Feynman Formulas As A Method of Averaging Random Hamiltonians
Feynman Formulas As A Method of Averaging Random Hamiltonians
Feynman Formulas As A Method of Averaging Random Hamiltonians
c Pleiades Publishing, Ltd., 2014.
Original Russian Text c Yu.N. Orlov, V.Zh. Sakbaev, O.G. Smolyanov, 2014, published in Trudy Matematicheskogo Instituta imeni
V.A. Steklova, 2014, Vol. 285, pp. 232–243.
Abstract—We propose a method for finding the mathematical expectation of random un-
bounded operators in a Hilbert space. The method is based on averaging random one-parameter
semigroups by means of the Feynman–Chernoff formula. We also consider an application of
this method to the description of various operations that assign quantum Hamiltonians to the
classical Hamilton functions.
DOI: 10.1134/S0081543814040154
INTRODUCTION
Feynman formulas give a representation of the Schrödinger semigroup exp(−tH), t ≥ 0, or
Schrödinger group exp(itH), t ∈ R, in terms of the limits of finitely multiple integrals over the
Cartesian powers of the configuration or phase space (as the multiplicity tends to infinity) of a
classical Hamiltonian system whose quantization yields the Hamilton operator H (this means that
H is a self-adjoint operator corresponding to the Hamilton function H of the classical system; in
particular, H may be a pseudodifferential operator whose symbol is the Hamilton function H).
When one uses integrals over the products of the configuration space, one speaks of Lagrangian
Feynman formulas; and when integrals over the products of the phase space are used, one speaks
of Hamiltonian Feynman formulas.
In this paper, we introduce a new method of averaging random one-parameter semigroups and
their generators by means of Feynman formulas with the use of Chernoff’s theorem [9]. By a random
semigroup we will mean a vector-valued random variable G with values in the set of one-parameter
semigroups of operators acting in some Banach space. If (E, A, μ) is a probability space on which
a random semigroup G is defined, then the generator of the random semigroup G is a random
variable HG on the same probability space that is defined by the following condition: for every
ε ∈ E, the value HG (ε) of the random variable HG is a generator of the semigroup G(ε). In
this paper, we show that the mathematical expectation of a random semigroup G is an operator-
valued function FG that is Chernoff equivalent (see [11]) to some semigroup UG . It is natural to
think of the generator of UG as the mathematical expectation of the random generator HG of the
random semigroup G. Chernoff equivalence means that the sequence of operator-valued functions
{(FG (t/n))n , t ≥ 0} converges in the strong operator topology to the semigroup UG (t), t ≥ 0,
uniformly on any interval [0, T ] of the half-axis t ≥ 0.
This definition of averaging of generators is an extension of the averaging procedure in the
space of operators, because when all values of a random generator are bounded, its mean value
a Keldysh Institute of Applied Mathematics, Russian Academy of Sciences, Miusskaya pl. 4, Moscow, 125047 Russia.
b Moscow Institute of Physics and Technology (State University), Institutskii per. 9, Dolgoprudnyi, Moscow oblast,
141700 Russia.
c Faculty of Mechanics and Mathematics, Moscow State University, Moscow, 119991 Russia.
222
FEYNMAN FORMULAS AS A METHOD OF AVERAGING 223
coincides with the ordinary mean value of elements of a Banach space (see Remark 2 to Theorem 2
below). When the values of a random generator are unbounded operators, the natural notion of the
mathematical expectation of the random variable HG may not be well defined, but the definition
introduced here is applicable (see Examples 1 and 2).
The proposed method of averaging random semigroup generators that are unbounded operators
is based on the study of averaging the exponential function of unbounded operators that takes values
in the space of bounded linear operators. The method is analogous to the method of summation
of unbounded operators by considering the sums of their quadratic forms (see [3]). In both cases,
instead of averaged operators themselves, one analyzes the values of some operator-valued function
on these operators. Of course, the method proposed in this paper is much more complicated due
to the application of Chernoff iterations and the notion of Chernoff equivalence of operator-valued
functions, which was introduced in [11].
One of the reasons for ambiguity in the quantum dynamics of a system (see [2]) is the am-
biguity in the choice of the classical Hamiltonian corresponding to a given law of motion of the
system in the coordinate space. For every law of motion of a classical system in the coordinate
space, there is a set of Hamilton functions to which this law of motion in the coordinate space
corresponds (Hamilton functions that define the same law of motion in the coordinate space are
called equivalent in [2]). Every classical Hamiltonian can be assigned a quantum Hamiltonian, but
the quantum Hamiltonians corresponding to equivalent (in the above sense) classical Hamiltonians
are different in the sense that the unitary transformation groups generated by these Hamiltonians
are different (see [2]). Now if some probability measure is defined on the set of equivalent classical
Hamiltonians (and hence on the set of the corresponding quantum Hamiltonians), then random
variables with values in the set of classical (quantum) Hamiltonians and with values in the set
of unitary groups arise. Another (possibly, the main) reason for the ambiguity in quantization is
the noncommutativity of the coordinate and momentum operators (see [1]). The relation between
this ambiguity and the averaging operation introduced above will be discussed in greater detail
below.
Chernoff’s theorem is one of the most general statements on the approximation of one-parameter
semigroups by operator-valued functions; it can be regarded as a far-reaching generalization of the
Euler theorem on the approximation of a solution to a differential equation by polygonal chains to
the infinite-dimensional case. In particular, this theorem is used for constructing approximations
of semigroups by Feynman formulas and approximations of equations with variable coefficients by
equations with piecewise constant coefficients.
One of the first applications of Chernoff’s theorem is (see [10, 6]) the derivation of the rep-
resentation of the unitary group generated by a Hamiltonian by means of the Feynman formula.
Namely, every classical Hamiltonian H is assigned a quantum Hamiltonian H = H, which is a
pseudodifferential operator in the Hilbert space H = L2 (R) with symbol H. In this case, the
t ∈ R, which
numerical function exp(itH), t ∈ R, is assigned an operator-valued function exp(itH),
is not a semigroup but, according to Chernoff’s theorem, is equivalent in the sense of the definition
from [11] to the unitary group exp(itH), t ∈ R.
In the present study, we apply a similar procedure in order to construct the mean value of a
random Hamiltonian. We assign the mean value in the sense defined above to every probability
measure on the set of quantum Hamiltonians corresponding to the same classical Hamilton function.
Thus, we define another quantization method, i.e., a method for associating an operator in the
Hilbert space with a classical Hamilton function. A similar situation occurs when the same Hamilton
function is considered as the τ -symbol (see [4]) of a family of pseudodifferential operators indexed
by the variable τ from the interval [0, 1]. In our case, the role of the parameter τ is played by a
measure on a set of operators in a Hilbert space.
Here the following question remains unstudied and is worth examining: What is the relationship
t ∈ R, the mean value of the ran-
between the mean value of the random operator function exp(itH),
dom group exp(itH), t ∈ R, and the operator function whose symbol is the mean value of the random
function exp(itH), t ∈ R (we conjecture that all three operator functions are Chernoff equivalent).
1. MAIN DEFINITIONS
A random variable is a measurable mapping ξ from a measure space (E, A, μ) to some Banach
space or to a linear topological space Z. In this context we will consider random semigroups; in
this case, the space Z is the Banach space Cs (R+ , B(X)) of strongly continuous mappings from the
half-axis R+ to the Banach space of linear bounded operators in a Banach space X, and the values
of the random variable ξ are one-parameter semigroups.
In particular, we consider the case when the algebra A is the maximal algebra 2E of all subsets of
the space E. If the measurable space E is a Hausdorff topological space and ε0 ∈ E, then we denote
by W (E, ε0 ) the class of all nonnegative normed finitely additive measures on the measurable space
(E, 2E ) that are concentrated in an arbitrary punctured neighborhood of the point ε0 in the sense
that the value of any measure in the class W (E, ε0 ) is equal to one on any punctured neighborhood
of the point ε0 (see [8, 7, 13]). We establish a relation between generalized limits for ε → ε0 of the
random semigroup G as a mapping G : E → Cs (R+ , B(X)) and the mathematical expectations of
the random semigroup G on the probability space (E, 2E , μ) with measures μ in the class W (E, ε0 ).
Random semigroups arose in [8] as a result of regularization that consists in approximating a
maximal symmetric operator which is not a generator of a semigroup by sequences of generators.
The ambiguity in the choice of approximations and the absence of convergence of the sequence of
regularized semigroups lead in this case to the necessity of studying random semigroups.
The source of randomness of the semigroups related to a quantum system is the ambiguity in
the choice of the classical Hamiltonian describing a family of trajectories in the classical coordinate
space (see [2]). By a classical Hamiltonian we mean a continuously differentiable function H on
the phase space Q × P of a system of classical mechanics. A classical Hamiltonian is said to be
nondegenerate if the function H is twice continuously differentiable and its Hessian is uniformly
bounded away from zero.
We say that two classical Hamiltonians are equivalent if the sets of trajectories corresponding
to these Hamiltonians in the coordinate space (i.e., the projections of the trajectories in the phase
space Q × P onto the coordinate space Q) coincide. In [2], it was shown that the class of equivalent
nondegenerate Hamiltonians is fairly wide and that any two representatives of a class of equivalent
Hamiltonians generate identical families of coordinate trajectories but different families of phase
trajectories. Moreover, different equivalent Hamiltonians generate different quantum dynamical
semigroups, i.e., different groups of unitary transformations of the Hilbert space H = L2 (Q).
Then the quantization of classical systems as a mapping from the set of classical Hamiltonians
to the set of quantum Hamiltonians (self-adjoint operators in the Hilbert space L2 (Q)) and then
to the set of unitary groups generated by quantum Hamiltonians is a random semigroup provided
that the set K of equivalent classical Hamiltonians with the algebra 2K of all subsets is equipped
with some nonnegative normed measure.
As the set E, we can take the set of all generators of strongly continuous semigroups G(X)
acting in the space X and equip this set with the topology τ of strong (or weak) graph convergence.
Let ε0 ∈ G(X) be a limit point of the set G(X) in the topology τ . Denote by W (E, ε0 ) the set of all
measures μ ∈ ba(E, 2E ) (i.e., elements of the Banach space of finitely additive measures of bounded
variation defined on the algebra 2E ) that are nonnegative, normed to unity, and concentrated in an
arbitrary punctured neighborhood of the point ε0 of the set E.
On the space Z, we define a family of functionals that act on an arbitrary element z ∈ Z
according to the rule ϕt,A,g (z) = z(t)A, g, t ∈ R+ , A ∈ X, g ∈ X∗ . Consider the topology τm
on Z generated by the family of functionals ϕt,A,g , t ∈ R+ , A ∈ X, g ∈ X∗ . Then the space Z
equipped with the structure of the algebra of Borel subsets is a measurable space, and the mapping
ξ : E → Z is a random variable.
By the mathematical expectation of a random variable ξ as a mapping from the measure space
(E, 2E , μ) to the topological space Z, we will mean the Pettis integral
M ξ = ξε dμ(ε),
E
holds for all t ∈ R+ , A ∈ X, and g ∈ X∗ . Since the latter integral is the Radon integral of a
numerical function with respect to a finitely additive measure, it is well defined and equality (2.1)
defines a linear bounded transformation M ξ(t) ∈ B(X) for any t ∈ R+ .
Theorem 1. Let μ be a nonnegative normed measure on the algebra 2E of subsets of the set E.
If a family of mappings ξ is uniformly bounded and there exists a linear dense subspace D in X such
that the family of mappings ξε (t)A ∈ C(R+ , X), ε ∈ E, is weakly (strongly) uniformly Lipschitz
continuous for every A ∈ D, then M ξ(t) ∈ Cw (R+ , B(X)) (M ξ(t) ∈ Cs (R+ , B(X))).
Proof. The weak uniform Lipschitz continuity means that for any A ∈ D and g ∈ X∗ there
exists a constant L > 0 such that
the strong Lipschitz continuity means that for any A ∈ D there exists a constant L > 0 such that
The uniform boundedness of a random semigroup ξ means that supε∈E, t∈R+ ξε (t)B(X) ≤ C for
some C > 0.
Therefore, by the uniform boundedness condition, for any t ≥ 0, A ∈ D, and g ∈ X∗ the
function ξε (t)A, g is bounded on the set E and, hence, integrable with respect to measure μ in the
Radon sense (see [8]), and the integral (2.1), as a function of the argument g,
is a linear continuous
functional on the space X∗ . Therefore, for any A ∈ D, the Pettis integral G(X) ξε A dμ(ε) ∈ X is
defined, and the mapping A → G(X) ξε A dμ(ε) is linear in A in view of the linearity of the Pettis
integral
and continuous in view of the uniform boundedness of the mapping ξ. Hence, the mapping
A → G(X) ξε A dμ(ε) can be extended by continuity from the linear subspace D to a linear bounded
transformation of the space X.
The equivalence relation introduced covers a fairly wide class Π of operator-valued functions.
It is symmetric, reflexive, and transitive on the class Π. Every function in Π is equivalent to some
semigroup; therefore, the class Π is factorized by this equivalence relation to the set of strongly
continuous semigroups of transformations of the space X.
To cover the set of all strongly continuous operator-valued functions satisfying the condition
F (0) = I by the equivalence relation, we introduce the following definition.
Definition 1 . Operator-valued functions F and G acting from some closed right half-neigh-
borhood of zero on the number axis to the Banach space B(X) of bounded linear operators on a
Banach space X are Chernoff equivalent if condition (2.2) holds for every T > 0 and every u ∈ X.
one. Suppose that there
exists a number m ∈ N such that for all n ≥ m + 1 the operators An are
bounded and the series ∞ k=m+1 μk Ak converges. Suppose also that there exists a linear
subspace
D ⊂ H that is an essential domain of each of the operators An , n ∈ 1, m, and Sn = nk=1 μk Ak ,
n ∈ 1, m.
Then the mean value of the random semigroup F(t) = n∈N exp(−itA
∞ n )μ(n), t ≥ 0, is Chernoff
equivalent to the unitary group U(t) = exp(−itS), t ∈ R, where S = k=1 μk Ak .
Proof. According to the hypotheses of the theorem, the series ∞ k=1 μk Ak converges in the
operator
m norm topology to the operator S and D is an essential domain of the operators Sm =
k=1 μk Ak for all m ∈ N and, hence, of the operator S.
Let us check that the function F satisfies all the hypotheses of Chernoff’s theorem.
Since exp(−itAn )|t=0 = I for all n ∈ N and μ(N) = 1, it follows that F(0) = I.
Let u ∈ H. In view of the strong continuity of the semigroups Un (t) = exp(−itAn ), t ∈ R,
the functions Un (t)u, t ∈ R, are continuous for all n ∈ N and uniformly bounded by the constant
uH . Since exp(−itA
n )B(H) = 1 for all n ∈ N and the measure μ is nonnegative and normed on
the set N, the series n∈N exp(−itAn )uμn converges uniformly in t ∈ R by the Weierstrass test to
a continuous function. Hence, the
operator function F(t), t ∈ R, is strongly continuous.
In addition, F(t)B(H) ≤ n∈N exp(−itAn )B(H) μn = μ(N) = 1; hence, the inequality
F(t)B(H) ≤ 1 holds.
Let us show that for any u ∈ D the function F(t)u = k∈N μk Uk (t)u is continuously differen-
tiable on [0, +∞), and find the limit value of the derivative F (+0)u.
Let u ∈ D. Then, for any k ∈ N,
the function vk (t, u) = exp(−itAk )u, t ∈ R, is continuously
d
differentiable. Moreover, the series k∈N μk dt vk (t) converges uniformly on [0, T ] for any t > 0,
because
d
m ∞
μk v (t)
dt k
= μ k A v
k k (t) = μ k A k u ≤ μ k Ak u + μk Ak · u < +∞.
k∈N k∈N k∈N k=1 k=m
Hence, for any u ∈ D, the function
F(t)u, t > 0, is continuously differentiable on [0, +∞); moreover,
d
its derivative has a limit equal to n∈N μn dt un (+0) = Su as t → +0.
Thus, the function F(t), t ≥ 0, satisfies all the hypotheses of Chernoff’s theorem, and so
limm→+∞ [(F(t/n))n − exp(−itS)]uH = 0; i.e., the function F is Chernoff equivalent to the
semigroup U. Theorem 2 is proved.
The last hypothesis of the theorem—the existence of an essential domain that is common for the
terms of the operator series and for its partial sums—is rather complicated and difficult to verify.
The following statement provides a sufficient condition for this hypothesis of the theorem to hold.
m
Corollary 1. Let A1 , . . . , Am be quantum Hamiltonians, and let j=1 pj = 1 and pj > 0.
If the quadratic form of the operator A1 majorizes the quadratic forms of the operators Ak , k =
definition D(A1 ) is an essential domain for the other operators
2, . . . , m (see [3]), then the domain of
and the averaged Hamiltonian A = m j=1 pj Aj , and the mean value of the semigroups is Chernoff
equivalent to the semigroup generated by the average Hamiltonian.
Remark 2. The assertion of Theorem 2 is nontrivial and interesting even in the particular
case when the operators Aj are bounded and self-adjoint for all j ∈ N.
Remark 3. The common domain of definition D of the operators An is an essential domain
of these operators rather than merely a dense subset of the space H, because the information on
the domain D determines uniquely a unitary semigroup, i.e., a self-adjoint operator, rather than a
symmetric operator that has a variety of self-adjoint extensions.
Remark 4. The assertion of the theorem is also of interest in the case of a finite series. It is
an additive analog of Trotter’s theorem and takes the following form.
Definition 2. We say that a maximal dissipative operator L is the sum of maximal dissipative
operators p1 L1 and p2 L2 for arbitrary p1 ≥ 0 and p2 ≥ 0, p1 + p2 = 1, in the exponential sense if
the operator-valued function G(t) = p1 exp(tL1 ) + p2 exp(tL2 ), t ≥ 0, is Chernoff equivalent to the
semigroup exp(tL).
It is obvious that the sum of operators is commutative and associative, the zero operator is the
zero element, and every maximal dissipative operator has its inverse.
Definition 3. Let L(ε), ε ∈ E, be a function that takes values in the set of maximal dissipative
operators. We say that a maximal dissipative operator L is the integral of the function L(ε), ε ∈ E,
with respect to a nonnegative normed
measure μ on the algebra 2E in the exponential sense if the
operator-valued function G(t) = E exp(tL(ε)) dμ(ε), t ≥ 0 (where the integral is understood in the
weak operator topology in the sense of Pettis), is Chernoff equivalent to the semigroup exp(tL).
Thus, the introduced procedure of averaging the generators of semigroups generalizes the aver-
aging procedure for bounded linear operators.
Remark 5. It seems impossible to give up the condition of nonnegativity
of the numbers μn
∞
and replace it even by the condition of absolute convergence of the series n=1 μn within the
proposed proof method, because then the boundedness condition for the norm of the operator F (t)
may fail for small values of t. Indeed, there exist unbounded self-adjoint operators A1 and A2 in
a Hilbert space H such that if μ1 + μ2 = 1, then, for any t > 0 and any δ > 0, there exists a
unit vector u ∈ H such that the inequality (μ1 exp(−itA1 ) + μ2 exp(−itA2 ))uH ≥ |μ1 | + |μ2 | − δ
holds. As the operators A1,2 , one can take commuting operators with an orthonormal basis of
(1,2)
eigenvectors {ek } corresponding to eigenvalues λk , respectively. Therefore,
(1) (2)
μ1 exp(−itA1 ) + μ2 exp(−itA2 ) ek = μ1 exp −itλk + μ2 exp −itλk ek
and, hence,
(2) (1)
μ1 exp(−itA1 ) + μ2 exp(−itA2 ) ek = μ1 + μ2 exp −it λk − λk .
H
Following [8], we will call such sequences {Ln } power regularizations of the symmetric operator L.
Proposition 2 (see [8]). Let a maximal symmetric operator L have a power regulariza-
tion of order q ∈ N. Let μ be a nonnegative normed purely finitely additive measure on the
set N. If the indices (n− , n+ ) of the operator L satisfy the condition n+ = 0, then the se-
quence of semigroups {exp(−itLn )}, t ≥ 0, converges in the strong operator topology to the
isometric semigroup
exp(−itL), t ≥ 0, uniformly on every finite interval, and the equality
exp(−itL) = N exp(−itLn ) dμ(n), t ≥ 0, holds. If, instead, n− = 0, then the sequence of semi-
groups {exp(−itLn )}, t ≥ 0, converges in the weak operator topology to the contraction semi-
∗ ∗
group exp(−itL ), t ≥ 0, uniformly on every finite interval, and the equality exp(−itL ) =
N exp(−itLn ) dμ(n), t ≥ 0, holds.
Example 2 (a sequence of Hamiltonians that converges in the topology of strong graph con-
vergence to a limit Hamiltonian).
Proposition 3 (see [8]). Let L be a maximal symmetric operator in a Hilbert space H. Let
μ be a nonnegative normed purely finitely additive measure on the set N. Suppose that a sequence
of self-adjoint operators {Ln } is such that its strong graph limit Γ contains the graph ΓL of the
operator L (coincides with it ). If the indices (n− , n+ ) of the operator L satisfy the condition
n+ = 0, then the sequence of semigroups {exp(−itLn )}, t ≥ 0, converges in the strong operator
topology to the isometric semigroup exp(−itL), t ≥ 0, uniformly on any finite interval, and the
equality exp(−itL) = N exp(−itLn ) dμ(n), t ≥ 0, holds. If, instead, n− = 0, then the sequence
of semigroups {exp(−itLn )}, t ≥ 0, converges in the weak operator topology to the contraction
∗ ∗
semigroup exp(−itL ), t ≥ 0, uniformly on every finite interval, and the equality exp(−itL ) =
N exp(−itLn ) dμ(n), t ≥ 0, holds.
Therefore, for any choice of a nonnegative normed purely finitely additive measure μ on the
set N, the following equalities hold in the sense of Definition 3: L = N Ln dμ(n) if n+ = 0, and
L∗ = N Ln dμ(n) if n+ > 0.
Remark 6. If a sequence of self-adjoint operators {Ln } is such that its strong graph limit Γ
contains the graph ΓL of the maximal symmetric operator L (coincides with it), then this does
not mean that the common domain of definition D = n∈N D(Ln ) is dense in the space H; the
variety D may even be trivial.
Thus, the mean values of random semigroups may be semigroups. There are examples when
the mean values of random strongly continuous semigroups are not semigroups but are strongly
continuous operator functions (see [5, 8, 12]).
of the limit function G in the strong operator topology and the equality G(0) = I. Moreover, it
follows from the estimate for the norm of the function F that
ACKNOWLEDGMENTS
This work was supported by the Russian Science Foundation (project no. 14-11-00687) and the
Russian Foundation for Basic Research (project no. 14-01-00516).
REFERENCES
1. F. A. Berezin, The Method of Second Quantization (Nauka, Moscow, 1986) [in Russian]; Engl. transl. of the
1st ed.: The Method of Second Quantization (Academic, New York, 1966), Pure Appl. Phys. 24.
2. V. V. Dodonov, V. I. Man’ko, and V. D. Skarzhinsky, “Nonuniquenesses of the variational description of classical
systems and the quantization problem,” Tr. Fiz. Inst. im. P.N. Lebedeva, Akad. Nauk SSSR 152, 37–89 (1983)
[Proc. Lebedev Phys. Inst. Acad. Sci. USSR 176 (Suppl.), 49–122 (1988)].
3. T. Kato, Perturbation Theory for Linear Operators (Springer, Berlin, 1966; Mir, Moscow, 1972).
4. V. P. Maslov, Operational Methods (Nauka, Moscow, 1973; Mir, Moscow, 1976).
5. J. O. Ogun, Yu. N. Orlov, and V. Zh. Sakbaev, “On a transformation of the initial-data space for a Cauchy
problem with blow-up solutions,” Preprint No. 87 (Keldysh Inst. Appl. Math., Moscow, 2012).
6. Yu. N. Orlov, V. Zh. Sakbaev, and O. G. Smolyanov, “Rate of convergence of Feynman approximations of
semigroups generated by the oscillator Hamiltonian,” Teor. Mat. Fiz. 172 (1), 122–137 (2012) [Theor. Math.
Phys. 172, 987–1000 (2012)].
7. V. Zh. Sakbaev, “Set-valued mappings specified by regularization of the Schrödinger equation with degeneration,”
Zh. Vychisl. Mat. Mat. Fiz. 46 (4), 683–699 (2006) [Comput. Math. Math. Phys. 46, 651–665 (2006)].
8. V. Zh. Sakbaev, Cauchy Problem for a Degenerate Linear Differential Equation and Averaging Its Approximating
Regularizations (Ross. Univ. Druzhby Narodov, Moscow, 2012), Sovrem. Mat., Fundam. Napr. 43.
9. P. R. Chernoff, “Note on product formulas for operator semigroups,” J. Funct. Anal. 2, 238–242 (1968).
10. O. G. Smolyanov, A. G. Tokarev, and A. Truman, “Hamiltonian Feynman path integrals via the Chernoff
formula,” J. Math. Phys. 43 (10), 5161–5171 (2002).
11. O. G. Smolyanov, H. von Weizsäcker, and O. Wittich, “Chernoff’s theorem and discrete time approximations of
Brownian motion on manifolds,” Potential Anal. 26 (1), 1–29 (2007).
12. L. Tartar, “Memory effects and homogenization,” Arch. Ration. Mech. Anal. 111 (2), 121–133 (1990).
13. K. Yosida and E. Hewitt, “Finitely additive measures,” Trans. Am. Math. Soc. 72, 46–66 (1952).
Translated by I. Nikitin