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

A uniqueness theorem for higher order anharmonic oscillators

Søren Fournais  and  Mikael Persson Sundqvist Aarhus University, Department of Mathematics, Ny Munkegade 181, 8000 Aarhus C, Denmark fournais@imf.au.dk Lund University, Department of Mathematical Sciences, Lund, Sweden mickep@maths.lth.se
Abstract.

We study for α\alpha\in\mathbb{R}, k{0}k\in{\mathbb{N}}\setminus\{0\} the family of self-adjoint operators

d2dt2+(tk+1k+1α)2-\frac{d^{2}}{dt^{2}}+\Bigl{(}\frac{t^{k+1}}{k+1}-\alpha\Bigr{)}^{2}

in L2()L^{2}(\mathbb{R}) and show that if kk is even then α=0\alpha=0 gives the unique minimum of the lowest eigenvalue of this family of operators. Combined with earlier results this gives that for any k1k\geq 1, the lowest eigenvalue has a unique minimum as a function of α\alpha.

Key words and phrases:
Eigenvalue estimation, Anharmonic oscillator, Spectral parameter.
1991 Mathematics Subject Classification:
47A75; 47E05, 34L15, 34B08

1. Introduction

1.1. Definition of 𝔔(k)(α)\mathfrak{Q}^{(k)}(\alpha) and main result

For any k{0}k\in{\mathbb{N}}\setminus\{0\} and α\alpha\in\mathbb{R} we define the operator

𝔔(k)(α)=d2dt2+(tk+1k+1α)2,\mathfrak{Q}^{(k)}(\alpha)=-\frac{d^{2}}{dt^{2}}+\Bigl{(}\frac{t^{k+1}}{k+1}-\alpha\Bigr{)}^{2},

as a self-adjoint operator in L2()L^{2}({\mathbb{R}}). This family of operators is connected with the study of Schrödinger operators with a magnetic field vanishing along a curve and with the Ginzburg-Landau theory of superconductivity. It first appeared in [9] (for k=1k=1) and was later studied in [7, 10, 6, 5, 8, 2, 3, 4].

We denote by {λj,𝔔(k)(α)}j=1\bigl{\{}\lambda_{j,\mathfrak{Q}^{(k)}(\alpha)}\bigr{\}}_{j=1}^{\infty} the increasing sequence of eigenvalues of 𝔔(k)(α)\mathfrak{Q}^{(k)}(\alpha). In particular, λ1,𝔔(k)(α)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} is the ground state eigenvalue, and we denote by uαu_{\alpha} the associated positive, L2L^{2}-normalized eigenfunction.

The main result of the present paper is the following theorem.

Theorem 1.1.

Assume that k2k\geq 2 is an even integer. Then λ1,𝔔(k)(α)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} attains a unique minimum at α=0\alpha=0. Moreover, this minimum is non-degenerate.

Remark 1.2.

This extends the previous results and discussions in [5, 8], where similar results were obtained for odd kk. The non-degeneracy was proved in [8]. In that paper it was also shown that Theorem 1.1 is valid for large even kk. The fact that the minimum is attained at α=0\alpha=0 was suggested by numerical computations done by V. Bonnaillie–Noël.

Combining our Theorem 1.1 with the results of  [5, 8] we get the following complete answer.

Theorem 1.3.

For any k{0}k\in{\mathbb{N}}\setminus\{0\}, the function αλ1,𝔔(k)(α)\alpha\mapsto\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} attains a unique minimum. Moreover, this minimum is non-degenerate.

The paper is organized as follows. In Section 2 we give several spectral bounds on the first two eigenvalues of 𝔔(k)(α)\mathfrak{Q}^{(k)}(\alpha). These estimates are used to prove Theorem 1.1 for 2k682\leq k\leq 68 in Section 3 and for k70k\geq 70 in Section 4.

2. Auxiliary results

2.1. Introduction

In this section we collect several spectral bounds that will help us in proving Theorem 1.1. In the following, we assume that kk denotes a positive even integer.

With the scaling s=α1/(k+1)ts=\alpha^{-1/(k+1)}t it becomes clear that the form domain of 𝔔(k)(α)\mathfrak{Q}^{(k)}(\alpha) is independent of α\alpha. Thus, we are allowed to use the machinery of analytic perturbation theory.

First we note that 𝔔(k)(α)\mathfrak{Q}^{(k)}(\alpha) and 𝔔(k)(α)\mathfrak{Q}^{(k)}(-\alpha) are unitarily equivalent (map ttt\mapsto-t along with αα\alpha\mapsto-\alpha). This implies that the function αλ1,𝔔(k)(α)\alpha\mapsto\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} is even, and hence has a critical point at α=0\alpha=0. It is proved in [8] that this critical point is a nondegenerate minimum. This also follows from our estimates below.

Lemma 2.1.

If αc\alpha_{c} is a critical point of λ1,𝔔(k)(α)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)}, then

+(tk+1k+1αc)uαc(t)2𝑑t=0\int_{-\infty}^{+\infty}\Bigl{(}\frac{t^{k+1}}{k+1}-\alpha_{c}\Bigr{)}u_{\alpha_{c}}(t)^{2}\,dt=0

and

+(tk+1k+1αc)2uαc(t)2𝑑t=λ1,𝔔(k)(αc)k+2.\int_{-\infty}^{+\infty}\Bigl{(}\frac{t^{k+1}}{k+1}-\alpha_{c}\Bigr{)}^{2}u_{\alpha_{c}}(t)^{2}\,dt=\frac{\lambda_{1,\mathfrak{Q}^{(k)}(\alpha_{c})}}{k+2}.
Sketch of proof.

The first identity, usually referred to as the Feynman–Hellmann formula, follows from first order perturbation theory,

αλ1,𝔔(k)(α)=2+(tk+1k+1α)uα(t)2𝑑t.\frac{\partial}{\partial\alpha}\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)}=-2\int_{-\infty}^{+\infty}\Bigl{(}\frac{t^{k+1}}{k+1}-\alpha\Bigr{)}u_{\alpha}(t)^{2}\,dt.

The second is a virial type identity and is proved by scaling. We refer to [8] for the details. ∎

2.2. Positive second derivative

A key element in our approach is the following Lemma 2.2, which can be used to rule out local maxima under appropriate estimates on the first eigenvalues.

Lemma 2.2 (Lemma 2.3 in [8]).

If αc\alpha_{c} is a critical point of λ1,𝔔(k)(α)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} and

k+2k+6λ2,𝔔(k)(αc)>λ1,𝔔(k)(αc)\frac{k+2}{k+6}\lambda_{2,\mathfrak{Q}^{(k)}(\alpha_{c})}>\lambda_{1,\mathfrak{Q}^{(k)}(\alpha_{c})}

then

2α2λ1,𝔔(k)(α)|α=αc>0.\frac{\partial^{2}}{\partial\alpha^{2}}\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)}\Big{|}_{\alpha=\alpha_{c}}>0.

We give a sketch of the proof for the sake of completeness.

Sketch of proof.

The proof is based on perturbation theory. The second derivative of λ1,𝔔(k)(α)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} is given by

2α2λ1,𝔔(k)(α)=24+(tk+1k+1α)uα(αuα)𝑑t.\frac{\partial^{2}}{\partial\alpha^{2}}\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)}=2-4\int_{-\infty}^{+\infty}\Bigl{(}\frac{t^{k+1}}{k+1}-\alpha\Bigr{)}u_{\alpha}\bigl{(}\partial_{\alpha}u_{\alpha}\bigr{)}\,dt.

Here

αuα=2(𝔔(k)(α)λ1,𝔔(k)(α))1(tk+1k+1α)uα,\partial_{\alpha}u_{\alpha}=-2(\mathfrak{Q}^{(k)}(\alpha)-\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)})^{-1}\Bigl{(}\frac{t^{k+1}}{k+1}-\alpha\Bigr{)}u_{\alpha},

where the inverse is the regularized resolvent. The rest of the proof uses Lemma 2.1, the bound

(𝔔(k)(αc)λ1,𝔔(k)(αc))1(λ2,𝔔(k)(αc)λ1,𝔔(k)(αc))1,\|(\mathfrak{Q}^{(k)}(\alpha_{c})-\lambda_{1,\mathfrak{Q}^{(k)}(\alpha_{c})})^{-1}\|\leq(\lambda_{2,\mathfrak{Q}^{(k)}(\alpha_{c})}-\lambda_{1,\mathfrak{Q}^{(k)}(\alpha_{c})})^{-1},

and the Cauchy-Schwarz inequality. ∎

To apply Lemma 2.2 we need good upper bounds on λ1,𝔔(k)(α)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} and lower bounds on λ2,𝔔(k)(α)\lambda_{2,\mathfrak{Q}^{(k)}(\alpha)}. These will be presented in the sections below.

2.3. Upper bounds

We will at several points need upper bounds on the first eigenvalue of 𝔔(k)(α)\mathfrak{Q}^{(k)}(\alpha). They are given in this section.

Lemma 2.3.

Assume that αc\alpha_{c} is a critical point of αλ1,𝔔(k)(α)\alpha\mapsto\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)}. Then, for all α\alpha\in\mathbb{R} it holds that

λ1,𝔔(k)(α)λ1,𝔔(k)(αc)+(ααc)2.\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)}\leq\lambda_{1,\mathfrak{Q}^{(k)}(\alpha_{c})}+(\alpha-\alpha_{c})^{2}.
Proof.

This follows by inserting the eigenfunction uαcu_{\alpha_{c}} corresponding to λ1,𝔔(k)(αc)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha_{c})} of 𝔔(k)(αc)\mathfrak{Q}^{(k)}(\alpha_{c}) into the quadratic form corresponding to 𝔔(k)(α)\mathfrak{Q}^{(k)}(\alpha) and using Lemma 2.1

Lemma 2.4.

For all α0\alpha\geq 0 it holds that

λ1,𝔔(k)(α)α2+Ak,\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)}\leq\alpha^{2}+A_{k},

with

Ak={23/29(4π6210π4+4410π2267757)1/4,k=2,π24k+2k+1(14(k+1)(2k+3)(2k+4)(2k+5))1/(k+2),k2.A_{k}=\begin{cases}\frac{2^{3/2}}{9}\bigl{(}\frac{4\pi^{6}-210\pi^{4}+4410\pi^{2}-26775}{7}\bigr{)}^{1/4},&k=2,\\ \frac{\pi^{2}}{4}\frac{k+2}{k+1}\bigl{(}\frac{1}{4}(k+1)(2k+3)(2k+4)(2k+5)\bigr{)}^{-1/(k+2)},&k\geq 2.\\ \end{cases}
Proof.

For k4k\geq 4 we refer to Lemma 3.1 in [8]. For k=2k=2 we use the same idea but with a different trial state. A calculation of the energy of the function

u(t)={23ρcos2(πt2ρ),|t|<ρ,0,|t|ρ,u(t)=\begin{cases}\frac{2}{\sqrt{3\rho}}\cos^{2}\bigl{(}\frac{\pi t}{2\rho}\bigr{)},&|t|<\rho,\\ 0,&|t|\geq\rho,\\ \end{cases}

gives (u=1\|u\|=1)

λ1,𝔔(2)(α)\displaystyle\lambda_{1,\mathfrak{Q}^{(2)}(\alpha)} +|u(t)|2+(t33α)2|u(t)|2dt\displaystyle\leq\int_{-\infty}^{+\infty}|u^{\prime}(t)|^{2}+\Bigl{(}\frac{t^{3}}{3}-\alpha\Bigr{)}^{2}|u(t)|^{2}\,dt
=α2+π23ρ2+4π6210π4+4410π226775252π6ρ6.\displaystyle=\alpha^{2}+\frac{\pi^{2}}{3\rho^{2}}+\frac{4\pi^{6}-210\pi^{4}+4410\pi^{2}-26775}{252\pi^{6}}\rho^{6}.

Minimizing in ρ\rho, we get the bound

λ1,𝔔(2)(α)α2+23/29(4π6210π4+4410π2267757)1/4α2+0.6642,\lambda_{1,\mathfrak{Q}^{(2)}(\alpha)}\leq\alpha^{2}+\frac{2^{3/2}}{9}\Bigl{(}\frac{4\pi^{6}-210\pi^{4}+4410\pi^{2}-26775}{7}\Bigr{)}^{1/4}\leq\alpha^{2}+0.6642,

attained for

ρ=21/4π(4π6210π4+4410π2267757)1/82.57.\rho=2^{1/4}\pi\Bigl{(}\frac{4\pi^{6}-210\pi^{4}+4410\pi^{2}-26775}{7}\Bigr{)}^{-1/8}\approx 2.57.

The upper bound given in Lemma 2.4 is graphed (for α=0\alpha=0 and 2k702\leq k\leq 70) in Figure 1 on page 1.

Lemma 2.5.

The function

kπ24k+2k+1(14(k+1)(2k+3)(2k+4)(2k+5))1/(k+2)k\mapsto\frac{\pi^{2}}{4}\frac{k+2}{k+1}\Bigl{(}\frac{1}{4}(k+1)(2k+3)(2k+4)(2k+5)\Bigr{)}^{-1/(k+2)}

appearing in Lemma 2.4 is increasing for k2k\geq 2. In particular it is always bounded from above by π2/4\pi^{2}/4.

Proof.

We will in the proof consider kk to be a real variable. Taking the logarithmic derivative of the expression, we get

a3k3+a2k2+a1k+a0(k+1)(k+2)2(2k+3)(2k+5)\frac{a_{3}k^{3}+a_{2}k^{2}+a_{1}k+a_{0}}{(k+1)(k+2)^{2}(2k+3)(2k+5)}

with (here we note that each term is increasing with kk and thus estimate from below with k=2k=2)

a3\displaystyle a_{3} =4log(2(k+1)(k+2)(2k+3)(2k+5))208log2\displaystyle=4\log(2(k+1)(k+2)(2k+3)(2k+5))-0-8\log 2
4log378203.73,\displaystyle\geq 4\log 78-0\geq 73,
a2\displaystyle a_{2} =20log(2(k+1)(k+2)(2k+3)(2k+5))10840log2\displaystyle=0\log(2(k+1)(k+2)(2k+3)(2k+5))-08-0\log 2
20log37810810.69,\displaystyle\geq 0\log 78-08\geq 069,
a1\displaystyle a_{1} =31log(2(k+1)(k+2)(2k+3)(2k+5))18962log2\displaystyle=1\log(2(k+1)(k+2)(2k+3)(2k+5))-89-2\log 2
31log3781895.02,\displaystyle\geq 1\log 78-89\geq-02,
a0\displaystyle a_{0} =15log(2(k+1)(k+2)(2k+3)(2k+5))10730log2\displaystyle=5\log(2(k+1)(k+2)(2k+3)(2k+5))-07-0\log 2
15log37810717.98.\displaystyle\geq 5\log 78-07\geq-798.

Now, the polynomial

p(k)=3.73k3+10.69k25.02k17.98p(k)=3.73k^{3}+10.69k^{2}-5.02k-17.98

satisfies

p(2)44.58andp(k)=11.19k2+21.38k5.02.p(2)\approx 44.58\quad\text{and}\quad p^{\prime}(k)=11.19k^{2}+21.38k-5.02.

Since p(k)>0p^{\prime}(k)>0 for k2k\geq 2 we find that pp is positive for k2k\geq 2. This implies that the function in the statement is increasing. The final part follows since the limit as k+k\to+\infty is π2/4\pi^{2}/4. ∎

2.4. Lower bounds

To be able to use Lemma 2.2 we need lower bounds on the second eigenvalue. The following function will appear in the bounds.

Lemma 2.6.

It holds that

h(a)\displaystyle h(a) :=max0<σ<1(1σ2)a/(a+2)σ2/(a+2)(a/2)4/(a+2)\displaystyle=\max_{0<\sigma<1}(1-\sigma^{2})^{a/(a+2)}\sigma^{2/(a+2)}(a/2)^{4/(a+2)} (2.1)
=24/(a+2)a(a+4)/(a+2)(a+1)1/(a+2)1.\displaystyle=2^{-4/(a+2)}a^{(a+4)/(a+2)}(a+1)^{1/(a+2)-1}.

Moreover, lima+h(a)=1\lim_{a\to+\infty}h(a)=1.

Proof.

Differentiating (1σ2)a/(a+2)σ2/(a+2)(a/2)4/(a+2)(1-\sigma^{2})^{a/(a+2)}\sigma^{2/(a+2)}(a/2)^{4/(a+2)} with respect to σ\sigma gives

2(a2)/(a+2)a4/(a+2)σa/(a+2)(1σ2)2/(a+2)(1(a+1)σ2)a+2,\frac{2^{(a-2)/(a+2)}a^{4/(a+2)}\sigma^{-a/(a+2)}\bigl{(}1-\sigma^{2}\bigr{)}^{-2/(a+2)}\bigl{(}1-(a+1)\sigma^{2}\bigr{)}}{a+2},

with the unique zero (in 0<σ<10<\sigma<1) at σ=1/a+1\sigma=1/\sqrt{a+1}. Since the function is zero at the endpoints and positive for 0<σ<10<\sigma<1 this must be the maximum. This proves (2.1)

The rest follows by a simple analysis of the right hand side of (2.1). The derivative equals

h(a)=24/(a+2)a2/(a+2)(a+1)1/(a+2)1[a(4+4log22logalog(a+1))+8](a+2)2.h^{\prime}(a)=\frac{2^{-4/(a+2)}a^{2/(a+2)}(a+1)^{1/(a+2)-1}\bigl{[}a\bigl{(}4+4\log 2-2\log a-\log(a+1)\bigr{)}+8\bigr{]}}{(a+2)^{2}}.

Lemma 2.7.

For all real α\alpha and all even k2k\geq 2 it holds that

𝔔(k)(α)h(k)[d2dt2+(tk/2k/2)2],\mathfrak{Q}^{(k)}(\alpha)\geq h(k)\biggl{[}-\frac{d^{2}}{dt^{2}}+\Bigl{(}\frac{t^{k/2}}{k/2}\Bigr{)}^{2}\biggr{]},

where hh is the function from Lemma 2.6.

Proof.

Let 𝔄=iddt\mathfrak{A}=-i\frac{d}{dt} and 𝔅=(tk+1k+1α)\mathfrak{B}=\bigl{(}\frac{t^{k+1}}{k+1}-\alpha\bigr{)}. Then the commutator [𝔄,𝔅][\mathfrak{A},\mathfrak{B}] equals

[𝔄,𝔅]=itk.[\mathfrak{A},\mathfrak{B}]=-it^{k}.

With the Cauchy–Schwarz inequality and the weighted arithmetic-geometric mean inequality, we find that (for all 0<σ<10<\sigma<1)

𝔔(k)(α)(1σ2)d2dt2+σtk=(1σ2)d2dt2+σ(k/2)2(tk/2k/2)2.\mathfrak{Q}^{(k)}(\alpha)\geq-(1-\sigma^{2})\frac{d^{2}}{dt^{2}}+\sigma t^{k}=-(1-\sigma^{2})\frac{d^{2}}{dt^{2}}+\sigma(k/2)^{2}\Bigl{(}\frac{t^{k/2}}{k/2}\Bigr{)}^{2}.

Scaling the variable and invoking Lemma 2.6 gives the result. ∎

Lemma 2.8.

Let hh be the function in Lemma 2.6. For all real α\alpha and all even k2k\geq 2 it holds that

λ2,𝔔(k)(α)Bk,\lambda_{2,\mathfrak{Q}^{(k)}(\alpha)}\geq B_{k},

with

Bk=h(k)32k/(k+2)(k+2)2(2k2)/(k+2)k(k+4)/(k+2)=32kk+2(k+2)22k+2k+2(k+1)k+1k+2.B_{k}=h(k)\frac{3^{2k/(k+2)}(k+2)}{2^{(2k-2)/(k+2)}k^{(k+4)/(k+2)}}=\frac{3^{\frac{2k}{k+2}}(k+2)}{2^{\frac{2k+2}{k+2}}(k+1)^{\frac{k+1}{k+2}}}.
Proof.

Let T>0T>0. We use the estimate

(tk/2k/2)22kTk2t22k4k2Tk,\Bigl{(}\frac{t^{k/2}}{k/2}\Bigr{)}^{2}\geq\frac{2}{k}T^{k-2}t^{2}-\frac{2k-4}{k^{2}}T^{k},

valid for all tt\in\mathbb{R}. Comparing with the harmonic oscillator, and using Lemma 2.7, we get the required estimate for the second eigenvalue. The optimal choice of TT is

T=(32k4)2/(k+2).T=\Bigl{(}\frac{3\sqrt{2k}}{4}\Bigr{)}^{2/(k+2)}. (2.2)

The lower bound of λ2,𝔔(k)(α)\lambda_{2,\mathfrak{Q}^{(k)}(\alpha)} in Lemma 2.8 will tend to 9/49/4 as k+k\to+\infty, which compared to the limit π2/4\pi^{2}/4 for the first eigenvalue is not good enough. Our next aim is to improve this lower bound on λ2,𝔔(k)(α)\lambda_{2,\mathfrak{Q}^{(k)}(\alpha)} for large kk.

Lemma 2.9.

Assume that k70k\geq 70 is even and α\alpha\in\mathbb{R}. Then

λ2,𝔔(k)(α)B~k,\lambda_{2,\mathfrak{Q}^{(k)}(\alpha)}\geq\widetilde{B}_{k},

with

B~k=512(πarctan((π/1.1)21.170(π/1.1)2)1.1)24.719.\widetilde{B}_{k}=\frac{\sqrt{5}-1}{2}\Biggl{(}\frac{\pi-\arctan\Bigl{(}\sqrt{\frac{(\pi/1.1)^{2}}{1.1^{70}-(\pi/1.1)^{2}}}\Bigr{)}}{1.1}\Biggr{)}^{2}\geq 4.719.
Proof.

We first do the commutator estimate

𝔔(k)(α)(1σ2)d2dt2+σtk=512(d2dt2+tk),\mathfrak{Q}^{(k)}(\alpha)\geq-(1-\sigma^{2})\frac{d^{2}}{dt^{2}}+\sigma t^{k}=\frac{\sqrt{5}-1}{2}\Bigl{(}-\frac{d^{2}}{dt^{2}}+t^{k}\Bigr{)},

where σ\sigma in the latter step is chosen to be 512\frac{\sqrt{5}-1}{2}. Next we note that the second eigenvalue of

d2dt2+tk-\frac{d^{2}}{dt^{2}}+t^{k}

in L2()L^{2}(\mathbb{R}) equals the first eigenvalue of the operator

d2dt2+tk-\frac{d^{2}}{dt^{2}}+t^{k}

in L2(+)L^{2}(\mathbb{R}^{+}) with Dirichlet condition at t=0t=0. Let T>1T>1. Then

d2dt2+tk𝔇(k):=d2dt2+Tkχ{t>T},-\frac{d^{2}}{dt^{2}}+t^{k}\geq\mathfrak{D}^{(k)}:=-\frac{d^{2}}{dt^{2}}+T^{k}\chi_{\{t>T\}},

where we, again, impose a Dirichlet condition at t=0t=0. Here χD\chi_{D} denotes the characteristic function of the set DD. Let us estimate the first eigenvalue λ1,𝔇(k)\lambda_{1,\mathfrak{D}^{(k)}} of 𝔇(k)\mathfrak{D}^{(k)}. Clearly

λ1,𝔇(k)(πT)2,\lambda_{1,\mathfrak{D}^{(k)}}\leq\Bigl{(}\frac{\pi}{T}\Bigr{)}^{2},

which is what one gets considering (0,T)(0,T) and imposing a Dirichlet condition at t=Tt=T. The ground state of 𝔇(k)\mathfrak{D}^{(k)} is given by (in the rest of this proof we write λ=λ1,𝔇(k)\lambda=\lambda_{1,\mathfrak{D}^{(k)}})

u(t)={c1sin(λt),0tT,c2eωt,tT,u(t)=\begin{cases}c_{1}\sin(\sqrt{\lambda}t),&0\leq t\leq T,\\ c_{2}e^{-\omega t},&t\geq T,\end{cases}

where

ω2+Tk=λ-\omega^{2}+T^{k}=\lambda

and where we have the gluing conditions at t=Tt=T:

c1sin(λT)\displaystyle c_{1}\sin(\sqrt{\lambda}T) =c2eωTand\displaystyle=c_{2}e^{-\omega T}\quad\text{and}
c1λcos(λT)\displaystyle c_{1}\sqrt{\lambda}\cos(\sqrt{\lambda}T) =c2ωeωT.\displaystyle=-c_{2}\omega e^{-\omega T}.

This gives the equation (in λ\sqrt{\lambda})

tan(λT)=λωi.e.tan(πλT)=λω,\tan(\sqrt{\lambda}T)=-\frac{\sqrt{\lambda}}{\omega}\quad\text{i.e.}\quad\tan(\pi-\sqrt{\lambda}T)=\frac{\sqrt{\lambda}}{\omega},

which has a unique solution in the interval π2T<λ<πT\frac{\pi}{2T}<\sqrt{\lambda}<\frac{\pi}{T}. We think of T>1T>1 and kk large, so that λ/ω\sqrt{\lambda}/\omega is small, and get

λω=λTkλ(π/T)2Tk(π/T)2.\frac{\sqrt{\lambda}}{\omega}=\sqrt{\frac{\lambda}{T^{k}-\lambda}}\leq\sqrt{\frac{(\pi/T)^{2}}{T^{k}-(\pi/T)^{2}}}.

And so by monotonicity

πλTarctan((π/T)2Tk(π/T)2),\pi-\sqrt{\lambda}T\leq\arctan\biggl{(}\sqrt{\frac{(\pi/T)^{2}}{T^{k}-(\pi/T)^{2}}}\biggr{)},

i.e.

λ(πarctan((π/T)2Tk(π/T)2)T)2.\lambda\geq\Biggl{(}\frac{\pi-\arctan\Bigl{(}\sqrt{\frac{(\pi/T)^{2}}{T^{k}-(\pi/T)^{2}}}\Bigr{)}}{T}\Biggr{)}^{2}.

Now, without optimizing, we find that with T=1.1T=1.1 and k70k\geq 70 it holds that

λ2,𝔔(k)(α)512λ512(πarctan((π/1.1)21.170(π/1.1)2)1.1)24.719.\lambda_{2,\mathfrak{Q}^{(k)}(\alpha)}\geq\frac{\sqrt{5}-1}{2}\lambda\geq\frac{\sqrt{5}-1}{2}\Biggl{(}\frac{\pi-\arctan\Bigl{(}\sqrt{\frac{(\pi/1.1)^{2}}{1.1^{70}-(\pi/1.1)^{2}}}\Bigr{)}}{1.1}\Biggr{)}^{2}\geq 4.719.

We will also need lower bounds on λ1,𝔔(k)(α)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} for large α\alpha. This is the content of the following two Lemmas.

Lemma 2.10.

For α3/2\alpha\geq 3/2 and even k2k\geq 2 it holds that

λ1,𝔔(k)(α)Ck,\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)}\geq C_{k}, (2.3)

with

Ck=min((321k+1)2,32(k+1)1(k+1)((32(k+1))1/(k+1)1)Θ0).C_{k}=\min\Biggl{(}\Bigl{(}\frac{3}{2}-\frac{1}{k+1}\Bigr{)}^{2},\frac{\frac{3}{2}(k+1)-1}{(k+1)\bigl{(}\bigl{(}\frac{3}{2}(k+1)\bigr{)}^{1/(k+1)}-1\bigr{)}}\Theta_{0}\Biggr{)}.

In particular, if 2k682\leq k\leq 68 it holds that

λ1,𝔔(k)(α)>λ1,𝔔(k)(0)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)}>\lambda_{1,\mathfrak{Q}^{(k)}(0)}

for all α3/2\alpha\geq 3/2.

Proof.

First we note that the potential (tk+1k+1α)2\bigl{(}\frac{t^{k+1}}{k+1}-\alpha\bigr{)}^{2} is decreasing for all t<1t<1 (in fact for all t<((k+1)α)1/(k+1)t<((k+1)\alpha)^{1/(k+1)}), and thus it is greater than (1/(k+1)3/2)2(1/(k+1)-3/2)^{2} for all t<1t<1 and all α3/2\alpha\geq 3/2.

For t1t\geq 1 and α3/2\alpha\geq 3/2, we estimate

(tk+1k+1α)2\displaystyle\Bigl{(}\frac{t^{k+1}}{k+1}-\alpha\Bigr{)}^{2} =1(k+1)2(j=0ktkj(α(k+1))1/(k+1))2(t(α(k+1))1/(k+1))2\displaystyle=\frac{1}{(k+1)^{2}}\Biggl{(}\sum_{j=0}^{k}t^{k-j}\bigl{(}\alpha(k+1)\bigr{)}^{1/(k+1)}\Biggr{)}^{2}\Bigl{(}t-\bigl{(}\alpha(k+1)\bigr{)}^{1/(k+1)}\Bigr{)}^{2}
1(k+1)2(32(k+1)1(32(k+1))1/(k+1)1)2(t(α(k+1))1/(k+1))2.\displaystyle\geq\frac{1}{(k+1)^{2}}\biggl{(}\frac{\frac{3}{2}(k+1)-1}{\bigl{(}\frac{3}{2}(k+1)\bigr{)}^{1/(k+1)}-1}\biggr{)}^{2}\Bigl{(}t-\bigl{(}\alpha(k+1)\bigr{)}^{1/(k+1)}\Bigr{)}^{2}.

Here we used that the expression in the big sum is increasing both in tt and in α\alpha, and then applied the formula for a geometric sum.

Thus, comparing with the minimum of the potential for t<1t<1 and with the de Gennes operator for t1t\geq 1 we conclude (2.3).

The last part follows by comparing the upper bound in Lemma 2.4 with the just obtained lower bound (and using the fact that Θ0>0.59\Theta_{0}>0.59 which is known from [1]). This is done in Figure 1.

Refer to caption
Figure 1. The disks are the upper bounds on λ1,𝔔(k)(0)\lambda_{1,\mathfrak{Q}^{(k)}(0)} from Lemma 2.4 as a function of kk. The squares are the lower bounds on λ1,𝔔(k)(α)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)}, α3/2\alpha\geq 3/2, as a function of kk, from Lemma 2.10.

We need a better bound for large kk than the one given in Lemma 2.10. We use instead α=2.8\alpha=2.8 as lower bound and find that

Lemma 2.11.

For α2.8\alpha\geq 2.8 it holds that

λ1,𝔔(k)(α)min((2.81k+1)2,2.8(k+1)1(k+1)((2.8(k+1))1/(k+1)1)Θ0).\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)}\geq\min\Biggl{(}\Bigl{(}2.8-\frac{1}{k+1}\Bigr{)}^{2},\frac{2.8(k+1)-1}{(k+1)\bigl{(}\bigl{(}2.8(k+1)\bigr{)}^{1/(k+1)}-1\bigr{)}}\Theta_{0}\Biggr{)}.

For k70k\geq 70 the first term is the smallest one, i.e.

λ1,𝔔(k)(α)(2.81k+1)2(2.8171)27.76.\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)}\geq\Bigl{(}2.8-\frac{1}{k+1}\Bigr{)}^{2}\geq\Bigl{(}2.8-\frac{1}{71}\Bigr{)}^{2}\geq 7.76.

In particular λ1,𝔔(k)(α)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} cannot obtain its global minimum for α2.8\alpha\geq 2.8.

Proof.

The proof is exactly the same as the proof of Lemma 2.10. The second statement follows from noticing that the second term in the minimum is increasing and that its value at k=70k=70 is

2.8(70+1)1(70+1)((2.8(70+1))1/(70+1)1)Θ021.2,\frac{2.8(70+1)-1}{(70+1)\bigl{(}\bigl{(}2.8(70+1)\bigr{)}^{1/(70+1)}-1\bigr{)}}\Theta_{0}\geq 21.2,

while the first term in the minimum is less than 2.82=7.842.8^{2}=7.84.

The last statement follows by using Lemma 2.5 to conclude that the upper bound on λ1,𝔔(k)(0)\lambda_{1,\mathfrak{Q}^{(k)}(0)} in Lemma 2.4 is less than π2/4\pi^{2}/4 for all kk. Since π2/4\pi^{2}/4 is less than 7.767.76 we are done. ∎

3. Proof of Theorem 1.1 for 2k682\leq k\leq 68

Lemma 3.1.

For each even kk, 2k682\leq k\leq 68, let

α=k+2k+6BkAk,\alpha^{*}=\sqrt{\frac{k+2}{k+6}B_{k}-A_{k}},

where AkA_{k} is the upper bound on λ1,𝔔(k)(0)\lambda_{1,\mathfrak{Q}^{(k)}(0)} from Lemma 2.4 and BkB_{k} is the lower bound on λ2,𝔔(k)(α)\lambda_{2,\mathfrak{Q}^{(k)}(\alpha)} from Lemma 2.8. Then, αλ1,𝔔(k)(α)\alpha\mapsto\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} has no critical point in the interval 0<α<α0<\alpha<\alpha^{*}.

Proof.

Assume, to get a contradiction, that 0<αc<α0<\alpha_{c}<\alpha^{*} is a critical point. Then, invoking Lemma 2.4 and the definition of α\alpha^{*} above, we find that

λ1,𝔔(k)(αc)Ak+αc2<Ak+(α)2=k+2k+6Bkk+2k+6λ2,𝔔(k)(αc),\lambda_{1,\mathfrak{Q}^{(k)}(\alpha_{c})}\leq A_{k}+\alpha_{c}^{2}<A_{k}+(\alpha^{*})^{2}=\frac{k+2}{k+6}B_{k}\leq\frac{k+2}{k+6}\lambda_{2,\mathfrak{Q}^{(k)}(\alpha_{c})},

which by Lemma 2.2 implies that αc\alpha_{c} is a non-degenerate local minimum. Hence all critical points in 0<α<α0<\alpha<\alpha^{*} must be non-degenerate local minimums. Now we know that zero is a non-degenerate local minimum. Since there cannot be more than one such in a row we get a contradiction. ∎

Lemma 3.2.

With kk and α\alpha^{*} as in the previous Lemma it holds that λ1,𝔔(k)(α)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} cannot attain its global minimal value in the interval [α,2α)[\alpha^{*},2\alpha^{*}).

Proof.

Assume, to get a contradiction, that we have one αc\alpha_{c} in this interval where we have have a global minimum. Then, in particular, λ1,𝔔(k)(αc)λ1,𝔔(k)(0)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha_{c})}\leq\lambda_{1,\mathfrak{Q}^{(k)}(0)}. Thus, combining again Lemmas 2.2 and 2.3 we find that any critical point in [α,αc)[\alpha^{*},\alpha_{c}) must be a non-degenerate minimum. However, by the previous Lemma we know that there are no critical points in (0,α)(0,\alpha^{*}), and so again we would have two non-degenerate minimums in a row. Since that is not possible we get a contradiction. ∎

Lemma 3.3.

Assume that 2k682\leq k\leq 68 is even. Denote by

α=32CkAk,\alpha^{**}=\frac{3}{2}-\sqrt{C_{k}-A_{k}},

where, again, AkA_{k} is the upper bound on λ1,𝔔(k)(0)\lambda_{1,\mathfrak{Q}^{(k)}(0)} from Lemma 2.4 and CkC_{k} is the lower bound on λ1,𝔔(k)(α)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} from Lemma 2.10.

If α>α\alpha>\alpha^{**} then λ1,𝔔(k)(α)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} cannot attain its global minimum.

Proof.

First we note that if α3/2\alpha\geq 3/2 then λ1,𝔔(k)(α)λ1,𝔔(k)(0)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)}\geq\lambda_{1,\mathfrak{Q}^{(k)}(0)} by Lemma 2.10. Assume, to get a contradiction, that λ1,𝔔(k)(α)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} attains its global minimum for a α<αc<3/2\alpha^{**}<\alpha_{c}<3/2. Then, by Lemma 2.3 it holds that

λ1,𝔔(k)(3/2)\displaystyle\lambda_{1,\mathfrak{Q}^{(k)}(3/2)} λ1,𝔔(k)(αc)+(αc3/2)2\displaystyle\leq\lambda_{1,\mathfrak{Q}^{(k)}(\alpha_{c})}+(\alpha_{c}-3/2)^{2}
λ1,𝔔(k)(0)+(α3/2)2\displaystyle\leq\lambda_{1,\mathfrak{Q}^{(k)}(0)}+(\alpha^{**}-3/2)^{2}
<Ak+CkAk=Ck.\displaystyle<A_{k}+C_{k}-A_{k}=C_{k}.

But this contradicts Lemma 2.10. ∎

The proof of Theorem 1.1 is completed for 2k682\leq k\leq 68 by ploting 2α2\alpha^{*} and α\alpha^{**} and noting that 2α>α2\alpha^{*}>\alpha^{**} for these kk. This is done in Figure 2.

Refer to caption
Figure 2. The disks are α\alpha^{**} from Lemma 3.3. The squares are 2α2\alpha^{*}, where α\alpha^{*} is defined in Lemma 3.1.

4. Proof of Theorem 1.1 for k70k\geq 70

Lemma 4.1.

Assume that k70k\geq 70. Then λ1,𝔔(k)(α)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} cannot have its global minimum for 0<α<2.830<\alpha<2.83.

Proof.

This follows the same lines as the proofs of Lemmas 3.1 and 3.2. We let

α=k+2k+6B~kAk,\alpha^{*}=\sqrt{\frac{k+2}{k+6}\widetilde{B}_{k}-A_{k}},

where AkA_{k} is the upper bound on λ1,𝔔(k)(0)\lambda_{1,\mathfrak{Q}^{(k)}(0)} from Lemma 2.4 (which is increasing in kk by Lemma 2.5) and B~k\widetilde{B}_{k} is the lower bound on λ2,𝔔(k)(α)\lambda_{2,\mathfrak{Q}^{(k)}(\alpha)} from Lemma 2.9. For k70k\geq 70 we note that

2α27276×4.719π242.83.2\alpha^{*}\geq 2\sqrt{\frac{72}{76}\times 4.719-\frac{\pi^{2}}{4}}\geq 2.83.

Combining this result with Lemma 2.11 we find that λ1,𝔔(k)(α)\lambda_{1,\mathfrak{Q}^{(k)}(\alpha)} cannot have its minimum attained for α>0\alpha>0. This proves Theorem 1.1.

Acknowledgements

SF was partially supported by the Lundbeck Foundation, the Danish Natural Science Research Council and the European Research Council under the European Community’s Seventh Framework Program (FP7/2007–2013)/ERC grant agreement 202859.

References

  • [1] V. Bonnaillie-Noël. Harmonic oscillators with Neumann condition of the half-line. Commun. Pure Appl. Anal., 11(6):2221–2237, 2012.
  • [2] N. Dombrowski and N. Raymond. Semiclassical analysis with vanishing magnetic fields.
  • [3] S. Fournais and M. Persson. Strong diamagnetism for the ball in three dimensions. Asymptot. Anal., 72(1-2):77–123, 2011.
  • [4] S. Fournais and M. Persson Sundqvist. Lack of diamagnetism and the Little–Parks effect. arXiv:XXXX:XXXX, 2013.
  • [5] B. Helffer. The Montgomery model revisited. Colloquium Mathematicum, volume in honor of A. Hulanicki, 118(2):391–400, 2010.
  • [6] B. Helffer and Y. A. Kordyukov. Spectral gaps for periodic Schrödinger operators with hypersurface magnetic wells. In Mathematical results in quantum mechanics, pages 137–154. World Sci. Publ., Hackensack, NJ, 2008.
  • [7] B. Helffer and A. Morame. Magnetic bottles in connection with superconductivity. J. Funct. Anal., 185(2):604–680, 2001.
  • [8] B. Helffer and M. Persson. Spectral properties of higher order anharmonic oscillators. Journal of Mathematical Sciences, 165(1), 2010.
  • [9] R. Montgomery. Hearing the zero locus of a magnetic field. Comm. Math. Phys., 168(3):651–675, 1995.
  • [10] X.-B. Pan and K.-H. Kwek. Schrödinger operators with non-degenerately vanishing magnetic fields in bounded domains. Trans. Amer. Math. Soc., 354(10):4201–4227 (electronic), 2002.