- 1Department of Mathematics, King’s College London, London, United Kingdom
- 2Department of Applied Mathematics and Theoretical Physics, University of Cambridge, Cambridge, United Kingdom
We discuss non-Lorentzian Lagrangian field theories in 2n − 1 dimensions that admit an SU(1, n) spacetime symmetry which includes a scaling transformation. These can be obtained by a conformal compactification of a 2n-dimensional Minkowskian conformal field theory. We discuss the symmetry algebra, its representations including primary fields and unitarity bounds. We also give various examples of free theories in a variety of dimensions and a discussion of how to reconstruct the parent 2n-dimensional theory.
1 Introduction
Lorentz symmetry plays a crucial role in many applications of Quantum Field Theory but it is not necessary. Indeed the condensed matter community more often than not looks at theories without it. This opens the door to additional spacetime symmetries such as the Bargmann, Carroll and Schrödinger groups. In particular non-Lorentzian conformal field theories have now received considerable attention and reveal many interesting features, for example see [1–5].
It is well-known that one way to construct non-Lorentzian theories with Schrödinger symmetry is to reduce a Lorentzian theory of one higher dimension on a null direction. From the higher dimensional perspective such null reductions are somewhat unphysical but that need not concern us if we are only interested in the features of the reduced theory. Indeed the (null) Kaluza-Klein momentum is often associated with particle number and, in contrast to traditional Kaluza-Klein theories, one need not truncate the action to the zero-modes but rather any given Fourier mode. The resulting theories are interesting in themselves and have applications in Condensed Matter Systems and DLCQ constructions (where one does have to try to make sense of a null reduction).
Here we will explore theories with novel spacetime SU(1, n) symmetry. These can be obtained by reducing a Lorentzian conformal field theory (CFT) along a null direction in conformally compactified Minkowski space. A key novelty here is that the conformal null reduction can be inverted so that the non-compact higher dimensional theory can in principle be reconstructed from the reduced theory provided all Kaluza-Klein modes are retained. The effect of such a reduction is to induce an Ω-deformation into the reduced theory. Since the SU(1, n) symmetry acts separately on each Fourier mode we can truncate our actions to any given Fourier mode number. Or we can keep them all and reconstruct the original theory.
We also comment that our interest in these models has arisen through an explicit class of supersymmetric non-Abelian gauge theories in five-dimensions with
In this paper we wish to illustrate some of general aspects of such theories. In Section 2 we will outline a construction by dimensional reduction of a CFT on conformally compactified Minkowski space and give the corresponding AdS interpretation. In Section 3 we discuss various properties of the SU(1, n) symmetry algebra such as primary fields, unitarity bounds and its relation to conventional non-relativistic conformal symmetry. In Section 4 we discuss a superconformal extension that is possible in the case of five-dimensions and construct some BPS bounds. In section 5 we will give explicit examples of theories with SU(1, n) symmetry. In the interest of simplicity we will only consider free theories here, although, as mentioned above, interacting theories can be constructed. In section 6 we will outline how, by retaining the entire Kaluza-Klein tower of fields, one can reconstruct the 2-point functions of the parent 2n-dimensional theory. Finally in section 7 we give our conclusions and comments.
2 Construction via Conformal Compactification
We start with 2n-dimensional Minkowski spacetime in lightcone coordinates with metric/
where μ ∈ { +, −,i}, i = 1, 2, … , 2n − 2, and perform the coordinate transformation1
Here Ωij is a constant anti-symmetric matrix that satisfies
Note, we can always perform a rotation in the xi directions so as to bring Ωij to a canonical form; in particular, one can always find orthogonal matrix M such that
This coordinate transformation leads to the metric
Following this we perform a Weyl transformation
Under such a conformal transformation a scalar operator
Note the range of x+ ∈ (−πR, πR) is finite. Thus we can conformally compactify the x+ direction of 2n-dimensional Minkowski space by x+ ∈ [−πR, πR]. In which case we can write
where for now we keep the range of k general, e.g. integer or half-integer. Lastly is helpful to note that the metric and inverse metric are
2.1 Dual AdS Slicing
As we have seen, the metric
Let Za, a = 0, 1, … , n be a set of (n + 1) complex coordinates, and ηab = diag (−1, 1, … , 1). Then, when constrained to
the Za provide coordinates on Lorentzian AdS2n+1, with metric given by
suitably pulled back to solutions of (2.10).
Next, we can parameterise solutions to the constraint (2.10) with 2n + 1 real coordinates (y, x+, x−, xi). We have2
where here M is the orthogonal matrix appearing in (2.4).
These coordinates provide a description for AdS2n+1 as a one-dimensional fibration over a non-compact form of n-dimensional complex projective space, sometimes denoted
where
can be identified as the metric on
To go to the conformal boundary, we now restrict to a surface of constant y, and take y large. It is then clear that as we do so, the metric approaches the form
thus recovering the form of the metric
Finally, let us discuss symmetries. Each isometry in the bulk, described by some Killing vector field, corresponds to a conformal symmetry on the boundary. The full set of such symmetries form the algebra
2.2 Symmetries Under Dimensional Reduction
Each continuous spacetime symmetry of a conformal field theory on Minkowski space is generated by an operator G, with the set of all such operators forming the algebra
Each operator G in turn correspond to a conformal Killing vector G∂ of the metric
with indices raised and lowered with the Minkowski metric ημν. Each of these vector fields G∂ then satisfies
Their non-vanishing commutators are
We can then perform the coordinate transformation (2.2) followed by the Weyl rescaling to arrive at the metric
It is now straightforward to see that translations along x+ are an isometry3 of the metric g. In terms of the original Minkowski symmetry generators, this is realised by the combination
Then, given some conformal field theory on 2n-dimensional Minkowski space, we can perform a Kaluza-Klein on the x+ interval. At the level of the symmetry algebra, this amounts to choosing a basis for the space of local operators which diagonalises P+. The resulting operators are Fourier modes on the x+ interval. They fall into representations of the centraliser of P+ within
A basis for the subalgebra
where the Jα are absent for n = 1, 2, and otherwise α = 1, … , n2−2n. Here, the
Then, these vector fields are indeed conformal Killing vector fields of
and vanishing for the other generators.
Let us identify the subalgebra of pure rotations within
So let us take n ≥ 3. We may, a priori, consider a general spatial rotation of the form
Thus, for n ≥ 3 the total rotation subalgebra is
Finally, let us state the commutation relations for the algebra
The remaining generators are sorted into “scalar” generators
All remaining commutators are found to be
where the coefficient in front of B in the commutator i [Gi, Pj] holds down to n = 2. Further, we denote by
One can show that this equation can always be uniquely solved for the
Following the discussion in Section 2.1, we identify
Then,
3 Primary Operators and Their Properties
So let us now consider a (2n − 1)-dimensional theory with SU(1, n) symmetry. Given some operator Φ(0) at the origin (x−, xi) = (0, 0), we say it has scaling dimension Δ if it satisfies [T, Φ(0)] = iΔΦ(0). Then, in direct analogy with the Schrödinger algebra of conventional non-relativistic conformal field theory, we can straightforwardly construct further states also with definite charge under T.
We find that {H, K} raise and lower scaling dimension by two units, respectively, so that if Φ(0) has scaling dimension Δ, then [H, Φ(0)] has scaling dimension (Δ + 2), while [K, Φ(0)] has (Δ−2). We have then also the pair {Pi, Gi}, which raise and lower scalig dimension by one unit, respectively.
Going further, we can generalise results from the n = 3 case [9], and define a primary operator at the origin (x−, xi) = (0, 0) by its transformation under the stabiliser of the origin within
Here,
The key property of such a primary is that it is annihilated by the lowering operators {K, Gi}, and thus sits at the bottom of a tower of states generated by the raising operators {H, Pi}, known as usual as descendants.
Given any operator Φ(0) at the origin, an operator at some point (x−, xi) is defined by
Then, requiring that at any point we have Φ(x + ϵ) − Φ(x) = ϵ−∂−Φ(x) + ϵi∂iΦ(x) fixes the action of H, Pi on Φ(x) [9]. Note, this is a somewhat more subtle computation than is encountered in relativistic conformal field theory, since the translation subalgebra span{H, Pi} is non-Abelian.
One can in particular apply the transformation rules (3.1) along with the algebra (2.23) to determine the transformation properties of a primary
3.1 Recovering Conventional Non-relativistic Conformal Field Theory
At the level of symmetries, the presence of conformal symmetry in the relativistic theory manifests itself as an enhancement of the Poincaré algebra to the conformal algebra. The analogous statement in non-relativistic theories is an enhancement of the inhomogeneous Galilean algebra—or rather, its central extension, the Bargmann algebra—to the Schrödinger algebra. Let us denote by Schr(d) the Schrödinger algebra governing the non-relativsitic conformal dynamics of a particle in d spatial dimensions.
Then, Schr(d) is realised precisely as the centraliser of a null translation within the conformal algebra
Recall, we defined the subalgebra
Hence, in the limit R → ∞, the subalgebra
Things therefore work smoothly at the level of the algebra. However, given a theory admitting the Ω-deformed non-relativistic conformal symmetry
A convenient way to arrive at this setup—which from the 2n-dimensional perspective coincides with that of DLCQ—is to first introduce an orbifold. In particular at finite R the orbifold restricts to operators that are periodic but with period 2πR/K along the x+ direction for some
Indeed, this precise DLCQ limit of a
3.2 State-Operator Map
A deep and powerful result tool in the study of relativistic conformal field theory is the operator-state map, relating on one hand conformal primary operators, and on the other, eigenstates of the Hamiltonian of the theory on a sphere. An analogous map exists in conventional non-relativistic conformal field theories [1], which relates primary operators—defined in a way entirely analogous to the above—to eigenstates of the Hamiltonian augmented by a harmonic potential.
We will now show that construction applies in an almost identical way to the
We approach the construction of our operator-state map from the perspective of automorphisms of the symmetry algebra, a well-established point of view in relativistic CFTs which has also recently been formulated for non-relativistic CFTs governed by the Schrödinger group [4].
Given some operator Φ(0) at the origin, we may define a state
Next, let us perform a Wick rotation in the symmetry algebra, defining D = −iT. Then, if Φ(0) has scaling dimension Δ under T, then
and thus |Φ⟩ has eigenvalue Δ under D. Then, just as with operators, we can use the ladder operators {H, K} and {Pi, Gi} to raise and lower the D eigenvalue of |Φ⟩. For instance, DH |Φ⟩ = (Δ + 2)H |Φ⟩, while DGi |Φ⟩ = (Δ−1)Gi |Φ⟩.
We can consider
Thus, we have on one hand primary operators and their descendants, all with definite scaling dimension, and on the other hand, eigenstates of the operator D = −iT. Let us now however explore an alternative frame, related by a similarity transform on the Hilbert space and space of operators. As we shall see, this transformation, which can be seen as a non-relativstic analogue of the operator-state map of relativistic CFT, relates the spectra of D with that of a combination of the form
So let us consider transformed states and operators given by
for some constant μ. Note, this transformation is clearly consistent with the identification (3.3). In particular, for a primary operator
as is familiar from the usual non-relativistic operator-state map [1]. Then, this defines an alternative map between on one hand the primary operators
Explicitly, the transformed operators under (3.5) are
while the remaining generators, the rotations and central charge, transform trivially as
while acting with
Up to normalisation these operators (3.7) take the same form as in conventional non-relativistic CFT [1, 5], and thus automatically satisfy the same algebra in the R → ∞ limit.
3.3 Implications of Unitarity
If we assume unitarity in the original Minkowskian theory, then all states will have non-negative norm. Just as is the case of Lorentzian CFTs, we can use this assumption to place constraints on the eigenvalues of certain operators. The original Minkowskian symmetry generators were all Hermitian operators, but since we are interested in states quantised in the analogue of radial quantisation, we should instead consider the barred generators of (3.7). Simple Hermiticity of the original generators implies for the barred generators the following reality conditions
So let us now consider the primary state
If the theory is unitary, all states have non-negative norm, implying that for primary states
Since
where i is not summed over. But let us now sum over i, so as to exploit the tracelessness of
where recall that
In particular, if
Note, we see that a scalar primary must have p+ ≤ 0. It is interesting to note that this condition appears to be manifestly realised in known supersymmetric interacting gauge theory examples of
We can use the positivity of Δ and M to improve our bound for Δ. In particular, for any primary with M > 0, consider the norm [2].
This leads to the inequality
the right hand side is manifestly semi-positive, and we have already shown Δ is too, so one arrives at
for any primary with M > 0. Since we have 2n − 2 spacial dimensions we see that, despite the Ω deformation, this bound agrees with the usual bound for theories with a Schrödinger symmetry algebra [2].
4 Superconformal Extension in Six Dimensions
We have thus far explored the reduction of symmetries of an even-dimensional conformal field theory when dimensionally reduced along a particular conformally-compactified direction. In six or fewer dimensions the conformal algebra admits extensions to various Lie superalgebras and thus it is natural to extend our analysis to determine the fate of supersymmetry under such dimensional reductions. In particular, any surviving supersymmetry constitutes a Lie superalgebra extension of
The dimensional reduction we have constructed is novel only for n ≥ 2, while the starting 2n-dimensional CFT can have supersymmetry only for n = 1, 2, 3. Motivated by a well-studied class of supersymmetric Lagrangian models with
4.1
In six-dimensions the only choices for relativistic superconformal algebras are D (4, 1) and D (4, 2) corresponding to
In Minkowski signature we choose conventions where all Bosonic generators are Hermitian, as before. Their commutation relations are the same as in Section 2.2. The R-symmetry generators have the standard form
with I ∈ {1, … , 5}. The Fermionic generators are six-dimensional symplectic-Majorana-Weyl Fermions. The reality condition as applied above is
and similar for
Again we wish to find the maximal subalgebra of all elements that commute with the element P+, defined in terms of the six-dimensional (hatted) operators as
We find that 3/4 of the supercharges commute with P+. Precisely which set of supercharges this is depends on whether Ωij is self-dual or anti-self-dual; without loss of generality, let us choose the latter case. Then, letting a ± subscript denote chirality under Γ05, the commuting supercharges are
The alternative case, where Ωij is self-dual, is found simply by swapping all Γ05 chiralities. Then, their commutation relations with the bosonic generators are
while we have anti-commutators
where we have defined the projectors Π± = 1/2 (1 ±Γ*).
Thus there are 50 = 1 + 15 + 10 + 24 Bosonic generators corresponding to the central extension,
The Fermionic generators can also be transformed by (3.6), which yields
Taking a symplectic-Majorana-Weyl reality condition for the six-dimensional spinors we find the following Hermiticity properties for the barred generators
Rather unusually for such algebras, along with a pair of Fermionic generators that raise and lower the eigenvalue of T, namely the Q− and S+, we also have generators that do not change this eigenvalue; Θ−. We can see that while
It follows inductively that any number of
which leads to the inequality
Summing again on α and A symmetrises on simultaneous exchange of α, β and A, B, allowing us to replace the product with the anticommutator. This then simply reproduces the earlier bound M ≥ 0.
A more interesting bound is found from the norm
which leads to
and implies
Since
where we defined
It is interesting to note that, up to a choice of real form for the respective algebras, the reduction of symmetry from the six-dimensional (2, 0) superalgebra down to centraliser of P+ is identical to the symmetry breaking pattern of the classical ABJM theory, which realises manifestly only a particular subalgebra of the full three-dimensional
5 Free Fields in Various Dimensions
In this section we want to discuss examples of field theories in (2n − 1)-dimensions with SU(1, n) symmetry. Our examples will be obtained by the conformal compactification of a 2n-dimensional free conformal theory. We will include the entire Kaluza-Klein tower in our discussion but as the SU(1, n) symmetry acts on each level independently one is also free to truncate the actions to only include fields of particular levels. Interacting versions of these theories can also be constructed by starting with an interacting conformal field theory, for example by considering the reduction of non-Abelian theories. In the interests of clarity we will not consider these here.
5.1 Scalars in 2n − 1 Dimesions
To begin we consider a free real scalar in (1 + 1)-dimensions, i.e. n = 1. As we will see this case is special, yet familiar. In particular we start with the action for a real scalar field:
where in this simple case
Since
Note that
Substituting into the action we find
By construction the SU(1, 1) symmetry separately on each of the fields ϕ(k) at fixed
The Liftshitz scaling T is simply
Finally the special conformal transformation K+ acts as:
One can readily check that these are indeed symmetries to first order.
However we see that they can be extended to
for any function f (x−). Taking κ constant, linear and quadratic leads to the H, T and K generators, respectively. In fact this is simply the action of one-dimensional diffeomorphisms and therefore yields an infinite-dimensional symmetry group with generators
These satisfy the Witt algebra
where H = L−1, T = L0, K = L1 form a finite dimensional subalgebra. However just as in the familiar case of the string worldsheet in the quantum theory, where we must normal order the operators ϕ(k), we will generate a central charge c = 1.
Let us now consider a free real scalar obtained from reduction from D = 2n:6
where
Next we expand
Note that we do not necessarily require that
As discussed this action admits an SU(1, n) spacetime symmetry acting on each level k independently.
5.2 Fermions in 2n − 1 Dimensions
Let us consider the reduction of a Fermion. Starting in 2n dimensions we have
Here
We see that
Note that we do not necessarily impose
This leads to the reduced action
where now γ−, γ+, γi are simply the γ-matrices of flat spacetime (i.e. the same as
We it is helpful to split
The action is then
Note that the last term essentially leads to a shift in k for some components of χ(k), depending on the eigenvalue of iγijΩij. It can also vanish if
Finally we observe that in one-dimension we simply find
One again the action has an infinite dimensional symmetry generated by Ln provided that the χ(k) are invariant. Furthermore we will encounter a central charge c = 1/2 once we normal order the fields in the quantum theory.
5.3 A 1-Form Gauge Field in 3-Dimensions
Let us start with a free four-dimensional Maxwell gauge field
where
Performing the integral over x+ we obtain
where
and we must identify
5.4 A 2-Form Gauge Field in 5-Dimensions
Finally we consider a free tensor in six-dimensions:
where
with
be a five-dimensional one-form with 2-form field-strength
Here
6 Recovering 2n-Dimensional Physics
In this section we would like to see how, by considering the entire Kaluza-Klein tower, we can reconstruct the correlation functions of the 2n-dimensional theory that we started with. Since there are additional complications that enter when the field has a non-trivial Lorentz transformation we will restrict our attention here to scalar fields.
6.1 From One to Two Dimensions
Let us start with a tower of scalar fields in one-dimension that are obtained from a two-dimensional scalar as given in (5.4). We can read off from the action (5.4) that the correlation functions are of the form (k > 0)
Let us try to compute a two-point function of the original two-dimensional theory. If we try to compute
We note that the sum over the Fourier modes is ill-defined. We can consider an iɛ prescription
This condition is of course familiar from the usual Hamiltonian treatment where ϕ(k) are the left moving oscillators. Thus we are left with
To evaluate this we note that
and differentiating gives
Continuing we find (setting ɛ = 0)
On the other hand we have
and hence
which in terms of the original coordinates is
which is the correct propagator for the two-dimensional theory.
It is clear that from this treatment we will never be able to reconstruct the right-moving sector as only
6.2 From 2n − 1 to 2n Dimensions
Now we want to repeat our analysis of 2-point functions but now in higher dimensions. For simplicity we use translational invariance to put one operator at the origin:
where
To this end, for spherically symmetric solutions, it is helpful to introduce
so that the equation reduces to
Ignoring the singularities at
for some constants dn,k. For n = 3 this agrees with the general form for a 2-point function in a five-dimensional theory with SU(1, 3) symmetry as constructed in [9].
We can now reconstruct the 2n-dimensional two-point function:
where
Here we again encounter the problem that the sum over all k will not be well-defined as |q| = 1 and introducing an iɛ prescription can only cure the convergence for large k or large − k but not both. To continue we require that positive modes Fourier modes of
which ensures that
Note that we encounter a problem if we quantize the theory using the action (5.14) with x− as “time” since we obtain the conjugate momentum
Thus [ϕ(k)(x−, xi), Π(k)(x−, 0)] = −2ikR−1 [ϕ(k)(x−, xi), ϕ(−k)(x−, 0)] is non-zero for k ≠ 0 and therefore we can’t simultaneously impose
which is potentially in contradiction with (6.19).
Let us look at this more closely on a case-by-case basis. For n = 1 there is no problem as only positive values of k appear in (6.19). For n = 2 we must take k to be half-integer so the smallest positive oscillator is ϕ(1/2) and the bound in (6.19) becomes k >−1/2 which also does not include any ϕ(k) with k < 0. At n = 3 we see that we require
To obtain the 2n-dimensional 2-point function we need
for some constant C ∼ g2/πVol (S2n−3). In particular for the two cases at hand this means that must have
In the following subsection we provide a derivation of this normalisation by requiring that we get the correct coefficient of the delta-function in (6.14).
We also see from (6.22) that indeed we require
Thus we recover the expected two-point function of the 2n-dimensional theory.
6.2.1 Green’s Function Normalisation
In this appendix we want to present an argument that the normalisation dn,k introduced in (6.15), which should be chosen to ensure the correct delta-function coefficient in (6.14), does indeed agree with the form (6.23). To do this we consider an arbitrary smooth function
with Gn,k given in (6.15). Here D is first quadrant of the z-plane (corresponding to
We therefore need to show that we can find coefficients dn,k such that
To this end we observe that, away from z = 0, we can write
where
Here γ is an arbitrary constant corresponding to the freedom to add a total derivative ω → ω + dΓ with
Next we switch to polar coordinates z = reiθ and observe that
Thus if we consider D as a wedge ranging between 0 and π/2 and
To compute this integral we observe that ωθ = ∂θφ with
and hence
Note that if k is in the range |k| > (n − 1)/2 then there is no value of l such that the denominators in φ vanish. For |k| ≤ (n − 1)/2 we must be more careful however, as discussed above, we are not interested in this case here.
The integral (6.34) depends on γ and yet γ should not affect the Green’s function Gk,n. In fact we find that φ(0) does not depend on γ but φ(π/2) does. Thus we need to impose a condition at θ = π/2 (corresponding to x− = 0 for any
Thus as a result we must take
which reproduces (6.23).
7 Conclusion and Comments
In this paper we have examined non-Lorentzian theories with SU(1, n) spacetime symmetry in (2n − 1)-dimensions. In particular we showed how one can construct such theories by reduction of a conformally invariant Lorentzian theory in 2n-dimensions. However other constructions may well exist. We showed that the novel operator-state map of the Schrödinger group extends straightforwardly to SU(1, n) theories and demonstrated how conventional non-relativistic conformal field theory is recovered in a particular limit. We also explored some unitarity bounds and a supersymmetric extension of the spacetime symmetry algebra in five dimensions, which has been explicitly realised in a class of gauge theory examples [6–8].
We then presented examples of free theories in a variety of dimensions with various field contents. Although we kept the Kaluza-Klein tower of fields this is not necessary for SU(1, n) symmetry and one can truncate the Lagrangians to a subset of Fourier modes. One can also consider including interactions (e.g. see [6–8]). We also discussed how to reconstruct the parent 2n-dimensional theory by keeping the entire Kaluza-Klein tower of operators. For this the role of the Ω-deformation is critical.
We note that in theories with SU(1, n) symmetry we have constructed there are terms with the ‘wrong-sign’ kinetic term induced by the Ω-deformation, when we view x− as time. However at the spatial origin such “wrong-sign” terms vanish. Given translational invariance this suggests that the SU(1, n) symmetry can be used to regain control of the theory. In particular, since there is a well-defined map to the original, non-compact, Minkowskian theory we believe that there should be a corresponding consistent treatment of the lower-dimensional theory which alleviates any such problems.
Data Availability Statement
The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.
Author Contributions
All authors listed have made a substantial, direct, and intellectual contribution to the work and approved it for publication.
Funding
NL is a co-investigator on the STFC grant ST/T000759/1, RM. was supported by David Tong’s Simons Investigator Grant, and TO was supported by the STFC studentship ST/S505468/1.
Conflict of Interest
The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.
Publisher’s Note
All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.
Acknowledgments
We would like to thank A. Lipstein and P. Richmond for initial collaboration on this work, and David Tong for helpful discussions. This paper is to be submitted to the Frontiers Special Edition on Non-Lorentzian Geometry and its Applications.7
Footnotes
1It is curious to note that this transformation is similar to the transformation used in [3] to convert to the so-called oscillator frame, along with an x+-dependent rotation by Ωij.
2As our focus is on continuous conformal symmetries on the boundary, it is sufficient for our purposes to consider this a local parameterisation of AdS2n+2, and thus neglect global features of this real coordinate choice.
3Translations in x+ are a conformal symmetry of the original metric
4We choose this sign for N, in line with the general NRCFT literature, since unitarity then requires N ≥ 0, as discussed in Section 3.3.
5Note that ΩAB should not be confused with Ωij which we used in the coordinate transformation. To ameliorate this problem we will always explicitly write the indices.
6There is also a coupling to the spacetime Ricci scalar but since we are working on a conformally flat metric, this term vanishes.
7www.frontiersin.org/research-topics/19214/non-lorentzian-geometry-and-its-applications
References
1. Nishida Y, Son DT. Nonrelativistic Conformal Field Theories. Phys Rev D (2007) 76:086004. [0706.3746]. doi:10.1103/physrevd.76.086004
2. Nishida Y, Son DT. Unitary Fermi Gas, $$\epsilon$$ Expansion, and Nonrelativistic Conformal Field Theories. Lect Notes Phys (2012) 836:233–75. [1004.3597]. doi:10.1007/978-3-642-21978-8_7
3. Goldberger WD, Khandker ZU, Prabhu S. OPE Convergence in Non-relativistic Conformal Field Theories. JHEP (2015) 12:04814128507. doi:10.1007/jhep12(2015)048
5. Doroud N, Tong D, Turner C. The Conformal Spectrum of Non-abelian Anyons. Scipost Phys (2018) 4:022. [1611.05848]. doi:10.21468/scipostphys.4.4.022
6. Lambert N, Lipstein A, Richmond P. Non-Lorentzian M5-Brane Theories from Holography. JHEP (2019) 08:060. [1904.07547]. doi:10.1007/jhep08(2019)060
7. Lambert N, Lipstein A, Mouland R, Richmond P. Bosonic Symmetries of (2, 0) DLCQ Field Theories. JHEP (2020) 01166. [1912.02638]. doi:10.1007/jhep01(2020)166
9. Lambert N, Lipstein A, Mouland R, Richmond P. Five-Dimensional Non-lorentzian Conformal Field Theories and Their Relation to Six-Dimensions. JHEP (2021) 03:053. [2012.00626]. doi:10.1007/jhep03(2021)053
10. Lambert N, Lipstein A, Mouland R, Richmond P. Instanton Worldlines in Five-Dimensional Ω-Deformed Gauge Theory (2105). p. 02008.
11. Lambert N, Lipstein A, Mouland R, Richmond P. Five-Dimensional Path Integrals for Six-Dimensional Conformal Field Theories (2109). p. 04829.
12. Pope CN, Sadrzadeh A, Scuro SR. Timelike Hopf Duality and Type IIA * String Solutions. Class Quan Grav. (2000) 17:623–41. [hep-th/9905161]. doi:10.1088/0264-9381/17/3/305
13. Mouland R. Non-Lorentzian Supersymmetric Models and M-Theory Branes. Ph.D Thesis (2021) 9. 2109.04416.
Keywords: symmetry, quantum field theory, non-lorentzian, conformal transformation, supersymmetry
Citation: Lambert N, Mouland R and Orchard T (2022) Non-Lorentzian SU(1, n) Spacetime Symmetry In Various Dimensions. Front. Phys. 10:864800. doi: 10.3389/fphy.2022.864800
Received: 28 January 2022; Accepted: 04 March 2022;
Published: 23 June 2022.
Edited by:
José Figueroa-O'Farrill, University of Edinburgh, United KingdomReviewed by:
Ioannis Papadimitriou, Beijing Institute for Mathematical Sciences and Applications - BIMSA, ChinaPietro Antonio Grassi, University of Eastern Piedmont, Italy
Copyright © 2022 Lambert, Mouland and Orchard. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.
*Correspondence: N. Lambert, bmVpbC5sYW1iZXJ0QGtjbC5hYy51aw==