Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Available online at www.sciencedirect.

com

ScienceDirect
Nuclear Physics B 958 (2020) 115117
www.elsevier.com/locate/nuclphysb

The momentum spaces of κ-Minkowski


noncommutative spacetime
Fedele Lizzi a,b,c , Mattia Manfredonia a,b , Flavio Mercati a,b,∗
a Dipartimento di Fisica “Ettore Pancini”, Università di Napoli Federico II, Napoli, Italy
b INFN, Sezione di Napoli, Italy
c Departament de Física Quàntica i Astrofísica, and Institut de Cíencies del Cosmos (ICCUB), Universitat de
Barcelona, Barcelona, Spain
Received 8 May 2020; received in revised form 16 July 2020; accepted 17 July 2020
Available online 23 July 2020
Editor: Stephan Stieberger

Abstract
A useful concept in the development of physical models on the κ-Minkowski noncommutative space-
time is that of a curved momentum space. This structure is not unique: several inequivalent momentum
space geometries have been identified. Some are associated to a different assumption regarding the signa-
ture of spacetime (i.e. Lorentzian vs. Euclidean), but there are inequivalent momentum spaces that can be
associated to the same signature and even the same group of symmetries. Moreover, in the literature there
are two approaches to the definition of these momentum spaces, one based on the right- (or left-)invariant
metrics on the Lie group generated by the κ-Minkowski algebra. The other is based on the construction
of 5-dimensional matrix representation of the κ-Minkowski coordinate algebra. Neither approach leads to
a unique construction. Here, we find the relation between these two approaches and introduce a unified
approach, capable of describing all momentum spaces, and identify the corresponding quantum group of
spacetime symmetries. We reproduce known results and get a few new ones. In particular, we describe the
three momentum spaces associated to the κ-Poincaré group, which are half of a de Sitter, anti-de Sitter or
Minkowski space, and we identify what distinguishes them. Moreover, we find a new momentum space
with the geometry of a light cone, associated to a κ-deformation of the Carroll group.
© 2020 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/). Funded by SCOAP3 .

* Corresponding author.
E-mail addresses: fedele.lizzi@na.infn.it (F. Lizzi), mattia.manfredonia91@gmail.com (M. Manfredonia),
flavio.mercati@gmail.com (F. Mercati).

https://doi.org/10.1016/j.nuclphysb.2020.115117
0550-3213/© 2020 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/). Funded by SCOAP3 .
2 F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117

1. Introduction

The κ-Minkowski noncommutative spacetime [1–3] is defined by the commutation relations:


i  
[x 0 , x i ] = x i , xi , xj = 0 , i, j = 1, . . . , 3, (1)
κ
where κ is a constant with the dimensions of an energy (in h̄ = 1 units). The above relations close
a Lie algebra, known as an(3). The commutation relations (1) can be generalized to [x μ , x ν ] =
i(v μ x ν − v ν x μ ), μ = 0, . . . , 3, where v μ is any set of four real numbers. However, all these
algebras are isomorphic and can be put in the form (1) by a linear redefinition of generators.
As we will see in Sec. 3 below, this isomorphism does not mean that all algebras with different
choices of v μ are physically equivalent. In fact, the generator x 0 is usually interpreted as a time
coordinate, and x i as a spatial one. This interpretation derives from the fact that the above algebra
can be derived as the “quantum homogeneous space” of a quantum-group deformation of the
Poincaré group known as κ-Poincaré [1–9] (for a review of fuzzy spacetimes see [10], for a
different example of quantum homogeneous space see [11]). This group is generated by the
elements a μ and μ ν , satisfying the following commutation and cocommutation rules:
[μν ] = μα ⊗ α ν , [μν , α β ] = 0,
i  μ α 
[a μ ] = μν ⊗ a ν + a μ ⊗ 1, [μν , a γ ] =  α δ 0 − δ μ 0 γ ν
κ 
+ αν δ 0 α − δ 0 ν ημγ , (2)
i i
S[] = −1 , S[a μ ] = −a μ , [a 0 , a i ] = a , [a i , a j ] = 0,
κ
ε[μν ] = δ μ ν , ε[a μ ] = 0, μα ν β ηαβ = ημν , ρ μ σ ν ηρσ = ημν ,
where ημν = diag(1, −1, −1, −1) is the usual Minkowski flat metric. Note that the commutation
and cocommutation rules involving the Lorentz sector are undeformed, the deformation being
limited to the translation sector, and the mixed part. The very last relation gives the appropri-
ate number of constraints so that the independent components of  are six. The κ-Minkowski
commutation relations (1) are left invariant by the following left co-action of κ-Poincaré:
L [x μ ] = μν ⊗ x ν + a μ ⊗ 1, (3)
which is an algebra homomorphism for the relations (1). This is the sense in which κ-Minkowski
is the quantum homogeneous space associated to κ-Poincaré, and, in this light, it is legitimate to
interpret x 0 as temporal and x i as spatial coordinates, because they transform as such, and the
separation between time and space indices in the generators of the κ-Poincaré group is deter-
mined by the form of the metric ημν appearing in its relations. In Sec. 3 we will show that any
algebra [x μ , x ν ] = i(v μ x ν − v ν x μ ) is covariant under a generalization of the quantum group (2),
with any choice of the matrix ημν . A linear redefinition of generators x μ , done to show that
all algebras with different values of v μ are isomorphic, induces a linear transformation of the
matrix ημν . Hence, to specify a model of quantum spacetime, the commutation relations of the
coordinates x μ (i.e. the choice of parameters v μ ) are not sufficient: one needs also to specify
the form of the matrix ημν , which can attribute timelike or lightlike nature to different linear
combinations of generators x μ . Two models, specified by (v μ , ημν ) and (v  μ , ημν  ) will then be

physically equivalent if there exists a linear transformation of the generators that sends v μ to v  μ
and, simultaneously, ημν to ημν  .
F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117 3

One unusual property of the κ-Minkowski spacetime is that it is associated to a curved mo-
mentum space, a momentum space that generalizes the vector space of Special/General Relativity
into a pseudo-Riemannian geometry. The idea of a curved momentum space has a long history.
It was formulated more than 20 years ago by S. Majid (see, e.g., [6,12,13]), and it can be mathe-
matically understood as dual in the sense of Hopf-algebraic duality of algebraic and coalgebraic
sectors describing quantum coordinates and quantum momenta [14]; the noncommutativity of
translation generators is a manifestation of the curvature of spacetime. Similarly, in a quantum
geometry, the noncommutativity of spacetime coordinates is a manifestation of the curvature of
momentum space, a phenomenon that Majid called ‘cogravity’.
The geometrical structure of the κ-Minkowski momentum space was first studied in [15], and
then further studied in a variety of works, including [15–21]. The simplest way to see this is to
consider the ordered plane waves built from the noncommutative coordinates (1):
 
exp ikμ x μ , k μ ∈ R4 , (4)
these are useful because they provide a basis in which we can expand functions, in order to
discuss field theories on κ-Minkowski (1) [22–29]. What is unusual is the fact that, because the
product (1) is noncommutative, these plane waves do not combine in a linear way:
   
exp ikμ x μ exp iqμ x μ =
 
(k0 + q0 )/κ ek0 /κ − 1 −k0 /κ e
q0 /κ − 1
exp i (k +q )/κ ki + e qi x + i(k0 + q0 )x ).
i 0
e 0 0 −1 k0 /κ q0 /κ
(5)
This fact, known since the 1990s [30], can be proven explicitly using only the commutation
relations [25,27]. Since the commutation rules are those of a Lie algebra, the exponentials are
closed under product, they form a subalgebra of the universal enveloping algebra of an(3). The
law:

(k0 + q0 )/κ ek0 /κ − 1 −k0 /κ e
q0 /κ − 1
(k, q) → p , pi = (k +q )/κ ki + e qi ,
e 0 0 −1 k0 /κ q0 /κ
p0 = k0 + q0 , (6)
generalizes in a nonlinear way the familiar composition law of “wave vectors” (or Fourier pa-
rameters) (k, q) → kμ + qμ , and reduces to it in the limit κ → ∞. It can be seen as a small
deformation of it, when the wave vectors are much smaller than κ [31,32]. There is a consensus
in the literature on the fact that this nonlinearity is a manifestation of the fact that the Fourier
parameters are coordinates on a nonlinear manifold. In fact, exponentiating the generators of a
Lie algebra like an(3), one obtains elements of the associated Lie group, which in our case is
AN (3) [19]. Then, since the algebra is not Abelian, the composition law between the parameters
in the exponentials is not linear, and they just codify the group product. As the theory of Lie
groups prescribes, these parameters are coordinate systems on the group manifold.
The expression of plane waves in (4) is not the only possible choice to represent a plane
wave, we had implicitly chosen an ordering prescription. There are other ordering choices, which
give rise to different factorizations of the group elements. For example, the time generator can
be ordered to the right, exp(iqi x i ) exp(iq0 x 0 ). Different orderings are related through nonlinear
relations between the real parameters appearing in the exponentials. For the two examples above:
  
  ek0 /κ − 1
exp ikμ x μ = exp i ki x i exp ik0 x 0 , (7)
k0 /κ
4 F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117

 k /κ 
this transformation, k0 → k0 , ki → e k0 /κ−1 ki is a general coordinate change, i.e., a diffeomor-
0

phism on the group manifold.


We interpret the group manifold associated to the Lie group AN (3) as the momentum space
of theories on κ-Minkowski that make use of noncommutative plane waves, e.g. (quantum) field
theories, in which ordered plane waves are a basis for scalar fields and solutions of the equations
of motion. Here we are interested in the geometry of this momentum space. In Lie group theory,
there is a natural way to define a metric on the group manifold: if there is a nondegenerate Killing
form, one can immediately define a bi-invariant metric. However, since the group AN (3) is not
semi-simple, the Killing form is degenerate and there is no bi-invariant metric. There is, however,
a basis of left-invariant forms and another one of right-invariant forms:

ki
θ0L = dk0 , θiL = dki + dk0 ,
κ (8)
θ0R = dk0 , θiR = ek0 /κ dki ,

which are related by a diffeomorphism implementing the group inverse (codified by the antipode
map at the Hopf algebra level): k0 → −k0 , ki → −ek0 /κ ki , which sends θμL → −θμR and vice-
versa. As observed in [33], any quadratic form built from the symmetrized tensor product of
right-invariant forms will give a right-invariant metric, and the same for left-invariant metrics.
In [33] it is stated that there is a unique right-invariant metric and a unique left-invariant one, but
the author is implicitly assuming that the signature is (+, −, −, −), and the rank is maximal (no
zero eigenvalues). Moreover, even under these conditions, there is a hidden assumption in [33]’s
proof that all metrics are diffeomorphic. In the following Section we will study in detail all the
possible right-invariant metrics, and classify them according to diffeomorphism classes.
In the literature on κ-Minkowski, the most common way used to introduce the curved mo-
mentum space is that used in [15], which was the first paper to observe that the momentum space
of κ-Minkowski may be described as a curved Riemannian manifold. [15] used a matrix repre-
sentation of an(3) (see also [34–36]). The algebra (1), in fact, can be seen as a subalgebra of the
five-dimensional Lorentz algebra so(4, 1) via the isomorphism:

x μ ∼ M 0μ + M 4μ , (9)

where M AB (A, B = 0, . . . , 4) are the Lorentz generators in the standard antisymmetric 5 × 5


matrix representation. This isomorphism induces the following five-dimensional representation
of the commutation relations (1):
⎛ ⎞ ⎛ ⎞
0 0 1 0 ei 0
i i
ρ(x 0 ) = − ⎝ 0 0̂ 0 ⎠ , ρ(x i ) = − ⎝ ei 0̂ ei ⎠ , (10)
κ κ
1 0 0 0 −ei 0

where eai = δia , three-dimensional vector quantities are in boldface (we do not distinguish be-
tween rows and columns, it should be clear form the position in the matrix), 0̂ is the zero
3 × 3 matrix. This is a ∗-representation under the involution compatible with the Lorentz group
(ρ α β )∗ = ηαλ ηγβ ρ γ λ (i.e. rising an index, flipping indices, complex conjugating and lowering
back the index), which leaves all generators ρ(x μ ) invariant. In this way, the plane waves/group
elements are represented as the matrices (in order to get simpler formulas we use the “time-to-
the-right” ordering):
F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117 5

⎛ p0
p 2 p p0
p 2

cosh pκ0 + e κ
2κ 2 κ sinh pκ0 + e κ
2κ 2
⎜ ⎟
⎜ ⎟
G∗ (pμ ) = eipi ρ(x ) eip0 ρ(x ) = ⎜
i 0 p0 p0
e κ
p
1 e κ
p
⎟. (11)
⎝ κ κ ⎠
p0 p0
p 2 p 2
sinh pκ0 − e κ
2κ 2
− κp cosh pκ0 − e κ
2κ 2
The idea is that, since the above representation is free, and the nondegenerate orbits have the
dimension of the group, these will be diffeomorphic to the group manifold.1 The group orbits
can be obtained by taking a fiducial vector uA in the five-dimensional vector space on which (11)
acts, and considering the points obtained by acting upon u with G∗ (pμ ) for all choices of p μ :

X A = X A (pμ ) = G∗ (pμ )A B uB . (12)


X A (pμ ) are the parametric representation of a four-dimensional submanifold embedded in a
five-dimensional Minkowski space. This submanifold is diffeomorphic to the group manifold of
AN (3), and to momentum space. Since all G∗ (pμ ) are elements of SO(4, 1), we have that

X A XA = X A (p)X B (p)ηAB = uA uB ηAB , ηAB = diag(1, −1, −1, −1, −1), (13)
for all p μ ∈ R4 . Choosing uA = (0, 0, 0, 0, 1), the above equation is that of de Sitter spacetime.
The conclusion in [15] or [34] is that the geometry of momentum space is de Sitter. And indeed
the metric on the orbit induced by the embedding X A(p):
 
∂X A ∂X B 1 
3
ds = −
2
ηAB dpμ dpν = 2 −dp0 + e 2 2p0 /κ 2
dpi , (14)
∂pμ ∂pν κ
i=1

is the same right-invariant metric found by [33]. Moreover, one can check that for uA =
(0, 0, 0, 0, 1) the relation X 0 + X 4 > 0 is verified for all choices of pμ , and therefore we are
actually dealing with half of de Sitter spacetime, the half covered by the flat slicing (the coordi-
nates pμ corresponding to time-to-the-right ordering of plane waves are what cosmologists call
comoving coordinates for de Sitter spacetime). This constraint makes the portion of momentum
space covered by the pμ coordinates non-Lorentz-invariant [17], and one has to choose a differ-
ent global topology for the ambient space (the elliptic topology [37]) in order to restore Lorentz
invariance [29,38].
The construction we just described, however, is not unique. For example, in [20] it was noticed
that a different fiducial vector [in particular a time-like one uA = (1, 0, 0, 0, 0)] gives rise to
a different momentum space. Since the Lorentz group has disconnected orbits, corresponding
to different fiducial vectors, their choice is not inconsequential. Moreover, as it turns out, the
representation (9) is not the only possible for an(3). In [39], the authors notice that x μ can be
represented as the four generators M̂ 4μ (times κ) of a five-dimensional orthogonal algebra
 
[M̂ AB , M̂ CD ] = i ĝ BC M̂ AD − ĝ AC M̂ BD − ĝ BD M̂ AC + ĝ AD M̂ BC , (15)

which preserves a 5D metric of the form ĝ 4A = ĝ A4 = δ A 0 , and ĝ μν is assumed to have eigenval-


ues (−1, 1, 1, 1). Then, according to the form of ĝ μν , one has different situations. The 5D matrix
can only have the following signatures: (−, +, +, +, +), in which case M̂ AB close the so(4, 1)

1 For example, exponentiating the standard representation of su(2) as 2 × 2 complex matrices acting on the vector
space of 2D spinors C 2 , one can prove that the nondegenerate orbits of the group are all 3-spheres embedded in R4
(under the canonical identification R4 ∼ C 2 ), and indeed the group manifold of su(2) is, topologically, a 3-sphere.
6 F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117

algebra, or (−, −, +, +, +), corresponding to the so(3, 2) algebra, or, finally, (−, +, +, +, 0),
which implies that M̂ AB generate the Poincaré algebra iso(3, 1). The form of ĝ AB is such that
the transformations generated by M̂ μν (which close a 4D Lorentz algebra) leave the direction
(0, 0, 0, 0, 1) in the ambient 5D space invariant, while those generated by M̂ 4μ change it. There-
fore, in [39], the momentum space is identified with the quotient of the 5D group with SO(3, 1),
which, according to the choice of matrix ĝ μν will be de Sitter [SO(4, 1)/SO(3, 1)], anti-de Sitter
de Sitter [SO(3, 2)/SO(3, 1)] or Minkowski [I SO(3, 1)/SO(3, 1)].
However, this method allows us only to know the local isometries of momentum space, and
does not reveal its global shape. For example, in the de Sitter case, it is necessary to explicitly
study the submanifold of R5 that is traced by the action of the generators x μ , as was illustrated
in Eqs. ((9)-(14)), in order to realize that it is only half of the de Sitter hyperboloid that can be
identified with momentum space.
The purpose of the present paper is to study all the momentum spaces that can be associated
to κ-Minkowski, and to find the relation between the different partial analyses of the past, e.g.
[15,20,33,39]. A synopsis of the main result is:

• In Sec. 2 we study all the possible right-invariant metrics on the AN (3) group, and their
equivalence classes under diffeomorphisms, which has not been done before.
• In Sec. 3 we clarify the sense in which different κ-Minkowski algebras (with different
choices of parameters v μ in the commutation relations) can be said to be physically equiva-
lent or inequivalent.
• In Sec. 4 we study all nondegenerate geometries of the κ-Momentum spaces, which can be
obtained with different choices of v μ parameters and spacetime metric. Many of these were
already known in the literature (de Sitter, anti-de Sitter, Riemannian hyperbolic). We find
two new ones:
– a hyperbolic space with signature (+, +, −, −), associated to a κ-Minkowski spacetime
symmetric under a κ-deformation of the I SO(2, 2) group,
– a cone, associated to a κ-Minkowski spacetime symmetric under a κ-deformation of the
Carroll group (or of a Lorentzian version of the Carroll group),
– half of Minkowski space.
Furthermore, we complete the classification of the momentum spaces associated to κ-
Minkowski spacetimes symmetric under the κ-Poincaré group: these can be (half of) de
Sitter, (half of) Minkowski or (half of) anti-de Sitter, depending on which coordinates the
spacetime metric g μν makes timelike: in the first case x 0 , in the second x 0 ± x i for some i,
and in the third x i for some i.

2. Right-invariant metrics on AN(3)

There is a certain freedom in choosing these metrics, and we need to classify them and find
the physical meaning of inequivalent classes.
Let us analyze the most generic right-invariant metric similarly to what was done in [33], but
without any initial assumption. One begins with the generic right-invariant line element:
ds 2 = −g μν θμR θνR , (16)
where is a generic symmetric 4 × 4 matrix. Calculating the Riemann and Ricci curvatures,
g μν
one can verify that, whatever g μν we choose, the curvatures satisfy the relation of maximally-
symmetric spaces:
F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117 7

1  
Rμνρσ = R gμρ gνσ − gμσ gνρ , (17)
12
and the scalar curvature is
12 det(g ij )
R= . (18)
det(g μν )
This result allows us to connect the approach based on right-invariant metrics to that of [39]
(based on a family of 5D representations of an(3)): both approaches agree that the local geometry
of the momentum spaces of κ-Minkowski is that of a maximally-symmetric space. Clearly, if the
momentum-space metric is assumed to be non-singular, the only possibilities are de Sitter, anti-de
Sitter and Minkowski space.
Let us study in more detail what classes of inequivalent metrics (under coordinate changes)
we have, without assuming anything on the signature or singularity of the 4D metric. We begin
by decomposing the indices in Eq. (16) in 0 and i = (1, 2, 3):
ds 2 = −g 00 dk02 − 2ek0 /κ g 0i dk0 dki − e2k0 /κ g ij dki dkj , (19)
we can introduce logarithmic coordinates as k0 = −κ log τ (k0 ∈ R so τ ∈ R+ ):
1  
ds 2 = − 2 g 00 dτ 2 − 2g 0i dτ dki + g ij dki dkj , (20)
τ
if g 0i = 0, the metric is already block-diagonal, and a simple linear transformation of the ki , i =
1, 2, 3 coordinates diagonalizes g ij and puts the line element in the form (25). If any of the three
components g 0i are nonzero, we can perform the following linear transformation: ki = ki + ci τ ,
and the line element takes the form
1  
ds 2 = − 2 (g 00 − 2g 0i ci + g ij ci cj )dτ 2 + 2(ci g ij − g 0j )dkj dτ + g ij dki dkj , (21)
τ
and the line element can be block-diagonalized if the following equations can be solved for ci :
g ij cj = g 0i . (22)
The above equation admits solutions only if the range of g ij
is equal to the range of the rectan-
gular matrix g μj , which means that the 3-vector g 0i lies within the range of g ij . This is always
true if g ij has maximum rank (i.e. it is invertible), but if g ij has some zero eigenvalues it might
not hold. Assuming that g 0i lies within the range of g ij , we can replace the solution of (22) into
the line element, using the fact that g 0i cj = g ij ci cj :
1  
ds 2 = − 2 (g 00 − g ij ci cj )dτ 2 + g ij dki dkj , (23)
τ
and, finally, we can perform a linear transformation of the ki coordinates that diagonalizes the
matrix g ij :
1    
ds 2 = − 2 λ0 dτ 2 + λ1 (dk1 )2 + λ2 (dk2 )2 + λ3 (dk3 )2 , (24)
τ
where λ0 = g 00 − λ1 (c1 )2 − λ2 (c2 )2 − λ3 (c3 )2 , and ci is the transformed version of ci . Going
back to the k0 variable:
 
ds 2 = −λ0 dk02 − e2k0 /κ λ1 (dk1 )2 + λ2 (dk2 )2 + λ3 (dk3 )2 , (25)

where the coefficients λμ are unconstrained real numbers.


8 F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117

Assume now that Eq. (22) is not solvable: then g ij has some zero eigenvalues, and we know
that the geometry is flat, because the Ricci scalar is proportional to det(g ij )/ det(g μν ) [Eq. (18)],
and sending one, two or three of the eigenvalues of g ij simultaneously to zero makes R vanish
(no matter what the eigenvalues of g μν do). Moreover, the fact that Eq. (22) is not solvable
implies that g 0i has nonzero components in the zero eigenspace of g ij . We need to consider two
distinct cases:

• If g 00 = 0, then the metric (20) takes the form

1  0i 
ds 2 = − 2g dτ dk i + g ij
dk i dk j , (26)
τ2
and transforming the spatial indices in order to diagonalize the spatial part:

1  i 

 2  2  2

ds 2 = − 2v dτ dk i + λ 1
(dk 1 ) + λ 2
(dk 2 ) + λ 3
(dk 3 ) , (27)
τ2
where some of the diagonal elements λi are zero (and the 3-vector v i = g  0i has at least one
nonzero component in the direction corresponding to a zero eigenvalue). The determinant of
this metric is:
 
−τ −8 λ2 λ3 (v 1 )2 + λ1 λ2 (v 3 )2 + λ1 λ3 (v 2 )2 , (28)

which is zero unless only one eigenvalue is zero. If, for example, λ1 = 0 (and consequently
v 1 = 0), and λ2 , λ3 = 0, then the determinant is −τ −8 v 1 λ2 λ3 = 0, and the 4D metric has
four nonzero eigenvalues.
If there are two zero λ’s, for example λ1 and λ2 , then the characteristic polynomial of the 4D
metric reduces to q 4 − λ3 q 3 − v 2 q 2 + λ3 [(v 1 )2 + (v 2 )2 ]q, and since v 1 , v 2 , λ3 = 0, there
are three nonzero eigenvalues.
Finally, if all three λi are zero, the 4D metric has two zero eigenvalues, and the remaining
two are ± v .
• If g 00 = 0, we can make the coordinate transformation τ = θ + g 0i /g 00 ki ,

1  
ds 2 = − g 00 dθ 2 + (g 0i g 0j /g 00 + g ij )dki dkj . (29)
(θ + g 0i k i /g 00 )2
 
The matrix g 0i g 0j /g 00 + g ij has determinant

λ2 λ3 (v 1 )2 + λ1 λ2 (v 2 )2 + λ1 λ3 (v 3 )2 − λ1 λ2 λ3 (30)

(where λi are the eigenvalues of g ij ). This, again, can be nonzero if only one of the λi is
zero. In the other two cases, just like before, one has one or two zero eigenvalues.

We conclude that, when Eq. (22) is not solvable, one has a flat 4D metric of arbitrary signature
with two, three or four nonzero eigenvalues.
Regarding the left-invariant metric, by what we said before they are all diffeomorphic to the
right-invariant ones, so they just correspond to the same geometries described here, written in
different coordinates.
F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117 9

3. Embeddings in 5D-metric-preserving groups

M̂ AB closing the algebra (15), which preserves a 5D metric of the form:


⎛ 00 ⎞
g ... g 03 1
⎜ .. .. .. ⎟
⎜ . . . 0 ⎟
⎜ ⎟
ĝ AB = ⎜ 0 ⎟, (31)
⎜ ⎟
⎝ g 03 . . . g 33 0⎠
1 0 0 0 0
introducing the representation ρ(x μ ) = κ1 M 4μ one can immediately verify that the com-
mutation relations (15) imply the κ-Minkowski commutation relations, [ρ(x μ ), ρ(x ν )] =
κ (δ 0 ρ(x ) − δ 0 ρ(x )), thanks to the particular form of ĝ
i μ ν ν μ AB . Similarly to what we illus-

trated in the introduction for the representation (10), we can now calculate the generic group
elements generated by ρ(x μ ), which are a four-parameter family of 5D matrices G∗ (pμ ) =
i 0
eipi ρ(x ) eip0 ρ(x ) .
Choosing a fiducial vector uA in the ambient space, we then derive its orbits as X A =
X (pμ ) = G∗ (pμ )A B uB . Now, assuming that ĝ AB is invertible and calling its inverse ĝAB , we
A

can find the metric that is induced on the orbit by the embedding XA(pμ ) as
∂X A ∂X B
ds 2 = −ĝAB dpμ dpν . (32)
∂pμ ∂pν
From the definition of right-invariant forms, we know that
(G∗ −1 )A B dG∗ B C = θμR ρ(x μ )A C , (33)
which implies that
∂X A
dpμ = dG∗ A B uB = θμR (G∗ −1 )A C ρ(x μ )C B uB . (34)
∂pμ
Now, assuming that ĝ AB is invertible, by definition its inverse is left invariant by the group
elements:
ĝAB (G∗ −1 )A C (G∗ −1 )B E = ĝCE , (35)
and so we can write
∂X A ∂X B
ds 2 = −ĝAB dpμ dpν
∂pμ ∂pν
(36)
= −θμR θνR ĝAB (G∗ −1 )A C (G∗ −1 )B E ρ(x μ )C D ρ(x ν )E F uD uF
= −θμR θνR ĝCE ρ(x μ )C D ρ(x ν )E F uD uF ,
and an explicit calculation reveals that
ds 2 = u4 u0 θ0R θ0R + u4 uμ θ0R θμR − (u4 )2 g μν θμR θνR . (37)
uA cannot have a zero 4 component, otherwise it would be left invariant by G∗ . So, any choice
of uA gives a line element ds 2 of the general form (16), which we studied in detail in the pre-
vious section. The particular choice uA = δ A 4 makes the matrix g μν used here and the one of
Equation (16) identical, otherwise, for other choices of fiducial vector, they are linearly related.
10 F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117

The matrix (31) is such that the submatrix g μν can be diagonalized with a linear redefinition
of the first four indices, and the result is a matrix of the form:
⎛ 0 ⎞
λ 0 0 0 v0
⎜ 0 λ1 0 0 v 1 ⎟
⎜ ⎟
ĝ = ⎜
AB
⎜ 0 0 λ
2 0 v2 ⎟ ⎟, (38)
⎝ 0 0 0 λ3 v 3 ⎠
v0 v1 v2 v3 0
where λμ are the eigenvalues of g μν , and v μ is the vector obtained by applying the similarity
that diagonalizes g μν to the vector (1, 0, 0, 0). However, now the transformed generators M̂  4B
do not provide a representation of the original “timpelike” κ-Minkowski algebra (1), but rather
of an algebra isomorphic to it, in which the role of the special x 0 coordinate is played by a linear
combination of the x μ coordinates. The commutation relations are those of the generalized κ-
Minkowski algebra:
 
[x μ , x ν ] = i v μ x ν − v ν x μ . (39)
We therefore proved that the representation (31) of κ-Minkowski, when the matrix g μν is not
chosen diagonal, is equivalent to a representation of the generalized κ-Minkowski algebra (39)
with diagonal invariant matrix. This clarifies the meaning of the four parameters v μ that ap-
pear in (39). In fact, the algebra (39) is isomorphic to the standard κ-Minkowski algebra (1): it
2
is sufficient to make the linear redefinition x i → v 0 x i − v i x 0 , x 0 → vi x i + 1−v 0v x 0 to show
it. This does not, however, mean that all the noncommutative spaces corresponding to different
choices of v μ are completely equivalent. One has, in fact, to consider the whole group of sym-
metry that leaves the relations (39) invariant under an inhomogeneous transformation of the form
x  μ = μ ν x ν + a μ , where μ ν and a μ commute with x μ . Any algebra of the following form:
 
[a μ , a ν ] = i v μ a ν − v ν a μ ,
[μ ν , ρ σ ] = 0 , (40)
    
[μ ν , a γ ] = i μα v α − v μ γ ν + α ν g̃αβ − g̃νβ v β g μγ ,
(where we make no assumptions regarding the matrices g and g̃ except that they are symmetric)
will leave the commutation relations (39) invariant. However, the Jacobi rules are not necessarily
satisfied by these relations. In fact:
[[μ ν , a ρ ]a σ ] + [a ρ , a σ ]μ ν ] + [[a σ , μ ν ], a ρ ]
  (41)
= g μ[σ v ρ] δ δ ν + g δ[ρ v σ ] μ ν + v μ g δ[σ ρ] ν + v δ g μ[σ ρ] ν g̃λδ v λ ,

where v μ = v μ − g μν g̃νρ v ρ and A[μν] = Aμν − Aνμ . A violation of the Jacobi rules would
indicate a failure of associativity, which we would not know how to handle. We must therefore
require (41) to vanish in all of our systems. An important case is when g μν is invertible and g̃μν
is its inverse. Then the above expression is identically zero, because all the terms v μ = 0, for
any choice of g μν .
We then understand better the meaning of our freedom to choose the parameters v μ : they
select one linear combination of the noncommutative coordinates that plays a special role in the
algebra, and we cannot simply re-label this direction “x 0 ” and reduce to the “timelike” case (1),
because this linear transformation would change the matrix g μν that is left invariant by the isome-
tries of our noncommutative spacetime. For example, depending on the initial choice of g μν , the
F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117 11

redefinition that puts the algebra in the form (1) might end up making x  0 into a spacelike co-
ordinate. The simultaneous choice of v μ and g μν determines a noncommutative geometry with
a particular set of symmetries, and these fall into families that can be classified by looking at
representations of the form (38), with a diagonal g μν and a generic v μ .2 Each choice of the four
eigenvalues λμ gives a 4-parameter family (dependent on v μ ) of inequivalent models. We can
then list these inequivalent models and describe the geometry of the corresponding momentum
spaces.

4. Nondegenerate geometries of momentum space

We are ready to list the possible nondegenerate geometries of momentum spaces one can
build for the κ-Minkowski noncommutative spacetime. There are also many degenerate cases, in
which the momentum space manifold is not topologically 4-dimensional, but we prefer to focus
on the nondegenerate ones, which have a chance to be useful in model building. Although all the
cases we list have topological dimension four, their geometry might be degenerate, in the sense
that the induced metric might have one or more zero eigenvalues.
Let us summarize, for clarity, the algorithm we use to write different representations of the
κ-Minkowksi coordinate algebra. Given an ambient metric of the form (38), we can write a
matrix representation of the 10-dimensional Lie algebra that preserves this matrix, in the sense
of satisfying commutation relations of the form (15), as
 
(M̂ AB )a b = i ĝ Aa δ B b − ĝ Ba δ A b , (42)

and then the components ρ(x μ ) = M̂ 4μ close a generalized κ-Minkowski algebra of the
form (39). Explicitly,
⎛ 0 ⎞ ⎛ ⎞
v 0 0 0 −λ0 0 v0 0 0 0
⎜ v1 0 0 0 0 ⎟ ⎜ 0 v 1 0 0 −λ1 ⎟
⎜ 2 ⎟ ⎜ ⎟

ρ(x ) = i ⎜ v 0 0 0
0
0 ⎟,⎟ ρ(x ) = i ⎜
1 2 0 ⎟
⎜0 v 0 0 ⎟,
⎝ v3 0 0 0 0 ⎠ ⎝ 0 v3 0 0 0 ⎠
0 0 0 0 −v 0 0 0 0 0 −v 1
⎛ ⎞ ⎛ ⎞ (43)
0 0 v0 0 0 0 0 0 v0 0
⎜ 0 0 v1 0 0 ⎟ ⎜ 0 0 0 v1 0 ⎟
⎜ ⎟ ⎜ ⎟
ρ(x ) = i ⎜
0
⎜ 0 0 v 2 0 −λ2 ⎟ ,
⎟ ρ(x 1
) = i ⎜ 0 0 0 v2
⎜ 0 ⎟ ⎟.
⎝ 0 0 v3 0 0 ⎠ ⎝ 0 0 0 v −λ ⎠
3 3

0 0 0 0 −v 2 0 0 0 0 −v 3
This is the most generic representation of the generalized κ-Minkowski commutation relations,
and it is easy to verify that it reduces to the previously-known cases for specific choices of λμ and
v ν (possibly after a permutation of the columns/rows). The values of λμ determine the geometry
of the associated momentum space.

2 Alternatively, one could say that a model is specified by a choice of parameters v μ and of matrix η
μν in Eqs. (40), but
there are equivalence classes of physically equivalent models. For instance, if two models are characterized by different
(v μ , gμν ) and (v  μ , gμν
 ), but there exists a linear redefinition of the generators x μ that transforms v μ into v  μ , and, at
the same time, g μν ) into g  μν , then the two models are the same, we are just labeling its coordinates in two different
ways.
12 F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117

The orbits Y A (pμ ) = G∗ (pμ )A B uB of the elements G∗ (pμ ) = exp[ipi ρ(x i )] exp[ip0 ρ(x 0 )]
are our momentum spaces, and, as we showed above, it doesn’t matter which point uA in the
ambient R5 we start from, we can always find a choice of λμ parameters for which the orbit
includes the point uA = (0, 0, 0, 0, 1), so we might as well always use this as our starting point.
By construction, the embedding coordinates Y A satisfy the quadratic equation

λ0 λ1 λ2 λ3
ĝAB Y A Y B = ĝAB uA uB = = const. , (44)
det ĝ AB
and, moreover, the fifth coordinate satisfies
μ
Y 4 = epμ v > 0 , (45)
which is the same as relation X 0 + X 4 > 0 from above, written in different coordinates. We can
now list the inequivalent cases one gets with different choices of λμ , and describe the geometry
of the corresponding momentum spaces.

4.1. Half of de Sitter momentum space

This case has already been sketched in the introduction, let us work it out again in our new
setup. With the following choice of ambient metric:
⎛ ⎞
−1 0 0 0 v 0
⎜ 0 1 0 0 v1 ⎟
⎜ ⎟

ĝ = ⎜ 0
AB
0 1 0 v2 ⎟ (46)

⎝ 0 0 0 1 v3 ⎠
v0 v1 v2 v3 0

the generators M̂ 4μ of (15) close the generalized κ-Minkowski algebra, and their exponentializa-
tion G∗ (p)A B acts on the fiducial vector uA = δ A 4 in such a way that the embedded hypersurface
it defines, Y A (p μ ) = G∗ (p)A B uB , is the Y 4 > 0 half of a de Sitter hyperboloid. The induced line
element on the embedded hypersurface is

ds 2 = −dp02 + e2p0 /κ (dp12 + dp22 + dp32 ) , (47)


which is the de Sitter metric in flat slicing, well-known to be a half-cover of the hyperboloid [40].
In Fig. 1 is a graphic representation of the embedded manifold in the 1+1-dimensional case.
The homogeneous isometries of this momentum space are given by the Lorentz group
SO(3, 1), to which the matrices μ ν of Eqs. (40) belong to in this case, because they leave
the Minkowski metric and its inverse invariant, i.e. μ ρ ν σ ημν = ηρσ and μ ρ ν σ ηρσ = ημν .
Therefore, in this case, we are dealing with a quantum group of isometries of the noncommu-
tative spacetime (39), of the form (40) with g̃μν = diag(−1, 1, 1, 1) = g̃ μν . This is a quantum
deformation of the Poincaré group I SO(3, 1).

4.2. Half of anti-de Sitter space

By placing the minus in one of the diagonal components corresponding to the 1, 2 or 3 axes,
for example:
F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117 13

Fig. 1. Half of de Sitter space covered by the p μ coordinates in 1+1 dimensions (in red). (For interpretation of the colors
in the figure(s), the reader is referred to the web version of this article.)

Fig. 2. Half of anti-de Sitter space covered by the p μ coordinates in 1+1 dimensions (in blue).
⎛ ⎞
1 0 0 0 v0
⎜ 0 −1 0 0 v1 ⎟
⎜ ⎟
ĝ AB =⎜
⎜ 0 0 1 0 v2 ⎟
⎟ (48)
⎝ 0 0 0 1 v3 ⎠
v0 v1 v2 v3 0
we get another Lorentzian momentum space manifold, whose induced metric has negative con-
stant curvature:
ds 2 = dp02 + e2k0 /κ (−dp12 + dp22 + dp32 ) , (49)
(the minus could be in front of any of the dp12 ,
or dp22 , dp32
terms). Again, the sum is positive Y4
for all values of p μ , so we are dealing with half a hyperboloid. (See Fig. 2.) Just like in the case
above, the region covered by these coordinates is not Lorentz-invariant, and one has to assume
a nontrivial global topology (like the elliptic topology of de Sitter) in order to recover Lorentz
invariance [29,38].
The local homogenous isometry group is again SO(3, 1). The quantum group of symmetries
is then a deformation of I SO(3, 1), but with the timelike direction being along the axis 1 (or 2,
14 F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117

Fig. 3. Riemannian hyperbolic momentum space in 1 + 1 dimensions. The p μ coordinates cover only the red sheet.

or 3, according to the choice of ĝ AB ). In other words, this model is identical to the one of the
previous point, except that the Lorentz matrices satisfy μ ρ ν σ g̃μν = g̃ρσ and μ ρ ν σ g̃ ρσ =
g̃ μν , where g̃μν = diag(1, −1, 1, 1) = g̃ μν , i.e., it has a negative 11, 22 or 33 component, as
opposed to the 00 component.

4.3. Riemannian hyperbolic momentum space

With a Euclidean signature for the 4D submatrix:


⎛ ⎞
1 0 0 0 v0
⎜ 0 1 0 0 v1 ⎟
⎜ ⎟
ĝ AB =⎜
⎜ 0 0 1 0 v2 ⎟
⎟ (50)
⎝ 0 0 0 1 v3 ⎠
v0 v1 v2 v3 0

we get a Euclidean-signature hyperboloid for momentum space. The condition Y 4 > 0 still holds,
but that includes the entirety of one of the two disconnected sheets of the hyperboloid. (See
Fig. 3.) The attractive feature of this momentum space is that the condition Y 4 > 0 does not
identify a region with boundary, and, what is more, the region it identifies is invariant under
‘Lorentz’ transformations, unlike what happens in the two cases listed above. We have a mo-
mentum space that is closed under 4-dimensional rotations, which might represent an appealing
simplicity feature for the study of Euclidean QFT.
The induced metric on this hyperboloid is:

ds 2 = dp02 + e2k0 /κ (dp12 + dp22 + dp32 ) . (51)

The homogeneous isometries are SO(4) rotations, and the quantum group associated to this
momentum space is (40) with g̃μν = diag(1, 1, 1, 1) = g̃ μν , which is a deformation of I SO(4),
the group of 4-dimensional Euclidean transformations.
F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117 15

4.4. Two-time hyperbolic space

With a 4D sub-matrix with two positive and two negative eigenvalues:


⎛ ⎞
−1 0 0 0 v0
⎜ 0 −1 0 0 v 1 ⎟
⎜ ⎟
ĝ = ⎜
AB
⎜ 0 0 1 0 v2 ⎟ ⎟ (52)
⎝ 0 0 0 1 v3 ⎠
v0 v1 v2 v3 0
we get a manifold with two temporal and two spatial directions, with metric:
ds 2 = −dp02 + e2p0 /κ (−dp12 + dp22 + dp32 ) , (53)
(again, we have three distinct cases, related by exchanges of the 1, 2, 3 axes). Whatever the values
of v μ , we can always rotate our axes Y i → Y  i so that this manifold corresponds to region Y 4 > 0
of the quadric
(Y 4 )2 + 2Y 0 Y 4 − (Y  1 )2 + (Y  2 )2 + (Y  3 )2 = 1 . (54)
This region has a finite boundary at, = 0 in correspondence of the submanifold
Y4 + (Y  2 )2
(Y  3 )2 = 1 + (Y  1 )2 .
It does not make sense to try and represent this momentum space in 1 + 1 dimensions as a
submanifold of R3 , so we won’t show, in this case, a pictorial representation.
The homogeneous isometries of this manifold close an SO(2, 2) group. Therefore this mo-
mentum space is associated to a noncommutative spacetime which is symmetric under a quantum
deformation of I SO(2, 2) (i.e. (40) with g̃μν = diag(−1, −1, 1, 1) = g̃ μν ).

4.5. Light cone momentum spaces

We get an interesting momentum space if we set λ0 = 0. The other λμ ’s can have either
Lorentzian or Euclidean signature:
⎛ ⎞
0 0 0 0 v0
⎜ 0 ±1 0 0 v 1 ⎟
⎜ ⎟
ĝ = ⎜
AB
⎜ 0 0 1 0 v2 ⎟ ⎟, (55)
⎝ 0 0 0 1 v3 ⎠
v0 v1 v2 v3 0
(any of the λ1 , λ2 or λ3 could be negative, or two of them could be negative and one positive,
without changing the result). In fact, the ĝAB -norm of uA is, in this case, zero, and we get a light
cone in the embedding space. Suppressing the 3 direction, we can see it represented in Fig. 4: the
condition Y 4 > 0 selects only the “future” half of the cone, and it further cuts a line off of it (see
Fig. 4). Notice how this makes it topologically equivalent to all the other momentum spaces we
encountered so far (although the metric of the AN(3) group manifold is not unique, the topology
is uniquely determined by its Lie group structure).
The induced metric is degenerate:
ds 2 = e2p0 /κ (±dp12 + dp22 + dp32 ) , (56)
and its homogeneous isometries are a contraction of the Lorentz group SO(3, 1) (or SO(2, 2)
in the case with the minus sign) into I SO(3) (or I SO(2, 1)). This is the contraction in which
16 F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117

Fig. 4. Light-cone momentum space. The p μ coordinates cover only the red sheet, with the exclusion of the white line.

the speed of light is sent to zero, and the corresponding inhomogeneous group is the Carroll
group [41,42] Carr(3, 1) (or, in the case with the minus sign, a version of the Carroll group in
which the spatial directions are Lorentzian). This quantum group is still of the form (40), but
with g̃ μν = diag(0, 1, 1, 1) and g̃μν = diag(1, 0, 0, 0) (this is a possible way to define the Carroll
group, as the group of inhomogeneous transformations x μ → μ ν x ν + a μ where the matrices
μ ν preserve a degenerate metric with one zero eigenvalue, and their transposes preserve a
‘complementary’ metric with three zero eigenvalues. The Galilei group can be obtained similarly,
by exchanging the roles of the two metrics). In this case, since the metric is not invertible, the
Jacobi relation (41) is not automatically satisfied. It turns out that for the Carroll case it holds
when the vector v is time-like, while the Galilei case requires it to be space-like.
We then found the momentum space associated to a noncommutative spacetime with the com-
mutation relations (39) of κ-Minkowski (generalized), and a deformation of the Carroll group as
quantum group of symmetry. Such a deformation has been already considered recently in [43],
and specifically in the κ-deformed case in [44]. We can add, to the discussion of [44], the obser-
vation that the momentum space associated to this Carrollian κ-deformed noncommutative space
is (part of) a cone.

4.6. Half Minkowski momentum space

As shown in Sec. 2, one can have a flat momentum space with a nondegenerate metric by
choosing a ĝ AB matrix of the form:
⎛ ⎞
0 1 0 0 v0
⎜ 1 0 0 0 v1 ⎟
⎜ ⎟
ĝ = ⎜
AB
⎜ 0 0 1 0 v ⎟,
2⎟ (57)
⎝ 0 0 0 1 v3 ⎠
v0 v1 v2 v3 0
modulo exchanges of 1, 2 3 axes. The induced metric in this case is:
F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117 17

ds 2 = 2 e2p0 /κ dp0 dp1 + e2p0 /κ (dp22 + dp32 ) , (58)


which is locally diffeomorphic to the Minkowski metric. However, the global geometry can’t
be that of Minkowski. Drawing an embedding diagram of the group orbits in this case won’t
help much, so we resort to studying the metric in the pμ coordinates. Begin with the Minkowski
metric:
ds 2 = −dt 2 + dx 2 + dy 2 + dz2 , (59)
and make the following coordinate transformation, consisting of a special conformal transforma-
tion followed by a translation:
√   2 
2 − a + 12 + b2 + p22 + p32 + a + 12
t =− ,
a+b
√   2 
2 a + 12 − b2 − p22 − p32 + b
x=− ,
a+b (60)

2p2
y=− ,
a+b

2p3
z=−
a+b
where
τ − p1 τ + p1
a= √ , b= √ . (61)
2 2
Then the metric (59)takes the form
1  
ds 2 = 2 2dτ dp1 + dp22 + dp32 , (62)
τ
which can be put in the form (58) with the redefinition τ = e−p0 /κ . Now, to find the boundary of
p0 /κ √ √
our coordinate patch, it is sufficient to observe that −(t + x) = e 2 + 2 > 2: we are dealing
with half of Minkowski space.
We found a third model whose quantum group is a deformation of Poincaré. The difference
between the first two, which had, respectively, a de Sitter and an anti-de Sitter momentum space,
was that, among the noncommutative coordinates x μ , the one corresponding to the time-like
direction was x 0 in the first case, and one of the x i in the second case. This third case is at the
interface between the two previous ones: the timelike coordinate is the difference between x 0
and one of the x i , while their sum is spacelike.
Notice that this has nothing to do with the form of the commutation relations (39), where
the parameters v μ are completely free and determine which linear combination of x μ ’s act as a
dilatation for the other ones (which close an Abelian subalgebra). One can have a model with,
say, anti-de Sitter momentum space and timelike x 1 coordinate, with commutation relations that
make x 0 the dilatation-like generator.

4.7. Singular cases

There are cases which correspond to quantum deformations of sensible Lie groups, in which
the matrix ĝ AB is singular. In particular, if two or more diagonal elements λμ are zero, the
18 F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117

matrix is singular, and one cannot write the momentum space manifold as a quadratic form in
the ambient space. We can however study it as an embedded hypersurface, by exponentializing
the representation (43). In the case λ0 = 1, λi = 0, we get an embedded manifold of the form
 
u2
u − u2v1 u2v2 u2v3
YA = 2w
, −  2 , −  2 , −  2 ,w (63)
v0 v 0 (2w) v 0 (2w) v 0 (2w)
0 μ
where u = ek0 v − 1 and w = ekμ v , which is clearly two-dimensional. This choice of parameters
λμ is dual to that of the Carroll group and therefore corresponds to a quantum deformation of the
Galileian group. We find that the corresponding momentum space is topologically degenerate, in
the sense that it is not four-dimensional. The same happens if we choose one of the λi to be the
only nonzero one (which corresponds to a sort of Galilei group in which the spatial rotations and
translations are replaced by a Lorentzian I SO(2, 1) subgroup).
The cases with two zero and two nonzero λμ ’s are all degenerate in the same sense, as can be
straightforwardly verified by considering the 5 × 4 matrix ∂kμ Y A , which has no 4 × 4 submatrix
with nonzero determinant.

5. Conclusions and outlook

We found that the κ-Minkowski noncommutative spacetime, or rather its four-parameter gen-
eralization (39), admits a plethora of nondegenerate momentum spaces, all with the same topol-
ogy but with different metric properties and different signatures (including degenerate ones).
These momentum spaces are each associated to a different quantum group describing the sym-
metries of the same (family of) noncommutative spacetimes (39). These groups are characterized
by the relations (40), where g̃μν and g̃ μν are numerical matrices which, when invertible, are
one the inverse of the other. In these invertible cases one has a quantum deformation of the
Poincaré group I SO(3, 1), the Euclidean group I SO(4), or the group I SO(2, 2). In the first
case (Poincaré), we can associate three different momentum spaces to the model, with very dif-
ferent geometries: a half de Sitter, half anti-de Sitter or a half Minkowski space. These three cases
are all associated with a Poincaré-invariant κ-Minkowski noncommutative spacetime, but they
are distinguished by which of the x μ coordinate (or rather, which of the four conjugate momenta)
corresponds to the timelike direction in momentum space. If it is x 0 (or a Lorentz transformation
thereof), we have a de Sitter momentum space, when it is connected by a Lorentz transformation
to one of the x 1 , x 2 , x 3 coordinates, we have anti-de Sitter, an if it is a 45◦ combination of x 0
with one of the x i , we have a Minkowski space. Notice that this has nothing to do with the choice
of parameters v μ , which determine which linear combination of the x μ acts as a dilatation upon
the other ones, which close an Abelian subalgebra. One can have any of the three momentum
spaces with any choice of v μ . It is also worth mentioning how the light-cone case represents
somehow an interface between the dS and AdS cases: these two momentum spaces have, re-
spectively, positive and negative curvature, while the light cone has zero curvature. Similarly,
the timelike combination of coordinates x μ corresponding to the light-cone momentum space is
the limit case between linear combinations of the form x 0 cosh τ + x i sinh τ for some (possibly
SO(3)-rotated) i and a real value of the parameter τ , which all admit a de Sitter momentum
space, and combinations of the form x 0 sinh τ + x i cosh τ , which correspond to an anti-de Sit-
ter momentum space. It would be interesting to consider κ-deformed models in regimes whose
spacetime symmetries are described by the Carroll group, i.e. in presence of strong gravity (for
an exploration of the phenomenology of such models see, for example, [45]).
F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117 19

Furthermore, we found a previously-unknown momentum space with the geometry of the


(forward) light-cone of a 5D Minkowski space (with one line removed). Strikingly, this momen-
tum space corresponds to a quantum group of symmetries based on the Carroll group, in which
the matrices g̃ μν and g̃μν left invariant by the homogeneous transformations μ ν have, respec-
tively, one and three zero eigenvalues (in orthogonal directions). There is also an associated case
in which the momentum space is still a cone, but the ambient space has signature (2, 2). This
corresponds to a Carroll-like group in which the matrix g̃μν has one zero, one negative and two
positive eigenvalues. We stress that momentum spaces for κ-Minkowski with a (quantum) Caroll
group of symmetry have never been studied before.
Finally, we explored the natural question: if we have a momentum space associated to the
κ-Carroll group, don’t we have one associated to a κ-Galileian group, which can be defined
through relations (40) with the roles of g̃ μν and g̃μν exchanged? The answer appears to be no,
because the corresponding momentum space manifold is not four-dimensional (it is degenerate
in a topological sense, not only metrically). The same holds for a group in which g̃ μν and g̃μν
have both two zero and two nonzero eigenvalues.
This, we believe, exhausts the study of all the interesting momentum spaces that can be
constructed on κ-Minkowski, and their associated quantum groups of space(time) symmetries.
Our analysis also unifies past approaches to κ-Momentum spaces based on right-invariant met-
rics [33] and matrix representations of κ-Minkowski [15,20,28,34], clarifying that there is a
right-invariant metric on the AN (3) group for each one of the momentum spaces that we iden-
tified, and an associated 5D matrix representation of the algebra (39) in terms of a subalgebra
of generators of a Lie group of homogeneous isometries of an ambient 5D metric. The linear
complement of this subalgebra always closes a subalgebra which coincides with the (Lie algebra
of the) homogeneous part of the quantum group of symmetries of κ-Minkowski associated to the
chosen momentum space.

CRediT authorship contribution statement

Fedele Lizzi: Conceptualization, Formal analysis, Funding acquisition, Investigation, Method-


ology, Project administration, Resources, Software, Supervision, Validation, Visualization, Writ-
ing - original draft, Writing - review & editing. Mattia Manfredonia: Conceptualization,
Formal analysis, Investigation, Methodology, Project administration, Resources, Software, Val-
idation, Visualization, Writing - original draft, Writing - review & editing. Flavio Mercati:
Conceptualization, Formal analysis, Funding acquisition, Investigation, Methodology, Project
administration, Resources, Software, Supervision, Validation, Visualization, Writing - original
draft, Writing - review & editing.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal rela-
tionships that could have appeared to influence the work reported in this paper.

Acknowledgements

We thank J. Kowalski-Glikman for pointing out Ref. [39] and the results therein, which partly
superposed with the previous version of our paper, and stimulated us to develop the new approach
described in the present version. FL and MM acknowledge support from the INFN Iniziativa
20 F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117

Specifica GeoSymQFT. FM the Action CA18108 QG-MM from the European Cooperation in
Science and Technology (COST) and partial support from the Foundational Questions Institute
(FQXi) grant SVCF 2018-190483. FL the Spanish MINECO under project MDM-2014-0369 of
ICCUB (Unidad de Excelencia ‘Maria de Maeztu’), grant FPA2016-76005-C2-1-P.

References

[1] J. Lukierski, A. Nowicki, H. Ruegg, Real forms of complex quantum anti-De Sitter algebra Uq (Sp(4, C)) and their
contraction schemes, Phys. Lett. B 271 (1991) 321–328, arXiv:hep-th/9108018 [hep-th].
[2] J. Lukierski, A. Nowicki, H. Ruegg, New quantum Poincaré algebra and k deformed field theory, Phys. Lett. B 293
(1992) 344–352, arXiv:hep-th/9108018 [hep-th].
[3] S. Majid, H. Ruegg, Bicrossproduct structure of kappa Poincaré group and noncommutative geometry, Phys. Lett.
B 334 (1994) 348–354, arXiv:hep-th/9405107 [hep-th].
[4] J. Lukierski, H. Ruegg, A. Nowicki, V.N. Tolstoy, q-deformation of Poincaré algebra, Phys. Lett. B 264 (3) (1991)
331–338, http://www.sciencedirect.com/science/article/pii/037026939190358W.
[5] S. Majid, Hopf algebras for physics at the Planck scale, Class. Quantum Gravity 5 (Dec. 1988) 1587–1606, https://
ui.adsabs.harvard.edu/abs/1988CQGra...5.1587M.
[6] S. Majid, Quantum groups and noncommutative geometry, J. Math. Phys. 41 (2000) 3892–3942, arXiv:hep-th/
0006167 [hep-th].
[7] J. Lukierski, Z. Skoda, M. Woronowicz, κ-deformed covariant quantum phase spaces as Hopf algebroids, Phys.
Lett. B 750 (2015) 401–406, arXiv:1507.02612 [hep-th].
[8] F. Lizzi, M. Manfredonia, F. Mercati, Localizability in κ-Minkowski spacetime, arXiv:1912.07098 [hep-th], 2019.
[9] F. Lizzi, M. Manfredonia, F. Mercati, T. Poulain, Localization and reference frames in κ-Minkowski spacetime,
Phys. Rev. D 99 (8) (2019) 085003, arXiv:1811.08409 [hep-th].
[10] F. Lizzi, P. Vitale, A. Zampini, The fuzzy disc: a review, J. Phys. Conf. Ser. 53 (2006) 830–842.
[11] G. Amelino-Camelia, G. Gubitosi, F. Mercati, Discreteness of area in noncommutative space, Phys. Lett. B
676 (4–5) (Jun 2009) 180–183, arXiv:0812.3663 [hep-th].
[12] S. Majid, Algebraic approach to quantum gravity. II. Noncommutative spacetime, arXiv:hep-th/0604130 [hep-th].
[13] S. Majid, Hopf algebras for physics at the Planck scale, Class. Quantum Gravity 5 (1988) 1587–1606.
[14] S. Majid, Principle of representation-theoretic self-duality, Phys. Essays 4 (1991) 395.
[15] J. Kowalski-Glikman, S. Nowak, Doubly special relativity and de Sitter space, Class. Quantum Gravity 20 (2003)
4799–4816, arXiv:hep-th/0304101 [hep-th].
[16] J. Kowalski-Glikman, De sitter space as an arena for doubly special relativity, Phys. Lett. B 547 (2002) 291–296,
arXiv:hep-th/0207279 [hep-th].
[17] L. Freidel, J. Kowalski-Glikman, S. Nowak, Field theory on κ-Minkowski space revisited: Noether charges and
breaking of Lorentz symmetry, Int. J. Mod. Phys. A 23 (2008) 2687–2718, arXiv:0706.3658 [hep-th].
[18] M. Arzano, J. Kowalski-Glikman, A. Walkus, Lorentz invariant field theory on κ-Minkowski space, Class. Quantum
Gravity 27 (2010) 025012, arXiv:0908.1974 [hep-th].
[19] M. Arzano, J. Kowalski-Glikman, Kinematics of a relativistic particle with de Sitter momentum space, Class. Quan-
tum Gravity 28 (2011) 105009, arXiv:1008.2962 [hep-th].
[20] M. Arzano, T. Trzesniewski, Diffusion on κ-Minkowski space, Phys. Rev. D 89 (12) (2014) 124024, arXiv:1404.
4762 [hep-th].
[21] I. Gutierrez-Sagredo, A. Ballesteros, G. Gubitosi, F.J. Herranz, Quantum groups, non-commutative Lorentzian
spacetimes and curved momentum spaces, in: C. Duston, M. Holman (Eds.), Spacetime Physics 1907 - 2017,
Minkowski Institute Press, Montreal, ISBN 978-1-927763-48-3, 2019, pp. 261–290, arXiv:1907.07979 [hep-th],
2019.
[22] P. Kosinski, J. Lukierski, P. Maslanka, Local field theory on κ-Minkowski space, star products and noncommutative
translations, Czechoslov. J. Phys. 50 (2000) 1283–1290, arXiv:hep-th/0009120.
[23] A. Agostini, G. Amelino-Camelia, M. Arzano, Dirac spinors for doubly special relativity and κ-Minkowski non-
commutative space-time, Class. Quantum Gravity 21 (2004) 2179–2202, arXiv:gr-qc/0207003 [gr-qc].
[24] A. Agostini, F. Lizzi, A. Zampini, Generalized Weyl systems and κ-Minkowski space, Mod. Phys. Lett. A 17 (2002)
2105–2126, arXiv:hep-th/0209174 [hep-th].
[25] A. Agostini, κ-Minkowski representations on Hilbert spaces, J. Math. Phys. 48 (2007) 052305, arXiv:hep-th/
0512114 [hep-th].
F. Lizzi et al. / Nuclear Physics B 958 (2020) 115117 21

[26] M. Arzano, M. Laudonio, Accelerated horizons and Planck-scale kinematics, Phys. Rev. D 97 (8) (2018) 085004,
arXiv:1711.05668 [gr-qc].
[27] F. Mercati, Quantum κ-deformed differential geometry and field theory, Int. J. Mod. Phys. D 25 (05) (2016)
1650053, arXiv:1112.2426 [math.QA].
[28] F. Mercati, M. Sergola, Pauli-Jordan function and scalar field quantization in κ-Minkowski noncommutative space-
time, Phys. Rev. D 98 (4) (2018) 045017, arXiv:1801.01765 [hep-th].
[29] F. Mercati, M. Sergola, Light cone in a quantum spacetime, Phys. Lett. B 787 (2018) 105–110, arXiv:1810.08134
[hep-th].
[30] G. Amelino-Camelia, S. Majid, Waves on noncommutative space-time and gamma-ray bursts, Int. J. Mod. Phys. A
15 (2000) 4301–4324, https://doi.org/10.1142/S0217751X00002777, arXiv:hep-th/9907110 [hep-th].
[31] J.M. Carmona, J.L. Cortes, D. Mazon, F. Mercati, About locality and the relativity principle beyond special relativ-
ity, Phys. Rev. D 84 (2011) 085010, arXiv:1107.0939 [hep-th].
[32] J.M. Carmona, J.L. Cortes, F. Mercati, Relativistic kinematics beyond special relativity, Phys. Rev. D 86 (2012)
084032, arXiv:1206.5961 [hep-th].
[33] D. Jurman, Fuzzy de Sitter space from κ-Minkowski space in matrix basis, Fortschr. Phys. 67 (4) (2019) 1800061,
arXiv:1710.01491 [math-ph].
[34] A. Ballesteros, G. Gubitosi, I. Gutierrez-Sagredo, F.J. Herranz, Curved momentum spaces from quantum (anti–)de
Sitter groups in (3+1) dimensions, Phys. Rev. D 97 (10) (2018) 106024, arXiv:1711.05050 [hep-th].
[35] A. Ballesteros, I. Gutierrez-Sagredo, F. Mercati, Coreductive Lie bialgebras and dual homogeneous spaces, arXiv:
1909.01000 [math-ph].
[36] A. Ballesteros, I. Gutierrez-Sagredo, F.J. Herranz, The κ-(A)dS noncommutative spacetime, Phys. Lett. B 796
(2019) 93–101, arXiv:1905.12358 [math-ph].
[37] M.K. Parikh, I. Savonije, E. Verlinde, Elliptic de Sitter space: dS/Z2 , Phys. Rev. D 67 (2003) 064005, arXiv:
hep-th/0209120.
[38] F. Mercati, M. Sergola, Physical constraints on quantum deformations of spacetime symmetries, Nucl. Phys. B 933
(2018) 320–339, arXiv:1802.09483 [hep-th].
[39] A. Blaut, M. Daszkiewicz, J. Kowalski-Glikman, S. Nowak, Phase spaces of doubly special relativity, Phys. Lett. B
582 (2004) 82–85, arXiv:hep-th/0312045 [hep-th].
[40] S. Pascu, Atlas of coordinate charts on de Sitter spacetime, arXiv:1211.2363 [gr-qc].
[41] J.M. Lévy-Leblond, Une nouvelle limite non-relativiste du groupe de Poincaré, Ann. IHP, Phys. Théor. 3 (1) (1965)
1–12, http://www.numdam.org/item/AIHPA_1965__3_1_1_0.
[42] C. Duval, G.W. Gibbons, P.A. Horvathy, P.M. Zhang, Carroll versus Newton and Galilei: two dual non-Einsteinian
concepts of time, Class. Quantum Gravity 31 (2014) 085016, arXiv:1402.0657 [gr-qc].
[43] M. Daszkiewicz, Canonical and Lie-algebraic twist deformations of Carroll, para-Galilei and static Hopf algebras,
Mod. Phys. Lett. A 34 (2019) 1950181.
[44] A. Ballesteros, G. Gubitosi, I. Gutierrez-Sagredo, F.J. Herranz, The k-Newtonian and k-Carrollian algebras and
their noncommutative spacetimes, arXiv:2003.03921v1.
[45] G. Amelino-Camelia, N. Loret, G. Mandanici, F. Mercati, Gravity in quantum spacetime, Int. J. Mod. Phys. D
19 (14) (Dec 2010) 2385–2392, arXiv:1007.0851 [gr-qc].

You might also like