0% found this document useful (0 votes)
20 views15 pages

2018 Extended Noncommutative Minkowski Spacetimes and Hybrid Gauge Symmetries

This paper studies Lie bialgebra structures that can extend the symmetries of noncommutative Minkowski spacetimes and allow for hybrid gauge-spacetime symmetries. It shows that centrally extending quantum deformations of spacetime symmetry groups, like the Poincaré algebra, can allow generators to mix when acting on tensor product states. This provides a new way to circumvent the Coleman-Mandula theorem and incorporate both spacetime and gauge symmetries in quantum field theories through noncommutative structures. The paper presents all possible Lie bialgebra structures with this type of hybrid action that are allowed on centrally-extended Poincaré and (A)dS algebras in (1+1) dimensions.

Uploaded by

WI TO
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
20 views15 pages

2018 Extended Noncommutative Minkowski Spacetimes and Hybrid Gauge Symmetries

This paper studies Lie bialgebra structures that can extend the symmetries of noncommutative Minkowski spacetimes and allow for hybrid gauge-spacetime symmetries. It shows that centrally extending quantum deformations of spacetime symmetry groups, like the Poincaré algebra, can allow generators to mix when acting on tensor product states. This provides a new way to circumvent the Coleman-Mandula theorem and incorporate both spacetime and gauge symmetries in quantum field theories through noncommutative structures. The paper presents all possible Lie bialgebra structures with this type of hybrid action that are allowed on centrally-extended Poincaré and (A)dS algebras in (1+1) dimensions.

Uploaded by

WI TO
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 15

Eur. Phys. J.

C (2018) 78:615
https://doi.org/10.1140/epjc/s10052-018-6097-1

Regular Article - Theoretical Physics

Extended noncommutative Minkowski spacetimes and hybrid


gauge symmetries
Angel Ballesteros1,a , Flavio Mercati2,b
1 Departamento de Física, Universidad de Burgos, 09001 Burgos, Spain
2 Dipartimento di Fisica, Università di Roma “La Sapienza”, P.le A. Moro 2, 00185 Roma, Italy

Received: 18 May 2018 / Accepted: 21 July 2018 / Published online: 1 August 2018
© The Author(s) 2018

Abstract We study the Lie bialgebra structures that can extent the structure of the possible Quantum Field Theory
be built on the one-dimensional central extension of the (QFT) candidates to describe the fundamental interactions
Poincaré and (A)dS algebras in (1 + 1) dimensions. These and fields. However, the application of standard QFT on
central extensions admit more than one interpretation, but Minkowski spacetime to quantum gravity encounters fun-
the simplest one is that they describe the symmetries of (the damental problems, and the assumption of local Poincaré
noncommutative deformation of) an Abelian gauge theory, (Lorentz) invariance might be at the core of the issue. It
U (1) or S O(2) on Minkowski or (A)dS spacetime. We show appears highly significant, in this context, that the Coleman-
that this highlights the possibility that the algebra of functions Mandula no-go theorem [1,2] implies that Poincaré symme-
on the gauge bundle becomes noncommutative. This is a new try cannot be generalized within the framework of QFT with
way in which the Coleman–Mandula theorem could be cir- Lie-group symmetries.
cumvented by noncommutative structures, and it is related to In the quest for the correct quantum theory of gravity we
a mixing of spacetime and gauge symmetry generators when were given a few hints by nature, which may point us towards
they act on tensor-product states. We obtain all Lie bialgebra the appropriate mathematical structures that are needed to
structures on centrally-extended Poincaré and (A)dS which solve the problem. One of the most interesting ones comes
are coisotropic w.r.t. the Lorentz algebra, and therefore admit from (2 + 1)-dimensional Quantum Gravity, a theory over
the construction of a noncommutative principal gauge bundle which we have good control, see [3–5]. Coupling this theory
on a quantum homogeneous spacetime. It is shown that sev- to matter fields and integrating away the gravitational degrees
eral different types of hybrid noncommutativity between the of freedom reveals that the effective background geometry
spacetime and gauge coordinates are allowed. In one of these that these fields see is not, as expected, Minkowski space
cases, an alternative interpretation of the central extension (see [6] and references therein). Instead, one finds something
leads to a new description of the well-known canonical non- that is best described as a ‘noncommutative spacetime’, in the
commutative spacetime as the quantum homogeneous space sense that the algebra of functions that intervene in the path
of a quantum Poincaré algebra. integral is nonabelian [6,7]. The symmetries of such non-
commutative backgrounds are not described by Lie groups of
isometries, but through a deformation thereof (the so-called
1 Introduction quantum groups [8–10]), whose semiclassical counterparts
(Poisson-Lie groups [9,11]) are also well-known to describe
Spacetime symmetries play an outstanding role in physics. classical phase spaces of (2 + 1) gravity coupled to point
Their effect on physical laws actually preceded the formu- particles (see [12–15] and references therein).
lation of the notion of four-dimensional spacetime, and ulti- Although in (3 + 1) dimensions it is not possible to repro-
mately motivated it. In modern physics, although we know duce the theoretical framework that allows to reach the above
that General Relativity implies the breaking at large scales conclusions in (2 + 1) dimensions, the notions of noncommu-
of Poincaré symmetry, the latter still plays a central role as tative spacetimes and their quantum-group symmetries are
the local symmetry group and it thus determines to a great still entirely valid, and we can study the possible consistent
choices for this noncommutativity. In particular, in the search
a e-mail: [email protected] for a ‘ground state’ of (3 + 1) quantum gravity which plays the
b e-mail: [email protected] same role as the effective noncommutative spacetime of (2 +

123
615 Page 2 of 15 Eur. Phys. J. C (2018) 78:615

1) quantum gravity, we are interested in maximally symmetric states are defined. In the simplest case of an Abelian inter-
noncommutative spacetimes, which admit a 10-dimensional nal symmetry, the spacetime symmetry group is enlarged
algebra of symmetries deforming the Poincaré algebra. A through a central extension Q. The essential result to be
physically-relevant deformation should be controlled by a emphasized is that centrally extending a quantum-group
parameter with the dimensions of a length, which should be deformation of a spacetime symmetry group gives, in gen-
related with the Planck scale [16]. From this perspective, eral, a different result than deforming the centrally-extended
quantum deformations of the Poincaré Lie algebra and their group in its entirety. In other words, the operations of quan-
Poisson-Lie counterparts have been thouroughly constructed tum deformation and central extension do not commute. The
and classified in the literature (for the (3 + 1) case and their same holds, more in general, for taking the direct product
associated noncommutative Minkowski spacetimes see [17– of two groups: their quantum group deformation does not
27] and references therein). have to be the direct product of two quantum groups. In this
In this paper we intend to investigate those features of case too it is the ‘coproduct’ structure that allows for more
quantum Poincaré groups and algebras that could allow to freedom on how algebra generators act on tensor products.
overcome the limitations imposed by the no-go theorems on As we will see in the sequel, [see Eq. (22)] under quantum
Lie-group symmetries of QFT. This is the fact, stated by deformations one can imagine a gauge generator Q mixing
the Coleman–Mandula theorem [1] (hereafter CMT), that in with Poincaré generators Ti when acting on tensor products,
ordinary QFT it is impossible to “combine space-time and i.e.
internal symmetries in any but a trivial way”. With “trivial
way to combine symmetries” Coleman and Mandula mean a Q  |ψ ⊗ |φ = Q  |ψ ⊗ |φ + |ψ ⊗ Q  |φ
direct product of groups, and thus this theorem precludes the (1) (2)
+ ci j Ti  |ψ ⊗ T j  |φ + ci Q  |ψ ⊗ Ti  |φ
existence of the so-called “hybrid symmetries” [28] in which (3)
the space-time and the gauge symmetries are intertwined. + ci Ti  |ψ ⊗ Q  |φ + c(4) Q  |ψ ⊗ Q  |φ + · · · ,
Supersymmetry offers a way to circumvent the no-go theo-
rem by making use of graded symmetry algebras which intro- or the opposite:
duce fermionic generators [29]. Quantum group and Yan-
gian symmetries arising from the Hopf algebras of nonlocal Ti  |ψ ⊗ |φ = Ti  |ψ) ⊗ |φ + |ψ ⊗ (Ti  |φ)
(1) (2)
conserved quantities of certain integrable (1 + 1) field theo- + di jk T j  |ψ ⊗ Tk  |φ + di j Q  |ψ ⊗ T j  |φ
ries were soon proven to avoid the Coleman-Mandula con- (3)
+ di j T j  |ψ ⊗ Q  |φ + d (4) Q  |ψ ⊗ Q  |φ + · · · ,
straint [30], although this result cannot be generalized to (3
+ 1) dimensions. On the other hand, the construction of QFT
(a) (a)
theories on twisted canonical noncommutative spacetimes where ci... and di... are numerical coefficients, and an expan-
constitutes another remarkable attempt in the same direction sion in powers of the generators Q and Ti is understood. In
(see [31–34] and references therein). Also, space-like extra fact, this paper presents all possible hybrid actions of this type
dimensions have been recently proposed as a novel mecha- that are allowed by quantum symmetries in (1 + 1) dimen-
nism to connect internal and spacetime symmetries [35]. sions.
In this work we adopt a purely kinematical perspective, This is an intriguing possibility, suggesting that in a non-
and we focus on the quantum group deformations of the commutative spacetime gauge transformations of composite
Poincaré symmetries in order to show that they could allow fields might involve a small component of e.g. translation, or
to circumvent the CMT from a different viewpoint. In fact Lorentz transformation. If the central generator Q is inter-
one of the fundamental assumptions of the CMT is a Leibniz preted as the generator of Abelian S O(2) gauge transforma-
action of the Poincaré algebra generators T on tensor product tions, then the dual coordinate q will represent a coordinate
states: on a neighbourhood of the identity of the S O(2) group mani-
fold, which is S 1 . Therefore, our approach intends to explore
T  |ψ ⊗ |φ = (T  |ψ) ⊗ |φ + |ψ ⊗ (T  |φ).
the possibility of making the manifold M × S 1 (the Carte-
One of the features of quantum groups is that, because of sian product of Minkowski space M with S 1 , usually inter-
the noncommutativity of the functions over the group, the preted as a trivial fibre bundle) noncommutative, and to get an
quantum algebra generators act in a nonlinear way on product explicit description of all possible roles played by the gauge
states coordinate in its interplay with the spacetime ones. In this
  (1)  
(2)
 sense, the possibility explored in the present paper can offer
T  |ψ ⊗ |φ = Ti  |ψ ⊗ Ti  |φ , a non-trivial way to unify gauge (‘internal’) and spacetime
i (‘external’) symmetries: the basic geometric object of field
as encoded by the ‘coproduct’ map  (see below) which theory, the (trivial) Cartesian product of Minkowski space
defines the way in which representations on tensor product with the internal space, could be replaced by a unified object

123
Eur. Phys. J. C (2018) 78:615 Page 3 of 15 615

in which it is not always possibe to distinguish the spatial the quantum homogeneous spaces arising from coisotropic
‘external’ directions from the gauge ‘internal’ ones. quantum deformations of the centrally extended Poincaré Lie
The natural starting point of such an investigation is the algebra. One of the cases so obtained is just the so-called
(1 + 1) dimensional Poincaré group with a one-dimensional canonical or θ -noncommutative spacetime (see [41–44] and
central extension, which in the commutative case repre- references therein), that can be thus interpreted as being the
sents the symmetries of Abelian gauge theory, but we stress quantum homogeneous space of a certain quantum Poincaré
that the approach here presented is fully applicable in any group. In Sect. 4 we consider the Lie bialgebra structures
dimension and also for higher dimensional (and non-abelian) for the central extension of the (1 + 1) dimensional (Anti-)
gauge groups. A systematic characterization of quantum de Sitter algebra. We find that the (A)dS case also allows
group deformations of the centrally extended Poincaré alge- the introduction of hybrid gauge symmetries for a nonvan-
bras is – to the best of our knowledge – so far an unex- ishing cosmological constant, but in general they turn out
plored issue. Some past works considered central extensions to be more restrictive than the Poincaré ones (for instance,
of the Poincaré algebra in the context of quantum groups, the θ -noncommutativity cannot be recovered). In Sect. 5 an
e.g. [36,37], but only some isolated cases were considered application of the approach here presented in higher dimen-
and the possible interpretation of central extensions as gauge sions is given by studying the possibility of extending the
symmetries was not developed. Moreover, in the classifica- interpretation of θ -noncommutativity as quantum homoge-
tion here presented we will take into account that Minkowski neous space to dimensions higher than (1 + 1) is analysed, but
spacetime is obtained as the homogeneous space associated the result is negative. Finally, in Sect. 6 we summarize and
to the quotient of the Poincaré group by the Lorentz sub- discuss the results presented in the paper, and we comment
group (which is the ‘isotropy subgroup’, i.e. the subgroup on several future research problems.
that leaves a point of the homogeneous spacetime invariant).
Since we are interested in the noncommutative spacetimes
whose symmetries are described by our quantum-deformed 2 From Lie bialgebras to noncommutative spacetimes
group, it turns out that requiring that the isotropy subgroup
closes a sub-Hopf algebra under deformation is too strong a In this section we recall how first-order noncommutative
requirement (see, for instance, [38] and references therein). spacetimes are obtained from their Lie bialgebra of sym-
In fact, the appropriate notion is that of coisotropy of the metries. The constraints imposed onto a Lie bialgebra by the
cocommutator of the Lie bialgebra on the isotropy subalge- request of underlying a quantum homogeneous space are also
bra [39,40] (see below). Essentially, this condition amounts summarized. A detailed description of quantum groups, Lie
to asking that the translation coordinates on the group close bialgebras and Poisson/quantum homogeneous spaces can
an algebra, and this algebra can be taken as the definition be found in [8–11,38–40] and references therein.
of a noncommutative algebra of coordinates on the quantum
homogeneous spacetime. The coisotropy condition will be 2.1 Lie bialgebras
used througout this paper as the essential constraint that limits
the number of possible Lie bialgebra structures (and conse- A Lie bialgebra (g, δ) is Lie algebra g with structure tensor
quently, of quantum deformations) for the centrally extended cikj
Poincaré algebra. Moreover, we will show that the very same
classification scheme can be applied to the centrally extended [X i , X j ] = cikj X k , (1)
(anti) de Sitter algebra and its associated extended noncom-
mutative spacetimes.
which is also endowed with a skew-symmetric “cocommu-
The structure and main results of the paper are the follow-
tator” map
ing. In Sect. 2 we present the basics of quantum groups, Lie
bialgebras and their associated first-order noncommutative δ :g→g∧g
spacetimes, which is illustrated with the example of the (non-
fulfilling the two following conditions:
extended) Poincaré algebra in (1 + 1) dimensions. In this case
the so-called ‘κ-Minkowski’ noncommutative spacetime is
selected by the coisotropy condition as the unique possible • 1. δ is a 1-cocycle, i.e.,
quantum homogeneous space. The detailed analysis of the
centrally extended Poincaré Lie bialgebras constitutes the δ([X, Y ]) = [δ(X ), Y ⊗ 1 + 1 ⊗ Y ]
core of the paper, and is given in Sect. 3. Here all possible +[X ⊗ 1 + 1 ⊗ X, δ(Y )], ∀ X, Y ∈ g.
types of (1 + 1) “hybrid gauge symmetries” and, therefore,
of their associated noncommutative algebras generated by • 2. The dual map δ ∗ : g ∗ ⊗ g ∗ → g ∗ is a Lie bracket on
the spacetime, x, and gauge, q coordinates are obtained as g∗.

123
615 Page 4 of 15 Eur. Phys. J. C (2018) 78:615

As a consequence, any cocommutator δ will be of the form can be explicitly found by solving the linear equations (5)
for the those of the Lie bialgebras here presented which are
jk
δ(X i ) = f i X j ∧ Xk , (2) coboundaries.

where f i is the structure tensor of the dual Lie algebra g ∗


jk
2.2 Quantum groups/algebras, Poisson-Lie groups and Lie
defined by bialgebras
jk
[ξ j , ξ k ] = f i ξ i , (3) A quantum group is an algebraic generalization of the notion
of Lie group where the algebra of functions on the group
where ξ l are the dual generators, defined by the pairing manifold is replaced by a noncommutative algebra for which
ξ j , X k  = δ j k . If, as usual in Lie group theory, we interpret the group multiplication is an algebra homomorphism. To
this pairing as the duality between a Lie algebra generator describe such a structure, we need the language of Hopf alge-
and its corresponding local coordinate around the identity of bras: a unital algebra H on which, in addition to a product
the Lie group G, then relations (3) invite the interpretation map · : H ⊗ H → H , a ‘coproduct’ map  : H → H ⊗ H ,
of ξi as the noncommutative group coordinates. In particular, together with an ‘antipode’ S : H → H and a ‘counit’
under certain conditions, a subset of these coordinates can be ε : H → C are defined. These three additional maps have
identified with the “local” coordinates of a given spacetime to be (anti) homomorphisms w.r.t. the algebra product, and
obtained as G/H with H being some isotropy subgroup. they are used to encode the group structures. For example, in
The problem of obtaining all possible Lie bialgebra struc- the case of the general linear group G L(n), the three maps
tures (g, δ) on a given Lie algebra g can be addressed by first read
solving the 1-cocycle condition, namely,
(M i j ) = M i k ⊗ M k j , S(M i j ) = (M −1 )i j ,
f kab cikj = f iak ckb j + f ikb cka j + f jak cik
b
+ f jkb cik
a
, (4) ε(M i j ) = δ i j . (6)
The three maps have to be consistent with the group axioms:
where the cikj tensor is known and linear equations (4) have to associativity, the existence of the identity and the existence
be solved for the f kab tensor. Second, the Jacobi identities of of the inverse. These translate into a coassociativity property
g ∗ impose quadratic equations onto the components of f kab of , and an identity involving the three maps and the mul-
that have not yet been fixed by (4). Therefore, a given finite tiplication: · ◦ (S ⊗ id) ◦  = · ◦ (id ⊗ S)◦ = 1 . These
dimensional Lie algebra g could admit a large number of rules, together with the homomorphism property of the three
Lie bialgebra structures δ, whose equivalence classes can be maps, make a consistent algebraic structure. We can describe
computed by making use of the automorphisms of g. Also, as in this way a standard Lie group in terms of the algebra of
it could be expected, for some Lie bialgebras the 1-cocycle functions on the group manifold, if the product · is assumed
δ is a coboundary commutative, but we can also generalize to the case in which
· is noncommutative: in that case the Hopf algebra still makes
δ(X ) = [X ⊗ 1 + 1 ⊗ X, r ], ∀ X ∈ g, (5) sense, but it does not describe an ordinary Lie group. We are
in this case dealing with a genuine quantum group.
where r is a skew-symmetric element of g ⊗ g given by To generalize the (extremely useful) notion of Lie algebra,
r = r ab X a ∧ X b , which in order to define an admissible δ it is sufficient to take the Hopf algebra dual H ∗ of a quantum
has to be a solution of the modified classical Yang–Baxter group, which is by definition a Hopf algebra. Here the algebra
equation (mCYBE) [9,10]. For semisimple Lie algebras all elements are not functions on a group manifold: they are
Lie bialgebra structures are coboundaries, and that is also the elements of the universal enveloping algebra U (g) of the
case for the Poincaré algebra in (2 + 1) and (3 + 1) dimensions, Lie algebra g associated to the group. The dual product map
despite these two Lie algebras are not semisimple [26]. In the in this case encodes the commutation rules of the Lie algebra.
case of Lie algebras with central generators, non-coboundary In the example of G L(n), the Lie algebra generators ti are
Lie bialgebra structures arise naturally, and only a fraction again n ×n matrices, but the Lie algebra product is defined by
of the Lie bialgebras that can be built on them admit a clas- the matrix commutator. The coproduct, antipode and counit
sical r -matrix (see for instance [45,46] where the case of the in this case define the Lie algebra generators as differential
centrally extended Galilei algebra is studied, or [47] where operators on the Lie group:
the Lie bialgebra structures on the Heisenberg-Weyl algebra
were classified and the corresponding quantum Hopf alge- (ti ) = ti ⊗ 1 + 1 ⊗ ti , S(ti ) = −ti , ε(ti ) = 0. (7)
bras were explicitly constructed). Since in our approach to
first-order noncommutative spacetimes we will not make use All of these maps will be deformed in a nonlinear way in
of r -matrices, we will omit them in the sequel, although they the noncommutative case. This is why the notion of quantum

123
Eur. Phys. J. C (2018) 78:615 Page 5 of 15 615

universal enveloping algebra, which admits nonlinear com- M = G/H, (9)


mutation rules between the generators, arises. In this context,
the first order deformation of a quantum algebra provides the where G is the group of isometries of the spacetime [e.g.
relevant information concerning the type of quantum defor- I S O(3, 1) (Poincaré), S O(4, 1) (AdS) or S O(3, 2) (AdS)]
mation we are constructing. Consider a quantum algebra gen- and H is the invariance subgroup of a point of the space
erated by ti , and an expansion of the commutators and the (the Lorentz subgroup S O(3, 1)). If we are dealing with a
coproducts in powers of the generators: quantum group, we need to generalize this notion so that the
isometry group leaves the noncommutative structures of the
[ti , t j ] = ci j k tk + ci j kl tk tl + O(t 3 ), homogeneous space invariant. In [40] such a generalization
(ti ) = ti ⊗ 1 + 1 ⊗ ti + f i jk t j ⊗ tk + O(t 2 ). (8) was discussed at the level of Poisson-Lie groups. A ‘Pois-
son homogenous space’ (M, π ) is a quotient M = G/H ,
The constants ci j k have to be structure constants of a Lie alge- together with a Poisson bracket π which is invariant under
bra: in fact the product of the universal enveloping algebra the action of G endowed with the Poisson-Lie structure .
is associative, and this implies that ci j k satisfies the Jacobi This implies that for a given the corresponding π has to be
identities. Similarly, since the coproduct  has to be coasso- found, and this turns out to be guaranteed if the Lie bialge-
ciative and it is a homomorphism with respect to the product, bra (g, δ) associated to fulfills the following ‘coisotropy’
the dual map defined by the antisymmetric (in j, k) part of condition:
f i jk satisfies the Jacobi identities and the 1-cocycle condi-
tion (4). Therefore the first-order term of the power expan- δ(Y ) ∈ Y ∧ X, Y ∈ h, X ∈ g, (10)
sion of a quantum algebra is a Lie bialgebra. The opposite
can usually be proved too: starting with a Lie bialgebra and where h is the Lie algebra of the isotropy subgroup H . Since
requiring the homomorphism and cocycle conditions order the duals of the generators of the Lie bialgebra ξ j , X k  =
by order, we can define an associated quantum algebra as a δ j k can be interpreted as ‘infinitesimal’ coordinates on a
power series in the generators. This highlights the importance neighbourhood of the origin of the quantum group, then their
of Lie bialgebras for quantum algebras. Lie brackets (3) provide a first-order expansion of the com-
The nature of the difficulties that one encounters in going mutation relations of the nonabelian algebra of functions on
from the Lie bialgebra to the quantum algebra are best illus- the group. Then, if we call y the generators dual to Y ∈ h,
trated by considering the relation with quantum groups. A Eq. (10) states that the y generators can only appear on the
quantum group can be understood as a quantization of a right-hand side of the Lie brackets between y and any other
Poisson-Lie group. This is a Lie group on which a Poisson generators. Therefore, if we identify a linear complement of
structure (i.e. a Poisson bracket) is defined, as an operation h ∗ , its generators x will close a Lie subalgebra. Moreover, we
on the algebra of functions on the group, which is compatible can identify this subalgebra as the noncommutative algebra
with the group operations (i.e. they must be Poisson maps). of coordinates on the quantum homogeneous space. This is
The Poisson bracket on the Poisson-Lie group is the classical the ultimate meaning of coisotropy: asking that the coordi-
precursor of the noncommutative product between functions nates on the noncommutative space close algebraically.
on the associated quantum group. The procedure of ‘quanti- Since we will be interested in constructing noncommu-
zation’ is not unique because of ordering ambiguities could tative spacetimes associated with quantum group deforma-
appear due to the fact that Poisson-Lie structures are, in gen- tions, the rest of the paper we will be dealing with coisotropic
eral, nonlinear in terms of the local coordinates on the group. Lie bialgebras, which will define for our purposes the rel-
Lie bialgebras enter this picture because they are the Lie evant subset of physically relevant quantum groups. Our
algebras of Poisson-Lie groups and are in one-to-one corre- Lie bialgebras will be based on a central extension of the
spondence with them: the first-order expansion of the Poisson Poincaré and (A)dS algebras in (1 + 1) dimensions, which
bracket gives the (structure constants of a) cocommutator of means g = iso(1, 1) ⊕ so(2) in the Poincaré case and
the Lie bialgebra, and the fact that the product is a Pois- g = so(2, 1) ⊕ so(2) in the case of (A)dS. As isotropy
son map implies the 1-cocycle condition. Therefore, given a subalgebras we can either choose the Lorentz subalgebra
Lie algebra g, the knowledge of its Lie bialgebra structures so(1, 1) or the direct sum of the Lorentz and so(2) subal-
(g, δ) provides the “directions” along which we can obtain gebras, so(1, 1) ⊕ so(2). In the interpretation of G repre-
its quantum group and quantum algebra deformations. senting the symmetries of a gauge bundle on a homogeneous
spacetime, the first case corresponds to quotenting down to
2.3 Poisson/quantum homogeneous spaces the bundle, while the second case corresponds to quotient-
ing out the gauge coordinate as well and projecting down to
In the commutative case, a homogeneous space M is obtained spacetime. We will see in what follows, however, that this
as the quotient interpretation is not tenable in all cases.

123
615 Page 6 of 15 Eur. Phys. J. C (2018) 78:615

2.4 (1 + 1) Poincaré Lie bialgebras the Lie algebra dual to the coisotropic solution (15) is
obtained by dualizing δ, namely:
As a warmup, we will compute the most generic Lie bialgebra
structure on the Poincaré algebra in (1 + 1) spacetime dimen- [x 0 , x 1 ] = a 1 x 0 − a 0 x 1 , [χ , x 0 ] = b x 0 + a 0 χ ,
sions, iso(1, 1), and we will impose onto it the coisotropy [χ , x 1 ] = b x 1 + a 1 χ . (17)
condition for the Lorentz subalgebra. The iso(1, 1) Lie alge-
bra and its quadratic Casimir are These expressions provide the most generic first-order non-
commutative Minkowski spacetime that can be obtained
[K , P0 ] = P1 , [K , P1 ] = P0 , [P0 , P1 ] = 0, (11) from quantum deformations of the (1 + 1) Poincaré algebra
C= P02 − P12 , (12) and satisfy the coisotropy condition. The algebra of func-
tions on this noncommutative spacetime is generated by the
where P0 and P1 are the generator of time- and space- subalgebra (x 0 , x 1 ):
like translations and K is the generator of boosts. To find
the most general Lie bialgebra structure (iso(1, 1), δ) we [x 0 , x 1 ] = a 1 x 0 + a 0 x 1 , (18)
write a generic (pre-)cocommutator (2), and impose the 1-
jk
cocycle (4) and Jacobi conditions on the structure tensor f i . and x 0 , x 1 admit the interpretation of noncommutative coor-
These equations admit two independent solutions: dinate functions. The parameters a 0 and a 1 represent the
⎧   components of a vector which generalizes the well-known

⎪ δ(K ) = K ∧ a 0 P0 + a 1 P1 ,
⎨ κ-Minkowski noncommutative spacetime [19–22], which is
δ(P0 ) = a 1 P0 ∧ P1 + b K ∧ P0 , straightforwardly recovered in the time-like case a 0 = 1/κ,


⎩ a 1 = 0 (the light-cone quantum deformation [24] is also
δ(P1 ) = a 0 P1 ∧ P0 + b K ∧ P1 ,
⎧   obtained when a 0 = ± a 1 ). Moreover, it is easy to check

⎪ δ(K ) = K ∧ a 0 P0 + a 1 P1 + c P1 ∧ P0 , that the algebra (18) is isomorphic to the a 1 = 0 case pro-

(13) vided that a 0 = 0. κ-Minkowski turns out to be the only
⎪ δ(P0 ) = a 1 P0 ∧ P1 ,

⎩ noncommutative Minkowski spacetime that can be obtained
δ(P1 ) = a 0 P1 ∧ P0 . as the homogeneous space of a quantum group. As we will
see in the sequel, this will drastically change when quantum
depending each of them on three real parameters: (a 0 , a 1 , b) deformations of a central extension of the Poincaré algebra
and (a 0 , a 1 , c). are considered.
Since we will be interested in the (1 + 1) noncommuta-
tive Minkowski spacetimes arising from H being the Lorentz
subgroup, the coisotropy condition w.r.t. the Lorentz subal-
3 Extended (1 + 1) noncommutative Minkowski
gebra reads:
spacetimes

δ(K ) ⊂ K ∧ X, X ∈ g. (14) The aim of this section is to perform the same Lie bialgebra
analysis for the case of the one-dimensional central extension
It is immediate to check that this condition generates the of the (1 + 1) Poincaré Lie algebra, which is defined by the
constraint c = 0 on the second solution, which now is Lie brackets and Casimir operators
converted into the b = 0 subcase of the first family. We
thus conclude that the most general bialgebra deformation of [K , P0 ] = P1 , [K , P1 ] = P0 , [P0 , P1 ] = 0,
iso(1, 1) which is coisotropic w.r.t. the Lorentz subalgebra [Q, · ] = 0, (19)
is C1 = Q, C2 = P02 − P12 . (20)
⎧  

⎪ δ(K ) = K ∧ a 0
P + a 1
P ,
⎨ 0 1 The generator Q defines a one-dimensional abelian subal-
δ(P0 ) = a 1 P0 ∧ P1 + b K ∧ P0 , (15) gebra, which is often called a pseudo-extension or ‘trivial’


⎩ central extension, due to the fact that the extended Lie alge-
δ(P1 ) = a 0 P1 ∧ P0 + b K ∧ P1 .
bra can be written as a direct sum of the initial Lie algebra
Introducing a dual basis for g ∗ through the pairing plus the extra generator (see [37] and references therein).
We also recall that the algebra (19) plays a relevant role in
x 0 , P0  = 1, x 1 , P0  = 0, χ , P0  = 1, (1 + 1) gravity and is sometimes called the Nappi-Witten
algebra (see [48,49]). Note that the dual Lie algebra g ∗
x , P1  = 0,
0
x , P1  = 1,
1
χ , P1  = 0, (16) contains a fourth generator q which, in the interpretation
x , K  = 0,
0
x , K  = 0,
1
χ , K  = 1, of (19) representing the symmetries of a U (1) gauge bundle

123
Eur. Phys. J. C (2018) 78:615 Page 7 of 15 615

on Minkowski spacetime, would stand for the gauge coordi- 3.2 Coisotropy with respect to the Lorentz subalgebra
nate. so(1, 1)

The coisotropy condition (14) with respect to the Lorentz


3.1 The general solution subalgebra generated by K implies that the terms P1 ∧ P0 ,
P1 ∧ Q and P1 ∧ Q in δ(K ) in (22) need to vanish, which is
The most general pre-cocommutator on (19) depends on 24 tantamount to say
real parameters



⎪ δ(K ) = k1 K ∧ P1 + k2 K ∧ P0 + k3 K ∧ Q + k4 P1 ∧ P0 + k5 P1 ∧ Q + k6 P0 ∧ Q,


⎨ δ(P ) = h K ∧ P + h K ∧ P + h K ∧ Q + h P ∧ P + h P ∧ Q + h P ∧ Q,
0 1 1 2 0 3 4 1 0 5 1 6 0
(21)

⎪ δ(P ) = p K ∧ P + p K ∧ P + p K ∧ Q + p P ∧ P + p P ∧ Q + p 6 P0 ∧ Q,


1 1 1 2 0 3 4 1 0 5 1
⎩ δ(Q) = m K ∧ P + m K ∧ P + m K ∧ Q + m P ∧ P + m P ∧ Q + m P ∧ Q.
1 1 2 0 3 4 1 0 5 1 6 0

By imposing the cocycle condition we go down to 14:



⎪ δ(K ) = k1 K ∧ P1 + k2 K ∧ P0 + k4 P1 ∧ P0 + k5 P1 ∧ Q + k6 P0 ∧ Q,


⎨ δ(P ) = h K ∧ P − k P ∧ P + h P ∧ Q + h P ∧ Q,
0 2 0 1 1 0 5 1 6 0
(22)

⎪ δ(P ) = h K ∧ P + k P ∧ P + h P ∧ Q + h 5 P0 ∧ Q,


1 2 1 2 1 0 6 1
⎩ δ(Q) = m P ∧ P ,
4 1 0

and by enlarging the dual basis (16) with the dual q of the
k4 = k5 = k6 = 0. (26)
central generator Q

As expected, these constraints ensure that {x 0 , x 1 , q} gener-


q, P0  = q, P1  = q, K  = 0, q, Q = 1, (23)
ate a Lie subalgebra within (24), which will be our generic
extended noncommutative spacetime, namely
the pre-cocommutator gives rise to the following Lie bracket:

[x 0 , x 1 ] = k1 x 0 − k2 x 1 − k4 χ − m 4 q, [x 1 , χ ] = −h 2 x 1 − k1 χ ,
[x 0 , χ ] = −h 2 x 0 − k2 χ , [x 1 , q] = h 5 x 0 + h 6 x 1 + k5 χ , (24)
[x , q] = h 5 x + h 6 x + k6 χ ,
0 1 0 [χ , q] = 0.

Finally, by imposing Jacobi identities onto (24) we get the


[x 0 , x 1 ] = k1 x 0 − k2 x 1 − m 4 q, [x 0 , q] = h 5 x 1 + h 6 x 0 ,
set of nonlinear equations:
[x 1 , q] = h 5 x 0 + h 6 x 1 ,
(27)
h 2 k4 = 0, h 6 k1 + h 5 k2 − h 2 k5 = 0, h 6 k4 = 0,
h 2 m 4 = 0, h 5 k1 + h 6 k2 − h 2 k6 = 0, h 6 m 4 = 0, which indeed is a non-trivial extension of the algebra (18)
(25) provided that the constrains (25) are taken into account.
Let us call:

which characterize the complete family of Lie bialgebra


structures on the centrally extended (1 + 1) Poincaré alge- k1 = a 1 , k2 = a 0 , m 4 q = 10 = −01 (28)
bra (19). A glimpse on (25) shows that many solutions exist,
but once the coisotropy conditions have been imposed we so that the first Lie bracket of the dual basis is of the form
will be able to single out a smaller number of solutions - [x μ , x ν ] = a μ x ν −a ν x μ +μν , where a μ is a constant vec-
those with potential physical interest. tor and μν is an antisymmetric (Lie-algebra-valued) matrix

123
615 Page 8 of 15 Eur. Phys. J. C (2018) 78:615

μν . Moreover, let’s call


[x 0 , x 1 ] = 0, [x 0 , q] = b0 x 0 + b1 x 1 , [x 1 , q] = b1 x 0 + b0 x 1 .
(35)
h 6 = b0 , h 5 = b1 , h 2 = ϕ, (29)

Then the (now coisotropic) pre-cocommutator (22) reads1 A.2 If a μ is lightlike, a μ aμ = 0, then the vector bμ is light-
like too and provides only one free parameter. Using
δ(Pμ ) = Pμ ∧ (a ν Pν ) + ϕ K ∧ Pμ + bμ P0 ∧ Q ‘light-cone coordinates’ v ± = √1 v 0 ± v 1 the two
2
ν
+ μ bν P1 ∧ Q, constraints can be written
δ(K ) = K ∧ (a μ Pμ ), δ(Q) = m 4 P1 ∧ P0 . (30)
a0
where the convention for the Minkowski metric is η00 = +1, b0 = − b1 , (36)
a1
η11 = −1 and of course η01 = η10 = 0. With this notation,
the Lie brackets of the extended noncommutative spacetime
so we see that a ± = 0 implies b∓ = 0. The two vectors
read
can be parametrized by their time component, and if we
[x 0 , x 1 ] = a 1 x 0 − a 0 x 1 + 01 , [x 0 , q] = b0 x 0 + b1 x 1 , write a μ = a 0 (1, ±1) that implies bμ = b0 (1, ∓1).
(31) Therefore we have the following couple (a 0 , b0 , ϕ) of
[x 1 , q] = b1 x 0 + b0 x 1 ,
3-parameter families of cocommutators:
and the rest of the brackets for the dual Lie algebra are

μ μ μ ⎪
⎪ δ(K ) = a 0 K ∧ (P0 ± P1 ),
[χ , x ] = ϕ x + a χ , [χ , q] = 0. (32) ⎪

⎨ δ(P ) = ϕ K ∧ P ∓ a 0 P ∧ P + b (P ∓ P ) ∧ Q,
0 0 1 0 0 0 1
The constraints (25), after imposing the coisotropy condi- ⎪
⎪ δ(P1 ) = ϕ K ∧ P1 + a P1 ∧ P0 + b0 (P1 ∓ P0 ) ∧ Q,
0


tions (26), can be written as ⎩
δ(Q) = 0,
b0 a 1 + b1 a 0 = 0, b1 a 1 + b0 a 0 = 0, (37)
ϕ 10 = 0, b0 10 = 0. (33)
each of which gives rise to a non-central extension by
We now study in full generality the solutions to these equa- q of a non-commutative Minkowskian spacetime:
tions. We have to distinguish two main cases (A and B), which
in turn divide into a few sub-cases. The numerical vector a μ
and the antisymmetric (Lie-algebra-valued) matrix μν will [x 0 , x 1 ] = a 0 (±x 0 + x 1 ), [x 0 , q] = b0 (x 0 ∓ x 1 ),
allow to interpret the dual algebra as, respectively, κ-vector- [x 1 , q] = b0 (∓x 0 + x 1 ),
like and canonical θ -noncommutative spacetimes. (38)

Case A. Vanishing 10 parameter.


A.3 Finally, if a μ is time- or space-like, then bμ = 0 and
we are left with this 3-parameter (a 0 , a 1 , ϕ) family of
If m 4 = 0 and therefore 10 = 0 we are only left with the
deformations:
first two of Eqs. (33). There are three sub-cases:
⎧ μ
A.1 If a μ = 0, then b0 and b1 are two free parameters, ⎨ δ(K ) = K ∧ (a Pμ ),

thus giving rise to a 3-parameter family of solutions δ(Pμ ) = Pμ ∧ (a ν Pν ) + ϕ K ∧ Pμ , (39)


(b0 , b1 , ϕ); δ(Q) = 0,

⎨ δ(K ) = 0,
⎪ whose associated noncommutative spacetime is a triv-
ν ial central extension of the κ-Minkowski spacetime:
δ(Pμ ) = ϕ K ∧ Pμ + bμ P0 ∧ Q + μ bν P1 ∧ Q,


δ(Q) = 0.
(34) [x 0 , x 1 ] = a 1 x 0 − a 0 x 1 , [x μ , q] = 0, (40)

We have a non-central extension by q of a commutative which would be the usual type of structure arising
Minkowski spacetime: within a quantum principal bundle construction, where
the quantum gauge group and the noncommutative
1 Notice that ν = ημρ ρν , so that ρb = δμ 0 b1 + δμ 1 b0 . base space commute between themselves.
μ μ ρ

123
Eur. Phys. J. C (2018) 78:615 Page 9 of 15 615

Case B. Nonzero ‘canonical noncommutativity parameter’ subgroup generated by I = Span{K , Q}. In this case the
10 coisotropy condition implies that

If m 4 = 0 and therefore 10 = 0, the constraints (33) imply δ(I ) ⊂ I ∧ X, I = Span{K , Q}, X ∈ g. (46)
that ϕ = b0 = 0. Then the remaining constraints reduce to
and this condition will guarantee that the space and time
b1 a 0 = 0, b1 a 1 = 0. (41) translation coordinates will close a Lie subalgebra indepen-
dently of the q generator (therefore, we will not have an
In this situation we have only the following two cases: extended noncommutative spacetime). The corresponding
conditions on the generic precocommutator (22) are
B.1 When a μ = 0, we have a 2-parameter family in which
only 10 and b1 are nonzero: k4 = 0, m 4 = 0, (47)


⎪ δ(K ) = 0,

⎪ and we obtain
⎨ δ(P ) = b P ∧ Q,
0 1 1
(42) ⎧

⎪ δ(P1 ) = b1 P0 ∧ Q, ⎪ δ(K ) = k1 K ∧ P1 + k2 K ∧ P0 + k5 P1 ∧ Q + k6 P0 ∧ Q,

⎪ ⎪

⎩ δ(Q) = m P ∧ P , ⎨ δ(P ) = h K ∧ P − k P ∧ P + h P ∧ Q + h P ∧ Q,
4 1 0 0 2 0 1 1 0 5 1 6 0

⎪ δ(P ) = h K ∧ P + k P ∧ P + h P ∧ Q + h 5 P0 ∧ Q,


1 2 1 2 1 0 6 1
and the associated noncommutative spacetime reads δ(Q) = 0,
(48)
[x 0 , x 1 ] = 01 , [x 0 , q] = b1 x 1 , [x 1 , q] = b1 x 0 .
(43) together with two further constrains coming from Jacobi
identities:
When both parameters are different from zero, the
Lie algebra (43) is isomorphic to the sl(2) Lie alge-
h 6 k1 + h 5 k2 − h 2 k5 = 0, h 5 k1 + h 6 k2 − h 2 k6 = 0.
bra. The b1 = 0 case leads to the well-known
θ -noncommutative spacetime in (1+1) dimensions (49)
(see [41–44] and references therein), which can be thus
considered as an extended noncommutative Minkowski Calling k6 = c0 and k5 = c1 we get
spacetime which is invariant under a certain quantum ⎧
deformation of the extended (1 + 1) Poincaré group. ⎪
⎪ δ(K ) = K ∧ (a μ Pμ ) + (cμ Pμ ) ∧ Q,


B.2 When b1 = 0, we obtain a 3-parameter family of defor- ⎨ δ(P ) = P ∧ (a ν P ) + ϕ K ∧ P
μ μ ν μ
mations ρσ (50)
⎪ + bμ P0 ∧ Q + ημρ
⎪ bσ P1 ∧ Q,
⎧ ⎪

μ ⎩ δ(Q) = 0,
⎨ δ(K ) = K ∧ (a Pμ ),

δ(Pμ ) = Pμ ∧ (a ν Pν ), (44)

⎩ and the commutators of the dual Lie algebra read
δ(Q) = m 4 P1 ∧ P0 ,
[x 0 , x 1 ] = a 1 x 0 − a 0 x 1 , [χ , x μ ] = ϕ x μ + a μ χ ,
which induces the most general kind of spacetime
[x 0 , q] = b0 x 0 + b1 x 1 + c0 χ , [χ , q] = 0, (51)
noncommutativity (vector-like κ term plus -term),
namely: [x 1 , q] = b1 x 0 + b0 x 1 + c1 χ ,
together with the constraints
[x 0 , x 1 ] = a 1 x 0 − a 0 x 1 + 01 , [x μ , q] = 0. (45)

Note that when a μ = 0 we have a non-trivial central b0 a 1 + b1 a 0 − ϕ c1 = 0, b1 a 1 + b0 a 0 − ϕ c0 = 0, (52)


extension of the vector-like κ-deformation, while in
the case a μ = 0 we recover again the θ -deformation. that by using light-cone coordinates can be written as

3.3 Coisotropy with respect to so(1, 1) ⊕ so(2) b+ a + = √1 ϕ c+ , b− a − = √1 ϕ c− . (53)


2 2

We could also consider as the isotropy subgroup for the The solutions of the above equations will be split into the
homogeneous space the central extension of the Lorentz cases when ϕ vanishes and when it doesn’t. Note that, as it

123
615 Page 10 of 15 Eur. Phys. J. C (2018) 78:615

was expected from the coisotropy condition, {x 0 , x 1 } always and the Jacobi identities lead to the conditions:
generates a Lie subalgebra, and its commutator
h 5 k2 − h 2 k5 = 0, h 5 k1 − h 2 k6 = 0,
[x , x ] = a x − a x ,
0 1 1 0 0 1
(54)  (k2 k6 − k1 k5 ) = 0, (59)

which are more restrictive than in the Poincaré case, as in the


is just the generalized κ-Minkowski spacetime.
 → 0 limit the last equation collapses to a tautology. For
 = 0, the equations above can be rewritten in the following
form:
4 Extended (1 + 1) noncommutative (A)dS spacetimes

⎨ a = (h 5 , h 2 )
It seems natural to wonder which kind of extended noncom- a ×b = 0, a ×  c = 0,
c = 0, b× b = (k5 , k2 ) , (60)
mutative spacetimes can be obtained when the cosmologi- ⎩
c = (k6 , k1 )
cal constant is non-vanishing, i.e., by considering centrally
extended (A)dS Lie bialgebras. This analysis can be per-
where the planar Euclidean vector product is defined as u ×
formed on the same footing as before, by taking into account
v = u 1 v2 − u 2 v1 , and its vanishing implies that the two
that the Lie brackets and Casimir operators for the centrally
2D vectors are parallel. Therefore the Eq. (60) imply that all
extended (A)dS algebra are given by
three vectors are nonzero and parallel to each other:
[K , P0 ] = P1 [K , P1 ] = P0 [P0 , P1 ] = − K
a = a (cos θ, sin θ ) , b = b (cos θ, sin θ ) ,
[Q, · ] = 0, (55)
c = c (cos θ, sin θ ) . (61)
C1 = P02 − P12 −  K 2 , C2 = M, (56)
Thus, they can be parametrized by their magnitudes a, b and
where the case of  > 0 corresponds to the dS case ( < 0
c (which could be zero), and by their common direction angle
is the AdS one), and the limit  → 0 leads to the centrally
θ.
extended Poincaré algebra (19).
In this case too, the extended (1 + 1) dimensional (A)dS
After imposing the cocycle condition on the general
spacetimes M are obtained as homogeneous spaces by taking
ansatz (21) we get
the Lorentz subgroup generated by K as the invariance sub-
⎧ group of the origin. Therefore, if we impose the coisotropy

⎪ δ(K ) = k1 K ∧ P1 + k2 K ∧ P0 + k5 P1 ∧ Q + k6 P0 ∧ Q,

⎪ condition with respect to K onto the cocommutator (57) we
⎨ δ(P0 ) = h 2 K ∧ P0 + h 5 P1 ∧ Q − k1 P1 ∧ P0 −  k6 K ∧ Q,
get that k5 = k6 = 0, which leads to
⎪ δ(P1 ) =  k5 K ∧ Q + k2 P1 ∧ P0 + h 2 K ∧ P1 + h 5 P0 ∧ Q,




δ(Q) = 0, h 5 k2 = h 5 k1 = 0, (62)
(57)
π
[which means that θ = 2 in (60)], and we have two cases:
therefore we immediately see that the central generator Q
always has vanishing cocommutator, which immediately pre-
C.1 When k1 = k2 = 0 and h 5 , h 2 are arbitrary we have
cludes the existence of the θ -noncommutativity in any of
the extended noncommutative (A)dS spacetimes. Notice also ⎧
that the solution above is more restrictive than the solu- ⎪
⎪ δ(K ) = 0,


tion one obtains in the Poincaré  = 0 case (22), because ⎨ δ(P ) = h K ∧ P + h P ∧ Q,
0 2 0 5 1
(63)
some equations following from the cocycle condition vanish ⎪
⎪ δ(P1 ) = h 2 K ∧ P1 + h 5 P0 ∧ Q,
altogether when  → 0. For instance, the solution above ⎪

⎩ δ(Q) = 0,
reduces to the solution (22) in the  → 0 limit when
m 4 = k4 = h 6 = 0.
Thus, the dual Lie algebra to the cocommutator (57) is which leads to the extended noncommutative spacetime

[x 0 , x 1 ] = 0, [x 1 , q] = h 5 x 0 , [x 0 , q] = h 5 x 1 . (64)
[x 0 , x 1 ] = k1 x 0 − k2 x 1 , [x 1 , χ ] = −k1 χ − h 2 x 1 ,
[x 1 , q] = h 5 x 0 + k5 χ , [x 0 , χ ] = −k2 χ + h 2 x 0 ,
When h 5 = 0 we have a non-central extension of a
[x 0 , q] = h 5 x 1 + k6 χ , [q, χ ] =  k6 x 0 −  k5 x 1 , commutative spacetime, that would correspond in the
(58) Poincaré case to the space A.1 with b0 = 1.

123
Eur. Phys. J. C (2018) 78:615 Page 11 of 15 615

C.2 When h 5 = 0 and k1 , k2 , h 2 are arbitrary we get tive Minkowski spacetime:



⎪ δ(K ) = k1 K ∧ P1 + k2 K ∧ P0 , [x μ , x ν ] = μν , (68)



⎨ δ(P ) = h K ∧ P − k P ∧ P ,
where μν commutes with x μ . In order for the right-hand
0 2 0 1 1 0
(65)

⎪ δ(P ) = ∧ + ∧ 1,


1 k P
2 1 P0 h 2 K P side of [x 0 , x 1 ] to take that form and to commute with x μ ,
⎩ δ(Q) = 0,
all the parameters of case B other than m 4 (recall that 01 =
m 4 q) have to be put to zero (so we are considering a sub-case
and we obtain a centrally extended noncommutative of B.1, when b1 = 0 or a sub-case of B.2, when a μ = 0).
(A)dS spacetime This is what the coalgebra and its dual Lie algebra look like:

⎨ δ(K ) = 0,
⎪ [x 0 , x 1 ] = 01 ,
[x 0 , x 1 ] = k1 x 0 − k2 x 1 , [x μ , q] = 0. (66)
δ(P0 ) = δ(P1 ) = 0, [x μ , q] = [χ , q] = 0, (69)

⎩ μ
δ(Q) = m 4 P1 ∧ P0 , [χ , x ] = 0.
which corresponds to the case A.3 in the Poincaré clas-
sification. The above Lie bialgebra can be interpreted as the infinitesi-
mal version of a quantum (1 + 1) extended Poincaré group,
It is also worth mentionining that if we impose coisotropy whose quotient with respect to its Lorentz subgroup gives
with respect to the extended isotropy subgroup (K , Q), this the noncommutative spacetime (68). We stress that the non-
condition provides no further constraints with respect to the coboundary nature of the cocommutator δ in (69) can be
general solution (57), and the noncommutative spacetime straightforwardly proven by taking into account that Q is
would be again central and, therefore, any possible classical r -matrix would
give δ(Q) = 0 when applying (5).
[x 0 , x 1 ] = k1 x 0 − k2 x 1 , (67) In fact, we know that δ provides the first order of the
deformed coproduct z , which in this case can be shown to
have no higher order terms. This means that the full coproduct
which is exactly the same result (54). Summarizing, the
of the quantum (1 + 1) extended Poincaré group is
(A)dS case turns out to be much more restrictive than the ⎧
Poincaré case. ⎪
⎪ z (K ) = 1 ⊗ K + K ⊗ 1,


At this point it could seem surprising that the noncom- ⎨  (P ) = 1 ⊗ P + P ⊗ 1,
z 0 0 0
mutative (A)dS spacetimes (64) and (66) do not depend on (70)
⎪ 
⎪ z 1(P ) = 1 ⊗ P + P1 ⊗ 1,
the cosmological constant  and, therefore, coincide with ⎪

1
⎩  (Q) = 1 ⊗ Q + Q ⊗ 1 + z P ∧ P ,
the corresponding noncommutative Minkowski spacetimes. z 1 0
Indeed, this is true only at first order, and higher order con- where z = m 4 , and (70) can be easily shown to be an algebra
tributions depending on  are expected to appear when the homomorphism with respect to the undeformed commutation
full quantum coproduct z is constructed and the all-orders rules (19). Therefore, since the coproduct (70) has only a first-
noncommutative spacetime is obtained by applying the full order deformation, the computation of the full Hopf algebra
Hopf algebra duality or, alternatively, by quantizing the all- duality cannot give rise to terms other than [x 0 , x 1 ] = 01 ,
orders Poisson-Lie group whose linearization corresponds and we can conclude that the (1 + 1) canonical nocommuta-
to the extended noncommutative spacetimes here presented tive spacetime is a true quantum homogeneous space for the
(see [50–52] for explicit examples, including the noncom- extended Poincare quantum group.
mutative κ-(A)dS spacetime in (2 + 1) dimensions, which This result seems to be quite interesting, because it is well-
turns out to be a nonlinear deformation of the κ-Minkowski known that the noncommutative spacetimes of the form (68)
spacetime). can be shown to be obtained when a twisting element of
the Poincaré algebra acts on the algebra of functions of the
commutative Minkowski spacetime (see [27,34,53–55] and
5 Canonical noncommutative spacetime as a quantum references therein), which acts covariantly on the algebra of
space functions generated by the relations (68). However, the latter
twisted symmetry does not allow any rigorous interpretation
In the case of deformations of the trivial central extension of the canonical noncommutative spacetime as a quantum
of the Poincaré algebra ( = 0), we have case B, in which homogenous space, in contradistinction to the one that we
the ‘canonical noncommutativity’ generator 10 is nonzero. have just presented.
Such case admits the interpretation of the bialgebra of sym- Our approach allows to understand the canonical noncom-
metries of the (1 + 1-dimensional) canonical noncommuta- mutative spacetime (68) as a Lie algebra of noncommutative

123
615 Page 12 of 15 Eur. Phys. J. C (2018) 78:615

coordinates, because the antisymmetric matrix μν on the product of Lie groups. However, the additional structures
right-hand side of its commutation relations is promoted to a introduced by quantum group symmetries allow for more
matrix of central generators. In this way we can see the alge- general ways in which gauge and spacetime symmetries may
bra of x μ and 01 as a subalgebra of the dual Lie algebra combine in a hybrid manner when they act on tensor prod-
to a Lie bialgebra of symmetries. In this sense we can show uct states. In this sense one can conceive, in the framework
that in (1 + 1) dimensions the canonical noncommutative of quantum groups, a more genuine unification of ‘internal’
Minkowski spacetime is genuinely a quantum homogeneous and ‘external’ symmetries. As part of a conjoined algebraic
space of a certain quantum extended Poincaré group, and the structure, gauge and spacetime transformations can ‘mix up’
central extension of the latter turns out to be essential since it when acting on tensor product states, making it impossible to
provides the additional noncommutative coordinate that can make a purely-gauge or a purely-spacetime transformation.
be identified with the 01 generator. Here we have explored this possibility in the case of (1 +
One could hope that the same structure generalizes to 1) spacetime dimensions, and of an Abelian, 1-dimensional
higher dimensions, provided that a dual Lie algebra of the gauge group [e.g. S O(2)]. In the commutative case, the CMT
form implies that the most general group of symmetries of an
S O(2) QFT on a Minkowski background is the direct product
[x μ , x ν ] = μν , [μν , x μ ] = [μν , ρσ ] = 0 (71) I S O(1, 1) × S O(2). We relaxed the assumption of having a
Lie algebra of symmetries, into having a Lie bialgebra which
can be found as a subalgebra of the dual to some Lie bialgebra is coisotropic w.r.t. the Lorentz subalgebra (and therefore
δ. This conjecture can be checked explicitly for Lie bialge- admits a homogeneous quantum Minkowski spacetime as a
bras δ of the higher dimensional centrally extended Poincaré quotient). The commutation rules for the spacetime, x, and
algebra. The generators μν of (71) will be dual to a number gauge, q coordinates that we obtain belong to the following
of new generators Tμν which are all assumed to be central five cases:
extensions of a Poincaré algebra with the suitable dimension:
A.1 [x, x] = 0, [x, q] ⊆ x: commutative spacetime, gauge
[Tμν , Pρ ] = [Tμν , Mρσ ] = [Tμν , Tρσ ] = 0. (72) coordinate acting on x as vector field;
A.2 [x, x] ⊆ x, [x, q] ⊆ x: noncommutative spacetime,
Now, the conditions (71) translate into the following assump- gauge coordinate acting on x as vector field;
tions on the cocommutators: A.3 [x, x] ⊆ x, [x, q] = 0: trivial quantum principal bun-
dle on noncommutative spacetime;
1. Terms of the form Pμ ∧ Pν can only appear in δ(Tμν ). B.1 [x, x] ⊆ q, [x, q] ⊆ x: canonical-like noncommuta-
2. Terms of the form Pμ ∧ Tρσ and Tμν ∧ Tρσ cannot appear tive spacetime, with right-hand-side non-primitive;
in any Lie bialgebra cocommutator. B.2 [x, x] ⊆ q + x, [x, q] = 0: mixture of canonical and
κ-like noncommutative spacetime.
Under such conditions we can compute the most general Lie
bialgebra in (2 + 1) and (3 + 1) dimensions which is generated If we want to interpret our starting point, the algebra
by Pμ , Mμν and Tμν . It turns out that these assumptions are iso(1, 1) ⊕ so(2), as the algebra of symmetries of a U (1)
strong enough that the cocycle conditions impose that Tμν gauge theory on Minkowski space, then it is natural to iden-
is primitive, i.e. δ(Tμν ) = 0, both in (2 + 1) and (3 + 1) tify the coordinates {x μ , q} as functions on the (trivial) prin-
dimensions. But then, since Pμ ∧ Pν appears only in δ(Tμν ), cipal fibre bundle M × U (1) (where M is Minkowski space),
this implies that [x μ , x ν ] = 0, and we end up with a trivial which indeed is the geometrical construction underlying a
solution. U (1) gauge theory. In the literature, one can find a notion of
quantum principal bundle [44,56–64], but this notion seems
to be too restrictive: the gauge group is assumed to close a
6 Outlook and discussion quantum group on its own [which can only be trivial in the
case of S O(2)], and the coordinate q on the gauge group is
In this paper we have shown that quantum group deforma- assumed to commute with the noncommutative coordinates
tions of centrally extended Lorentzian symmetries can be of the base space. This situation is realized only in our case
considered as a possible way to circumvent the no-go CMT, A.3, in which the base space is (generalized) κ-Minkowski,
since they allow relaxing the assumption of Leibniz action and the fibre is a standard commutative U (1). But our con-
of symmetry generators on tensor product states. The main struction allows for more exotic possibilities, since q can be
conclusion of the CMT is that the spacetime symmetry group non-commuting with respect to the coordinates x: in case
and the gauge group, when acting on a QFT, can be unified A.2 we have a κ-Minkowski spacetime with a gauge coordi-
into a combined structure only in a trivial way, i.e. as a direct nate that acts on the spacetime coordinates, [x, q] ⊆ x, and

123
Eur. Phys. J. C (2018) 78:615 Page 13 of 15 615

in case A.1 we have the same thing, this time on a commu- commutative spacetime arising from a quantum group sym-
tative spacetime. Cases B are stranger yet: the q coordinate metry, one needs to accept a form of “Hopf-algebrization”
appears on the right-hand-side of the commutation relation of the standard formulation of field theory, in which scalar
of the x, in a term that reminds of the canonical noncom- fields are considered as the element of an algebra which, in
mutative spacetimes (see below). However, in the case B.1 the commutative case, is the (Abelian) algebra of functions
the coordinate q does not commute with x and therefore the on the spacetime manifols, and, in the noncommutative case,
algebra has nothing to do with the canonical spacetimes. The turns into a non-Abelian one.
last case, B.2 is the only one which can be interpreted in Finally, our last exploration involved considering the pos-
some sense as a generalization of the canonical spacetimes sibility that the additional coordinate q has nothing to do
(in which the commutation relations are a linear combination with a coordinate on (a neighbourhood of the identity of) the
of those of κ-Minkowski and the canonical ones). gauge group, and instead it could be interpreted in the context
After the main analysis described above, we also consid- of canonical noncommutative spacetimes. These are well-
ered the possibility of requiring coisotropy w.r.t. the sum of studied noncommutative spacetimes [27,34,53–55] in which
Lorentz and so(2) subalgebras. This ensures that the space- the commutator between two spacetime coordinates is equal
time coordinates always close a generalized κ-Minkowski to a constant times the identity operator: [x μ , x ν ] = θ μν 1
algebra [x 0 , x 1 ] = a 1 x 0 − a 0 x 1 , while the commutation (much like the Heisenberg commutation relations [ p, q] =
rules of x with q in principle can contain the Lorentz group i h̄). Such spacetimes have interesting properties, e.g. the
coordinate χ . These commutation rules depend on five real coordinates respect a Heisenberg uncertainty principle [65],
parameters which satisfy two quadratic constraints. In each the corresponding QFTs present the phenomenon of IR/UV
case the commutators [x, q] include a term proportional to χ . mixing, and the only known example of QFT defined at all
The dual statement is that the cocommutator of the Lorentz scales, the Grosse–Wulkenhaar model [66–69], is defined on
generator K contains terms of the kind Q ∧ Pμ , which mix such a noncommutative background. However, the commu-
spacetime and gauge generators. tation relations [x μ , x ν ] = θ μν 1 are not usually considered
It was then also natural to extend the previous analysis to in a Lie algebra framework, thus precluding the interpreta-
centrally extended (1 + 1)-dimensional (A)dS groups with tion of θ -spacetimes as quantum homogeneous spaces under
nonvanishing cosmological constant . We found that this a given quantum-group.
(as is often the case) is more restrictive than Poincaré, and In this paper, we presented an alternative possibility that
only two classes of solutions emerged: allows to interpret the (1 + 1)-dimensional canonical space-
time as the homogeneous space of a quantum Poincaré group.
C.1 [x, x] = 0, [x, q] ⊆ x, This can be achieved by considering the right-hand side of
C.2 [x, x] ⊆ x, [x, q] ⊆ 0, [x μ , x ν ] = θ μν as the dual noncommutative coordinate asso-
ciated to central generator of the extended (1 + 1) Poincaré
which tend, in the  → 0 limit, to, respectively, a subcase Lie algebra. In this way, the canonical commutation rule can
of A.1 and case A.3. In other words, we found that only A.1 be obtained as one of the brackets of a dual extended Poincaré
and A.3 admit a generalization to curved spacetime. Lie bialgebra. We then constructed the quantum group asso-
In summary, with the five extended noncommutative ciated to this Lie bialgebra, and the coproducts are all primi-
Minkowski spacetimes above described, we revealed a new tive except for that of the central generator Q. This approach
quantum-algebraic possibility to have putative quantum- to describing the symmetries of the canonical spacetime is
gravitational effects that circumvent the limitations of the an alternative one to that of twisting the Poincaré group pre-
Coleman–Mandula theorem, which might work even on a sented in [27,54,55], and, unlike it, allows to understand
commutative spacetime (case A.1). The main message is the noncommutative spacetime as a quantum homogeneous
that the full algebra of functions on the principal bundle space. Finally, we tested whether a similar approach could
M × U (1) can be noncommutative by assuming quantum be used in more than (1 + 1) dimensions. Unfortunately we
group Poincaré symmetries to hold, and can be so in such a encountered an obstruction: for d > 1, there is no bialge-
way that the simultaneous determination of the gauge coor- bra deformation of the (d + 1)-dimensional Poicaré algebra
dinate q and spacetime coordinates x is subject to quantum extended with d(d + 1)/2 central generators, in which said
limitations (because [x, q] = 0). This is highly suggestive of generators have a nonzero cocommutator. This means that the
situations in which we have to renounce either to the perfect dual Lie algebra can never be of the form [x μ , x ν ] = μν
knowledge of the value of some charged field or of the space- with [x ρ , μν ] = [μν , ρσ ] = 0. This shows that only in
time coordinate at which it is calculated. However such an (1 + 1) dimensions the canonical noncommutative spacetime
interpretation would require a suitable formulation of non- can be obtained as a quantum space for a centrally extended
commutative gauge theory on fibre bundles with “hybrid” Poincaré algebra. Nevertheless this does not exclude the pos-
symmetries. Indeed, in order to talk about QFT on a non-

123
615 Page 14 of 15 Eur. Phys. J. C (2018) 78:615

sibility of recovering θ -noncommutative spacetimes as duals 22. J. Lukierski, H. Ruegg, Phys. Lett. B 329, 189 (1994)
of Lie bialgebra structures associated to other Lie algebras. 23. A. Ballesteros, F.J. Herranz, M.A. del Olmo, M. Santander, J. Math.
Phys 35, 4928 (1994)
Also, we would like to stress that the approach here pre- 24. A. Ballesteros, F.J. Herranz, M.A. del Olmo, C.M. Pereña, M. San-
sented for the construction of hybrid gauge symmetries can tander, J. Phys. A 28, 7113–7125 (1995)
be generalized to both higher dimensional kinematical Lie 25. A. Ballesteros, F.J. Herranz, M.A. del Olmo, M. Santander, Phys.
algebras t of spacetime symmetries and higher dimensional Lett. B 351, 137 (1995)
26. S. Zakrzewski, Commun. Math. Phys. 185, 285–311 (1997)
(non-abelian) gauge symmetries with gauge Lie algebra g. 27. J. Lukierski, M. Woronowicz, Phys. Lett. B 633, 116–124 (2006)
In such cases, the Lie bialgebra structures (b, δ) of the direct 28. J.E. Mandula, Scholarpedia 10(2), 7476 (2015)
sum Lie algebra b = t ⊕ g will in general intertwine both 29. R. Haag, J.T. Lopuszanski, M. Sohnius, Nucl. Phys. B 88, 257
symmetries, and for those cocommutators δ that fulfil the (1975)
30. D. Bernard, A. Le Clair, Commun. Math. Phys. 142, 99–138 (1991)
coisotropy condition, the corresponding dual algebra δ ∗ will 31. D.V. Vassilevich, Mod. Phys. Lett. A 21, 1279–1284 (2006)
contain as a subalgebra the (first order) noncommutative 32. M. Chaichian, A. Tureanu, Phys. Lett. B 637, 199 (2006)
coordinates in the associated noncommutative bundle. Work 33. M. Chaichian, A. Tureanu, G. Zet, Phys. Lett. B 651, 319–323
along these lines is in progress. (2007)
34. P. Aschieri, M. Dimitrijevic, F. Meyer, S. Schraml, J. Wess, Lett.
Math. Phys. 78, 61–71 (2006)
Acknowledgements A.B. has been partially supported by Ministerio
35. M. Lindner, S. Ohmer, Phys. Lett. B 773, 231–235 (2017)
de Economía, Industria y Competitividad (MINECO, Spain) under grant
36. A. Ballesteros, F.J. Herranz, M.A. del Olmo, M. Santander, J. Phys.
MTM2016-79639-P (AEI/FEDER, UE), by Junta de Castilla y León
A 26, 5801–5823 (1993)
(Spain) under grant VA057U16 and by the Action MP1405 QSPACE
37. J.A. de Azcárraga, J.C. Pérez Bueno, J. Math. Phys. 36, 6879–6896
from the European Cooperation in Science and Technology (COST).
(1995)
F.M. was funded by the European Union and the Istituto Italiano di
38. M.S. Dijkhuizen, T.H. Koornwinder, Geom. Dedicata 52, 291
Alta Matematica under a Marie Curie COFUND action.
(1994)
39. S. Zakrzewszki, Poisson homogeneous spaces, in: J. Lukierski, Z.
Open Access This article is distributed under the terms of the Creative
Popowicz, J. Sobczyk (eds.), Quantum Groups (Karpacz 1994),
Commons Attribution 4.0 International License (http://creativecomm
PWN Warszaw (1995) 629
ons.org/licenses/by/4.0/), which permits unrestricted use, distribution,
40. A. Ballesteros, C. Meusburger, P. Naranjo, J. Phys. A 50, 395202
and reproduction in any medium, provided you give appropriate credit
(2017)
to the original author(s) and the source, provide a link to the Creative
41. S. Doplicher, K. Fredenhagen, J.E. Roberts, Phys. Lett. B 331, 39
Commons license, and indicate if changes were made.
(1994)
Funded by SCOAP3 .
42. S. Doplicher, K. Fredenhagen, J.E. Roberts, Commun. Math. Phys.
172, 187 (1995)
43. R.J. Szabo, Phys. Rept. 378, 207–299 (2003)
44. P. Aschieri, P. Bieliavsky, C. Pagani, A. Schenkel, Commun. Math.
References Phys. 352, 287 (2017)
45. A. Ballesteros, E. Celeghini, F.J. Herranz, J. Phys. A 33, 3431
1. S. Coleman, J. Mandula, Phys. Rev. 159, 1251 (1967) (2000)
2. C.J. Fewster, Commun. Math. Phys. 357, 353–378 (2018) 46. A. Opanowicz, J. Phys. A 31, 2887 (1998)
3. S. Carlip, Quantum gravity in 2+1 dimensions (Cambridge Uni- 47. A. Ballesteros, F.J. Herranz, P. Parashar, J. Phys. A 30, L149 (1997)
versity Press, Cambridge, 2003) 48. C. Nappi, E. Witten, Phys. Rev. Lett. 71, 3751 (1993)
4. A. Achucarro, P.K. Townsend, Phys. Lett. B180, 89 (1986) 49. D. Cangemi, R. Jackiw, Ann. Phys. 225, 229–263 (1993)
5. E. Witten, Nucl. Phys. B311, 46 (1988) 50. A. Ballesteros, F.J. Herranz, N.R. Bruno, Phys. Lett. B 574, 276
6. L. Freidel, E.R. Livine, Phys. Rev. Lett. 96, 221301 (2006) (2003)
7. H.J. Matschull, M. Welling, Class. Quant. Grav. 15, 2981–3030 51. A. Ballesteros, F.J. Herranz, C. Meusburger, P. Naranjo, SIGMA
(1998) 10, 052 (2014)
8. V.G. Drinfel’d , Proc. Int. Congress of Math. 1, 798 (1987), ed A. 52. A. Ballesteros, N.R. Bruno, F.J. Herranz, Adv. High Energy Phys.
V. Gleason, Berkeley, AMS 7876942 (2017)
9. V. Chari, A. Pressley, A Guide to Quantum Groups (Cambridge 53. J. Wess. arXiv:hep-th/0408080 (2004)
University Press, Cambridge, 1994) 54. M. Chaichian, P. Kulish, K. Nishijima, A. Tureanu, Phys. Lett. B
10. S. Majid, Foundations of quantum group theory (Cambridge Uni- 604, 98–102 (2004)
versity Press, Cambridge, 2000) 55. F. Koch, E. Tsouchnika, Nucl. Phys. B 717, 387–403 (2005)
11. V. Drinfel’d, Soviet Math. Dokl. 27, 68–71 (1983) 56. T. Brzezinski, S. Majid, Phys. Lett. B 298, 339 (1993)
12. A.Y. Alekseev, A.Z. Malkin, Commun. Math. Phys. 169, 99 (1995) 57. T. Brzezinski, S. Majid, Commun. Math. Phys. 157, 591 (1993)
13. V.V. Fock, A.A. Rosly, Am. Math. Soc. Transl. 191, 67 (1999) 58. T. Brzezinski, S. Majid, Commun. Math. Phys. 167, 235 (1995).
14. C. Meusburger, B. Schroers, J. Math. Phys. 49, 083510 (2008) (erratum)
15. C. Meusburger, B. Schroers, Nucl. Phys. B 806, 462 (2009) 59. T. Brzezinski, J. Geom. Phys. 20, 349 (1996)
16. F. Mercati, M. Sergola. arXiv:1802.09483 (2018) 60. M. Pflaum, Commun. Math. Phys. 166, 279 (1994)
17. J. Lukierski, H. Ruegg, A. Nowicki, V.N. Tolstoy, Phys. Lett. B 61. M. Durdevic. arXiv:hep-th/9311029 (1993)
264, 331 (1991) 62. M. Durdevic, J. Phys. A 30, 2027–2054 (1997)
18. J. Lukierski, A. Nowicky, H. Ruegg, Phys. Lett. B 271, 321 (1991) 63. P.M. Hajac, Commun. Math. Phys. 182, 579 (1996)
19. P. Maslanka, J. Phys. A 26, L1251 (1993) 64. S.B. Sontz, Principal bundles: the quantum case (Springer, New
20. S. Majid, H. Ruegg, Phys. Lett. B 334, 348 (1994) York, 2015)
21. S. Zakrzewski, J. Phys. A 27, 2075 (1994)

123
Eur. Phys. J. C (2018) 78:615 Page 15 of 15 615

65. G. Amelino-Camelia, G. Gubitosi, F. Mercati, Phys. Lett. B 676, 68. H. Grosse, R. Wulkenhaar, Lett. Math. Phys. 71, 13 (2005)
180–183 (2009) 69. M. Buric, M. Wohlgenannt, JHEP 1003, 053 (2010)
66. H. Grosse, R. Wulkenhaar, JHEP 0312, 019 (2003)
67. H. Grosse, R. Wulkenhaar, Commun. Math. Phys. 254, 91–127
(2005)

123

You might also like