This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Two-level Adiabatic Transition Probability for Small Avoided Crossings generated by tangential intersections

Kenta Higuchi and Takuya Watanabe
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 h,ε+0h,\varepsilon\to+0 of solutions to the time-dependent Schrödinger equation

ihddtψ(t)=H(t;ε)ψ(t),t.ih\frac{d}{dt}\psi(t)=H(t;\varepsilon)\psi(t),\quad t\in\mathbb{R}. (1.1)

Here, the Hamiltonian H(t;ε)H(t;\varepsilon) is given as a 2×22\times 2 matrix-valued function

H(t;ε):=(V(t)εεV(t)),H(t;\varepsilon):=\begin{pmatrix}V(t)&\varepsilon\\ \varepsilon&-V(t)\end{pmatrix}, (1.2)

where V(t)V(t) is a real-valued smooth function and h,εh,\varepsilon are small positive parameters. Its two eigenvalues are

E±(t;ε)=±V(t)2+ε2.E_{\pm}(t;\varepsilon)=\pm\sqrt{V(t)^{2}+\varepsilon^{2}}.

In this model, the ratio t/ht/h is interpreted as the time variable, E±(t;ε)E_{\pm}(t;\varepsilon) are the two energy-levels of H(t;ε)H(t;\varepsilon), and the adiabatic limit h0h\to 0 corresponds to the slow variation of the Hamiltonian H(t;ε)H(t;\varepsilon) compared with the time.

According to the adiabatic theorem, one expects that for a solution ψ(t)\psi(t) to (1.1), the projection Π(t;ε)ψ(t)\Pi_{-}(t;\varepsilon)\psi(t) onto the eigenspace associated with E(t;ε)E_{-}(t;\varepsilon) is “small” for every tt\in\mathbb{R} if ψ(t0)\psi(t_{0}) belongs to the eigenspace associated with E+(t0;ε)E_{+}(t_{0};\varepsilon) at some t0t_{0}\in\mathbb{R}. More simply, we can say that the adiabatic theorem asserts the smallness of the transition probability. Here, we call

P(ε,h):=limt+Π(t;ε)J+(t)22P(\varepsilon,h):=\lim_{t\to+\infty}\left\|\Pi_{-}(t;\varepsilon)J_{\ell}^{+}(t)\right\|_{\mathbb{C}^{2}}^{2} (1.3)

the transition probability, where J+J_{\ell}^{+} is the normalized solution to (1.1) such that limtΠ(t;ε)J+(t)2=0\lim_{t\to-\infty}\left\|\Pi_{-}(t;\varepsilon)J_{\ell}^{+}(t)\right\|_{\mathbb{C}^{2}}=0 with J+(t)2=1\left\|J_{\ell}^{+}(t)\right\|_{\mathbb{C}^{2}}=1 (this solution will be introduced in Appendix A.1). These limits exist under suitable conditions on VV near infinity (Condition A in this paper). Note that ψ(t)22=±Π±(t;ε)ψ(t)22\left\|\psi(t)\right\|_{\mathbb{C}^{2}}^{2}=\sum_{\pm}\left\|\Pi_{\pm}(t;\varepsilon)\psi(t)\right\|^{2}_{\mathbb{C}^{2}} is constant in tt for any solution ψ\psi.

As long as ε>0\varepsilon>0, the two energy-levels E±(t;ε)E_{\pm}(t;\varepsilon) are smooth functions of tt, and never intersect with each other:

inft|E+(t;ε)E(t;ε)|=inft2V(t)2+ε22ε>0.\inf_{t\in\mathbb{R}}\left|E_{+}(t;\varepsilon)-E_{-}(t;\varepsilon)\right|=\inf_{t\in\mathbb{R}}2\sqrt{V(t)^{2}+\varepsilon^{2}}\geq 2\varepsilon>0. (1.4)

This quantity called the energy-gap is bounded from below by 2ε2\varepsilon even if VV vanishes at some point. This phenomenon occurring near each zero of VV is called an avoided crossing. The simplest case V(t)=vtV(t)=vt with a positive constant vv is investigated individually by L.D. Landau and C. Zener in 30’s [16, 23]. The transition probability

P(ε,h)=exp(πε2vh)P(\varepsilon,h)=\exp\left(-\frac{\pi\varepsilon^{2}}{vh}\right) (1.5)

for this case is known as the Landau-Zener formula. This is exact and true for any positive ε,h\varepsilon,h. For fixed ε>0\varepsilon>0, this formula implies that the transition probability is exponentially small with respect to h>0h>0. 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 hh 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 V(t)V(t) vanishes at some tt and if ε\varepsilon (see (1.4)) is sufficiently small compared with hh. One observes from Landau-Zener formula (1.5) that the transition probability is small and the adiabatic approximation is reasonable if εh1/2\varepsilon\gg h^{1/2}. However, one also observe that it is almost one if εh1/2\varepsilon\ll h^{1/2}. 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 vv with |V(0)||V^{\prime}(0)| when V(t)V(t) vanishes only at t=0t=0 and V(0)0V^{\prime}(0)\neq 0, namely, the situation that V(t)V(t) and V(t)-V(t) intersect transversely at t=0t=0 (see [12] and also its microlocal version [3]). From the viewpoint of the energy-levels, the approaching/receding |E+(t;ε)E(t;ε)|2ε=2(V(t)2+ε2ε)|E_{+}(t;\varepsilon)-E_{-}(t;\varepsilon)|-2\varepsilon=2(\sqrt{V(t)^{2}+\varepsilon^{2}}-\varepsilon) near a transversal crossing of ±V\pm V is of order |t||t|.

In the tangential case V(0)=0V^{\prime}(0)=0, the transition probability is studied by one of the authors under the condition εhm/(m+1)\varepsilon\gg h^{m/(m+1)} corresponding to the adiabatic regime, where mm stands for the vanishing order of VV at t=0t=0 as in [20, 21] (equivalently, |E+(t;ε)E(t;ε)|2ε|E_{+}(t;\varepsilon)-E_{-}(t;\varepsilon)|-2\varepsilon is of order |t|m|t|^{m}). In this case, transition probability is exponentially small as hε(m+1)/mh\varepsilon^{-(m+1)/m} tends to 0. The analyticity of VV 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 E+(t;ε)E(t;ε)E_{+}(t;\varepsilon)-E_{-}(t;\varepsilon) 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 εhm/(m+1)\varepsilon\ll h^{m/(m+1)} 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 mm is odd, and that it is still small of order εhm/(m+1)\varepsilon h^{-m/(m+1)} when mm 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, εhm/(m+1)\varepsilon\ll h^{m/(m+1)} with mm 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 εhm/(m+1)\varepsilon h^{-m/(m+1)} when the number is even. The effect of the quantum interference appears in the coefficient of the term of order εhm/(m+1)\varepsilon h^{-m/(m+1)}. 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 mm. 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 V(t)V(t) and V(t)-V(t). 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 ε,h\varepsilon,h continuously without changing V(t)V(t) as in Example 2.11. Note that the situation neither adiabatic nor non-adiabatic regime, namely, εhm/(m+1)\varepsilon\sim h^{m/(m+1)} for some m2m\geq 2, has not been treated yet, although the case for m=1m=1 was treated by [7].

Our proof is based on the classical method. We first introduce the Jost solutions J±=J±(t;ε,h)J_{\ell}^{\pm}=J_{\ell}^{\pm}(t;\varepsilon,h) and Jr±=Jr±(t;ε,h)J_{r}^{\pm}=J_{r}^{\pm}(t;\varepsilon,h) admitting the asymptotic behavior (2.1) at infinity, and in particular, J+J_{\ell}^{+} satisfies (1.3) (see Appendix A.1 for the construction). Then the total transition probability amplitude and the transition probability are s21(ε,h)s_{21}(\varepsilon,h) and the square of its modulus, where s21(ε,h)s_{21}(\varepsilon,h) stands for the (2,1)(2,1)-entry of the scattering matrix S(ε,h)S(\varepsilon,h) defined by

(J+(t;ε,h),J(t;ε,h))=(Jr+(t;ε,h),Jr(t;ε,h))S(ε,h).(J_{\ell}^{+}(t;\varepsilon,h),J_{\ell}^{-}(t;\varepsilon,h))=(J_{r}^{+}(t;\varepsilon,h),J_{r}^{-}(t;\varepsilon,h))S(\varepsilon,h).

Note that one has

Π(t;ε)J+(t;ε,h)s21(ε,h)Jr(t;ε,h)0as t+.\Pi_{-}(t;\varepsilon)J_{\ell}^{+}(t;\varepsilon,h)-s_{21}(\varepsilon,h)J_{r}^{-}(t;\varepsilon,h)\to 0\quad\text{as }t\to+\infty.

To study the entries of S(ε,h)S(\varepsilon,h), we continue the solutions J±J_{\ell}^{\pm} from -\infty to ++\infty. More precisely, we construct solutions which approximately belong to the eigenspece associated with E±(t;ε)E_{\pm}(t;\varepsilon) 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 TkT_{k} across each avoided crossing near tkt_{k} is crucial to obtain the transition probability. The four entries of TkT_{k} are the probability amplitudes of the local transition at the vanishing point tkt_{k}.

The asymptotic behavior of TkT_{k} around each avoided crossing near tkt_{k} 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 (1,2)(1,2) and (2,1)(2,1)-entries of T(ε,h)T(\varepsilon,h) correspond to the local transition probability amplitude from E+E_{+} to EE_{-} and from EE_{-} to E+E_{+} when the vanishing order mm is odd and V(t)(ttk)0V(t)(t-t_{k})\geq 0 near tkt_{k}. 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 20tV(r)𝑑r\mp 2\int_{0}^{t}V(r)dr off-course has a zero of the same order as VV.

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 2n+12n+1 matrices of SU(2)\mathrm{\,SU}(2), 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 CC^{\infty}-category without any assumption on the analyticity. We notice that the assumption on V(t)V(t) 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 V(t)C(;)V(t)\in C^{\infty}(\mathbb{R};\mathbb{R}) has a limit Vr{0}V_{r}\in\mathbb{R}\setminus\{0\} (resp. V{0}V_{\ell}\in\mathbb{R}\setminus\{0\}) as t+t\to+\infty (resp. -\infty), and satisfies

VVrL1([0,+)),VVL1((,0]),VL1().V-V_{r}\in L^{1}([0,+\infty)),\quad V-V_{\ell}\in L^{1}((-\infty,0]),\quad V^{\prime}\in L^{1}(\mathbb{R}).

For simplicity, we assume Vr>0V_{r}>0. Based on the argument in Appendix A.1 under Condition A, one sees the unique existence of Jost solutions J±(t)J_{\bullet}^{\pm}(t) ({,r}\bullet\in\{\ell,r\}) which satisfy the asymptotic conditions:

Jr+(t)exp[ithVr2+ε2](cosθrsinθr)\displaystyle J_{r}^{+}(t)\sim\exp\left[-\frac{it}{h}\sqrt{V_{r}^{2}+\varepsilon^{2}}\right]\left(\begin{array}[]{c}\cos{\theta_{r}}\\ \sin{\theta_{r}}\end{array}\right) ast+,\displaystyle\textrm{as}\,\,\,t\to+\infty, (2.1)
Jr(t)exp[+ithVr2+ε2](sinθrcosθr)\displaystyle J_{r}^{-}(t)\sim\exp\left[+\frac{it}{h}\sqrt{V_{r}^{2}+\varepsilon^{2}}\right]\left(\begin{array}[]{c}-\sin{\theta_{r}}\\ \cos{\theta_{r}}\end{array}\right) ast+,\displaystyle\textrm{as}\,\,\,t\to+\infty,
J+(t)exp[ithV2+ε2](cosθsinθ)\displaystyle J_{\ell}^{+}(t)\sim\exp\left[-\frac{it}{h}\sqrt{V_{\ell}^{2}+\varepsilon^{2}}\right]\left(\begin{array}[]{c}\cos{\theta_{\ell}}\\ \sin{\theta_{\ell}}\end{array}\right) ast,\displaystyle\textrm{as}\,\,\,t\to-\infty,
J(t)exp[+ithV2+ε2](sinθcosθ)\displaystyle J_{\ell}^{-}(t)\sim\exp\left[+\frac{it}{h}\sqrt{V_{\ell}^{2}+\varepsilon^{2}}\right]\left(\begin{array}[]{c}-\sin{\theta_{\ell}}\\ \cos{\theta_{\ell}}\end{array}\right) ast,\displaystyle\textrm{as}\,\,\,t\to-\infty,

where tan2θ=ε/V\tan{2\theta_{\bullet}}=\varepsilon/V_{\bullet} with 0<θ<π/20<\theta_{\bullet}<\pi/2 (equivalently determined by θ=arctan(ε1(V2+ε2V))\theta_{\bullet}=\arctan(\varepsilon^{-1}(\sqrt{V_{\bullet}^{2}+\varepsilon^{2}}-V_{\bullet}))). The pairs (Jr+,Jr)(J_{r}^{+},J_{r}^{-}) and (J+,J)(J_{\ell}^{+},J_{\ell}^{-}) form bases of the solution space. Each of them corresponds to one of the eigenvalues ±Vr2+ε2\pm\sqrt{V_{r}^{2}+\varepsilon^{2}} and ±V2+ε2\pm\sqrt{V_{\ell}^{2}+\varepsilon^{2}} of H(t,ε)H(t,\varepsilon) at the infinity. Note that a function ψ=(ψ1,ψ2)t\psi={}^{t}(\psi_{1},\psi_{2}) is a solution to (1.1) if and only if (ψ2¯,ψ1¯)t{}^{t}(-\overline{\psi_{2}},\overline{\psi_{1}}) is so. This implies that (Jr+(t),Jr(t))(J_{r}^{+}(t),J_{r}^{-}(t)) and (J+(t),J(t))(J_{\ell}^{+}(t),J_{\ell}^{-}(t)) are orthonormal bases on 2{\mathbb{C}}^{2} at each tt\in\mathbb{R}. Then we can introduce the scattering matrix S(ε,h)S(\varepsilon,h) as the change of basis between the pairs of Jost solutions:

(J+,J)=(Jr+,Jr)S(ε,h),S(ε,h)=(s11(ε,h)s12(ε,h)s21(ε,h)s22(ε,h)).\left(J_{\ell}^{+},J_{\ell}^{-}\right)=\left(J_{r}^{+},J_{r}^{-}\right)S(\varepsilon,h),\quad S(\varepsilon,h)=\left(\begin{array}[]{cc}s_{11}(\varepsilon,h)&s_{12}(\varepsilon,h)\\ s_{21}(\varepsilon,h)&s_{22}(\varepsilon,h)\end{array}\right). (2.2)

This matrix is unitary. In particular, one has |s11|=|s22||s_{11}|=|s_{22}|, |s12|=|s21||s_{12}|=|s_{21}|, and |s11|2+|s21|2=1|s_{11}|^{2}+|s_{21}|^{2}=1.

Definition 2.1.

The transition probability P(ε,h)P(\varepsilon,h) is defined by

P(ε,h):=|s21(ε,h)|2.P(\varepsilon,h):=|s_{21}(\varepsilon,h)|^{2}.
Remark 2.2.

The above definition of the transition probability is equivalent to (1.3). In fact, one has J±(t)2=1\|J_{\bullet}^{\pm}(t)\|_{\mathbb{C}^{2}}=1 for any tt, and

limtΠ±J±(t)2=1,\displaystyle\lim_{t\to-\infty}\left\|\Pi_{\pm}J_{\ell}^{\pm}(t)\right\|_{\mathbb{C}^{2}}=1,\quad limtΠJ±(t)2=0,\displaystyle\lim_{t\to-\infty}\left\|\Pi_{\mp}J_{\ell}^{\pm}(t)\right\|_{\mathbb{C}^{2}}=0,
limt+Π±Jr±(t)2=1,\displaystyle\lim_{t\to+\infty}\left\|\Pi_{\pm}J_{r}^{\pm}(t)\right\|_{\mathbb{C}^{2}}=1,\quad limt+ΠJr±(t)2=0.\displaystyle\lim_{t\to+\infty}\left\|\Pi_{\mp}J_{r}^{\pm}(t)\right\|_{\mathbb{C}^{2}}=0.
Condition B.

The function V(t)V(t) has a finite number of zeros t1>>tnt_{1}>\cdots>t_{n} on \mathbb{R}, where each zero tkt_{k} for k=1,nk=1,\ldots n is of finite order denoted by mkm_{k}.

This assumption implies that for k=1,,nk=1,\ldots,n,

V(l)(tk)=0(1l<mk),vk:=V(mk)(tk)0.V^{(l)}(t_{k})=0\quad(1\leq l<m_{k}),\quad v_{k}:=V^{(m_{k})}(t_{k})\neq 0. (2.3)

Let mm_{*} denote the maximal order of the zeros:

m=maxj{1,2,,n}mjm_{*}=\max_{j\in\{1,2,\ldots,n\}}m_{j} (2.4)

and let Λ\Lambda_{*} denote the index set of k{1,2,,n}k\in\{1,2,\ldots,n\} which attains mm_{*} (i.e., mk=mkΛm_{k}=m_{*}\iff k\in\Lambda_{*}). Put

σk:=j=1kmj\sigma_{k}:=\sum_{j=1}^{k}m_{j} (2.5)

for k=1,2,,nk=1,2,\ldots,n. Then Vr=limt+V(t)>0V_{r}=\lim_{t\to+\infty}V(t)>0 implies that σk\sigma_{k} determines the sign of V(t)V(t) on each interval (tk+1,tk)(t_{k+1},t_{k}), and in particular σn\sigma_{n} determines the sign of VV_{\ell}, namely (1)σkV(t)>0(-1)^{\sigma_{k}}V(t)>0 for tk+1<t<tkt_{k+1}<t<t_{k} and (1)σnV>0(-1)^{\sigma_{n}}V_{\ell}>0.

As we mentioned in the introduction, the ratio of ε\varepsilon and (a specific power of) hh is crucial. We set

μ:=μm,\mu_{*}:=\mu_{m_{*}}, (2.6)

where

μm=μm(ε,h):=εhmm+1\mu_{m}=\mu_{m}(\varepsilon,h):=\varepsilon h^{-\frac{m}{m+1}} (2.7)

for each mm\in\mathbb{N}. We focus on the regime μ1\mu_{*}\ll 1. In the case where there exists at least one avoided crossing generated by a tangential intersection, that is m2m_{*}\geq 2, we obtain the following result.

Refer to caption
Figure 1: An example of V(t)V(t) and energies E±(ε,h)E_{\pm}(\varepsilon,h)
Theorem 1.

Assume Conditions A, B and m2m_{*}\geq 2. Then there exist μ0>0\mu_{0}>0 and h0>0h_{0}>0 such that, for any ε\varepsilon and hh with μ(ε,h)(0,μ0]\mu_{*}(\varepsilon,h)\in(0,\mu_{0}] and h(0,h0]h\in(0,h_{0}], the transition probability P(ε,h)P(\varepsilon,h) has the asymptotic expansions:

P(ε,h)={1C(h)μ2+𝒪(μ2(μ+h1m(m+1)))ifσnis odd,C(h)μ2+𝒪(μ2(μ+h1m(m+1)))ifσnis even,P(\varepsilon,h)=\left\{\begin{array}[]{lll}1-C_{*}(h)\,\mu_{*}^{2}&+&{\mathcal{O}}\left(\mu_{*}^{2}\left(\mu_{*}+h^{\frac{1}{m_{*}(m_{*}+1)}}\right)\right)\qquad\textrm{if}\ \sigma_{n}\ \textrm{is odd},\\[12.0pt] {\quad}C_{*}(h)\,\mu_{*}^{2}&+&{\mathcal{O}}\left(\mu_{*}^{2}\left(\mu_{*}+h^{\frac{1}{m_{*}(m_{*}+1)}}\right)\right)\qquad\textrm{if}\ \sigma_{n}\ \textrm{is even},\end{array}\right.

where the coefficient C(h)C_{*}(h) consists of the product of two factors γ\gamma_{*} and δ(h)\delta_{*}(h), that is C(h)=γδ(h)C_{*}(h)=\gamma_{*}\delta_{*}(h), which are given by

γ=4((m+1)!2)2m+1Γ(m+2m+1)2(11+(1)m2sin2(π2(m+1))),\displaystyle\gamma_{*}=4\left(\frac{(m_{*}+1)!}{2}\right)^{\frac{2}{m_{*}+1}}\!\!\varGamma\left(\frac{m_{*}+2}{m_{*}+1}\right)^{2}\!\!\left(1-\frac{1+(-1)^{m_{*}}}{2}\sin^{2}\left(\frac{\pi}{2(m_{*}+1)}\right)\right),
δ(h)=jΛ|vj|2m+1+2j,kΛj<k|vjvk|1m+1cos(2htktjV(t)𝑑t+θmj,k),\displaystyle\delta_{*}(h)=\sum_{j\in\Lambda_{*}}|v_{j}|^{-\frac{2}{m_{*}+1}}+2\!\!\sum_{\begin{subarray}{c}j,k\in\Lambda_{*}\\ j<k\end{subarray}}\!\!|v_{j}v_{k}|^{-\frac{1}{m_{*}+1}}\,\cos\left(\frac{2}{h}\int_{t_{k}}^{t_{j}}V(t)dt+\theta_{m_{*}}^{j,k}\right), (2.8)

with

θmj,k={(sgnvj)πm+1if m is odd and sgnvj=sgnvk,0otherwise.\theta_{m_{*}}^{j,k}=\left\{\begin{aligned} &(\operatorname{sgn}v_{j})\frac{\pi}{m_{*}+1}&&\text{if $m_{*}$ is odd and }\ \operatorname{sgn}v_{j}=-\operatorname{sgn}v_{k},\\ &0&&\text{otherwise}.\end{aligned}\right.

Here, Γ\varGamma stands for the standard Gamma function Γ(z)=0+tz1et𝑑t\varGamma(z)=\int_{0}^{+\infty}t^{z-1}e^{-t}dt.

Remark 2.3.

When every avoided crossing is generated by a transversal intersection, that is, m=1m_{*}=1, Theorem 1 is proven in [22] under an additional assumption that VV is analytic near the real line. Our method also deduces the same asymptotic formula under Conditions A, B and the additional condition that μ~1:=(log(1/h))1/2εh1/2\tilde{\mu}_{1}:=(\log(1/h))^{{1/2}}\varepsilon h^{-1/2}, replaced with μ1\mu_{1}, is sufficiently small (see also [10, Remark 1.2]).

Remark 2.4.

The factor γ\gamma_{*} depends only on the highest order mm_{*} of the zeros and never vanishes while the factor δ(h)\delta_{*}(h) depends also on the behavior of VV not only the local property at zeros and may vanish. This vanishing phenomenon corresponds to the destructive quantum interference. Suppose, for example, that |vj||v_{j}| among jΛj\in\Lambda_{*} are the same. Put N:=#ΛN_{*}:=\#\Lambda_{*} and n:=minΛn_{*}:=\min\Lambda_{*}. Then the condition for δ(h)\delta_{*}(h) to vanish is given by

N+2(jΛ{n}cos𝒱j+j,kΛ{n}j<kcos(𝒱j𝒱k))=0,N_{*}+2\left(\sum_{j\in\Lambda_{*}\setminus\{n_{*}\}}\cos{\mathcal{V}}_{j}+\sum_{\begin{subarray}{c}j,k\in\Lambda_{*}\setminus\{n_{*}\}\\ j<k\end{subarray}}\cos({\mathcal{V}}_{j}-{\mathcal{V}}_{k})\right)=0, (2.9)

where

𝒱j:=2htntjV(t)𝑑t+1(1)m2(sgnvj)π2(m+1).{\mathcal{V}}_{j}:=\frac{2}{h}\int_{t_{n_{*}}}^{t_{j}}V(t)dt+\frac{1-(-1)^{m_{*}}}{2}(\operatorname{sgn}v_{j})\frac{\pi}{2(m_{*}+1)}. (2.10)

The algebraic curve (2.9) in (N1)(N_{*}-1)-variables {𝒱j}jΛ{n}\{{\mathcal{V}}_{j}\}_{j\in\Lambda_{*}\setminus\{n_{*}\}} appears as so-called Fermi surface in the context of the discrete Laplacian on the (N1)(N_{*}-1)-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 k=1,2,,n1k=1,2,\ldots,n-1, we denote the area enclosed by V(t)V(t) and V(t)-V(t) between tk+1t_{k+1} and tkt_{k} by

𝒜k:=2tk+1tk|V(t)|𝑑t.\mathcal{A}_{k}:=2\int_{t_{k+1}}^{t_{k}}\left|V(t)\right|dt. (2.11)
Refer to caption
Figure 2: Cases (b) (left) and (c) (right) in Example 2.5
Example 2.5 (Two avoided crossings).

Let the number nn of avoided crossings be two. Then the transition probability P(ε,h)P(\varepsilon,h) is 11 (resp. 0) modulo 𝒪(μ2){\mathcal{O}}(\mu_{*}^{2}) if the sum σ2=m1+m2\sigma_{2}=m_{1}+m_{2} of the order of zeros is odd (resp. even). In particular, when the two zeros have the same order, one sees that P(ε,h)=𝒪(μ2)P(\varepsilon,h)={\mathcal{O}}(\mu_{*}^{2}) independent of the parity of the order. We give the coefficient C(h)C_{*}(h) attached to μ2\mu_{*}^{2} in each situation:

  1. (a).

    m1>m2m_{1}>m_{2};

    C(h)=γm1|v1|2m1+1.\displaystyle C_{*}(h)=\gamma_{m_{1}}|v_{1}|^{-\frac{2}{m_{1}+1}}. (2.12)
  2. (b).

    m1=m221m_{1}=m_{2}\in 2\mathbb{Z}-1 and |v1|=|v2||v_{1}|=|v_{2}|;

    C(h)=4γm1|v1|2m1+1cos2(𝒜12hπ2(m1+1)).C_{*}(h)=4\gamma_{m_{1}}|v_{1}|^{-\frac{2}{m_{1}+1}}\cos^{2}\left(\frac{\mathcal{A}_{1}}{2h}-\frac{\pi}{2(m_{1}+1)}\right). (2.13)
  3. (c).

    m1=m22m_{1}=m_{2}\in 2\mathbb{Z} and |v1|=|v2||v_{1}|=|v_{2}|;

    C(h)=4γm1|v1|2m1+1cos2𝒜12h.C_{*}(h)=4\gamma_{m_{1}}|v_{1}|^{-\frac{2}{m_{1}+1}}\cos^{2}\frac{\mathcal{A}_{1}}{2h}. (2.14)
Remark 2.6.

In Cases (b) and (c) of Example 2.5, we see that C(h)C_{*}(h) may vanish and the order of the transition probability varies due to the destructive quantum interference under the Bohr-Sommerfeld type quantization rule

{𝒜1h+m1πm1+12πCase (b),𝒜1h+π2πCase (c).\left\{\begin{aligned} &\frac{{\mathcal{A}}_{1}}{h}+\frac{m_{1}\pi}{m_{1}+1}\in 2\pi\mathbb{Z}&&\text{Case {\rm(b)}},\\ &\frac{{\mathcal{A}}_{1}}{h}+\pi\in 2\pi\mathbb{Z}&&\text{Case {\rm(c)}}.\end{aligned}\right. (2.15)

This condition is a generalization of that shown in [22] (for m1=1m_{1}=1).

Example 2.7 (Three avoided crossings).

Let n=3n=3. The transition probability is determined modulo 𝒪(μ2){\mathcal{O}}(\mu_{*}^{2}) by the sum (m1+m2+m3)(m_{1}+m_{2}+m_{3}) whereas the coefficient C(h)C_{*}(h) attached to μ2\mu_{*}^{2} is determined by zeros tjt_{j} only for jΛj\in\Lambda_{*} and by integrals of VV between them. In particular, when #Λ2\#\Lambda_{*}\leq 2 and Λ{1,3}\Lambda_{*}\neq\{1,3\}, the coefficient C(h)C_{*}(h) is given by the same formula as a model with two avoided crossings.

  1. (a).

    Λ={1,3}\Lambda_{*}=\{1,3\} and |v1|=|v2|\left|v_{1}\right|=\left|v_{2}\right|;

    C(h)=4γm1|v1|2m1+1cos2(𝒜1+(1)m2𝒜2h).C_{*}(h)=4\gamma_{m_{1}}\left|v_{1}\right|^{-\frac{2}{m_{1}+1}}\cos^{2}\left(\frac{\mathcal{A}_{1}+(-1)^{m_{2}}\mathcal{A}_{2}}{h}\right). (2.16)
  2. (b).

    m1=m2=m321m_{1}=m_{2}=m_{3}\in 2\mathbb{Z}-1 and |v1|=|v2|=|v3||v_{1}|=|v_{2}|=|v_{3}|;

    C(h)\displaystyle C_{*}(h) =γm1|v1|2m1+1[3+2(cos(𝒜1hπm1+1)\displaystyle=\gamma_{m_{1}}|v_{1}|^{-\frac{2}{m_{1}+1}}\Bigl{[}3+2\Bigl{(}\cos\Bigl{(}\frac{{\mathcal{A}}_{1}}{h}-\frac{\pi}{m_{1}+1}\Bigr{)} (2.17)
    +cos(𝒜2hπm1+1)+cos(𝒜1𝒜2h))].\displaystyle\quad\qquad\qquad\qquad+\cos\Bigl{(}\frac{{\mathcal{A}}_{2}}{h}-\frac{\pi}{m_{1}+1}\Bigr{)}+\cos\Bigl{(}\frac{\mathcal{A}_{1}-\mathcal{A}_{2}}{h}\Bigr{)}\Bigr{)}\Bigr{]}.
  3. (c).

    m1=m2=m32m_{1}=m_{2}=m_{3}\in 2\mathbb{Z} and |v1|=|v2|=|v3||v_{1}|=|v_{2}|=|v_{3}|;

    C(h)=γm1|v1|2m1+1[3+2(cos𝒜1h+cos𝒜2h+cos(𝒜1+𝒜2h))].\displaystyle C_{*}(h)=\gamma_{m_{1}}|v_{1}|^{-\frac{2}{m_{1}+1}}\left[3+2\left(\cos\frac{{\mathcal{A}}_{1}}{h}+\cos\frac{{\mathcal{A}}_{2}}{h}+\cos\left(\frac{\mathcal{A}_{1}+\mathcal{A}_{2}}{h}\right)\right)\right].
Remark 2.8.

While the destructive quantum interference condition in the case n=2n=2 is that the area on the phase space is quantized (i.e. discretized) as in (2.15), that condition in n=3n=3 is that two areas lie along the Fermi curve.

Refer to caption
Refer to caption
Figure 3: Cases b (above) and c (below) in Example 2.7

2.2 Coexistence of the two parameter regimes

Refer to caption
Figure 4: Adiabatic and non-adiabatic regimes (A)m(A)_{m} and (N)m(N)_{m} for m=3,11m=3,11 (logarithmic scale, 10150ε,h110^{-150}\leq\varepsilon,h\leq 1, (A)m={(ε,h);μm100}(A)_{m}=\{(\varepsilon,h);\,\mu_{m}\geq 100\}, (N)m={(ε,h); 0<μm0.01}(N)_{m}=\{(\varepsilon,h);\,0<\mu_{m}\leq 0.01\}).

Recall that the quantum dynamics around each avoided crossing near t=tkt=t_{k} depends principally on the magnitude of the parameter μmk\mu_{m_{k}}. More precisely, μmk1\mu_{m_{k}}\ll 1 and μmk1\mu_{m_{k}}\gg 1 correspond to the non-adiabatic and adiabatic regimes (note that the regime μmk1\mu_{m_{k}}\sim 1 is studied only for the transversal case mk=1m_{k}=1 [3]). This parameter is different for two zeros of V(t)V(t) with different order, thus the transition problem with several avoided crossings generated by tangential intersections admits various regimes.

Note that μm\mu_{m} obeys the algebraic order relation:

m<mμm<μm.m<m^{\prime}\iff\mu_{m}<\mu_{m^{\prime}}. (2.18)

The regime μm1\mu_{m_{*}}\ll 1 considered in Theorem 1 corresponds to non-adiabatic regime μmk1\mu_{m_{k}}\ll 1 for every k{1,,n}k\in\{1,\ldots,n\}. Conversely, the regime μm1\mu_{m_{\circledast}}\gg 1 (with mm_{\circledast} standing for the minimum order mink{1,,n}mk\min_{k\in\{1,\ldots,n\}}m_{k}) considered in [20] corresponds to adiabatic regime μmk1\mu_{m_{k}}\gg 1 for every kk.

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

{1,2,,n}=Λ¯Λ¯\{1,2,\ldots,n\}=\overline{\Lambda_{\sharp}}\sqcup\overline{\Lambda_{\flat}}

such that

μmk1(kΛ¯),μmk1(kΛ¯).\mu_{m_{k}}\gg 1\quad(\forall k\in\overline{\Lambda_{\sharp}}),\qquad\mu_{m_{k}}\ll 1\quad(\forall k\in\overline{\Lambda_{\flat}}).

Again by (2.18), this corresponds to

μ:=μm1,μ:=μm1,\mu_{\sharp}:=\mu_{m_{\sharp}}\gg 1,\quad\mu_{\flat}:=\mu_{m_{\flat}}\ll 1,

where we put m:=maxkΛ¯mkm_{\flat}:=\max_{k\in\overline{\Lambda_{\flat}}}m_{k} and m:=minkΛ¯mkm_{\sharp}:=\min_{k\in\overline{\Lambda_{\sharp}}}m_{k}. We also put

Λ:={k;mk=m},Λ:={k;mk=m}.\Lambda_{\flat}:=\{k;\,m_{k}=m_{\flat}\},\quad\Lambda_{\sharp}:=\{k;\,m_{k}=m_{\sharp}\}.

Figure 4 illustrates the regimes for m=3,11m=3,11. When each zero of VV is either of order 3 or 11, we here study the regime (N)3(A)11(N)_{3}\cap(A)_{11} while Theorem 1 and [20] concern the regime (N)11(N)_{11} and (A)3(A)_{3}, respectively. In Figure 5, the problem here corresponds to (N)1(A)2(N)_{1}\cap(A)_{2} or (N)2(A)3(N)_{2}\cap(A)_{3}. Note also that these figures are displayed with a logarithmic scale. Hence the borders between regimes are straight lines. Indeed, the border μm=c\mu_{m}=c for some c>0c>0 is rewritten as logε=logc+mm+1logh\log\varepsilon=\log c+\frac{m}{m+1}\log h.

Refer to caption
Figure 5: Adiabatic and non-adiabatic regimes (A)m(A)_{m} and (N)m(N)_{m} for m=1,2,3m=1,2,3.

In the study of adiabatic regime, one of the authors employed the exact-WKB method [20] which requires the function VV to be analytic. Hence we also suppose the additional condition.

Condition C.

V(t)V(t) is real-analytic on an interval containing [tn,t1][t_{n},t_{1}].

Under this condition, when ε\varepsilon is small enough, there exist 2mk2m_{k} zeros of V(t)2+ε2V(t)^{2}+\varepsilon^{2} near each t=tkt=t_{k} 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 ζk,1(ε),ζk,mk(ε)\zeta_{k,1}(\varepsilon),\zeta_{k,m_{k}}(\varepsilon), which behave like

ζk,j(ε)tk+(mk!vkε)1/mkexp[2j12mkπi]\zeta_{k,j}(\varepsilon)\sim t_{k}+\left(\frac{m_{k}!}{v_{k}}\varepsilon\right)^{1/m_{k}}\exp\left[\frac{2j-1}{2m_{k}}\pi i\right]

as ε0\varepsilon\to 0. As in [22], the action integral Ak,j(ε)A_{k,j}(\varepsilon) for j=1,mkj=1,m_{k} is given by

Ak,j:=2tkζk,j(ε)V(t)2+ε2𝑑t,A_{k,j}:=2\int_{t_{k}}^{\zeta_{k,j}(\varepsilon)}\sqrt{V(t)^{2}+\varepsilon^{2}}\,dt,

where the path is the segment from tkt_{k} to ζk,j(ε)\zeta_{k,j}(\varepsilon) and the branch of the square root of the integrand is ε\varepsilon at t=tkt=t_{k}. Note that ImAk,j>0{\rm Im}\,A_{k,j}>0 on this branch and there exist ak,j>0a_{k,j}>0 such that

ImAk,j=ak,jε(mk+1)/mk+𝒪(ε(mk+2)/mk){\rm Im}\,A_{k,j}=a_{k,j}\varepsilon^{(m_{k}+1)/m_{k}}+{\mathcal{O}}\left(\varepsilon^{(m_{k}+2)/m_{k}}\right)

as ε0\varepsilon\to 0.

Roughly speaking, the absolute value of the “probability amplitude of the transition around an avoided crossing near tkt_{k}” is small in the limit μmk+\mu_{m_{k}}\to+\infty. Contrary to the non-adiabatic case Λ\Lambda_{\flat}, this fact is independent of the parity of mkm_{k}. The probability amplitude has the same order as

exp[akμmk(mk+1)/mk]1,ak:=min{Imak,1,Imak,mk}.\exp\left[-a_{k}\mu_{m_{k}}^{(m_{k}+1)/m_{k}}\right]\ll 1,\quad a_{k}:=\min\{{\rm Im}\hskip 1.0pta_{k,1},{\rm Im}\hskip 1.0pta_{k,m_{k}}\}.

From the sake of distinguishing this difference, we introduce

Λodd¯={kΛ¯;mk:odd}={k(1),k(2),,k(N)},\displaystyle\overline{\Lambda_{\sharp}^{\rm odd}}=\{k\in\overline{\Lambda_{\sharp}}\,;\,m_{k}:\,\text{odd}\}=\{k(1),k(2),\ldots,k(N)\},

where N=#Λodd¯N=\#\overline{\Lambda_{\sharp}^{\rm odd}} stands for the number of the elements of Λodd¯\overline{\Lambda_{\sharp}^{\rm odd}}, and the elements are labeled in ascending order k(1)<k(2)<<k(N)k(1)<k(2)<\cdots<k(N).

We also introduce the effective energy V~(t)=V~(t;m,m)\tilde{V}(t)=\tilde{V}(t;m_{\flat},m_{\sharp}) in this regime by

V~(t)={V(t)on (tk(2l1),tk(2l)),V(t)otherwise,\tilde{V}(t)=\left\{\begin{aligned} -V(t)&\quad\text{on $(t_{k(2l-1)},t_{k(2l)})$},\\ V(t)&\quad\text{otherwise},\end{aligned}\right. (2.19)

where tk(2l)t_{k(2l)} is taken as -\infty when tk(2l1)=tk(N)t_{k(2l-1)}=t_{k(N)} (see also (2.21)).

Putting a:=minkΛaka:=\min_{k\in\Lambda_{\sharp}}a_{k} and introducing two functions

ϵ1=ϵ1(m,m,a)\displaystyle\epsilon_{1}=\epsilon_{1}(m_{\flat},m_{\sharp},a) =μ+exp[aμ(m+1)/m],\displaystyle=\mu_{\flat}+\exp\left[-a\mu_{\sharp}^{(m_{\sharp}+1)/m_{\sharp}}\right],
ϵ2=ϵ2(m,m,a)\displaystyle\epsilon_{2}=\epsilon_{2}(m_{\flat},m_{\sharp},a) =μ(μ+h1m(m+1))+μ(m+1)/mexp[aμ(m+1)/m],\displaystyle=\mu_{\flat}\left(\mu_{\flat}+h^{\frac{1}{m_{\flat}(m_{\flat}+1)}}\right)+\mu_{\sharp}^{-(m_{\sharp}+1)/m_{\sharp}}\exp\left[-a\mu_{\sharp}^{(m_{\sharp}+1)/m_{\sharp}}\right],

we state the asymptotic expansion of the transition probability in this intermediate regime:

Theorem 2.

Assume Conditions A, B and C. Then there exist 0<μ0<10<\mu_{0}<1 and h0>0h_{0}>0 such that, for any ε\varepsilon and hh satisfying μ<μ0<μ01<μ\mu_{\flat}<\mu_{0}<\mu_{0}^{-1}<\mu_{\sharp}, and h(0,h0]h\in(0,h_{0}], the transition probability P(ε,h)P(\varepsilon,h) has the asymptotic expansions:

P(ε,h)={1(ε,h)+(ε,h)if(σn+N)is odd,(ε,h)+(ε,h)if(σn+N)is even,P(\varepsilon,h)=\left\{\begin{array}[]{lll}1-{\mathcal{L}}(\varepsilon,h)&+&{\mathcal{E}}(\varepsilon,h)\qquad\textrm{if}\ (\sigma_{n}+N)\ \textrm{is odd},\\[12.0pt] {\quad}{\mathcal{L}}(\varepsilon,h)&+&{\mathcal{E}}(\varepsilon,h)\qquad\textrm{if}\ (\sigma_{n}+N)\ \textrm{is even},\end{array}\right.

where the leading term (ε,h)=𝒪(ϵ12){\mathcal{L}}(\varepsilon,h)={\mathcal{O}}(\epsilon_{1}^{2}) and the error therm (ε,h)=𝒪(ϵ1ϵ2){\mathcal{E}}(\varepsilon,h)={\mathcal{O}}(\epsilon_{1}\epsilon_{2}).

Remark 2.9.

The parity which characterizes the transition probability depends not only on σn\sigma_{n} determined by VV but also on NN determined by the regime. This implies that the switch of P(ε,h)P(\varepsilon,h) occurs with changing the regime without doing the energy VV (see Figure 6).

As we mentioned in Section 2.2, Theorem 2 covers the range of the pair of the parameters (ε,h)(\varepsilon,h) included in the parameter regime determined by mm_{\flat} and mm_{\sharp}. Regarding ε\varepsilon as a function of hh like a one-parameter problem, we find the typical cases, which realize the intermediate regime μ\mu_{\sharp}\to\infty and μ0\mu_{\flat}\to 0.

Polynomial case:

If εhα\varepsilon\sim h^{\alpha} with

mm+1<α<mm+1,\frac{m_{\flat}}{m_{\flat}+1}<\alpha<\frac{m_{\sharp}}{m_{\sharp}+1},

the contribution coming from Λ\Lambda_{\sharp} is exponentially small.

Logarithmic case:

If ε=(hlog(1/hρ))m/(m+1)\varepsilon=(h\log(1/h^{\rho}))^{m_{\sharp}/(m_{\sharp}+1)} with some positive constant ρ\rho, the contribution coming from Λ\Lambda_{\sharp} must be taken into account, since exp[aμ(m+1)/m]=haρ\exp[-a\mu_{\sharp}^{(m_{\sharp}+1)/m_{\sharp}}]=h^{a\rho}.

In the former case, the leading term is similar to that in Theorem 1 and is given by

(ε,h)\displaystyle{\mathcal{L}}(\varepsilon,h) =μ2(jΛγ|vj+1|2m+1+2j,kΛj<kReCj,k(ε,h)cos[1htktjV~(t)𝑑t]),\displaystyle=\mu_{\flat}^{2}\left(\sum_{j\in\Lambda_{\flat}}\gamma_{\flat}|v_{j+1}|^{-\frac{2}{m_{\flat}+1}}+2\!\!\!\!\sum_{\begin{subarray}{c}j,k\in\Lambda_{\flat}\\ j<k\end{subarray}}\!\!\!\!{\rm Re}\,C_{j,k}^{\flat\flat}(\varepsilon,h)\cos\left[\frac{1}{h}\int_{t_{k}}^{t_{j}}\tilde{V}(t)dt\right]\right), (2.20)

where the factor Cj,k(ε,h)C_{j,k}^{\flat\flat}(\varepsilon,h) is of 𝒪(1){\mathcal{O}}(1) 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 (ε,h){\mathcal{L}}(\varepsilon,h) is of the form:

μ2(jΛγ|vj+1|2m+1+2j,kΛj<kReCj,k(ε,h)cos[1htktjV~(t)𝑑t])\displaystyle\mu_{\flat}^{2}\left(\sum_{j\in\Lambda_{\flat}}\gamma_{\flat}|v_{j+1}|^{-\frac{2}{m_{\flat}+1}}+2\!\!\!\!\sum_{\begin{subarray}{c}j,k\in\Lambda_{\flat}\\ j<k\end{subarray}}\!\!\!\!{\rm Re}\,C_{j,k}^{\flat\flat}(\varepsilon,h)\cos\left[\frac{1}{h}\int_{t_{k}}^{t_{j}}\tilde{V}(t)dt\right]\right)
+kΛexp[2akμ(m+1)/m]\displaystyle\quad+\sum_{k\in\Lambda_{\sharp}}\exp\left[-2a_{k}\mu_{\sharp}^{(m_{\sharp}+1)/m_{\sharp}}\right]
+2jΛ,kΛj<kReCj,k(ε,h)μexp[akμ(m+1)/m]cos[1htktjV~(t)𝑑t]\displaystyle\quad+2\!\!\!\!\sum_{\begin{subarray}{c}j\in\Lambda_{\flat},k\in\Lambda_{\sharp}\\ j<k\end{subarray}}\!\!\!\!{\rm Re}\,C_{j,k}^{\flat\sharp}(\varepsilon,h)\mu_{\flat}\exp\left[-a_{k}\mu_{\sharp}^{(m_{\sharp}+1)/m_{\sharp}}\right]\cos\left[\frac{1}{h}\int_{t_{k}}^{t_{j}}\tilde{V}(t)dt\right]
+2j,kΛj<kReCj,k(ε,h)exp[(aj+ak)μ(m+1)/m]cos[1htktjV~(t)𝑑t],\displaystyle\quad+2\!\!\!\sum_{\begin{subarray}{c}j,k\in\Lambda_{\sharp}\\ j<k\end{subarray}}\!\!\!{\rm Re}\,C_{j,k}^{\sharp\sharp}(\varepsilon,h)\exp\left[-(a_{j}+a_{k})\mu_{\sharp}^{(m_{\sharp}+1)/m_{\sharp}}\right]\cos\left[\frac{1}{h}\int_{t_{k}}^{t_{j}}\tilde{V}(t)dt\right],

where Cj,k(ε,h)C_{j,k}^{\flat\sharp}(\varepsilon,h) and Cj,k(ε,h)C_{j,k}^{\sharp\sharp}(\varepsilon,h) are of 𝒪(1){\mathcal{O}}(1) and referred in (5.17) and (5.18) respectively.

Remark 2.10.

The mixed terms coming from ϵ12\epsilon_{1}^{2} correspond to quantum interference terms referred in Remark 2.4. The phase shift caused by the integral of the energy VV changes into the phase shift done by that of the effective energy V~\tilde{V} as in Figure 6.

In our method, we represent the Jost solution J+J_{\ell}^{+} by several bases. As we mentioned in the introduction, each basis is consists of solutions corresponding to an eigenvector associated with E±(t,ε)E_{\pm}(t,\varepsilon) in each region between two avoided crossings. Consequently, the absolute value of the coefficients gives Π±(t;ε)J+2\|\Pi_{\pm}(t;\varepsilon)J_{\ell}^{+}\|_{\mathbb{C}^{2}}. Moreover, one observes from our proof that

1Πσ(t)(t;ε)J+(t)2>Πσ(t)(t;ε)J+(t)20,1\sim\|\Pi_{\sigma(t)}(t;\varepsilon)J_{\ell}^{+}(t)\|_{\mathbb{C}^{2}}>\|\Pi_{-\sigma(t)}(t;\varepsilon)J_{\ell}^{+}(t)\|_{\mathbb{C}^{2}}\sim 0, (2.21)

for any tt\in\mathbb{R}, where σ(t):=(1)σn+NsgnV~(t)\sigma(t):=(-1)^{\sigma_{n}+N}\operatorname{sgn}\tilde{V}(t). In this sense, (the square of) the modulus of the probability amplitude that the energy follows the curve τ=Eσ(t)(t;ε)\tau=E_{\sigma(t)}(t;\varepsilon) is almost 1. In Figure 6, we draw this curve in green.

Example 2.11.

Figure 6 shows an example case with mk=2k1m_{k}=2k-1 (k=1,2,3k=1,2,3).

Refer to caption
Figure 6: Change of the effective energy of Example 2.11 for each regime

3 Construction of exact solutions

In this section, we construct exact solutions which form a local basis near each vanishing point t=tkt=t_{k} by means of a method of successive approximations due to [6]. While the equation treated there is a second order 2×22\times 2 system of time-independent Schrödinger equations, our equation in this paper is a first order 2×22\times 2 system.

3.1 Estimates of fundamental solutions

For simplicity, we assume tk=0t_{k}=0, and let mm denote mkm_{k}. Let II be a small interval containing the vanishing point 0 in its interior. We fix

u±(t)=exp(ih0tV(s)𝑑s)u^{\pm}(t)=\exp\left(\mp\frac{i}{h}\int_{0}^{t}V(s)ds\right)

as a particular solution to

(hDt±V(t))u=0on I\left(hD_{t}\pm V(t)\right)u=0\quad\text{on }\ I (3.1)

respectively, where DtD_{t} stands for id/dt-id/dt. For any point aIa\in I, we define an integral operator Ka±K^{\pm}_{a} by

Ka±f(t):=ihu±(t)atf(s)u±(s)𝑑sfor fC(I),K^{\pm}_{a}f(t):=\frac{i}{h}u^{\pm}(t)\int_{a}^{t}\frac{f(s)}{u^{\pm}(s)}ds\quad\text{for }\ f\in C(I), (3.2)

where C(I)C(I) is the Banach space of continuous functions on II equipped with the norm fC(I):=supxI|f(x)|\|f\|_{C(I)}:=\sup_{x\in I}|f(x)|. Since

(hDt±V(t))Ka±f=ffor fC(I),\left(hD_{t}\pm V(t)\right)K^{\pm}_{a}f=f\quad\text{for }\ f\in C(I), (3.3)

for any aIa\in I, the integral operator Ka±:C(I)C(I)K^{\pm}_{a}:C(I)\to C(I) is well-defined as a fundamental solution of hDt±V(t)hD_{t}\pm V(t) respectively for the signs ±\pm.

Using these fundamental solutions Ka±±K^{\pm}_{a^{\pm}} with base points a±Ia^{\pm}\in I respectively, our equation (1.1) turns into the integral system with arbitrary constants c+,cc^{+},c^{-}\in\mathbb{C}:

{ψ1(t)=εKa++ψ2(t)+c+u+(t),ψ2(t)=εKaψ1(t)+cu(t).\left\{\begin{aligned} &\psi_{1}(t)=-\varepsilon K^{+}_{a^{+}}\psi_{2}(t)+c^{+}u^{+}(t),\\ &\psi_{2}(t)=-\varepsilon K^{-}_{a^{-}}\psi_{1}(t)+c^{-}u^{-}(t).\end{aligned}\right. (3.4)

Depending on the choice of the base points a+a^{+} and aa^{-}, the initial value for ψ1\psi_{1} and ψ2\psi_{2} at these points are determined:

ψ1(a+)=c+u+(a+),ψ2(a)=cu(a).\psi_{1}(a^{+})=c^{+}u^{+}(a^{+}),\quad\psi_{2}(a^{-})=c^{-}u^{-}(a^{-}).

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 ε\varepsilon, and this estimate gives the critical rate μm=εhm/(m+1)\mu_{m}=\varepsilon h^{-m/(m+1)}.

Let q\|\cdot\|_{q} for qq\in\mathbb{R} be a norm on the space of continuously differentiable functions C1(I)C^{1}(I) defined by

fq:=supI|f|+hqsupI|f|fC1(I).\left\|f\right\|_{q}:=\sup_{I}|f|+h^{q}\sup_{I}|f^{\prime}|\quad f\in C^{1}(I). (3.5)
Proposition 3.1.

For any a±Ia^{\pm}\in I, there exists C>0C>0 such that

(u±)1Ka±±(uf)1m+1Chmm+1f1m+1\left\|(u^{\pm})^{-1}K_{a^{\pm}}^{\pm}(u^{\mp}f)\right\|_{\frac{1}{m+1}}\leq Ch^{-\frac{m}{m+1}}\|f\|_{\frac{1}{m+1}} (3.6)

for h>0h>0 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 II\subset\mathbb{R}, consider the integral

I(h):=If(t)exp(2iht0tV(s)𝑑s)𝑑t,\mathcal{I}_{I}(h):=\int_{I}f(t)\exp\left(\frac{2i}{h}\int_{t_{0}}^{t}V(s)ds\right)dt, (3.7)

with a continuously differentiable function fC1(I)f\in C^{1}(I) possibly depending on hh. Then there exists a constant C>0C>0 independent of ff (but depending on VV) such that

|I(h)|ChsupI(|f|+|f|),\left|\mathcal{I}_{I}(h)\right|\leq Ch\sup_{I}(|f|+|f^{\prime}|), (3.8)

for h>0h>0 small enough when VV does not vanish on II. If t0t_{0} is the unique zero in II of VV, one has

|I(h)|C(h1m+1supI|f|+h2m+1supI|f|)=Ch1m+1f1m+1,\left|\mathcal{I}_{I}(h)\right|\leq C\left(h^{\frac{1}{m+1}}\sup_{I}|f|+h^{\frac{2}{m+1}}\sup_{I}|f^{\prime}|\right)=Ch^{\frac{1}{m+1}}\|f\|_{\frac{1}{m+1}}, (3.9)

where mm denotes the order of vanishing at t0t_{0}. Moreover if ff is independent of hh, we have

I(h)=f(t0)ωmh1m+1+𝒪(h2m+1).\mathcal{I}_{I}(h)=f(t_{0})\omega_{m}h^{\frac{1}{m+1}}+{\mathcal{O}}(h^{\frac{2}{m+1}}). (3.10)

Here, the constant ωm\omega_{m} is given by

ωm=2((m+1)!2|V(m)(t0)|)1m+1Γ(m+2m+1)ηm,\omega_{m}=2\left(\frac{(m+1)!}{2|V^{(m)}(t_{0})|}\right)^{\frac{1}{m+1}}\varGamma\left(\frac{m+2}{m+1}\right)\,\eta_{m}, (3.11)

with

ηm:={cos(π2(m+1))m:even,exp(sgn(V(m)(t0))iπ2(m+1))m:odd.\eta_{m}:=\left\{\begin{aligned} &\cos\left(\frac{\pi}{2(m+1)}\right)\quad&&m:\text{even},\\ &\exp\left(\frac{\operatorname{sgn}(V^{(m)}(t_{0}))i\pi}{2(m+1)}\right)\quad&&m:\text{odd}.\end{aligned}\right. (3.12)

Proof of Lemma 3.2.

Suppose that VV does not vanish on II, that is, a non-stationary case. Then we have for tIt\in I

h2iV(t)ddtexp(2iht0tV(s)𝑑s)=exp(2iht0tV(s)𝑑s).\frac{h}{2iV(t)}\frac{d}{dt}\exp\left(\frac{2i}{h}\int_{t_{0}}^{t}V(s)ds\right)=\exp\left(\frac{2i}{h}\int_{t_{0}}^{t}V(s)ds\right). (3.13)

This with an integration by parts and the compactness of II implies the estimate from above of |I(h)||\mathcal{I}_{I}(h)| by ChsupI(|f|+|f|)Ch\sup_{I}(|f|+|f^{\prime}|).

Suppose next that t0t_{0} is the unique zero of VV in II. Take a smooth cut-off function χ\chi such that χ(t)=1\chi(t)=1 for |tt0|<Ch1/(m+1)|t-t_{0}|<Ch^{1/(m+1)} and χ(t)=0\chi(t)=0 for |tt0|>2Ch1/(m+1)|t-t_{0}|>2Ch^{1/(m+1)}. On the support of 1χ1-\chi, one has the estimate

|ddt(f(t)2iV(t))|Ctm+1(supI|f|+tsupI|f|).\left|\frac{d}{dt}\left(\frac{f(t)}{2iV(t)}\right)\right|\leq\frac{C}{t^{m+1}}\left(\sup_{I}|f|+t\sup_{I}|f^{\prime}|\right).

This with a similar argument as above implies that contribution coming from the support of 1χ1-\chi to the integral is estimated by Ch1/(m+1)f1m+1Ch^{1/(m+1)}\|f\|_{\frac{1}{m+1}}. The other part is estimated by h1/(m+1)supI|f|h^{1/(m+1)}\sup_{I}|f| since the support of χ\chi is 𝒪(h1/(m+1)){\mathcal{O}}(h^{1/(m+1)}), and the estimate (3.9) follows.

We then suppose also that ff is independent of hh. Let g=g(t)g=g(t) be the smooth function defined near t0t_{0} such that

2t0tV(s)𝑑s=(tt0)m+1g(t),g(t)=2V(m)(t0)(m+1)!+𝒪(tt0).2\int_{t_{0}}^{t}V(s)ds=(t-t_{0})^{m+1}g(t),\quad g(t)=\frac{2V^{(m)}(t_{0})}{(m+1)!}+{\mathcal{O}}(t-t_{0}).

Take a smooth cut-off function χ\chi whose value is 1 near t0t_{0} and supported only on a small neighborhood where the change of the variable τ=(tt0)|g(t)|1/(m+1)\tau=(t-t_{0})|g(t)|^{1/(m+1)} is valid. Then one has

Iχ(t)f(t)exp(2iht0tV(s))𝑑s=|g(t0)|1m+1χ(t(τ))f~(τ)eσiτm+1/h𝑑τ\int_{I}\chi(t)f(t)\exp\left(\frac{2i}{h}\int_{t_{0}}^{t}V(s)\right)ds=|g(t_{0})|^{-\frac{1}{m+1}}\int_{\mathbb{R}}\chi(t(\tau))\tilde{f}(\tau)e^{\sigma i\tau^{m+1}/h}d\tau

with σ=sgng(t0)\sigma=\operatorname{sgn}g(t_{0}) and a smooth function f~=f~(τ)\tilde{f}=\tilde{f}(\tau) satisfying f~(0)=f(t0)\tilde{f}(0)=f(t_{0}). 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 1χ1-\chi.

Based on this lemma, let us prove Proposition 3.1.

Proof of Proposition 3.1.

For fC1(I)f\in C^{1}(I), we have by definition

(u±)1Ka±±(uf)(t)=iha±texp(2ih0sV(r)𝑑r)f(s)𝑑s.(u^{\pm})^{-1}K^{\pm}_{a^{\pm}}(u^{\mp}f)(t)=\frac{i}{h}\int_{a^{\pm}}^{t}\exp\left(\mp\frac{2i}{h}\int_{0}^{s}V(r)dr\right)f(s)ds.

According to (3.9) of Lemma 3.2, this integral is estimated by Ch1m+1f1m+1Ch^{\frac{1}{m+1}}\|f\|_{\frac{1}{m+1}}. For the derivative, we have

ddt[(u±)1Ka±±(uf)(t)]=ihexp(2ih0tV(r)𝑑r)f(t).\frac{d}{dt}\left[(u^{\pm})^{-1}K^{\pm}_{a^{\pm}}(u^{\mp}f)(t)\right]=\frac{i}{h}\exp\left(\mp\frac{2i}{h}\int_{0}^{t}V(r)dr\right)f(t).

This is clearly bounded by h1supI|f|h^{-1}\sup_{I}|f|. ∎

Remark 3.3.

The argument in the proof of Lemma 3.2 shows that the estimate (3.6) becomes better as 𝒪(h){\mathcal{O}}(h) if the integral interval does not contain t=0t=0.

3.2 Method of successive approximations (MSA)

From Proposition 3.1, it follows that for each (c+,c)t2{}^{t}(c^{+},c^{-})\in\mathbb{C}^{2}, 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 c1w1(t)+c2w2(t)c_{1}w_{1}(t)+c_{2}w_{2}(t) of the solutions w1(t)w_{1}(t) and w2(t)w_{2}(t) corresponding to the choices 𝒆1=(1,0)t\bm{e}_{1}={}^{t}(1,0) and 𝒆2=(0,1)t\bm{e}_{2}={}^{t}(0,1) for (c+,c)t{}^{t}(c^{+},c^{-}). Moreover, the solution can be constructed by MSA:

w1(t)=w1(t;a,a+):=(k0(ε2Ka++Ka)ku+εKak0(ε2Ka++Ka)ku+)w_{1}(t)=w_{1}(t;a^{-},a^{+}):=\begin{pmatrix}\displaystyle\sum_{k\geq 0}(\varepsilon^{2}K^{+}_{a^{+}}K^{-}_{a^{-}})^{k}u^{+}\\ \displaystyle-\varepsilon K^{-}_{a^{-}}\sum_{k\geq 0}(\varepsilon^{2}K^{+}_{a^{+}}K^{-}_{a^{-}})^{k}u^{+}\end{pmatrix} (3.14)
(resp.w2(t)=w2(t;a+,a):=(εKa++k0(ε2KaKa++)kuk0(ε2KaKa++)ku))\left(\text{resp.}\ w_{2}(t)=w_{2}(t;a^{+},a^{-}):=\begin{pmatrix}\displaystyle-\varepsilon K^{+}_{a^{+}}\sum_{k\geq 0}(\varepsilon^{2}K^{-}_{a^{-}}K^{+}_{a^{+}})^{k}u^{-}\\ \displaystyle\sum_{k\geq 0}(\varepsilon^{2}K^{-}_{a^{-}}K^{+}_{a^{+}})^{k}u^{-}\end{pmatrix}\right) (3.15)

in II for a fixed small εhmm+1\varepsilon h^{-\frac{m}{m+1}}. These iteration formulas imply that the solutions admit the asymptotic expansions when μm:=εhmm+10\mu_{m}:=\varepsilon h^{-\frac{m}{m+1}}\to 0 as follows:

w1(t;a,a+)\displaystyle w_{1}(t;a^{-},a^{+}) =(u+(t)+ε2Ka++Kau+(t)εKau+(t))+(𝒪(μm4)𝒪(μm3)),\displaystyle=\begin{pmatrix}u^{+}(t)+\displaystyle\varepsilon^{2}K^{+}_{a^{+}}K^{-}_{a^{-}}u^{+}(t)\\[5.0pt] \displaystyle-\varepsilon K^{-}_{a^{-}}u^{+}(t)\end{pmatrix}+\begin{pmatrix}\displaystyle{\mathcal{O}}(\mu_{m}^{4})\\[5.0pt] \displaystyle{\mathcal{O}}(\mu_{m}^{3})\end{pmatrix}, (3.16)
w2(t;a+,a)\displaystyle w_{2}(t;a^{+},a^{-}) =(εKa++u(t)u(t)+ε2KaKa++u(t))+(𝒪(μm3)𝒪(μm4)).\displaystyle=\begin{pmatrix}\displaystyle-\varepsilon K^{+}_{a^{+}}u^{-}(t)\\[5.0pt] u^{-}(t)+\displaystyle\varepsilon^{2}K^{-}_{a^{-}}K^{+}_{a^{+}}u^{-}(t)\end{pmatrix}+\begin{pmatrix}\displaystyle{\mathcal{O}}(\mu_{m}^{3})\\[5.0pt] \displaystyle{\mathcal{O}}(\mu_{m}^{4})\end{pmatrix}.

Notice that Proposition 3.1 allows us to choose arbitrarily the base points a±a^{\pm} of the fundamental solutions in this construction. As we mentioned in Remark 3.3, we obtain better asymptotic formulas

w1(t;a,a+)=(u+(t)+𝒪(ε2/h)𝒪(ε))on I{±t>0}w_{1}(t;a^{-},a^{+})=\begin{pmatrix}u^{+}(t)+{\mathcal{O}}(\varepsilon^{2}/h)\\ {\mathcal{O}}(\varepsilon)\end{pmatrix}\quad\text{on }\ I\cap\{\pm t>0\} (3.17)

if ±a>0\pm a^{-}>0, that is, the integral interval [a,t][a^{-},t] of KaK^{-}_{a^{-}} does not contain the zero, and likewise

w2(t;a+,a)=(𝒪(ε)u(t)+𝒪(ε2/h))on I{±t>0}w_{2}(t;a^{+},a^{-})=\begin{pmatrix}{\mathcal{O}}(\varepsilon)\\ u^{-}(t)+{\mathcal{O}}(\varepsilon^{2}/h)\end{pmatrix}\quad\text{on }\ I\cap\{\pm t>0\} (3.18)

if ±a+>0\pm a^{+}>0, that is, the integral interval [a+,t][a^{+},t] of Ka++K^{+}_{a^{+}} does not.


Take (ε,h)(\varepsilon,h)-independent constants rr and \ell such that <0<r\ell<0<r and [,r]I[\ell,r]\subset I. We define the four MSA solutions w1.r,w2,r,w1,w_{1.r},w_{2,r},w_{1,\ell}, and w2,w_{2,\ell} in II as

w1,r(t):=w1(t;r,r),\displaystyle w_{1,r}(t):=w_{1}(t;r,r), w2,r(t):=w2(t;r,r),\displaystyle w_{2,r}(t):=w_{2}(t;r,r), (3.19)
w1,(t):=w1(t;,),\displaystyle w_{1,\ell}(t):=w_{1}(t;\ell,\ell), w2,(t):=w2(t;,).\displaystyle w_{2,\ell}(t):=w_{2}(t;\ell,\ell).

According to Proposition 3.1 and the asymptotic formula (3.16), one sees that the asymptotic behaviors of these MSA solutions as μm0\mu_{m}\to 0 are clear on [,r][\ell,r], and also that wj,r(t)w_{j,r}(t) (resp. wj,(t)w_{j,\ell}(t)) (j=1,2)(j=1,2) behave like the initial data uj(t)u_{j}(t) near t=rt=r (resp. t=t=\ell) for a fixed small μm\mu_{m}. Moreover the pairs (w1,r(t),w2,r(t))(w_{1,r}(t),w_{2,r}(t)) and (w1,(t),w2,(t))(w_{1,\ell}(t),w_{2,\ell}(t)) form bases of the space of solutions.

4 Connection formulas

In this section, we also treat a zero of V(t)V(t) as tk=0t_{k}=0 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 (w1,r(t),w2,r(t))(w_{1,r}(t),w_{2,r}(t)) and (w1,(t),w2,(t))(w_{1,\ell}(t),w_{2,\ell}(t)) introduced by (3.19) in the previous section. The claim of this subsection is the asymptotic behavior of the transfer matrix T(ε,h)T(\varepsilon,h) as follows:

Theorem 3.

For the solutions defined by (3.19), we have

(w1,(t)w2,(t))=(w1,r(t)w2,r(t))T(ε,h),\begin{pmatrix}w_{1,\ell}(t)&w_{2,\ell}(t)\end{pmatrix}=\begin{pmatrix}w_{1,r}(t)&w_{2,r}(t)\end{pmatrix}T(\varepsilon,h), (4.1)

where the 2×22\times 2-matrix T=T(ε,h)T=T(\varepsilon,h) admits the following asymptotic formula:

T(ε,h)=I2iμmTsub+𝒪(μm2+μmh1m+1)T(\varepsilon,h)=I_{2}-i\mu_{m}T_{\operatorname{sub}}+{\mathcal{O}}(\mu_{m}^{2}+\mu_{m}h^{\frac{1}{m+1}}) (4.2)

as (ε,h)(0,0)(\varepsilon,h)\to(0,0) with μm=εhmm+10\mu_{m}=\varepsilon h^{-\frac{m}{m+1}}\to 0 and

Tsub=(0ωmω¯m0),T_{\operatorname{sub}}=\begin{pmatrix}0&\omega_{m}\\ \overline{\omega}_{m}&0\end{pmatrix}, (4.3)

where ωm\omega_{m} is given by (3.11) with t0=0t_{0}=0.

Remark 4.1.

By construction and the symmetry Ka±f¯=Kaf¯\overline{K_{a}^{\pm}f}=-K_{a}^{\mp}\overline{f}, we have

(0110)w2,¯=w1,(=r,).\begin{pmatrix}0&1\\ -1&0\end{pmatrix}\overline{w_{2,\bullet}}=w_{1,\bullet}\quad(\bullet=r,\ell). (4.4)

This makes TT symmetrical in the following sense:

T=(τ1τ2τ2¯τ1¯)SU(2),T=\begin{pmatrix}\tau_{1}&-\tau_{2}\\ \overline{\tau_{2}}&\overline{\tau_{1}}\end{pmatrix}\in\mathrm{\,SU}(2), (4.5)

where SU(2)\mathrm{\,SU}(2) is the special unitary group of degree 22 (see §A.2). Namely, TT is unitary (|τ1|2+|τ2|2=1)(|\tau_{1}|^{2}+|\tau_{2}|^{2}=1). This is a consequence that the time evolution by HH is unitary and that for =,r\bullet=\ell,r, the basis

(w1,(),w2,())=((u+()0),(0u()))(w_{1,\bullet}(\bullet),w_{2,\bullet}(\bullet))=\left(\begin{pmatrix}u^{+}(\bullet)\\ 0\end{pmatrix},\begin{pmatrix}0\\ u^{-}(\bullet)\end{pmatrix}\right)

of 2\mathbb{C}^{2} is orthonormal.

Proof of Theorem 3.

Since the pair (w1,r,w2,r)(w_{1,r},\,w_{2,r}) forms a basis of the space of solutions, the matrix (w1,r(t)w2,r(t))(w_{1,r}(t)\,w_{2,r}(t)) is invertible for any tIt\in I. Then (4.1) is rewritten as

T(ε,h)=(w1,r(t)w2,r(t))1(w1,(t)w2,(t)),T(\varepsilon,h)=(w_{1,r}(t)\,w_{2,r}(t))^{-1}(w_{1,\ell}(t)\,w_{2,\ell}(t)),

where the right-hand side is also independent of tt. Substituting rr for tt, we have

T(ε,h)\displaystyle T(\varepsilon,h) =(u(r)00u+(r))(w1,(r)w2,(r))\displaystyle=\begin{pmatrix}u^{-}(r)&0\\ 0&u^{+}(r)\end{pmatrix}(w_{1,\ell}(r)\,w_{2,\ell}(r)) (4.6)
=Id(0u(r)εK+u(r)u+(r)εKu+(r)0)+𝒪(μm2).\displaystyle=\mathrm{Id}-\begin{pmatrix}0&u^{-}(r)\varepsilon K_{\ell}^{+}u^{-}(r)\\ u^{+}(r)\varepsilon K_{\ell}^{-}u^{+}(r)&0\end{pmatrix}+{\mathcal{O}}(\mu_{m}^{2}).

Here, we have used u+u=1u^{+}u^{-}=1 and (3.16).

In order to prove Theorem 3, it is enough to compute the asymptotic behavior of the quantity u(r)εK+u(r)u^{-}(r)\varepsilon K_{\ell}^{+}u^{-}(r) as (ε,h)(0,0)(\varepsilon,h)\to(0,0) with μm0\mu_{m}\to 0. The computation of

u(r)εK+u(r)\displaystyle u^{-}(r)\varepsilon K_{\ell}^{+}u^{-}(r) =iεhrexp(2ih0tV(s)𝑑s)𝑑t\displaystyle=\frac{i\varepsilon}{h}\int_{\ell}^{r}\exp\left(\frac{2i}{h}\int_{0}^{t}V(s)ds\right)\,dt (4.7)

is carried out based on Lemma 3.2. In fact, since the integral interval [,r][\ell,r] of (4.7) contains the zero of V(t)V(t), we have

u(r)εK+u(r)\displaystyle u^{-}(r)\varepsilon K_{\ell}^{+}u^{-}(r) =iεh(ωmh1m+1+𝒪(h2m+1))\displaystyle=-\frac{i\varepsilon}{h}\left(\omega_{m}h^{\frac{1}{m+1}}+{\mathcal{O}}(h^{\frac{2}{m+1}})\right)
=iμmωm+𝒪(μmh1m+1).\displaystyle=-i\mu_{m}\omega_{m}+{\mathcal{O}}(\mu_{m}h^{\frac{1}{m+1}}). (4.8)

Combining (4.6) and (4.8), we obtain Theorem 3.

Hence this theorem implies that the transfer matrix TkT_{k} in (5.1) is given by

Tk(ε,h)=(1iωmkμmkiω¯mkμmk1)+𝒪(μmk2+μmkh1mk+1)T_{k}(\varepsilon,h)=\begin{pmatrix}1&-i\omega_{m_{k}}\mu_{m_{k}}\\ -i\overline{\omega}_{m_{k}}\mu_{m_{k}}&1\end{pmatrix}+{\mathcal{O}}(\mu_{m_{k}}^{2}+\mu_{m_{k}}h^{\frac{1}{m_{k}+1}}) (4.9)

as (ε,h)(0,0)(\varepsilon,h)\to(0,0) with μmk=εhmkmk+10\mu_{m_{k}}=\varepsilon h^{-\frac{m_{k}}{m_{k}+1}}\to 0.

4.2 Between the vanishing points

In addition to Theorem 3, which gives the connection formula around the vanishing point of V(t)V(t), a similar argument yields the following proposition, which gives the connection formula between two consecutive zeros of V(t)V(t).

Let tk+1<tkt_{k+1}<t_{k} be two consecutive zeros of V(t)V(t) (i.e. V0V\neq 0 on ]tk+1,tk[]t_{k+1},t_{k}[) with multiplicities mk,mk+1m_{k},m_{k+1}, and let j,rj\ell_{j},r_{j} (j=k,k+1)(j=k,k+1) be base points such that k+1<tk+1<rk+1<k<tk<rk\ell_{k+1}<t_{k+1}<r_{k+1}<\ell_{k}<t_{k}<r_{k}. The configuration of these points implies that each interval Ij:=[j,rj]I_{j}:=[\ell_{j},r_{j}] includes tjt_{j} and does not intersect with each other. We set m=max{mk,mk+1}m_{*}=\max\{m_{k},m_{k+1}\}. We can consider two bases (w1,k,w2,k)(w_{1,\ell_{k}},w_{2,\ell_{k}}) and (w1,rk+1,w2,rk+1)(w_{1,r_{k+1}},w_{2,r_{k+1}}), which are similarly given by the formulas (3.19) with t0=tkt_{0}=t_{k} and t0=tk+1t_{0}=t_{k+1} respectively.

Proposition 4.2.

The change of basis Tk,k+1(ε,h)T_{k,k+1}(\varepsilon,h) given by

(w1,rk+1,w2,rk+1)=(w1,k,w2,k)Tk.k+1(ε,h)(w_{1,r_{k+1}},w_{2,r_{k+1}})=(w_{1,\ell_{k}},w_{2,\ell_{k}})T_{k.k+1}(\varepsilon,h) (4.10)

admits the following asymptotic behavior as (ε,h)(0,0)(\varepsilon,h)\to(0,0) with μm0\mu_{m_{*}}\to 0:

Tk,k+1(ε,h)=(exp(ihtk+1tkV(t)𝑑t)00exp(ihtk+1tkV(t)𝑑t))+𝒪(ε2h).T_{k,k+1}(\varepsilon,h)=\begin{pmatrix}\exp\left(-\frac{i}{h}\int_{t_{k+1}}^{t_{k}}V(t)dt\right)&0\\ 0&\exp\left(\frac{i}{h}\int_{t_{k+1}}^{t_{k}}V(t)dt\right)\end{pmatrix}+{\mathcal{O}}\left(\frac{\varepsilon^{2}}{h}\right). (4.11)
Remark 4.3.

The error term in (4.11) is rewritten as 𝒪(ε2/h)=𝒪(μ12){\mathcal{O}}(\varepsilon^{2}/h)={\mathcal{O}}(\mu_{1}^{2}). From the order relation (2.18) of μm\mu_{m} with respect to mm, this error is smaller than that in (4.2) when m>1m>1.

Proof of Proposition 4.2.

We can derive from (4.10) the expression of Tk,k+1T_{k,k+1} by a similar way to the proof of Theorem 3. We set uk±=exp(itktV(s)𝑑s/h)u_{k}^{\pm}=\exp\left(\mp i\int_{t_{k}}^{t}V(s)ds/h\right), and compute the matrix Tk,k+1T_{k,k+1} by using the value at t=kt=\ell_{k}:

Tk,k+1(ε,h)\displaystyle\quad T_{k,k+1}(\varepsilon,h)
=(w1,kw2,k)1(w1,rk+1w2,rk+1)|t=k\displaystyle=(w_{1,\ell_{k}}\ w_{2,\ell_{k}})^{-1}(w_{1,r_{k+1}}\,w_{2,r_{k+1}})\bigr{|}_{t=\ell_{k}}
=(uk(k)00uk+(k))((uk+1+(k)00uk+1(k))+𝒪(ε2h)).\displaystyle=\begin{pmatrix}u^{-}_{k}(\ell_{k})&0\\ 0&u^{+}_{k}(\ell_{k})\end{pmatrix}\left(\begin{pmatrix}u^{+}_{k+1}(\ell_{k})&0\\ 0&u^{-}_{k+1}(\ell_{k})\end{pmatrix}+{\mathcal{O}}\left(\frac{\varepsilon^{2}}{h}\right)\right).

Here, we have used the asymptotic formulas (3.17) and (3.18). Note that the integral interval [rk+1,k][r_{k+1},\ell_{k}] of the fundamental solutions for the construction does not contain any zero of V(t)V(t). We deduce (4.11) from this with the identity

uk±(k)uk+1(k)=exp(±ihtk+1tkV(t)𝑑t).u^{\pm}_{k}(\ell_{k})u^{\mp}_{k+1}(\ell_{k})=\exp\left(\pm\frac{i}{h}\int_{t_{k+1}}^{t_{k}}V(t)dt\right). (4.12)

5 End of the proofs

By using the transfer matrices TkT_{k} (k=1,2,,n)(k=1,2,\ldots,n), Tk,k+1T_{k,k+1} (k=1,2,,n1)(k=1,2,\ldots,n-1) introduced in the previous section and TrT_{r}, TT_{\ell} in Appendix A.1, the scattering matrix S=S(ε,h)S=S(\varepsilon,h) is represented as

S=Tr1T1T1,2T2T2,3T3Tn1,nTnT.S=T_{r}^{-1}T_{1}T_{1,2}T_{2}T_{2,3}T_{3}\cdots T_{n-1,n}T_{n}T_{\ell}. (5.1)

Here, the matrix TrT_{r} (resp. TT_{\ell}) has a similar form to Tk,k+1T_{k,k+1} which connects Jost solutions Jr±J_{r}^{\pm} (resp. J±J_{\ell}^{\pm}) and local solutions near t1t_{1} (resp. tnt_{n}). The previous section shows the asymptotic behaviors of TkT_{k} and Tk,k+1T_{k,k+1} (see (4.9), (4.11)), and the appendix does those of TrT_{r} and TT_{\ell} (see (A.17), (A.18), (A.20)). The formula of TT_{\ell} depending on the sign of VV_{\ell} can be rewritten as

T={Tn,n+1(σn is even)Tn,n+1J(σn is odd),T_{\ell}=\left\{\begin{aligned} &T_{n,n+1}&&\qquad(\text{$\sigma_{n}$ is even})\\ &T_{n,n+1}J&&\qquad(\text{$\sigma_{n}$ is odd}),\end{aligned}\right. (5.2)

by means of σn=k=1nmk\sigma_{n}=\sum_{k=1}^{n}m_{k} and

Tn,n+1:=(exp(ihR)00exp(+ihR)),J:=(0110),T_{n,n+1}:=\begin{pmatrix}\exp\left(-\frac{i}{h}R_{\ell}\right)&0\\ 0&\exp\left(+\frac{i}{h}R_{\ell}\right)\end{pmatrix},\qquad J:=\begin{pmatrix}0&-1\\ 1&0\end{pmatrix}, (5.3)

where JJ is a complex structure on 2\mathbb{C}^{2}, that is, J2=IdJ^{2}=-\mathrm{Id}. Note that all of these transfer matrices are elements of SU(2)\mathrm{\,SU}(2), which is mentioned in §A.2. By using the notation (A.23), the scattering matrix SS is expressed as S=Tr1𝒯nS=T_{r}^{-1}{\mathcal{T}}_{n} (resp. S=Tr1𝒯nJS=T_{r}^{-1}{\mathcal{T}}_{n}J) if σn\sigma_{n} is even (resp. odd). This implies that, when σn\sigma_{n} is even (resp. odd), the off-daiagnoal entry s21s_{21} equals eiRr/hτ21ne^{-iR_{r}/h}\tau_{21}^{n} (resp. eiRr/hτ22ne^{-iR_{r}/h}\tau_{22}^{n}). From Lemma A.4 alone, it is complicated to examine the asymptotics of τ22n\tau_{22}^{n} up to the coefficient of 𝒪(μ2){\mathcal{O}}(\mu^{2}). However, thanks to the unitarity of the scattering matrix SS, that is, |s11|2+|s21|2=1|s_{11}|^{2}+|s_{21}|^{2}=1, the computation of the asymptotic behavior of |s21|2|s_{21}|^{2} can be reduced to that of |τ21n|2|\tau_{21}^{n}|^{2} even if σn\sigma_{n} is odd.

5.1 Proof of Theorem 1

Let us demonstrate the proof of Theorem 1. Taking the complex numbers αk\alpha_{k}, βk(μ)\beta_{k}(\mu) and νk\nu_{k} in Lemma A.4 as

αk=1,βk(μ)=iωmk¯μmk,νk=exp(ihtk+1tkV(t)𝑑t),\displaystyle\alpha_{k}=1,\qquad\beta_{k}(\mu)=-i\overline{\omega_{m_{k}}}\mu_{m_{k}},\qquad\nu_{k}=\exp\left(-\frac{i}{h}\int_{t_{k+1}}^{t_{k}}V(t)dt\right),

and noting μmkμ\mu_{m_{k}}\ll\mu_{*} for any kk, we have from the algebraic formula (A.28) the asymptotic behavior of |τ21n|2|\tau_{21}^{n}|^{2} as follows:

|τ21n|2\displaystyle|\tau_{21}^{n}|^{2} =μ2jΛ|ωmj|2+2μ2Rej,kΛj<kωmj¯ωmke2ihtktjV(t)𝑑t\displaystyle=\mu_{*}^{2}\sum_{j\in\Lambda_{*}}|\omega_{m_{j}}|^{2}+2\mu_{*}^{2}\,{\rm Re}\hskip 1.0pt\,\sum_{\begin{subarray}{c}j,k\in\Lambda_{*}\\ j<k\end{subarray}}\overline{\omega_{m_{j}}}\omega_{m_{k}}e^{-\frac{2i}{h}\int_{t_{k}}^{t_{j}}V(t)dt} (5.4)
+𝒪(μ3)+𝒪(μμ1),\displaystyle\quad+{\mathcal{O}}(\mu_{*}^{3})+{\mathcal{O}}(\mu_{*}\mu_{*-1}),

where μ1\mu_{*-1} stands for μm1\mu_{m_{*}-1}. Concerning the error terms in (5.4), the former one, i.e. 𝒪(μ3){\mathcal{O}}(\mu_{*}^{3}), is a higher order error term coming from Λ\Lambda_{*} 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 𝒪(μμ1){\mathcal{O}}(\mu_{*}\mu_{*-1}). If mm_{*} is odd, ωmj\omega_{m_{j}} is not real for jΛj\in\Lambda_{*}, and the following phase shift term may arise from the product ωmj¯ωmk\overline{\omega_{m_{j}}}\omega_{m_{k}} depending on the sign of vjv_{j} and vkv_{k}:

arg(ωmj¯ωmk)=((sgnvk)(sgnvj))π2(m+1)={(sgnvj)π2(m+1)ifsgnvj=sgnvk,0otherwise.\operatorname{arg}(\overline{\omega_{m_{j}}}\omega_{m_{k}})=\frac{((\operatorname{sgn}v_{k})-(\operatorname{sgn}v_{j}))\pi}{2(m_{*}+1)}=\left\{\begin{aligned} -&\frac{(\operatorname{sgn}v_{j})\pi}{2(m_{*}+1)}&&\text{if}\ \operatorname{sgn}v_{j}=-\operatorname{sgn}v_{k},\\ 0&&&\text{otherwise}.\end{aligned}\right. (5.5)

Therefore, when mm_{*} is odd, the quantity |τ21|2|\tau_{21}|^{2} behaves like

μ2(ωo)2(jΛ|vj|2m+1+2Rej,kΛj<k|vjvk|1m+1cos(2htktjV(t)𝑑t+θj,k))\displaystyle\mu_{*}^{2}(\omega_{*}^{o})^{2}\left(\sum_{j\in\Lambda_{*}}|v_{j}|^{-\frac{2}{m_{*}+1}}+2{\rm Re}\hskip 1.0pt\,\sum_{\begin{subarray}{c}j,k\in\Lambda_{*}\\ j<k\end{subarray}}|v_{j}v_{k}|^{-\frac{1}{m_{*}+1}}\cos\left(\frac{2}{h}\int_{t_{k}}^{t_{j}}V(t)dt+\theta^{j,k}\right)\right)
+𝒪(μ3)+𝒪(μμ1),\displaystyle\quad+{\mathcal{O}}(\mu_{*}^{3})+{\mathcal{O}}(\mu_{*}\mu_{*-1}), (5.6)

where

ωo\displaystyle\omega_{*}^{o} =2((m+1)!2)1m+1Γ(m+2m+1),\displaystyle=2\left(\frac{(m_{*}+1)!}{2}\right)^{\frac{1}{m_{*}+1}}\!\!\varGamma\left(\frac{m_{*}+2}{m_{*}+1}\right),
θj,k\displaystyle\theta^{j,k} ={(sgnvj)π2(m+1)ifsgnvj=sgnvk,0otherwise.\displaystyle=\left\{\begin{aligned} &(\operatorname{sgn}v_{j})\frac{\pi}{2(m_{*}+1)}&&\text{if}\ \operatorname{sgn}v_{j}=-\operatorname{sgn}v_{k},\\ &0&&\text{otherwise.}\end{aligned}\right.

On the other hand, ωmj\omega_{m_{j}} is real when mm_{*} is even. In this case, the quantity |τ21|2|\tau_{21}|^{2} can be computed similarly as the formula (5.6) by replacing ωo\omega_{*}^{o} with ωocos(π/2(m+1))\omega_{*}^{o}\cos(\pi/2(m_{*}+1)) and θj,k\theta^{j,k} with 0. Therefore we have completed the proof of Theorem 1.

\Box

5.2 Proof of Theorem 2

The intermediate regime where μ0\mu_{\flat}\to 0 and μ\mu_{\sharp}\to\infty 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 TkT_{k} is replaced as

Tk(ε,h)={Tk(N)(ε,h)kΛ¯,Tk(A)(ε,h)kΛ¯,T_{k}(\varepsilon,h)=\left\{\begin{aligned} &T_{k}^{(N)}(\varepsilon,h)&&k\in\overline{\Lambda_{\flat}},\\ &T_{k}^{(A)}(\varepsilon,h)&&k\in\overline{\Lambda_{\sharp}},\end{aligned}\right. (5.7)

where Tk(N)T_{k}^{(N)} is equal to (4.9) and Tk(A)T_{k}^{(A)} 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 ψ±\psi_{\bullet}^{\pm} (=,r)(\bullet=\ell,r) 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 (ψ+,ψ)(\psi_{\ell}^{+},\psi_{\ell}^{-}) and (ψr+,ψr)(\psi_{r}^{+},\psi_{r}^{-}). By matching the asymptotic behaviors of the exact WKB solutions ψ±\psi_{\bullet}^{\pm} and the MSA solutions wj,w_{j,\bullet} (j{1,2},{,r})(j\in\{1,2\},\bullet\in\{\ell,r\}) on their semiclassical wave front sets referred in [10], we have the connection formula between them. Consequently, in the case where both mkm_{k} and σk1\sigma_{k-1} are even, the transfer matrix Tk(A)T_{k}^{(A)} is of the form:

Tk(A):=Tkw=(αkβ¯kβkα¯k)SU(2),T_{k}^{(A)}:=T_{k}^{\rm w}=\begin{pmatrix}\alpha_{k}&-\bar{\beta}_{k}\\ \beta_{k}&\bar{\alpha}_{k}\end{pmatrix}\in\mathrm{\,SU}(2),

where αk,βk\alpha_{k},\beta_{k} have the asymptotic expansions as μmk\mu_{m_{k}}\to\infty:

αk\displaystyle\alpha_{k} =exp[i2h(Ak,1Ak,mk)]\displaystyle=\exp\left[-\frac{i}{2h}(A_{k,1}-A_{k,m_{k}})\right]
+(1)mkexp[i2h(Ak,12Ak,1¯+Ak,mk)]+𝒪(μmkmk+1mk),\displaystyle\quad+(-1)^{m_{k}}\exp\left[-\frac{i}{2h}(A_{k,1}-2\overline{A_{k,1}}+A_{k,m_{k}})\right]+{\mathcal{O}}\left(\mu_{m_{k}}^{-\frac{m_{k}+1}{m_{k}}}\right), (5.8)
βk\displaystyle\beta_{k} =(1)σk{exp[i2h(Ak,1Ak,mk¯)]\displaystyle=(-1)^{\sigma_{k}}\Bigl{\{}\exp\left[-\frac{i}{2h}(A_{k,1}-\overline{A_{k,m_{k}}})\right]
(1)mkexp[i2h(Ak,12Ak,1¯+Ak,mk¯)]}\displaystyle\hskip 56.9055pt-(-1)^{m_{k}}\exp\left[-\frac{i}{2h}(A_{k,1}-2\overline{A_{k,1}}+\overline{A_{k,m_{k}}})\right]\Bigr{\}}
+𝒪(μmkmk+1mkexp[aμmk(mk+1)/mk]),\displaystyle\qquad+{\mathcal{O}}\left(\mu_{m_{k}}^{-\frac{m_{k}+1}{m_{k}}}\exp\left[-a\mu_{m_{k}}^{(m_{k}+1)/m_{k}}\right]\right), (5.9)

with the notations given in §2.2. Actually αk\alpha_{k} is of 𝒪(1){\mathcal{O}}(1) and βk\beta_{k} is exponentially small of 𝒪(exp[aμmk(mk+1)/mk]){\mathcal{O}}(\exp[-a\mu_{m_{k}}^{(m_{k}+1)/m_{k}}]) as μmk\mu_{m_{k}}\to\infty. Notice that the above case where both mkm_{k} and σk1\sigma_{k-1} are even implies that the eigenvalue of H(t,ε)H(t,\varepsilon) (i.e. the energy of the system: V(t)2+ε2\sqrt{V(t)^{2}+\varepsilon^{2}}) and the energy without an interaction (i.e. V(t))V(t)) have the same sign before and behind the vanishing point t=tkt=t_{k}. Conversely, in the case where the sign of V(t)2+ε2\sqrt{V(t)^{2}+\varepsilon^{2}} does not coincide with that of V(t)V(t) near the vanishing point, the correspondence of the exact WKB solutions to the MSA solutions varies. In fact, Tk(A)T_{k}^{(A)} for kΛ¯k\in\overline{\Lambda_{\sharp}} depends on mkm_{k} and σk1\sigma_{k-1} as follows:

Tk(A)={Tkwif(mk,σk1)=(even,even),𝒞σk1Tkwif(mk,σk1)=(even,odd),iQ(i00i)Tkwif(mk,σk1)=(odd,even),iQ(i00i)𝒞σk1Tkwif(mk,σk1)=(odd,odd),T_{k}^{(A)}=\left\{\begin{aligned} &T_{k}^{\rm w}&&\qquad\text{if}\ (m_{k},\sigma_{k-1})=(\text{even},\text{even}),\\[5.0pt] &{\mathcal{C}}^{\sigma_{k-1}}T_{k}^{\rm w}&&\qquad\text{if}\ (m_{k},\sigma_{k-1})=(\text{even},\text{odd}),\\[5.0pt] &iQ\begin{pmatrix}-i&0\\ 0&i\end{pmatrix}T_{k}^{\rm w}&&\qquad\text{if}\ (m_{k},\sigma_{k-1})=(\text{odd},\text{even}),\\[5.0pt] &iQ\begin{pmatrix}i&0\\ 0&-i\end{pmatrix}{\mathcal{C}}^{\sigma_{k-1}}T_{k}^{\rm w}&&\qquad\text{if}\ (m_{k},\sigma_{k-1})=(\text{odd},\text{odd}),\end{aligned}\right. (5.10)

where the matrix Q=(0110)Q=\begin{pmatrix}0&1\\ 1&0\end{pmatrix} equipped with useful properties:

Q(abcd)=(dcba)Q,Q2=Id,Q\begin{pmatrix}a&b\\ c&d\end{pmatrix}=\begin{pmatrix}d&c\\ b&a\end{pmatrix}Q,\qquad Q^{2}={\rm Id}, (5.11)

and 𝒞{\mathcal{C}} is an operator of taking a complex conjugate, that is 𝒞(abcd)=(a¯b¯c¯d¯){\mathcal{C}}\begin{pmatrix}a&b\\ c&d\end{pmatrix}=\begin{pmatrix}\bar{a}&\bar{b}\\ \bar{c}&\bar{d}\end{pmatrix}. Notice that each matrix in (5.10) from which are removed the factor iQiQ if it exists belongs to SU(2)\mathrm{\,SU}(2). Introducing the notation of the transfer matrix TkT_{k}^{\prime} belonging to SU(2)\mathrm{\,SU}(2) as follows:

Tk={Tk(N)kΛ¯TkwkΛ¯(mk,σk1)=(even,even),𝒞σk1TkwkΛ¯(mk,σk1)=(even,odd),(i00i)TkwkΛ¯(mk,σk1)=(odd,even),(i00i)𝒞σk1TkwkΛ¯(mk,σk1)=(odd,odd),T_{k}^{\prime}=\left\{\begin{aligned} &T_{k}^{(N)}&&k\in\overline{\Lambda_{\flat}}\\ &T_{k}^{\rm w}&&k\in\overline{\Lambda_{\sharp}}\quad(m_{k},\sigma_{k-1})=(\text{even},\text{even}),\\[5.0pt] &{\mathcal{C}}^{\sigma_{k-1}}T_{k}^{\rm w}&&k\in\overline{\Lambda_{\sharp}}\quad(m_{k},\sigma_{k-1})=(\text{even},\text{odd}),\\[5.0pt] &\begin{pmatrix}-i&0\\ 0&i\end{pmatrix}T_{k}^{\rm w}&&k\in\overline{\Lambda_{\sharp}}\quad(m_{k},\sigma_{k-1})=(\text{odd},\text{even}),\\[5.0pt] &\begin{pmatrix}i&0\\ 0&-i\end{pmatrix}{\mathcal{C}}^{\sigma_{k-1}}T_{k}^{\rm w}&&k\in\overline{\Lambda_{\sharp}}\quad(m_{k},\sigma_{k-1})=(\text{odd},\text{odd}),\end{aligned}\right. (5.12)

and recalling the commutative property (5.11), we can express the scattering matrix in the intermediate regime by

S\displaystyle S =Tr1k=1nTkTk,k+1=Tr1(k=1nTk~Tk,k+1~)(iQ)N,\displaystyle=T_{r}^{-1}\prod_{k=1}^{n}T_{k}T_{k,k+1}=T_{r}^{-1}\left(\prod_{k=1}^{n}\widetilde{T_{k}^{\prime}}\widetilde{T_{k,k+1}}\right)\left(iQ\right)^{N}, (5.13)

where the notation of Aj~\widetilde{A_{j}} for the 2×22\times 2 matrix AjA_{j} depending on jj stands for

Aj~\displaystyle\widetilde{A_{j}} ={QAjQ(k(2l1)j<k(2l)),Ajotherwise,\displaystyle=\left\{\begin{aligned} &QA_{j}Q&&\ (k(2l-1)\leq j<k(2l)),\\ &A_{j}&&\qquad\text{otherwise},\end{aligned}\right. (5.14)

with the same notations as in §2.2. Remark that if AjSU(2)A_{j}\in\mathrm{\,SU}(2) then QAjQSU(2)QA_{j}Q\in\mathrm{\,SU}(2). 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 N=#Λodd¯N=\#\overline{\Lambda_{\sharp}^{\rm odd}}.

As a sequel to this computation of the product in (5.13), we can derive the dependence on μ,μ\mu_{\flat},\mu_{\sharp} more precisely. Denoting the (1,1)(1,1)-entry of Tk~\widetilde{T_{k}^{\prime}} by αk~\widetilde{\alpha_{k}^{\prime}}, we see that αk~\widetilde{\alpha_{k}^{\prime}} is of 𝒪(1){\mathcal{O}}(1) and, in particular αk~=1\widetilde{\alpha_{k}^{\prime}}=1 for kΛk\in\Lambda_{\sharp}. Setting, similarly, the (2,1)(2,1)-entry of Tk~\widetilde{T_{k}^{\prime}} by βk~\widetilde{\beta_{k}^{\prime}}, we can rewrite βk~\widetilde{\beta_{k}^{\prime}} as

βk~={pkμ(kΛ),qk(mk,σk1)exp[akμ(m+1)/m](kΛ),\widetilde{\beta_{k}^{\prime}}=\left\{\begin{aligned} &p_{k}\mu_{\flat}&&\qquad(k\in\Lambda_{\flat}),\\ &q_{k}(m_{k},\sigma_{k-1})\exp[-a_{k}\mu_{\sharp}^{(m_{\sharp}+1)/m_{\sharp}}]&&\qquad(k\in\Lambda_{\sharp}),\end{aligned}\right. (5.15)

where pkp_{k} and qk(mk,σk1)q_{k}(m_{k},\sigma_{k-1}) are uniquely determined by (5.9), (5.12) and (5.14). Notice that pkp_{k} and qk(mk,σk1)q_{k}(m_{k},\sigma_{k-1}) are of 𝒪(1){\mathcal{O}}(1) in each regime. On the other hand, Tk,k+1~\widetilde{T_{k,k+1}} can be regarded as the matrix Tk,k+1T_{k,k+1} by replacing V~(t){\tilde{V}}(t) (see (2.19)) with V(t)V(t). From this fact, it is deduced that the asymptotic of the transition probability in the intermediate regime is determined by the effective energy V~\tilde{V}. Hence, we can obtain the asymptotic behavior of |τ21|2|\tau_{21}|^{2} as follows:

μ2(jΛγ|vj+1|2m+1+2j,kΛj<kReCj,k(ε,h)cos[1htktjV~(t)𝑑t])\displaystyle\mu_{\flat}^{2}\left(\sum_{j\in\Lambda_{\flat}}\gamma_{\flat}|v_{j+1}|^{-\frac{2}{m_{\flat}+1}}+2\!\!\!\!\sum_{\begin{subarray}{c}j,k\in\Lambda_{\flat}\\ j<k\end{subarray}}\!\!\!\!{\rm Re}\,C_{j,k}^{\flat\flat}(\varepsilon,h)\cos\left[\frac{1}{h}\int_{t_{k}}^{t_{j}}\tilde{V}(t)dt\right]\right)
+kΛexp[2akμ(m+1)/m]\displaystyle\quad+\sum_{k\in\Lambda_{\sharp}}\exp\left[-2a_{k}\mu_{\sharp}^{(m_{\sharp}+1)/m_{\sharp}}\right]
+2jΛ,kΛj<kReCj,k(ε,h)μexp[akμ(m+1)/m]cos[1htktjV~(t)𝑑t]\displaystyle\quad+2\!\!\!\!\sum_{\begin{subarray}{c}j\in\Lambda_{\flat},k\in\Lambda_{\sharp}\\ j<k\end{subarray}}\!\!\!\!{\rm Re}\,C_{j,k}^{\flat\sharp}(\varepsilon,h)\mu_{\flat}\exp\left[-a_{k}\mu_{\sharp}^{(m_{\sharp}+1)/m_{\sharp}}\right]\cos\left[\frac{1}{h}\int_{t_{k}}^{t_{j}}\tilde{V}(t)dt\right]
+2j,kΛj<kReCj,k(ε,h)exp[(aj+ak)μ(m+1)/m]cos[1htktjV~(t)𝑑t]\displaystyle\quad+2\!\!\!\sum_{\begin{subarray}{c}j,k\in\Lambda_{\sharp}\\ j<k\end{subarray}}\!\!\!{\rm Re}\,C_{j,k}^{\sharp\sharp}(\varepsilon,h)\exp\left[-(a_{j}+a_{k})\mu_{\sharp}^{(m_{\sharp}+1)/m_{\sharp}}\right]\cos\left[\frac{1}{h}\int_{t_{k}}^{t_{j}}\tilde{V}(t)dt\right]
+𝒪(ϵ1ϵ2),\displaystyle\quad+{\mathcal{O}}(\epsilon_{1}\epsilon_{2}),

where ϵ1,ϵ2\epsilon_{1},\epsilon_{2} are given in §2.2 and

Cj,k(ε,h)\displaystyle C_{j,k}^{\flat\flat}(\varepsilon,h) =pjpk¯,\displaystyle=p_{j}\,\overline{p_{k}}, (5.16)
Cj,k(ε,h)\displaystyle C_{j,k}^{\flat\sharp}(\varepsilon,h) =(κ=j+1k1ακ~2)αk~pjqk(mk,σk1)¯,\displaystyle=\left(\prod_{\kappa=j+1}^{k-1}\widetilde{\alpha_{\kappa}^{\prime}}^{2}\right)\widetilde{\alpha_{k}^{\prime}}\,p_{j}\,\overline{q_{k}(m_{k},\sigma_{k-1})}, (5.17)
Cj,k(ε,h)\displaystyle C_{j,k}^{\sharp\sharp}(\varepsilon,h) =αk~(κ=j+1k1ακ~2)αk~qj(mj,σj1)qk(mk,σk1)¯.\displaystyle=\widetilde{\alpha_{k}^{\prime}}\left(\prod_{\kappa=j+1}^{k-1}\widetilde{\alpha_{\kappa}^{\prime}}^{2}\right)\widetilde{\alpha_{k}^{\prime}}\,q_{j}(m_{j},\sigma_{j-1})\,\overline{q_{k}(m_{k},\sigma_{k-1})}. (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 hh is not required for the argument here.

We first consider the Jost solutions Jr±J_{r}^{\pm} near ++\infty. A discussion for Jl±J_{l}^{\pm} is done similarly but the difference is that we are assuming that VrV_{r} is positive (Condition A). Let HrH_{r} denote the limiting Hamiltonian at ++\infty:

Hr:=(VrεεVr).H_{r}:=\begin{pmatrix}V_{r}&\varepsilon\\ \varepsilon&-V_{r}\end{pmatrix}. (A.1)

The functions defined by

φr+(t)=eiλrt/h(cosθrsinθr),φr(t)=e+iλrt/h(sinθrcosθr),\varphi_{r}^{+}(t)=e^{-i\lambda_{r}t/h}\begin{pmatrix}\cos\theta_{r}\\ \sin\theta_{r}\end{pmatrix},\quad\varphi_{r}^{-}(t)=e^{+i\lambda_{r}t/h}\begin{pmatrix}-\sin\theta_{r}\\ \cos\theta_{r}\end{pmatrix}, (A.2)

where λr=Vr2+ε2\lambda_{r}=\sqrt{V_{r}^{2}+\varepsilon^{2}} and tan2θr=ε/Vr\tan 2\theta_{r}=\varepsilon/V_{r} (0<θr<π/4)(0<\theta_{r}<\pi/4), are particular solutions to hDtψ+Hrψ=0hD_{t}\psi+H_{r}\psi=0 and form a basis of 2\mathbb{C}^{2} for each tt\in\mathbb{R}.

Proposition A.1.

There uniquely exists a pair of solutions (ϕr+,ϕr)(\phi_{r}^{+},\phi_{r}^{-}) to the system (1.1) such that

limt+(ϕr±(t)φr±(t))=0.\lim_{t\to+\infty}\left(\phi_{r}^{\pm}(t)-\varphi_{r}^{\pm}(t)\right)=0. (A.3)
Proof.

Let U(t)U(t) be a 2×22\times 2-matrix valued C1C^{1}-function. We have

ddt(Φr(t)U(t))=Φr(t)U(t)+Φr(t)U(t)=1ihHrΦr(t)U(t)+Φr(t)U(t)\frac{d}{dt}(\Phi_{r}(t)U(t))=\Phi_{r}^{\prime}(t)U(t)+\Phi_{r}(t)U^{\prime}(t)=\frac{1}{ih}H_{r}\Phi_{r}(t)U(t)+\Phi_{r}(t)U^{\prime}(t)

with Φr:=(φr+,φr)\Phi_{r}:=(\varphi_{r}^{+},\varphi_{r}^{-}). Thus, if U(t)U(t) satisfies

U=1ihΦr1(HHr)ΦrU,U^{\prime}=\frac{1}{ih}\Phi_{r}^{-1}(H-H_{r})\Phi_{r}U, (A.4)

each column of the matrix-valued function ΦrU\Phi_{r}U is a solution to the equation (1.1). Put Ar(t):=Φr1(HHr)ΦrA_{r}(t):=\Phi_{r}^{-1}(H-H_{r})\Phi_{r}. From the identity

HHr=(V(t)Vr)(1001)H-H_{r}=(V(t)-V_{r})\begin{pmatrix}1&0\\ 0&-1\end{pmatrix} (A.5)

and Condition A, the matrix-valued function Ar(t)A_{r}(t):

Ar(t)=(V(t)Vr)(cos2θre+2iλrt/hsin2θre2iλrt/hsin2θrcos2θr),A_{r}(t)=(V(t)-V_{r})\begin{pmatrix}\cos 2\theta_{r}&-e^{+2i\lambda_{r}t/h}\sin 2\theta_{r}\\ -e^{-2i\lambda_{r}t/h}\sin 2\theta_{r}&-\cos 2\theta_{r}\end{pmatrix}, (A.6)

is integrable on the half-line [0,[[0,\infty[. Then the function

Ur(t):=exp(ih+tAr(s)𝑑s),U_{r}(t):=\exp\left(-\frac{i}{h}\int_{+\infty}^{t}A_{r}(s)ds\right), (A.7)

is well-defined and solves the equation (A.4) with the boundary condition

limt+Ur(t)=Id.\lim_{t\to+\infty}U_{r}(t)=\mathrm{Id}.

Here we recall that Id{\rm Id} stands for the 2×22\times 2 unit matrix. We finally obtain the solutions ϕr±(t)\phi_{r}^{\pm}(t) with the asymptotic behavior (A.3):

(ϕr+(t),ϕr(t)):=Φr(t)Ur(t).(\phi_{r}^{+}(t),\phi_{r}^{-}(t)):=\Phi_{r}(t)U_{r}(t). (A.8)

From Proposition A.1 and the trace-free property of H(t;ε)H(t;\varepsilon), the pair of (ϕr+,ϕr)(\phi_{r}^{+},\phi_{r}^{-}) forms a basis. Similarly, this fact implies that ϕr+\phi_{r}^{+} (resp. ϕr\phi_{r}^{-}) coincides with the Jost solution Jr+J_{r}^{+} (resp. JrJ_{r}^{-}).

Next, we give the asymptotic behaviors of ϕr±\phi^{\pm}_{r} as ε0\varepsilon\to 0 near some fixed point trt_{r}. Take tr>t1t_{r}>t_{1} (recall that t1=max{t;V=0}t_{1}=\max\{t\in\mathbb{R}\,;\,V=0\} is the first zero of VV) satisfying

+tr(V(t)Vr)𝑑t0.\int_{+\infty}^{t_{r}}(V(t)-V_{r})\,dt\neq 0. (A.9)

We introduce some kind of the action integral taking into account of contributions from the infinity as

Rr=Vrtr++tr(V(s)Vr)𝑑s,\displaystyle R_{r}=V_{r}t_{r}+\int_{+\infty}^{t_{r}}(V(s)-V_{r})\,ds,

and put

ur±=exp(ihtrtV(s)𝑑s).u_{r}^{\pm}=\exp\left(\mp\frac{i}{h}\int_{t_{r}}^{t}V(s)ds\right).
Proposition A.2.

We have

ϕr+(t)=eiRr/h(ur++𝒪(ε2/h)𝒪(ε)),ϕr(t)=e+iRr/h(𝒪(ε)ur+𝒪(ε2/h))\displaystyle\phi_{r}^{+}(t)=e^{-iR_{r}/h}\begin{pmatrix}u_{r}^{+}+{\mathcal{O}}(\varepsilon^{2}/h)\\ {\mathcal{O}}(\varepsilon)\end{pmatrix},\quad\phi_{r}^{-}(t)=e^{+iR_{r}/h}\begin{pmatrix}{\mathcal{O}}(\varepsilon)\\ u_{r}^{-}+{\mathcal{O}}(\varepsilon^{2}/h)\end{pmatrix}

as (ε2/h,ε)(0,0)(\varepsilon^{2}/h,\varepsilon)\to(0,0) uniformly for tt in a small neighborhood of trt_{r}.

Before proving Proposition A.2, we prepare the following.

Lemma A.3.

Let AA be a matrix of the form:

A=ih(abb¯a)A=\frac{i}{h}\begin{pmatrix}-a&b\\ \bar{b}&a\end{pmatrix}

with a{0}a\in\mathbb{R}\setminus\{0\} and bb\in\mathbb{C}. For |b/a|1|b/a|\ll 1, one has

eA\displaystyle e^{A} =(eia/h+𝒪((b/a)2)𝒪(b/a)𝒪(b/a)e+ia/h+𝒪((b/a)2)).\displaystyle=\begin{pmatrix}e^{-ia/h}+{\mathcal{O}}((b/a)^{2})&{\mathcal{O}}(b/a)\\ {\mathcal{O}}(b/a)&e^{+ia/h}+{\mathcal{O}}((b/a)^{2})\end{pmatrix}.
Proof.

Since A2=h2(a2+|b|2)IdA^{2}=-h^{-2}(a^{2}+|b|^{2})\mathrm{Id}, an algebraic computation gives

eA=(coszh)Id+iz(sinzh)(abb¯a)e^{A}=\left(\cos\frac{z}{h}\right)\operatorname{Id}+\frac{i}{z}\left(\sin\frac{z}{h}\right)\begin{pmatrix}-a&b\\ \bar{b}&a\end{pmatrix} (A.10)

where z=a2+|b|2z=\sqrt{a^{2}+|b|^{2}}. We have z=sgn(a)a(1+𝒪(b2/a2))z=\operatorname{sgn}(a)a(1+{\mathcal{O}}(b^{2}/a^{2})) under |b/a|1|b/a|\ll 1 and a0a\neq 0. This gives the following asymptotic formula:

eA=(eia/h+𝒪((b/a)2)00e+ia/h+𝒪((b/a)2))+iz(sinzh)(0bb¯0).e^{A}=\begin{pmatrix}e^{-ia/h}+{\mathcal{O}}((b/a)^{2})&0\\ 0&e^{+ia/h}+{\mathcal{O}}((b/a)^{2})\end{pmatrix}+\frac{i}{z}\left(\sin\frac{z}{h}\right)\begin{pmatrix}0&b\\ \bar{b}&0\end{pmatrix}.

The lemma follows from b/z=𝒪(b/a)b/z={\mathcal{O}}(b/a). ∎

Proof of Proposition A.2.

From the expression (A.6), UrU_{r} defined by (A.7) is written as

Ur(t;ε,h)=exp(ih(r(t)cos2θr(ε)𝒥r(t;h)sin2θr(ε)𝒥r(t;h)¯sin2θr(ε)r(t)cos2θr(ε))),U_{r}(t;\varepsilon,h)=\exp\left(\frac{i}{h}\begin{pmatrix}-\mathcal{I}_{r}(t)\cos 2\theta_{r}(\varepsilon)&\mathcal{J}_{r}(t;h)\sin 2\theta_{r}(\varepsilon)\\[7.0pt] \overline{\mathcal{J}_{r}(t;h)}\sin 2\theta_{r}(\varepsilon)&\mathcal{I}_{r}(t)\cos 2\theta_{r}(\varepsilon)\end{pmatrix}\right), (A.11)

where

r(t)=+t(V(s)Vr)𝑑s,\displaystyle\mathcal{I}_{r}(t)=\int_{+\infty}^{t}(V(s)-V_{r})ds,\quad 𝒥r(t;h)=+t(V(s)Vr)e+2isλr/h𝑑s.\displaystyle\mathcal{J}_{r}(t;h)=\int_{+\infty}^{t}(V(s)-V_{r})e^{+2is\lambda_{r}/h}ds.

Apply Lemma A.3 with

a=a(t,ε)=r(t)cos2θr(ε),b=b(t,ε,h)=𝒥r(t;h)sin2θr(ε).a=a(t,\varepsilon)=\mathcal{I}_{r}(t)\cos 2\theta_{r}(\varepsilon),\quad b=b(t,\varepsilon,h)=\mathcal{J}_{r}(t;h)\sin 2\theta_{r}(\varepsilon). (A.12)

By the choice of trt_{r} with the condition (A.9), a=r(t)cos2θr(ε)a=\mathcal{I}_{r}(t)\cos 2\theta_{r}(\varepsilon) never vanishes for tt near trt_{r}. By definition, we have θr(ε)=𝒪(ε)\theta_{r}(\varepsilon)={\mathcal{O}}(\varepsilon), and consequently |b/a|=𝒪(ε)|b/a|={\mathcal{O}}(\varepsilon). Then Lemma A.3 shows

Ur(t;ε,h)=(eir(t)/h+𝒪(ε2/h)𝒪(ε)𝒪(ε)e+ir(t)/h+𝒪(ε2/h))U_{r}(t;\varepsilon,h)=\begin{pmatrix}e^{-i\mathcal{I}_{r}(t)/h}+{\mathcal{O}}(\varepsilon^{2}/h)&{\mathcal{O}}(\varepsilon)\\ {\mathcal{O}}(\varepsilon)&e^{+i\mathcal{I}_{r}(t)/h}+{\mathcal{O}}(\varepsilon^{2}/h)\end{pmatrix} (A.13)

Note that we have the following decomposition of r(t)\mathcal{I}_{r}(t):

r(t)=r(tr)+trt(V(s)Vr)𝑑s=Rr+trtV(s)𝑑sVrt.\mathcal{I}_{r}(t)=\mathcal{I}_{r}(t_{r})+\int_{t_{r}}^{t}(V(s)-V_{r})ds=R_{r}+\int_{t_{r}}^{t}V(s)ds-V_{r}t. (A.14)

Since Φr(t)\Phi_{r}(t) admits the asymptotic formula

Φr(t)=(φr+φr)=(1+𝒪(ε2h))(eiVrt/h𝒪(ε)𝒪(ε)e+iVrt/h),\Phi_{r}(t)=\begin{pmatrix}\varphi_{r}^{+}&\varphi_{r}^{-}\end{pmatrix}=\left(1+{\mathcal{O}}\left(\frac{\varepsilon^{2}}{h}\right)\right)\begin{pmatrix}e^{-iV_{r}t/h}&{\mathcal{O}}(\varepsilon)\\ {\mathcal{O}}(\varepsilon)&e^{+iV_{r}t/h}\end{pmatrix}, (A.15)

as (ε2/h,ε)(0,0)(\varepsilon^{2}/h,\varepsilon)\to(0,0), Proposition A.2 follows from (A.13) and (A.14). ∎

The asymptotic formula

(Jr+Jr)=(ur++𝒪(ε2/h)𝒪(ε)𝒪(ε)ur+𝒪(ε2/h))(eiRr/h00e+iRr/h)\begin{pmatrix}J_{r}^{+}&J_{r}^{-}\end{pmatrix}=\begin{pmatrix}u_{r}^{+}+{\mathcal{O}}(\varepsilon^{2}/h)&{\mathcal{O}}(\varepsilon)\\ {\mathcal{O}}(\varepsilon)&u_{r}^{-}+{\mathcal{O}}(\varepsilon^{2}/h)\end{pmatrix}\begin{pmatrix}e^{-iR_{r}/h}&0\\ 0&e^{+iR_{r}/h}\end{pmatrix} (A.16)

is directly deduced from Propositions A.1 and A.2. Therefore, we obtain

Tr=(eiRr/h+𝒪(ε2/h)𝒪(ε2)𝒪(ε2)e+iRr/h+𝒪(ε2/h)).T_{r}=\begin{pmatrix}e^{-iR_{r}/h}+{\mathcal{O}}(\varepsilon^{2}/h)&{\mathcal{O}}(\varepsilon^{2})\\ {\mathcal{O}}(\varepsilon^{2})&e^{+iR_{r}/h}+{\mathcal{O}}(\varepsilon^{2}/h)\end{pmatrix}. (A.17)

For the Jost solutions J±(t)J_{\ell}^{\pm}(t), the same argument as above works when V>0V_{\ell}>0 and a similar one induces the existences and the asymptotic behaviors when V<0V_{\ell}<0.

In the case where V>0V_{\ell}>0, by exchanging the sub-index rr for \ell, one sees

T=(eiR/h+𝒪(ε2/h)𝒪(ε2)𝒪(ε2)e+iR/h+𝒪(ε2/h)).T_{\ell}=\begin{pmatrix}e^{-iR_{\ell}/h}+{\mathcal{O}}(\varepsilon^{2}/h)&{\mathcal{O}}(\varepsilon^{2})\\ {\mathcal{O}}(\varepsilon^{2})&e^{+iR_{\ell}/h}+{\mathcal{O}}(\varepsilon^{2}/h)\end{pmatrix}. (A.18)

Here

R=Vt+t(V(s)V)𝑑sR_{\ell}=V_{\ell}t_{\ell}+\int_{-\infty}^{t_{\ell}}(V(s)-V_{\ell})\,ds

with t<tn=min{t;V=0}t_{\ell}<t_{n}=\min\{t\in\mathbb{R}\,;\,V=0\} satisfying that the second integral term in the right-hand side does not vanish.

In the case where V<0V_{\ell}<0, we choose instead of (A.2) particular solutions φ±(t)\varphi_{\ell}^{\pm}(t) to hDtψ+Hψ=0hD_{t}\psi+H_{\ell}\psi=0 as

φ+(t)=eiλt/h(sinηcosη),φ(t)=e+iλt/h(cosηsinη),\varphi_{\ell}^{+}(t)=e^{-i\lambda_{\ell}t/h}\begin{pmatrix}\sin\eta_{\ell}\\ \cos\eta_{\ell}\end{pmatrix},\quad\varphi_{\ell}^{-}(t)=e^{+i\lambda_{\ell}t/h}\begin{pmatrix}-\cos\eta_{\ell}\\ \sin\eta_{\ell}\end{pmatrix}, (A.19)

where λ=V2+ε2\lambda_{\ell}=\sqrt{V_{\ell}^{2}+\varepsilon^{2}} and tan2η=ε/(V)\tan 2\eta_{\ell}=\varepsilon/(-V_{\ell}) (0<η<π/4)(0<\eta_{\ell}<\pi/4). They coincide with the leading terms of Jost solutions J±(t)J_{\ell}^{\pm}(t) when V<0V_{\ell}<0 and satisfy the asymptotic formulas:

φ+e+iVt/h(𝒪(ε)1+𝒪(ε2/h)),φeiVt/h(1+𝒪(ε2/h)𝒪(ε))\varphi_{\ell}^{+}\sim e^{+iV_{\ell}t/h}\begin{pmatrix}{\mathcal{O}}(\varepsilon)\\ 1+{\mathcal{O}}(\varepsilon^{2}/h)\end{pmatrix},\quad\varphi_{\ell}^{-}\sim e^{-iV_{\ell}t/h}\begin{pmatrix}-1+{\mathcal{O}}(\varepsilon^{2}/h)\\ {\mathcal{O}}(\varepsilon)\end{pmatrix}

as (ε2/h,ε)(0,0)(\varepsilon^{2}/h,\varepsilon)\to(0,0) for each tt. One sees that, with (A.19), Proposition A.1 also holds. One also have similar asymptotic formulas to those of Proposition A.2:

ϕ+(t)=e+iR/h(𝒪(ε)u++𝒪(ε2/h)),ϕ(t)=eiR/h(u+𝒪(ε2/h)𝒪(ε))\displaystyle\phi_{\ell}^{+}(t)=e^{+iR_{\ell}/h}\begin{pmatrix}{\mathcal{O}}(\varepsilon)\\ u_{\ell}^{+}+{\mathcal{O}}(\varepsilon^{2}/h)\end{pmatrix},\quad\phi_{\ell}^{-}(t)=e^{-iR_{\ell}/h}\begin{pmatrix}-u_{\ell}^{-}+{\mathcal{O}}(\varepsilon^{2}/h)\\ {\mathcal{O}}(\varepsilon)\end{pmatrix}

as (ε2/h,ε)(0,0)(\varepsilon^{2}/h,\varepsilon)\to(0,0) uniformly in a small neighborhood of t=tt=t_{\ell}, where u±=exp(ittV(s)𝑑s/h)u_{\ell}^{\pm}=\exp(\mp i\int_{t_{\ell}}^{t}V(s)ds/h). We obtain

T=(𝒪(ε)eiR/h+𝒪(ε2/h)e+iR/h+𝒪(ε2/h)𝒪(ε)).T_{\ell}=\begin{pmatrix}{\mathcal{O}}(\varepsilon)&-e^{-iR_{\ell}/h}+{\mathcal{O}}(\varepsilon^{2}/h)\\ e^{+iR_{\ell}/h}+{\mathcal{O}}(\varepsilon^{2}/h)&{\mathcal{O}}(\varepsilon)\end{pmatrix}. (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:

Tk(μ)=(αk βk(μ)βk(μ)αk¯),Tk,k+1=(νk00νk¯),T_{k}(\mu)=\begin{pmatrix}\alpha_{k}&-\vbox{\hrule height=0.5pt\kern 1.07639pt\hbox{\kern-1.00006pt$\beta_{k}(\mu)$\kern-1.00006pt}}\\ \beta_{k}(\mu)&\overline{\alpha_{k}}\end{pmatrix},\qquad T_{k,k+1}=\begin{pmatrix}\nu_{k}&0\\ 0&\overline{\nu_{k}}\end{pmatrix},

where αk\alpha_{k}, βk\beta_{k}, and νk\nu_{k} are complex numbers such that detTk=detTk,k+1=1\det T_{k}=\det T_{k,k+1}=1, namely |αk|2+|βk|2=|νk|2=1|\alpha_{k}|^{2}+|\beta_{k}|^{2}=|\nu_{k}|^{2}=1, and βk(μ)=𝒪(μ)\beta_{k}(\mu)={\mathcal{O}}(\mu) as μ0\mu\to 0. Notice that, in our context (see (4.2)), TkT_{k} and Tk,k+1T_{k,k+1} have this form with the numbers given modulo 𝒪(μ2){\mathcal{O}}(\mu^{2}) by

αk1,βkiωk¯μmk=𝒪(μmk),νkexp(ihtk+1tkV(t)𝑑t).\displaystyle\alpha_{k}\equiv 1,\quad\beta_{k}\equiv-i\overline{\omega_{k}}\mu_{m_{k}}={\mathcal{O}}(\mu_{m_{k}}),\quad\nu_{k}\equiv\exp\left(-\frac{i}{h}\int_{t_{k+1}}^{t_{k}}V(t)dt\right).

In this subsection we give an algebraic formula by means of these notations αk\alpha_{k}, βk(μ)\beta_{k}(\mu) and νk\nu_{k} for simplicity. We know that the product of them is of the form

TkTk,k+1=(αkνkβkνk¯βkνkαkνk¯).T_{k}T_{k,k+1}=\begin{pmatrix}\alpha_{k}\nu_{k}&-\overline{\beta_{k}\nu_{k}}\\ \beta_{k}\nu_{k}&\overline{\alpha_{k}\nu_{k}}\end{pmatrix}. (A.21)

Let SU(2)\mathrm{\,SU}(2) be the special unitary group of degree 22 given by

SU(2)={TM2();T=(ab¯ba¯),a,b,detT=1}.\mathrm{\,SU}(2)=\left\{T\in M_{2}(\mathbb{C})\,;\,T=\begin{pmatrix}a&-\bar{b}\\ b&\bar{a}\end{pmatrix},a,b\in\mathbb{C},\ \det T=1\right\}. (A.22)

One sees that all of the above matrices belong to SU(2)\mathrm{\,SU}(2). Denoting the products of these matrices by

𝒯n=T1T1,2T2TnTn,n+1=(τ11nτ12nτ21nτ22n)SU(2),{\mathcal{T}}_{n}=T_{1}T_{1,2}T_{2}\cdots T_{n}T_{n,n+1}=\begin{pmatrix}\tau_{11}^{n}&\tau_{12}^{n}\\ \tau_{21}^{n}&\tau_{22}^{n}\end{pmatrix}\in\mathrm{\,SU}(2), (A.23)

we get the following lemma:

Lemma A.4.

As μ0\mu\to 0, the following asymptotic formulas hold.

τ11n(μ)=j=1nαjνj+𝒪(μ2),\displaystyle\tau_{11}^{n}(\mu)=\prod_{j=1}^{n}\alpha_{j}\nu_{j}+{\mathcal{O}}(\mu^{2}), (A.24)
τ21n(μ)=j=1n(κ=1j1ακ¯)βj(μ)(k=j+1nαk)(κ=1j1νκ¯)(k=jnνk)+𝒪(μ2),\displaystyle\tau_{21}^{n}(\mu)=\sum_{j=1}^{n}\left(\prod_{\kappa=1}^{j-1}\overline{\alpha_{\kappa}}\right)\beta_{j}(\mu)\left(\prod_{k=j+1}^{n}\alpha_{k}\right)\left(\prod_{\kappa=1}^{j-1}\overline{\nu_{\kappa}}\right)\left(\prod_{k=j}^{n}\nu_{k}\right)+{\mathcal{O}}(\mu^{2}), (A.25)
τ12n(μ)=τ21n¯(μ),τ22n(μ)=τ11n¯(μ),\displaystyle\tau_{12}^{n}(\mu)=-\overline{\tau_{21}^{n}}(\mu),\qquad\tau_{22}^{n}(\mu)=\overline{\tau_{11}^{n}}(\mu), (A.26)

with the convention that κ=10ακ¯=κ=10νκ¯=1\prod_{\kappa=1}^{0}\overline{\alpha_{\kappa}}=\prod_{\kappa=1}^{0}\overline{\nu_{\kappa}}=1.

The proof of this lemma is based on the mathematical induction for the product of the matrices (A.21). A simple computation of |τ21n|2|\tau_{21}^{n}|^{2}, which corresponds to the transition probability, gives

|τ21n|2\displaystyle|\tau_{21}^{n}|^{2} =j=1n|βj(μ)|2+2Re[1j<knβj(μ)αj(κ=j+1k1ακ2)αkβk¯(μ)(κ=jk1νκ2)]\displaystyle=\sum_{j=1}^{n}|\beta_{j}(\mu)|^{2}+2{\rm Re}\hskip 1.0pt\,\left[\sum_{1\leq j<k\leq n}\beta_{j}(\mu)\alpha_{j}\left(\prod_{\kappa=j+1}^{k-1}\alpha_{\kappa}^{2}\right)\alpha_{k}\overline{\beta_{k}}(\mu)\left(\prod_{\kappa=j}^{k-1}\nu_{\kappa}^{2}\right)\right]
+𝒪(μ3),\displaystyle\qquad+{\mathcal{O}}(\mu^{3}), (A.27)

with the convention that κ=j+1jακ2=κ=jj1νκ=1\prod_{\kappa=j+1}^{j}\alpha_{\kappa}^{2}=\prod_{\kappa=j}^{j-1}\nu_{\kappa}=1. Note that we used |αk|=1+𝒪(μ2)|\alpha_{k}|=1+{\mathcal{O}}(\mu^{2}) and |νk|=1|\nu_{k}|=1 in the above computation. In particular, when αk=1+𝒪(μ2)\alpha_{k}=1+{\mathcal{O}}(\mu^{2}), we have

|τ21n|2=j=1n|βj(μ)|2+2Re[1j<knβj(μ)βk(μ)¯(κ=jk1νκ2)]+𝒪(μ3).|\tau_{21}^{n}|^{2}=\sum_{j=1}^{n}|\beta_{j}(\mu)|^{2}+2{\rm Re}\hskip 1.0pt\,\left[\sum_{1\leq j<k\leq n}\beta_{j}(\mu)\overline{\beta_{k}(\mu)}\left(\prod_{\kappa=j}^{k-1}\nu_{\kappa}^{2}\right)\right]+{\mathcal{O}}(\mu^{3}). (A.28)

Remark that the factor γn\gamma_{n} 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.