Acta Polytechnica doi:10.14311/AP.2016.56.0173 Acta Polytechnica 56(3):173–179, 2016 © Czech Technical University in Prague, 2016 available online at http://ojs.cvut.cz/ojs/index.php/ap COVARIANT INTEGRAL QUANTIZATIONS AND THEIR APPLICATIONS TO QUANTUM COSMOLOGY Jean Pierre Gazeau Astroparticle and Cosmology, Univ Paris Diderot, Sorbonne Paris Cité, Paris 75205, France correspondence: gazeau@apc.univ-paris7.fr Abstract. We present a general formalism for giving a measure space paired with a separable Hilbert space a quantum version based on a normalized positive operator-valued measure. The latter are built from families of density operators labeled by points of the measure space. We especially focus on group representation and probabilistic aspects of these constructions. Simple phase space examples illustrate the procedure: plane (Weyl-Heisenberg symmetry), half-plane (affine symmetry). Interesting applications to quantum cosmology (“smooth bouncing”) for Friedmann-Robertson-Walker metric are presented and those for Bianchi I and IX models are mentioned. Keywords: integral quantization; covariance; POVM; affine group; Weyl-Heisenberg group; coherent states; FRW model; smooth bouncing. 1. Introduction Group theory is one of the favorite domains of J. Pa- tera and P. Winternitz. This contribution to the cele- bration of the Jiri and Pavel 80th birthday emphasizes the role of group representation theory in a certain type of quantization procedures. These methods [1–9] are named (covariant) integral quantizations and offer a large scope of possibilities for mapping a classical mathematical model of a physical system to a quan- tum model. They are based on normalized positive operator-valued measures (POVM) and present at least two attractive advantages: (1.) By employing all resources of integration and distributions, they are most appropriate when we have to deal with some geometric singularities. (2.) They afford a semi-classical phase space portrait of quantum states and quantum dynamics together with a probabilistic interpretation. Moreover, their recent applications to quantum cos- mology [10–15] has yielded interesting results as: • consistent quantum dynamics of isotropic, aniso- tropic non-oscillatory and anisotropic models of Universe; • singularity resolution; • unitary dynamics without boundary conditions; • consistent semi-classical description of involved quantum dynamics. The aim of the present article is to give an overview of the methods, their main characteristics, and their promising applications to early cosmology. In Sec- tion 2 we give a rapid survey of canonical quantiza- tion with the problems raised by this procedure, and we define a few basic requirements that any quan- tization procedure should fullfill. In Section 3 we define POVM-based integral quantizations of func- tions defined on a measure space, their subsequent semi-classical counterparts (i.e., lower symbols) and discuss their possibilities when they are compared with other methods. Covariant integral quantizations are then considered when the measure space is homo- geneous for some group, and we give the example of the half-plane viewed as the affine group of the real line. Section 4 is devoted to another example, the integral quantization of functions on the plane viewed as a coset of the Weyl-Heisenberg group. In Section 5 we illustrate the method with recent applications to quantum cosmology where the initial singularity is regularized by using the affine integral quantization. Finally we sketch in Section 6 a more philosophical point of view on the approaches described in this contribution. A large part of the content has already been pub- lished, as exemplified by the above citations. On the other hand the original side of the paper lies in the synthetic presentation of the method together with some new ideas like the general (and promising) con- struction sketched in Section 3.2. 2. What is quantization? 2.1. Canonical quantization The basic procedure of quantization of a classical me- chanical model is well known and is named “canonical" (Heisenberg, Born, Jordan, Dirac, Weyl, etc). Starting from a phase space or symplectic manifold, e.g. R2, e.g. the phase space for the motion of particle on the line, it consists in the maps R2 3 (q,p), {q,p} = 1 7→ self-adjoint (Q,P), (1) with the canonical commutation rule (ccr) [Q,P] = i~I, and f(q,p) 7→ f(Q,P) 7→ (Sym f)(Q,P). (2) 173 http://dx.doi.org/10.14311/AP.2016.56.0173 http://ojs.cvut.cz/ojs/index.php/ap Jean Pierre Gazeau Acta Polytechnica Beyond the ordering problem [16–18] raised by the sec- ond step (symmetrisation), one should keep in mind that [Q,P] = i~I holds true with self-adjoint Q, P, only if both have continuous spectrum (−∞, +∞), and there is uniqueness of the solution, up to unitary equivalence (von Neumann). But then what about singular f, e.g., the angle arctan(p/q)? What about barriers or other impassable boundaries? The mo- tion on a circle? In a bounded interval? On the half-line? Despite their elementary aspects, these ex- amples leave open many questions both on mathemat- ical and physical levels, irrespective of the manifold quantization procedures, like Path Integral Quantiza- tion (Feynman, thesis, 1942), Geometric Quantization with Weyl (1927), Groenewold (1946), Kirillov (1961), Souriau (1966), Kostant (1970), Deformation Quan- tization with Moyal (1947), Bayen, Flato, Fronsdal, Lichnerowicz, Sternheimer (1978), Fedosov (1985), Kontsevich (2003), Coherent state or anti-Wick or Toeplitz quantization with Klauder (1961), Berezin (1974). Of course, the canonical procedure is universally accepted in view of its numerous experimental valida- tions, one of the most famous and simplest one going back to the early period of Quantum Mechanics with the quantitative prediction of the isotopic effect in vibrational spectra of diatomic molecules [19]. These data validated the canonical quantization, contrary to the Bohr-Sommerfeld ansatz (which predicts no isotopic effect). Nevertheless this does not prove that another method of quantization fails to yield the same prediction [3]. Moreover, as already mentioned above, the canonical quantization is difficult, if not impos- sible, in many circumstances. As a matter of fact, the canonical or the Weyl-Wigner integral quantiza- tion maps f(q) to f(Q) (resp. f(p) to f(P)), and so are unable to cure any kind of classical singularity. Clearly, quantization procedures may be intractable when different phase space geometries and/or topolo- gies are considered. Self-adjointness of basic operators is not guaranteed anymore. 2.2. Quantization: minimal requirements In our viewpoint, any quantization of a set X (e.g., a phase space or something else) and functions on it should meet four basic requirements: (1.) Linearity — it is a linear map Q : C(X) 7→A(H), Q(f) ≡ Af, (3) where: • C(X) is a vector space of complex-valued func- tions f(x) on set X, i.e., a “classical” mathemat- ical model; • A(H) is a vector space of linear operators (ignor- ing domain limitations) in some complex Hilbert space H, i.e, a “quantum” mathematical model. (2.) Unity — The map (3) is such that the function f = 1 is mapped to the identity operator I on H. (3.) Reality — A real f is mapped to a self-adjoint operator Af in H or, at least, a symmetric operator, (4.) Covariance — If X is endowed with symmetry, the latter should be preserved in a “certain” sense. Of course, further requirements are necessarily added, depending on the mathematical structures equipping X and C(X) (e.g., measure, topology, man- ifold, closure under algebraic operations, time evolu- tion or dynamics, etc). Moreover, a physical inter- pretation should be advanced about measurement of spectra of classical f ∈C(X) or quantum Af ∈A(H) to which are given the status of observables. Finally, an unambiguous classical limit of the quantum physi- cal quantities should exist, the limit operation being associated to a change of scale. 3. Integral Quantization(s) 3.1. Integral quantization: general setting and POVM Let (X,ν) be a measure space. Suppose that there exists an X-labeled family of bounded operators on a Hilbert space H resolving the identity I: X 3 x 7→ M(x) ∈L(H), ∫ X M(x) dν(x) = I, (4) where the equality holds in a weak sense. The case we are specially interested in occurs when the M(x)’s are positive semi-definite and unit trace, i.e., when M(x) ≡ ρ(x) (density operator). (5) Then, if X is a space with suitable topology, the map B(X) 3 ∆ 7→ ∫ ∆ ρ(x) dν(x) (6) may define a normalized positive operator-valued mea- sure (POVM) on the σ-algebra B(X) of Borel sets. The quantization of complex-valued functions f(x) on X is the linear map: f 7→ Af = ∫ X M(x)f(x) dν(x), (7) understood as the sesquilinear form, Bf (ψ1,ψ2) = ∫ X 〈ψ1|M(x)|ψ2〉f(x) dν(x), (8) defined on a dense subspace of H. If f is real and at least semi-bounded, and if the M(x)’s are posi- tive operators, then the Friedrich’s extension of Bf univocally defines a self-adjoint operator. If f is not semi-bounded, there is no natural choice of a self- adjoint operator associated with Bf. We then need more information on H to solve this subtlety. Covariance implies symmetry. Symmetry suggests a mathematical group G. A symmetry structure on X presupposes that some group G acts on X. Our hypothesis here is that X is an homogeneous space for G, which means that X ∼ G/H for H a subgroup of G. We now examine two possible cases. 174 vol. 56 no. 3/2016 Covariant Integral Quantizations and Quantum Cosmology 3.1.1. X = G: quantizing the group Let G be a Lie group with left invariant Haar measure dµ(g), and let g 7→ U(g) be a unitary irreducible representation (UIR) of G in a Hilbert space H. Let M be a bounded operator on H. Suppose that the operator R := ∫ G M(g) dµ(g), M(g) := U(g)MU(g)†, (9) is defined in a weak sense. From the left invariance of dµ(g) the operator R commutes with all operators U(g), g ∈ G, and so, from Schur’s Lemma, R = cMI with cM = ∫ G tr ( ρ0M(g) ) dµ(g), (10) where the unit trace positive operator ρ0 is chosen in order to make the integral convergent. If cM is finite and not zero, then we get the resolution of the identity:∫ G M(g) dν(g) = I, dν(g) := dµ(g)/cM. (11) For this reason, the operator M to be U transported will be designated (admissible) fiducial operator. The method becomes more apparent when the representa- tion U is square-integrable in the following sense (for example, it is the case for the affine group of the real line, see below). Suppose that there exists a density operator ρ which is admissible in the sense that cρ = ∫ G dµ(g) tr ( ρρ(g) ) < ∞, (12) with ρ(g) = U(g)ρU†(g). Then the resolution of the identity (11) holds with M(g) and cM = cρ. This allows a covariant integral quantization of complex- valued functions on the group f 7→ Af = ∫ G ρ(g)f(g) dν(g), dν(g) := dµ(g) cρ , (13) U(g)AfU†(g) = AU(g)f (covariance), (14) where ( U(g)f ) (g′) := f(g−1g′) (15) is the regular representation if f ∈ L2(G, dµ(g)). This leads to a non-commutative version of the original manifold structure for G. Example: affine CS integral quantization In this example, the group G is the affine group of the real line and is also viewed as the phase space for the motion on the half-line, i.e., the half-plane {(q,p) ∈ R∗+ ×R}. The group law is defined as (q,p)(q0,p0) = ( qq0, p0 q + p ) , (16) and the left invariant measure is the symplectic dq dp. This group has two UIR’s which are both square integrable. We choose the following one. L2(R∗+, dx) 3 ψ(x) 7→ ( U(q,p)ψ ) (x) = eipx √ q ψ (x q ) (17) As a density operator we choose the rank one pro- jector ρ = |ψ〉〈ψ| where the fiducial vector, i.e., the ‘wavelet” ψ is in L2(R+, dx) ∩L2(R+, dx/x) and its affine transport produces the overcomplete family of affine coherent states (ACS) |q,p〉 = U(q,p)|ψ〉. (18) The ACS integral quantization reads as f(q,p) 7→ Af = ∫ R+×R dq dp cρ f(q,p)|q,p〉〈q,p| (19) with cρ = ∫∞ 0 |ψ(x)| 2 dx x . More details and new devel- opments are given in the recent [20]. Applications of this method to the resolution of the singularity prob- lem on a quantum level for a series of cosmological models (FRW, Bianchi I, Bianchi IX), for which the volume — expansion pair variables form the above half-plane, are given in the recent papers [10–15] and will be summarised in Section 5. 3.1.2. X = G/H: quantizing a non-trivial group coset In the absence of square-integrability over G (the trivial illustration is the Weyl-Heisenberg group which underlies in particular the canonical quantization, see Section 4), there exists a definition of square-integrable representations with respect to a left coset manifold X = G/H, with H a closed subgroup of G (it is the center in the Weyl-Heisenberg case), equipped with a quasi-invariant measure ν [2]. For a global Borel section σ : X → G of the group, let νσ be the unique quasi-invariant measure defined by dνσ(x) = λ ( σ(x),x ) dν(x), (20) where λ(g,x)dν(x) = dν(g−1x) for g ∈ G with λ(g,x) obeying the cocycle condition λ(g1g2,x) = λ(g1,x)λ(g2,g−11 x). Let U be a UIR which is square integrable mod(H) with an admissible ρ, i.e., cρ := ∫ X tr(ρρσ(x)) dνσ(x) < ∞, with ρσ(x) = U(σ(x))ρU(σ(x))†. Then we have the resolution of the identity and the resulting quantization f 7→ Af = 1 cρ ∫ X f(x)ρσ(x) dνσ(x) (21) Covariance holds here too in a certain sense (see [2, Chapter 11]). Besides the Weyl-Heisenberg group, another exam- ple concerns the motion on the circle for which G is the group of Euclidean displacements in the plane, i.e., the semi-direct product R2 o SO(2), and the subgroup H is isomorphic to R [22]. Other examples involve the relativity groups, Galileo, Poincaré [2], 1 + 1 Anti de Sitter (unit disk and SU(1, 1)) [21], 1 + 1 and 3 + 1 de Sitter [23]. 175 Jean Pierre Gazeau Acta Polytechnica 3.2. An example of construction of original operator M or ρ Let U be a UIR of G and $(x) be a function on the coset X = G/H. Suppose that it allows to define a bounded operator M$σ on H through the operator- valued integral M$σ = ∫ X $(x)U(σ(x)) dνσ(x). (22) Then, under appropriate conditions on the “weight” function $(σ(x)) such that U be a UIR which is square integrable mod(H) and M$σ is admissible in the above sense, the family of transported operators M$σ (x) := U(σ(x))M$σ U†(σ(x)) resolves the identity. 3.3. Semi-classical portraits Some quantization features, e.g., spectral properties of Af, may be derived or at least well grasped from functional properties of the lower (Lieb) or covariant (Berezin) symbol (it generalizes Husimi function or Wigner function) Af 7→ f̌(x) := tr ( M(x)Af ) , (23) When M = ρ (density operator) this new function is the local average of the original f with respect to the probability distribution tr(ρ(x)ρ(x′)) f(x) 7→ f̌(x) = ∫ X f(x′) tr ( ρ(x)ρ(x′) ) dν(x′). (24) The Bargmann-Segal-like map f 7→ f̌ which, gen- eralizes the Berezin or heat kernel transform when X = G is a Lie group, is in general a regularization of the original, possibly extremely singular, f, and the classical limit itself means that, given one or more scale parameter(s) �(i) and a distance d(f, f̌), we have d(f, f̌) → 0 as �(i) → 0. (25) 3.4. Comments Beyond the freedom (think of analogy with Signal Analysis where different techniques are complemen- tary) allowed by integral quantization, the advantages of the method with regard to other quantization pro- cedures in use are of four types. (i)The minimal amount of constraints imposed on the classical objects to be quantized. (ii)Once a choice of (positive) operator-valued mea- sure has been made, which must be consistent with experiment, there is no ambiguity in the issue, con- trarily to other method(s) in use (think in particular of the ordering problem). To one classical model corresponds one and only one quantum model. Of course different choices of (P)OVM are requested to be physically equivalent, e.g., leading to the same experimental predictions, if they are aimed to de- scribe one specific system. (iii)The method produces in essence a regularizing effect, at the exception of certain choices, like the Weyl-Wigner (i.e., canonical) integral quantization. (iv)The method, through POVM choices, offers the possibility to keep a fully probabilistic content. As a matter of fact, the Weyl-Wigner integral quanti- zation does not rest on a POVM. 4. Weyl-Heisenberg covariant integral quantization(s) The Weyl-Heisenberg group is defined as GWH = {(s,z) | s ∈ R, z ∈ C} with multiplication law (s,z)(s′,z′) = ( s + s′ + Im(zz̄′),z + z′ ) (26) Let H be a separable (complex) Hilbert space with orthonormal basis e0,e1, . . . ,en ≡ |en〉, . . . , and cor- responding lowering and raising operators defined in the usual way a|en〉 = √ n|en−1〉, a|e0〉 = 0, a†|en〉 = √ n + 1|en+1〉, [a,a†] = I, It is well known that the Weyl-Heisenberg group has a unique non-trivial UIR U, up to unitary equivalence, each element in this equivalence class corresponding to a non-zero real number like the Planck constant. Consider the center C = {(s, 0) | s ∈ R} of GWH. Then, the set X is the coset X = GWH/C ∼ C with measure d2z/π. Choosing the trivial section C 3 z 7→ σ(z) = (0,z), to each z corresponds the (unitary) displacement (∼ Weyl) operator D(z) := U(σ(z)) = eza †−z̄a, D(−z) = (D(z))−1 = D(z)†. (27) The noncommutativity of quantum mechanics is en- coded in the relation D(z)D(z′) = e 1 2 (zz̄ ′−z̄z′)D(z + z′) = e(zz̄′−z̄z ′)D(z′)D(z), (28) i.e., z 7→ D(z) is a projective representation of the abelian group C. The standard (i.e., Schrödinger- Klauder-Glauber-Sudarshan) CS are defined as |z〉 = D(z)|e0〉. (29) We now adopt the construction sketched in Section 3.2. Let $(z) be a function on the complex plane obeying $(0) = 1. Suppose that it allows to define a fidu- cial operator M$ on H through the operator-valued integral M$ = ∫ C $(z)D(z) d2z π . (30) Then, the family of displaced M$(z) := D(z)M$D(z)† under the unitary action D(z) resolves the identity ∫ C M$(z) d2z π = I. (31) 176 vol. 56 no. 3/2016 Covariant Integral Quantizations and Quantum Cosmology It is a direct consequence of D(z)D(z′)D(z)† = ezz̄ ′−z̄z′D(z′), of ∫ C e zξ̄−z̄ξ d2ξ π = πδ2(z), and of $(0) = 1 with D(0) = I. The resulting quantiza- tion map is given by f 7→ A$f = ∫ C M$(z)f(z) d2z π = ∫ C $(z)D(z)fs[f](z) d2z π , (32) where are involved the symplectic Fourier transforms fs and its space reverse fs fs[f](z) = ∫ C ezξ̄−z̄ξf(ξ) d2ξ π , fs[f](z) = fs[f](−z) Both are unipotent fs[fs[f]] = f and fs[fs[f]] = f. We have the following covariance properties • Translation covariance: A$f(z−z0) = D(z0)A $ f(z)D(z0) †. (33) • Parity covariance A$f(−z) = PA $ f(z)P ∀f ⇐⇒ $(z) = $(−z) ∀z, (34) where P = ∑∞ n=0(−1) n|en〉〈en| is the parity opera- tor. • Complex conjugation covariance A$ f(z) = ( A$f(z) )† ∀f ⇐⇒ $(−z) = $(z) ∀z, (35) • For rotational covariance we define the unitary rep- resentation θ 7→ UT(θ) of S1 on the Hilbert space H as the diagonal operator UT(θ)|en〉 = ei(n+ν)θ|en〉, where ν is arbitrary real. From the matrix elements of D(z) one proves easily the rotational covariance property UT(θ)D(z)UT(θ)† = D(eiθz), (36) and its immediate consequence on the nature of M$ and the covariance of A$f : the equality UT(θ)A$f UT(−θ) = A $ T(θ)f, (37) with T(θ)f(z) := f ( e−iθz ) , holds true if and only if $(eiθz) = $(z) ∀z,θ, i.e., if and only if M$ is diagonal. 5. Applications of ACS quantization in quantum cosmology Let us consider the Friedmann-Robertson-Walker model filled with barotropic fluid with equation of state p = wρ. The line element is given by: ds2 = −N(t)2dt2 + a(t)2δijωiωj, (38) where a(t) is the scale factor and the ωi’s are right- invariant dual vectors. The resolving of the Hamilto- nian constraint leads to a model of singular universe analogous to a particle moving on the half-line (0,∞). Indeed, canonical coordinates (q,p) can be introduced in terms of the volume V ∼ a3 and the expansion rate V̇ /V ∼ ȧ/a through the relations V = q2/(1−w) and K = (3/8)(1 −w)pq−(1+w)/(1−w). Clearly, q > 0 and p ∈ R. Then the reduced Hamiltonian reads as {q,p} = 1, h(q,p) = α(w)p2 + 6k̃qβ(w), q > 0. with k̃ = ( ∫ dω)2/3k, α(w) = 3(1−w)2/32 and β(w) = 2(3w + 1)/(3(1 −w)). k = 0,−1 or 1 (in suitable unit of inverse area) depending on whether the universe is flat, open or closed. Assuming a closed universe with radiation content w = 1/3 and k = +1, the classical isotropy singularity is cured on the quantum level by using affine CS quantization [10]. Indeed the latter regularizes the kinetic term by adding a repulsive centrifugal potential which prevents the system from reaching the value q = 0. Precisely, with the choice of a fiducial vector ψ, h = 1 24 p2 + 6q2 7→ Ah = 1 24 P 2 + K(ψ) 24 1 Q2 + 6M(ψ)Q2, (39) P ≡−i d dx , Qφ(x) ≡ xφ(x). The constants K(ψ) and M(ψ) are > 0 for all ψ and their appearance is due to the affine CS quantization. Moreover, for a ψ such that K(ψ) ≥ 34 the quantum Hamiltonian Ah is essentially self-adjoint, giving a unique unitary evolution: there is no need of boundary conditions. The semi-classical dynamics is ruled by the ACS mean value of the quantum Hamiltonian : 〈q,p|Ah|q,p〉 = c(ψ) ∫ R+×R dq′dp′ ∣∣〈q′,p′|q,p〉∣∣2h(q′,p′), (40) with a displacement of the equilibrium point of the potential at Q4eq = 1 144 K M . The constant c(ψ) is > 0 for all ψ. In Figure 1 contour plots of phase space trajectories for classical Hamiltonian h(q,p) (left) and semi-classical Hamiltonian (right) defined by (40) are compared. As the result of affine quantization and restoring arbitrariness on w and k, we obtain two corrections to the Friedmann equation which reads in its semi- classical version as(ȧ a )2 + c2a2P (1 −w) 2A 1 V 2 + B kc2 a2 = 8πG 3c2 ρ, (41) where A and B are positive factor dependent on ψ and can be adjusted at will in consistence with (so far very hypothetical!) observations. The first correction is the repulsive potential, which depends on the volume, 177 Jean Pierre Gazeau Acta Polytechnica Figure 1. Contour plots of the classical and semiclas- sical Hamiltonians for the closed Friedmann universe. The singularity q = 0 on the left is replaced with a bounce on the right. and this excludes non-compact universes from quan- tum modelling. We notice that as the singularity is approached a → 0, this potential grows faster (∼ a−6) than the density of fluid (∼ a−3(1+w)) and therefore at some point the contraction must come to a halt. Second, the curvature becomes dressed by the factor B. This effect could in principle be observed far away from the quantum phase. However, we do not observe the intrinsic curvature neither in the geometry nor in the dynamics of space. Nevertheless, for a convenient choice of ψ, this factor ≈ 1 Also note that the form of the repulsive potential does not depend on the state of fluid filling the uni- verse: the origin of singularity avoidance is quantum geometrical. The applications of the same quantization methods to more involved cosmological models like Bianchi I and Bianchi IX are described in [11–15]. For other approaches to quantum cosmology based on affine symmetry see [24–26]. 6. Conclusion To conclude with an allusion to more “philosophical" perspectives, we think that integral quantization is just a part of a world of mathematical models for one phenomenon. Indeed, the physical laws are expressed in terms of combinations of mathematical symbols, and these combinations make sense for people who have learnt these elements of mathematical language. This language is in constant development, enhance- ment, since the set of phenomenons which are acces- sible to our understanding is constantly broadening. These combinations take place within a mathemat- ical model. A model is usually scale dependent. It depends on a ratio of physical (i.e., measurable) quan- tities, like lengths, time(s), sizes, impulsions, actions, energies, etc. Changing scale for a model amounts to “quantize” or “de-quantize” toward (semi-)classicality. One changes perspective like one would change glasses to examine one’s environment. Acknowledgements The author is indebted to Jiri Patera and Pavel Winter- nitz for inviting him at many occasions to share their enthusiasm in search for symmetry. He is also indebted to H. Bergeron, E. Czuchry, and P. Małkiewicz, for having developed with them a large part of the material exposed in the present contribution. References [1] J.-P. Gazeau, Coherent States in Quantum Physics, Wiley-VCH, Berlin, 2009. [2] S. T. Ali, J.-P Antoine, and J.-P. Gazeau, Coherent States, Wavelets and their Generalizations 2d edition, Theoretical and Mathematical Physics, Springer, New York (2013), specially Chapter 11. [3] H. Bergeron, J.-P. Gazeau, and A. Youssef, Are the Weyl and coherent state descriptions physically equivalent? Phys. Lett. A 377 598(2013); arXiv:1102.3556 [4] H Bergeron and J.-P Gazeau, Integral quantizations with two basic examples, Annals of Physics (NY) 344 43 (2014); arXiv:1308.2348 [5] H. Bergeron, E. M. F. Curado, J.-P. Gazeau, and Ligia M. C. S. Rodrigues, Quantizations from (P)OVM’s, J. Phys.: Conf. Ser. (2014); arXiv:1310.3304 [6] M. Baldiotti, R. Fresneda, and J.-P. Gazeau, Three examples of covariant integral quantization, Proceedings of Science (2014). [7] J.-P. Gazeau and B. Heller, Positive-Operator Valued Measure (POVM) quantization, Axioms (Special issue on Quantum Statistical Inference) 4 1 (2015); http://www.mdpi.com/2075-1680/4/1/1 [8] M. Baldiotti, R. Fresneda, and J.-P. Gazeau, About Dirac&Dirac constraint quantizations, Invited comment, Phys. Scr. 90 074039 (2015). [9] R. Fresneda and J.-P Gazeau, Integral quantizations with POVM and some applications, in Proceedings of the 30th International Colloquium on Group Theoretical Methods eds. J. Van der Jeugt et al, Journal of Physics: Conference Series 597 (2015) 012037. [10] H. Bergeron, A. Dapor, J.-P. Gazeau, and P. Małkiewicz, Smooth big bounce from affine quantization, Phys. Rev. D 89 083522 (2014); arXiv:1305.0653 [11] H Bergeron, A Dapor, J.-P. Gazeau, and P Małkiewicz, Smooth Bounce in Affine Quantization of Bianchi I, Phys. Rev D 91 124002 (2015); arXiv:1501.07718 [12] H. Bergeron, E. Czuchry, J.-P. Gazeau, P. Małkiewicz, and W. Piechocki, Smooth quantum dynamics of the Mixmaster universe, Phys. Rev. D 92 061302(R) (2015); arXiv:1501.02174 [13] H. Bergeron, E. Czuchry, J.-P. Gazeau, P. Małkiewicz, and W. Piechocki, Singularity avoidance in quantum Mixmaster universe, Phys. Rev D 92 124018 (2015); arXiv:1501.07871 [14] H. Bergeron, E. Czuchry, J.-P. Gazeau, and P. Małkiewicz, Inflationary aspects of quantum Mixmaster universe, submitted; arXiv:1511.05790 [15] H. Bergeron, E. Czuchry, J.-P. Gazeau, and P. Małkiewicz, Vibronic framework for quantum Mixmaster universe, Phys. Rev D 93 064080 (2016); arXiv:1512.00304 178 http://arxiv.org/abs/1102.3556 http://arxiv.org/abs/1308.2348 http://arxiv.org/abs/1310.3304 http://www.mdpi.com/2075-1680/4/1/1 http://arxiv.org/abs/1305.0653 http://arxiv.org/abs/1501.07718 http://arxiv.org/abs/1501.02174 http://arxiv.org/abs/1501.07871 http://arxiv.org/abs/1511.05790 http://arxiv.org/abs/1512.00304 vol. 56 no. 3/2016 Covariant Integral Quantizations and Quantum Cosmology [16] M. Born and P. Jordan, On quantum mechanics, Zs. f. Phys. 34 858 (1925). [17] B. S. Agarwal and E. Wolf, Calculus for Functions of Noncommuting Operators and General Phase-Space Methods in Quantum Mechanics, Phys. Rev. D 2 2161; 2187; 2206 (1970). [18] F. A. Berezin and M. A. Shubin, The Schrödinger Equation, Kluwer Academic Publishers, Dordrecht, 1991. [19] G. Herzberg, Molecular Spectra and Molecular Structure: Spectra of Diatomic Molecules, 2nd. ed., Krieger Pub., Malabar, FL, 1989. [20] J.-P. Gazeau and R. Murenzi, Covariant Affine Integral Quantization(s), accepted to J. Math. Phys. (2016), arXiv:1512.08274 [21] J.-P. Gazeau and M. del Olmo, Unit disk & SU(1,1) integral quantizations, in preparation. [22] R. Fresneda, J.-P. Gazeau, and D. Noguera, Covariant integral quantization of the motion on the circle, in preparation. [23] J.-P. Gazeau, A. Izadi, A. Rabei, and M. Takook, Covariant Integral Quantization of the motion on 1 + 1 de Sitter space-time, in preparation. [24] M. Fanuel and S. Zonetti, Affine Quantization and the Initial Cosmological Singularity, Eur. Phys. Lett. 101, 10001 (2013); [25] J. R. Klauder, An Affinity for Affine Quantum Gravity, Proc. Steklov Institute of Mathematics 272, 169-176 (2011); gr-qc/1003.261 [26] J. R. Klauder, Enhanced Quantization, Particles, Fields & Gravity, World Scientific, Singapore, 2015. 179 http://arxiv.org/abs/1512.08274 Acta Polytechnica 56(3):173–179, 2016 1 Introduction 2 What is quantization? 2.1 Canonical quantization 2.2 Quantization: minimal requirements 3 Integral Quantization(s) 3.1 Integral quantization: general setting and POVM 3.1.1 X=G: quantizing the group 3.1.2 X=G/H: quantizing a non-trivial group coset 3.2 An example of construction of original operator M or rho 3.3 Semi-classical portraits 3.4 Comments 4 Weyl-Heisenberg covariant integral quantization(s) 5 Applications of ACS quantization in quantum cosmology 6 Conclusion Acknowledgements References