n-1}|-\partial_{n})\end{aligned}
𝖢|V(𝖢,,ω,𝖢,,𝖢,ω)=U(𝖢,𝖢,,ω,𝖢,,𝖢)|ωV(y|1,,n)U(1|2,,n,y)\!\begin{aligned} \big{\langle}\mathsf{C}|V(\mathsf{C},\ldots,\omega,\mathsf{C},\ldots,\mathsf{C},\omega)\big{\rangle}&=-\big{\langle}U(\mathsf{C},\mathsf{C},\ldots,\omega,\mathsf{C},\ldots,\mathsf{C})|\omega\big{\rangle}\\ V(y|\partial_{1},\ldots,\partial_{n})&\mapsto-U(-\partial_{1}|\partial_{2},\ldots,\partial_{n},y)\end{aligned}
Table 1: Some natural pairings in chiral-FDA.

For instance, to obtain U𝖢,,𝖢,ω,,𝖢U(𝖢,,𝖢,ω,,𝖢)U_{\mathsf{C},\ldots,\mathsf{C},\omega,\ldots,\mathsf{C}}\equiv U(\mathsf{C},\ldots,\mathsf{C},\omega,\ldots,\mathsf{C}) from Vω,,𝖢,ω,,𝖢V_{\omega,\ldots,\mathsf{C},\omega,\ldots,\mathsf{C}}, we replace yny\mapsto-\partial_{n}, 1y\partial_{1}\mapsto y and ii1\partial_{i}\mapsto\partial_{i-1} (for ini\neq n), while to obtain U𝖢,,ω,,𝖢U_{\mathsf{C},\ldots,\omega,\ldots,\mathsf{C}} from V𝖢,,ω,𝖢,,𝖢,ωV(𝖢,,ω,𝖢,,𝖢,ω)V_{\mathsf{C},\ldots,\omega,\mathsf{C},\ldots,\mathsf{C},\omega}\equiv V(\mathsf{C},\ldots,\omega,\mathsf{C},\ldots,\mathsf{C},\omega), we replace y1y\mapsto-\partial_{1}, ii+1\partial_{i}\mapsto\partial_{i+1} (for ini\leq n) and ny\partial_{n}\mapsto y with an overall minus sign. Since there are (n2)=n(n2)2\binom{n}{2}=\frac{n(n-2)}{2} number of VV’s sub-vertices compared to nn number of UU’s sub-vertices at a given order nn, this replacement rule always works. Therefore, as a matter of convenience, one only needs to construct all AA_{\infty} structure maps V(n)V^{(n)}, and everything will follow.

Note that the pairing |\langle-|-\rangle not only facilitates the cyclic reordering of fields and vertices, but is also defined to eliminate all auxiliary yy and y~\tilde{y} variables. This ensures that all possible contractions between the stand-alone field ω\omega or 𝖢\mathsf{C} with the multilinear maps or sub-vertices VV’s and UU’s are fully captured.

3 Light-cone ansatz and exact solutions

Armed with the above chiral FDA data, we now propose a family of explicit representations of (ω,𝖢)(\omega,\mathsf{C}) and show that they form exact harmonic solutions in chiral HSGRA.

Recall that since the positive-helicity fields CC’s can be generated from ω\omega via linear chiral FDA, cf. (2.10), we can construct all positive-helicity data starting from the physical fields 𝒜α(2s2)\mathcal{A}_{\alpha(2s-2)}. Furthermore, we show below that 𝒜α(2)\mathcal{A}_{\alpha(2)} is the field that generate a self-dual background. It is also worth stressing before running our analysis that Bα(2s)B^{\alpha(2s)} and their \nabla-generated thereof are not derived objects from the positive helicity fields 𝒜α(2s2)\mathcal{A}_{\alpha(2s-2)}. Rather, they should be regarded simply as linearized fields that propagate on the self-dual background generated by 𝒜α(2)ωα(2)\mathcal{A}_{\alpha(2)}\equiv\omega_{\alpha(2)}.

3.1 Cartan’s structure equations for empty self-dual background

One of the nice features of 𝔥𝔰Λ\mathfrak{hs}_{\Lambda} is that its 𝔰𝔬(2,3)\mathfrak{so}(2,3) subalgebra can be defined via the generators:

Lαα=12yαyα,Pαα˙=12yαy~α˙,L~α˙α˙=12y~α˙y~α˙.\displaystyle L_{\alpha\alpha}=-\frac{1}{2}y_{\alpha}y_{\alpha}\,,\qquad P_{\alpha\dot{\alpha}}=-\frac{1}{2}y_{\alpha}\tilde{y}_{\dot{\alpha}}\,,\qquad\widetilde{L}_{\dot{\alpha}\dot{\alpha}}=-\frac{1}{2}\tilde{y}_{\dot{\alpha}}\tilde{y}_{\dot{\alpha}}\,. (3.1)

As usual, Pαα˙P_{\alpha\dot{\alpha}} is translation generator, and (Lαα,L~α˙α˙)(L_{\alpha\alpha},\widetilde{L}_{\dot{\alpha}\dot{\alpha}}) are Lorentz’s. The algebra which describes AdS4 is [45]

[Lαβ,Lγδ]\displaystyle[L_{\alpha\beta},L_{\gamma\delta}]_{\star} =Λ(ϵβγLαγ+ϵαγLβδ+ϵβδLαγ+ϵαγLβδ),\displaystyle=\Lambda\big{(}\epsilon_{\beta\gamma}L_{\alpha\gamma}+\epsilon_{\alpha\gamma}L_{\beta\delta}+\epsilon_{\beta\delta}L_{\alpha\gamma}+\epsilon_{\alpha\gamma}L_{\beta\delta}\big{)}\,, (3.2a)
[L~α˙β˙,L~γ˙δ˙]\displaystyle[\widetilde{L}_{\dot{\alpha}\dot{\beta}},\widetilde{L}_{\dot{\gamma}\dot{\delta}}]_{\star} =(ϵβ˙γ˙Lα˙γ˙+ϵα˙γ˙Lβ˙δ˙+ϵβ˙δ˙Lα˙γ˙+ϵα˙γ˙Lβ˙δ˙),\displaystyle=\big{(}\epsilon_{\dot{\beta}\dot{\gamma}}L_{\dot{\alpha}\dot{\gamma}}+\epsilon_{\dot{\alpha}\dot{\gamma}}L_{\dot{\beta}\dot{\delta}}+\epsilon_{\dot{\beta}\dot{\delta}}L_{\dot{\alpha}\dot{\gamma}}+\epsilon_{\dot{\alpha}\dot{\gamma}}L_{\dot{\beta}\dot{\delta}}\big{)}\,, (3.2b)
[Lαβ,Pγγ˙]\displaystyle[L_{\alpha\beta},P_{\gamma\dot{\gamma}}]_{\star} =Λ(ϵβγPαγ˙+ϵαγPβγ˙),\displaystyle=\Lambda\Big{(}\epsilon_{\beta\gamma}P_{\alpha\dot{\gamma}}+\epsilon_{\alpha\gamma}P_{\beta\dot{\gamma}}\Big{)}\,, (3.2c)
[L~α˙β˙,Pγγ˙]\displaystyle[\widetilde{L}_{\dot{\alpha}\dot{\beta}},P_{\gamma\dot{\gamma}}]_{\star} =(ϵβ˙γ˙Pγα˙+ϵα˙γ˙Pγβ˙),\displaystyle=\big{(}\epsilon_{\dot{\beta}\dot{\gamma}}P_{\gamma\dot{\alpha}}+\epsilon_{\dot{\alpha}\dot{\gamma}}P_{\gamma\dot{\beta}}\big{)}\,, (3.2d)
[Pαα˙,Pββ˙]\displaystyle[P_{\alpha\dot{\alpha}},P_{\beta\dot{\beta}}]_{\star} =(ϵα˙β˙Lαβ+ΛϵαβL~α˙β˙).\displaystyle=\big{(}\epsilon_{\dot{\alpha}\dot{\beta}}L_{\alpha\beta}+\Lambda\epsilon_{\alpha\beta}\widetilde{L}_{\dot{\alpha}\dot{\beta}}\big{)}\,. (3.2e)

It is easy to notice that LαβL_{\alpha\beta} becomes central of the algebra (3.2) in the flat limit Λ0\Lambda\rightarrow 0. That is

[Pαα˙,Pββ˙]|Λ=0=ϵα˙β˙Lαβ,\displaystyle[P_{\alpha\dot{\alpha}},P_{\beta\dot{\beta}}]_{\star}\big{|}_{\Lambda=0}=\epsilon_{\dot{\alpha}\dot{\beta}}L_{\alpha\beta}\,, (3.3)

instead of being trivial. For this reason, the algebra generated by (3.1) in the flat limit is referred to as a deformed chiral Poincaré algebra [56].

The Cartan structure equations associated to empty flat spacetime read:

dΩ12[Ω,Ω]=0,Ω=12ϖαβLαβ+hαα˙Pαα˙+12ϖ~α˙β˙L~α˙β˙.\displaystyle d\Omega-\frac{1}{2}[\Omega,\Omega]=0\,,\qquad\Omega=\frac{1}{2}\varpi_{\alpha\beta}L^{\alpha\beta}+h_{\alpha\dot{\alpha}}P^{\alpha\dot{\alpha}}+\frac{1}{2}\widetilde{\varpi}_{\dot{\alpha}\dot{\beta}}\widetilde{L}^{\dot{\alpha}\dot{\beta}}\,. (3.4)

where hαα˙h_{\alpha\dot{\alpha}} is the background vierbein and ϖαβ\varpi_{\alpha\beta}, ϖ~α˙β˙\widetilde{\varpi}_{\dot{\alpha}\dot{\beta}} are the (A)SD components of the spin connection. In terms of components, (3.4) unfolds into

Rαβ:\displaystyle R_{\alpha\beta}: =dϖαα+Hαα=0,\displaystyle=d\varpi_{\alpha\alpha}+H_{\alpha\alpha}=0\,, (3.5a)
Tαα˙:\displaystyle T_{\alpha\dot{\alpha}}: =dhαα˙+ϖ~α˙γ˙hαγ˙hαα˙=0,\displaystyle=dh_{\alpha\dot{\alpha}}+\widetilde{\varpi}_{\dot{\alpha}\dot{\gamma}}\wedge h_{\alpha}{}^{\dot{\gamma}}\equiv\nabla h_{\alpha\dot{\alpha}}=0\,, (3.5b)
Rα˙β˙:\displaystyle R_{\dot{\alpha}\dot{\beta}}: =dϖ~α˙α˙+ϖ~α˙γ˙ϖ~α˙=γ˙0.\displaystyle=d\widetilde{\varpi}_{\dot{\alpha}\dot{\alpha}}+\widetilde{\varpi}_{\dot{\alpha}\dot{\gamma}}\wedge\widetilde{\varpi}_{\dot{\alpha}}{}^{\dot{\gamma}}=0\,. (3.5c)

The reader may notice the difference with the Cartan structure equations associated with the conventional Poincaré algebra where [P,P]=0[P,P]=0. In particular, for SD background, the ASD component of the spin connection does not participate in the torsion constraint (3.5b). Furthermore, the Riemann two-form (3.5a) reveals an intriguing feature for SD background. That is, the spin connection ϖαα\varpi_{\alpha\alpha} is the main player of this game since the vierbein and the SD component of the spin connection can be solved for once we know ϖαα.\varpi_{\alpha\alpha}.

3.2 Deformed self-dual pp-wave background

This subsection proposes a class of SD pp-wave backgrounds in which chiral HSGRA is well-defined. This class of backgrounds is generated by a fluctuation of the ASD component of the spin connection, which is characterized by a function 𝙷2(v,z,z~)\mathtt{H}_{2}(v,z,\tilde{z}) that is harmonic in the complex 2-plane (z,z~)(z,\tilde{z}), and two principal spinors (λ,λ~)(\lambda,\tilde{\lambda}).444As is well-known, any symmetric tensor of the form Sα(s)Sα1αsS_{\alpha(s)}\equiv S_{\alpha_{1}\ldots\alpha_{s}} can be factorized into a symmetrized product of one-index spinors [57], i.e. Sα1αs=τ(α1ζαs),\displaystyle S_{\alpha_{1}\ldots\alpha_{s}}=\tau_{(\alpha_{1}}\ldots\zeta_{\alpha_{s})}\,, (3.6) where each spinor {τ,,ζ}\{\tau,\ldots,\zeta\} is referred to as principal spinor of Sα(s)S_{\alpha(s)}. Note that these principal spinors can be used to classify the Weyl tensors in classical GR, see e.g. [58, 57].

Deformed Cartan’s structure equation.

Suppose we want to construct a non-trivial vacuum solution which does not describe an empty flat space. Then, the new vacuum solution whose geometrical information is encoded by the traceless part of the Riemann tensor, i.e. the Weyl tensor, can be regarded as a deformation away from empty flat space. In this case, we shall modify the Cartan’s structure equations (3.5) to

𝑹αα:\displaystyle\boldsymbol{R}_{\alpha\alpha}: =dϖαα+𝑯αα=0,\displaystyle=d\boldsymbol{\varpi}_{\alpha\alpha}+\boldsymbol{H}_{\alpha\alpha}=0\,, (3.7a)
D𝒆αα˙\displaystyle D\boldsymbol{e}_{\alpha\dot{\alpha}} =d𝒆αα˙+ϖ~α˙β˙𝒆α=β˙0,\displaystyle=d\boldsymbol{e}_{\alpha\dot{\alpha}}+\widetilde{\boldsymbol{\varpi}}_{\dot{\alpha}\dot{\beta}}\wedge\boldsymbol{e}_{\alpha}{}^{\dot{\beta}}=0\,, (3.7b)
𝑹~α˙α˙:\displaystyle\widetilde{\boldsymbol{R}}_{\dot{\alpha}\dot{\alpha}}: =dϖ~α˙β˙+ϖ~α˙γ˙ϖ~β˙=γ˙𝑯~β˙β˙Cβ˙β˙α˙α˙,\displaystyle=d\widetilde{\boldsymbol{\varpi}}_{\dot{\alpha}\dot{\beta}}+\widetilde{\boldsymbol{\varpi}}_{\dot{\alpha}\dot{\gamma}}\wedge\widetilde{\boldsymbol{\varpi}}_{\dot{\beta}}{}^{\dot{\gamma}}=\widetilde{\boldsymbol{H}}^{\dot{\beta}\dot{\beta}}C_{\dot{\beta}\dot{\beta}\dot{\alpha}\dot{\alpha}}\,, (3.7c)

where

ϖαβ=ϖαα+ωαα,𝒆αα˙=hαα˙+ωαα˙,ϖ~α˙β˙=ϖ~α˙β˙+ωα˙β˙,\displaystyle\quad\boldsymbol{\varpi}_{\alpha\beta}=\varpi_{\alpha\alpha}+\omega_{\alpha\alpha}\,,\quad\boldsymbol{e}_{\alpha\dot{\alpha}}=h_{\alpha\,\dot{\alpha}}+\omega_{\alpha\dot{\alpha}},\quad\widetilde{\boldsymbol{\varpi}}_{\dot{\alpha}\dot{\beta}}=\widetilde{\varpi}_{\dot{\alpha}\dot{\beta}}+\omega_{\dot{\alpha}\dot{\beta}}\,, (3.8)

are the deformed vierbein and spin connections with ωαα,ωαα˙,ωα˙α˙\omega_{\alpha\alpha},\omega_{\alpha\dot{\alpha}},\omega_{\dot{\alpha}\dot{\alpha}} being the associated fluctuations.555All the bold symbols are geometrical datum of the deformed background. Here, DD stands for the connection of the deformed SD background, and (3.7) contains only the SD component Cα˙(4)C_{\dot{\alpha}(4)} of the Weyl tensor due to SD condition, cf. [39].

Tetrad and dyad bases.

In 4d4d, we can associate to each point xx\in\mathcal{M} a tetrad of complex null vectors ζa=(a,na,pa,p~a)\zeta^{a}=(\ell^{a},n^{a},p^{a},\tilde{p}^{a}) (see e.g. [57]) where

ana=1=pap~a,apa=ap~a=napa=nap~a=0.\displaystyle\ell_{a}n^{a}=-1=p_{a}\tilde{p}^{a}\,,\qquad\qquad\ell_{a}p^{a}=\ell_{a}\tilde{p}^{a}=n_{a}p^{a}=n_{a}\tilde{p}^{a}=0\,. (3.9)

Then, a vector VaV^{a} may be written in the tetrad basis as

Va=ua+vna+zpa+z~p~a.\displaystyle V^{a}=u\ell^{a}+vn^{a}+zp^{a}+\tilde{z}\tilde{p}^{a}\,. (3.10)

Here, we may refer to (u,v,z,z~)(u,v,z,\tilde{z}) as spacetime coordinates.

Since the local Lorentz group of a 4d4d complexified \mathcal{M} is SL(2,)×SL(2,)SL(2,\mathbb{C})\times SL(2,\mathbb{C}), we can also associate to each point xx\in\mathcal{M} the dyad bases (λα,λ~α˙)(\lambda^{\alpha},\tilde{\lambda}^{\dot{\alpha}}) and (μα,μ~α˙)(\mu^{\alpha},\tilde{\mu}^{\dot{\alpha}}), cf. [57], where (μ,μ~)(\mu,\tilde{\mu}) is another pair of principal spinors such that

μλ=1=[μ~λ~].\displaystyle\langle\mu\,\lambda\rangle=1=[\tilde{\mu}\,\tilde{\lambda}]\,. (3.11)

Using the same normalization in (3.9), we can fix unambiguously

a=λαλ~α˙,na=μαμ~α˙,pa=λαμ~α˙,p~a=μαλ~α˙.\displaystyle\ell_{a}=\lambda_{\alpha}\tilde{\lambda}_{\dot{\alpha}}\,,\qquad n_{a}=\mu_{\alpha}\tilde{\mu}_{\dot{\alpha}}\,,\qquad p_{a}=\lambda_{\alpha}\tilde{\mu}_{\dot{\alpha}}\,,\qquad\tilde{p}_{a}=\mu_{\alpha}\tilde{\lambda}_{\dot{\alpha}}\,. (3.12)

In constructing the proposed harmonic solutions for chiral HSGRA, we will use (λ,λ~)(\lambda,\tilde{\lambda}) to characterize positive-helicity fields and (μ,μ~)(\mu,\tilde{\mu}) to describe the negative-helicity fields as they are related in the sense of (2.7).

Note that we will define the spinors (λ,λ~)(\lambda,\tilde{\lambda}) as well as (μ,μ~)(\mu,\tilde{\mu}) to be covariantly constant wrt. the connection associated to empty flat background, i.e.

λα=0,λ~α˙=0,μα=0,μ~α˙=0.\displaystyle\nabla\lambda_{\alpha}=0\,,\qquad\nabla\tilde{\lambda}_{\dot{\alpha}}=0\,,\qquad\nabla\mu_{\alpha}=0\,,\qquad\nabla\tilde{\mu}_{\dot{\alpha}}=0\,. (3.13)
Light-cone ansatz.

Typically, in standard GR, one can simplify the process of finding exact solutions for Einstein equations by considering the Kerr-Schild ansatz [59]

𝒈ab=gab+𝙷ab,𝒈ab=gab𝙷ab\displaystyle\boldsymbol{g}_{ab}=g_{ab}+\mathtt{H}\ell_{a}\ell_{b}\,,\qquad\qquad\boldsymbol{g}^{ab}=g^{ab}-\mathtt{H}\ell^{a}\ell^{b}\, (3.14)

where gabg_{ab} is the metric of the background and 𝙷\mathtt{H} is a scalar profile function whose role is to generate a deformed spacetime (,𝒈)(\mathcal{M},\boldsymbol{g}). However, in SDGR, the situation is slightly different. In particular, as the ASD component of the spin connection is the main object where all other geometrical data are derived from, cf. [60], one must consider a modified version of the Kerr-Schild ansatz, adapted to accommodate the ASD component of the spin connection. For our purpose of constructing exact solutions in SD setting, we can consider

ϖαα=ϖαα+𝙷2λ|h|λ]λαλα,λ|h|λ]:=λγhγγ˙λ~γ˙.\displaystyle\boldsymbol{\varpi}_{\alpha\alpha}=\varpi_{\alpha\alpha}+\mathtt{H}_{2}\langle\lambda|h|\lambda]\lambda_{\alpha}\lambda_{\alpha}\,,\qquad\langle\lambda|h|\lambda]:=\lambda^{\gamma}h_{\gamma\dot{\gamma}}\tilde{\lambda}^{\dot{\gamma}}\,. (3.15)

Note that the above ansatz is inspired from the recent work [46, 47]. For the sake of generality, we have introduced a subscript “2” to the profile function 𝙷\mathtt{H} to indicate that it corresponds to the ansatz for the spin-2 sector. For spin-ss sector, we may denote the corresponding profile function as 𝙷s\mathtt{H}_{s} (see below).

Harmonicity of 𝙷2\mathtt{H}_{2}.

There is a few properties that we want to impose on the profile function 𝙷2\mathtt{H}_{2}. Namely, 𝙷2\mathtt{H}_{2} must be uu-independent and harmonic in (z,z~)(z,\tilde{z})-plane spanned by (pa,p~a)(p^{a},\tilde{p}^{a}) [61], i.e.

aa𝙷2=0,pap~bab𝙷2=0.\displaystyle\ell^{a}\nabla_{a}\mathtt{H}_{2}=0\,,\qquad p^{a}\tilde{p}^{b}\nabla_{a}\nabla_{b}\mathtt{H}_{2}=0\,. (3.16)

This allows us to write 𝙷2(v,z,z~)\mathtt{H}_{2}(v,z,\tilde{z}) as the sum of two functions

𝙷2(v,z,z~)=𝙷2(v,z)+𝙷~2(v,z~),\displaystyle\mathtt{H}_{2}(v,z,\tilde{z})=\mathtt{H}_{2}(v,z)+\widetilde{\mathtt{H}}_{2}(v,\tilde{z})\,, (3.17)

so that 𝙷2\mathtt{H}_{2} effectively behaves as a 2d2d off-shell scalar field. Nevertheless, as shown below, the self-duality condition will set one of them, specifically 𝙷~2(v,z~)\widetilde{\mathtt{H}}_{2}(v,\tilde{z}), to zero.

The harmonicity of 𝙷2\mathtt{H}_{2} can be recast into the following pde:

𝙷2=hαα˙(λα𝙷~2α˙+λ~α˙𝙷2α),\displaystyle\nabla\mathtt{H}_{2}=h^{\alpha\dot{\alpha}}(\lambda_{\alpha}\tilde{\mathtt{H}}_{2\,\dot{\alpha}}+\tilde{\lambda}_{\dot{\alpha}}\mathtt{H}_{2\,\alpha})\,, (3.18)

where 𝙷2α\mathtt{H}_{2\,\alpha} and 𝙷~2α˙\tilde{\mathtt{H}}_{2\,\dot{\alpha}} are first-order derivatives of 𝙷2\mathtt{H}_{2} in the yy- and y~\tilde{y}-sector, respectively. Furthermore, all derivatives of 𝙷2\mathtt{H}_{2}, i.e. αα˙αα˙𝙷2𝙷2α(k)α˙(k)\nabla^{\alpha\dot{\alpha}}\ldots\nabla^{\alpha\dot{\alpha}}\mathtt{H}_{2}\equiv\mathtt{H}_{2}^{\alpha(k)\,\dot{\alpha}(k)} can be factorized as

𝙷α(k)α˙(k)\displaystyle\mathtt{H}^{\alpha(k)\,\dot{\alpha}(k)} =λα(k)𝙷~2α˙(k)+λ~α˙(k)𝙷2α(k).\displaystyle=\lambda^{\alpha(k)}\widetilde{\mathtt{H}}_{2}^{\dot{\alpha}(k)}+\tilde{\lambda}^{\dot{\alpha}(k)}\mathtt{H}_{2}^{\alpha(k)}\,. (3.19)

This special property of 𝙷2\mathtt{H}_{2} can be used to generate the linear FDA for the spin-2 sector. However, let us first specify the deformed data in (3.7).

Solving deformed Cartan’s structure equations.

Assuming Cα˙(4)=ϕ(0,4)λ~α˙(4)C_{\dot{\alpha}(4)}=\phi_{(0,4)}\tilde{\lambda}_{\dot{\alpha}(4)}, where ϕ(0,4)\phi_{(0,4)} is a scalar function, cf. (3.20e), and plugging the ansatz (3.15) to (3.7), we obtain

[λ𝙷2]\displaystyle[\lambda\,\mathtt{H}_{2}] =0,\displaystyle=0\,, (3.20a)
ϖαα\displaystyle\boldsymbol{\varpi}_{\alpha\alpha} =ϖαα+𝙷2λ|h|λ]λαλα,\displaystyle=\varpi_{\alpha\alpha}+\mathtt{H}_{2}\langle\lambda|h|\lambda]\lambda_{\alpha}\lambda_{\alpha}\,, (3.20b)
𝒆αα˙\displaystyle\boldsymbol{e}_{\alpha\dot{\alpha}} =hαα˙+λ𝙷2λ|h|λ]λαλ~α˙,\displaystyle=h_{\alpha\dot{\alpha}}+\langle\lambda\,\mathtt{H}_{2}\rangle\langle\lambda|h|\lambda]\lambda_{\alpha}\tilde{\lambda}_{\dot{\alpha}}\,, (3.20c)
ϖ~α˙β˙\displaystyle\widetilde{\boldsymbol{\varpi}}_{\dot{\alpha}\dot{\beta}} =ϖ~α˙β˙+λγ(2)𝙷2γ(2)λ|h|λ]λ~α˙λ~α˙,\displaystyle=\widetilde{\varpi}_{\dot{\alpha}\dot{\beta}}+\langle\lambda^{\gamma(2)}\,\mathtt{H}_{2\,\gamma(2)}\rangle\langle\lambda|h|\lambda]\tilde{\lambda}_{\dot{\alpha}}\tilde{\lambda}_{\dot{\alpha}}\,, (3.20d)
Cα˙(4)\displaystyle C_{\dot{\alpha}(4)} =ϕ(0,4)λ~α˙(4),ϕ(0,4)=12λγ(3)𝙷2γ(3),\displaystyle=\phi_{(0,4)}\tilde{\lambda}_{\dot{\alpha}(4)}\,,\qquad\phi_{(0,4)}=\frac{1}{2}\langle\lambda^{\gamma(3)}\,\mathtt{H}_{2\,\gamma(3)}\rangle\,, (3.20e)

where

[λ𝙷2]:=λ~α˙𝙷2α˙,λα(n)𝙷2,α(n):=λα(n)𝙷2α(n),λα(n)λα1λαn.\displaystyle[\lambda\,\mathtt{H}_{2}]:=\tilde{\lambda}^{\dot{\alpha}}\,\mathtt{H}_{2\,\dot{\alpha}}\,,\qquad\langle\lambda^{\alpha(n)}\,\mathtt{H}_{2\,,\alpha(n)}\rangle:=\lambda^{\alpha(n)}\,\mathtt{H}_{2\,\alpha(n)}\,,\qquad\lambda^{\alpha(n)}\equiv\lambda^{\alpha_{1}}\ldots\lambda^{\alpha_{n}}\,. (3.21)

In deriving the above results, the following relations are useful:

hαα˙λ|h|λ]\displaystyle h_{\alpha\dot{\alpha}}\langle\lambda|h|\lambda] =12(Hαβλβλ~α˙+H~α˙β˙λ~β˙λα),\displaystyle=-\frac{1}{2}\Big{(}H_{\alpha\beta}\lambda^{\beta}\tilde{\lambda}_{\dot{\alpha}}+\widetilde{H}_{\dot{\alpha}\dot{\beta}}\tilde{\lambda}^{\dot{\beta}}\lambda_{\alpha}\Big{)}\,, (3.22a)
dλ|h|λ]𝙷2\displaystyle d\langle\lambda|h|\lambda]\,\mathtt{H}_{2} =+12([λ𝙷2]+~λ𝙷2),\displaystyle=+\frac{1}{2}\Big{(}\mathcal{H}[\lambda\,\mathtt{H}_{2}]+\widetilde{\mathcal{H}}\langle\lambda\,\mathtt{H}_{2}\rangle\Big{)}\,, (3.22b)
dλ|h|λ]λα(k)𝙷2α(k)\displaystyle d\langle\lambda|h|\lambda]\langle\lambda^{\alpha(k)}\,\mathtt{H}_{2\,\alpha(k)}\rangle =+12~λα(k+1)𝙷2α(k+1),\displaystyle=+\frac{1}{2}\widetilde{\mathcal{H}}\langle\lambda^{\alpha(k+1)}\,\mathtt{H}_{2\,\alpha(k+1)}\rangle\,, (3.22c)

where

:=Hα(2)λα(2),~:=H~α˙(2)λ~α˙(2).\displaystyle\mathcal{H}:=H_{\alpha(2)}\lambda^{\alpha(2)}\,,\quad\widetilde{\mathcal{H}}:=\widetilde{H}_{\dot{\alpha}(2)}\tilde{\lambda}^{\dot{\alpha}(2)}\,. (3.23)

Observe that (3.20a) implies the vanishing of all higher-order derivatives of 𝙷2\mathtt{H}_{2} in the y~\tilde{y}-sector. Thus, our analysis will focus solely on the derivative of 𝙷2\mathtt{H}_{2} in the yy-sector.

From (3.20), it is easy to see that the self-dual connection takes the form

D=12𝙷2λ|h|λ]λαλαLαα12λγ(2)𝙷2γ(2)λ|h|λ]λ~α˙λ~α˙L~α˙α˙.\displaystyle D=\nabla-\frac{1}{2}\mathtt{H}_{2}\langle\lambda|h|\lambda]\lambda_{\alpha}\lambda_{\alpha}L^{\alpha\alpha}-\frac{1}{2}\langle\lambda^{\gamma(2)}\,\mathtt{H}_{2\,\gamma(2)}\rangle\langle\lambda|h|\lambda]\tilde{\lambda}_{\dot{\alpha}}\tilde{\lambda}_{\dot{\alpha}}\widetilde{L}^{\dot{\alpha}\dot{\alpha}}\,. (3.24)

Remarkably, the principle spinors λα,λ~α˙\lambda_{\alpha},\tilde{\lambda}_{\dot{\alpha}} are again covariantly constant wrt. the deformed connection DD, i.e.

Dλα=0,Dλ~α˙=0.\displaystyle D\lambda_{\alpha}=0\,,\qquad D\tilde{\lambda}_{\dot{\alpha}}=0\,. (3.25)

Therefore, DD only acts non-trivially on the profile function 𝙷2\mathtt{H}_{2} and its derivative thereof, while “effortlessly” passing through (λ,λ~)(\lambda,\tilde{\lambda}). This fact is essential for us to obtain a simple exact solution for chiral HSGRA with the above setting.

Note, however, that

Dμα=𝙷2λ|h|λ]λα,D2μα=0.\displaystyle D\mu^{\alpha}=-\mathtt{H}_{2}\langle\lambda|h|\lambda]\lambda^{\alpha}\,,\qquad D^{2}\mu^{\alpha}=0\,. (3.26)

Thus, unlike (λ,λ~)(\lambda,\tilde{\lambda}), the principal spinors (μ,μ~)(\mu,\tilde{\mu}) are not covariantly constant wrt. DD.

Metric of self-dual pp-wave background.

Remarkably, the SD background metric generated by ωαα\omega_{\alpha\alpha} only slightly deviates from the conventional solutions in GR, cf. [62, 63, 64, 61, 65], by being self-dual and well-defined in Euclidean or split signature.

- Recall that the line element for Brinkmann metric in the conventional GR is

ds2=dudv+𝙷(v,z,z¯)dv2+dzdz¯.\displaystyle ds^{2}=-du\,dv+\mathtt{H}(v,z,\bar{z})dv^{2}+dzd\bar{z}\,. (3.27)

Here, 𝙷\mathtt{H} is the profile function, which is harmonic in (z,z¯)(z,\bar{z})-plane, that generates a generic Petrov type-N spacetime given that a=λαλ~α˙\ell_{a}=\lambda_{\alpha}\tilde{\lambda}_{\dot{\alpha}}. As 𝙷(v,z,z¯)\mathtt{H}(v,z,\bar{z}) is harmonic in the (z,z¯)(z,\bar{z}) plane, we can write

𝙷(v,z,z¯)=f1(v,z)+f2(v,z¯),\displaystyle\mathtt{H}(v,z,\bar{z})=f_{1}(v,z)+f_{2}(v,\bar{z})\,, (3.28)

for generic functions f1(v,z)f_{1}(v,z), f2(v,z¯)f_{2}(v,\bar{z}). For real solutions, then it is necessary that (f2)=f1(f_{2})^{*}=f_{1}.

- In SD setting, one can deduce from the deformed vierbein (3.20c) that

gμνSD=ημν+λ𝙷2μν.\displaystyle g^{SD}_{\mu\nu}=\eta_{\mu\nu}+\langle\lambda\,\mathtt{H}_{2}\rangle\ell_{\mu}\ell_{\nu}\,. (3.29)

That is, the metric is now generated by the first-derivative of 𝙷2\mathtt{H}_{2} in the zz-direction.666The derivative is represented by the operator paap^{a}\nabla_{a}. In such a setting, we have

dsSD2\displaystyle ds^{2}_{SD} =dudv+λ𝙷2dv2+dzdz~,λ𝙷2=j(v,z),\displaystyle=-du\,dv+\langle\lambda\,\mathtt{H}_{2}\rangle dv^{2}+dzd\tilde{z}\,,\qquad\langle\lambda\,\mathtt{H}_{2}\rangle=j(v,z)\,, (3.30a)
ϕ(0,4)\displaystyle\phi_{(0,4)} =λα(3)𝙷2α(3)z3j(v,z).\displaystyle=\langle\lambda^{\alpha(3)}\,\mathtt{H}_{2\,\alpha(3)}\rangle\sim\partial_{z}^{3}j(v,z)\,. (3.30b)

That is, the Brinkmann-like profile function is now dependent solely on (v,z)(v,z), consistent with the self-duality of the solution.

Intriguingly, even in the SD case, it is possible to place multiple pp-waves at different locations viv_{i} along the vv-direction so that the total profile function for the self-dual pp-wave metric can be written as λ𝙷2=iλ𝙷2i(vi)\langle\lambda\,\mathtt{H}_{2}\rangle=\sum_{i}\langle\lambda\,\mathtt{H}_{2}^{i}(v_{i})\rangle.

3.3 Linear chiral-FDA

This subsection studies the linearized FDA for chiral HSGRA. Let us begin by stating our light-cone ansatz proposal for all positive-helicity spinning fields.

3.3.1 Light-cone ansatz for all spin

Following the discussion around (2.9), it is reasonable to consider the following ansatz for all spins

ωα(2s2):\displaystyle\omega_{\alpha(2s-2)}: =λ|h|λ]λα(2s2)𝙷s,\displaystyle=\langle\lambda|h|\lambda]\lambda_{\alpha(2s-2)}\mathtt{H}_{s}\,, (3.31a)
Bα(2s)\displaystyle B^{\alpha(2s)} :=ψsμα(2s),ψs:=λα(2s)Bα(2s),μλ=1,\displaystyle:=\psi_{s}\mu^{\alpha(2s)}\,,\qquad\psi_{s}:=\lambda_{\alpha(2s)}B^{\alpha(2s)}\,,\quad\langle\mu\,\lambda\rangle=1\,, (3.31b)

where 𝙷s\mathtt{H}_{s} is a harmonic scalar function associated with a positive-helicity massless spin-ss field, i.e.

aa𝙷s=0,pap~bab𝙷s=0,\displaystyle\ell^{a}\nabla_{a}\mathtt{H}_{s}=0\,,\qquad p^{a}\tilde{p}^{b}\nabla_{a}\nabla_{b}\mathtt{H}_{s}=0\,, (3.32)

as in the case of 𝙷2\mathtt{H}_{2}, cf. (3.16). Furthermore, for the negative-helicity fields Bα(2s)B^{\alpha(2s)}, we will assume that

μα=0,aaψs=0,pap~babψs=0.\displaystyle\nabla\mu^{\alpha}=0\,,\qquad\ell^{a}\nabla_{a}\psi_{s}=0\,,\qquad p^{a}\tilde{p}^{b}\nabla_{a}\nabla_{b}\psi_{s}=0\,. (3.33)

This guarantees that the free action of the model is (2.7). Note that on the proposed SD pp-wave background

DβBβα(2s1)α˙=0,\displaystyle D_{\beta}{}^{\dot{\alpha}}B^{\beta\alpha(2s-1)}=0\,, (3.34)

if paaψs=0p^{a}\nabla_{a}\psi_{s}=0, or ψ=ψ(v,z~)\psi=\psi(v,\tilde{z}). This is in accordance with the fact that ψs\psi_{s} are harmonic, cf. (3.33), and Bα(2s)B^{\alpha(2s)} are linear fluctuations on the proposed self-dual background. Thus, (3.33) can serve as a suitable ansatz for the negative-helicity fields.

3.3.2 𝙷\mathtt{H}-FDA

Let us now cast the system (3.18) into generating function form. Since we have assumed 𝙷s\mathtt{H}_{s} to be harmonic, cf. (3.32), all 𝙷sα(k)α˙(k)\mathtt{H}_{s}^{\alpha(k)\,\dot{\alpha}(k)} will admit the same factorization form as in (3.19). Then, we can write the factorization condition for 𝙷sα(k)α˙(k)\mathtt{H}_{s}^{\alpha(k)\,\dot{\alpha}(k)} using some generating functions (𝙷s(y),𝙷~s(y~))(\mathtt{H}_{s}(y),\widetilde{\mathtt{H}}_{s}(\tilde{y})) as

𝙷s(y)\displaystyle\nabla\mathtt{H}_{s}(y) =θαα𝙷s(y),θα:=λ~β˙hαβ˙,\displaystyle=\theta^{\alpha}\partial_{\alpha}\mathtt{H}_{s}(y)\,,\qquad\theta^{\alpha}:=\tilde{\lambda}_{\dot{\beta}}h^{\alpha\dot{\beta}}\,, (3.35a)
𝙷~s(y)\displaystyle\nabla\widetilde{\mathtt{H}}_{s}(y) =θ~α˙α˙𝙷~s(y~),θ~α˙:=λβhβα˙.\displaystyle=\tilde{\theta}^{\dot{\alpha}}\partial_{\dot{\alpha}}\widetilde{\mathtt{H}}_{s}(\tilde{y})\,,\qquad\tilde{\theta}^{\dot{\alpha}}:=\lambda_{\beta}h^{\beta\dot{\alpha}}\,. (3.35b)

One can check that the above system satisfies the Bianchi identities

2𝙷s(y)=0,2𝙷~s(y~)=0.\displaystyle\nabla^{2}\mathtt{H}_{s}(y)=0\,,\qquad\nabla^{2}\widetilde{\mathtt{H}}_{s}(\tilde{y})=0\,. (3.36)

On the deformed background sourced by ωαα\omega_{\alpha\alpha}, cf. (3.15), it can be shown that

D𝙷sα(k)\displaystyle D\mathtt{H}_{s}^{\alpha(k)} =θβ𝙷sβα(k)+k𝙷2λ|h|λ]λαλγ𝙷sγα(k1),\displaystyle=\theta_{\beta}\mathtt{H}_{s}^{\beta\alpha(k)}+k\mathtt{H}_{2}\langle\lambda|h|\lambda]\lambda^{\alpha}\lambda_{\gamma}\mathtt{H}_{s}^{\gamma\alpha(k-1)}\,, (3.37a)
D𝙷~sα˙(k)\displaystyle D\widetilde{\mathtt{H}}_{s}^{\dot{\alpha}(k)} =θ~β˙𝙷~sβ˙α˙(k)+k𝙷2λγ(2)𝙷2γ(2)λ|h|λ]λ~α˙λ~γ˙𝙷~sγ˙α˙(k1),\displaystyle=\tilde{\theta}_{\dot{\beta}}\widetilde{\mathtt{H}}_{s}^{\dot{\beta}\dot{\alpha}(k)}+k\mathtt{H}_{2}\langle\lambda^{\gamma(2)}\mathtt{H}_{2\,\gamma(2)}\rangle\langle\lambda|h|\lambda]\tilde{\lambda}^{\dot{\alpha}}\tilde{\lambda}_{\dot{\gamma}}\widetilde{\mathtt{H}}_{s}^{\dot{\gamma}\dot{\alpha}(k-1)}\,, (3.37b)

or in terms of generating function

D𝙷s(y)\displaystyle D\mathtt{H}_{s}(y) =θα(α+λαλyλ)𝙷s(y),\displaystyle=\theta^{\alpha}\big{(}\partial_{\alpha}+\lambda_{\alpha}\langle\lambda\,y\rangle\langle\lambda\,\partial\rangle\big{)}\mathtt{H}_{s}(y)\,,
D𝙷~s(y~)\displaystyle D\widetilde{\mathtt{H}}_{s}(\tilde{y}) =θ~α˙(~α˙+λ~α˙λγ(2)𝙷2γ(2)[λy][λ])𝙷~s(y~).\displaystyle=\tilde{\theta}^{\dot{\alpha}}\big{(}\tilde{\partial}_{\dot{\alpha}}+\tilde{\lambda}_{\dot{\alpha}}\langle\lambda^{\gamma(2)}\mathtt{H}_{2\,\gamma(2)}\rangle[\lambda\,y][\lambda\,\partial]\big{)}\widetilde{\mathtt{H}}_{s}(\tilde{y})\,. (3.38a)

Note that

D2𝙷s(y)=0,D2𝙷~s(y)=0.\displaystyle D^{2}\mathtt{H}_{s}(y)=0\,,\qquad D^{2}\widetilde{\mathtt{H}}_{s}(y)=0\,. (3.39)

Thus, as long as 𝙷s\mathtt{H}_{s} is harmonic, cf. (3.32), the above system can generate all auxiliary scalars which are derivative of 𝙷s\mathtt{H}_{s}.

3.3.3 Linear FDA for spin-2 sector

Let us now consider the FDA for the harmonic spin-2 solution. As observed earlier in (3.20a), where [λ𝙷2]=0[\lambda\,\mathtt{H}_{2}]=0 due to self-duality constraint, it is natural to impose

𝙷~2α˙(k1)=0,\displaystyle\widetilde{\mathtt{H}}_{2}^{\dot{\alpha}(k\geq 1)}=0\,, (3.40)

which turns out to be a consistent condition for constructing a SD pp-wave background.

In the spin-2 sector, all data we need to generate to complete a linear-FDA are

{ωαα,ωαα˙,ωα˙α˙}k=0{Cα(k)α˙(k+4)}m=0{Bα(m+4)α˙(m)}.\displaystyle\{\omega_{\alpha\alpha},\omega_{\alpha\dot{\alpha}},\omega_{\dot{\alpha}\dot{\alpha}}\}\bigcup_{k=0}^{\infty}\{C^{\alpha(k)\,\dot{\alpha}(k+4)}\}\bigcup_{m=0}^{\infty}\{B^{\alpha(m+4)\,\dot{\alpha}(m)}\}\,. (3.41)

Using (3.37), and covariant constancy of (λ,λ~)(\lambda,\tilde{\lambda}) wrt. DD, cf. (3.25), we obtain

Cα(k)α˙(4+k)=12𝙷2α(3+k)λα(3)λ~α˙(4+k)\displaystyle C^{\alpha(k)\,\dot{\alpha}(4+k)}=\frac{1}{2}\langle\mathtt{H}_{2}^{\alpha(3+k)}\lambda_{\alpha(3)}\rangle\tilde{\lambda}^{\dot{\alpha}(4+k)}\, (3.42)

from solving iteratively Cα(k)α˙(4+k)=hββ˙Cα(k)βα˙(4+k)β˙\nabla C_{\alpha(k)\,\dot{\alpha}(4+k)}=h^{\beta\dot{\beta}}C_{\alpha(k)\beta\,\dot{\alpha}(4+k)\dot{\beta}}. The solution can be written in terms of generating function as

C4(y)=1(4+Ny)![yλ]4+Nyλ3𝙷2(y)|y=y,[yλ]=y~α˙λ~α˙,Ny=yαα.\displaystyle C_{4}(y)=\frac{1}{(4+N_{y})!}[y\,\lambda]^{4+N_{y}}\langle\lambda\,\partial^{\prime}\rangle^{3}\mathtt{H}_{2}(y^{\prime})\Big{|}_{y^{\prime}=y}\,,\qquad[y\,\lambda]=\tilde{y}^{\dot{\alpha}}\tilde{\lambda}_{\dot{\alpha}}\,,\quad N_{y}=y^{\alpha}\partial_{\alpha}\,. (3.43)

For the negative-helicity auxiliary fields Bα(m+4)α˙(m)B^{\alpha(m+4)\,\dot{\alpha}(m)}, we observe that Bα(4+m)α˙(m)B^{\alpha(4+m)\,\dot{\alpha}(m)} for m1m\geq 1 is non-linear in 𝙷2\mathtt{H}_{2} and their derivatives even for the case ψ2=const\psi_{2}=const. This is due to the fact that Dμα0D\mu^{\alpha}\neq 0 when the deformation sourced by ωαα\omega_{\alpha\alpha} occurs. Although we cannot find a closed form for Bα(m+4)α˙(m)B^{\alpha(m+4)\,\dot{\alpha}(m)}, we can still prove that the set of linearized data (3.41) can solve the non-linear EOMs for chiral HSGRA, cf. (2.3). Therefore, they provide us exact solutions for the spin-2 sector.

3.3.4 Linear FDA for all spins

Proceeding analogously with the case of spin-2, we will assume that

𝙷~sα˙(k1)=0,\displaystyle\widetilde{\mathtt{H}}_{s}^{\dot{\alpha}(k\geq 1)}=0\,, (3.44)

and find the following linear data

s{j=02s2{ωα(2s2j)α˙(j)}k=0{Cα(k)α˙(2s+k)}m=0{Bα(2s+m)α˙(m)}}.\displaystyle\bigcup_{s}\left\{\bigcup_{j=0}^{2s-2}\{\omega_{\alpha(2s-2-j)\,\dot{\alpha}(j)}\}\bigcup_{k=0}^{\infty}\{C^{\alpha(k)\,\dot{\alpha}(2s+k)}\}\bigcup_{m=0}^{\infty}\{B^{\alpha(2s+m)\,\dot{\alpha}(m)}\}\right\}\,. (3.45)
one-form sector.

In analogy with the spin-2 sector, we start with the light-cone ansatz (3.31a), solve iteratively the FDA equation ωα(2s2k)α˙(k)=hαωα(2s3k)α˙(k)β˙β\nabla\omega_{\alpha(2s-2-k)\,\dot{\alpha}(k)}=h_{\alpha}{}^{\beta}\omega_{\alpha(2s-3-k)\,\dot{\alpha}(k)\dot{\beta}} and obtain

ωα(2s2k)α˙(k)=λγ(k)𝙷sγ(k)λ|h|λ]λα(2s2k)λ~α˙(k),0k2s2.\displaystyle\omega_{\alpha(2s-2-k)\,\dot{\alpha}(k)}=\langle\lambda^{\gamma(k)}\,\mathtt{H}_{s\,\gamma(k)}\rangle\langle\lambda|h|\lambda]\lambda_{\alpha(2s-2-k)}\tilde{\lambda}_{\dot{\alpha}(k)}\,,\qquad 0\leq k\leq 2s-2\,. (3.46)

This can be written in terms of generating function as

ωs=λ|h|λ]Γ[2s1](yλ+[yλ]λ)2s2𝙷s(y)|y=0.\displaystyle\omega_{s}=\frac{\langle\lambda|h|\lambda]}{\Gamma[2s-1]}\Big{(}\langle y\,\lambda\rangle+[y\,\lambda]\langle\lambda\,\partial^{\prime}\rangle\Big{)}^{2s-2}\mathtt{H}_{s}(y^{\prime})\Big{|}_{y^{\prime}=0}\,. (3.47)

Here, we keep 𝙷s3𝙷2\mathtt{H}_{s\geq 3}\neq\mathtt{H}_{2} for generality. However, it is intriguing noting that when {𝙷s}=𝙷2\{\mathtt{H}_{s}\}=\mathtt{H}_{2} for s\forall s, we can subsume the full generating function ω\omega of the one-form sector as

ω|𝙷s=𝙷2=sωs1|𝙷s=𝙷2=cosh(yλ+[yλ]λ)𝙷2(y)|y=0.\displaystyle\omega\big{|}_{\mathtt{H}_{s}=\mathtt{H}_{2}}=\sum_{s}\omega_{s\geq 1}\big{|}_{\mathtt{H}_{s}=\mathtt{H}_{2}}=\cosh\Big{(}\langle y\,\lambda\rangle+[y\,\lambda]\langle\lambda\,\partial^{\prime}\rangle\Big{)}\mathtt{H}_{2}(y^{\prime})\Big{|}_{y^{\prime}=0}\,. (3.48)
zero-form sector.

Similarly to the spin-2 case, we obtain the generating function for a spin-ss 0-form field CC as

Cα(k)α˙(2s+k)=12𝙷2α(2s1+k)λα(2s1)λ~α˙(2s+k)\displaystyle C^{\alpha(k)\,\dot{\alpha}(2s+k)}=\frac{1}{2}\langle\mathtt{H}_{2}^{\alpha(2s-1+k)}\lambda_{\alpha(2s-1)}\rangle\tilde{\lambda}^{\dot{\alpha}(2s+k)}\, (3.49)

or

Cs(y)=1(2s+Ny)![yλ]2s+Nyλ2s1𝙷s(y)|y=y,[yλ]=y~α˙λ~α˙,Ny=yαα.\displaystyle C_{s}(y)=\frac{1}{(2s+N_{y})!}[y\,\lambda]^{2s+N_{y}}\langle\lambda\,\partial^{\prime}\rangle^{2s-1}\mathtt{H}_{s}(y^{\prime})\Big{|}_{y^{\prime}=y}\,,\qquad[y\,\lambda]=\tilde{y}^{\dot{\alpha}}\tilde{\lambda}_{\dot{\alpha}}\,,\quad N_{y}=y^{\alpha}\partial_{\alpha}\,. (3.50)

As before, Bα(2s+m)α˙(m)B^{\alpha(2s+m)\,\dot{\alpha}(m)} for m1m\geq 1 is non-linear in 𝙷2\mathtt{H}_{2} and their derivatives. However, we observe in general that

Bs=Bs(yλ,[yλ],yμ,𝙷2,ψs).\displaystyle B_{s}=B_{s}(\langle y\,\lambda\rangle,[y\,\lambda],\langle y\,\mu\rangle,\mathtt{H}_{2},\psi_{s})\,. (3.51)

This fact is sufficient to prove that the set in (3.45), whose 1-form subset consists of ωs\omega_{s} generating functions (cf. (3.47)), and the 0-form subset consists of 𝖢s=(Cs,Bs)\mathsf{C}_{s}=(C_{s},B_{s}) fields, indeed solves the full non-linear equations of motion of chiral HSGRA.

Plane-wave solutions.

In the case where [66]

𝙷(y)=exp(±xαα˙καλ~α˙+yκ),λκ=1,\displaystyle\mathtt{H}(y)=\exp\Big{(}\pm x^{\alpha\dot{\alpha}}\kappa_{\alpha}\tilde{\lambda}_{\dot{\alpha}}+\langle y\,\kappa\rangle\Big{)}\,,\qquad\langle\lambda\,\kappa\rangle=1\,, (3.52)

where κ\kappa is a spinor associated with the momentum of a plane-wave. Then, it can be computed from (3.50) and (3.47) that

C\displaystyle C exp(±xαα˙καλ~α˙+[yλ]),\displaystyle\sim\exp\Big{(}\pm x^{\alpha\dot{\alpha}}\kappa_{\alpha}\tilde{\lambda}_{\dot{\alpha}}+[y\,\lambda]\Big{)}\,, (3.53a)
ω\displaystyle\omega λ|h|λ]exp(±xαα˙καλ~α˙+yλ+[yλ]).\displaystyle\sim\langle\lambda|h|\lambda]\exp\Big{(}\pm x^{\alpha\dot{\alpha}}\kappa_{\alpha}\tilde{\lambda}_{\dot{\alpha}}+\langle y\,\lambda\rangle+[y\,\lambda]\Big{)}\,. (3.53b)

These represent a family of plane-wave solutions for chiral HSGRA.

3.4 Non-linear chiral-FDA

Typically, when finding solutions for the non-linear equations of motion (2.3), one often consider the following expansion of fields:

ω\displaystyle\omega =ω(1)+ω(2)+ω(3)+=𝛀+ω¯+ω(3),\displaystyle=\omega^{(1)}+\omega^{(2)}+\omega^{(3)}+\ldots=\boldsymbol{\Omega}+\underline{\omega}+\omega^{(3)}\ldots\,, (3.54a)
𝖢\displaystyle\mathsf{C} =𝖢(1)+𝖢(2)+𝖢(3)+=𝖢¯+𝖢(2)+𝖢(3)+,\displaystyle=\mathsf{C}^{(1)}+\mathsf{C}^{(2)}+\mathsf{C}^{(3)}+\ldots=\underline{\mathsf{C}}+\mathsf{C}^{(2)}+\mathsf{C}^{(3)}+\ldots\,, (3.54b)

where 𝛀=12ϖαβLαβ+𝒆αα˙Pαα˙+12ϖ~α˙β˙L~α˙β˙,\boldsymbol{\Omega}=\frac{1}{2}\boldsymbol{\varpi}_{\alpha\beta}L^{\alpha\beta}+\boldsymbol{e}_{\alpha\dot{\alpha}}P^{\alpha\dot{\alpha}}+\frac{1}{2}\widetilde{\boldsymbol{\varpi}}_{\dot{\alpha}\dot{\beta}}\widetilde{L}^{\dot{\alpha}\dot{\beta}}, and solve (2.3) order by order as in e.g. [67, 68]. However, it is not always the case one can solve for (ω,𝖢)(\omega,\mathsf{C}) to all orders and sum them up, since the computations quickly become quite involved.

Remarkably, it turns out that the 0-form 𝖢=(C,B)\mathsf{C}=(C,B) in (3.50) and the 11-form ω\omega in (3.48) solve the full non-linear equations of motion of chiral HSGRA. Thus, they form a family of exact solutions in chiral HSGRA. Note that we have denoted ω(2)ω¯=(3.47)\omega^{(2)}\equiv\underline{\omega}=\eqref{eq:omega-s-generating} as well as 𝖢(1)=𝖢¯=(C¯,B¯)\mathsf{C}^{(1)}=\underline{\mathsf{C}}=(\underline{C},\underline{B}), cf. (3.50).

Free equations of motion.

As a consistency check, let us first derive free EOMs from (2.3). Observe that for any generating function f(y,y~)f(y,\tilde{y}), we have

f𝛀|Λ=0=(1+[12])ey1+y2+[y1]+[y2]f(y1,y~1)𝛀(y2,y~2)|yi=0y~i=0.\displaystyle f\star\boldsymbol{\Omega}\Big{|}_{\Lambda=0}=(1+[\partial_{1}\,\partial_{2}])e^{\langle y\,\partial_{1}\rangle+\langle y\,\partial_{2}\rangle+[y\,\partial_{1}]+[y\,\partial_{2}]}f(y_{1},\tilde{y}_{1})\boldsymbol{\Omega}(y_{2},\tilde{y}_{2})\Big{|}_{\begin{subarray}{c}y_{i}=0\\ \tilde{y}_{i}=0\end{subarray}}\,. (3.55)

Thus, at linear order in fluctuation, we have

dω¯\displaystyle d\underline{\omega} =𝒱(𝛀,ω¯)+𝒱(𝛀,𝛀,𝖢¯),\displaystyle=\mathcal{V}(\boldsymbol{\Omega},\underline{\omega})+\mathcal{V}(\boldsymbol{\Omega},\boldsymbol{\Omega},\underline{\mathsf{C}})\,, (3.56a)
d𝖢¯\displaystyle d\underline{\mathsf{C}} =𝒰(𝛀,𝖢¯),\displaystyle=\mathcal{U}(\boldsymbol{\Omega},\underline{\mathsf{C}})\,, (3.56b)

where 𝒰(ω,𝖢¯)=Uω,𝖢¯+U𝖢¯,ω\mathcal{U}(\omega,\underline{\mathsf{C}})=U_{\omega,\underline{\mathsf{C}}}+U_{\underline{\mathsf{C}},\omega} and

Uω,𝖢¯\displaystyle U_{\omega,\underline{\mathsf{C}}} =+122exp(y2+12)ω1(y1,y~1)𝖢¯2(y2,y~2),\displaystyle=+\frac{1}{2}\mho_{2}\exp\Big{(}\langle y\,\partial_{2}\rangle+\langle\partial_{1}\,\partial_{2}\rangle\Big{)}\omega_{1}(y_{1},\tilde{y}_{1})\underline{\mathsf{C}}_{2}(y_{2},\tilde{y}_{2})\,, (3.57a)
U𝖢¯,ω\displaystyle U_{\underline{\mathsf{C}},\omega} =122exp(y112)𝖢¯1(y1,y~1)ω2(y2,y~2),\displaystyle=-\frac{1}{2}\mho_{2}\exp\Big{(}\langle y\,\partial_{1}\rangle-\langle\partial_{1}\,\partial_{2}\rangle\Big{)}\underline{\mathsf{C}}_{1}(y_{1},\tilde{y}_{1})\omega_{2}(y_{2},\tilde{y}_{2})\,, (3.57b)

can be obtained from the natural pairings,777recall that 2\mho_{2} is defined in (2.20). cf. Table 1. It can be then checked that (3.56) reproduces (2.10).

(ω¯,𝖢¯)(\underline{\omega},\underline{\mathsf{C}}) are exact solutions.

It is useful noticing that as 𝒱\mathcal{V}’s and 𝒰\mathcal{U}’s vertices are constructed from the set of {yi,ij,[yi],[ij]}\left\{\langle y\,\partial_{i}\rangle\,,\langle\partial_{i}\,\partial_{j}\rangle\,,[y\,\partial_{i}]\,,[\partial_{i}\,\partial_{j}]\right\} operators, the building blocks for fields will not change, i.e. fields are always functions of {yλ,[yλ],yμ,λ,𝙷s,ψs}\{\langle y\,\lambda\rangle,[y\,\lambda],\langle y\,\mu\rangle,\langle\lambda\,\partial^{\prime}\rangle,\mathtt{H}_{s},\psi_{s}\}.

Proposition 3.1.

Let

𝛀and𝖢¯=𝖢¯(yλ,yμ,[yλ],Ny,λ,𝙷2,ψ2)\boldsymbol{\Omega}\quad\text{and}\quad\underline{\mathsf{C}}=\underline{\mathsf{C}}(\langle y\,\lambda\rangle,\langle y\,\mu\rangle,[y\,\lambda],N_{y},\langle\lambda\,\partial^{\prime}\rangle,\mathtt{H}_{2},\psi_{2})

be the field content of the spin-2 sector. Then, all higher-order vertices 𝒱n4(𝛀,𝛀,𝖢¯,,𝖢¯)\mathcal{V}_{n\geq 4}(\boldsymbol{\Omega},\boldsymbol{\Omega},\underline{\mathsf{C}},\ldots,\underline{\mathsf{C}}) and 𝒰n3(𝛀,𝖢¯,,𝖢¯)\mathcal{U}_{n\geq 3}(\boldsymbol{\Omega},\underline{\mathsf{C}},\ldots,\underline{\mathsf{C}}) vanish. Thus, the spin-2 sector characterized by (𝛀,𝖢¯)(\boldsymbol{\Omega},\underline{\mathsf{C}}) forms a family of exact solutions in chiral HSGRA.

Proof.

Let us first consider 𝒱n4(𝛀,𝛀,𝖢¯,,𝖢¯)\mathcal{V}_{n\geq 4}(\boldsymbol{\Omega},\boldsymbol{\Omega},\underline{\mathsf{C}},\ldots,\underline{\mathsf{C}}). From (2.18), it is easy to notice that there are forced contractions between the background connection 1-form 𝛀\boldsymbol{\Omega}’s in the yy-sector via the operations ijn2\langle\partial_{i}\,\partial_{j}\rangle^{n-2} with nn being the number of fields entering the sub-vertices VV. It is easy to notice that ijn2𝛀i𝛀j=0\langle\partial_{i}\,\partial_{j}\rangle^{n-2}\boldsymbol{\Omega}_{i}\boldsymbol{\Omega}_{j}=0 whenever n>4n>4. For the case n=4n=4, the contraction ij2\langle\partial_{i}\,\partial_{j}\rangle^{2} give888The \wedge-product is suppressed temporally for simplicity.

ϖααϖαα=(ϖ+ω)αα(ϖ+ω)αα=0.\displaystyle\boldsymbol{\varpi}^{\alpha\alpha}\boldsymbol{\varpi}_{\alpha\alpha}=(\varpi+\omega)^{\alpha\alpha}(\varpi+\omega)_{\alpha\alpha}=0\,. (3.58)

Therefore, 𝒱n4(𝛀,𝛀,𝖢¯,,𝖢¯)=0\mathcal{V}_{n\geq 4}(\boldsymbol{\Omega},\boldsymbol{\Omega},\underline{\mathsf{C}},\ldots,\underline{\mathsf{C}})=0. Note that the cubic vertex 𝒱(𝛀,𝛀,𝖢¯)\mathcal{V}(\boldsymbol{\Omega},\boldsymbol{\Omega},\underline{\mathsf{C}}) is needed for consistency as it produces the SD components of the Weyl tensor cf. (3.7c).

Next, we turn our focus to 𝒰n3(𝛀,𝖢¯,,𝖢¯)\mathcal{U}_{n\geq 3}(\boldsymbol{\Omega},\underline{\mathsf{C}},\ldots,\underline{\mathsf{C}}) vertices, which can be obtained via the 𝒰\mathcal{U}-𝒱\mathcal{V} duality maps, cf. (B.9). For n>5n>5, each sub-vertex UU consists of a forced contraction of type yin2\langle y\,\partial_{i}\rangle^{n-2} where ii is the position of 𝛀\boldsymbol{\Omega} in U𝖢¯,,𝖢¯,𝛀i,,𝖢¯U_{\underline{\mathsf{C}},\ldots,\underline{\mathsf{C}},\boldsymbol{\Omega}_{i},\ldots,\underline{\mathsf{C}}}. These sub-vertices clearly vanish for n5n\geq 5. Therefore, we only need to pay special attention to the cases where n=3,4n=3,4.

\bullet For n=3n=3, we have 𝒰(𝛀,𝖢¯,𝖢¯)=U𝛀,𝖢¯,𝖢¯+U𝖢¯,𝛀,𝖢¯+U𝖢¯,𝖢¯,𝛀\mathcal{U}(\boldsymbol{\Omega},\underline{\mathsf{C}},\underline{\mathsf{C}})=U_{\boldsymbol{\Omega},\underline{\mathsf{C}},\underline{\mathsf{C}}}+U_{\underline{\mathsf{C}},\boldsymbol{\Omega},\underline{\mathsf{C}}}+U_{\underline{\mathsf{C}},\underline{\mathsf{C}},\boldsymbol{\Omega}} where [51]

U𝛀,𝖢¯,𝖢¯\displaystyle U_{\boldsymbol{\Omega},\underline{\mathsf{C}},\underline{\mathsf{C}}} =+3y1e+(1u)y3+(1v)13+uy2+v12,\displaystyle=+\mho_{3}\langle y\,\partial_{1}\rangle\int e^{+(1-u)\langle y\,\partial_{3}\rangle+(1-v)\langle\partial_{1}\,\partial_{3}\rangle+u\langle y\,\partial_{2}\rangle+v\langle\partial_{1}\,\partial_{2}\rangle}\,, (3.59a)
U𝖢¯,𝛀,𝖢¯\displaystyle U_{\underline{\mathsf{C}},\boldsymbol{\Omega},\underline{\mathsf{C}}} =3y2e(1u)12+(1v)y1+u23+vy3+(vu),\displaystyle=-\mho_{3}\langle y\,\partial_{2}\rangle\int e^{-(1-u)\langle\partial_{1}\,\partial_{2}\rangle+(1-v)\langle y\,\partial_{1}\rangle+u\langle\partial_{2}\,\partial_{3}\rangle+v\langle y\,\partial_{3}\rangle}+(v\leftrightarrow u)\,, (3.59b)
U𝖢¯,𝖢¯,𝛀\displaystyle U_{\underline{\mathsf{C}},\underline{\mathsf{C}},\boldsymbol{\Omega}} =+3y3e(1v)13+(1u)y1v23+uy2.\displaystyle=+\mho_{3}\langle y\,\partial_{3}\rangle\int e^{-(1-v)\langle\partial_{1}\,\partial_{3}\rangle+(1-u)\langle y\,\partial_{1}\rangle-v\langle\partial_{2}\,\partial_{3}\rangle+u\langle y\,\partial_{2}\rangle}\,. (3.59c)

Here, 3=exp([y1]+[y2]+[y3]+[12]+[13]+[23])\mho_{3}=\exp\big{(}[y\,\partial_{1}]+[y\,\partial_{2}]+[y\,\partial_{3}]+[\partial_{1}\,\partial_{2}]+[\partial_{1}\,\partial_{3}]+[\partial_{2}\,\partial_{3}]\big{)} (cf., (2.20)). We notice that there will be no contraction between 𝖢¯\underline{\mathsf{C}}’s in the y~\tilde{y} sector because the \star-product (2.2.2) acts trivially on 𝖢¯i\underline{\mathsf{C}}_{i} due to contractions between the principal spinors λ~\tilde{\lambda}’s.999This is how one imposes locality into the formally consistent FDA (2.3) making it becomes a non-trivial system that describes chiral HSGRA, cf. [41, 42, 51]. Note that the prefactors yi\langle y\,\partial_{i}\rangle in each of the UU sub-vertices act on 𝛀\boldsymbol{\Omega}, so only ϖαα\boldsymbol{\varpi}_{\alpha\alpha} and 𝒆αα˙\boldsymbol{e}_{\alpha\dot{\alpha}} will give non-trivial contributions.

- ϖ\boldsymbol{\varpi}’s contributions. In this case, 3\mho_{3} acts trivially. Thus, all UU’s vertices reduce to

Uϖ,𝖢¯,𝖢¯\displaystyle U_{\boldsymbol{\varpi},\underline{\mathsf{C}},\underline{\mathsf{C}}} =+e+(1u)y3+(1v)13+uy2+v12(ϖ1ααyαy1α,𝖢¯2(y2,y~),𝖢¯3(y3,y~)),\displaystyle=+\int e^{+(1-u)\langle y\,\partial_{3}\rangle+(1-v)\langle\partial_{1}\,\partial_{3}\rangle+u\langle y\,\partial_{2}\rangle+v\langle\partial_{1}\,\partial_{2}\rangle}(\boldsymbol{\varpi}_{1\alpha\alpha}y^{\alpha}y_{1}^{\alpha},\underline{\mathsf{C}}_{2}(y_{2},\tilde{y}),\underline{\mathsf{C}}_{3}(y_{3},\tilde{y}))\,,
U𝖢¯,ϖ,𝖢¯\displaystyle U_{\underline{\mathsf{C}},\boldsymbol{\varpi},\underline{\mathsf{C}}} =e(1u)12+(1v)y1+u23+vy3(𝖢¯1(y1,y~),ϖ2ααyαy2α,𝖢¯3(y3,y~))\displaystyle=-\int e^{-(1-u)\langle\partial_{1}\,\partial_{2}\rangle+(1-v)\langle y\,\partial_{1}\rangle+u\langle\partial_{2}\,\partial_{3}\rangle+v\langle y\,\partial_{3}\rangle}(\underline{\mathsf{C}}_{1}(y_{1},\tilde{y}),\boldsymbol{\varpi}_{2\alpha\alpha}y^{\alpha}y_{2}^{\alpha},\underline{\mathsf{C}}_{3}(y_{3},\tilde{y}))
e(1v)12+(1u)y1+v23+uy3(𝖢¯1(y1,y~),ϖ2ααyαy2α,𝖢¯3(y3,y~)),\displaystyle\quad-\int e^{-(1-v)\langle\partial_{1}\,\partial_{2}\rangle+(1-u)\langle y\,\partial_{1}\rangle+v\langle\partial_{2}\,\partial_{3}\rangle+u\langle y\,\partial_{3}\rangle}(\underline{\mathsf{C}}_{1}(y_{1},\tilde{y}),\boldsymbol{\varpi}_{2\alpha\alpha}y^{\alpha}y_{2}^{\alpha},\underline{\mathsf{C}}_{3}(y_{3},\tilde{y}))\,,
U𝖢¯,𝖢¯,ϖ\displaystyle U_{\underline{\mathsf{C}},\underline{\mathsf{C}},\boldsymbol{\varpi}} =+e(1v)13+(1u)y1v23+uy2(𝖢¯1(y1,y~),𝖢¯2(y2,y~),ϖ3ααyαy3α).\displaystyle=+\int e^{-(1-v)\langle\partial_{1}\,\partial_{3}\rangle+(1-u)\langle y\,\partial_{1}\rangle-v\langle\partial_{2}\,\partial_{3}\rangle+u\langle y\,\partial_{2}\rangle}(\underline{\mathsf{C}}_{1}(y_{1},\tilde{y}),\underline{\mathsf{C}}_{2}(y_{2},\tilde{y}),\boldsymbol{\varpi}_{3\alpha\alpha}y^{\alpha}y_{3}^{\alpha})\,.

To see the pattern of cancellation, we can bring ϖi\boldsymbol{\varpi}_{i} in to the first position in UU. Since CC’s are 0-forms, the act of swapping positions of fields only provides minus signs in the exponential. Let the canonical ordering of fields in 𝒰\mathcal{U} be ϖ,𝖢¯,𝖢¯\boldsymbol{\varpi},\underline{\mathsf{C}},\underline{\mathsf{C}}, we get

𝒰(ϖ,𝖢¯,𝖢¯)=\displaystyle\mathcal{U}(\boldsymbol{\varpi},\underline{\mathsf{C}},\underline{\mathsf{C}})= +e(1u)y3+uy2[2+(1v)13+v12](ϖ1,𝖢¯2,𝖢¯3)\displaystyle+\int e^{(1-u)\langle y\,\partial_{3}\rangle+u\langle y\,\partial_{2}\rangle}\big{[}2+(1-v)\langle\partial_{1}\,\partial_{3}\rangle+v\langle\partial_{1}\,\partial_{2}\rangle\big{]}(\boldsymbol{\varpi}_{1},\underline{\mathsf{C}}_{2},\underline{\mathsf{C}}_{3})
e(1v)y1+vy3[2+(1u)21+u23](ϖ2,𝖢¯1,𝖢¯3)\displaystyle-\int e^{(1-v)\langle y\,\partial_{1}\rangle+v\langle y\,\partial_{3}\rangle}\big{[}2+(1-u)\langle\partial_{2}\,\partial_{1}\rangle+u\langle\partial_{2}\,\partial_{3}\rangle\big{]}(\boldsymbol{\varpi}_{2},\underline{\mathsf{C}}_{1},\underline{\mathsf{C}}_{3})
e(1u)y1+uy3[2+(1v)21+v23](ϖ2,𝖢¯1,𝖢¯3)\displaystyle-\int e^{(1-u)\langle y\,\partial_{1}\rangle+u\langle y\,\partial_{3}\rangle}\big{[}2+(1-v)\langle\partial_{2}\,\partial_{1}\rangle+v\langle\partial_{2}\,\partial_{3}\rangle\big{]}(\boldsymbol{\varpi}_{2},\underline{\mathsf{C}}_{1},\underline{\mathsf{C}}_{3})
+e(1u)y1+uy2[2+(1v)31+v32](ϖ3,𝖢¯1,𝖢¯2).\displaystyle+\int e^{(1-u)\langle y\,\partial_{1}\rangle+u\langle y\,\partial_{2}\rangle}\big{[}2+(1-v)\langle\partial_{3}\,\partial_{1}\rangle+v\langle\partial_{3}\,\partial_{2}\rangle\big{]}(\boldsymbol{\varpi}_{3},\underline{\mathsf{C}}_{1},\underline{\mathsf{C}}_{2})\,.

By simple change in variables, say uvu\leftrightarrow v in the first line of (3.4), we see that all sub-vertices of 𝒰\mathcal{U} add up to zero since they cancel each other pairwise. We conclude that 𝒰(ϖ,𝖢¯,𝖢¯)=0\mathcal{U}(\boldsymbol{\varpi},\underline{\mathsf{C}},\underline{\mathsf{C}})=0, which is a beautiful consequence of chiral HSGRA being a Lorentz-invariant theory.

- 𝒆\boldsymbol{e} contributions. In the case where 𝒰(𝛀,𝖢¯,𝖢¯)𝒰(𝒆,𝖢¯,𝖢¯)\mathcal{U}(\boldsymbol{\Omega},\underline{\mathsf{C}},\underline{\mathsf{C}})\sim\mathcal{U}(\boldsymbol{e},\underline{\mathsf{C}},\underline{\mathsf{C}}), we obtain

U𝒆,𝖢¯,𝖢¯\displaystyle U_{\boldsymbol{e},\underline{\mathsf{C}},\underline{\mathsf{C}}} =+Ξ3(2+[12]+[13])e+(1u)y3+uy2(𝒆1αα˙yαy~1α˙,𝖢¯2,𝖢¯3),\displaystyle=+\Xi_{3}\big{(}2+[\partial_{1}\,\partial_{2}]+[\partial_{1}\,\partial_{3}]\big{)}\int e^{+(1-u)\langle y\,\partial_{3}\rangle+u\langle y\partial_{2}\rangle}(\boldsymbol{e}_{1\alpha\dot{\alpha}}y^{\alpha}\tilde{y}_{1}^{\dot{\alpha}},\underline{\mathsf{C}}_{2},\underline{\mathsf{C}}_{3})\,,
U𝖢¯,𝒆,𝖢¯\displaystyle U_{\underline{\mathsf{C}},\boldsymbol{e},\underline{\mathsf{C}}} =Ξ3(2[21]+[23])e+(1v)y1+vy3(𝒆2αα˙yαy~2α˙,𝖢¯1,𝖢¯3)+(vu),\displaystyle=-\Xi_{3}\big{(}2-[\partial_{2}\,\partial_{1}]+[\partial_{2}\,\partial_{3}]\big{)}\int e^{+(1-v)\langle y\,\partial_{1}\rangle+v\langle y\partial_{3}\rangle}(\boldsymbol{e}_{2\alpha\dot{\alpha}}y^{\alpha}\tilde{y}_{2}^{\dot{\alpha}},\underline{\mathsf{C}}_{1},\underline{\mathsf{C}}_{3})+(v\leftrightarrow u)\,,
U𝖢¯,𝖢¯,𝒆\displaystyle U_{\underline{\mathsf{C}},\underline{\mathsf{C}},\boldsymbol{e}} =+Ξ3(2[31][32])e+(1u)y1+uy2(𝒆3αα˙yαy~3α˙,𝖢¯1,𝖢¯2).\displaystyle=+\Xi_{3}\big{(}2-[\partial_{3}\,\partial_{1}]-[\partial_{3}\,\partial_{2}]\big{)}\int e^{+(1-u)\langle y\,\partial_{1}\rangle+u\langle y\,\partial_{2}\rangle}(\boldsymbol{e}_{3\alpha\dot{\alpha}}y^{\alpha}\tilde{y}_{3}^{\dot{\alpha}},\underline{\mathsf{C}}_{1},\underline{\mathsf{C}}_{2})\,.

where Ξ3=exp([y1]+[y2]+[y3])\Xi_{3}=\exp\big{(}[y\,\partial_{1}]+[y\,\partial_{2}]+[y\,\partial_{3}]\big{)}. Observe that all U𝒆,𝖢¯,𝖢¯U_{\boldsymbol{e},\underline{\mathsf{C}},\underline{\mathsf{C}}} vertices cancel pairwise.

\bullet For n=4n=4 case, where

𝒰(𝛀,𝖢¯,𝖢¯,𝖢¯)=Uϖ,𝖢¯,𝖢¯,𝖢¯+U𝖢¯,ϖ,𝖢¯,𝖢¯+U𝖢¯,𝖢¯,ϖ,𝖢¯+U𝖢¯,𝖢¯,𝖢¯,ϖ,\displaystyle\mathcal{U}(\boldsymbol{\Omega},\underline{\mathsf{C}},\underline{\mathsf{C}},\underline{\mathsf{C}})=U_{\boldsymbol{\varpi},\underline{\mathsf{C}},\underline{\mathsf{C}},\underline{\mathsf{C}}}+U_{\underline{\mathsf{C}},\boldsymbol{\varpi},\underline{\mathsf{C}},\underline{\mathsf{C}}}+U_{\underline{\mathsf{C}},\underline{\mathsf{C}},\boldsymbol{\varpi},\underline{\mathsf{C}}}+U_{\underline{\mathsf{C}},\underline{\mathsf{C}},\underline{\mathsf{C}},\boldsymbol{\varpi}}\,,

and [51]

Uϖ,𝖢¯,𝖢¯,𝖢¯\displaystyle U_{\boldsymbol{\varpi},\underline{\mathsf{C}},\underline{\mathsf{C}},\underline{\mathsf{C}}} =+y12e(1u1u2)y4+(1v1v2)14+u1y2+u2y3+v112+v213\displaystyle=+\langle y\partial_{1}\rangle^{2}\int e^{(1-u_{1}-u_{2})\langle y\partial_{4}\rangle+(1-v_{1}-v_{2})\langle\partial_{1}\partial_{4}\rangle+u_{1}\langle y\partial_{2}\rangle+u_{2}\langle y\partial_{3}\rangle+v_{1}\langle\partial_{1}\partial_{2}\rangle+v_{2}\langle\partial_{1}\partial_{3}\rangle} (3.62a)
U𝖢¯,ϖ,𝖢¯,𝖢¯\displaystyle U_{\underline{\mathsf{C}},\boldsymbol{\varpi},\underline{\mathsf{C}},\underline{\mathsf{C}}} =y22e(1u1u2)y4+(1v1v2)24+u2y1+u1y3v212+v123\displaystyle=-\langle y\partial_{2}\rangle^{2}\int e^{(1-u_{1}-u_{2})\langle y\partial_{4}\rangle+(1-v_{1}-v_{2})\langle\partial_{2}\partial_{4}\rangle+u_{2}\langle y\partial_{1}\rangle+u_{1}\langle y\partial_{3}\rangle-v_{2}\langle\partial_{1}\partial_{2}\rangle+v_{1}\langle\partial_{2}\partial_{3}\rangle}
y22e(1u1u2)y4+(1v1v2)24+u1y1+u2y3v112+v223\displaystyle\quad-\langle y\partial_{2}\rangle^{2}\int e^{(1-u_{1}-u_{2})\langle y\partial_{4}\rangle+(1-v_{1}-v_{2})\langle\partial_{2}\partial_{4}\rangle+u_{1}\langle y\partial_{1}\rangle+u_{2}\langle y\partial_{3}\rangle-v_{1}\langle\partial_{1}\partial_{2}\rangle+v_{2}\langle\partial_{2}\partial_{3}\rangle}
y22e(1uRvL)y4+(1uLvR)24+vLy1+uRy3uL12+vR23\displaystyle\quad-\langle y\partial_{2}\rangle^{2}\int e^{(1-u^{R}-v^{L})\langle y\partial_{4}\rangle+(1-u^{L}-v^{R})\langle\partial_{2}\partial_{4}\rangle+v^{L}\langle y\partial_{1}\rangle+u^{R}\langle y\partial_{3}\rangle-u^{L}\langle\partial_{1}\partial_{2}\rangle+v^{R}\langle\partial_{2}\partial_{3}\rangle} (3.62b)
U𝖢¯,𝖢¯,ϖ,𝖢¯\displaystyle U_{\underline{\mathsf{C}},\underline{\mathsf{C}},\boldsymbol{\varpi},\underline{\mathsf{C}}} =+y32e(1u1u2)y4+(1v1v2)34+u2y1+u1y2v213v123\displaystyle=+\langle y\partial_{3}\rangle^{2}\int e^{(1-u_{1}-u_{2})\langle y\partial_{4}\rangle+(1-v_{1}-v_{2})\langle\partial_{3}\partial_{4}\rangle+u_{2}\langle y\partial_{1}\rangle+u_{1}\langle y\partial_{2}\rangle-v_{2}\langle\partial_{1}\partial_{3}\rangle-v_{1}\langle\partial_{2}\partial_{3}\rangle}
+y32e(1v1v2)y4+(1u1u2)34+v1y1+v2y2u113u223\displaystyle\quad+\langle y\partial_{3}\rangle^{2}\int e^{(1-v_{1}-v_{2})\langle y\partial_{4}\rangle+(1-u_{1}-u_{2})\langle\partial_{3}\partial_{4}\rangle+v_{1}\langle y\partial_{1}\rangle+v_{2}\langle y\partial_{2}\rangle-u_{1}\langle\partial_{1}\partial_{3}\rangle-u_{2}\langle\partial_{2}\partial_{3}\rangle}
+y32e(1uRvL)y4+(1uLvR)34+vLy1+uRy2uL13vR23\displaystyle\quad+\langle y\partial_{3}\rangle^{2}\int e^{(1-u^{R}-v^{L})\langle y\partial_{4}\rangle+(1-u^{L}-v^{R})\langle\partial_{3}\partial_{4}\rangle+v^{L}\langle y\partial_{1}\rangle+u^{R}\langle y\partial_{2}\rangle-u^{L}\langle\partial_{1}\partial_{3}\rangle-v^{R}\langle\partial_{2}\partial_{3}\rangle} (3.62c)
U𝖢¯,𝖢¯,𝖢¯,ϖ\displaystyle U_{\underline{\mathsf{C}},\underline{\mathsf{C}},\underline{\mathsf{C}},\boldsymbol{\varpi}} =+y42e(1v1v2)14+(1u1u2)y1v224+u2y2v134+u1y3\displaystyle=+\langle y\partial_{4}\rangle^{2}\int e^{-(1-v_{1}-v_{2})\langle\partial_{1}\partial_{4}\rangle+(1-u_{1}-u_{2})\langle y\partial_{1}\rangle-v_{2}\langle\partial_{2}\partial_{4}\rangle+u_{2}\langle y\partial_{2}\rangle-v_{1}\langle\partial_{3}\partial_{4}\rangle+u_{1}\langle y\partial_{3}\rangle} (3.62d)

with (ui,vi,uL,uR,vL,vR)(u_{i},v_{i},u^{L},u^{R},v^{L},v^{R}) are certain coefficients that are not crucial in the followings.

Again, let the canonical ordering of the quartic in the spin-2 sector be (ϖ,𝖢¯,𝖢¯,𝖢¯)(\boldsymbol{\varpi},\underline{\mathsf{C}},\underline{\mathsf{C}},\underline{\mathsf{C}}). We can ignore the pre-factors yi2\langle y\,\partial_{i}\rangle^{2} of each UU’s sub-vertices since they only produce ϖααyα(2)\boldsymbol{\varpi}_{\alpha\alpha}y^{\alpha(2)}. It can then be shown in a simple manner that all UU sub-vertices cancel pairwise after some repositioning.

Therefore, 𝒱n4(𝛀,𝛀,C,,C)=0\mathcal{V}_{n\geq 4}(\boldsymbol{\Omega},\boldsymbol{\Omega},C,\ldots,C)=0 and 𝒰n3(𝛀,C,,C)=0\mathcal{U}_{n\geq 3}(\boldsymbol{\Omega},C,\ldots,C)=0 as claimed. ∎

Let us now state our main results.

Theorem 3.2.

Given ω¯=ω¯([yλ],yλ,λ,𝙷)\underline{\omega}=\underline{\omega}([y\,\lambda],\langle y\,\lambda\rangle,\langle\lambda\,\partial^{\prime}\rangle,\mathtt{H}), then 𝒱n3(ω¯,ω¯,𝖢¯,,𝖢¯)=0\mathcal{V}_{n\geq 3}(\underline{\omega},\underline{\omega},\underline{\mathsf{C}},\ldots,\underline{\mathsf{C}})=0.

Proof.

Recall that 𝒱n(ω¯,ω¯,𝖢¯,,𝖢¯)\mathcal{V}_{n}(\underline{\omega},\underline{\omega},\underline{\mathsf{C}},\ldots,\underline{\mathsf{C}}) vertices have ijn2\langle\partial_{i}\,\partial_{j}\rangle^{n-2} as prefactors where iith and jjth are the positions of ω¯i,ω¯j\underline{\omega}_{i},\underline{\omega}_{j} in the sub-vertices VV’s. As ω¯i\underline{\omega}_{i} are functions of yλi\langle y\,\lambda_{i}\rangle, it is clear that 𝒱n=0\mathcal{V}_{n}=0 for n3n\geq 3. ∎

Theorem 3.3.

For (ω¯,𝖢¯)(\underline{\omega},\underline{\mathsf{C}}) defined as above, all vertices 𝒰n2(ω¯,𝖢¯,,𝖢¯)=0\mathcal{U}_{n\geq 2}(\underline{\omega},\underline{\mathsf{C}},\ldots,\underline{\mathsf{C}})=0.

Proof.

Consider 𝒰n2(ω¯,𝖢¯,,C)\mathcal{U}_{n\geq 2}(\underline{\omega},\underline{\mathsf{C}},\ldots,C) whose sub-vertices UU’s carry yin2\langle y\,\partial_{i}\rangle^{n-2} as pre-factors. Then, all UU’s will have the same pre-factors which are proportional to (yλn2×)(\langle y\,\lambda\rangle^{n-2}\times\ldots). At this stage, we can repeat the proof of Proposition 3.1 where we take the canonical ordering of fields in each UU’s sub-vertices to be (ω¯,𝖢¯,,𝖢¯)(\underline{\omega},\underline{\mathsf{C}},\ldots,\underline{\mathsf{C}}). The claim is that all UU’s vertices will cancel pairwise based on how vertices of chiral HSGRA were designed [41, 42, 51]. This statement can be checked explicitly with 𝒰(ω¯,𝖢¯)=Uω¯,𝖢¯+U𝖢¯,ω¯\mathcal{U}(\underline{\omega},\underline{\mathsf{C}})=U_{\underline{\omega},\underline{\mathsf{C}}}+U_{\underline{\mathsf{C}},\underline{\omega}}, cf. (3.57). It is analogous for the cubic (3.59) (where we replace 𝛀\boldsymbol{\Omega} by ω¯\underline{\omega}), and the quartic (3.62). ∎

With the above results, it is clear that the complete non-linear FDA is solved by (ω¯,𝖢¯)(\underline{\omega},\underline{\mathsf{C}}). Thus, the non-linear FDA solved by (ω¯,𝖢¯)(\underline{\omega},\underline{\mathsf{C}}) truncates to (3.56) as it is not necessary to solve for higher-order fluctuations.

3.5 Effective action for chiral HSGRA on SD pp-wave background

We now show that the effective action associated to the above exact solutions is precisely the usual kinetic action for chiral higher-spin fields on any SD background.

In what follows, we denote

(𝒜):=(dω𝒱(𝛀,ω))|y~=0=d𝒜,\displaystyle\mathcal{F}(\mathcal{A}):=\Big{(}d\omega-\mathcal{V}(\boldsymbol{\Omega},\omega)\Big{)}\Big{|}_{\tilde{y}=0}=d\mathcal{A}\,, (3.63)

to be the ‘effective curvature tensor’ for chiral HSGRA, since we have shown all higher-order vertices vanish on the solutions. Then, the effective action associated to the above field strength reads

SχHSGRASDppwave=𝖢|𝓗d𝒜y,𝓗:=𝒆αα˙𝒆αyαα˙yα.\displaystyle S_{\chi HSGRA}^{SD\,pp-wave}=\int_{\mathcal{M}}\langle\mathsf{C}\,|\,\boldsymbol{\mathcal{H}}\wedge d\mathcal{A}\rangle_{y}\,,\qquad\boldsymbol{\mathcal{H}}:=\boldsymbol{e}_{\alpha\dot{\alpha}}\boldsymbol{e}_{\alpha}{}^{\dot{\alpha}}y^{\alpha}y^{\alpha}\,. (3.64)

where y\langle\,\rangle_{y} stands for the operation of integrating out yy to produce appropriate contractions, cf. [39]. We obtain

S[B,A]=12s=1;𝓗d𝒜y=12s=1gBα(2s)DαAα(2s1)α˙α˙.\displaystyle S[B,A]=\frac{1}{2}\sum_{s=1}^{\infty}\int\left\langle\mathcal{B};\boldsymbol{\mathcal{H}}\wedge d\mathcal{A}\right\rangle_{y}=\frac{1}{2}\sum_{s=1}^{\infty}\int\sqrt{g}B^{\alpha(2s)}D_{\alpha}{}^{\dot{\alpha}}A_{\alpha(2s-1)\,\dot{\alpha}}\,. (3.65)

Thus, the proposed action provides the kinetic terms for all chiral higher-spin fields on a generic SD background, as promised.

4 Discussion

In this paper, we obtain a class of exact self-dual pp-wave solutions, in chiral HSGRA by considering a SD modification of the Kerr-Schild ansatz, cf. (3.15). We then generalize our analysis to the case of higher spins. For completeness, we show that the spacetime action of the obtained solutions reduces to the standard kinetic action for free fields on any SD background.

Given that there are other self-dual solutions such as self-dual Taub-NUT black hole (see e.g. [69, 70, 71]), BPST instantons [72], or ADHM instantons [73], we expect that there will be more exact solutions in chiral HSGRA, which are higher-spin generalizations of the aforementioned self-dual solutions.

The reason we believe there should be more classes of exact solutions in chiral HSGRA is that this theory is integrable in the sense of Bardeen [74], see e.g. [48] and [49] for relevant discussion. Therefore, chiral HSGRA should be solvable. This speculation is backed by the trivialization of the scattering amplitudes of chiral HSGRA in flat space, cf. [75, 76, 77], and some recent twistor construction of the theory or some closed subsectors of it, see e.g. [78, 79, 80, 81].

It will be interesting to check whether all self-dual/chiral higher-spin theories possess this property. Note that, even if some of these theories may have quantum anomalies, as discussed in [49], it is still possible to make them quantum integrable through a Green-Schwarz-like mechanism, as shown in [82, 83]. In particular, by introducing suitable axionic interactions in twistor space [84, 85], one should be able to cure any quantum anomalies that arise in self-dual/chiral theories. A beautiful discussion related to this story in spacetime can also be found, e.g., in [86].

While the analysis of chiral HSGRA is relatively straightforward in Λ=0\Lambda=0 case, it is more intricate in the case of (A)dS, where higher-order vertices will arise, as observed in [18, 19, 42]. (See also [87] for a higher-dimensional and [88] for a double-copy generalization of the theory). This fact makes the task of construction a covariant action for chiral HSGRA difficult. Nevertheless, one may hope that the light-cone ansatz introduced in [46, 47] will provide further insight in guessing the correct action for chiral HSGRA due to the resulting simple form of higher-order vertices after substituting them in. See [89, 90] for some work along this line.

Acknowledgement

Many discussions with Yasha Neiman and Zhenya Skvortsov are gratefully acknowledged. This work is supported by the Young Scientist Training (YST) program at the Asia Pacific Center of Theoretical Physics (APCTP) through the Science and Technology Promotion Fund and Lottery Fund of the Korean Government, and also the Korean Local Governments – Gyeongsangbuk-do Province and Pohang City. The author also appreciates the support from the Fonds de la Recherche Scientifique under grant number F.4503.20 (HighSpinSymm), grant number F.4544.21 (HigherSpinGraWave) and funding from the European Research Council (ERC) under grant number 101002551, when part of this work took place in Mons, Belgium.

Appendix A Convention and useful identities

In the main text, we have raised and lowered spinorial indices with the following convention:

hαα˙μhμββ˙=2ϵαβϵα˙β˙,vα=vβϵβα,vα=vβϵαβ,vα˙=vβ˙ϵβ˙α˙,vα˙=vβ˙ϵα˙β˙.\displaystyle h^{\mu}_{\alpha\dot{\alpha}}h_{\mu\beta\dot{\beta}}=2\epsilon_{\alpha\beta}\epsilon_{\dot{\alpha}\dot{\beta}}\,,\quad v_{\alpha}=v^{\beta}\epsilon_{\beta\alpha}\,,\quad v^{\alpha}=v_{\beta}\epsilon^{\alpha\beta}\,,\quad v_{\dot{\alpha}}=v^{\dot{\beta}}\epsilon_{\dot{\beta}\dot{\alpha}}\,,\quad v^{\dot{\alpha}}=v_{\dot{\beta}}\epsilon^{\dot{\alpha}\dot{\beta}}\,. (A.1)

Here, ϵαβ\epsilon^{\alpha\beta} is the 𝔰𝔭(2)\mathfrak{sp}(2)-invariant tensor with the properties ϵαβ=ϵαβ=ϵβα\epsilon^{\alpha\beta}=\epsilon_{\alpha\beta}=-\epsilon^{\beta\alpha} and ϵ01=1\epsilon^{01}=1. Furthermore, the following identities are useful

hαα˙hββ˙\displaystyle h^{\alpha\dot{\alpha}}\wedge h^{\beta\dot{\beta}} =+12Hαβϵα˙β˙+12H~α˙β˙ϵαβ,\displaystyle=+\frac{1}{2}H^{\alpha\beta}\epsilon^{\dot{\alpha}\dot{\beta}}+\frac{1}{2}\widetilde{H}^{\dot{\alpha}\dot{\beta}}\epsilon^{\alpha\beta}\,,\qquad\qquad HααH~α˙α˙\displaystyle H^{\alpha\alpha}\wedge\widetilde{H}^{\dot{\alpha}\dot{\alpha}} =0,\displaystyle=0\,, (A.2a)
HααHαα\displaystyle H_{\alpha\alpha}\wedge H^{\alpha\alpha} =H~α˙α˙H~α˙α˙=Vol,\displaystyle=-\widetilde{H}_{\dot{\alpha}\dot{\alpha}}\wedge\widetilde{H}^{\dot{\alpha}\dot{\alpha}}=\text{Vol}\,, (A.2b)
hαα˙Hββ\displaystyle h^{\alpha\dot{\alpha}}\wedge H^{\beta\beta} =13ϵαβ3βα˙,\displaystyle=-\frac{1}{3}\epsilon^{\alpha\beta}\mathcal{H}_{3}^{\beta\dot{\alpha}}\,,\qquad\qquad hαα˙H~β˙β˙\displaystyle h^{\alpha\dot{\alpha}}\wedge\widetilde{H}^{\dot{\beta}\dot{\beta}} =+13ϵα˙β˙3αβ˙,\displaystyle=+\frac{1}{3}\epsilon^{\dot{\alpha}\dot{\beta}}\mathcal{H}_{3}^{\alpha\dot{\beta}}\,, (A.2c)

where

hαγ˙hαγ˙=Hαα,hγα˙hγα˙=H~α˙α˙,hαα˙Hβ˙α˙=3αβ˙.\displaystyle h^{\alpha}{}_{\dot{\gamma}}\wedge h^{\alpha\dot{\gamma}}=H^{\alpha\alpha}\,,\qquad\qquad h_{\gamma}{}^{\dot{\alpha}}\wedge h^{\gamma\dot{\alpha}}=\widetilde{H}^{\dot{\alpha}\dot{\alpha}}\,,\qquad\qquad h^{\alpha}{}_{\dot{\alpha}}\wedge H^{\dot{\beta}\dot{\alpha}}=\mathcal{H}_{3}^{\alpha\dot{\beta}}\,. (A.3)

Appendix B Strong homotopy algebras

This appendix provides some basic facts about homotopy algebras adapted to the notations used in the main text.

Recall that the QQ-manifold (𝒩,Q)(\mathcal{N},Q) is a \mathbb{Z}-graded target space 𝒩=n𝒩n\mathcal{N}=\bigoplus_{n\in\mathbb{Z}}\mathcal{N}_{n} whose homogeneous elements ana_{n} has degree |an|=n|a_{n}|=n. Have we chosen to work with a degree kk-shifted target space 𝒩\mathcal{N}, the corresponding subspaces would be

𝒩[k]n:=𝒩n+k.\displaystyle\mathcal{N}[k]_{n}:=\mathcal{N}_{n+k}\,. (B.1)

Hereinafter, we use Koszul sign convention where ab=()|a||b|baab=(-)^{|a||b|}ba for two graded objects (a,b)(a,b) of degrees |a||a| and |b||b|, respectively.

A/LA_{\infty}/L_{\infty}-algebras.

Denote C(𝒩)C^{\bullet}(\mathcal{N}) as the space of homomorphisms Hom(T𝒩,𝒩)\mathrm{Hom}_{\mathbb{C}}(T\mathcal{N},\mathcal{N}), where T𝒩:=n=1𝒩nT\mathcal{N}:=\bigoplus_{n=1}^{\infty}\mathcal{N}^{\otimes n} is the tensor algebra associated with 𝒩\mathcal{N}, whose elements 𝖥mCm(𝒩)\mathsf{F}_{m}\in C^{m}(\mathcal{N}) are 𝖥m\mathsf{F}_{m} maps with degree 1. We define the Gerstenhaber \circ-product as

𝖥n𝖥m:=k=0n1()|𝖥m|(|Φ1|++|Φk|)𝖥n(Φ1,,Φk,𝖥m(Φk+1,,Φk+m),,Φm+n1)\displaystyle\mathsf{F}_{n}\circ\mathsf{F}_{m}:=\sum_{k=0}^{n-1}(-)^{|\mathsf{F}_{m}|(|\Phi_{1}|+\ldots+|\Phi_{k}|)}\mathsf{F}_{n}(\Phi_{1},\ldots,\Phi_{k},\mathsf{F}_{m}(\Phi_{k+1},\ldots,\Phi_{k+m}),\ldots,\Phi_{m+n-1})\, (B.2)

where Φ\Phi’s are elements of 𝒩0\mathcal{N}_{0} or 𝒩1\mathcal{N}_{1}.

Definition B.1.

The pair (𝒩,𝖥)(\mathcal{N},\mathsf{F}) where 𝖥=i=1𝖥i\mathsf{F}=\sum_{i=1}^{\infty}\mathsf{F}_{i} is called AA_{\infty}-algebra if

𝖥𝖥=0.\displaystyle\mathsf{F}\circ\mathsf{F}=0\,. (B.3)

Following from the above definition of AA_{\infty}-algebra, it is clear that 𝖥12=0\mathsf{F}_{1}^{2}=0, and therefore 𝖥1\mathsf{F}_{1} defines a differential. The next structure map 𝖥2\mathsf{F}_{2} is a Leibniz-compatible binary operation. Then, the homotopy given by 𝖥3\mathsf{F}_{3} measures the failure of the associativity of 𝖥2\mathsf{F}_{2}, and so on. For more detail, see e.g. [91].

Note that an AA_{\infty}-algebra can induce an LL_{\infty}-algebra defined by a pair (𝒩,𝖫)\mathcal{N},\mathsf{L}), where 𝖫=𝖫i\mathsf{L}=\sum\mathsf{L}_{i} with |𝖫|=1|\mathsf{L}|=1 called LL_{\infty}-structure maps with suitable symmetrization among the arguments, i.e.

𝖫n(Φ1,,Φn)=σSn()K𝖥n(Φσ(1),,Φσ(n)).\displaystyle\mathsf{L}_{n}(\Phi_{1},\ldots,\Phi_{n})=\sum_{\sigma\in S_{n}}(-)^{K}\mathsf{F}_{n}(\Phi_{\sigma(1)},\ldots,\Phi_{\sigma(n)})\,. (B.4)

Here, ()K(-)^{K} is the Koszul sign resulting from σ\sigma permutation under the symmetric group SnS_{n}. For instance,

𝖫n(Φ1,,Φi,Φi+1,,Φn)=()|Φi||Φi+1|𝖫n(Φ1,,Φi+1,Φi,,Φn).\displaystyle\mathsf{L}_{n}(\Phi_{1},\ldots,\Phi_{i},\Phi_{i+1},\ldots,\Phi_{n})=(-)^{|\Phi_{i}||\Phi_{i+1}|}\mathsf{L}_{n}(\Phi_{1},\ldots,\Phi_{i+1},\Phi_{i},\ldots,\Phi_{n})\,. (B.5)

As the reader may notice, 𝖫\mathsf{L} can be identified with the homological vector field QQ on 𝒩\mathcal{N}. Furthermore, following from (B.2) and Definition B.1, 𝖫\mathsf{L}’s must obey:

m+n1=constσSn()K𝖫m(𝖫n(Φσ(1),,Φσ(n)),Φσ(n+1),,Φσ(n+m1))=0.\displaystyle\sum_{m+n-1=\text{const}}\sum_{\sigma\in S_{n}}(-)^{K}\mathsf{L}_{m}(\mathsf{L}_{n}(\Phi_{\sigma(1)},\ldots,\Phi_{\sigma(n)}),\Phi_{\sigma(n+1)},\ldots,\Phi_{\sigma(n+m-1)})=0\,. (B.6)

As usual, 𝖫1d\mathsf{L}_{1}\equiv d is the defining differential on C(𝒩)C^{\bullet}(\mathcal{N}). Note that an LL_{\infty}-algebra with 𝖫n3=0\mathsf{L}_{n\geq 3}=0 is referred to as a dgLa.

The LL_{\infty}-structure maps (𝒱,𝒰)(\mathcal{V},\mathcal{U}) for chiral HSGRA can be obtained by restricting the super coordinates Φ\Phi’s to their appropriate components, i.e. to ω\omega’s or 𝖢\mathsf{C}’s. In particular, when there are strictly two ω\omega arguments in 𝖫\mathsf{L}, then 𝖫(ω,ω,𝖢,,𝖢)𝒱(ω,ω,𝖢,,𝖢)\mathsf{L}(\omega,\omega,\mathsf{C},\ldots,\mathsf{C})\equiv\mathcal{V}(\omega,\omega,\mathsf{C},\ldots,\mathsf{C}), and when there is only one ω\omega we have the identification 𝖫(ω,𝖢,,𝖢)𝒰(ω,𝖢,,𝖢)\mathsf{L}(\omega,\mathsf{C},\ldots,\mathsf{C})\equiv\mathcal{U}(\omega,\mathsf{C},\ldots,\mathsf{C}).

Although 𝒱\mathcal{V}’s and 𝒰\mathcal{U}’s are LL_{\infty}-structure maps, it will be more convenient to work with their sub-vertices which originate from a cyclic AA_{\infty}-algebra where the signs and permutations can be kept in track in a simple manner, cf. [41]. This can be done by letting ω\omega and 𝖢\mathsf{C} take values in some matrix algebra 𝙼𝚊𝚝(n,)\mathtt{Mat}(n,\mathbb{C}) as alluded to in the main text. Furthermore, since 𝒱\mathcal{V}’s and 𝒰\mathcal{U}’s originated from the same LL_{\infty}-algebra, we can relate them together through the following nature pairing.

Definition B.2.

A degree-22 cyclic LL_{\infty}-algebra is an LL_{\infty}-algebra endowed with a non-degenerate \mathbb{C}-linear pairing |:𝒩𝒩[1][2]\langle-|-\rangle:\mathcal{N}\oplus\mathcal{N}^{\vee}[-1]\rightarrow\mathbb{C}[-2] such that

a|b=()(|a|+1)(|b|+1)b|a.\displaystyle\langle a|b\rangle=(-)^{(|a|+1)(|b|+1)}\langle b|a\rangle\,. (B.7)

From the above definition, we have for instance

𝖫n(Φ1,,Φn),Φn+1\displaystyle\langle\mathsf{L}_{n}(\Phi_{1},\ldots,\Phi_{n}),\Phi_{n+1}\rangle =()|Φ1|(k=2n+1|Φk|)+|Φ1|+|Φn+2|𝖫n(Φ2,,Φn+2),Φ1),\displaystyle=(-)^{|\Phi_{1}|\big{(}\sum_{k=2}^{n+1}|\Phi_{k}|\big{)}+|\Phi_{1}|+|\Phi_{n+2}|}\langle\mathsf{L}_{n}(\Phi_{2},\ldots,\Phi_{n+2}),\Phi_{1})\,, (B.8a)
𝖫n(Φ1,,Φn)|Φn+1\displaystyle\langle\mathsf{L}_{n}(\Phi_{1},\ldots,\Phi_{n})|\Phi_{n+1}\rangle =()k=1n|Φk|Φ1|𝖫n(Φ2,,Φn+1),\displaystyle=(-)^{\sum_{k=1}^{n}|\Phi_{k}|}\langle\Phi_{1}|\mathsf{L}_{n}(\Phi_{2},\ldots,\Phi_{n+1})\rangle\,, (B.8b)

Upon restricting Φ\Phi’s to ω\omega and 𝖢\mathsf{C}, we obtain

V(ω,𝖢,,ω,𝖢,,𝖢)|C\displaystyle\big{\langle}V(\omega,\mathsf{C},\ldots,\omega,\mathsf{C},\ldots,\mathsf{C})|C\big{\rangle} =+ω|U(𝖢,,ω,𝖢,,𝖢),\displaystyle=+\big{\langle}\omega|U(\mathsf{C},\ldots,\omega,\mathsf{C},\ldots,\mathsf{C})\big{\rangle}\,, (B.9a)
𝖢|V(𝖢,,ω,𝖢,,𝖢,ω)\displaystyle\big{\langle}\mathsf{C}|V(\mathsf{C},\ldots,\omega,\mathsf{C},\ldots,\mathsf{C},\omega)\big{\rangle} =U(𝖢,,ω,𝖢,,𝖢)|ω,\displaystyle=-\big{\langle}U(\mathsf{C},\ldots,\omega,\mathsf{C},\ldots,\mathsf{C})|\omega\big{\rangle}\,, (B.9b)

where VV’s and UU’s are sub-vertices originated from AA_{\infty}-algebra of 𝒱\mathcal{V} and 𝒰\mathcal{U}. The maps (B.9) are called the 𝒱\mathcal{V}-𝒰\mathcal{U} duality maps [41, 42, 51], which can be used to extract 𝒰\mathcal{U}’s vertices from 𝒱\mathcal{V}’s (at a fixed order nn), and vice versa.

References

  • [1] X. Bekaert, N. Boulanger, A. Campoleoni, M. Chiodaroli, D. Francia, M. Grigoriev, E. Sezgin and E. Skvortsov, Snowmass White Paper: Higher Spin Gravity and Higher Spin Symmetry, 2205.01567.
  • [2] S. Weinberg, Photons and Gravitons in SS-Matrix Theory: Derivation of Charge Conservation and Equality of Gravitational and Inertial Mass, Phys. Rev. 135 (1964) B1049–B1056.
  • [3] S. R. Coleman and J. Mandula, All Possible Symmetries of the S Matrix, Phys. Rev. 159 (1967) 1251–1256.
  • [4] J. Maldacena and A. Zhiboedov, Constraining Conformal Field Theories with A Higher Spin Symmetry, J. Phys. A 46 (2013) 214011 [1112.1016].
  • [5] X. Bekaert, J. Erdmenger, D. Ponomarev and C. Sleight, Quartic AdS Interactions in Higher-Spin Gravity from Conformal Field Theory, JHEP 11 (2015) 149 [1508.04292].
  • [6] C. Sleight and M. Taronna, Higher-Spin Gauge Theories and Bulk Locality, Phys. Rev. Lett. 121 (2018), no. 17 171604 [1704.07859].
  • [7] T. Tran, Constraining higher-spin S-matrices, JHEP 02 (2023) 001 [2212.02540].
  • [8] Y. Neiman, Quartic locality of higher-spin gravity in de Sitter and Euclidean anti-de Sitter space, Phys. Lett. B 843 (2023) 138048 [2302.00852].
  • [9] A. Y. Segal, Conformal higher spin theory, Nucl. Phys. B 664 (2003) 59–130 [hep-th/0207212].
  • [10] A. A. Tseytlin, On limits of superstring in AdS(5) x S**5, Theor. Math. Phys. 133 (2002) 1376–1389 [hep-th/0201112].
  • [11] X. Bekaert, E. Joung and J. Mourad, Effective action in a higher-spin background, JHEP 02 (2011) 048 [1012.2103].
  • [12] P. Hähnel and T. McLoughlin, Conformal higher spin theory and twistor space actions, J. Phys. A 50 (2017), no. 48 485401 [1604.08209].
  • [13] T. Adamo, P. Hähnel and T. McLoughlin, Conformal higher spin scattering amplitudes from twistor space, JHEP 04 (2017) 021 [1611.06200].
  • [14] T. Basile, M. Grigoriev and E. Skvortsov, Covariant action for conformal higher spin gravity, J. Phys. A 56 (2023), no. 38 385402 [2212.10336].
  • [15] R. R. Metsaev, Poincare invariant dynamics of massless higher spins: Fourth order analysis on mass shell, Mod. Phys. Lett. A 6 (1991) 359–367.
  • [16] R. R. Metsaev, S matrix approach to massless higher spins theory. 2: The Case of internal symmetry, Mod. Phys. Lett. A 6 (1991) 2411–2421.
  • [17] D. Ponomarev and E. D. Skvortsov, Light-Front Higher-Spin Theories in Flat Space, J. Phys. A 50 (2017), no. 9 095401 [1609.04655].
  • [18] R. R. Metsaev, Light-cone gauge cubic interaction vertices for massless fields in AdS(4), Nucl. Phys. B 936 (2018) 320–351 [1807.07542].
  • [19] E. Skvortsov, Light-Front Bootstrap for Chern-Simons Matter Theories, JHEP 06 (2019) 058 [1811.12333].
  • [20] R. R. Metsaev, Cubic interaction vertices for N=1 arbitrary spin massless supermultiplets in flat space, JHEP 08 (2019) 130 [1905.11357].
  • [21] R. R. Metsaev, Cubic interactions for arbitrary spin 𝒩\mathcal{N} -extended massless supermultiplets in 4d flat space, JHEP 11 (2019) 084 [1909.05241].
  • [22] M. Tsulaia and D. Weissman, Supersymmetric quantum chiral higher spin gravity, JHEP 12 (2022) 002 [2209.13907].
  • [23] M. P. Blencowe, A Consistent Interacting Massless Higher Spin Field Theory in DD = (2+1), Class. Quant. Grav. 6 (1989) 443.
  • [24] E. Bergshoeff, M. P. Blencowe and K. S. Stelle, Area Preserving Diffeomorphisms and Higher Spin Algebra, Commun. Math. Phys. 128 (1990) 213.
  • [25] A. Campoleoni, S. Fredenhagen, S. Pfenninger and S. Theisen, Asymptotic symmetries of three-dimensional gravity coupled to higher-spin fields, JHEP 11 (2010) 007 [1008.4744].
  • [26] M. Henneaux and S.-J. Rey, Nonlinear WinfinityW_{infinity} as Asymptotic Symmetry of Three-Dimensional Higher Spin Anti-de Sitter Gravity, JHEP 12 (2010) 007 [1008.4579].
  • [27] C. N. Pope and P. K. Townsend, Conformal Higher Spin in (2+1)-dimensions, Phys. Lett. B 225 (1989) 245–250.
  • [28] E. S. Fradkin and V. Y. Linetsky, A Superconformal Theory of Massless Higher Spin Fields in DD = (2+1), Mod. Phys. Lett. A 4 (1989) 731.
  • [29] M. Grigoriev, I. Lovrekovic and E. Skvortsov, New Conformal Higher Spin Gravities in 3d3d, JHEP 01 (2020) 059 [1909.13305].
  • [30] M. Grigoriev, K. Mkrtchyan and E. Skvortsov, Matter-free higher spin gravities in 3D: Partially-massless fields and general structure, Phys. Rev. D 102 (2020), no. 6 066003 [2005.05931].
  • [31] A. Sharapov, E. Skvortsov and A. Sukhanov, Matter-coupled higher spin gravities in 3d: no- and yes-go results, 2409.12830.
  • [32] K. Alkalaev and X. Bekaert, Towards higher-spin AdS2/CFT1 holography, JHEP 04 (2020) 206 [1911.13212].
  • [33] M. Sperling and H. C. Steinacker, Covariant cosmological quantum space-time, higher-spin and gravity in the IKKT matrix model, JHEP 07 (2019) 010 [1901.03522].
  • [34] H. C. Steinacker, Gravity as a quantum effect on quantum space-time, Phys. Lett. B 827 (2022) 136946 [2110.03936].
  • [35] H. C. Steinacker and T. Tran, Soft limit of higher-spin interactions in the IKKT model, 2311.14163.
  • [36] H. C. Steinacker and T. Tran, Quantum hs-Yang-Mills from the IKKT matrix model, Nucl. Phys. B 1005 (2024) 116608 [2405.09804].
  • [37] A. Manta, H. C. Steinacker and T. Tran, 𝔥𝔰\mathfrak{hs}-extended gravity from the IKKT matrix model, 2411.02598.
  • [38] E. Skvortsov and Y. Yin, Low spin solutions of higher spin gravity: BPST instanton, JHEP 07 (2024) 032 [2403.17148].
  • [39] K. Krasnov, E. Skvortsov and T. Tran, Actions for self-dual Higher Spin Gravities, JHEP 08 (2021) 076 [2105.12782].
  • [40] R. Penrose, Twistor algebra, J. Math. Phys. 8 (1967) 345.
  • [41] A. Sharapov, E. Skvortsov, A. Sukhanov and R. Van Dongen, Minimal model of Chiral Higher Spin Gravity, 2205.07794.
  • [42] A. Sharapov and E. Skvortsov, Chiral Higher Spin Gravity in (A)dS4 and secrets of Chern-Simons Matter Theories, 2205.15293.
  • [43] D. Sullivan, Infinitesimal computations in topology, Publications Mathématiques de l’IHÉS 47 (1977) 269–331.
  • [44] M. Alexandrov, A. Schwarz, O. Zaboronsky and M. Kontsevich, The Geometry of the master equation and topological quantum field theory, Int. J. Mod. Phys. A 12 (1997) 1405–1429 [hep-th/9502010].
  • [45] K. Krasnov and E. Skvortsov, Flat Self-dual Gravity, 2106.01397.
  • [46] T. Adamo and T. Tran, Higher-spin Yang-Mills, amplitudes and self-duality, 2210.07130.
  • [47] Y. Neiman, Self-dual gravity in de Sitter space: Light-cone ansatz and static-patch scattering, Phys. Rev. D 109 (2024), no. 2 024039 [2303.17866].
  • [48] D. Ponomarev, Chiral Higher Spin Theories and Self-Duality, JHEP 12 (2017) 141 [1710.00270].
  • [49] R. Monteiro, From Moyal deformations to chiral higher-spin theories and to celestial algebras, JHEP 03 (2023) 062 [2212.11266].
  • [50] A. Sharapov, E. Skvortsov and R. Van Dongen, Chiral Higher Spin Gravity and Convex Geometry, 2209.01796.
  • [51] A. Sharapov, E. Skvortsov, A. Sukhanov and R. Van Dongen, More on Chiral Higher Spin Gravity and convex geometry, Nucl. Phys. B 990 (2023) 116152 [2209.15441].
  • [52] P. Van Nieuwenhuizen, Free graded differential superalgebras, in Group Theoretical Methods in Physics: Proceedings of the XIth International Colloquium Held at Boğaziçi University, Istanbul, Turkey August 23–28, 1982, pp. 228–247, Springer, 2005.
  • [53] R. D’Auria, P. Fre and T. Regge, Graded Lie Algebra Cohomology and Supergravity, Riv. Nuovo Cim. 3N12 (1980) 1.
  • [54] M. A. Vasiliev, Consistent Equations for Interacting Massless Fields of All Spins in the First Order in Curvatures, Annals Phys. 190 (1989) 59–106.
  • [55] J. Stasheff, Homotopy associativity, H-spaces from a Homotopy Point of View (2006) 27–30.
  • [56] D. Ponomarev, Towards higher-spin holography in flat space, JHEP 01 (2023) 084 [2210.04035].
  • [57] R. Penrose and W. Rindler, SPINORS AND SPACE-TIME. VOL. 2: SPINOR AND TWISTOR METHODS IN SPACE-TIME GEOMETRY. Cambridge Monographs on Mathematical Physics. Cambridge University Press, 4, 1988.
  • [58] E. Newman and R. Penrose, An Approach to gravitational radiation by a method of spin coefficients, J. Math. Phys. 3 (1962) 566–578.
  • [59] R. P. Kerr and A. Schild, Republication of: A new class of vacuum solutions of the einstein field equations, General Relativity and Gravitation 41 (2009), no. 10 2485–2499.
  • [60] K. Krasnov, Self-Dual Gravity, Class. Quant. Grav. 34 (2017), no. 9 095001 [1610.01457].
  • [61] J. M. Stewart, Advanced general relativity. Cambridge Monographs on Mathematical Physics. Cambridge University Press, 4, 1994.
  • [62] H. Brinkmann, Einstein spaces which are mapped conformally on each other, Mathematische Annalen 94 (1925), no. 1 119–145.
  • [63] P. C. Aichelburg and R. U. Sexl, On the Gravitational field of a massless particle, Gen. Rel. Grav. 2 (1971) 303–312.
  • [64] A. Bilge and M. Gürses, Generalized kerr-schild transformation, in Group Theoretical Methods in Physics, pp. 252–255. Springer, 1983.
  • [65] H. Stephani, D. Kramer, M. A. H. MacCallum, C. Hoenselaers and E. Herlt, Exact solutions of Einstein’s field equations. Cambridge Monographs on Mathematical Physics. Cambridge Univ. Press, Cambridge, 2003.
  • [66] E. Skvortsov and R. Van Dongen, Minimal models of field theories: Chiral Higher Spin Gravity, 2204.10285.
  • [67] V. E. Didenko and A. V. Korybut, Planar solutions of higher-spin theory. Part I. Free field level, JHEP 08 (2021) 144 [2105.09021].
  • [68] V. E. Didenko and A. V. Korybut, Planar solutions of higher-spin theory. Nonlinear corrections, JHEP 01 (2022) 125 [2110.02256].
  • [69] S. W. Hawking, Gravitational Instantons, Phys. Lett. A 60 (1977) 81.
  • [70] C. P. Boyer and J. D. Finley, III, Killing Vectors in Selfdual, Euclidean Einstein Spaces, J. Math. Phys. 23 (1982) 1126.
  • [71] T. Adamo, G. Bogna, L. Mason and A. Sharma, Scattering on self-dual Taub-NUT, Class. Quant. Grav. 41 (2024), no. 1 015030 [2309.03834].
  • [72] A. A. Belavin, A. M. Polyakov, A. S. Schwartz and Y. S. Tyupkin, Pseudoparticle Solutions of the Yang-Mills Equations, Phys. Lett. B 59 (1975) 85–87.
  • [73] M. F. Atiyah, N. J. Hitchin, V. G. Drinfeld and Y. I. Manin, Construction of Instantons, Phys. Lett. A 65 (1978) 185–187.
  • [74] W. A. Bardeen, Selfdual Yang-Mills theory, integrability and multiparton amplitudes, Prog. Theor. Phys. Suppl. 123 (1996) 1–8.
  • [75] E. D. Skvortsov, T. Tran and M. Tsulaia, Quantum Chiral Higher Spin Gravity, Phys. Rev. Lett. 121 (2018), no. 3 031601 [1805.00048].
  • [76] E. Skvortsov, T. Tran and M. Tsulaia, More on Quantum Chiral Higher Spin Gravity, Phys. Rev. D 101 (2020), no. 10 106001 [2002.08487].
  • [77] E. Skvortsov and T. Tran, One-loop Finiteness of Chiral Higher Spin Gravity, JHEP 07 (2020) 021 [2004.10797].
  • [78] T. Tran, Twistor constructions for higher-spin extensions of (self-dual) Yang-Mills, JHEP 11 (2021) 117 [2107.04500].
  • [79] T. Tran, Toward a twistor action for chiral higher-spin gravity, 2209.00925.
  • [80] Y. Herfray, K. Krasnov and E. Skvortsov, Higher-spin self-dual Yang-Mills and gravity from the twistor space, JHEP 01 (2023) 158 [2210.06209].
  • [81] Y. Neiman, Higher-spin self-dual General Relativity: 6d and 4d pictures, covariant vs. lightcone, JHEP 07 (2024) 178 [2404.18589].
  • [82] K. Costello and S. Li, Quantization of open-closed BCOV theory, I, 1505.06703.
  • [83] K. Costello and S. Li, Anomaly cancellation in the topological string, Adv. Theor. Math. Phys. 24 (2020), no. 7 1723–1771 [1905.09269].
  • [84] K. J. Costello, Quantizing local holomorphic field theories on twistor space, 2111.08879.
  • [85] R. Bittleston, D. Skinner and A. Sharma, Quantizing the Non-linear Graviton, Commun. Math. Phys. 403 (2023), no. 3 1543–1609 [2208.12701].
  • [86] R. Monteiro, R. Stark-Muchão and S. Wikeley, Anomaly and double copy in quantum self-dual Yang-Mills and gravity, JHEP 09 (2023) 030 [2211.12407].
  • [87] T. Basile, Massless chiral fields in six dimensions, 2409.12800.
  • [88] D. Ponomarev, Chiral higher-spin double copy, 2409.19449.
  • [89] C. Chowdhury, G. Doran, A. Lipstein, R. Monteiro, S. Nagy and K. Singh, Light-cone actions and correlators of self-dual theories in AdS4, 2411.04172.
  • [90] K. Krasnov and A. Lipstein, Pure connection formalism and Plebanski’s second heavenly equation, 2412.15779.
  • [91] H. Kajiura, Noncommutative homotopy algebras associated with open strings, Rev. Math. Phys. 19 (2007) 1–99 [math/0306332].