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

1-D Schrödinger operators with Coulomb-like potentials

Yuriy Golovaty Department of Mechanics and Mathematics, Ivan Franko National University of Lviv
1 Universytetska str., 79000 Lviv, Ukraine
yuriy.golovaty@lnu.edu.ua
Abstract.

We study the convergence of 1D Schrödinger operators HεH_{\varepsilon} with the potentials which are regularizations of a class of pseudo-potentials having in particular the form

αδ(x)+βδ(x)+γ/|x|orαδ(x)+βδ(x)+γ/x.\alpha\delta^{\prime}(x)+\beta\delta(x)+\gamma/|x|\quad\text{or}\quad\alpha\delta^{\prime}(x)+\beta\delta(x)+\gamma/x.

The limit behaviour of HεH_{\varepsilon} in the norm resolvent topology, as ε0\varepsilon\to 0, essentially depends on a way of regularization of the Coulomb potential and the existence of zero-energy resonances for δ\delta^{\prime}-like potential. All possible limits are described in terms of point interactions at the origin. As a consequence of the convergence results, different kinds of L()L^{\infty}(\mathbb{R})-approximations to the even and odd Coulomb potentials, both penetrable and impenetrable in the limit, are constructed.

Key words and phrases:
1D Schrödinger operator, Coulomb potential, one-dimensional hydrogen atom, δ\delta^{\prime}-potential, scattering problem, penetrability of potential, point interaction
2000 Mathematics Subject Classification:
Primary 34L40, 34B09; Secondary 81Q10

1. Introduction and main results

One-dimensional Schrödinger operators with the Coulomb potentials, the structure of their spectra and the question of penetrability of the Coulomb potentials have been the subject of several mathematical discussions [8, 9, 10], [11, 12, 13, 14], starting with the work of Loudon [1]. These studies are related to the one-dimensional models of the hydrogen atom

d2ψdx2γ|x|ψ=Eψ,d2ψdx2+γxψ=Eψ,x.-\frac{d^{2}\psi}{dx^{2}}-\frac{\gamma}{|x|}\,\psi=E\psi,\qquad-\frac{d^{2}\psi}{dx^{2}}+\frac{\gamma}{x}\,\psi=E\psi,\quad x\in\mathbb{R}. (1.1)

Since the potentials have singularities at the origin, the first derivative of wave function ψ\psi also has in general singularities as x0x\to 0, and therefore the wave function should be subject to some additional conditions at x=0x=0. For these formal differential expressions, mathematics gives a large enough set of the boundary conditions associated with self-adjoint operators in L2()L^{2}(\mathbb{R}) [11, 15, 16]. The main issue here is a physically motivated choice of such conditions. We noticed that this problem has many common features with the problem of δ\delta^{\prime}-potential [17, 18, 19, 20, 21, 22, 23, 24]. First of all, both the Coulomb potential and the δ\delta^{\prime}-potential are very sensitive to a way of their regularization. From a physical point of view, this means that there is no unique one-dimensional model of the hydrogen atom described by the pseudo-Hamiltonians in (1.1). However there are many different quantum systems with the Coulomb-like potentials that exhibit different physical properties.

We study the norm resolvent convergence of Hamiltonians with the Coulomb-like potentials perturbed by localized singular potentials. Assume that real-valued function QQ is locally integrable outside the origin and has an interior singularity at x=0x=0, namely

Q(x)={qx,if a<x<0,q+x,if 0<x<aQ(x)=\begin{cases}\frac{q_{-}}{x},&\text{if \ }-a<x<0,\\ \frac{q_{+}}{x},&\text{if \ }\kern 12.0pt0<x<a\end{cases} (1.2)

for some real constants qq_{-}, q+q_{+} and a>0a>0. We also suppose that QQ is bounded from below if |x|>a|x|>a. Set

Qε(x)={Q(x),if |x|>ε,lnεεϰ(xε),if |x|<ε,Q_{\varepsilon}(x)=\begin{cases}\phantom{\frac{\ln\varepsilon}{\varepsilon}}Q(x),&\text{if \ }|x|>\varepsilon,\\ \frac{\ln\varepsilon}{\varepsilon}\,\varkappa\left(\frac{x}{\varepsilon}\right),&\text{if \ }|x|<\varepsilon,\end{cases} (1.3)

where ϰ\varkappa is a function belonging to L(1,1)L^{\infty}(-1,1). Also let UU and VV be real-valued, measurable and bounded functions with compact supports. In additional, we suppose that their supports are contained in interval =(1,1)\mathcal{I}=(-1,1). We study the convergence of Schrödinger operators

Hε=d2dx2+Qε(x)+1ε2U(xε)+1εV(xε),H_{\varepsilon}=-\frac{d^{2}}{dx^{2}}+Q_{\varepsilon}(x)+\frac{1}{\varepsilon^{2}}\,U\left(\frac{x}{\varepsilon}\right)+\frac{1}{\varepsilon}\,V\left(\frac{x}{\varepsilon}\right), (1.4)

as the positive parameter ε\varepsilon tends to zero. We hereafter interpret ε2U(ε1)\varepsilon^{-2}U(\varepsilon^{-1}\,\cdot) and ε1V(ε1)\varepsilon^{-1}V(\varepsilon^{-1}\,\cdot) as δ\delta^{\prime}-like and δ\delta-like potentials respectively, because

ε2U(ε1x)αδ(x),ε1V(ε1x)βδ(x)\varepsilon^{-2}U(\varepsilon^{-1}x)\to\alpha\delta^{\prime}(x),\qquad\varepsilon^{-1}V(\varepsilon^{-1}x)\to\beta\delta(x)

in the sense of distributions as ε0\varepsilon\to 0, provided UU is a function of zero-mean. In general, the potentials of HεH_{\varepsilon} diverge, because we do not assume that U𝑑x=0\int_{\mathbb{R}}U\,dx=0.

Before stating our main result we introduce some notation. We say that the Schrödinger operator d2dt2+U-\frac{d^{2}}{dt^{2}}+U possesses a zero-energy resonance if there exists a non-trivial solution hh of the equation h′′+Uh=0-h^{\prime\prime}+Uh=0 that is bounded on the whole line. We call hh the half-bound state. We will also simply say that the potential UU is resonant and it possesses a half-bound state hh. We set

θ=h(+)h(),\theta=\frac{h(+\infty)}{h(-\infty)}, (1.5)

where h(±)=limx±h(x)h(\pm\infty)=\lim\limits_{x\to\pm\infty}h(x). These limits exist, because the half-bound state is constant outside the support of UU as a bounded solution of equation h′′=0h^{\prime\prime}=0. Moreover, both the values h(±)h(\pm\infty) are different from zero. Since a half-bound state is defined up to a scalar factor, we fix half-bound state h0h_{0} so that

h0()=1,h0(+)=θ.h_{0}(-\infty)=1,\qquad h_{0}(+\infty)=\theta. (1.6)

Let us set

μ=Vh02𝑑x.\mu=\int_{\mathcal{I}}Vh_{0}^{2}\,dx. (1.7)

We also introduce the spaces

𝒬±={ψL2(±):ψ,ψACloc(±),ψ′′+QψL2(±)}\mathcal{Q}_{\pm}=\left\{\psi\in L^{2}(\mathbb{R}_{\pm})\colon\psi,\,\psi^{\prime}\in AC_{loc}(\mathbb{R}_{\pm}),\>-\psi^{\prime\prime}+Q\psi\in L^{2}(\mathbb{R}_{\pm})\right\}

and denote by 𝒬\mathcal{Q} the space of L2()L^{2}(\mathbb{R})-functions ϕ\phi such that ϕ|±𝒬±\phi|_{\mathbb{R}_{\pm}}\in\mathcal{Q}_{\pm}. Here ACloc(±)AC_{loc}(\mathbb{R}_{\pm}) denotes the set of functions ψ\psi on ±\mathbb{R}_{\pm} which are absolutely continuous on every compact subset of ±\mathbb{R}_{\pm}. Note that the first derivative of ϕ𝒬\phi\in\mathcal{Q} is in general undefined at the origin and has a logarithmic singularity at this point [11, 8, 12].

We say self-adjoint operators HεH_{\varepsilon} converge as ε0\varepsilon\to 0 in the norm resolvent sense if the resolvents (Hεζ)1(H_{\varepsilon}-\zeta)^{-1} converge in the uniform operator topology for all ζ\zeta\in\mathbb{C}\setminus\mathbb{R}.

Our main result reads as follows.

Theorem 1.

The operator family HεH_{\varepsilon} given by (1.4) converges as ε0\varepsilon\to 0 in the norm resolvent sense. If potential UU has a zero-energy resonance, the corresponding half-bound state h0h_{0} satisfies (1.6) and

θ2q+q=ϰh02𝑑x,\theta^{2}q_{+}-q_{-}=\int_{\mathcal{I}}\varkappa h_{0}^{2}\,dx, (1.8)

then HεH_{\varepsilon} converge to operator \mathcal{H} that is defined by ϕ=ϕ′′+Qϕ\mathcal{H}\phi=-\phi^{\prime\prime}+Q\phi on functions ϕ\phi in 𝒬\mathcal{Q}, subject to the coupling conditions

ϕ(+0)=θϕ(0),limx+0(θϕ(x)ϕ(x)(θ2q+q)ϕ(0)lnx)=μϕ(0).\begin{gathered}\phi(+0)=\theta\phi(-0),\\ \lim_{x\to+0}\big{(}\theta\phi^{\prime}(x)-\phi^{\prime}(-x)-(\theta^{2}q_{+}-q_{-})\phi(-0)\ln x\big{)}=\mu\phi(-0).\end{gathered} (1.9)

Otherwise, that is, if either (1.8) does not hold or else UU is not resonant, operators HεH_{\varepsilon} converge to the direct sum =𝒟𝒟+\mathcal{H}=\mathcal{D}_{-}\oplus\mathcal{D}_{+} of the Dirichlet half-line Schrödinger operators 𝒟±=d2dx2+Q\mathcal{D}_{\pm}=-\frac{d^{2}}{dx^{2}}+Q with domains dom𝒟±={ψ𝒬±:ψ(0)=0}\mathop{\rm dom}\mathcal{D}_{\pm}=\{\psi\in\mathcal{Q}_{\pm}\colon\psi(0)=0\}.

Moreover, in both the cases we have

(Hεζ)1(ζ)1Cε1/4.\|(H_{\varepsilon}-\zeta)^{-1}-(\mathcal{H}-\zeta)^{-1}\|\leq C\varepsilon^{1/4}. (1.10)
Remark 1.

If a half-bound state hh is not normalized to unity at x=x=-\infty as in (1.6), then (1.7) and (1.8) transform to read

μ=1|h()|2Vh2𝑑x,θ2q+q=1|h()|2ϰh2𝑑x.\mu=\frac{1}{|h(-\infty)|^{2}}\int_{\mathcal{I}}Vh^{2}\,dx,\qquad\theta^{2}q_{+}-q_{-}=\frac{1}{|h(-\infty)|^{2}}\int_{\mathcal{I}}\varkappa h^{2}\,dx. (1.11)
Remark 2.

Take note that point interactions (1.9) involve implicitly the regularizing function ϰ\varkappa via condition (1.8), which describes a certain interaction of the δ\delta^{\prime}-like and the Coulomb-like potentials.

If we introduce notation b±(ϕ)=limx±0(ϕ(x)q±ϕ(±0)ln|x|)b_{\pm}(\phi)=\lim_{x\to\pm 0}\big{(}\phi^{\prime}(x)-q_{\pm}\phi(\pm 0)\ln|x|\big{)}, then (1.9) can be written in the form

ϕ(+0)=θϕ(0),θb+(ϕ)b(ϕ)=μϕ(0).\phi(+0)=\theta\phi(-0),\qquad\theta b_{+}(\phi)-b_{-}(\phi)=\mu\phi(-0). (1.12)

Taking into account the jump condition for ϕ\phi, we see that

θb+(ϕ)b(ϕ)=θlimx+0(ϕ(x)q+ϕ(+0)ln|x|)limx0(ϕ(x)qϕ(0)ln|x|)=limx+0(θϕ(x)ϕ(x)(θq+ϕ(+0)qϕ(0))ln|x|)=limx+0(θϕ(x)ϕ(x)(θ2q+q)ϕ(0)lnx).\theta b_{+}(\phi)-b_{-}(\phi)=\theta\lim_{x\to+0}\big{(}\phi^{\prime}(x)-q_{+}\phi(+0)\ln|x|\big{)}-\lim_{x\to-0}\big{(}\phi^{\prime}(x)-q_{-}\phi(-0)\ln|x|\big{)}\\ =\lim_{x\to+0}\Big{(}\theta\phi^{\prime}(x)-\phi^{\prime}(-x)-\big{(}\theta q_{+}\phi(+0)-q_{-}\phi(-0)\big{)}\ln|x|\Big{)}\\ =\lim_{x\to+0}\big{(}\theta\phi^{\prime}(x)-\phi^{\prime}(-x)-(\theta^{2}q_{+}-q_{-})\phi(-0)\ln x\big{)}.
Remark 3.

In the case when q=q+=0q_{-}=q_{+}=0 and ϰ=0\varkappa=0, i.e., QQ has no singularity at the origin, the results of this article coincide with the results obtained in [17, 18, 19, 20], where the convergence of Hamiltonians with (αδ+βδ)(\alpha\delta^{\prime}+\beta\delta)-like potentials was discussed.

Now we give some consequences for scattering problems. Let us agree to say that the potentials in (1.4) are penetrable in the limit as ε0\varepsilon\to 0 if the corresponding Schrödinger operators HεH_{\varepsilon} converge to operator \mathcal{H} associated with point interaction (1.9). If the operators converge to the direct sum 𝒟𝒟+\mathcal{D}_{-}\oplus\mathcal{D}_{+}, we say the potentials are opaque in the limit or asymptotically opaque.

Theorem 1 asserts that potentials Qε+ε2U(ε1)+ε1V(ε1)Q_{\varepsilon}+\varepsilon^{-2}U(\varepsilon^{-1}\,\cdot)+\varepsilon^{-1}V(\varepsilon^{-1}\,\cdot) are generally asymptotically opaque. However, for each potential UU that possesses a zero-energy resonance there exists a regularization of QQ having the form (1.3) such that condition (1.8) is fulfilled and hence the potentials are penetrable in the limit. It is also worth noting that resonant potentials are not something exotic, because for any UU of compact support there exists a discrete infinite set of real coupling constants α\alpha for which potential αU\alpha U has a zero-energy resonance.

Coming back to the problem of penetrability of the Coulomb potentials, let us suppose that the potentials of HεH_{\varepsilon} do not contain the δ\delta^{\prime}-like component, i.e., Hε=d2dx2+Qε+ε1V(ε1)H_{\varepsilon}=-\frac{d^{2}}{dx^{2}}+Q_{\varepsilon}+\varepsilon^{-1}V(\varepsilon^{-1}\,\cdot). We left the δ\delta-like potential in the Hamiltonian, because, as shown in the following theorem, VV has no direct influence on the penetrability in the limit.

Theorem 2.

Potentials Qε+ε1V(ε1)Q_{\varepsilon}+\varepsilon^{-1}V(\varepsilon^{-1}\,\cdot) are penetrable in the limit as ε0\varepsilon\to 0 if and only if QεQ_{\varepsilon} converge in the sense of distributions. This is in turn true if and only if the condition

q+q=ϰ𝑑xq_{+}-q_{-}=\int_{\mathcal{I}}\varkappa\,dx (1.13)

holds. In the penetrable case, HεH_{\varepsilon} converge to operator \mathcal{H} associated with point interactions

ϕ(+0)=ϕ(0),limx+0(ϕ(x)ϕ(x)(q+q)ϕ(0)lnx)=βϕ(0),\phi(+0)=\phi(-0),\quad\lim_{x\to+0}\big{(}\phi^{\prime}(x)-\phi^{\prime}(-x)-(q_{+}-q_{-})\phi(0)\ln x\big{)}=\beta\phi(0), (1.14)

where β\beta is the mean value of VV.

2. Coulomb-like potentials:
penetrability and opaqueness in the limit

In this section we will prove Theorem 2 and give some examples of the Coulomb-like potentials QεQ_{\varepsilon} that are penetrable and opaque in the limit.

2.1. Convergence of Coulomb-like potentials

Function QQ of the form (1.2) near the origin is nonintegrable and therefore mapping C0()ψQψ𝑑xC^{\infty}_{0}(\mathbb{R})\ni\psi\mapsto\int_{\mathbb{R}}Q\psi\,dx is not a distribution. However we can find infinitely many functionals q𝒟()q\in\mathcal{D}^{\prime}(\mathbb{R}) which coincide with QQ outside the origin, i.e.,

q(ψ)=Qψ𝑑xfor all ψC0({0}).q(\psi)=\int_{\mathbb{R}}Q\psi\,dx\qquad\text{for all }\psi\in C_{0}^{\infty}(\mathbb{R}\setminus\{0\}).

Among such functionals there exists the family Q\mathcal{F}_{Q} of distributions with the lowest order of singularity. It is easy to check that each qQq\in\mathcal{F}_{Q} is continuous in space C00,γ()C^{0,\gamma}_{0}(\mathbb{R}) of Hölder continuous functions of compact support, but qq is not continuous in C00()C^{0}_{0}(\mathbb{R}). In this sense, qq is more singular than Dirac’s δ\delta-function, but less singular than δ\delta^{\prime}-function. Moreover, if q1q_{1} and q2q_{2} belong to Q\mathcal{F}_{Q}, then q2q1=cδ(x)q_{2}-q_{1}=c\delta(x) for some complex constant cc, and therefore Q={q1+cδ(x):c}\mathcal{F}_{Q}=\{q_{1}+c\delta(x)\colon c\in\mathbb{C}\}.

A word of explanation is necessary with regard to regularization of QQ given by (1.3). Here we use the analogy with formula (ln|x|)=𝒫1x(\ln|x|)^{\prime}=\mathcal{P}\frac{1}{x}. Suppose that GG is an antiderivative of QQ such that G(x)=qln(x)G(x)=q_{-}\ln(-x) for x(a,0)x\in(-a,0) and G(x)=q+lnxG(x)=q_{+}\ln x for x(0,a)x\in(0,a). Function GG specifies a regular distribution on the line, because it belongs to Lloc1()L_{loc}^{1}(\mathbb{R}). We set g=Gg=G^{\prime}, where GG^{\prime} is the derivative in the sense of distributions. Indeed, gg coincides with QQ outside the origin and gQg\in\mathcal{F}_{Q}. Let us now approximate GG in 𝒟()\mathcal{D}^{\prime}(\mathbb{R}) by the sequence of continuous functions

Gε(x)={G(x),if |x|>ε,a(xε)lnε,if |x|<ε,G_{\varepsilon}(x)=\begin{cases}G(x),&\text{if \ }|x|>\varepsilon,\\ a\left(\tfrac{x}{\varepsilon}\right)\ln\varepsilon,&\text{if \ }|x|<\varepsilon,\end{cases}

where aa is a C1C^{1}-function such that a(1)=qa(-1)=q_{-} and a(1)=q+a(1)=q_{+} (see Fig. 1). Then distribution gg admits a regularization by Lloc1()L_{loc}^{1}(\mathbb{R})-functions having the form

Qε(x):=Gε(x)={Q(x),if |x|>ε,lnεεa(xε),if |x|<ε.Q_{\varepsilon}(x):=G_{\varepsilon}^{\prime}(x)=\begin{cases}Q(x),&\text{if \ }|x|>\varepsilon,\\ \frac{\ln\varepsilon}{\varepsilon}\,a^{\prime}\left(\tfrac{x}{\varepsilon}\right),&\text{if \ }|x|<\varepsilon.\end{cases}
Refer to caption
Figure 1. Plots of GεG_{\varepsilon} and QεQ_{\varepsilon}
Lemma 1.

A sequence QεQ_{\varepsilon}, given by (1.3), converges in 𝒟()\mathcal{D}^{\prime}(\mathbb{R}) if and only if condition (1.13) holds. Moreover, the limit distribution, if it exists, belongs to Q\mathcal{F}_{Q}.

Proof.

For each ψC0()\psi\in C_{0}^{\infty}(\mathbb{R}), we have

Qε(x)ψ(x)𝑑x=lnεεεεϰ(xε)ψ(x)𝑑x+q+εaψ(x)x𝑑x+qaεψ(x)x𝑑x+|x|>aQ(x)ψ(x)𝑑x.\int_{\mathbb{R}}Q_{\varepsilon}(x)\psi(x)\,dx=\frac{\ln\varepsilon}{\varepsilon}\int_{-\varepsilon}^{\varepsilon}\varkappa\left(\tfrac{x}{\varepsilon}\right)\psi(x)\,dx\\ +q_{+}\int_{\varepsilon}^{a}\frac{\psi(x)}{x}\,dx+q_{-}\int_{-a}^{-\varepsilon}\frac{\psi(x)}{x}\,dx+\int_{|x|>a}Q(x)\psi(x)\,dx.

If ψ(0)0\psi(0)\neq 0, then all integrals on the right hand side are of the order O(lnε)O(\ln\varepsilon) as ε0\varepsilon\to 0, except for the last one. Indeed, we have

lnεεεεϰ(xε)ψ(x)𝑑xψ(0)lnεϰ𝑑t=lnε11ϰ(t)(ψ(εt)ψ(0))𝑑t,\displaystyle\frac{\ln\varepsilon}{\varepsilon}\int_{-\varepsilon}^{\varepsilon}\varkappa\left(\tfrac{x}{\varepsilon}\right)\psi(x)\,dx-\psi(0)\ln\varepsilon\>\int_{\mathcal{I}}\varkappa\,dt=\ln\varepsilon\>\int_{-1}^{1}\varkappa(t)(\psi(\varepsilon t)-\psi(0))\,dt,
εaψ(x)x𝑑x+ψ(0)lnε=ψ(0)lna+εaψ(x)ψ(0)x𝑑x,\displaystyle\int_{\varepsilon}^{a}\frac{\psi(x)}{x}\,dx+\psi(0)\ln\varepsilon=\psi(0)\ln a+\int_{\varepsilon}^{a}\frac{\psi(x)-\psi(0)}{x}\,dx,
aεψ(x)x𝑑xψ(0)lnε=ψ(0)lna+aεψ(x)ψ(0)x𝑑x.\displaystyle\int^{-\varepsilon}_{-a}\frac{\psi(x)}{x}\,dx-\psi(0)\ln\varepsilon=-\psi(0)\ln a+\int^{-\varepsilon}_{-a}\frac{\psi(x)-\psi(0)}{x}\,dx.

The right-hand sides have finite limits as ε0\varepsilon\to 0, since ψ(x)ψ(0)=O(x)\psi(x)-\psi(0)=O(x) as x0x\to 0. We obtain then

Qε(x)ψ(x)𝑑x=(ϰ𝑑tq++q)ψ(0)lnε+fε(ψ),\int_{\mathbb{R}}Q_{\varepsilon}(x)\psi(x)\,dx=\left(\int_{\mathcal{I}}\varkappa\,dt-q_{+}+q_{-}\right)\psi(0)\ln\varepsilon+f_{\varepsilon}(\psi),

where {fε}ε>0\{f_{\varepsilon}\}_{\varepsilon>0} is a sequence of continuous functionals which converges in 𝒟()\mathcal{D}^{\prime}(\mathbb{R}) as ε0\varepsilon\to 0. Therefore sequence QεQ_{\varepsilon} converges in the space of distributions if and only if q+q=ϰ𝑑tq_{+}-q_{-}=\int_{\mathcal{I}}\varkappa\,dt.

Note that we have actually proved that condition (1.13) is necessary and sufficient for the convergence of functionals QεQ_{\varepsilon} in Hölder space C00,γ()C^{0,\gamma}_{0}(\mathbb{R}), γ(0,1)\gamma\in(0,1). In fact, for any ψC00,γ()\psi\in C^{0,\gamma}_{0}(\mathbb{R}) we have ψ(x)ψ(0)=O(xγ)\psi(x)-\psi(0)=O(x^{\gamma}) as x0x\to 0, and this is sufficient for the convergence of fεf_{\varepsilon}. Therefore if QεQ_{\varepsilon} converge in 𝒟()\mathcal{D}^{\prime}(\mathbb{R}), then the limit distribution belongs to Q\mathcal{F}_{Q}. ∎

2.2. Proof of Theorem 2

The proof deals with the convergence of operators

Hε=d2dx2+Qε+ε1V(ε1),H_{\varepsilon}=-\frac{d^{2}}{dx^{2}}+Q_{\varepsilon}+\varepsilon^{-1}V(\varepsilon^{-1}\,\cdot), (2.1)

and so we have the partial case of Theorem 1 when U=0U=0. First of all, note that the trivial potential U=0U=0 possesses a zero-energy resonance with half-bound state h0=1h_{0}=1. Since θ=1\theta=1, (1.7) and (1.8) become μ=Vdx=:β\mu=\int_{\mathbb{R}}V\,dx=:\beta and

q+q=ϰ𝑑xq_{+}-q_{-}=\int_{\mathcal{I}}\varkappa\,dx

respectively. Hence, if the last condition holds, then operators HεH_{\varepsilon}, given by (2.1), converge to operator \mathcal{H} associated with non-trivial point interactions (1.14), according to Theorem 1. In this case we obtain a partial transparency of the potentials Qε+ε1V(ε1)Q_{\varepsilon}+\varepsilon^{-1}V(\varepsilon^{-1}\,\cdot) in the limit. Next, in view of Lemma 1, the penetrability of these potentials is equivalent to the convergence of QεQ_{\varepsilon} in the space of distributions.

2.3. Examples of Coulomb-like potentials

Following are some examples of potentials QεQ_{\varepsilon}, illustrating the penetrability and impenetrability of the Coulomb-like potentials in the limit. Let us consider two regularizations of the classic Coulomb potential Q(x)=|x|1Q(x)=-|x|^{-1}:

Q0,ε(x)={1|x|,if |x|>ε,0,if |x|<ε;Q1,ε(x)={1|x|,if |x|>ε,ε1|lnε|,if |x|<εQ_{0,\varepsilon}(x)=\begin{cases}-\dfrac{1}{|x|},&\text{if \ }|x|>\varepsilon,\\ \qquad 0,&\text{if \ }|x|<\varepsilon;\end{cases}\qquad Q_{1,\varepsilon}(x)=\begin{cases}\enspace-\dfrac{1}{|x|},&\text{if \ }|x|>\varepsilon,\\ \varepsilon^{-1}|\ln\varepsilon|,&\text{if \ }|x|<\varepsilon\end{cases}

(see Fig. 2). Both of the sequences Qj,εQ_{j,\varepsilon} converge to |x|1-|x|^{-1} pointwise, but Q1,εQ_{1,\varepsilon} only converges in the sense of distributions. In the case of Q1,εQ_{1,\varepsilon}, we have q=1q_{-}=1, q+=1q_{+}=-1 and ϰ=1\varkappa=-1, and therefore condition (1.13) holds. In view of Theorem 2, potentials Q0,εQ_{0,\varepsilon} are asymptotically opaque; whereas Q1,εQ_{1,\varepsilon} are penetrable in the limit as ε0\varepsilon\to 0. In other words, the transition probability |Tε(k)|2|T_{\varepsilon}(k)|^{2} calculated for Q0,εQ_{0,\varepsilon} tends to zero ε0\varepsilon\to 0 for all kk, but the corresponding probability for Q1,εQ_{1,\varepsilon} has a limit |T(k)|2|T(k)|^{2}, which is a non-zero function of kk.

Refer to caption
Figure 2. Impenetrable and penetrable regularizations of the even Coulomb potential

For the odd Coulomb potential Q(x)=x1Q(x)=x^{-1}, we can also provide two different regularizations, plotted in Fig. 3, as follows:

Q2,ε(x)={1x,if |x|>εε1|lnε|,if |x|<ε,Q3,ε(x)={1x,if |x|>ε,ε2|lnε|x,if |x|<ε.Q_{2,\varepsilon}(x)=\begin{cases}\qquad\dfrac{1}{x},&\text{if \ }|x|>\varepsilon\\ \varepsilon^{-1}|\ln\varepsilon|,&\text{if \ }|x|<\varepsilon\end{cases},\qquad Q_{3,\varepsilon}(x)=\begin{cases}\qquad\dfrac{1}{x},&\text{if \ }|x|>\varepsilon,\phantom{\int\limits^{N}_{N}}\\ \varepsilon^{-2}|\ln\varepsilon|\,x,&\text{if \ }|x|<\varepsilon\end{cases}.

In this case, q=q+=1q_{-}=q_{+}=1 and so condition (1.13) holds for potentials Q3,εQ_{3,\varepsilon} only, when ϰ(t)=t\varkappa(t)=-t. Therefore Q3,εQ_{3,\varepsilon} are penetrable in the limit as ε0\varepsilon\to 0, unlike the potentials Q2,εQ_{2,\varepsilon}, which are asymptotically opaque.

Refer to caption
Figure 3. Impenetrable and penetrable regularizations of the odd Coulomb potential

Many authors have regularized the Coulomb potentials by so-called truncated ones of the form

Rε(x)={Q(x),if |x|>ε,aε(x),if |x|<ε,R_{\varepsilon}(x)=\begin{cases}Q(x),&\text{if \ }|x|>\varepsilon,\\ a_{\varepsilon}(x),&\text{if \ }|x|<\varepsilon,\end{cases}

where |aε|cε1|a_{\varepsilon}|\leq c\varepsilon^{-1} (see Fig. 4). From asymptotical point of view, RεR_{\varepsilon} can be regarded as potentials QεQ_{\varepsilon} with ϰ=0\varkappa=0. It follows from the proof of Lemma 1 that RεR_{\varepsilon} can converge in 𝒟()\mathcal{D}^{\prime}(\mathbb{R}) if and only if q=q+q_{-}=q_{+}, i.e., Q(x)Q(x) is the odd Coulomb potential near the origin. Moshinsky [8] was the first who noticed the penetrability of potential γ/x\gamma/x. On the other hand, a regularization of the even Coulomb potential γ/|x|\gamma/|x| by the truncated potentials is always asymptotically opaque and leads to the Dirichlet condition in the limit [1, 3, 2, 7]. The same assertion is also valid for modified Coulomb interactions having the form Mε(x)=1|x|+εM_{\varepsilon}(x)=-\dfrac{1}{|x|+\varepsilon}; this potentials also diverge in 𝒟()\mathcal{D}^{\prime}(\mathbb{R}). Such regularizations were considered in [1, 2, 4, 5, 6].

Refer to caption
Figure 4. Truncated and modified potentials

It should be noted that the equivalence of penetrability in the limit and the convergence in the space of distributions for potentials Qε+ε1V(ε1)Q_{\varepsilon}+\varepsilon^{-1}V(\varepsilon^{-1}\,\cdot) is consistent with Kurasov’s results [12, 14]. Kurasov has interpreted the formal differential expression d2dx2γx-\frac{d^{2}}{dx^{2}}-\frac{\gamma}{x} in \mathbb{R} as a map from some Hilbert space to the space of distributions. This operator has been defined in the principal value sense

H=v.p.(d2dx2γx)+βδ(x)H=v.p.\left(-\frac{d^{2}}{dx^{2}}-\frac{\gamma}{x}\right)+\beta\delta(x)

on the whole line. As shown in [14], HH is the self-adjoint operator that is defined by Hϕ=ϕ′′γϕ/xH\phi=-\phi^{\prime\prime}-\gamma\phi/x on functions from W22((ε,ε))W_{2}^{2}(\mathbb{R}\setminus(-\varepsilon,\varepsilon)) for every positive ε>0\varepsilon>0 and satisfying the boundary conditions

ϕ(+0)=ϕ(0),b+(ϕ)b(ϕ)=βϕ(0),\phi(+0)=\phi(-0),\qquad b_{+}(\phi)-b_{-}(\phi)=\beta\phi(0),

where b±(ϕ)=limx±0(ϕ(x)γϕ(±0)ln|x|)b_{\pm}(\phi)=\lim_{x\to\pm 0}\big{(}\phi^{\prime}(x)-\gamma\phi(\pm 0)\ln|x|\big{)}. Indeed, such considerations implicitly presupposed the existence of a regularization Qε+ε1V(ε1)Q_{\varepsilon}+\varepsilon^{-1}V(\varepsilon^{-1}\,\cdot) of the pseudopotential 𝒫1x+βδ(x)-\mathcal{P}\frac{1}{x}+\beta\delta(x), which converges in 𝒟()\mathcal{D}^{\prime}(\mathbb{R}). These coupling conditions agree with (1.14), if γ=q+=q\gamma=q_{+}=q_{-}.

Returning to the question of penetrability of the one-dimensional Coulomb potentials, it is probably worth considering that this question has no unambiguous answer. One should agree with the authors of [11] that mathematics alone cannot tell which boundary conditions for the wave function at the origin should be chosen to model a given experimental situation.

3. Proof of Theorem 1

3.1. Formal construction of limit operator

Let us consider the equation

y′′+(Qζ)y=f,x{0},-y^{\prime\prime}+(Q-\zeta)y=f,\quad x\in\mathbb{R}\setminus\{0\}, (3.1)

for given fL2()f\in L^{2}(\mathbb{R}) and ζ\zeta\in\mathbb{C}. In one-sided neighbourhoods of the origin the last equation becomes

y′′+(qxζ)y=f,x(a,0);y′′+(q+xζ)y=f,x(0,a).-y^{\prime\prime}+\left(\frac{q_{-}}{x}-\zeta\right)y=f,\quad x\in(-a,0);\qquad-y^{\prime\prime}+\left(\frac{q_{+}}{x}-\zeta\right)y=f,\quad x\in(0,a).

The following proposition was proved in [14].

Proposition 1.

Let yy be a solution of (3.1) such that yL2()y\in L^{2}(\mathbb{R}). Then there exist the finite limits y(±0)=limx±0y(x)y(\pm 0)=\lim_{x\to\pm 0}y(x) and

y(x)=y(±0)+O(|x|1/2)as x±0.y(x)=y(\pm 0)+O(|x|^{1/2})\quad\text{as \ }x\to\pm 0.

For the derivative of the solution we have asymptotics

y(x)=q±y(±0)ln|x|+b±(y)+o(1)as x±0,y^{\prime}(x)=q_{\pm}y(\pm 0)\ln|x|+b_{\pm}(y)+o(1)\quad\text{as }x\to\pm 0,

where bb_{-} and b+b_{+} are some constants depending on yy.

We will use this proposition for some formal considerations. To proof the norm resolvent convergence of HεH_{\varepsilon} we do really need more subtle estimates of the remainder terms; a stronger version of these asymptotics is presented in Lemma 2. Set yε=(Hεζ)1fy_{\varepsilon}=(H_{\varepsilon}-\zeta)^{-1}f for fL2()f\in L^{2}(\mathbb{R}) and ζ\zeta\in\mathbb{C}\setminus\mathbb{R}. We look for the formal asymptotics of yεy_{\varepsilon}, as ε0\varepsilon\to 0, having the form

yε(x){u(x),if |x|>ε,v0(xε)+v1(xε)εlnε+v2(xε)ε,if |x|<ε.y_{\varepsilon}(x)\sim\begin{cases}u(x),&\text{if }|x|>\varepsilon,\\ v_{0}\left(\tfrac{x}{\varepsilon}\right)+v_{1}\left(\tfrac{x}{\varepsilon}\right)\varepsilon\ln\varepsilon+v_{2}\left(\tfrac{x}{\varepsilon}\right)\varepsilon,&\text{if }|x|<\varepsilon.\end{cases} (3.2)

We also assume that the coupling conditions

[yε]±ε=0,[yε]±ε=0[y_{\varepsilon}]_{\pm\varepsilon}=0,\qquad[y^{\prime}_{\varepsilon}]_{\pm\varepsilon}=0 (3.3)

hold, where []x[\,\cdot\,]_{x} is the jump of a function at point xx. Function yεy_{\varepsilon} is a unique solution of equation

yε′′+(Qε(x)+ε2U(ε1x)+ε1V(ε1x))yε=ζyε+f-y_{\varepsilon}^{\prime\prime}+\big{(}Q_{\varepsilon}(x)+\varepsilon^{-2}U(\varepsilon^{-1}x)+\varepsilon^{-1}V(\varepsilon^{-1}x)\big{)}y_{\varepsilon}=\zeta y_{\varepsilon}+f (3.4)

belonging to the domain of HεH_{\varepsilon}. Since the interval on which the (αδ+βδ)(\alpha\delta^{\prime}+\beta\delta)-like perturbation is localized shrinks to a point, uu must solve the equation

u′′+Qu=ζu+fin {0}.-u^{\prime\prime}+Qu=\zeta u+f\qquad\text{in }\mathbb{R}\setminus\{0\}.

This solution can not be uniquely determined without additional conditions at the origin. One naturally expects that these conditions depend on the perturbation.

Suppose that |x|<ε<a|x|<\varepsilon<a. Then we can as follows rewrite equation (3.4) in the terms of new variable t=x/εt=x/\varepsilon. If we set vε(t)=yε(εt)v^{\varepsilon}(t)=y_{\varepsilon}(\varepsilon t), then

d2vεdt2+(U(t)+εlnεϰ(t)+εV(t))vε=ε2ζvε+ε2f,t.-\frac{d^{2}v^{\varepsilon}}{dt^{2}}+\big{(}U(t)+\varepsilon\ln\varepsilon\,\varkappa(t)+\varepsilon V(t)\big{)}v^{\varepsilon}=\varepsilon^{2}\zeta v^{\varepsilon}+\varepsilon^{2}f,\quad t\in\mathcal{I}. (3.5)

Furthermore, in view of Proposition 1, matching conditions (3.3) imply

v0(±1)+O(εlnε)\displaystyle v_{0}(\pm 1)+O(\varepsilon\ln\varepsilon) =u(±0)+O(ε1/2),\displaystyle=u(\pm 0)+O(\varepsilon^{1/2}),
ε1v0(±1)+v1(±1)lnε+v2(±1)\displaystyle\varepsilon^{-1}v_{0}^{\prime}(\pm 1)+v_{1}^{\prime}(\pm 1)\ln\varepsilon+v_{2}^{\prime}(\pm 1) =q±u(±0)lnε+b±(u)+o(1).\displaystyle=q_{\pm}u(\pm 0)\ln\varepsilon+b_{\pm}(u)+o(1).

In particular, we have

v0(±1)=u(±0),v0(±1)=0,v1(±1)=q±u(±0),v2(±1)=b±(u).v_{0}(\pm 1)=u(\pm 0),\quad v_{0}^{\prime}(\pm 1)=0,\quad v_{1}^{\prime}(\pm 1)=q_{\pm}u(\pm 0),\quad v_{2}^{\prime}(\pm 1)=b_{\pm}(u). (3.6)

Substituting (3.2) for |x|<ε|x|<\varepsilon into (3.5) and applying (3.6) yield

v0′′+Uv0=0,t,\displaystyle-v_{0}^{\prime\prime}+Uv_{0}=0,\;\;t\in\mathcal{I}, v0(1)=0,v0(1)=0;\displaystyle v_{0}^{\prime}(-1)=0,\quad v_{0}^{\prime}(1)=0; (3.7)
v1′′+Uv1=ϰv0,t,\displaystyle-v_{1}^{\prime\prime}+Uv_{1}=-\varkappa v_{0},\;\;t\in\mathcal{I}, v1(1)=qu(0),v1(1)=q+u(+0);\displaystyle v_{1}^{\prime}(-1)=q_{-}u(-0),\;\;v_{1}^{\prime}(1)=q_{+}u(+0); (3.8)
v2′′+Uv2=Vv0,t,\displaystyle-v_{2}^{\prime\prime}+Uv_{2}=-Vv_{0},\;\;t\in\mathcal{I}, v2(1)=b(u),v2(1)=b+(u).\displaystyle v_{2}^{\prime}(-1)=b_{-}(u),\;\;v_{2}^{\prime}(1)=b_{+}(u). (3.9)

Let us first suppose that potential UU is resonant. Since the supports of UU is contained in \mathcal{I}, a half-bound state hh is constant outside \mathcal{I} and its restriction to \mathcal{I} is a non-trivial solution of the boundary value problem

h′′+Uh=0,t,h(1)=0,h(1)=0.-h^{\prime\prime}+Uh=0,\quad t\in\mathcal{I},\qquad h^{\prime}(-1)=0,\quad h^{\prime}(1)=0. (3.10)

Moreover h(±)=h(±1)h(\pm\infty)=h(\pm 1) and hence h0(1)=1h_{0}(-1)=1 and h0(1)=θh_{0}(1)=\theta. Then the equation in (3.7) has a one-parameter family of solutions v0=ch0v_{0}=ch_{0}. But owing to v0(1)=u(0)v_{0}(-1)=u(-0), we have

v0=u(0)h0.v_{0}=u(-0)h_{0}. (3.11)

Hence v0(1)=u(0)h0(1)=θu(0)v_{0}(1)=u(-0)h_{0}(1)=\theta u(-0). On the other hand, v0(1)=u(+0)v_{0}(1)=u(+0) by (3.6). From this we deduce

u(+0)=θu(0).u(+0)=\theta u(-0). (3.12)

Next, problem (3.8) is solvable if and only if

θq+u(+0)qu(0)=u(0)ϰh02𝑑x,\theta q_{+}u(+0)-q_{-}u(-0)=u(-0)\int_{\mathcal{I}}\varkappa h_{0}^{2}\,dx, (3.13)

because the corresponding homogeneous problem possesses non-trivial solutions. The last condition can be easy obtained by multiplying the equation in (3.8) by h0h_{0} and integrating by parts. Combining (3.12) and (3.13) gives us

{u(+0)θu(0)=0,θq+u(+0)(q+ϰh02𝑑x)u(0)=0.\begin{cases}\phantom{\theta q_{+}}u(+0)-\kern 6.0pt\theta u(-0)=0,\\ \theta q_{+}u(+0)-\left(q_{-}+\int_{\mathcal{I}}\varkappa h_{0}^{2}\,dx\right)u(-0)=0.\end{cases} (3.14)

The linear system admits a nonzero solution (u(0),u(+0))(u(-0),u(+0)) if and only if

θ2q+q=ϰh02𝑑x.\theta^{2}q_{+}-q_{-}=\int_{\mathcal{I}}\varkappa h_{0}^{2}\,dx. (3.15)

If (3.13) holds, then (3.8) has a one-parameter family of solutions v1=v1+c1h0v_{1}=v_{1}^{*}+c_{1}h_{0}. Let us fix v1v_{1} such that v1(1)=0v_{1}(-1)=0; this is possible, because h0(1)0h_{0}(-1)\neq 0.

We at last turn to problem (3.9). Multiplying the equation in (3.9) by half-bound state h0h_{0} and integrating by parts, we can similarly compute the solvability condition

θb+(u)b(u)=Vv0h0𝑑t.\theta b_{+}(u)-b_{-}(u)=\int_{\mathcal{I}}Vv_{0}h_{0}\,dt. (3.16)

We can choose v2v_{2} to satisfy v2(1)=0v_{2}(-1)=0. Recalling now (3.11), we can rewrite (3.16) in the form

θb+(u)b(u)=u(0)Vh02𝑑t.\theta b_{+}(u)-b_{-}(u)=u(-0)\int_{\mathcal{I}}Vh_{0}^{2}\,dt. (3.17)

Therefore if potential UU is resonant and (3.15) holds, then the leading term uu of asymptotics (3.2) must solve the problem

u′′+Qu=ζu+fin {0},u(+0)θu(0)=0,θb+(u)b(u)=μu(0),\begin{gathered}-u^{\prime\prime}+Qu=\zeta u+f\qquad\text{in }\mathbb{R}\setminus\{0\},\\ u(+0)-\theta u(-0)=0,\qquad\theta b_{+}(u)-b_{-}(u)=\mu u(-0),\end{gathered} (3.18)

where μ\mu is given by (1.7). The coupling conditions at the origin agree with (1.9) in view of Remark 2.

In the case when either UU has no zero-energy resonance or else UU is resonant, but (3.15) does not hold, both the values u(0)u(-0) and u(+0)u(+0) equal zero. Indeed, if UU is not resonant, then problem (3.7) has only trivial solution v0v_{0} and then the first condition in (3.6) implies u(0)=0u(0)=0. On the other hand, if UU is resonant, but (3.15) does not hold, then system (3.14) has a unique solution u(0)=u(+0)=0u(-0)=u(+0)=0. Hence uu should be a solution of the problem

u′′+Qu=ζu+fin {0},u(0)=0.-u^{\prime\prime}+Qu=\zeta u+f\quad\text{in }\mathbb{R}\setminus\{0\},\quad u(0)=0. (3.19)

3.2. Improvement of asymptotics

We will again focus our attention on the case of resonant potential UU. Our aim is to construct an element uεdomHεu_{\varepsilon}\in\mathop{\rm dom}H_{\varepsilon} that approximates yε=(Hεζ)1fy_{\varepsilon}=(H_{\varepsilon}-\zeta)^{-1}f.

From now on, W2k(Ω)W_{2}^{k}(\Omega) and W2k,loc(Ω)W_{2}^{k,loc}(\Omega) stand for the Sobolev spaces and f\|f\| stands for L2()L^{2}(\mathbb{R})-norm of a function ff. To obtain the uniform approximation of yεy_{\varepsilon} in L2()L^{2}(\mathbb{R}) with respect to ff, we will refine asymptotics (3.2). Let zεz_{\varepsilon} be a solution of the Cauchy problem

z′′+U(t)z=f(εt),z(1)=0,z(1)=0.-z^{\prime\prime}+U(t)z=f(\varepsilon t),\qquad z(-1)=0,\quad z^{\prime}(-1)=0. (3.20)

We introduce the function

wε(x)={u(x),if |x|>ε,v0(xε)+v1(xε)εlnε+v2(xε)ε+ε2zε(xε),if |x|<ε.w_{\varepsilon}(x)=\begin{cases}u(x),&\text{if }|x|>\varepsilon,\\ v_{0}\left(\tfrac{x}{\varepsilon}\right)+v_{1}\left(\tfrac{x}{\varepsilon}\right)\varepsilon\ln\varepsilon+v_{2}\left(\tfrac{x}{\varepsilon}\right)\varepsilon+\varepsilon^{2}z_{\varepsilon}\left(\tfrac{x}{\varepsilon}\right),&\text{if }|x|<\varepsilon.\end{cases} (3.21)

Note that wεw_{\varepsilon} is not in general smooth enough to belong to the domain of HεH_{\varepsilon}; by construction, approximation wεw_{\varepsilon} belongs to W2,loc2({ε,ε})W_{2,loc}^{2}(\mathbb{R}\setminus\{-\varepsilon,\varepsilon\}) and has jump discontinuities at the points x=±εx=\pm\varepsilon. We will show that the jumps of wεw_{\varepsilon} and its first derivative are small enough uniformly on ff, and therefore there exists a corrector ρε\rho_{\varepsilon} with the infinitesimal L2L^{2}-norm, as ε0\varepsilon\to 0, such that wε+ρεdomHεw_{\varepsilon}+\rho_{\varepsilon}\in\mathop{\rm dom}H_{\varepsilon}.

We introduce two cut-functions ξ\xi and η\eta that are smooth outside the origin and have compact supports contained in [0,12a][0,\frac{1}{2}a], where aa is the same as in (1.2). In addition, ξ(+0)=1\xi(+0)=1, ξ(+0)=0\xi^{\prime}(+0)=0, η(+0)=0\eta(+0)=0 and η(+0)=1\eta^{\prime}(+0)=1. Let us set

ρε(x)=[wε]εξ(xε)[wε]εη(xε)[wε]εξ(xε)[wε]εη(xε).\rho_{\varepsilon}(x)=[w_{\varepsilon}]_{-\varepsilon}\,\xi(-x-\varepsilon)-[w_{\varepsilon}^{\prime}]_{-\varepsilon}\,\eta(-x-\varepsilon)-[w_{\varepsilon}]_{\varepsilon}\,\xi(x-\varepsilon)-[w_{\varepsilon}^{\prime}]_{\varepsilon}\,\eta(x-\varepsilon). (3.22)

It is easy to check that [ρε(k)]±ε=[wε(k)]±ε[\rho_{\varepsilon}^{(k)}]_{\pm\varepsilon}=-[w_{\varepsilon}^{(k)}]_{\pm\varepsilon} for k=0,1k=0,1. Moreover, ρε(x)=0\rho_{\varepsilon}(x)=0 for x(ε,ε)x\in(-\varepsilon,\varepsilon). From this we conclude that wε+ρεW2,loc2()w_{\varepsilon}+\rho_{\varepsilon}\in W_{2,loc}^{2}(\mathbb{R}), hence that wε+ρεdomHεw_{\varepsilon}+\rho_{\varepsilon}\in\mathop{\rm dom}H_{\varepsilon}.

3.3. Some uniform bounds

Recall that in Subsection 3.1 we have actually derived u=(ζ)1fu=(\mathcal{H}-\zeta)^{-1}f. Hence uW2,loc2((b,b))u\in W^{2}_{2,loc}(\mathbb{R}\setminus(-b,b)) for any b>0b>0. By the Sobolev imbedding theorems, function uu is continuously differentiable on {0}\mathbb{R}\setminus\{0\}. In addition, the estimates hold

uc1f,uC1(K)c2(K)f\|u\|\leq c_{1}\|f\|,\qquad\|u\|_{C^{1}(K)}\leq c_{2}(K)\|f\| (3.23)

for any compact set KK that does not contain the origin.

Lemma 2.

The following estimates

|u(x)u(±0)|C1f|xln|x||,\displaystyle|u(x)-u(\pm 0)|\leq C_{1}\|f\|\,|x\ln|x||, (3.24)
|u(x)q±u(±0)ln|x|b±(u)|C2f|x|1/2\displaystyle\big{|}u^{\prime}(x)-q_{\pm}u(\pm 0)\ln|x|-b_{\pm}(u)\big{|}\leq C_{2}\|f\|\,|x|^{1/2} (3.25)

hold as x±0x\to\pm 0, where b±b_{\pm} are linear bounded functionals on dom\mathop{\rm dom}\mathcal{H}. In addition,

|b±(u)|C3f.|b_{\pm}(u)|\leq C_{3}\|f\|. (3.26)

The constants CkC_{k} do not depend on ff.

Proof.

We will prove (3.24)–(3.26) on the positive half-line only. For the case x0x\to-0 the proof is similar. In view of (1.2), we have

u′′=q+xuζufu^{\prime\prime}=\frac{q_{+}}{x}\,u-\zeta u-f (3.27)

for x(0,a)x\in(0,a). Temporarily write fζ=ζu+ff_{\zeta}=\zeta u+f. Consequently

u(x)=q+xau(s)s𝑑s+xafζ(s)𝑑s+u(a),\displaystyle u^{\prime}(x)=-q_{+}\int_{x}^{a}\frac{u(s)}{s}\,ds+\int_{x}^{a}f_{\zeta}(s)\,ds+u^{\prime}(a), (3.28)
u(x)=q+xasxsu(s)𝑑sxa(sx)fζ(s)𝑑s+u(a)+u(a)(xa).\displaystyle u(x)=q_{+}\int_{x}^{a}\frac{s-x}{s}\,u(s)\,ds-\int_{x}^{a}(s-x)f_{\zeta}(s)\,ds+u(a)+u^{\prime}(a)(x-a). (3.29)

From this we see in particular that there exists the finite limit value

u(+0)=q+0au(s)𝑑s+x0afζ(s)𝑑s+u(a)au(a)u(+0)=q_{+}\int_{0}^{a}u(s)\,ds+x\int_{0}^{a}f_{\zeta}(s)\,ds+u(a)-au^{\prime}(a) (3.30)

not only for an element of dom\mathop{\rm dom}\mathcal{H}, but for any L2(+)L^{2}(\mathbb{R}_{+})-solution of (3.27). In fact, the most singular (as x+0x\to+0) integral

xasxsu(s)𝑑s\int_{x}^{a}\frac{s-x}{s}\,u(s)\,ds

converges to 0au(s)𝑑s\int_{0}^{a}u(s)\,ds by Lebesgue’s dominated convergence theorem, because

|sxsχ(x,a)(s)|1for s(0,a).\left|\frac{s-x}{s}\,\chi_{(x,a)}(s)\right|\leq 1\qquad\text{for \ }s\in(0,a).

Here χ(x,a)\chi_{(x,a)} is the characteristic function of interval (x,a)(x,a). Combining (3.30) and the second inequality in (3.23), we discover

uC0([0,a])cf.\|u\|_{C^{0}([0,a])}\leq c\|f\|. (3.31)

Subtracting (3.30) from (3.29), we can represent the difference as

u(x)u(+0)=q+xxau(s)s𝑑s+0x(sfζ(s)q+u(s))𝑑s+xxafζ(s)𝑑s+u(a)x.u(x)-u(+0)=-q_{+}x\int_{x}^{a}\frac{u(s)}{s}\,ds\\ +\int_{0}^{x}\big{(}sf_{\zeta}(s)-q_{+}u(s)\big{)}\,ds+x\int_{x}^{a}f_{\zeta}(s)\,ds+u^{\prime}(a)x.

Since

xau(s)s𝑑s=u(+0)x(lnalnx)+xau(s)u(+0)s𝑑s,\int_{x}^{a}\frac{u(s)}{s}\,ds=u(+0)\,x(\ln a-\ln x)+\int_{x}^{a}\frac{u(s)-u(+0)}{s}\,ds, (3.32)

we finally have

u(x)u(+0)=q+u(+0)x(lnxlna)+0x(sfζ(s)q+u(s))𝑑s+xxafζ(s)𝑑s+u(a)xq+xxau(s)u(+0)s𝑑s.u(x)-u(+0)=q_{+}u(+0)\,x(\ln x-\ln a)+\int_{0}^{x}\big{(}sf_{\zeta}(s)-q_{+}u(s)\big{)}\,ds\\ +x\int_{x}^{a}f_{\zeta}(s)\,ds+u^{\prime}(a)x-q_{+}x\int_{x}^{a}\frac{u(s)-u(+0)}{s}\,ds. (3.33)

Hence

|u(x)u(+0)|c1fx|lnx|+|q+|xxa|u(s)u(+0)|s𝑑s,|u(x)-u(+0)|\leq c_{1}\|f\|\,x|\ln x|+|q_{+}|\,x\int_{x}^{a}\frac{|u(s)-u(+0)|}{s}\,ds,

where we employed (3.23) and (3.31) to obtain the estimates

|u(+0)|\displaystyle|u(+0)| +|u(a)|c2f,fζ=ζu+fc3f,\displaystyle+|u^{\prime}(a)|\leq c_{2}\,\|f\|,\qquad\|f_{\zeta}\|=\|\zeta u+f\|\leq c_{3}\,\|f\|,
|0xu(s)𝑑s|xsups(0,x)|u(s)|c4xf,\displaystyle\left|\int_{0}^{x}u(s)\,ds\right|\leq x\sup_{s\in(0,x)}|u(s)|\leq c_{4}x\|f\|,
|0xsfζ(s)𝑑s|x0x|fζ(s)|𝑑sx3/2f.\displaystyle\left|\int_{0}^{x}sf_{\zeta}(s)\,ds\right|\leq x\int_{0}^{x}|f_{\zeta}(s)|\,ds\leq x^{3/2}\|f\|.

Consequently Gronwall’s inequality implies

|u(x)u(+0)|c1fe|q+|x(lnalnx)x|lnx|C1fx|lnx|,|u(x)-u(+0)|\leq c_{1}\|f\|\,e^{|q_{+}|\,x(\ln a-\ln x)}\,x|\ln{x}|\leq C_{1}\|f\|\,x|\ln{x}|, (3.34)

as x+0x\to+0, which establishes (3.24). Applying (3.32) to (3.28), we find

u(x)=q+u(+0)(lnxlna)q+xau(s)u(+0)s𝑑s+xafζ(s)𝑑s+u(a)=q+u(+0)lnx+b+(u)+r(x,u),u^{\prime}(x)=q_{+}u(+0)(\ln x-\ln a)-q_{+}\int_{x}^{a}\frac{u(s)-u(+0)}{s}\,ds\\ +\int_{x}^{a}f_{\zeta}(s)\,ds+u^{\prime}(a)=q_{+}u(+0)\ln x+b_{+}(u)+r(x,u),

where

b+(u)=u(a)q+u(+0)lna+0afζ(s)𝑑sq+0au(s)u(+0)s𝑑s,\displaystyle b_{+}(u)=u^{\prime}(a)-q_{+}u(+0)\ln a+\int_{0}^{a}f_{\zeta}(s)\,ds-q_{+}\int_{0}^{a}\frac{u(s)-u(+0)}{s}\,ds,
r(x,u)=q+0xu(s)u(+0)s𝑑s0xfζ(s)𝑑s.\displaystyle r(x,u)=q_{+}\int_{0}^{x}\frac{u(s)-u(+0)}{s}\,ds-\int_{0}^{x}f_{\zeta}(s)\,ds.

Thus formulas (3.23), (3.24) and (3.31) provide the bounds

|r(x,u)||q+|0x|u(s)u(+0)|s𝑑s+|ζ|0x|u|𝑑s+0x|f|𝑑sC1f0x|lns|𝑑s+c4(u+f)x1/2C2fx1/2,\displaystyle\begin{aligned} |r(x,u)|&\leq|q_{+}|\int_{0}^{x}\frac{|u(s)-u(+0)|}{s}\,ds+|\zeta|\int_{0}^{x}|u|\,ds+\int_{0}^{x}|f|\,ds\\ &\leq C_{1}\,\|f\|\int_{0}^{x}|\ln s|\,ds+c_{4}(\|u\|+\|f\|)x^{1/2}\leq C_{2}\|f\|x^{1/2},\end{aligned}
|b+(u)|c5(|u(a)|+|u(+0)|)+c6fζ+|q+|0a|u(s)u(+0)|s𝑑sc7f+c8f0a|lns|𝑑sC3f,\displaystyle\begin{aligned} |b_{+}(u)|&\leq c_{5}(|u^{\prime}(a)|+|u(+0)|)+c_{6}\|f_{\zeta}\|\\ &+|q_{+}|\int_{0}^{a}\frac{|u(s)-u(+0)|}{s}\,ds\leq c_{7}\,\|f\|+c_{8}\|f\|\int_{0}^{a}|\ln s|\,ds\leq C_{3}\|f\|,\end{aligned}

which establishes (3.25) and (3.26). ∎

By construction functions vkv_{k} in (3.2) belong to W22()W_{2}^{2}(\mathcal{I}). We will show that their W22W_{2}^{2}-norms can be estimated by the L2L^{2}-norm of ff.

Lemma 3.

Assume that v0v_{0}, v1v_{1} and v2v_{2} are solunions of (3.7), (3.8) and (3.9) respectively. Suppose that these solutions are chosen so that v0(1)=u(0)v_{0}(-1)=u(-0), v1(1)=0v_{1}(-1)=0 and v2(1)=0v_{2}(-1)=0. Then

vkW22()C1f\|v_{k}\|_{W_{2}^{2}(\mathcal{I})}\leq C_{1}\|f\| (3.35)

for all fL2()f\in L^{2}(\mathbb{R}) and k=0,1,2k=0,1,2, the constant C1C_{1} being independent of ff.

Let zεz_{\varepsilon} be the solution of (3.20). Then zεz_{\varepsilon} also belongs to W22()W_{2}^{2}(\mathcal{I}), with the estimate

zεW22()C2ε1/2f,\|z_{\varepsilon}\|_{W_{2}^{2}(\mathcal{I})}\leq C_{2}\varepsilon^{-1/2}\|f\|, (3.36)

where C2C_{2} does not depend of ff and ε\varepsilon.

Proof.

It is evident from (3.11) and Lemma 2 that v0W22()c|u(0)|C1f\|v_{0}\|_{W_{2}^{2}(\mathcal{I})}\leq c|u(-0)|\leq C_{1}\|f\|. To prove this estimate for v1v_{1} and v2v_{2}, we construct below representations for the desired solutions. Let ω\omega be a solution of the Cauchy problem

ω′′+Uω=ϰh0,t,ω(1)=0,ω(1)=q.-\omega^{\prime\prime}+U\omega=-\varkappa h_{0},\;\;t\in\mathcal{I},\qquad\omega(-1)=0,\quad\omega^{\prime}(-1)=q_{-}.

We set v1=u(0)ωv_{1}=u(-0)\omega. This function solves the equation in (3.8) and v1(1)=qu(0)v_{1}^{\prime}(-1)=q_{-}u(-0). The boundary condition at t=1t=1 also holds, because multiplying the equation for ω\omega by half-bound state h0h_{0} and integrating by parts twice yield

θω(1)=q+ϰh02𝑑x.\theta\omega^{\prime}(1)=q_{-}+\int_{\mathcal{I}}\varkappa h_{0}^{2}\,dx.

From this we have

v1(1)=u(0)ω(1)=θ1u(0)(q+ϰh02𝑑x)=q+θu(0)=q+u(+0),v_{1}^{\prime}(1)=u(-0)\omega^{\prime}(1)=\theta^{-1}u(-0)\left(q_{-}+\int_{\mathcal{I}}\varkappa h_{0}^{2}\,dx\right)=q_{+}\theta u(-0)=q_{+}u(+0),

by (1.8). Next, solution v2v_{2} of (3.9) can be written as

v2(t)=b(u)ω(t)+u(0)Ω(t),v_{2}(t)=b_{-}(u)\omega(t)+u(-0)\Omega(t), (3.37)

where Ω\Omega solves the problem

Ω′′+UΩ=Vh0,t,Ω(1)=0,Ω(1)=0.-\Omega^{\prime\prime}+U\Omega=-Vh_{0},\;\;t\in\mathcal{I},\qquad\Omega(-1)=0,\quad\Omega^{\prime}(-1)=0.

Note that Ω(1)=θ1Vh02𝑑t\Omega^{\prime}(1)=\theta^{-1}\int_{\mathbb{R}}Vh_{0}^{2}\,dt. This equality can be obtained by multiplying equation y′′+Uy=Vh0-y^{\prime\prime}+Uy=-Vh_{0} by half-bound state h0h_{0} and integrating by parts. So we have v2(1)=b(u)ω(1)+u(0)Ω(1)=0v_{2}(-1)=b_{-}(u)\omega(-1)+u(-0)\Omega(-1)=0, v2(1)=b(u)ω(1)+u(0)Ω(1)=b(u)v_{2}^{\prime}(-1)=b_{-}(u)\omega^{\prime}(-1)+u(-0)\Omega^{\prime}(-1)=b_{-}(u) and

v2(1)=b(u)ω(1)+u(0)Ω(1)=θ1(b(u)+u(0)Vh02𝑑t)=b+(u)v_{2}^{\prime}(1)=b_{-}(u)\omega^{\prime}(1)+u(-0)\Omega^{\prime}(1)=\theta^{-1}\left(b_{-}(u)+u(-0)\int_{\mathcal{I}}Vh_{0}^{2}\,dt\right)=b_{+}(u)

in view of coupling condition (3.17). Hence v2v_{2} of the form (3.37) is a solution of (3.9) such that v2(1)=0v_{2}(-1)=0. Estimate (3.35) for k=1,2k=1,2 follows from the explicit form of v1v_{1}, v2v_{2}, bounds (3.23) and Lemma 2.

Since UL()U\in L^{\infty}(\mathbb{R}), solution zεz_{\varepsilon} of the Cauchy problem satisfies

zεW22()c1f(ε)L2().\|z_{\varepsilon}\|_{W_{2}^{2}(\mathcal{I})}\leq c_{1}\|f(\varepsilon\,\cdot)\|_{L^{2}(\mathcal{I})}.

We also have

11|f(εt)|2𝑑tc2ε1εε|f(τ)|2𝑑τc3ε1f2.\int_{-1}^{1}|f(\varepsilon t)|^{2}\,dt\leq c_{2}\varepsilon^{-1}\int_{-\varepsilon}^{\varepsilon}|f(\tau)|^{2}\,d\tau\leq c_{3}\varepsilon^{-1}\|f\|^{2}.

Therefore (3.36) follows from the last bound. ∎

Lemma 4.

Assume that function ρε\rho_{\varepsilon} is given by (3.22). There exist constants C1C_{1} and C2C_{2} being independent of ff such that

sup|x|>ε(|ρε(x)|+|ρε′′(x)|)C1ε1/2f,\displaystyle\sup_{|x|>\varepsilon}(|\rho_{\varepsilon}(x)|+|\rho^{\prime\prime}_{\varepsilon}(x)|)\leq C_{1}\varepsilon^{1/2}\|f\|, (3.38)
QρεC2ε1/4f.\displaystyle\|Q\rho_{\varepsilon}\|\leq C_{2}\varepsilon^{1/4}\|f\|. (3.39)
Proof.

To prove (3.38) it suffices to show

|[wε]ε|+|[wε]ε|+|[wε]ε|+|[wε]ε|cε1/2f,\big{|}[w_{\varepsilon}]_{-\varepsilon}\big{|}+\big{|}[w_{\varepsilon}]_{\varepsilon}\big{|}+|[w^{\prime}_{\varepsilon}]_{-\varepsilon}\big{|}+\big{|}[w^{\prime}_{\varepsilon}]_{\varepsilon}\big{|}\leq c\varepsilon^{1/2}\|f\|,

since functions ξ\xi and η\eta in (3.22) are smooth and bounded together with all their derivatives, if |x|>ε|x|>\varepsilon. Combining Lemmas 2, 3 and the continuity of embedding W22()C1()W_{2}^{2}(\mathcal{I})\subset C^{1}(\mathcal{I}), we conclude that

|[wε]ε|=|v0(1)u(ε)|=|u(0)u(ε)|c1fε|lnε|,\displaystyle\big{|}[w_{\varepsilon}]_{-\varepsilon}\big{|}=|v_{0}(-1)-u(-\varepsilon)|=|u(-0)-u(-\varepsilon)|\leq c_{1}\|f\|\,\varepsilon|\ln\varepsilon|,
|[wε]ε|=|u(ε)qu(0)lnεb(u)|c2fε1/2,\displaystyle\big{|}[w^{\prime}_{\varepsilon}]_{-\varepsilon}\big{|}=|u^{\prime}(-\varepsilon)-q_{-}u(-0)\ln\varepsilon-b_{-}(u)|\leq c_{2}\|f\|\,\varepsilon^{1/2},
|[wε]ε|=|u(ε)u(+0)v1(1)εlnεv2(1)εzε(ε)ε2|c3fε|lnε|,\displaystyle\big{|}[w_{\varepsilon}]_{\varepsilon}\big{|}=|u(\varepsilon)-u(+0)-v_{1}(1)\,\varepsilon\ln\varepsilon-v_{2}(1)\,\varepsilon-z_{\varepsilon}(\varepsilon)\,\varepsilon^{2}|\leq c_{3}\|f\|\,\varepsilon|\ln\varepsilon|,
|[wε]ε|=|u(ε)q+u(+0)lnεb+(u)zε(ε)ε|c4fε1/2,\displaystyle\big{|}[w^{\prime}_{\varepsilon}]_{\varepsilon}\big{|}=|u^{\prime}(\varepsilon)-q_{+}u(+0)\ln\varepsilon-b_{+}(u)-z^{\prime}_{\varepsilon}(\varepsilon)\,\varepsilon|\leq c_{4}\|f\|\,\varepsilon^{1/2},

which establishes (3.38).

Next, let us fix γ(0,12)\gamma\in(0,\frac{1}{2}). Since |η(x)|c|x||\eta(x)|\leq c|x| as |x|0|x|\to 0, we have

supε<|x|<εγ|ρε(x)|c5(|[wε]ε|+|[wε]ε|+(|[wε]ε|+|[wε]ε|)εγ)c6(ε|lnε|+εγ+1/2)fc7εγ+1/2f\sup_{\varepsilon<|x|<\varepsilon^{\gamma}}|\rho_{\varepsilon}(x)|\leq c_{5}\big{(}|[w_{\varepsilon}]_{-\varepsilon}|+|[w_{\varepsilon}]_{\varepsilon}|+(|[w^{\prime}_{\varepsilon}]_{-\varepsilon}|+|[w^{\prime}_{\varepsilon}]_{\varepsilon}|)\,\varepsilon^{\gamma}\big{)}\\ \leq c_{6}(\varepsilon|\ln\varepsilon|+\varepsilon^{\gamma+1/2})\|f\|\leq c_{7}\varepsilon^{\gamma+1/2}\|f\| (3.40)

for ε<|x|<εγ\varepsilon<|x|<\varepsilon^{\gamma}. Recall that ρε(x)=0\rho_{\varepsilon}(x)=0 for |x|<ε|x|<\varepsilon and |x|>a|x|>a, provided ε\varepsilon is small enough. Then utilizing estimates (3.38) and (3.40), we obtain the bound

Qρε2\displaystyle\allowdisplaybreaks\|Q\rho_{\varepsilon}\|^{2} =ε<|x|<aQ2|ρε|2𝑑xmax{|q|,|q+|}ε<|x|<ax2|ρε|2𝑑x\displaystyle=\int\limits_{\varepsilon<|x|<a}\kern-4.0ptQ^{2}|\rho_{\varepsilon}|^{2}\,dx\leq\max\{|q_{-}|,|q_{+}|\}\kern-3.0pt\int\limits_{\varepsilon<|x|<a}\kern-4.0ptx^{-2}|\rho_{\varepsilon}|^{2}\,dx
c8(ε<|x|<εγx2|ρε|2𝑑x+εγ<|x|<ax2|ρε|2𝑑x)\displaystyle\leq c_{8}\left(\int\limits_{\varepsilon<|x|<\varepsilon^{\gamma}}\kern-4.0ptx^{-2}|\rho_{\varepsilon}|^{2}\,dx+\kern-3.0pt\int\limits_{\varepsilon^{\gamma}<|x|<a}\kern-4.0ptx^{-2}|\rho_{\varepsilon}|^{2}\,dx\right)
c8supε<|x|<εγ|ρε(x)|2ε<|x|<εγx2𝑑x+c8supεγ<|x|<ax2|ρε(x)|2εγ<|x|<a𝑑x\displaystyle\leq c_{8}\sup_{\varepsilon<|x|<\varepsilon^{\gamma}}|\rho_{\varepsilon}(x)|^{2}\kern-3.0pt\int\limits_{\varepsilon<|x|<\varepsilon^{\gamma}}\kern-6.0ptx^{-2}\,dx+c_{8}\sup_{\varepsilon^{\gamma}<|x|<a}x^{-2}|\rho_{\varepsilon}(x)|^{2}\kern-3.0pt\int\limits_{\varepsilon^{\gamma}<|x|<a}\kern-6.0ptdx
c9ε2γ+1fε<|x|<εγx2𝑑x+c10ε12γfc11(ε2γ+ε12γ)f.\displaystyle\leq c_{9}\varepsilon^{2\gamma+1}\|f\|\kern-3.0pt\int\limits_{\varepsilon<|x|<\varepsilon^{\gamma}}\kern-6.0ptx^{-2}\,dx+c_{10}\varepsilon^{1-2\gamma}\|f\|\leq c_{11}(\varepsilon^{2\gamma}+\varepsilon^{1-2\gamma})\|f\|.

Assertion (3.39) follows from this inequality, provided γ=1/4\gamma=1/4. ∎

3.4. End of the proof

We showed above that uε=wε+ρεu_{\varepsilon}=w_{\varepsilon}+\rho_{\varepsilon} belongs to the domain of HεH_{\varepsilon}. We will now prove that uεu_{\varepsilon} solves the equation

(Hεζ)uε=f+gε,(H_{\varepsilon}-\zeta)u_{\varepsilon}=f+g_{\varepsilon}, (3.41)

in which remainder term gεg_{\varepsilon} is small in L2L_{2}-norm uniformly with respect to ff. Let us compute gεg_{\varepsilon}. If |x|>ε|x|>\varepsilon, then we have

gε(x)=(d2dx2+Q(x)ζ)(u(x)+ρε(x))f(x)=ρε′′(x)+(Q(x)ζ)ρε(x),g_{\varepsilon}(x)=\big{(}-\tfrac{d^{2}}{dx^{2}}+Q(x)-\zeta\big{)}\big{(}u(x)+\rho_{\varepsilon}(x)\big{)}-f(x)=-\rho_{\varepsilon}^{\prime\prime}(x)+(Q(x)-\zeta)\rho_{\varepsilon}(x),

by (3.18). If |x|<ε|x|<\varepsilon, then

gε(x)\displaystyle g_{\varepsilon}(x) =d2dx2uε(xε)+(ε2U(xε)+ε1lnεϰ(xε)+ε1V(xε)ζ)uε(xε)f(x)\displaystyle=-\frac{d^{2}}{dx^{2}}\,u_{\varepsilon}\left(\tfrac{x}{\varepsilon}\right)+\big{(}\varepsilon^{-2}U\left(\tfrac{x}{\varepsilon}\right)+\varepsilon^{-1}\ln\varepsilon\,\varkappa\left(\tfrac{x}{\varepsilon}\right)+\varepsilon^{-1}V\left(\tfrac{x}{\varepsilon}\right)-\zeta\big{)}u_{\varepsilon}\left(\tfrac{x}{\varepsilon}\right)-f(x)
=ε2(v0′′(xε)+U(xε)v0(xε))\displaystyle=\varepsilon^{-2}\big{(}-v_{0}^{\prime\prime}\left(\tfrac{x}{\varepsilon}\right)+U\left(\tfrac{x}{\varepsilon}\right)v_{0}\left(\tfrac{x}{\varepsilon}\right)\big{)}
+ε1lnε(v1′′(xε)+U(xε)v1(xε)+ϰ(xε)v0(xε))\displaystyle+\varepsilon^{-1}\ln\varepsilon\,\big{(}-v_{1}^{\prime\prime}\left(\tfrac{x}{\varepsilon}\right)+U\left(\tfrac{x}{\varepsilon}\right)v_{1}\left(\tfrac{x}{\varepsilon}\right)+\varkappa\left(\tfrac{x}{\varepsilon}\right)v_{0}\left(\tfrac{x}{\varepsilon}\right)\big{)}
+ε1(v2′′(xε)+U(xε)v2+V(xε)v0(xε))\displaystyle+\varepsilon^{-1}\big{(}-v_{2}^{\prime\prime}\left(\tfrac{x}{\varepsilon}\right)+U\left(\tfrac{x}{\varepsilon}\right)v_{2}+V\left(\tfrac{x}{\varepsilon}\right)v_{0}\left(\tfrac{x}{\varepsilon}\right)\big{)}
zε′′(xε)+U(xε)zε(xε)f(x)\displaystyle-z^{\prime\prime}_{\varepsilon}\left(\tfrac{x}{\varepsilon}\right)+U\left(\tfrac{x}{\varepsilon}\right)z_{\varepsilon}\left(\tfrac{x}{\varepsilon}\right)-f(x)
+lnεϰ(xε)(v1(xε)lnε+v2(xε)+εzε(xε))\displaystyle+\ln\varepsilon\>\varkappa\left(\tfrac{x}{\varepsilon}\right)\big{(}v_{1}\left(\tfrac{x}{\varepsilon}\right)\ln\varepsilon+v_{2}\left(\tfrac{x}{\varepsilon}\right)+\varepsilon z_{\varepsilon}\left(\tfrac{x}{\varepsilon}\right)\big{)}
+V(xε)(v1(xε)lnε+v2(xε)+εzε(xε))ζuε(xε)\displaystyle+V\left(\tfrac{x}{\varepsilon}\right)\big{(}v_{1}\left(\tfrac{x}{\varepsilon}\right)\ln\varepsilon+v_{2}\left(\tfrac{x}{\varepsilon}\right)+\varepsilon z_{\varepsilon}\left(\tfrac{x}{\varepsilon}\right)\big{)}-\zeta u_{\varepsilon}\left(\tfrac{x}{\varepsilon}\right)
=(ϰ(xε)lnε+V(xε))(v1(xε)lnε+v2(xε)+zε(xε)ε)ζuε(xε)\displaystyle=\big{(}\varkappa\left(\tfrac{x}{\varepsilon}\right)\ln\varepsilon+V\left(\tfrac{x}{\varepsilon}\right)\big{)}\big{(}v_{1}\left(\tfrac{x}{\varepsilon}\right)\ln\varepsilon+v_{2}\left(\tfrac{x}{\varepsilon}\right)+z_{\varepsilon}\left(\tfrac{x}{\varepsilon}\right)\varepsilon\big{)}-\zeta u_{\varepsilon}\left(\tfrac{x}{\varepsilon}\right)

by (3.7)–(3.9) and (3.20). Hence we have

gερε′′+ζρε+Qρε+supt(|U(t)||lnε|+|V(t)|)×v1(ε1)lnε+v2(ε1)+εzε(ε1)L2(ε,ε)+|ζ|uε(ε1)L2(ε,ε)c1(ε1/2+ε1/4)f+c2ε1/2|lnε|v1lnε+v2+εzεL2()+|ζ|ε1/2uεL2()cε1/4f\|g_{\varepsilon}\|\leq\|\rho_{\varepsilon}^{\prime\prime}+\zeta\rho_{\varepsilon}\|+\|Q\rho_{\varepsilon}\|+\sup_{t\in\mathcal{I}}\big{(}|U(t)||\ln\varepsilon|+|V(t)|\big{)}\\ \times\|v_{1}(\varepsilon^{-1}\cdot)\ln\varepsilon+v_{2}(\varepsilon^{-1}\cdot)+\varepsilon z_{\varepsilon}(\varepsilon^{-1}\cdot)\|_{L^{2}(-\varepsilon,\varepsilon)}+|\zeta|\,\|u_{\varepsilon}(\varepsilon^{-1}\cdot)\|_{L^{2}(-\varepsilon,\varepsilon)}\\ \leq c_{1}(\varepsilon^{1/2}+\varepsilon^{1/4})\|f\|+c_{2}\varepsilon^{1/2}|\ln\varepsilon|\|v_{1}\ln\varepsilon+v_{2}+\varepsilon z_{\varepsilon}\|_{L^{2}(\mathcal{I})}\\ +|\zeta|\varepsilon^{1/2}\,\|u_{\varepsilon}\|_{L^{2}(\mathcal{I})}\leq c\varepsilon^{1/4}\|f\|

in view of Lemmas 3 and 4. Here we also used inequality

εε|m(xε)|2𝑑xε11|m(t)|2𝑑t=εmL2()2\int_{-\varepsilon}^{\varepsilon}|m\left(\tfrac{x}{\varepsilon}\right)|^{2}\,dx\leq\varepsilon\int_{-1}^{1}|m(t)|^{2}\,dt=\varepsilon\,\|m\|^{2}_{L^{2}(\mathcal{I})}

for any mL2()m\in L^{2}(\mathcal{I}). Therefore (3.41) implies

uεyε=(Hεζ)1gε|ζ|1gεcε1/4f.\|u_{\varepsilon}-y_{\varepsilon}\|=\|(H_{\varepsilon}-\zeta)^{-1}g_{\varepsilon}\|\leq|\zeta|^{-1}\|g_{\varepsilon}\|\leq c\varepsilon^{1/4}\|f\|. (3.42)

Now let us consider the difference

uε(x)u(x)={ρε(x)if |x|>ε,v0(xε)+v1(xε)εlnε+v2(xε)ε+zε(xε)ε2u(x)if |x|<ε.u_{\varepsilon}(x)-u(x)=\begin{cases}\rho_{\varepsilon}(x)&\text{if }|x|>\varepsilon,\\ v_{0}\left(\tfrac{x}{\varepsilon}\right)+v_{1}\left(\tfrac{x}{\varepsilon}\right)\varepsilon\ln\varepsilon+v_{2}\left(\tfrac{x}{\varepsilon}\right)\varepsilon+z_{\varepsilon}\left(\tfrac{x}{\varepsilon}\right)\varepsilon^{2}-u(x)&\text{if }|x|<\varepsilon.\end{cases}

We can as before invoke bound (3.23), Lemmas 3 and 4 to derive

uεuρε+ε1/2v0+v1εlnε+v2ε+zεε2L2()+uL2(ε,ε)c1ε1/2(f+max|x|ε|u(x)|)c2ε1/2f.\|u_{\varepsilon}-u\|\leq\|\rho_{\varepsilon}\|+\varepsilon^{1/2}\|v_{0}+v_{1}\varepsilon\ln\varepsilon+v_{2}\varepsilon+z_{\varepsilon}\varepsilon^{2}\|_{L^{2}(\mathcal{I})}\\ +\|u\|_{L_{2}(-\varepsilon,\varepsilon)}\leq c_{1}\varepsilon^{1/2}(\|f\|+\max_{|x|\leq\varepsilon}|u(x)|)\leq c_{2}\varepsilon^{1/2}\|f\|. (3.43)

Recalling the definitions of yεy_{\varepsilon} and uu, we estimate

(Hεζ)1f(ζ)1f=yεuyεuε+uεuCε1/4f,\|(H_{\varepsilon}-\zeta)^{-1}f-(\mathcal{H}-\zeta)^{-1}f\|=\|y_{\varepsilon}-u\|\leq\|y_{\varepsilon}-u_{\varepsilon}\|+\|u_{\varepsilon}-u\|\leq C\varepsilon^{1/4}\|f\|,

by (3.42) and (3.43). The last bound establishes the norm resolvent convergence of HεH_{\varepsilon} to the operator \mathcal{H} and estimate (1.10), which is the desired conclusion for the case when potential VV is resonant.

If VV is not resonant, function uu in asymptotics (3.2) solves problem (3.19). Since both the value u(0)u(-0) and u(+0)u(+0) are equal zero, uu^{\prime} has no logarithmic singularity at the origin in view of Lemma 2. From this reason the uniform approximation to yεy_{\varepsilon} has the form

uε(x)={u(x)+ρε(x),if |x|>ε,εv2(xε)+ε2zε(xε),if |x|<ε,u_{\varepsilon}(x)=\begin{cases}u(x)+\rho_{\varepsilon}(x),&\text{if }|x|>\varepsilon,\\ \varepsilon v_{2}\left(\tfrac{x}{\varepsilon}\right)+\varepsilon^{2}z_{\varepsilon}\left(\tfrac{x}{\varepsilon}\right),&\text{if }|x|<\varepsilon,\end{cases}

where v2v_{2} solves the problem

v2′′+Uv2=0,t,v2(1)=u(0),v2(1)=u(+0).-v_{2}^{\prime\prime}+Uv_{2}=0,\quad t\in\mathcal{I},\qquad v_{2}^{\prime}(-1)=-u^{\prime}(-0),\quad v_{2}^{\prime}(1)=u^{\prime}(+0).

The rest of the proof is similar to the proof for the previous case.

References

  • [1] R. Loudon. One-dimensional hydrogen atom. American Journal of Physics 27(9) (1959), 649-655.
  • [2] L. K. Haines, D. H. Roberts. One-dimensional hydrogen atom. American Journal of Physics 37(11), (1969), 1145-1154.
  • [3] Andrews, M. Singular potentials in one dimension. American Journal of Physics 44(11), (1976), 1064-1066.
  • [4] C. H. Mehta, S. H. Patil. Bound states of the potential V(r)=Z(r+β)V(r)=-\frac{Z}{(r+\beta)}. Physical Review A 17(1), (1978), 43-46.
  • [5] F. Gesztesy. On the one-dimensional Coulomb Hamiltonian. Journal of Physics A: Mathematical and General 13(3), (1980), 867.
  • [6] M. Klaus. Removing cut-offs from one-dimensional Schrödinger operators. Journal of Physics A: Mathematical and General 13(9), (1980), L295.
  • [7] U. Oseguera and M. de Llano, Two singular potentials: the space-splitting effect. Journal of Mathematical Physics 34, 4575 (1993).
  • [8] M. Moshinsky. Penetrability of a one-dimensional Coulomb potential. Journal of Physics A: Mathematical and General 26 (1993), 2445-2450.
  • [9] R. G. Newton. Comment on ’Penetrability of a one-dimensional Coulomb potential’ by M Moshinsky. Journal of Physics A: Mathematical and General 27 (1994), 4717-4718.
  • [10] M. Moshinsky. Response to “Comment on ’Penetrability of a one-dimensional Coulomb potential’ ” by Roger G Newton. Journal of Physics A: Mathematical and General 27 (1994), 4719-4721.
  • [11] W. Fischer, H. Leschke and P. Müller. The functional-analytic versus the functional-integral approach to quantum Hamiltonians. The one-dimensional hydrogen atom. Journal of Physics A: Mathematical and General 36 (1995), 2313-2323.
  • [12] P. Kurasov. On the Coulomb potential in one dimension. Journal of Physics A: Mathematical and General 29 (1996) 1767-1771.
  • [13] W. Fischer, H. Leschke and P. Muller. Comment on ’On the Coulomb potential in one dimension’ by P Kurasov. Journal of Physics A: Mathematical and General 30 (1997) 5579-5581.
  • [14] P. Kurasov. Response to “Comment on ’On the Coulomb potential in one dimension’ ” by Fischer, Leschke and Muller. Journal of Physics A: Mathematical and General 30 (1997) 5583-5589.
  • [15] C. R. de Oliveira and A. A. Verri, Self-adjoint extensions of Coulomb systems in 11, 22 and 33 dimensions. Annals of Physics 324 (2009) 251-266.
  • [16] B. Bodenstorfer, A. Dijksma and H. Langer. Dissipative eigenvalue problems for a Sturm-Liouville operator with a singular potential. Proceedings of the Royal Society of Edinburgh Section A: Mathematics 130(6) (2000), 1237-1257.
  • [17] Yu. D. Golovaty, R. O. Hryniv. On norm resolvent convergence of Schrödinger operators with δ\delta^{\prime}-like potentials. Journal of Physics A: Mathematical and Theoretical 43 (2010) 155204 (14pp) (A Corrigendum: 2011 J. Phys. A: Math. Theor. 44 049802)
  • [18] Yu. Golovaty. Schrödinger operators with (αδ+βδ)(\alpha\delta^{\prime}+\beta\delta)-like potentials: norm resolvent convergence and solvable models, Methods of Funct. Anal. Topology (3) 18 (2012), 243–255.
  • [19] Yu. D. Golovaty and R. O. Hryniv. Norm resolvent convergence of singularly scaled Schrödinger operators and δ\delta^{\prime}-potentials. Proceedings of the Royal Society of Edinburgh: Section A Mathematics 143 (2013), 791-816.
  • [20] Yu. Golovaty, 1D Schrödinger Operators with Short Range Interactions: Two-Scale Regularization of Distributional Potentials. Integral Equations and Operator Theory 75(3) (2013), 341-362.
  • [21] A. V. Zolotaryuk. Two-parametric resonant tunneling across the δ(x)\delta^{\prime}(x) potential. Adv. Sci. Lett. 1 (2008), 187-191.
  • [22] A. V. Zolotaryuk. Point interactions of the dipole type defined through a three-parametric power regularization. Journal of Physics A: Mathematical and Theoretical 43 (2010), 105302.
  • [23] Yu. Golovaty. Two-parametric δ\delta^{\prime}-interactions: approximation by Schrödinger operators with localized rank-two perturbations. Journal of Physics A: Mathematical and Theoretical 51(25) (2018), 255202.
  • [24] Yu. Golovaty. Schrödinger operators with singular rank-two perturbations and point interactions. Integr. Equ. Oper. Theory 90:57 (2018).