<<

arXiv:2107.07879v1 [gr-qc] 23 May 2021 itnt rbesoeo hc sscesu n h te n saf a is one other the procedures. and quantization successful simple, the is two of which and both of using procedures, one quantization problems different distinct, two with begin We Introduction 1 ino of tion ∗ hseecs evsa rld oavldadsrihfradqua straightforward and valid a to prelude a as serves exercise This klauder@ufl.edu eateto hsc n eateto Mathematics of Department and Physics of Department htehbttepoon ieecsbtenapcso seek of gravity. aspects mo Einstein’s between of a f differences treatment typical opens quantum profound some analysis the from That exhibit gleaned gravity that quantum understood. of be p treatment can simple est two procedure for each quantizations affine so W and coordinate. canonical positive using strictly sion a e.g., orientation, gravi one augmen for in quantization quantization canonical correct like the is which find quantization, to effort an in tization nvriyo lrd,Gievle L32611-8440 FL Gainesville, Florida, of University opQatmGaiyi ieydvlpduigcnnclqu canonical using developed widely is Gravity Quantum Loop n ffieQatmGravity Quantum Affine and e opQatmGravity Quantum Loop Let xmn ahOther Each Examine onR Klauder R. John Abstract 1 ∗ pndiscus- open e e bounded ted y Affine ty. roblems n the ing eatures an- d- ntiza- ailure but 1.1 Choosing a canonical quantization The classical variables, p&q, that are elements of a constant zero , better known as Cartesian variables, such as those featured by Dirac [1], are promoted to self-adjoint quantum operators P (= P †) and Q (= Q†), ranged as −∞

1.1.1 First canonical example Our example is just the familiar harmonic oscillator, for which −∞

i~(∂ψ(x, t)/∂t)=(−~2(∂2/∂x2)+ x2)/2 ψ(x, t) . (1)

Solutions to Eq. (1) for our example are well known. In particular, for the harmonic oscillator: the eigenvalues are given by En = ~(n + 1/2) for n = 0, 1, 2, , ...; the eigenfunctions (with ~ = 1) are given by ψn(x) = −x2/2 NHn(x) e with n = 0, 1, 2, ...; Hn(x) are the Hermite polynomials [2]; and N is a normalization factor.

1.1.2 Second canonical example For our next example we keep the same classical Hamiltonian, and we retain −∞

1In particular, in [1], the mid-page of 114, Dirac wrote “However, if the system does have a classical analogue, its connexion with classical is specially close and one can usually assume that the Hamiltonian is the same function of the canonical coordinates and momenta in the quantum theory as in the classical theory.† ” Footnote “† This assumption is found in practice to be successful only when applied with the dynamical coordinates and momenta referring to a Cartesian system of axes and not to more general curvilinear coordinates.”

2 when ~ → 0. Hence, there are two valid sets of eigenvalues, specifically E0 = ~[(0, 2, 4, ...)+1/2] and E1 = ~[(1, 3, 5, ...)+1/2]. Moreover, these two eigenvalues can be weaved together, such as E = ~[(0, 2, 3, 5, 7, 8, ...)+1/2], which implies that there are infinitely many versions of the spectrum. That judgement renders canonical quantization of the half-harmonic os- cillator to be an invalid quantization.

1.1.3 First affine example The traditional classical affine variables are d ≡ pq and q > 0, and they have a Poisson bracket given by {q,d} = q. Although the harmonic oscillator extends over the whole real line, we choose q to satisfy −b 0. For very large b we can approximate a full-line harmonic oscillator and even see what happens if we choose b →∞ to mimic the full-line story. The classical affine variables now are −∞ < d ≡ p(q + b) < ∞ and 0 < (q + b) < ∞, while the classical harmonic oscillator Hamiltonian is given by H′(d, q) = [d2 (q+b)−2 +q2]/2, an expression that obeys H(p, q)=(p2 +q2)/2 albeit that −b 0. The quantum partial-harmonic oscillator is now given by

H′(D, Q) = [D(Q + b)−2D + Q2]/2 , (2)

′ † 2 which, when b → ∞, we find limb→∞ H (D, Q)=(P P + Q )/2), for which † P = P now since −∞

1.1.4 Second affine example The common canonical operator expression, [Q, P ] = i~, 11, directly implies [Q, (P Q + QP )/2] = i~ Q, the basic affine operator!. To confirm the affine expression, multiply i~11= [Q, P ] by Q, which gives i~ Q =(Q2P −P Q2)/2= (Q2P + QP Q − QP Q − P Q2)/2, i.e., i~ Q = [Q, (QP + P Q)/2], which is the

3 basic affine expression, [Q, D] = i~ Q, where D ≡ (P Q + QP )/2. This derivation demands that Q > 0 and/or Q < 0. Canonical quantization implies affine quantization, but adds a limit on coordinates.2 Regarding our problem, now b = 0, and so the classical affine variables are d ≡ pq and q > 0, which lead to the half-harmonic oscillator H′(d, q) = (d2 q−2 + q2)/2. The basic affine quantum operators are D and Q> 0, where (P †Q + QP )/2 ≡ D (= D†) and Q> 0(= Q† > 0). These quantum variables lead to [Q, D]= i~ Q. The half-harmonic oscillator quantum Hamiltonian is given by H′(D, Q)=(DQ−2D +Q2)/2. Schr¨odinger’s representation is given by Q = x> 0 and

D = −i~[x(∂/∂x)+(∂/∂x)x)]/2= −i~[x(∂/∂x)+1/2] . (3)

Finally, Schr¨odinger’s equation is given by

i~(∂ψ(x, t)/∂t = (4) = [−~2(x(∂/∂x)+1/2) x−2 (x(∂/∂x)+1/2)+ x2]/2 ψ(x, t) = [−~2 (∂2/∂x2)+(3/4)~2/x2 + x2]/2 ψ(x, t) .

Solutions of (4) have been provided by L. Gouba [3]. Her solutions for the half-harmonic oscillator contain eigenvalues that are equally spaced as are the eigenvalues of the full-harmonic oscillator, although the spacing itself differs in the two cases. The relevant differential equation in (4) is known as a spiked harmonic oscillator, and its solutions are based on confluent hypergeometric 3/2 functions. It is noteworthy that every eigenfunction ψn(x) ∝ x , when 0 < x ≪ 1, for all n =0, 1, 2, .... Finally, an affine quantization of the half-harmonic oscillator can be con- sidered a correctly solved problem.

1.2 Lessons from canonical and affine quantization procedures An important lesson from the foregoing set of examples is that canonical quantization requires classical variables, i.e., p& q, to be promoted to quan- tum operators must satisfy −∞

2A multiplication by Q2 − 1, instead of Q, could create a finite range of coordinates.

4 Our analysis of gravity will exploit the fact that the metric (the analog 2 a b of q) is required to be strictly positive, i.e., ds(x) = gab(x) dx dx > 0, a provided {dx } ≡/ 0. In addition, the metric itself, gab(x), is dimensionless. These important properties are well adapted to the possibilities offered by affine quantization.

2 A Loop for Einstein’s Gravity

To begin this section the author has chosen to offer the initial paragraph re- garding “” from Wikipedia in order to have a common text provided by proponents of that topic.3

2.1 The initial paragraph from Wikipedia’s “Loop Quantum Gravity” “Beyond the CMS Higgs-.jpg Simulated CMS particle detector data depicting a produced by col- liding protons decaying into hadron jets and electrons Standard Model show Evidence show Theories show hide Quantum gravity False Loop quantum gravity Causal dy- namical triangulation Causal systems Canonical quan- tum gravity Superfluid vacuum theory show Experi- ments vte Loop quantum gravity (LQG)[1][2][3][4][5] is a theory of quantum gravity, which aims to merge and , in- corporating matter of the Standard Model into the framework established for the pure quantum gravity case. As a candidate for quantum gravity, LQG competes with string theory.[6] Loop quantum gravity is an attempt to develop a quantum theory of gravity based directly on Einstein’s geo- metric formulation rather than the treatment of gravity as a force. To do this, in LQG theory space and time are quantized analogously to the way quantities like and are quantized in quantum mechanics. The theory gives a physical picture of where space and time are

3The references within this paragraph are those belonging to Wikipedia and are not references chosen by the author for the rest of this article.

5 granular and discrete directly because of quantization just like in the quantum theory of and the discrete energy levels of atoms. An implication of a quantized space is that a minimum distance ex- ists. LQG postulates that the structure of space is composed of finite loops woven into an extremely fine fabric or network. These networks of loops are called spin networks. The evolution of a spin network, or spin foam, has a scale on the order of a , approximately 10−35 metres, and smaller scales are meaningless. Consequently, not just matter, but space it- self, prefers an atomic structure. The areas of research, which involves about 30 research groups worldwide, [7] share the basic physical assumptions and the mathematical description of quantum space. Research has evolved in two directions: the more traditional canonical loop quantum gravity, and the newer covariant loop quantum gravity, called spin foam theory. The most well-developed theory that has been advanced as a direct result of loop quantum gravity is called loop quantum cosmology (LQC). LQC advances the study of the early universe, incorporating the concept of the Big Bang into the broader theory of the Big Bounce, which envisions the Big Bang as the beginning of a period of expansion that follows a period of contraction, which one could talk of as the Big Crunch.”4

2.2 The author’s view of selected properties of Loop Quantum Gravity The story that follows may not correctly follow the history, but the author offers a reasonable attempt, especially regarding the most important features. The basic elements for classical gravity are the metric gab(x)= gba(x) > 0 and the momentum πcd(x) = πdc(x) [6]. The metric positivity complicates i 9 canonical quantization so it is modified by introducing Ea(x) ∈ IR , with i j a, b, c, .. = i, j, k, .. = 1, 2, 3, and set gab(x) = Ea(x) δij Eb (x) ≥ 0. This is i 9 k not strictly positive because Ea(x) ∈ IR , and Ec (x) = 0 is not excluded as i canonical quantization requires. It is evident that Ea(x) is dimensionless to ensure that the metric is also dimensionless. a i A partner expression Ai (x) is introduced to pair with Ea(x), and these two variables lead to a Poisson bracket, effectively of unity, which is a a key property for canonical quantization. The expression Ai (x) is not di- 4This paragraph is a direct quotation of “Loop Quantum Gravity” from Wikipedis’s contribution to that topic.

6 a mensionless, but, as canonical quantization requires, Ai (x) has the dimen- sion of ~. Indeed, these expressions, when quantized, basically lead to ˆi ˆb ′ ~ i b 3 ′ [Ea(x), Aj(x )] = i δj δa δ (x, x ) 11. While we have begun the quantization study, it is essential to first as- certain whether or not we have found physically correct quantum operators or not. This issue arises because while there are many different classical variables that are acceptable for some expression, e.g., the pair of vari- ables (skipping the usual indices) F (E, A)= F (E, A), but, when quantized, Fˆ(E,ˆ Aˆ) =6 Fˆ (E,ˆ Aˆ). This fact implies, generally, there is only one set of quantum operators that is physically correct, and some rule must decide which are the classical functions that promote to physically correct quantum functions. For canonical quantization, Dirac [1] suggested that physically correct quantum operators are those that were promoted from Cartesian co- ordinates. Dirac proposed this requirement, but did not prove his claim. Recently, the present author [7] has proved that Dirac was correct, namely: using canonical quantization, ‘constant zero curvature’ classical coordinates are those that lead to physically correct quantum operators. i The relevance of this paragraph is since the classical variables Ea(x) and b Aj(x) are not Cartesian coordinates [8], the proposed canonical quantization ˆi does not have a physically correct quantization for the operators Ea(x) and ˆb Aj(x). The present canonical gravity quantization leads to a very tiny, but not zero, discrete behavior for space and time [4]. The author attributes this proposed discrete behavior to a non-physical choice of basic quantum oper- ators. On the other hand, the present affine gravity quantization, as we will show, has no discrete behavior, but instead, has a smooth and continuous behavior of space and time.

2.2.1 Quantum gravity efforts The topic of quantum gravity has gathered numerous suggestions offered in Wikipedia [4], but no mention of affine quantum gravity has appeared there.

7 3 An Affine Quantization of Einstein’s Gravity

3.1 Favored variables for affine quantization

In an affine quantization our favored variables are the metric gab(x), which c ce is dimensionless, and the momentric πd(x) (≡ π (x) gde(x)), which has the of ~.5 The Poisson brackets become

a c a c c a 3 {πb (x), πd(y)} = (1/2) [δd πb (x) − δb πd (x)] δ (x, y) c c c 3 {gab(x), πd(y)} = (1/2)[ δagbd(x)+ δb gad(x)] δ (x, y)

{gab(x),gcd(y)} =0 . (5)

In addition, we can promote these expressions to quantum operators, which then become

a c ~ a c c c 3 [ˆπb (x), πˆd(y)] = i(1/2) [δd πˆb (x) − δb πˆa(x)] δ (x, y) c ~ c c 3 [ˆgab(x), πˆd(y)] = i(1/2) [ δagˆbd(x)+ δbgˆad(x)]δ (x, y)

[ˆgad(x), gˆcd(y)]=0 . (6)

Observe that the middle line of (5) – and thus also that of (6) – are still true if the metric variables change sign. This implies that using affine vari- ables can lead to either strictly positive metrics, which we accept, or it could also lead to strictly negative metrics, which we reject. This important fea- ture does not hold with analogous canonical classical and canonical quantum equations. In the next section we will establish the fact that the classical Poisson brackets in (5) match the quantum commutation relations in (6).

3.2 Affine coherent states for quantum gravity

c The basic affine quantum operators,g ˆab(x) andπ ˆd(x), are fundamental to the required coherent states, which are chosen [5] as

ab 3 c d 3 (i/~) π (x)ˆgab(x) d x −(i/~) η (x)ˆπc (x) d x |π; gi ≡ e R e R d |αi , (7) 5Named for the momentum and the metric.

8 where {η(x)} is a 3×3 symmetric, arbitrary matrix, and {g(x)} ≡ e{η(x)} > 0. c ~ The fiducial vector |αi satisfies the equation [(ˆgab(x) − δab 11) + iπˆd(x)/α(x) ] c |αi = 0, which establishes that hα|gˆab(x)|αi = δab and hα|πˆd(x)|αi = 0. For clarity, we offer a further description for a metric example given by

w1(x) e 0 0 w1(x)0 0  w2(x)    0 e 0 = exp 0 w2(x) 0 w3(x)  0 0 e   0 0 w3(x)  that we denote as W (x) = exp(w(x)). Next, we introduce O(x) and its transpose O(x)T which are arbitrary, 3 × 3, orthogonal matrices such that O(x)T O(x)= O(x) O(x)T ≡ 11. A typical 3 × 3 orthogonal matrix, given as an example, is 10 0   0 cos(θ1(x)) − sin(θ1(x))  0 sin(θ1(x)) cos(θ1(x))  while changing 2 more positions of “1” leads to 3 angle parameters and to add 3 more parameters from W (x), leads to 6 variables. As expected, 6 is the number of parameters in gab(x)= gba(x) > 0.

3.3 The semi-classical gravity expression The Hamiltonian expressions for gravity are the most difficult part of the theory. According to ADM [6], the classical canonical expression is given, cd ce using metric gab(x) and momentum π (x) variables (with π (x)gde(x) = c πd(x)) as

H(π,g)= {g−1/2(x)[πa(x))πb(x) − 1 πa(x)πb(x)] + g1/2(x) R(x)} d3x , (8) R b a 2 a b where g(x) = det[gab(x)], and R(x) is the Ricci scalar term. Indeed, (8) can also be considered to be written directly in classical affine variables, namely c gab(x) and πd(x). In this case we rename the canonical expression, H(π,g), to an affine expression, H′(π,g). It is noteworthy that the momentum aspects of the classical Hamiltonian can be negative as well as positive. To simplify this aspect we adopt 2 × 2 1 2 1 metrics. In that case, the first factors of (8) could become π1 = π2 =0, π2 = 2 a b 1, and π1 = −1. In that case, πb πa = −2. For this example, the second a b factors of (8) πa πb = 0. In addition, the Ricci scalar can be negative as well

9 as positive. These properties carry over to the quantized elements, which could imply, a simple , roughly like H = P1P2 + Q1Q2. The affine quantized version of expression (8) using metric,g ˆab(x), and c momnentric operators,π ˆd(x), is given by

H′(ˆπ, gˆ) (9) = {[ˆπa(x)ˆg−1/2(x)ˆπb (x) − 1 πˆa(x)ˆg−1/2(x)ˆπb(x)]+ˆg1/2(x) Rˆ(x)} d3x . R b a 2 a b Moreover, the connection between the classical and quantum terms arises from

H′(π,g)= hπ; g|H′(ˆπ, gˆ)|π; gi (10) ′ a ac η/2 e η/2 d η/2 a η/2 b = hα|H (ˆπb + π [e ]cgˆed[e ]b , [e ]c gˆab[e ]d)]|αi , ′ a ac η/2 e η/2 d η/2 a η/2 b = H (hα|πˆb |αi + π [e ]chα|gˆed|αi[e ]b , [e ]c hα|gˆab|αi[e ]d) + O(~; π,g) , which becomes

′ a ′ a ~ H (πb ,gcd)= H (πb ,gcd)+ O( ; π,g) (11) leading to the usual classical limit as ~ → 0, or when O is tiny enough to ignore. This expression confirms the equality of (5) and (6), as to the similarity of classical and quantum expressions.

4 Schr¨odinger’s Representation and Equation

In the previous section we developed the quantum Hamiltonian in (9), and now we develop a specific formulation wherein the dimensionless metric oper- ator is given byg ˆab(x) → gab(x) and the dimensionless metric determinant is given byg ˆ(x) ≡ det[ˆgab(x)] → g(x) ≡ det[gab(x)], as well as the momentric, which has the of ~, and offers a formal expression given by

c ~ πˆd(x) → −i(1/2) [gde(x)(∂/∂gce(x))+(∂/∂gce(x))gde(x)] . (12)

It is noteworthy that

a −1/2 πˆb (x) g(x) =0 , (13)

10 a 1 a which is obtained byπ ˆb F (g) = 0, and implies that [gbc (∂/∂gac)+ 2 δb ]F (g)= ac 1 a 0. Next, we find gbc g g dF (g)/dg + 2 δb F (g) = 0, which requires that 1 −1/2 gdF (g)/dg + 2 F (g) = 0; hence F (g) ∝ g . These expressions permit us to offer Schr¨odinger’s equation, which is

i~(∂ Ψ(x, t)/∂t) = {{ [[ˆπa(x, t)g(x, t)−1/2πˆb(x, t) − (1/2)ˆπa(x, t)g(x, t)−1/2πˆb(x, t) R b a a b +g(x, t)1/2 R(x, t)]] d3x}} Ψ(x, t) . (14)

4.1 Possible solutions to Schr¨oedinger equation

a −1/2 The relation in (13) also leads toπ ˆb (x) Πyg(y) = 0 for any and all y. This property offers eigenfunctions having the general form Ψ(x) = −1/2 V (x) Πyg(y) ; this expression is suggested here because in selected ultralo- cal procedures [9], such behavior is required. The positivity and negativity of the Hamiltonian operator points to continuum solutions rather than dis- crete levels of eigenfunctions. The Hamiltonian is really a constraint that requires solutions of the form H′(ˆπ(x), gˆ(x))Ψ(x) = 0. In particular, the constraint expressions can also be correctly extracted using a suitable func- tional integration [10]. This procedure points toward the physical features that are the most difficult with the Hamiltonian operator, and simpler issues can complete the story which have been examined before in [11].

References

[1] P.A.M. Dirac, The Principles of Quantum Mechanics, (Claredon Press, Oxford, 1958), page 114, in a footnote.

[2] “quantum harmonic oscillator”; Wikipedia.

[3] L. Gouba, “Affine Quantization on the Half Line”, Journal of High Energy Physics, Gravitation and Cosmology, 7, 352-365 (2021); DOI:10.4236/jhepgc.2021.71019.

[4] “Quantum Gravity”; Wikipedia.

[5] J.R. Klauder, “Using Affine Quantization to Analyze Non- Renormalizable Scalar Fields and the Quantization of Einstein’s

11 Gravity“, Journal of High Energy Physics, Gravitation and Cosmology 6, 802-816 (2020); DOI:10.4236/jhepgc.2020.64053.

[6] R. Arnowitt, S. Deser, and C. Misner, “The Dynamics of General Rela- tivity”, Gravitation: An Introduction to Current Research, Ed. L. Wit- ten, (Wiley & Sons, New York, 1962), p. 227; arXiv:gr-qc/0405109.

[7] J.R. Klauder, “Using Coherent States to Make Physically Correct Classical-to-Quantum Procedures that Help Resolve Nonrenomalizable Fields Including Einstein’s Gravity”; arXiv:2105.03206.

[8] J.R. Klauder, “Is Loop Quantum Gravity a Physically Correct Quanti- zation?”: Journal of High Energy Physics, Gravitation and Cosmology, 6, 49-51 (2020); DOI:10.4236/jhepgc.2020.61006.

[9] J.R. Klauder, “An Ultralocal Classical and Quantum Gravity Theory”, Journal of High Energy Physics, Gravitation and Cosmology, 6, 656-662 (2020): DOI:10.4236/jhepgc.2020.64044.

[10] J.R. Klauder, “Universal Procedure for Enforcing Quantum Con- straints”, Nuclear Phys. B 547, 397-412 (1999).

[11] J.R. Klauder, “Using Affine Quantization to Analyze Non- renormalizable Scalar Fields and the Quantization of Einstein’s Gravity”, Journal of High Energy Physics, Gravitation and Cosmology, 6, 802-816 (2020); DOI:10.4236/jhepgc.2020.64053.

12