Hamiltonian constraint

Last updated

The Hamiltonian constraint arises from any theory that admits a Hamiltonian formulation and is reparametrisation-invariant. The Hamiltonian constraint of general relativity is an important non-trivial example.

Contents

In the context of general relativity, the Hamiltonian constraint technically refers to a linear combination of spatial and time diffeomorphism constraints reflecting the reparametrizability of the theory under both spatial as well as time coordinates. However, most of the time the term Hamiltonian constraint is reserved for the constraint that generates time diffeomorphisms.

Simplest example: the parametrized clock and pendulum system

Parametrization

In its usual presentation, classical mechanics appears to give time a special role as an independent variable. This is unnecessary, however. Mechanics can be formulated to treat the time variable on the same footing as the other variables in an extended phase space, by parameterizing the temporal variable(s) in terms of a common, albeit unspecified parameter variable. Phase space variables being on the same footing.

We introduce
t
{\displaystyle \tau }
as an unphysical parameter labeling different possible correlations between the time reading
t
{\displaystyle t}
of the clock and the elongation
x
{\displaystyle x}
of the pendulum.
t
{\displaystyle \tau }
is unphysical parameter and there are many different choices for it. Parametrized Harmonic Oscillator.jpg
We introduce as an unphysical parameter labeling different possible correlations between the time reading of the clock and the elongation of the pendulum. is unphysical parameter and there are many different choices for it.

Say our system comprised a pendulum executing a simple harmonic motion and a clock. Whereas the system could be described classically by a position x=x(t), with x defined as a function of time, it is also possible to describe the same system as x() and t() where the relation between x and t is not directly specified. Instead, x and t are determined by the parameter , which is simply a parameter of the system, possibly having no objective meaning in its own right.

The system would be described by the position of a pendulum from the center, denoted , and the reading on the clock, denoted . We put these variables on the same footing by introducing a fictitious parameter ,

whose 'evolution' with respect to takes us continuously through every possible correlation between the displacement and reading on the clock. Obviously the variable can be replaced by any monotonic function, . This is what makes the system reparametrisation-invariant. Note that by this reparametrisation-invariance the theory cannot predict the value of or for a given value of but only the relationship between these quantities. Dynamics is then determined by this relationship.

Dynamics of this reparametrization-invariant system

The action for the parametrized Harmonic oscillator is then

where and are canonical coordinates and and are their conjugate momenta respectively and represent our extended phase space (we will show that we can recover the usual Newton's equations from this expression). Writing the action as

we identify the as

Hamilton's equations for are

which gives a constraint,

is our Hamiltonian constraint! It could also be obtained from the Euler–Lagrange equation of motion, noting that the action depends on but not its derivative. Then the extended phase space variables , , , and are constrained to take values on this constraint-hypersurface of the extended phase space. We refer to as the `smeared' Hamiltonian constraint where is an arbitrary number. The 'smeared' Hamiltonian constraint tells us how an extended phase space variable (or function thereof) evolves with respect to :

(these are actually the other Hamilton's equations). These equations describe a flow or orbit in phase space. In general we have

for any phase space function . As the Hamiltonian constraint Poisson commutes with itself, it preserves itself and hence the constraint-hypersurface. The possible correlations between measurable quantities like and then correspond to `orbits' generated by the constraint within the constraint surface, each particular orbit differentiated from each other by say also measuring the value of say along with and at one -instant; after determining the particular orbit, for each measurement of we can predict the value will take.

Deparametrization

The other equations of Hamiltonian mechanics are

Upon substitution of our action these give,

These represent the fundamental equations governing our system.

In the case of the parametrized clock and pendulum system we can of course recover the usual equations of motion in which is the independent variable:

Now and can be deduced by

We recover the usual differential equation for the simple harmonic oscillator,

We also have or

Our Hamiltonian constraint is then easily seen as the condition of constancy of energy! Deparametrization and the identification of a time variable with respect to which everything evolves is the opposite process of parametrization. It turns out in general that not all reparametrisation-invariant systems can be deparametrized. General relativity being a prime physical example (here the spacetime coordinates correspond to the unphysical and the Hamiltonian is a linear combination of constraints which generate spatial and time diffeomorphisms).

Reason why we could deparametrize here

The underlining reason why we could deparametrize (aside from the fact that we already know it was an artificial reparametrization in the first place) is the mathematical form of the constraint, namely,

Substitute the Hamiltonian constraint into the original action we obtain

which is the standard action for the harmonic oscillator. General relativity is an example of a physical theory where the Hamiltonian constraint isn't of the above mathematical form in general, and so cannot be deparametrized in general.

Hamiltonian of classical general relativity

In the ADM formulation of general relativity one splits spacetime into spatial slices and time, the basic variables are taken to be the induced metric, , on the spatial slice (the metric induced on the spatial slice by the spacetime metric), and its conjugate momentum variable related to the extrinsic curvature, , (this tells us how the spatial slice curves with respect to spacetime and is a measure of how the induced metric evolves in time). [1] These are the metric canonical coordinates.

Dynamics such as time-evolutions of fields are controlled by the Hamiltonian constraint.

The identity of the Hamiltonian constraint is a major open question in quantum gravity, as is extracting of physical observables from any such specific constraint.

In 1986 Abhay Ashtekar introduced a new set of canonical variables, Ashtekar variables to represent an unusual way of rewriting the metric canonical variables on the three-dimensional spatial slices in terms of a SU(2) gauge field and its complementary variable. [2] The Hamiltonian was much simplified in this reformulation. This led to the loop representation of quantum general relativity and in turn loop quantum gravity. [3]

Within the loop quantum gravity representation Thiemann formulated a mathematically rigorous operator as a proposal as such a constraint. [4] Although this operator defines a complete and consistent quantum theory, doubts have been raised[ by whom? ] as to the physical reality of this theory due to inconsistencies with classical general relativity (the quantum constraint algebra closes, but it is not isomorphic to the classical constraint algebra of GR, which is seen as circumstantial evidence of inconsistencies definitely not a proof of inconsistencies), and so variants have been proposed.

Metric formulation

The idea was to quantize the canonical variables and , making them into operators acting on wavefunctions on the space of 3-metrics, and then to quantize the Hamiltonian (and other constraints). However, this program soon became regarded as dauntingly difficult for various reasons, one being the non-polynomial nature of the Hamiltonian constraint:

where is the scalar curvature of the three metric . Being a non-polynomial expression in the canonical variables and their derivatives it is very difficult to promote to a quantum operator.

Expression using Ashtekar variables

The configuration variables of Ashtekar's variables behave like an gauge field or connection . Its canonically conjugate momentum is is the densitized "electric" field or triad (densitized as ). What do these variables have to do with gravity? The densitized triads can be used to reconstruct the spatial metric via

The densitized triads are not unique, and in fact one can perform a local in space rotation with respect to the internal indices . This is actually the origin of the gauge invariance. The connection can be used to reconstruct the extrinsic curvature. The relation is given by

where is related to the spin connection, , by and .

In terms of Ashtekar variables the classical expression of the constraint is given by,

where field strength tensor of the gauge field . Due to the factor this is non-polynomial in the Ashtekar's variables. Since we impose the condition

we could consider the densitized Hamiltonian instead,

This Hamiltonian is now polynomial the Ashtekar's variables. This development raised new hopes for the canonical quantum gravity programme. [5] Although Ashtekar variables had the virtue of simplifying the Hamiltonian, it has the problem that the variables become complex. When one quantizes the theory it is a difficult task ensure that one recovers real general relativity as opposed to complex general relativity. Also there were also serious difficulties promoting the densitized Hamiltonian to a quantum operator.

A way of addressing the problem of reality conditions was noting that if we took the signature to be , that is Euclidean instead of Lorentzian, then one can retain the simple form of the Hamiltonian for but for real variables. One can then define what is called a generalized Wick rotation to recover the Lorentzian theory. [6] Generalized as it is a Wick transformation in phase space and has nothing to do with analytical continuation of the time parameter .

Expression for real formulation of Ashtekar variables

Thomas Thiemann addressed both the above problems. [4] He used the real connection

In real Ashtekar variables the full Hamiltonian is

where the constant is the Barbero–Immirzi parameter. [7] The constant is -1 for Lorentzian signature and +1 for Euclidean signature. The have a complicated relationship with the densitized triads and causes serious problems upon quantization. Ashtekar variables can be seen as choosing to make the second more complicated term was made to vanish (the first term is denoted because for the Euclidean theory this term remains for the real choice of ). Also we still have the problem of the factor.

Thiemann was able to make it work for real . First he could simplify the troublesome by using the identity

where is the volume,

The first term of the Hamiltonian constraint becomes

upon using Thiemann's identity. This Poisson bracket is replaced by a commutator upon quantization. It turns out that a similar trick can be used to teat the second term. Why are the given by the densitized triads ? It actually come about from the Gauss Law

We can solve this in much the same way as the Levi-Civita connection can be calculated from the equation ; by rotating the various indices and then adding and subtracting them. The result is complicated and non-linear. To circumvent the problems introduced by this complicated relationship Thiemann first defines the Gauss gauge invariant quantity

where , and notes that

We are then able to write

and as such find an expression in terms of the configuration variable and . We obtain for the second term of the Hamiltonian

Why is it easier to quantize ? This is because it can be rewritten in terms of quantities that we already know how to quantize. Specifically can be rewritten as

where we have used that the integrated densitized trace of the extrinsic curvature is the "time derivative of the volume".

Related Research Articles

<span class="mw-page-title-main">Loop quantum gravity</span> Theory of quantum gravity, merging quantum mechanics and general relativity

Loop quantum gravity (LQG) is a theory of quantum gravity, which aims to merge quantum mechanics and general relativity, incorporating matter of the Standard Model into the framework established for the pure quantum gravity case. It is an attempt to develop a quantum theory of gravity based directly on Einstein's geometric formulation rather than the treatment of gravity as a force. As a theory LQG postulates that the structure of space and time 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 above the order of a Planck length, approximately 10−35 meters, and smaller scales are meaningless. Consequently, not just matter, but space itself, prefers an atomic structure. The areas of research, which involve about 30 research groups worldwide, 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.

In physics, a Langevin equation is a stochastic differential equation describing how a system evolves when subjected to a combination of deterministic and fluctuating ("random") forces. The dependent variables in a Langevin equation typically are collective (macroscopic) variables changing only slowly in comparison to the other (microscopic) variables of the system. The fast (microscopic) variables are responsible for the stochastic nature of the Langevin equation. One application is to Brownian motion, which models the fluctuating motion of a small particle in a fluid.

<span class="mw-page-title-main">Weierstrass elliptic function</span> Class of mathematical functions

In mathematics, the Weierstrass elliptic functions are elliptic functions that take a particularly simple form. They are named for Karl Weierstrass. This class of functions are also referred to as ℘-functions and they are usually denoted by the symbol ℘, a uniquely fancy script p. They play an important role in the theory of elliptic functions. A ℘-function together with its derivative can be used to parameterize elliptic curves and they generate the field of elliptic functions with respect to a given period lattice.

In the ADM formulation of general relativity, spacetime is split into spatial slices and a time axis. The basic variables are taken to be the induced metric on the spatial slice and the metric's conjugate momentum , which is related to the extrinsic curvature and is a measure of how the induced metric evolves in time. These are the metric canonical coordinates.

A first class constraint is a dynamical quantity in a constrained Hamiltonian system whose Poisson bracket with all the other constraints vanishes on the constraint surface in phase space. To calculate the first class constraint, one assumes that there are no second class constraints, or that they have been calculated previously, and their Dirac brackets generated.

In physics, the Polyakov action is an action of the two-dimensional conformal field theory describing the worldsheet of a string in string theory. It was introduced by Stanley Deser and Bruno Zumino and independently by L. Brink, P. Di Vecchia and P. S. Howe in 1976, and has become associated with Alexander Polyakov after he made use of it in quantizing the string in 1981. The action reads:

Variational Bayesian methods are a family of techniques for approximating intractable integrals arising in Bayesian inference and machine learning. They are typically used in complex statistical models consisting of observed variables as well as unknown parameters and latent variables, with various sorts of relationships among the three types of random variables, as might be described by a graphical model. As typical in Bayesian inference, the parameters and latent variables are grouped together as "unobserved variables". Variational Bayesian methods are primarily used for two purposes:

  1. To provide an analytical approximation to the posterior probability of the unobserved variables, in order to do statistical inference over these variables.
  2. To derive a lower bound for the marginal likelihood of the observed data. This is typically used for performing model selection, the general idea being that a higher marginal likelihood for a given model indicates a better fit of the data by that model and hence a greater probability that the model in question was the one that generated the data.
<span class="mw-page-title-main">Linear time-invariant system</span> Mathematical model which is both linear and time-invariant

In system analysis, among other fields of study, a linear time-invariant (LTI) system is a system that produces an output signal from any input signal subject to the constraints of linearity and time-invariance; these terms are briefly defined below. These properties apply (exactly or approximately) to many important physical systems, in which case the response y(t) of the system to an arbitrary input x(t) can be found directly using convolution: y(t) = (xh)(t) where h(t) is called the system's impulse response and ∗ represents convolution (not to be confused with multiplication). What's more, there are systematic methods for solving any such system (determining h(t)), whereas systems not meeting both properties are generally more difficult (or impossible) to solve analytically. A good example of an LTI system is any electrical circuit consisting of resistors, capacitors, inductors and linear amplifiers.

Nondimensionalization is the partial or full removal of physical dimensions from an equation involving physical quantities by a suitable substitution of variables. This technique can simplify and parameterize problems where measured units are involved. It is closely related to dimensional analysis. In some physical systems, the term scaling is used interchangeably with nondimensionalization, in order to suggest that certain quantities are better measured relative to some appropriate unit. These units refer to quantities intrinsic to the system, rather than units such as SI units. Nondimensionalization is not the same as converting extensive quantities in an equation to intensive quantities, since the latter procedure results in variables that still carry units.

In general relativity, a geodesic generalizes the notion of a "straight line" to curved spacetime. Importantly, the world line of a particle free from all external, non-gravitational forces is a particular type of geodesic. In other words, a freely moving or falling particle always moves along a geodesic.

<span class="mw-page-title-main">Complex torus</span>

In mathematics, a complex torus is a particular kind of complex manifold M whose underlying smooth manifold is a torus in the usual sense. Here N must be the even number 2n, where n is the complex dimension of M.

<span class="mw-page-title-main">Canonical quantum gravity</span> A formulation of general relativity

In physics, canonical quantum gravity is an attempt to quantize the canonical formulation of general relativity. It is a Hamiltonian formulation of Einstein's general theory of relativity. The basic theory was outlined by Bryce DeWitt in a seminal 1967 paper, and based on earlier work by Peter G. Bergmann using the so-called canonical quantization techniques for constrained Hamiltonian systems invented by Paul Dirac. Dirac's approach allows the quantization of systems that include gauge symmetries using Hamiltonian techniques in a fixed gauge choice. Newer approaches based in part on the work of DeWitt and Dirac include the Hartle–Hawking state, Regge calculus, the Wheeler–DeWitt equation and loop quantum gravity.

Linear dynamical systems are dynamical systems whose evolution functions are linear. While dynamical systems, in general, do not have closed-form solutions, linear dynamical systems can be solved exactly, and they have a rich set of mathematical properties. Linear systems can also be used to understand the qualitative behavior of general dynamical systems, by calculating the equilibrium points of the system and approximating it as a linear system around each such point.

In mathematics, Reidemeister torsion is a topological invariant of manifolds introduced by Kurt Reidemeister for 3-manifolds and generalized to higher dimensions by Wolfgang Franz (1935) and Georges de Rham (1936). Analytic torsion is an invariant of Riemannian manifolds defined by Daniel B. Ray and Isadore M. Singer as an analytic analogue of Reidemeister torsion. Jeff Cheeger and Werner Müller (1978) proved Ray and Singer's conjecture that Reidemeister torsion and analytic torsion are the same for compact Riemannian manifolds.

In mathematics, particularly linear algebra, the Schur–Horn theorem, named after Issai Schur and Alfred Horn, characterizes the diagonal of a Hermitian matrix with given eigenvalues. It has inspired investigations and substantial generalizations in the setting of symplectic geometry. A few important generalizations are Kostant's convexity theorem, Atiyah–Guillemin–Sternberg convexity theorem, Kirwan convexity theorem.

An affine term structure model is a financial model that relates zero-coupon bond prices to a spot rate model. It is particularly useful for deriving the yield curve – the process of determining spot rate model inputs from observable bond market data. The affine class of term structure models implies the convenient form that log bond prices are linear functions of the spot rate.

In the ADM formulation of general relativity one splits spacetime into spatial slices and time, the basic variables are taken to be the induced metric, , on the spatial slice, and its conjugate momentum variable related to the extrinsic curvature, ,. These are the metric canonical coordinates.

<span class="mw-page-title-main">Loop representation in gauge theories and quantum gravity</span> Description of gauge theories using loop operators

Attempts have been made to describe gauge theories in terms of extended objects such as Wilson loops and holonomies. The loop representation is a quantum hamiltonian representation of gauge theories in terms of loops. The aim of the loop representation in the context of Yang–Mills theories is to avoid the redundancy introduced by Gauss gauge symmetries allowing to work directly in the space of physical states. The idea is well known in the context of lattice Yang–Mills theory. Attempts to explore the continuous loop representation was made by Gambini and Trias for canonical Yang–Mills theory, however there were difficulties as they represented singular objects. As we shall see the loop formalism goes far beyond a simple gauge invariant description, in fact it is the natural geometrical framework to treat gauge theories and quantum gravity in terms of their fundamental physical excitations.

In quantum mechanics, a Dirac membrane is a model of a charged membrane introduced by Paul Dirac in 1962. Dirac's original motivation was to explain the mass of the muon as an excitation of the ground state corresponding to an electron. Anticipating the birth of string theory by almost a decade, he was the first to introduce what is now called a type of Nambu–Goto action for membranes.

Tau functions are an important ingredient in the modern mathematical theory of integrable systems, and have numerous applications in a variety of other domains. They were originally introduced by Ryogo Hirota in his direct method approach to soliton equations, based on expressing them in an equivalent bilinear form.

References

  1. Misner, Charles W.; Thorne, Kip S.; Wheeler, John Archibald. Gravitation . New York: W. H. Freeman and company.
  2. Ashtekar, Abhay (1986-11-03). "New Variables for Classical and Quantum Gravity". Physical Review Letters. American Physical Society (APS). 57 (18): 2244–2247. doi:10.1103/physrevlett.57.2244. ISSN   0031-9007.
  3. Rovelli, Carlo; Smolin, Lee (1988-09-05). "Knot Theory and Quantum Gravity". Physical Review Letters. American Physical Society (APS). 61 (10): 1155–1158. doi:10.1103/physrevlett.61.1155. ISSN   0031-9007.
  4. 1 2 Thiemann, T. (1996). "Anomaly-free formulation of non-perturbative, four-dimensional Lorentzian quantum gravity". Physics Letters B. Elsevier BV. 380 (3–4): 257–264. arXiv: gr-qc/9606088 . doi:10.1016/0370-2693(96)00532-1. ISSN   0370-2693.
  5. See the book Lectures on Non-Perturbative Canonical Gravity for more details on this and the subsequent development. First published in 1991. World Scientific Publishing Co. Pte. LtD.
  6. Thiemann, T (1996-06-01). "Reality conditions inducing transforms for quantum gauge field theory and quantum gravity". Classical and Quantum Gravity. IOP Publishing. 13 (6): 1383–1403. arXiv: gr-qc/9511057 . doi:10.1088/0264-9381/13/6/012. ISSN   0264-9381.
  7. Barbero G., J. Fernando (1995-05-15). "Real Ashtekar variables for Lorentzian signature space-times". Physical Review D. American Physical Society (APS). 51 (10): 5507–5510. arXiv: gr-qc/9410014 . doi:10.1103/physrevd.51.5507. ISSN   0556-2821.