Two-level Adiabatic Transition Probability for Small Avoided Crossings generated by tangential intersections
Abstract
In this paper, the asymptotic behaviors of the transition probability for two-level avoided crossings are studied under the limit where two parameters (adiabatic parameter and energy gap parameter) tend to zero. This is a continuation of our previous works where avoided crossings are generated by tangential intersections and obey a non-adiabatic regime. The main results elucidate not only the asymptotic expansion of transition probability but also a quantum interference caused by several avoided crossings and a coexistence of two-parameter regimes arising from different vanishing orders.
1 Introduction
In quantum mechanics, especially in the quantum chemistry, the adiabatic approximation and the Born-Oppenheimer approximation are widely used. The adiabatic theorem, the motivation of these approximations, asserts that in the slowly varying Hamiltonian the quantum effect like the transition between the energy-levels hardly occurs. From this point of view, it is important to accurately describe how much slowing down the variation shrinks the transition probability.
In this paper, we study a mathematical model such that the transition probability is not always small even in case of the adiabatic approximation. Since the transition probability intuitively depends on the size of the smallest gap between energy-levels, the approaching (resp. receding) speed to (resp. from) the smallest gap, and the quantum interference, we consider asymptotic behavior in a two-parameter singular limit of solutions to the time-dependent Schrödinger equation
(1.1) |
Here, the Hamiltonian is given as a matrix-valued function
(1.2) |
where is a real-valued smooth function and are small positive parameters. Its two eigenvalues are
In this model, the ratio is interpreted as the time variable, are the two energy-levels of , and the adiabatic limit corresponds to the slow variation of the Hamiltonian compared with the time.
According to the adiabatic theorem, one expects that for a solution to (1.1), the projection onto the eigenspace associated with is “small” for every if belongs to the eigenspace associated with at some . More simply, we can say that the adiabatic theorem asserts the smallness of the transition probability. Here, we call
(1.3) |
the transition probability, where is the normalized solution to (1.1) such that with (this solution will be introduced in Appendix A.1). These limits exist under suitable conditions on near infinity (Condition A in this paper). Note that is constant in for any solution .
As long as , the two energy-levels are smooth functions of , and never intersect with each other:
(1.4) |
This quantity called the energy-gap is bounded from below by even if vanishes at some point. This phenomenon occurring near each zero of is called an avoided crossing. The simplest case with a positive constant is investigated individually by L.D. Landau and C. Zener in 30’s [16, 23]. The transition probability
(1.5) |
for this case is known as the Landau-Zener formula. This is exact and true for any positive . For fixed , this formula implies that the transition probability is exponentially small with respect to . There are many results generalizing the Landau-Zener formula. Under some analyticity condition, such an exponential decay estimate is obtained even in case of more general Hamiltonian, for example operator-valued unbounded Hamiltonians as in [4, 5, 12, 13, 17], while a smoothness condition without an analyticity yields nothing but a polynomial decay with respect to as in [15]. Note that in the general setting, the condition of the energy-gap is replaced with the gap condition, which mandates that the spectrum is decomposed into a disjoint union of two subsets and that the distance between them is positive. The history of these generalizations can be consulted in the survey [9] and in the books [8, 19].
The transition probability may become larger when the energy-gap is also small. In our model, this situation occurs if vanishes at some and if (see (1.4)) is sufficiently small compared with . One observes from Landau-Zener formula (1.5) that the transition probability is small and the adiabatic approximation is reasonable if . However, one also observe that it is almost one if . The former situation is called the adiabatic regime, and the latter the non-adiabatic regime, which was discussed in [22] and also in [3, 18].
The leading term of the transition probability is given by the same formula by replacing with when vanishes only at and , namely, the situation that and intersect transversely at (see [12] and also its microlocal version [3]). From the viewpoint of the energy-levels, the approaching/receding near a transversal crossing of is of order .
In the tangential case , the transition probability is studied by one of the authors under the condition corresponding to the adiabatic regime, where stands for the vanishing order of at as in [20, 21] (equivalently, is of order ). In this case, transition probability is exponentially small as tends to 0. The analyticity of and the adiabatic regime condition are necessary for applying the exact WKB method. In fact, the “complex crossing points” of the energy-levels, which are the zeros of on the complex plane and are called turning points in the WKB method, are essential for this case. The adiabatic regime condition implies that these complex crossing points are not too close to each other.
On the other hand, the situation corresponding to the non-adiabatic regime is studied by the other author [10]. He applied other classical method (also used in [2]) to a little bit more general setting. The transition probability is almost one as in the Landau-Zener formula only when is odd, and that it is still small of order when is even.
One of other generalizations is the existence of several avoided crossings. Following the classical probability theory, one may think that the transition probability is obtained by multiplying and summing the non-negative “local transition probability” around each avoided crossing. However, as well as other quantum situations, only a complex-valued probability amplitude is associated with each avoided crossing. Then the “total” probability amplitude is given by multiplying and summing them, and the transition probability is its absolute square. This phenomenon is treated in [14, 21, 22].
This paper is a continuation of the authors’ previous works in the viewpoint of dealing with several avoided crossings generated by tangential intersections with different vanishing orders in the non-adiabatic regime. Our first result, Theorem 1, concerns several tangential avoided crossings in the non-adiabatic regime, that is, with the maximum among the avoided crossings. It shows that the transition probability is almost one when the number of odd avoided crossings is odd and that it is small of order when the number is even. The effect of the quantum interference appears in the coefficient of the term of order . In Formula (2.8), the second term describes the quantum interference while the first term is given by the sum of absolute square of the local transition probability amplitudes. In particular, this coefficient vanishes in some cases. We also show some concrete models (see Remark 2.4 and Examples 2.5 and 2.7).
One notices that the border of the parameter regimes for each avoided crossing depends on the vanishing order . Consequently, there are parameter regimes which is adiabatic for some avoided crossings and non-adiabatic for the others when there are several tangential intersections of and . Our second result , Theorem 2, concerns this situation, and shows that the leading term of the transition probability depends on the parity of the number of odd avoided crossings in the non-adiabatic regime. Since the local probability amplitude around an avoided crossing in the non-adiabatic and adiabatic regime has already been computed in Theorem 1 and in [21], Theorem 2 is obtained by combining them. The novelties of this paper are to examine precisely the transition probability in the intermediate regime, where the non-adiabatic regime and the adiabatic one coexist, and to elucidate a possibility of “switching of the transition probability” by varying two parameters continuously without changing as in Example 2.11. Note that the situation neither adiabatic nor non-adiabatic regime, namely, for some , has not been treated yet, although the case for was treated by [7].
Our proof is based on the classical method. We first introduce the Jost solutions and admitting the asymptotic behavior (2.1) at infinity, and in particular, satisfies (1.3) (see Appendix A.1 for the construction). Then the total transition probability amplitude and the transition probability are and the square of its modulus, where stands for the -entry of the scattering matrix defined by
Note that one has
To study the entries of , we continue the solutions from to . More precisely, we construct solutions which approximately belong to the eigenspece associated with away from any avoided crossings, and compute the transfer matrices between the bases consisting of such solutions. The transfer matrix is almost diagonal when there is no avoided crossing between two points. Thus, the transfer matrix across each avoided crossing near is crucial to obtain the transition probability. The four entries of are the probability amplitudes of the local transition at the vanishing point .
The asymptotic behavior of around each avoided crossing near is given in Theorem 3. As we mentioned above, the exact WKB solutions used in the previous work [21] concerning avoided crossings generated by tangential intersection are no longer valid in the non-adiabatic regime. The solutions are constructed in Section 3 by the method of successive approximations (MSA for short) due to [2, 10]. For example, the and -entries of correspond to the local transition probability amplitude from to and from to when the vanishing order is odd and near . The leading term of them is given by applying the degenerate stationary phase method (Lemma 3.2) to the oscillatory integral (4.7), where the derivative of the phase function off-course has a zero of the same order as .
This paper is organized as follows. In Section 2, we make precise the definitions and settings, and state our main results Theorems 1 and 2. We construct the solutions by the method of successive approximations (MSA) in Section 3, and prove the connection formulas Theorem 3 and Proposition 4.2 by using these solutions in Section 4. Finally, we will complete the proofs in Section 5. To obtain the product of matrices of , we employ an algebraic formula shown in Appendix A.2.
2 Results
2.1 Assumptions and main result
As mentioned in the introduction, we focus on the non-adiabatic regime and work under the -category without any assumption on the analyticity. We notice that the assumption on and the setting of the problem are sightly different from the previous work [22] but the definitions of the transition probability in the series of our works are the same. We first assume the following:
Condition A.
The function has a limit (resp. ) as (resp. ), and satisfies
For simplicity, we assume . Based on the argument in Appendix A.1 under Condition A, one sees the unique existence of Jost solutions () which satisfy the asymptotic conditions:
(2.1) | ||||||
where with (equivalently determined by ). The pairs and form bases of the solution space. Each of them corresponds to one of the eigenvalues and of at the infinity. Note that a function is a solution to (1.1) if and only if is so. This implies that and are orthonormal bases on at each . Then we can introduce the scattering matrix as the change of basis between the pairs of Jost solutions:
(2.2) |
This matrix is unitary. In particular, one has , , and .
Definition 2.1.
The transition probability is defined by
Remark 2.2.
The above definition of the transition probability is equivalent to (1.3). In fact, one has for any , and
Condition B.
The function has a finite number of zeros on , where each zero for is of finite order denoted by .
This assumption implies that for ,
(2.3) |
Let denote the maximal order of the zeros:
(2.4) |
and let denote the index set of which attains (i.e., ). Put
(2.5) |
for . Then implies that determines the sign of on each interval , and in particular determines the sign of , namely for and .
As we mentioned in the introduction, the ratio of and (a specific power of) is crucial. We set
(2.6) |
where
(2.7) |
for each . We focus on the regime . In the case where there exists at least one avoided crossing generated by a tangential intersection, that is , we obtain the following result.

Theorem 1.
Remark 2.3.
When every avoided crossing is generated by a transversal intersection, that is, , Theorem 1 is proven in [22] under an additional assumption that is analytic near the real line. Our method also deduces the same asymptotic formula under Conditions A, B and the additional condition that , replaced with , is sufficiently small (see also [10, Remark 1.2]).
Remark 2.4.
The factor depends only on the highest order of the zeros and never vanishes while the factor depends also on the behavior of not only the local property at zeros and may vanish. This vanishing phenomenon corresponds to the destructive quantum interference. Suppose, for example, that among are the same. Put and . Then the condition for to vanish is given by
(2.9) |
where
(2.10) |
The algebraic curve (2.9) in -variables appears as so-called Fermi surface in the context of the discrete Laplacian on the -dimensional diamond lattice, which is a generalization of the hexagonal lattice (see [1]).
The rest of this subsection is devoted to the concrete expression of the transition probability in Theorem 1 for typical models by means of the following geometric quantity on the (time-energy) phase space. For each , we denote the area enclosed by and between and by
(2.11) |

Example 2.5 (Two avoided crossings).
Let the number of avoided crossings be two. Then the transition probability is (resp. ) modulo if the sum of the order of zeros is odd (resp. even). In particular, when the two zeros have the same order, one sees that independent of the parity of the order. We give the coefficient attached to in each situation:
-
(a).
;
(2.12) -
(b).
and ;
(2.13) -
(c).
and ;
(2.14)
Remark 2.6.
Example 2.7 (Three avoided crossings).
Let . The transition probability is determined modulo by the sum whereas the coefficient attached to is determined by zeros only for and by integrals of between them. In particular, when and , the coefficient is given by the same formula as a model with two avoided crossings.
-
(a).
and ;
(2.16) -
(b).
and ;
(2.17) -
(c).
and ;
Remark 2.8.
While the destructive quantum interference condition in the case is that the area on the phase space is quantized (i.e. discretized) as in (2.15), that condition in is that two areas lie along the Fermi curve.


2.2 Coexistence of the two parameter regimes

Recall that the quantum dynamics around each avoided crossing near depends principally on the magnitude of the parameter . More precisely, and correspond to the non-adiabatic and adiabatic regimes (note that the regime is studied only for the transversal case [3]). This parameter is different for two zeros of with different order, thus the transition problem with several avoided crossings generated by tangential intersections admits various regimes.
Note that obeys the algebraic order relation:
(2.18) |
The regime considered in Theorem 1 corresponds to non-adiabatic regime for every . Conversely, the regime (with standing for the minimum order ) considered in [20] corresponds to adiabatic regime for every .
Here, we consider the case that the two different regimes coexist, that is, the set of indices is decomposed into a disjoint union of two parts
such that
Again by (2.18), this corresponds to
where we put and . We also put
Figure 4 illustrates the regimes for . When each zero of is either of order 3 or 11, we here study the regime while Theorem 1 and [20] concern the regime and , respectively. In Figure 5, the problem here corresponds to or . Note also that these figures are displayed with a logarithmic scale. Hence the borders between regimes are straight lines. Indeed, the border for some is rewritten as .

In the study of adiabatic regime, one of the authors employed the exact-WKB method [20] which requires the function to be analytic. Hence we also suppose the additional condition.
Condition C.
is real-analytic on an interval containing .
Under this condition, when is small enough, there exist zeros of near each like the power roots. We call these zeros turning points and denote the nearest two turning points to the real axis on the upper half-plane by , which behave like
as . As in [22], the action integral for is given by
where the path is the segment from to and the branch of the square root of the integrand is at . Note that on this branch and there exist such that
as .
Roughly speaking, the absolute value of the “probability amplitude of the transition around an avoided crossing near ” is small in the limit . Contrary to the non-adiabatic case , this fact is independent of the parity of . The probability amplitude has the same order as
From the sake of distinguishing this difference, we introduce
where stands for the number of the elements of , and the elements are labeled in ascending order .
We also introduce the effective energy in this regime by
(2.19) |
where is taken as when (see also (2.21)).
Putting and introducing two functions
we state the asymptotic expansion of the transition probability in this intermediate regime:
Theorem 2.
Remark 2.9.
The parity which characterizes the transition probability depends not only on determined by but also on determined by the regime. This implies that the switch of occurs with changing the regime without doing the energy (see Figure 6).
As we mentioned in Section 2.2, Theorem 2 covers the range of the pair of the parameters included in the parameter regime determined by and . Regarding as a function of like a one-parameter problem, we find the typical cases, which realize the intermediate regime and .
- Polynomial case:
-
If with
the contribution coming from is exponentially small.
- Logarithmic case:
-
If with some positive constant , the contribution coming from must be taken into account, since .
In the former case, the leading term is similar to that in Theorem 1 and is given by
(2.20) |
where the factor is of and consulted in (5.16). In other cases including the latter case, the leading term is more complicated than (2.20). In fact, the leading term is of the form:
where and are of and referred in (5.17) and (5.18) respectively.
Remark 2.10.
In our method, we represent the Jost solution by several bases. As we mentioned in the introduction, each basis is consists of solutions corresponding to an eigenvector associated with in each region between two avoided crossings. Consequently, the absolute value of the coefficients gives . Moreover, one observes from our proof that
(2.21) |
for any , where . In this sense, (the square of) the modulus of the probability amplitude that the energy follows the curve is almost 1. In Figure 6, we draw this curve in green.
Example 2.11.
Figure 6 shows an example case with ().

3 Construction of exact solutions
In this section, we construct exact solutions which form a local basis near each vanishing point by means of a method of successive approximations due to [6]. While the equation treated there is a second order system of time-independent Schrödinger equations, our equation in this paper is a first order system.
3.1 Estimates of fundamental solutions
For simplicity, we assume , and let denote . Let be a small interval containing the vanishing point in its interior. We fix
as a particular solution to
(3.1) |
respectively, where stands for . For any point , we define an integral operator by
(3.2) |
where is the Banach space of continuous functions on equipped with the norm . Since
(3.3) |
for any , the integral operator is well-defined as a fundamental solution of respectively for the signs .
Using these fundamental solutions with base points respectively, our equation (1.1) turns into the integral system with arbitrary constants :
(3.4) |
Depending on the choice of the base points and , the initial value for and at these points are determined:
In the next subsection, we show a construction of the unique solution to the system by an iteration. For this purpose, we give the following estimates for the fundamental solutions. Note that they are independent of , and this estimate gives the critical rate .
Let for be a norm on the space of continuously differentiable functions defined by
(3.5) |
Proposition 3.1.
For any , there exists such that
(3.6) |
for small enough.
For the sake of the proof of Proposition 3.1, we introduce the the following lemma which plays an important role in this paper.
Lemma 3.2.
On a compact interval , consider the integral
(3.7) |
with a continuously differentiable function possibly depending on . Then there exists a constant independent of (but depending on ) such that
(3.8) |
for small enough when does not vanish on . If is the unique zero in of , one has
(3.9) |
where denotes the order of vanishing at . Moreover if is independent of , we have
(3.10) |
Here, the constant is given by
(3.11) |
with
(3.12) |
Proof of Lemma 3.2.
Suppose that does not vanish on , that is, a non-stationary case. Then we have for
(3.13) |
This with an integration by parts and the compactness of implies the estimate from above of by .
Suppose next that is the unique zero of in . Take a smooth cut-off function such that for and for . On the support of , one has the estimate
This with a similar argument as above implies that contribution coming from the support of to the integral is estimated by . The other part is estimated by since the support of is , and the estimate (3.9) follows.
We then suppose also that is independent of . Let be the smooth function defined near such that
Take a smooth cut-off function whose value is 1 near and supported only on a small neighborhood where the change of the variable is valid. Then one has
with and a smooth function satisfying . The resulting asymptotic formula is obtained from this integral by applying the method of degenerate stationary phase (see e.g. [11]). The non-stationary estimate (3.8) is applicable on the support of .
∎
Based on this lemma, let us prove Proposition 3.1.
Proof of Proposition 3.1.
3.2 Method of successive approximations (MSA)
From Proposition 3.1, it follows that for each , there uniquely exists a solution to the integral system (3.4). By a linearity of the system, the solution is given by the linear combination of the solutions and corresponding to the choices and for . Moreover, the solution can be constructed by MSA:
(3.14) |
(3.15) |
in for a fixed small . These iteration formulas imply that the solutions admit the asymptotic expansions when as follows:
(3.16) | ||||
Notice that Proposition 3.1 allows us to choose arbitrarily the base points of the fundamental solutions in this construction. As we mentioned in Remark 3.3, we obtain better asymptotic formulas
(3.17) |
if , that is, the integral interval of does not contain the zero, and likewise
(3.18) |
if , that is, the integral interval of does not.
Take -independent constants and such that and . We define the four MSA solutions , and in as
(3.19) | ||||||
According to Proposition 3.1 and the asymptotic formula (3.16), one sees that the asymptotic behaviors of these MSA solutions as are clear on , and also that (resp. ) behave like the initial data near (resp. ) for a fixed small . Moreover the pairs and form bases of the space of solutions.
4 Connection formulas
In this section, we also treat a zero of as for simplicity. The purpose of this section is to establish the two connection formulas. One of them connects across the vanishing point and the other does between the consecutive vanishing points.
4.1 Across the vanishing point
The crucial point of this proof is the connection formula between the two bases and introduced by (3.19) in the previous section. The claim of this subsection is the asymptotic behavior of the transfer matrix as follows:
Theorem 3.
Remark 4.1.
By construction and the symmetry , we have
(4.4) |
This makes symmetrical in the following sense:
(4.5) |
where is the special unitary group of degree (see §A.2). Namely, is unitary . This is a consequence that the time evolution by is unitary and that for , the basis
of is orthonormal.
Proof of Theorem 3.
Since the pair forms a basis of the space of solutions, the matrix is invertible for any . Then (4.1) is rewritten as
where the right-hand side is also independent of . Substituting for , we have
(4.6) | ||||
Here, we have used and (3.16).
In order to prove Theorem 3, it is enough to compute the asymptotic behavior of the quantity as with . The computation of
(4.7) |
is carried out based on Lemma 3.2. In fact, since the integral interval of (4.7) contains the zero of , we have
(4.8) |
∎
4.2 Between the vanishing points
In addition to Theorem 3, which gives the connection formula around the vanishing point of , a similar argument yields the following proposition, which gives the connection formula between two consecutive zeros of .
Let be two consecutive zeros of (i.e. on ) with multiplicities , and let be base points such that . The configuration of these points implies that each interval includes and does not intersect with each other. We set . We can consider two bases and , which are similarly given by the formulas (3.19) with and respectively.
Proposition 4.2.
The change of basis given by
(4.10) |
admits the following asymptotic behavior as with :
(4.11) |
Remark 4.3.
Proof of Proposition 4.2.
We can derive from (4.10) the expression of by a similar way to the proof of Theorem 3. We set , and compute the matrix by using the value at :
Here, we have used the asymptotic formulas (3.17) and (3.18). Note that the integral interval of the fundamental solutions for the construction does not contain any zero of . We deduce (4.11) from this with the identity
(4.12) |
∎
5 End of the proofs
By using the transfer matrices , introduced in the previous section and , in Appendix A.1, the scattering matrix is represented as
(5.1) |
Here, the matrix (resp. ) has a similar form to which connects Jost solutions (resp. ) and local solutions near (resp. ). The previous section shows the asymptotic behaviors of and (see (4.9), (4.11)), and the appendix does those of and (see (A.17), (A.18), (A.20)). The formula of depending on the sign of can be rewritten as
(5.2) |
by means of and
(5.3) |
where is a complex structure on , that is, . Note that all of these transfer matrices are elements of , which is mentioned in §A.2. By using the notation (A.23), the scattering matrix is expressed as (resp. ) if is even (resp. odd). This implies that, when is even (resp. odd), the off-daiagnoal entry equals (resp. ). From Lemma A.4 alone, it is complicated to examine the asymptotics of up to the coefficient of . However, thanks to the unitarity of the scattering matrix , that is, , the computation of the asymptotic behavior of can be reduced to that of even if is odd.
5.1 Proof of Theorem 1
Let us demonstrate the proof of Theorem 1. Taking the complex numbers , and in Lemma A.4 as
and noting for any , we have from the algebraic formula (A.28) the asymptotic behavior of as follows:
(5.4) | ||||
where stands for . Concerning the error terms in (5.4), the former one, i.e. , is a higher order error term coming from and the latter is a cross term between the largest vanishing order and the second largest one. This error coming from a cross term is at most . If is odd, is not real for , and the following phase shift term may arise from the product depending on the sign of and :
(5.5) |
Therefore, when is odd, the quantity behaves like
(5.6) |
where
On the other hand, is real when is even. In this case, the quantity can be computed similarly as the formula (5.6) by replacing with and with . Therefore we have completed the proof of Theorem 1.
5.2 Proof of Theorem 2
The intermediate regime where and requires making use of the transfer matrix based on the exact WKB method in [21] under Condition C at the vanishing points governed by an adiabatic regime. Namely, in the intermediate regime, the transfer matrix is replaced as
(5.7) |
where is equal to (4.9) and will be given below. The key of the proof of Theorem 2 is to reduce the computation of the product including the different kinds of the transfer matrices in (5.7) to the same computation as in the proof of Theorem 1.
According to the exact WKB method in [21], we obtained the existences of the exact WKB solutions near a vanishing point and their asymptotic behaviors away from the vanishing point under an adiabatic regime. Moreover we got the change of basis between and . By matching the asymptotic behaviors of the exact WKB solutions and the MSA solutions on their semiclassical wave front sets referred in [10], we have the connection formula between them. Consequently, in the case where both and are even, the transfer matrix is of the form:
where have the asymptotic expansions as :
(5.8) | ||||
(5.9) |
with the notations given in §2.2. Actually is of and is exponentially small of as . Notice that the above case where both and are even implies that the eigenvalue of (i.e. the energy of the system: ) and the energy without an interaction (i.e. have the same sign before and behind the vanishing point . Conversely, in the case where the sign of does not coincide with that of near the vanishing point, the correspondence of the exact WKB solutions to the MSA solutions varies. In fact, for depends on and as follows:
(5.10) |
where the matrix equipped with useful properties:
(5.11) |
and is an operator of taking a complex conjugate, that is . Notice that each matrix in (5.10) from which are removed the factor if it exists belongs to . Introducing the notation of the transfer matrix belonging to as follows:
(5.12) |
and recalling the commutative property (5.11), we can express the scattering matrix in the intermediate regime by
(5.13) |
where the notation of for the matrix depending on stands for
(5.14) |
with the same notations as in §2.2. Remark that if then . Hence the expression (5.13) implies that the algebraic lemma (A.27) can be applied directly to the computation of the off-diagonal entry of the product, and that the transition probability depends also on the parity of the number .
As a sequel to this computation of the product in (5.13), we can derive the dependence on more precisely. Denoting the -entry of by , we see that is of and, in particular for . Setting, similarly, the -entry of by , we can rewrite as
(5.15) |
where and are uniquely determined by (5.9), (5.12) and (5.14). Notice that and are of in each regime. On the other hand, can be regarded as the matrix by replacing (see (2.19)) with . From this fact, it is deduced that the asymptotic of the transition probability in the intermediate regime is determined by the effective energy . Hence, we can obtain the asymptotic behavior of as follows:
where are given in §2.2 and
(5.16) | ||||
(5.17) | ||||
(5.18) |
The proof of Theorem2 have been completed.
Appendix A Appendix
A.1 Jost solutions
In this subsection A.1, we give the existence of the Jost solutions for the definition of the scattering matrix. We remark that the smallness of is not required for the argument here.
We first consider the Jost solutions near . A discussion for is done similarly but the difference is that we are assuming that is positive (Condition A). Let denote the limiting Hamiltonian at :
(A.1) |
The functions defined by
(A.2) |
where and , are particular solutions to and form a basis of for each .
Proposition A.1.
There uniquely exists a pair of solutions to the system (1.1) such that
(A.3) |
Proof.
Let be a -matrix valued -function. We have
with . Thus, if satisfies
(A.4) |
each column of the matrix-valued function is a solution to the equation (1.1). Put . From the identity
(A.5) |
and Condition A, the matrix-valued function :
(A.6) |
is integrable on the half-line . Then the function
(A.7) |
is well-defined and solves the equation (A.4) with the boundary condition
Here we recall that stands for the unit matrix. We finally obtain the solutions with the asymptotic behavior (A.3):
(A.8) |
∎
From Proposition A.1 and the trace-free property of , the pair of forms a basis. Similarly, this fact implies that (resp. ) coincides with the Jost solution (resp. ).
Next, we give the asymptotic behaviors of as near some fixed point . Take (recall that is the first zero of ) satisfying
(A.9) |
We introduce some kind of the action integral taking into account of contributions from the infinity as
and put
Proposition A.2.
We have
as uniformly for in a small neighborhood of .
Before proving Proposition A.2, we prepare the following.
Lemma A.3.
Let be a matrix of the form:
with and . For , one has
Proof.
Since , an algebraic computation gives
(A.10) |
where . We have under and . This gives the following asymptotic formula:
The lemma follows from . ∎
Proof of Proposition A.2.
From the expression (A.6), defined by (A.7) is written as
(A.11) |
where
Apply Lemma A.3 with
(A.12) |
By the choice of with the condition (A.9), never vanishes for near . By definition, we have , and consequently . Then Lemma A.3 shows
(A.13) |
Note that we have the following decomposition of :
(A.14) |
Since admits the asymptotic formula
(A.15) |
The asymptotic formula
(A.16) |
is directly deduced from Propositions A.1 and A.2. Therefore, we obtain
(A.17) |
For the Jost solutions , the same argument as above works when and a similar one induces the existences and the asymptotic behaviors when .
In the case where , by exchanging the sub-index for , one sees
(A.18) |
Here
with satisfying that the second integral term in the right-hand side does not vanish.
In the case where , we choose instead of (A.2) particular solutions to as
(A.19) |
where and . They coincide with the leading terms of Jost solutions when and satisfy the asymptotic formulas:
as for each . One sees that, with (A.19), Proposition A.1 also holds. One also have similar asymptotic formulas to those of Proposition A.2:
as uniformly in a small neighborhood of , where . We obtain
(A.20) |
A.2 Algebraic lemma
In order to know the asymptotic behavior of the scattering matrix (5.1), it suffices to compute the products of the matrices of the following forms:
where , , and are complex numbers such that , namely , and as . Notice that, in our context (see (4.2)), and have this form with the numbers given modulo by
In this subsection we give an algebraic formula by means of these notations , and for simplicity. We know that the product of them is of the form
(A.21) |
Let be the special unitary group of degree given by
(A.22) |
One sees that all of the above matrices belong to . Denoting the products of these matrices by
(A.23) |
we get the following lemma:
Lemma A.4.
As , the following asymptotic formulas hold.
(A.24) | |||
(A.25) | |||
(A.26) |
with the convention that .
The proof of this lemma is based on the mathematical induction for the product of the matrices (A.21). A simple computation of , which corresponds to the transition probability, gives
(A.27) |
with the convention that . Note that we used and in the above computation. In particular, when , we have
(A.28) |
Remark that the factor does not appear explicitly in the leading term in the formulas (A.27) and (A.28).
Acknowledgements
The authors gratefully express our thanks to M. Zerzeri for useful discussions. This research is supported by the Grant-in-Aid for JSPS Fellows Grant Number JP22J00430, and Grant-in-Aid for Scientific Research (C) JP18K03349, JP21K03282. The authors are grateful to the support by the Research Institute for Mathematical Sciences, an International Joint Usage/Research Center located in Kyoto University.
References
- [1] K. Ando, H. Isozaki, H. Morioka. Spectral properties of Schrödinger operators on perturbed lattices. Ann. Henri Poincaré 17 (2016), no.8, 2103–2171.
- [2] M. Assal, S. Fujiié and K. Higuchi, Semiclassical resonance asymptotics for systems with degenerate crossings of classical trajectories. Int. Math. Res. Not. 2024 (2024), no.8, 6879–6905.
- [3] Y. Colin de Verdière, M. Lombardi and J. Pollet, The microlocal Landau-Zerner formula, Ann. Inst. H. Poincaré, 71(1999), no. 1, 95–127.
- [4] Y. Colin de Verdière, The level crossing problem in semi-classical analysis. I. The symmetric case. Proceedings of the International Conference in Honor of Frédéric Pham (Nice, 2002). Ann. Inst. Fourier (Grenoble) 53(2003), no. 4, 1023–1054.
- [5] Y. Colin de Verdière, The level crossing problem in semi-classical analysis. II. The Hermitian case. Ann. Inst. Fourier (Grenoble) 54(2004), no. 5, 1423–1441.
- [6] S. Fujiié, A. Martinez and T. Watanabe, Widths of resonances above an energy-level crossing, J. Funct. Anal. 280(2021), no. 6, 48 pages.
- [7] G.-A. Hagedorn, Proof of the Landau-Zener formula in an adiabatic limit with small eigenvalue gaps, Commun. Math. Phys. 136(4)(1991), 33–49.
- [8] G.-A. Hagedorn, Molecular propagation through electron energy level crossings. Mem. Amer. Math. Soc. 111(1994), no. 536.
- [9] G.-A. Hagedorn and A. Joye, Recent results on non-adiabatic transitions in Quantum mechanics, Recent advances in differential equations and mathematical physics, Contemp. Math., 412, Amer. Math. Soc., Providence, RI, 2006, 183–198.
- [10] K. Higuchi, Local scattering matrix for a degenerate avoided-crossing in the non-coupled regime, Lett. Math. Phys., to appear.
- [11] L. Hörmander, The analysis of linear partial differential operators. I: Distribution theory and Fourier analysis, Grundlehren der Mathematischen Wissenschaften, 256(1983). Berlin Heidelberg-New York -Tokyo: Springer-Verlag. IX
- [12] A. Joye, Proof of the Landau-Zener Formula, Asympt. Anal. 9(1994), 209–258.
- [13] A. Joye, H. Kunz and C.-E. Pfister, Exponential Decay and Geometric Aspect of Transition Probabilities in the Adiabatic Limit. Ann. Phys. 208 (1991), 299–332.
- [14] A. Joye, G. Mileti and C.-E. Pfister, Interferences in adiabatic transition probabilities mediated by Stokes lines, Phys. Rev. A. 44(1991), 4280–4295.
- [15] T. Kato, On the adiabatic theorem of quantum mechanics, J. Phys. Soc. Japan 5(1950), 435–439.
- [16] L.D. Landau, Collected papers of L. D. Landau, Pergamon Press, 1965.
- [17] A. Martinez, Precise exponential estimates in adiabatic theory, J. Math. Phys. 35(1994), 3889–3915.
- [18] V. Rousse, Landau-Zener transitions for eigenvalue avoided crossings, Asympt. Anal. 37(2004) no. 3-4, 293–328.
- [19] S. Teufel, Adiabatic perturbation theory in quantum dynamics, Lecture Notes in Mathematics 1821. Springer-Verlag, Berlin, Heidelberg, New York, 2003.
- [20] T. Watanabe, Adiabatic transition probability for a tangential crossing, Hiroshima Math. J. 36(2006), no. 3, 443–468.
- [21] T. Watanabe, Exact WKB approach to 2-level adiabatic transition problems with a small spectral gap, Tohoku Mathematical Publications 37(2012). 1–68.
- [22] T. Watanabe and M. Zerzeri, Landau-Zener formula in a “non-adiabatic” regime for avoided crossings, Anal. Math. Phys. 11(2021), no. 82, 47.
- [23] C. Zener, Non-Adiabatic Crossing of Energy Levels, Proc. R. Soc. London A, 137(1932), 696–702.