1 Introduction: Quantitative Bounds Versus Weakly Coupled Bound States

In this paper, we study operators of the form

$$\begin{aligned} T(p) + V \end{aligned}$$

where \(p=-i\nabla \) is the quantum-mechanical momentum operator and the kinetic energy symbol \(T:{\mathbb {R}}^d\rightarrow [0,\infty )\) is a measurable function. See Sect. 3 for the precise definition of \(T(p)+V\) and the condition we need on the kinetic energy symbol T and the interaction potential V. There has been an enormous amount of interest in the study of bound states for such operators. Usually, in standard quantum mechanics the symbol is given by \(T(\eta )=\eta ^2\), that is, the kinetic energy is given by the the Laplacian \(p^2=-\Delta \). In this case, it is well-known that quantum mechanics in three and more dimensions is quite different from one and two dimensions. In three and more dimensions the perturbed operator \(p^2+V\) can be unitarily equivalent to the free operator \(p^2\) for potentials V which are small in some sense [40, Theorem XIII.27], and for a suitable class of potentials the famous Cwikel–Lieb–Rozenblum bound (CLR) holds, which gives a bound on the number of bound states in terms of a semi-classical phase-space volume, see [11, 30, 41, 42]. The works of Cwikel and Lieb had been motivated by Simon [48], the work of Rozenblum by the St. Petersburg school of mathematical physics around Birman and Solomyak, whose work had been virtually unnoticed in the west until the mid 1970s.Footnote 1 Rozenblum’s paper [41] was an announcement of his result and, typically for the journal, did not contain any proofs. A version of his result also appeared in the summer school lecture notes by Birman and Solomyak [7]. The version with full proofs was published in [42]. Similarly, Lieb’s paper [29] is announcement of his result and the details of his proof had been published later in [30, 51]

In one and two dimensions, arbitrarily small attractive potentials produce a bound state, see Problem 2 on page 156 in [25], or [10, 60] and also [38]. More precisely, it was rigorously shownFootnote 2 by Barry Simon many years ago, see [50], that it suffices that V is not identically zero and \(V \in L^1\) with \(\int V\, \mathrm {d}x\le 0\), together with some mild moment condition on V, so that \(-\Delta +\lambda V\) has a strictly negative bound state in one and two dimensions, for any \(\lambda >0\). This had been generalized in [37, 59] to higher order Schrödinger-type operators.

Recently renewed interest in weakly coupled bound states arose due to the observation that such states can be found in many other physically interesting cases and they are responsible for different physical behavior of these systems compared to what one is used to from usual quantum mechanics in high dimensions. These systems include quantum wave guides, systems with homogenous or increasing magnetic fields, the Bardeen–Cooper–Schrieffer (BCS) model for superconductivity. These examples are not necessarily one or two-dimensional, but they are described by Schrödinger type operators with strongly degenerate kinetic energies [16, 18, 26, 38], that is, the kinetic energy T can be degenerate, \(T(\eta )=0\), not only at a single point but on a “large” set in momentum space. For example, the kinetic energy could have a zero set which is an embedded hypersurface in \({\mathbb {R}}^d\). At this point it is important to emphasize that the results of [16] concern the special BCS Hamiltonian in three dimensions where, in particular, the zero set of the kinetic energy is a two-dimensional sphere in \({\mathbb {R}}^3\). The works [18, 26, 38] require that the kinetic energy T is locally bounded, satisfying some growth conditions at infinity, the zero set of the kinetic energy T is a smooth co-dimension one submanifold of \({\mathbb {R}}^d\), in [38] only locally, and that \(\int V\, {\mathrm{d}}x <0\) or, even stronger, \(V\le 0\) and \(V\ne 0\). These last two conditions are stronger than the conditions on the potential in the original work of Simon. In particular, they leave open the question what happens if \(V\ne 0\) and \(\int V\, {\mathrm{d}}x =0\) or if the co-dimension of the zero set of the kinetic energy T is larger than one.

Motivated by the above questions, we consider a very general class of kinetic energies and potentials. More importantly, we have a sharp existence result for weakly coupled bound states: We give conditions under which weakly coupled bound states exist for any non-trivial attractive potential, \(\int V\, {\mathrm{d}}x\le 0\), and if our conditions are not met, then for any strictly negative but sufficiently small potential weakly coupled bound states exist. Moreover, in the second case, we prove a quantitative bound on the number of negative bound states.

We are able to do this by identifying the mechanism which is responsible for the creation of weakly coupled bound states: Roughly speaking \(\eta \mapsto T(\eta )^{-1}\) being integrable or not in a small neighborhood of the zero set of T distinguishes between having a quantitative bound on the number of negative bound states in the first case or having weakly coupled bound states for arbitrarily small attractive potentials in the second, see Theorems 1.1 and 1.3, for the precise conditions.Footnote 3

In the following we will always assume, without further mentioning it, that the potential V is relatively form small with respect to the kinetic energy T(p). That is, there exists \(0<a<1\) and \(b>0\) such that

$$\begin{aligned} |\langle \varphi , V\varphi \rangle |\le \Vert \sqrt{T(p)} \varphi \Vert ^2+ b\Vert \varphi \Vert ^2 \end{aligned}$$

for all \(\varphi \in \mathcal {D}(\sqrt{T(p)})= \mathcal {Q}(T(p))\), the form domain of T(p). Here we identify, for simplicity, \(\langle \varphi , V\varphi \rangle = \langle \sqrt{|V|}\varphi , {\mathrm {sgn}(V)}\sqrt{|V|}\varphi \rangle \) for the quadratic form of the potential V and also assume that its quadratic form domain \(\mathcal {Q}(V)\) contains \(\mathcal {Q}(T(p))\). In this case, the famous KLMN theorem guarantees that one can define the generalized Schrödinger operator as a sum of the quadratic forms corresponding to T(p) and V, see [40, 56]. By a slight abuse of notation, we denote this operator by \(T(p)+V\).

Our first theorem concerns the existence of bound states. We write \(V\ne 0\) if V is not the zero function, more precisely, if \(V\ne 0\) on a set of positive Lebesgue measure.

Theorem 1.1

(Weakly coupled bound states) Let \(T:{\mathbb {R}}^d\rightarrow [0,\infty )\) be measurable. Assume that there exists a compact set \(M\subset {\mathbb {R}}^d\) such that

$$\begin{aligned} \int _{M_\delta } T(\eta )^{-1}\, \mathrm {d}\eta = \infty \text { for all } \delta >0 , \end{aligned}$$
(1.1)

where \(M_\delta :=\{\eta \in {\mathbb {R}}^d:\, {\text {dist}}(\eta ,M)\le \delta \}\). Then \(\inf \sigma (T(p))=0\) and if the potential \(V \ne 0\), obeys the basic assumptions from Sect. 2 and \(V\le 0\), then \(T(p)+V\), defined in the quadratic form sense, has at least one strictly negative bound state.

Moreover, if T is locally bounded this also holds for sign indefinite potentials in the sense that if \(V\in L^1({\mathbb {R}}^d)\), \(V\ne 0\), obeys the basic assumptions from Sect. 2 and \(\int V\,\mathrm {d}x\le 0\), then the operator \(T(p)+V\) has again at least one strictly negative eigenvalue.

Remarks 1.2

  1. (i)

    Our theorem poses rather weak conditions on the zero set of the kinetic energy symbol T. Moreover, T does not have to satisfy any growth conditions at infinity. Of course, if T does not satisfy a growth condition at infinity, then, even if \(V\in L^1\), it does not have to be relatively form bounded with respect to T(p). Assuming this and \(\sigma _{\mathrm{ess}}(T(p)+V)=\sigma (T(p))\), one can formulate a version of Theorem 1.1 which still guarantees the existence of some negative spectrum, but not necessarily discrete, we leave the details to the interested reader.

  2. (ii)

    If the potential is negative, \(V\le 0\), and the zero set of the kinetic energy is somewhat ‘thick’, there can be even be infinitely many bound states below the essential spectrum, see Theorem 3.4 and Corollaries 3.6 and 3.7 below. Moreover, if the kinetic energy symbol T is continuous, then the compact set M above can be chosen to be a subset of the zero set of T. In this case, the behaviour of T near its zero set determines whether (1.1) holds or not.

  3. (iii)

    We would like to emphasize that our result also holds when \(\int V\, {\mathrm{d}}x =0\). This makes our criterion applicable in several cases, when previous results [16, 18, 26, 38] fall short. One example is when the Fourier transform \(\widehat{V}\) of the potential is zero on a large ball centered at the origin.

  4. (iv)

    Our method in the proof of Theorem 1.1 relies on a very simple and natural variational calculation, which also works in the critical case where \(\int V\, \mathrm {d}x=0\). It does not require a detailed analysis of the generalized Schrödinger operator \(T(p)+V\). Its main advantage is its simplicity and its wide range of applications.

The situation discussed in Theorem 1.1 changes drastically when \(\eta \mapsto T(\eta )^{-1}\) is integrable near the zero set of T, more precisely, when \(T^{-1}\mathbf {1}_{\{T<\delta \}}\in L^1({\mathbb {R}}^d)\) for some \(\delta >0\). Not only do weakly coupled bound states cease to exist but we have even a quantitative bound on the number of negative eigenvalues in this case. More precisely, introduce the function \(G:[0,\infty ]\rightarrow [0,\infty ]\) by

$$\begin{aligned} G(u):=u\int _{T(\eta )<u}T(\eta )^{-1}\, \frac{\mathrm {d}\eta }{(2\pi )^d} \end{aligned}$$
(1.2)

for \(u\ge 0\). It is clear from the definition that \(G(u)<\infty \) if and only if \(\int _{T<u} T(\eta )^{-1}\, \mathrm {d}\eta <\infty \) and if \(G(u_0)<\infty \) for some \(u_0>0\), then \(G(u)<\infty \) for all \(0\le u\le u_0\) and in this case \(\lim _{u\rightarrow 0} G(u)=0\). The function G is the central object in our quantitative bound on the number of bound states for the Schrödinger-type operator \(T(p)+V\).

Theorem 1.3

(Quantitative bound) Let \(T:{\mathbb {R}}^d\rightarrow [0,\infty )\) be measurable and assume that the potential V obeys the basic assumptions from Sect. 2. Then for \(T(p)+V\), defined in the quadratic form sense, we have the bound

$$\begin{aligned} N(T(p)+V)\le \frac{\alpha ^2}{(1-2\alpha )^2}\int G(V_-(x)/\alpha ^2)\, \mathrm {d}x , \end{aligned}$$
(1.3)

for all \(0<\alpha <\tfrac{1}{2}\). Here \(V_-=\max (0,-V)\) is the negative part of V and \(N(T(p)+V)\) is the number of eigenvalues of \(T(p)+V\) which are strictly negative.

Moreover, if for a given potential V there exists a perturbation \(W>0\) with \(G(V_-+W)<\infty \), then the above bound also includes zero-energy eigenvalues, that is,

$$\begin{aligned} \begin{aligned} N_0(T(p)+V)&:=\#\{\text {eigenvalues of } T(P)+V\le 0\}\\&\le \frac{\alpha ^2}{(1-2\alpha )^2}\int G(V_-(x)/\alpha ^2)\, \mathrm {d}x , \end{aligned} \end{aligned}$$
(1.4)

Remarks 1.4

  1. (i)

    As we will see in Section A, in many practical cases, even when the kinetic energy symbol T is not homogeneous, the function G from Theorem 1.3 can be straightforwardly evaluated and in most cases the result of this evaluation agrees with the precise semi-classical guess up to a small factor. In particular, when \(T(\eta )=\eta ^2\), we recover Cwikel’s version of the CLR bound.

  2. (ii)

    A straightforward argument shows that if for all \(u>0\) the sublevel sets \(\{T<u\}\) have finite Lebesgue measure, then

    $$\begin{aligned}&\int _{T< u}T(\eta )^{-1}\, \mathrm {d}\eta<\infty \text { for some } u>0\\&\qquad \Longleftrightarrow \int _{T< u }T(\eta )^{-1}\, \mathrm {d}\eta <\infty \text { for all } u>0 . \end{aligned}$$

    In this case, \(G(u)<\infty \) for all \(u\ge 0\) is equivalent to \(G(u_0)<\infty \) for some \(u_0>0\). Moreover, in the case that the sublevel sets \(\{T<u\}\) have finite Lebesgue measure, Lemma B.2 in the appendix yields a non-trivial relative form compactness criterium of a potential V which does not require that T diverges to infinity at infinity.

    Of course, in all the applications we know of one usually has \(\lim _{\eta \rightarrow \infty }T(\eta )=\infty \) or, in the case of discrete Schrödinger operators, the range of possible momenta \(\eta \) is a bounded subset of \({\mathbb {R}}^d\). Thus in these applications one always has \(G(u)<\infty \) for all \(u>0\) once \(G(u_0)<\infty \) for some \(u_0>0\).

  3. (iii)

    The function G above has a nice semi-classical interpretation. We note

    $$\begin{aligned} G(u)&= \int \limits _{T/u<1} (T(\eta )/u)^{-1}\, \frac{\mathrm {d}\eta }{(2\pi )^d} = \int \limits _{T/u<1} \frac{\mathrm {d}\eta }{(2\pi )^d} \int _0^\infty s^{-2}\mathbf {1}_{\{ T/u<s\}}\,\mathrm {d}s \\&= \int _0^\infty \int _{{\mathbb {R}}^d} \mathbf {1}_{\{ T(\eta ) -\min (1,s)u<0\}}\,\frac{\mathrm {d}\eta }{(2\pi )^d} \, \frac{\mathrm {d}s}{s^2} \\&= \int _{{\mathbb {R}}^d} \mathbf {1}_{\{ T(\eta ) -u<0\}}\,\frac{\mathrm {d}\eta }{(2\pi )^d} + \int _0^1 \int _{{\mathbb {R}}^d} \mathbf {1}_{\{ T(\eta ) -\min (1,s)u<0\}}\,\frac{\mathrm {d}\eta }{(2\pi )^d} \, \frac{\mathrm {d}s}{s^2} . \end{aligned}$$

    Thus with the classical phase-space volume, given by

    $$\begin{aligned} N^{\mathrm {cl}}(T+V)&:=\iint \limits _{{\mathbb {R}}^d\times {\mathbb {R}}^d} \mathbf {1}_{\{ T(\eta )+V(x)<0\}}\, \frac{\mathrm {d}\eta \mathrm {d}x}{(2\pi )^d}\\ {}&= \int _{{\mathbb {R}}^d} \left| \left\{ \eta \in {\mathbb {R}}^d:\, T(\eta ) +V(x)<0 \right\} \right| \,\frac{\mathrm {d}x}{(2\pi )^d} , \end{aligned}$$

    a reformulation of the bound in (1.3) is

    $$\begin{aligned} \begin{aligned} N(T(p)+V)&\le \frac{\alpha ^2}{(1-2\alpha )^2} N^{\mathrm {cl}}(T+V/\alpha ^2) \\&\quad + \frac{\alpha ^2}{(1-2\alpha )^2}\int _0^1 N^{\mathrm {cl}}\left( T +sV(x)/\alpha ^2 \right) \, \frac{\mathrm {d}s}{s^2} \end{aligned} \end{aligned}$$
    (1.5)

    for all \(0<\alpha <1/2\). The first part on the right hand side above is clearly related to the classical phase space volume guess suggested by the uncertainty principle and the second part can be considered as a quantum correction. We would like to emphasize that for the usual Schrödinger operator one has

    $$\begin{aligned} N^{\mathrm {cl}}(|\eta |^2+V/\alpha ^2) = \frac{|B_1^d|}{(2\pi )^d} \int _{{\mathbb {R}}^d} V_-(x)^{d/2}\, {\mathrm{d}}x \end{aligned}$$
    (1.6)

    which for a large class of potentials, e.g., \(V\in \mathcal {C}^\infty _0({\mathbb {R}}^d)\) is finite in all dimensions. On the other hand, it is well known that in dimensions one and two the Cwikel–Lieb–Rozenblum bound does not hold. Thus even in this well understood case, finiteness of the classical phase space volume does not imply the existence of a quantitative bound of CLR-type. Our quantum correction, the second term in (1.5) takes care of this. Theorem 1.3 shows that a simple general quantitative bound on the number of bound states holds, exactly when the contribution from the quantum correction, i.e., the second term in (1.5), is finite.

To show that under some slightly stronger conditions on T,  Theorems 1.1 and 1.3 are complementary, we provide

Theorem 1.5

Let \(Z:=\{T=0\}\) be the zero set of the kinetic energy symbol \(T:{\mathbb {R}}^d\rightarrow [0,\infty )\). Assume that Z is compact and that T is small only close to its zero set, more precisely, we assume that for all \(\delta >0\) there exists \(u>0\) with

$$\begin{aligned} \int _{\{T<u\}\cap Z_\delta ^c} \frac{1}{T(\eta )}\, \mathrm {d}\eta <\infty , \end{aligned}$$
(1.7)

where \(Z_\delta ^c\) is the complement of \(Z_\delta := \{\eta \in {\mathbb {R}}^d:\, {\text{ dist }}(\eta ,Z)\le \delta \}\). Then

  1. (a)
    $$\begin{aligned} \int _{Z_\delta } \frac{1}{T(\eta )}\,\mathrm {d}\eta =\infty \quad \text {for some } \delta >0 \end{aligned}$$
    (1.8)

    is equivalent to \(T(p)+V\) having weakly coupled bound states for any non-trivial attractive potential V. That is, for any \(V\le 0\), \(V\ne 0\), which obeys the basic assumptions from Sect. 2 the operator \(T(p)+V\) has at least one strictly negative eigenvalue.

    Moreover, if in addition the kinetic energy symbol T is locally bounded, then (1.8) is also equivalent to \(T(p)+V\) having a strictly negative eigenvalue for non-trivial sign changing potentials V in the sense that if \(V\in L^1({\mathbb {R}}^d)\), \(V\ne 0\), obeys the basic assumptions from Sect. 2 and \(\int V\,\mathrm {d}x\le 0\), then the operator \(T(p)+V\) has again at least one strictly negative eigenvalue.

  2. (b)
    $$\begin{aligned} \int _{Z_\delta } \frac{1}{T(\eta )}\, \mathrm {d}\eta <\infty \quad \text {for some } \delta >0 \end{aligned}$$
    (1.9)

    is equivalent to the existence of a quantitative bound on the number of bound states in the following sense: There exists a function \(G_0:[0,\infty ]\rightarrow [u,\infty ]\) with \(G_0(u)<\infty \) for all small enough \(u>0\) and \(\lim _{u\rightarrow 0+}G_0(u)=0\) such that for any potential V which is relatively form compact with respect to T(p) one has the bound

    $$\begin{aligned} N(T(p)+V)\le \int G_0(V_-(x))\, \mathrm {d}x , \end{aligned}$$

    where \(V_-=\max (0,-V)\) is the negative part of V and \(N(T(p)+V)\) is the number eigenvalues of \(T(p)+V\) which are strictly negative. Moreover, in this case one can take \(G_0(u)= \frac{\alpha ^2}{(1-2\alpha )^2}G(\alpha ^{-2}u)\) with G defined in (1.2) and \(0<\alpha <1/2\).

Remarks 1.6

  1. (i)

    To see that (1.7) is, indeed, a rather weak growth and regularity condition on T we note that (1.7) is fulfilled under the following two conditions on T:

    $$\begin{aligned} 1)\quad \text{ There } \text{ exists } 0<\varepsilon ,R<\infty \text{ with } T(\eta )\ge \varepsilon \text{ for } \text{ all } |\eta |> R , \end{aligned}$$
    (1.10)

    which is, for example, the case if \(T(\eta )\rightarrow \infty \) for

    $$\begin{aligned} 2) \quad T \text{ is } \text{ lower } \text{ semi-continuous }. \end{aligned}$$
    (1.11)

    Indeed, under the above two conditions one has

    $$\begin{aligned} \text {for any } \delta>0 \text { there exists }u>0 \text { with }\{T<u\}\cap Z_\delta ^c = \emptyset , \end{aligned}$$
    (1.12)

    which clearly implies (1.7). To see (1.12), define \(r(u):=\sup \{{\text {dist}}(\eta ,Z):\, T(\eta )<u\}\). Then \(\{T<u\}\subset Z_\delta \) is equivalent to \(r(u)\le \delta \), so it is enough to show \(\lim _{u\rightarrow 0+} r(u)=0\). Clearly, \(0<u\mapsto r(u)\) is increasing. Assume that there exists \(\varepsilon _0>0\) with \(r(u)\ge 2\varepsilon _0\) for all \(u>0\). Taking \(u=1/n\), this yields the existence of a sequence \(\xi _n\in {\mathbb {R}}^d\) with \({\text {dist}}(\xi _n,Z)>\varepsilon _0\) and \(T(\xi _n)<1/n\). Because of (1.10) we have \(\xi _n\le R\) for all large enough n. Thus we can take a subsequence \(\xi _{n_l}\) such that \(\eta =\lim _{l\rightarrow \infty }\xi _{n_l}\) exists. Because of the lower semicontinuity of T one has \(0\le T(\eta )\le \liminf _{l\rightarrow \infty } T(\xi _{n_l})=0\), so \(\eta \in Z\), which contradicts \({\text {dist}}(\eta ,Z) = \lim _{l\rightarrow \infty }{\text {dist}}(\xi _{n_l},Z)\ge \varepsilon _0>0\). So \(r(u)\rightarrow 0\) as \(u\rightarrow 0\), which proves (1.12).

  2. (ii)

    The existence of a quantitative bound on the number of strictly negative eigenvalues of \(T(p)+V\) clearly implies that weakly coupled bound states do not exist, see Remark 1.4.i. On the other hand, we do not know of any result in the literature which shows, without assuming strong additional conditions on the kinetic energy T(P), that the absence of weakly coupled bound states implies the existence of a quantitative semi-classical type bound on the number of strictly negative bound states of Schrödinger-type operators \(T(P)+V\). These strong additional assumptions refer to Markov properties [27], positivity-preservation of the associated semigroup [12, 30, 43] or the inverse of the kinetic energy being in a weak \(L^p\)-space, thus imposing a kind of homogeneity condition, [14, 15], see also [46, 47]. The main point of our Theorem 1.5 is that under very weak regularity assumptions on T near its zero set, which are fulfilled in all physically relevant cases, the two phenomena of weakly coupled bound states and the existence of quantitative semi-classical type bound on the number of strictly negative bound states are indeed complementary.

  3. (iii)

    Condition (1.7) ensures that we have the equivalences

    $$\begin{aligned} \int _{Z_\delta } \frac{1}{T(\eta )}\,\mathrm {d}\eta= & {} \infty \text { for all } \delta>0\nonumber \\&\Longleftrightarrow \int _{Z_\delta } \frac{1}{T(\eta )}\, \mathrm {d}\eta =\infty \text { for some } \delta>0 \nonumber \\&\Longleftrightarrow \int _{T<u} \frac{1}{T(\eta )}\, \mathrm {d}\eta =\infty \text { for all } u>0 . \end{aligned}$$
    (1.13)

    This clearly ensures that conditions (1.8) and (1.9) are complementary. The equivalence (1.13) follows from Lemma 6.4 in Sect. 6.2.

To put our work into perspective: Previously, the weakest condition on the potential which guaranteed existence of at least one strictly negative eigenvalue is due to Pankrashkin [38] who showed that a strictly negative eigenvalue exists if \(V\in L^1 \) and \(\int V\, {\mathrm{d}}x <0\). This condition is weaker than the condition of [16, 18] where the authors have to assume that the Fourier transform \(\widehat{V}\) is non-vanishing within a large enough ball centered at the origin.Footnote 4 In [16, 18, 26], they adapt the method of Simon, using the Birman–Schwinger principle [5, 6, 55], to identify a singular piece of the Birman–Schwinger operator. This approach needs global assumptions on the zero set of the kinetic energy, see also Remark 1.2.iii . The work [38] uses a construction of appropriate trial functions, as such the assumptions on the zero set of the kinetic energy in [38] are local. More precisely, there is an open set such that locally within this set the zero set of T is a smooth submanifold of \({\mathbb {R}}^d\). The work [26] establishes the precise asymptotic rate of the negative eigenvalues, but for this they need strong assumptions.

It is easy to construct examples of potentials \(V\in L^1\) such that its Fourier transform \(\widehat{V}\) is zero on a large centered ball. Simply take any spherically symmetric Schwartz function \(\widehat{V}\) which is supported on a large enough centered annulus in Fourier space and let V be the inverse Fourier transform of \(\widehat{V}\). In this case \(\int V\, {\mathrm{d}}x =\widehat{V}(0)=0\) and our Theorem 1.1 shows that there exists at least one strictly negative eigenvalue under suitable conditions on the kinetic energy T, whereas the previous results in [18, 26, 38] are not applicable. The only exception is the pioneering work of Simon, which for the Laplace operator \(T(p)=p^2\) in one and two dimensions, shows that there are strictly negative eigenvalues if \(\int V\, {\mathrm{d}}x\le 0\), V does not vanish identically, and, for some technical reasons, some high enough moments of V are finite.

Some additional remarks concerning Theorem 1.3 are

Remarks 1.7

  1. (i)

    As already mentioned, from the definition of G together with a simple monotonicity argument it is clear that if for some \(u_0\) one has \(G(u_0)<\infty \) then \(G(u)<\infty \) for all \(0\le u\le u_0\) and \(\lim _{u\rightarrow 0+}G(u)=0\). Thus, if \(G(u_0)<\infty \), or equivalently, \(T\mathbf {1}_{T<u_0}\in L^1({\mathbb {R}}^d)\), for some \(u_0>0\), a simple construction yields a potential \(V<0\) such thatFootnote 5\(\int _{{\mathbb {R}}^d} G(V_-(x)/\alpha ^2)\, \mathrm {d}x<\infty \). But then, replacing V by \(\lambda V\) for \(0<\lambda \le 1\), the monotone convergence theorem gives \(\lim _{\lambda \rightarrow 0+} \int _{{\mathbb {R}}^d} G(\lambda V_-(x)/\alpha ^2)\, \mathrm {d}x =0\) and the bound provided by (1.3) shows

    $$\begin{aligned} N(T(p)+\lambda V)&\le \frac{\alpha ^2}{(1-2\alpha )^2} \int _{{\mathbb {R}}^d} G(\lambda V_-(x)/\alpha ^2)\, \mathrm {d}x<1 \end{aligned}$$

    for all small enough \(\lambda >0\). So in this case there exists a strictly negative potential for which \(T(p)+V\) has no strictly negative eigenvalues. So Theorems 1.1 and 1.3 appear to be complementary. More precisely, as Theorem 1.5 below shows this is, indeed, the case under some slight additional global assumptions on the kinetic energy T, which seem to be fulfilled in all physically relevant applications.

  2. (ii)

    Theorem 1.1 shows that strictly negative eigenvalues of \(T(P)+V\) exist once the integral of \(T(\eta )^{-1}\) diverges in a neighborhood of a compact set. On the other hand, Theorem 1.3 yields a quantitative bound on the number of negative eigenvalues under the condition that \(T^{-1}\mathbf {1}_{\{T<\delta \}}\) is integrable for some \(\delta >0\). Naturally, one can ask the question what could happen if \(T^{-1}\) is integrable over every compact set, but diverges globally. As an example of such a situation, we consider the operator

    $$\begin{aligned} H_\lambda = (p_1^2+p_2^2)^{1/2} - \lambda U(x_1,x_2,x_3) \quad \text {on } L^2({\mathbb {R}}^3) , \end{aligned}$$

    for some \(0\le U\in \mathcal {C}_0^\infty ({\mathbb {R}}^3)\), \(\lambda \ge 0 \). For sufficiently small \(\lambda \) this operator does not have negative spectrum (no weakly coupled bound states). On the other hand, after some critical \(\lambda _{\mathrm {cr}}\) the infimum of the essential spectrum immediately goes down and discrete eigenvalues still do not exist. More precisely, by construction we have

    $$\begin{aligned} \begin{aligned} \sigma (H_\lambda )\cap (-\infty ,0)&=\sigma _{\mathrm{ess}}(H)\cap (-\infty ,0) \\&= \bigcup _{x_3\in {\mathbb {R}}} \big (\sigma ((p_1^2+p_2^2)^{1/2} -\lambda U(\cdot ,x_3))\cap (-\infty ,0) \big ) . \end{aligned} \end{aligned}$$
    (1.14)

    For fixed \(x_3\in {\mathbb {R}}\) consider the operator \((p_1^2+p_2^2)^{1/2} -\lambda U(\cdot ,x_3)\) as an operator on \(L^2({\mathbb {R}}^2)\). Its quadratic form is monotonically decreasing in \(\lambda >0\) and for fixed \( x_{3} \in {\mathbb {R}}\) we can apply Theorem 1.3 to show that for small enough \(\lambda \) the operator \((p_1^2+p_2^2)^{1/2} -\lambda U(\cdot ,x_3)\) is positive. This example shows that for the existence of weakly coupled bound states one needs that the kinetic energy goes to zero fast enough near its zero set.

We would like to stress that in all or our results, the assumptions on the kinetic energy symbol T are very weak and fulfilled in all physically interesting cases. Our theorems have several applications, discussed in Section A, including Schrödinger operators with fractional Laplacians, different types of Schrödinger type operators with degenerate kinetic energies such as pseudo-relativistic Schrödinger operators with positive mass and two-particle pseudo-relativistic Schrödinger operators with different masses, including very different masses, BCS-type operators, and discrete Schrödinger operators.

Our paper is organized as follows: We first address the question of existence of negative bound states. The main idea in the proof of Theorem 1.1 is first shown in a simple model case in Sect. 4. In Sect. 5 we give the proof of Theorem 1.1 and its refinement Theorem 3.4 and their corollaries. In Sect. 6 we give the proof of Theorem 1.3 and in Sect. 6.2 the proof of Theorem 1.5. The applications of Theorems 1.1 and 1.3 are discussed in Appendix A.

2 Basic Assumptions

We consider operators of the form

$$\begin{aligned} H = T(p) + V \end{aligned}$$
(2.1)

where \(p = -i\nabla \) is the quantum-mechanical momentum operator and the symbol of the kinetic energy T is a measurable nonnegative function on \({\mathbb {R}}^d\). We define T(p) as the Fourier multiplier,

$$\begin{aligned} T(p)\varphi :=\mathcal {F}^{-1}[T(\cdot )\widehat{\varphi }(\cdot )] \end{aligned}$$
(2.2)

where we use the convention

$$\begin{aligned} \widehat{f}(\eta ):=\mathcal {F}(f)(\eta )&:=\frac{1}{(2\pi )^{d/2}} \int _{{\mathbb {R}}^d} e^{-i\eta \cdot x} f(x)\, {\mathrm{d}}x \end{aligned}$$

and

for the Fourier transform and its inverse. A-priori the above expressions are only defined when fg are Schwartz class, but they extend to unitary operators to all of \(L^2({\mathbb {R}}^d)\) by density of Schwartz functions in \(L^2({\mathbb {R}})\), see [31, 52].

For a positive self-adjoint operator A we denote by \(\mathcal {Q}(A)\) its form domain and by \(\mathcal {D}(A)\) its domain. Thus \(\mathcal {Q}(A)= \mathcal {D}(\sqrt{A})\). In particular,

$$\begin{aligned} \mathcal {Q}(T(p)) = \left\{ f\in L^2({\mathbb {R}}^d):\, \int _{{\mathbb {R}}^d}T(\eta )|\widehat{f}(\eta )|^2\, \mathrm{d}\eta <\infty \right\} \end{aligned}$$
(2.3)

The potential \(V:{\mathbb {R}}^d\rightarrow {\mathbb {R}}\) is a Borel-measurable function. In order to define the Schrödinger type operator \(T(P)+V\) as a sum of quadratic forms, it is enough to assume, for simplicity, that its modulus |V| is form small with respect to T(p), that is, for some \(0<a<1\) and \(b>0\) we have

$$\begin{aligned} \langle \varphi , |V|\varphi \rangle \le \Vert \sqrt{T(p)}\varphi \Vert ^2+ b\Vert \varphi \Vert ^2 \end{aligned}$$
(2.4)

for all \(\varphi \in \mathcal {D}(\sqrt{T(p)})= \mathcal {Q}(T(p))\), the form domain of T(p). In this case the quadratic form domain \(\mathcal {Q}(V)\) of V is given by the domain of the multiplication operator \(|V|^{1/2}\), \(\mathcal {Q}(V)=\mathcal {D}(|V|^{1/2})\), and we identify, for simplicity, \(\langle \varphi , V\varphi \rangle = \langle \sqrt{|V|}\varphi , {\mathrm {sgn}(V)}\sqrt{|V|}\varphi \rangle \) for the quadratic form of the potential V. We also assume that \(\mathcal {Q}(T(p))\subset \mathcal {Q}(V)\).

More generally, let \(V_+= \max (V,0)\), respectively \(V_-=\max (-V,0)\), be the positive, respectively negative, part of the potential V and assume that \(\mathcal {Q}(T(p)\cap \mathcal {Q}(V_+)\) is dense in \(L^2({\mathbb {R}}^d)\) and that \(V_-\) is relatively form small with respect to \(T(p)+V_+\). That is, \(\mathcal {Q}(T(p)\cap \mathcal {Q}(V_+)\subset \mathcal {Q}(V_-)\) and there exists \(0\le \alpha <1\) and \(0\le \beta <\infty \) such that

$$\begin{aligned} \Vert \sqrt{V_-}f\Vert ^2 \le \alpha \Big ( \Vert \sqrt{T(p)}f\Vert ^2+ \Vert \sqrt{V_+}f\Vert ^2 \Big ) +\beta \Vert f\Vert ^2 \end{aligned}$$
(2.5)

Under either conditions (2.4) or (2.5), the famous KLMN theorem, see, e.g., [56, Theorem 6.24] or [53, Theorem 7.5.7], shows that the natural quadratic form corresponding to \(T(p)+V\) is closed on \(\mathcal {Q}(T(p))\cap \mathcal {Q}(V_+)\) and defines a lower bounded self-adjoint operator, which we will denote by \(T(p)+V\), for simplicity.

Since we have the form smallness condition on |V|, or we split V into its positive and negative parts, we do not discuss highly singular oscillating potentials in this work.

Since T(p) is a Fourier multiplier, i.e., multiplication by a function on the Fourier side, it has purely essential spectrum, \(\sigma (T(p))=\sigma _{\mathrm{ess}}(T(p))=\mathrm {essrange}(T) \subset [0,\infty )\). Here the essential range of T is given by

$$\begin{aligned} \mathrm {essrange}(T)= \{E\in {\mathbb {R}}:\, |T^{-1}((E-\delta ,E+\delta ))|>0 \text { for all } \delta >0\} . \end{aligned}$$

We also need a condition on the essential spectrum of \( T(p)+V\). We will always assume that

$$\begin{aligned} \sigma _{\mathrm{ess}} (T(p)+V)\subset \sigma _{\mathrm{ess}}(T(p)) . \end{aligned}$$
(2.6)

For sufficient conditions, which imply (2.6) and (2.5), respectively (2.4), see Remark 2.1.

For the quantitative bound on the number of bund states we need a slightly stronger assumption, where in (2.5) we replace \(V_+\) by zero and in (2.6) we replace V by \(-V_-\).

Remarks 2.1

  1. (i)

    The conditions (2.5) and (2.6) are, in particular, fulfilled when |V| is relatively form compact with respect to T(p), i.,e, \( (T(p)+1)^{-1/2} |V| (T(p)+1)^{-1/2}\) is a compact operator on \(L^2({\mathbb {R}}^d)\). This is well-known, see [56, Section 6.3] or [53, Section 7.5] for example. In this case, V is automatically relatively form small with respect to T(p) with relative bound zero, [56, Lemma 6.26]. Moreover, by the relative form compactness \(\sigma _{\mathrm{ess}}(T(p)+V)=\sigma _{\mathrm{ess}}(T(p))= \sigma (T(p))\), [56, Lemma 6.26] or [53, Theorem 7.8.4],

  2. (ii)

    Necessary and sufficient conditions are given in [35] for a potential V to be relatively form bounded, respectively relative form compact, with respect to the usual kinetic energy \( T(p)=p^2\). These conditions cover, in particular, examples where the potential is highly oscillatory. Since a characterization of relatively form bound and form compact potentials V in the spirit of [35] is not known for the general class of kinetic energies we are interested in, we provide in Appendix B two sufficient conditions for the invariance of the essential spectrum of T(p) under perturbation by a potential V which is relative form bounded with respect to T(p). These condition are far from optimal but are easy to apply in a wide variety of cases. For example, assume that V is relative form small with respect to T(p). Then one has \(\sigma _{\text {{ess}}}(T(p)+V) = \sigma _{\text {{ess}}}(T(p))\) as soon as \(|V|^{1/2}(T(p)+1)^{-1}\) is a compact operator or if \(V\in L^1({\mathbb {R}}^d)\), assuming in addition that \(\lim _{\eta \rightarrow \infty }T(\eta )=\infty \), see Lemma B.1 in the appendix. A different criterium, which still needs that T diverges at infinity, is discussed in Lemma B.2.

3 Existence of Bound States: The General Setup

Physical heuristic suggests that a weak attractive potential can create a bound state if T is small close to its zero set. To make this precise we introduce a local version of this.

Definition 3.1

Let \(T:{\mathbb {R}}^d\rightarrow [0,\infty )\) be measurable. T has a thick zero set near \(\omega \in Z\) if

$$\begin{aligned} \int _{B_\delta (\omega )} T(\eta )^{-1}\, \mathrm{d}\eta = \infty \text { for all } \delta >0, \end{aligned}$$
(3.1)

where \(B_\delta (\omega )= \{\eta \in {\mathbb {R}}^d:\, {\text {dist}}(\eta ,\omega )<\delta \}\) is the open ball of radius \(\delta \) centered at \(\omega \).

In the following, we will assume, without mentioning all the time, that the assumptions of Theorem 1.1 hold and \(T(p)+V\) is defined in the quadratic form sense. A local version of Theorem 1.1 is given by

Theorem 3.2

Suppose that \(V\le W\ne 0\) obey the basic assumptions from Sect. 2, \(W\in L^1\), \(\int W\, {\mathrm{d}}x\le 0\), and T has a thick zero set near some point \(\omega \in Z\). Then \(T(p)+V\) has at least one strictly negative eigenvalue.

Remark 3.3

The sole role of the comparison potential W is to be able to easily include potentials V obey the basic assumptions from Sect. 2, but are not integrable. For example, if the positive part \(V_+\in L^1\) and \(V_-\not \in L^1\), and V obeys the basic assumptions from Sect. 2, we can choose

$$\begin{aligned} W^m_{R} = V_+ -\min (V_-, m)\mathbf {1}_{B_R} \end{aligned}$$

where \(B_R\) is a centered ball of radius \(R>0\). Then \(V\le W^m_R\in L^1\), \(W^m_R\) obeys the basic assumptions from Sect. 2, and \(\int W^m_R\, {\mathrm{d}}x <0\) for mR large enough. Thus Theorem 3.2 yields the existence of a negative eigenvalue of \(T(p)+V\) in this case.

In particular, the existence of a strictly negative eigenvalue of \(T(p) +V\) for potentials \(V\ne 0\), which obey the basic assumptions from Sect. 2 and which are sign definite, that is, \(V\le 0\), follows at once from Theorem 3.2. We do not, however, consider sign changing potentials which are not in \(L^1({\mathbb {R}}^d)\).

A refinement of Theorem 3.2, when the zero set of the kinetic energy T has many disjoint thick parts is given by the next theorem, which also yields an easy criterion for the infinitude of weakly coupled bound states.

Theorem 3.4

Assume that \(V, W\ne 0\) obey the basic assumptions from Sect. 2 and that for some \(k\in {\mathbb {N}}\) the kinetic energy T has a thick zero set near k pairwise distinct points \(\omega _1,\ldots ,\omega _k\in Z\).

  1. (a)

    If \(V\le 0\), then \(T(p)+V\) has at least k strictly negative eigenvalues.

  2. (b)

    If \(V\le W\), \(W\in L^1\), and the \(k\times k\) matrix \(M=(\widehat{W}(\omega _l-\omega _m))_{l,m=1,\ldots ,k }\), where \(\widehat{W}\) is the Fourier transform of W, is strictly negative definite, then \(T(p)+V\) has at least k strictly negative eigenvalues.

  3. (c)

    If \(V\le W\), \(W\in L^1\), the \(k\times k\) matrix \(M=(\widehat{W}(\omega _l-\omega _m))_{l,m=1,\ldots ,k }\), is negative semi-bounded, the eigenvalue 0 of this matrix is non-degenerate, and the function T is locally bounded, then \(T(p)+V\) has at least k strictly negative eigenvalues.

Remarks 3.5

  1. (i)

    In the spirit of Theorem 3.4, one can formulate a condition under which the operator \(T(p)+V\) has at least k eigenvalues when \(V\le W\), for a semi-bounded matrix \(M=(\widehat{W}(\omega _l-\omega _m))_{l,m=1,\ldots ,k }\) with a degenerate eigenvalue zero. We are not doing this for the sake of simplicity, but leave it to the interested reader.

  2. (ii)

    Similar to part (a) of Theorem 3.4, in [9, 38] the authors also had the condition that the matrix \(M=(\widehat{V}(\omega _l-\omega _m))_{l,m=1,\ldots ,k }\) is negative definite, but the conditions on the zero set of the kinetic energy are much stronger than ours. Moreover, we can also handle the case of a non-degenerate zero eigenvalue of M. In the following we use for two real-valued functions fg the notation \(f\lesssim g\) if there exists a constant \(C>0 \) such that \(f\le Cg\). We also write \(f\sim g\) if \(f\lesssim g\) and \(g\lesssim f\).

Useful corollaries of Theorems 3.2 and 3.4 are

Corollary 3.6

Assume that \(V\le W\ne 0\) obey the basic assumptions from Sect. 2, that there are k isolated points \(\omega _1,\ldots ,\omega _k\) and that near \(F=\{\omega _1,\ldots ,\omega _k\}\), i.e., in an open neighborhood \(\mathcal {O}\) containing F, the kinetic energy symbol obeys the bound

$$\begin{aligned} T(\eta )\lesssim {\text {dist}}(\eta ,F)^\gamma \quad \text {for all } \eta \in \mathcal {O}\text { and some } \gamma \ge d . \end{aligned}$$
(3.2)
  1. (a)

    If \(V\le 0\) or if \(W\in L^1({\mathbb {R}}^d)\) and the matrix \(M= (\widehat{W}(\omega _l-\omega _m))_{l,m=1,\ldots ,k }\) is negative definite, then \(T(p)+V\) has at least k strictly negative eigenvalues.

  2. (b)

    If \(W\in L^1({\mathbb {R}}^d)\) and \(\int W\, {\mathrm{d}}x\le 0\), then \(T(p)+V\) has at least one strictly negative eigenvalue.

Corollary 3.7

Assume that \(V\le W\ne 0\) obey the basic assumptions from Sect. 2and that there is a \(\mathcal {C}^2\) submanifold \(\Sigma \) of codimension \(1\le m\le d-1\) such that near \(\Sigma \), i.e., in an open neighborhood \(\mathcal {O}\) containing \(\Sigma \), the kinetic energy symbol obeys the bound

$$\begin{aligned} T(\eta )\lesssim {\text {dist}}(\eta ,\Sigma )^\gamma \quad \text {for all } \eta \in \mathcal {O}\text { and some } \gamma \ge m . \end{aligned}$$
(3.3)
  1. (a)

    If \(V\le 0\) then \(T(p)+V\) has infinitely many strictly negative eigenvalues.

  2. (b)

    If \(V\le W\), \(W\in L^1({\mathbb {R}}^d)\), and \(\int W\, {\mathrm{d}}x\le 0\), then \(T(p)+V\) has at least one strictly negative eigenvalue.

Remark 3.8

In most applications T is continuous and the zero set of T is either a point, a collection of points, or a smooth submanifold in \({\mathbb {R}}^d\). So Corollaries 3.6 and 3.7 would be enough to cover all applications we can think of. However, we find that the proof for the general case is so simple, that adding further structure to its assumptions only obscures the simplicity of the proof. So we prefer to state Theorem 1.1 and its local versions, Theorems 3.2 and 3.4, in their generality.

Some of the most interesting applications are considered in Appendix A.

4 Existence of Bound States: A Simple Model Case

We want to construct a test function \(\varphi \) such that \(\langle \varphi , (T(p)+V) \varphi \rangle <0 \). Once one has such a state together with \(\sigma _{\mathrm{ess}} (T(p)+V)\subset [0,\infty )\), the Rayleigh–Ritz variational principle shows that strictly negative discrete spectrum exists. Of course, the catch is how to guess such a variational state \(\varphi \) in a systematic way.

To motivate our construction for our general set-up, we will discuss here the simple model case, where \(T(\eta )= |\eta |^\gamma \), i.e., \(T(p)= (-\Delta )^{\gamma /2}\) is a fractional Laplacian.

4.1 The Case \(\int V\,\mathrm {d}x<0\): Learning from Failure

We will work mainly in Fourier-space and, for simplicity, consider \(\int V\, {\mathrm{d}}x <0\) first. In order to make the kinetic energy small, a natural first guess for the Fourier transform of the test function would be

$$\begin{aligned} \widetilde{w}_\delta :=\mathbf {1}_{A_\delta } , \end{aligned}$$

for a suitably chosen set \(A_\delta \) of finite positive measure which concentrates around zero, since this makes the kinetic energy small. However, it turns out that this is not a good ansatz and it is instructive to see why. In order to use the assumption \(\int V\, {\mathrm{d}}x <0\), we want our test function to converge to a constant, so its Fourier transform should converge to a delta-function at zero. Thus we need to normalize \(\widetilde{w}_\delta \) and are led to consider

$$\begin{aligned} \widetilde{\varphi }_\delta :=\frac{\widetilde{w}_\delta }{\Vert \widetilde{w}_\delta \Vert _{L^1_\eta }} . \end{aligned}$$

Note that this choice fulfills two crucial assumptions: Let

$$\begin{aligned} \kappa _\delta (x):=\mathcal {F}^{-1}(\widetilde{\varphi }_\delta )(x) =\frac{1}{(2\pi )^{d/2}}\int _{{\mathbb {R}}^d} e^{i\eta \cdot x} \widetilde{\varphi }_\delta (\eta )\, \mathrm{d}\eta \end{aligned}$$

be the inverse Fourier transform of \(\widetilde{\varphi }_\delta \). We always have

$$\begin{aligned} |\kappa _\delta (x)|\le \frac{\Vert \widetilde{\varphi }_\delta \Vert _{L^1_\eta }}{(2\pi )^{d/2}} = \frac{1}{(2\pi )^{d/2}} \end{aligned}$$

by our normalization of \(\widetilde{\varphi }_\delta \). Moreover, as long as \(A_\delta \) concentrates to the single point zero in a suitable way, we also have

$$\begin{aligned} \lim _{\delta \rightarrow 0} \kappa _\delta (x) = \frac{1}{(2\pi )^{d/2}} = \frac{1}{(2\pi )^{d/2}} \end{aligned}$$

for all \(x\in {\mathbb {R}}^d\). Since, by assumption the potential V is integrable, we conclude with Lebesgue’s dominated convergence theorem

$$\begin{aligned} \lim _{\delta \rightarrow 0}\langle \kappa _\delta , V\kappa _\delta \rangle = \frac{1}{(2\pi )^d} \int V\, {\mathrm{d}}x. \end{aligned}$$
(4.1)

So if \( \int V\, {\mathrm{d}}x <0\), the choice for \(\varphi _\delta \) yields a test function which makes the potential contribution strictly negative.

It only remains to see whether the kinetic energy vanishes and, since the set \(A_\delta \) concentrates near zero, this should be the case, but there is a catch: Note that

$$\begin{aligned} \langle \kappa _\delta , (-\Delta )^{\gamma /2}\kappa _\delta \rangle&= \langle \widetilde{\varphi }_\delta , |\eta |^\gamma \widetilde{\varphi }_\delta \rangle = \frac{1}{\Vert \widetilde{w}_\delta \Vert _{L^1_\eta }^2} \int _{A_\delta } |\eta |^\gamma \,\mathrm{d}\eta . \end{aligned}$$

Since \(\Vert \widetilde{w}_\delta \Vert _{L^1_\eta }= |A_\delta |\), the Lebesgue measure of the set \(A_\delta \), we can use rearrangement inequalities, see, e.g., [31], to make the kinetic energy smallest by chosing \(A_\delta \) to be centered ball of radius \(\delta \), say. In this case \(|A_\delta |\sim \delta ^d\) and thus

$$\begin{aligned} \langle \kappa _\delta , (-\Delta )^{\gamma /2}\kappa _\delta \rangle \sim \frac{1}{\delta ^{2d}} \int _{0}^\delta r^{\gamma +d-1}\, \mathrm{d}r \sim \delta ^{\gamma -d} \end{aligned}$$
(4.2)

and this goes to zero as \(\delta \rightarrow 0\) only if \(\gamma >d\) and it misses the critical case where \(\gamma =d\).

So we have to modify the test functions. A better choice, which also works for \(\gamma =d\), turns outFootnote 6 to be given by

$$\begin{aligned} \widehat{w}_\delta (\eta ) :=|\eta |^{-\gamma }\,\mathbf {1}_{A_\delta }(\eta ) \end{aligned}$$
(4.3)

where now the set \(A_\delta \) has to stay away from zero to make \(\widehat{w}_\delta (\eta )\) normalizable. Note that \(|\eta |^{-\gamma }\) is just the inverse of the symbol \(T(\eta )=|\eta |^\gamma \).

We further set, as before,

$$\begin{aligned} \widehat{\varphi }_\delta :=\frac{\widehat{w}_\delta }{\Vert \widehat{w}_\delta \Vert _{L^1_\eta }} . \end{aligned}$$

With this choice

$$\begin{aligned} \langle w_\delta , (-\Delta )^{\gamma /2} w_\delta \rangle = \langle \widehat{w}_\delta , |\eta |^\gamma \widehat{w}_\delta \rangle = \int _{A_\delta } |\eta |^{-\gamma }\, \mathrm{d}\eta = \Vert \widehat{w}_\delta \Vert _{L^1_\eta }, \end{aligned}$$

hence

$$\begin{aligned} \langle \varphi _\delta , (-\Delta )^{\gamma /2}\varphi _\delta \rangle = \frac{1}{\Vert \widehat{w}_\delta \Vert _{L^1_\eta }} . \end{aligned}$$
(4.4)

As before, we still have

$$\begin{aligned} \langle \varphi _\delta , V\varphi _\delta \rangle \rightarrow \frac{1}{(2\pi )^d} \int V\, {\mathrm{d}}x \quad \text {as } \delta \rightarrow 0 \end{aligned}$$

as soon as \(A_\delta \) concentrates near zero in the limit \(\delta \rightarrow 0\). Since the function \({\mathbb {R}}^d\ni \eta \mapsto |\eta |^{-\gamma }\) has a non-integrable singularity near zero for \(\gamma \ge d\), we can make \(A_\delta \) concentrate near zero, thus having \(\Vert \widehat{w}_\delta \Vert _{L^1_\eta }\) blow up and, because of (4.4), we get \(\lim _{\delta \rightarrow 0} \langle \varphi _\delta , (-\Delta )^{\gamma /2}\varphi _\delta \rangle =0\), i.e., the kinetic energy vanishes in the limit \(\delta \rightarrow 0\) as soon as \(\gamma \ge d\).

Explicitly, choosing \(A_\delta \) to be the annulus

$$\begin{aligned} A_\delta :=\{ r_{1,\delta }<|\eta |<r_{2,\delta }\} \end{aligned}$$

we have

$$\begin{aligned} \Vert \widehat{w}_\delta \Vert _{L^1_\eta } \sim \int _{r_{1,\delta }}^{r_{2,\delta }} r^\gamma r^{d-1}\, \mathrm {d}r = \left\{ \begin{array}{ll} \ln (\frac{r_{2,\delta }}{r_{1,\delta }}) &{}{}\quad \text{ if }\;\, \gamma =d\\ \frac{1}{\gamma -d}\left[ r_{1,d}^{-(\gamma -d)}- r_{2,\delta }^{-(\gamma -d)}\right] &{}{}\quad \text{ if }\;\, \gamma >d \end{array} \right. \end{aligned}$$

and choosing \(r_{1,\delta }=\delta ^2\) and \(r_{2,\delta }= \delta \) we see \(\lim _{\delta \rightarrow 0} \Vert \widehat{w}_\delta \Vert _{L^1_\eta } =\infty \). With (4.4)

$$\begin{aligned} \lim _{\delta \rightarrow 0}\langle \varphi _\delta , ((-\Delta )^{\gamma /2}+V) \varphi _\delta \rangle = \frac{1}{(2\pi )^d} \int V\, {\mathrm{d}}x , \end{aligned}$$

follows. So bound states with strictly negative energy exist once \(\int V\, {\mathrm{d}}x<0\).

4.2 The Case of \(\int V\, {\mathrm{d}}x =0\)

To include the case where V does not vanish identically but \(\int V\, {\mathrm{d}}x=0\), we have to further modify the test function. Second order perturbation theory suggest that the test function should be modified by adding a suitable multiple of the potential V. This suggests the ansatz

$$\begin{aligned} \varphi _\delta +\alpha \phi \end{aligned}$$
(4.5)

for some \(\alpha \in {\mathbb {R}}\) and a suitably nice function \(\phi \), to be determined later, as a trial state for the computation of the energy. Using this we get, with \(T(p)=|p|^\gamma = (-\Delta )^{\gamma /2}\),

$$\begin{aligned} E(\delta ,\alpha )&:=\langle \varphi _\delta +\alpha \phi , (T(p)+V)(\varphi _\delta +\alpha \phi ) \rangle \\&= \langle \varphi _\delta , T(p)\varphi _\delta \rangle + \langle \varphi _\delta , V\varphi _\delta \rangle + 2\alpha \mathrm {Re}(\langle \varphi _\delta , T(p)\phi \rangle ) + 2\alpha \mathrm {Re}(\langle \varphi _\delta , V\phi \rangle ) \\&\quad + \alpha ^2 \langle \phi , T(p)\phi \rangle + \alpha ^2 \langle \phi , V\phi \rangle . \end{aligned}$$

From the discussion above we know

$$\begin{aligned} \lim _{\delta \rightarrow 0} \langle \varphi _\delta , T(p)\varphi _\delta \rangle= & {} 0, \nonumber \\ \lim _{\delta \rightarrow 0} \langle \varphi _\delta , V\varphi _\delta \rangle= & {} \frac{1}{(2\pi )^d} \int V\,{\mathrm{d}}x =0, \nonumber \\ \lim _{\delta \rightarrow 0} \langle \varphi _\delta , V\phi \rangle= & {} \frac{1}{(2\pi )^{d/2}} \int V\phi \,{\mathrm{d}}x , \end{aligned}$$
(4.6)

and, since T(p) is a positive operator, we also have

$$\begin{aligned} |\langle \varphi _\delta , T(p)\phi \rangle | \le \langle \varphi _\delta , T(p)\varphi _\delta \rangle ^{1/2} \langle \phi , T(p)\phi \rangle ^{1/2} \rightarrow 0 \quad \text {as }\delta \rightarrow 0. \end{aligned}$$
(4.7)

Thus

$$\begin{aligned} E(\alpha ):=\lim _{\delta \rightarrow 0} E(\delta ,\alpha ) = 2\alpha \frac{1}{(2\pi )^{d/2}} \mathrm {Re}\int V\phi \,{\mathrm{d}}x + \alpha ^2 \langle \phi , T(p)\phi \rangle + \alpha ^2 \langle \phi , V\phi \rangle \end{aligned}$$

and

$$\begin{aligned} \lim _{\alpha \rightarrow 0} \frac{E(\alpha )}{\alpha } = \frac{2}{(2\pi )^{d/2}} \mathrm {Re}\int V\phi \,{\mathrm{d}}x . \end{aligned}$$

This shows that we will have \(E(\delta ,\alpha )<0\) for some finite \(\delta >0\) and \(\alpha >0\), if we can find a Schwartz function \(\phi \in \mathcal {Q}(T(p))\) such that \(\int V\phi \,{\mathrm{d}}x <0\).

Split \(V= V_+-V_-\), the positive and negative parts of V. By assumption, \(\int V_-\, {\mathrm{d}}x = \int V_+\, {\mathrm{d}}x>0\). Take a big centered ball B such that \(\int _B V_-\, {\mathrm{d}}x >0\) and consider the set

$$\begin{aligned} D:=B\cap \{ V_->0 \} . \end{aligned}$$

Let \(\kappa _\varepsilon \in \mathcal {C}^\infty _0({\mathbb {R}}^d)\) be an approximate delta-function and set

$$\begin{aligned} \phi _\varepsilon :=\kappa _\varepsilon *\mathbf {1}_{D} . \end{aligned}$$

This is a nice infinitely often differentiable function with compact support, in particular a Schwartz function, so its Fourier transform decays rapidly, hence hence \( \phi _\varepsilon \) is in the form domains of T(p) and \(V_\pm \). By the properties of convolutions [31], we have \(0\le \phi _\varepsilon \le 1\) and \(\phi _\varepsilon \rightarrow \mathbf {1}_{D}\) in \(L^1\) for \(\varepsilon \rightarrow 0\), hence, after taking a subsequence, also pointwise almost everywhere. With slight abuse of notation we denote this subsequence still by \(\phi _\varepsilon \). With the help of Lebesgue’s dominated convergence theorem one sees

$$\begin{aligned} \lim _{\varepsilon \rightarrow 0} \int V\phi _\varepsilon \,{\mathrm{d}}x = -\int _B V_-\, {\mathrm{d}}x <0 , \end{aligned}$$

so using \(\phi _\varepsilon \) instead of \(\phi \) for some small enough \(\varepsilon >0\) in the above argument shows that there are \(\alpha ,\delta ,\varepsilon >0\) such that

$$\begin{aligned} \langle \varphi _\delta +\alpha \phi _\varepsilon , (T(p)+V)(\varphi _\delta +\alpha \phi _\varepsilon ) \rangle <0. \end{aligned}$$

Hence the variational principle shows that we have a strictly negative eigenvalue of \(T(p)+V\) also in the case where V does not vanish identically but \(\int V\, {\mathrm{d}}x=0\).

5 Existence of Bound States: Proof of the General Case

In this section we give the proof of Theorems 1.1 and 3.4 and Colloraries 3.6 and 3.7. To prepare for this, we give a lemma first, which is a convenient replacement for (4.4). We have to be a little bit careful here, to ensure that the constructed function is normalizable. The construction in (4.3) worked, because there the kinetic energy was bounded away from zero in any open set not containing zero. In the general case, where T is just measurable, this is not so clear. As an easy way out we simply cut the kinetic energy close to zero.

Lemma 5.1

Let \(T,\chi :{\mathbb {R}}^d\rightarrow [0,\infty )\) be measurable and \(0\le \chi \le 1\). For \(\tau >0\) define the function \(\widehat{w}_\tau \) by

$$\begin{aligned} \widehat{w}_\tau (\eta )= \max (T(\eta ),\tau )^{-1} \chi (\eta ) \quad \text {for }\eta \in {\mathbb {R}}^d. \end{aligned}$$

Then \(w_\tau :=\mathcal {F}^{-1}(\widehat{w}_\tau )\in \mathcal {Q}(T(p))\), the quadratic form domain of T(p), and we have the bound

$$\begin{aligned} \langle w_\tau , T(p) w_\tau \rangle \le \Vert \widehat{w}_\tau \Vert _{L^1_\eta } \, \end{aligned}$$

for its kinetic energy.

Remark 5.2

At first sight the bound provided by Lemma 5.1 seems surprising, since the left hand side of the bound scales quadratically in w but the right hand side is linear in \(\widehat{w}\). This is not a contradiction, though, since we assume that \(\widehat{w}_\tau = \max (T,\tau )^{-1}\chi \) and \(0\le \chi \le 1\), which breaks the scaling.

Proof of Lemma 5.1

This is a simple calculation. Since T is positive and by Plancherel,

$$\begin{aligned} \langle w_\tau , T(p) w_\tau \rangle&=\langle \sqrt{T(p)}w_\tau , \sqrt{T(p)} w_\tau \rangle = \langle \widehat{w}_\tau , T \widehat{w}_\tau \rangle \\&= \int _{{\mathbb {R}}^d} T(\eta ) \max (T(\eta ),\tau )^{-2}\chi (\eta )^2\, \mathrm{d}\eta \\&\le \int _{{\mathbb {R}}^d} \max (T(\eta ),\tau )^{-1}\chi (\eta )^2\, \mathrm{d}\eta \\&\le \int _{{\mathbb {R}}^d} \max (T(\eta ),\tau )^{-1}\chi (\eta )\, \mathrm{d}\eta = \Vert \widehat{w}_\tau \Vert _{L^1_\eta } , \end{aligned}$$

since, by assumption \(0\le \chi \le 1\), thus also \(0\le \chi ^2\le \chi \). \(\square \)

Now we come to the

Proof of Theorem 3.2

Since \(V\le W\), we have \(\langle \varphi , V_+\varphi \rangle \le \langle \varphi , W_+\varphi \rangle <\infty \) for all \(\varphi \in \mathcal {Q}(W_+)\), i.e., \(\mathcal {Q}(W_+)\subset \mathcal {Q}(V_+)\). Hence \(\mathcal {Q}(T(p)+W)= \mathcal {Q}(T(p))\cap \mathcal {Q}(W_+)\subset \mathcal {Q}(T(p)+V)\) and \(\langle \varphi , (T(p) +V) \varphi \rangle \le \langle \varphi , (T(p) +W)\varphi \rangle \) for all \(\varphi \in \mathcal {Q}(T(p)+W)\). The variational principle shows that \(T(p)+V\) has at least as many negative eigenvalues as \(T(p)+W\) has, [4, 8, 53]. So replacing V with W, if necessary, we can, without loss of generality, assume \(V\in L^1({\mathbb {R}}^d)\) and \(\int V\, {\mathrm{d}}x \le 0\).

Let \(\omega \in {\mathbb {R}}^d\) be such that

$$\begin{aligned} \int _{B_{1/n}(\omega )} T(\eta )^{-1}\, \mathrm{d}\eta = \infty \end{aligned}$$

for every \(n\in {\mathbb {N}}\). By monotone convergence, we have

$$\begin{aligned} \lim _{\tau \rightarrow 0} \int _{B_{1/n}(\omega )} \max (T(\eta ),\tau )^{-1} \, \mathrm {d}\eta = \int _{B_{1/n}(\omega )} T(\eta )^{-1}\, \mathrm{d}\eta = \infty , \end{aligned}$$

so there exists a sequence \(\tau _{n+1}<\tau _n\rightarrow 0\), for \(n\rightarrow \infty \), with

$$\begin{aligned} \lim _{n\rightarrow \infty } \int _{B_{1/n}(\omega )} \max (T(\eta ),\tau _n)^{-1}\,\mathrm {d}\eta = \infty . \end{aligned}$$
(5.1)

Define the functions \(\widehat{w}_n\) and \(\widehat{\varphi }_n\) by

$$\begin{aligned} \widehat{w}_n(\eta ):=\max (T(\eta ),\tau _n)^{-1} \mathbf {1}_{B_{1/n}(\eta )} \text { and } \widehat{\varphi }_n(\eta ) :=\frac{\widehat{w}_n(\eta )}{\Vert \widehat{w}_n\Vert _{L^1_\eta }} \end{aligned}$$

for every \(\eta \in {\mathbb {R}}^d\). Note that \(\widehat{w}_n\in L^1_\eta \), so \(\varphi _n\) is non-trivial. Because of Lemma 5.1, \(w_n\in \mathcal {Q}(T(p))\subset \mathcal {Q}(V)\). By construction,

$$\begin{aligned} \Vert \widehat{w}_n\Vert _{L^1_\eta } \rightarrow \infty \text { as }n\rightarrow \infty . \end{aligned}$$

In addition, since the sets \(B_{1/n}(\omega )\) concentrate around \(\omega \) and \(\widehat{\varphi }_n\) is \(L^1\) normalized, we also have that \(\widehat{\varphi }_n\) is a sequence of approximate delta-functions which concentrates at \(\omega \). Thus we have the uniform bound

$$\begin{aligned} |\varphi _n(x)|\le \frac{1}{(2\pi )^{d/2}}\Vert \widehat{\varphi }\Vert _{L^1_\eta } = \frac{1}{(2\pi )^{d/2}} \end{aligned}$$

and the pointwise limit

$$\begin{aligned} \lim _{n\rightarrow \infty } \varphi _n(x) = \frac{1}{(2\pi )^{d/2}} e^{i\omega \cdot x} \text { for all } x\in {\mathbb {R}}^d . \end{aligned}$$

Using Lebesgue’s dominated convergence theorem this shows

$$\begin{aligned} \lim _{n\rightarrow \infty } \langle \varphi _n, V\varphi _n \rangle = \frac{1}{(2\pi )^d} \int _{{\mathbb {R}}^d} V\, {\mathrm{d}}x . \end{aligned}$$

For the kinetic energy we simply note that Lemma 5.1 yields

$$\begin{aligned} \langle \varphi _n, T(p)\varphi _n \rangle = \frac{1}{\Vert \widehat{w}_n\Vert _{L^1_\eta }^2} \langle w_n, T(p)w_n \rangle \le \frac{1}{\Vert \widehat{w}_n\Vert _{L^1_\eta }} \rightarrow 0 \text { as } n\rightarrow \infty . \end{aligned}$$

So if \(\int V\, {\mathrm{d}}x<0\) we can immediately conclude

$$\begin{aligned} \lim _{n\rightarrow \infty } \langle \varphi _n, (T(p)+V)\varphi _n \rangle = \frac{1}{(2\pi )^d} \int _{{\mathbb {R}}^d} V\, {\mathrm{d}}x <0 \end{aligned}$$

and the variational principle implies that there is a strictly negative eigenvalue of \(T(p)+V\).

In the case \(\int V_+\, {\mathrm{d}}x = \int V_-\, {\mathrm{d}}x >0 \), so \(\int V\, {\mathrm{d}}x =0\), we use the construction of Sect. 4.2 to see that there exists a positive function \(\phi \in \mathcal {C}^\infty _0({\mathbb {R}}^d)\) with

$$\begin{aligned} \int _{{\mathbb {R}}^d} V\phi \, {\mathrm{d}}x <0. \end{aligned}$$

Similarly to the discussion in Sect. 4.2, we modify the trial state to the form

$$\begin{aligned} \varphi (x)= \varphi _n(x) + \alpha e^{i\omega \cdot x} \phi (x) \text { for } x\in {\mathbb {R}}^d. \end{aligned}$$

Setting \(\widetilde{\phi }(x)= e^{i\omega \cdot x} \phi (x)\) we have, analogously to the calculation in Sect. 4.2,

$$\begin{aligned} \lim _{\alpha \rightarrow 0 } \alpha ^{-1}&\lim _{n\rightarrow \infty } \langle \varphi _n + \alpha \widetilde{\phi }, (T(p)+V)(\varphi _n + \alpha \widetilde{\phi })\rangle \\&= \frac{2}{(2\pi )^{d/2}} \mathrm {Re}\int e^{-i\omega \cdot x}V(x)\widetilde{\phi }(x)\,{\mathrm{d}}x \\&= \frac{2}{(2\pi )^{d/2}} \int V(x)\phi (x)\,{\mathrm{d}}x <0 . \end{aligned}$$

So for all large enough \(n\in {\mathbb {N}}\) and small enough \(\alpha >0\)

$$\begin{aligned} \langle \varphi _n + \alpha \widetilde{\phi }, (T(p)+V)(\varphi _n + \alpha \widetilde{\phi })\rangle <0 , \end{aligned}$$

which, by the variational principle, implies the existence of at least one negative eigenvalue for \(T(p)+V\). \(\square \)

Proof of Theorem 3.4

We will first prove part b: Assume that there are k distinct points \(\omega _1,\ldots ,\omega _k\) such that

$$\begin{aligned} \int _{B_\delta (\omega _l)} T(\eta )^{-1}\, \mathrm{d}\eta = \infty \end{aligned}$$
(5.2)

for all \(r=1,\ldots ,k\) and all \(\delta >0\). Using the previous construction, we see that for each \(r=1,\ldots ,k\) there exist functions \(\varphi _{r,n}\) where the support of \(\widehat{\varphi }_{r,n}\) concentrates in Fourier space near \(\omega _r\). Then

$$\begin{aligned}&\lim _{n\rightarrow \infty } \left\langle \sum _{r=1}^k c_r\varphi _{r,n}, (T(p)+W)\left( \sum _{r=1}^k c_r\varphi _{r,n}\right) \right\rangle \\ {}&\quad = \lim _{n\rightarrow \infty } \sum _{r,s=1}^k \overline{c_r} c_s \langle \varphi _{r,n}, (T(p)+W)\varphi _{s,n}\rangle \\ {}&\quad = \frac{1}{(2\pi )^d} \sum _{r,s=1}^k\overline{c_r} c_s\int _{{\mathbb {R}}^d} W(x) e^{-ix(\omega _r-\omega _s)}\, \mathrm {d}x = \frac{1}{(2\pi )^{d/2}} \sum _{r,s=1}^k\widehat{W}(\omega _r-\omega _s) \overline{c_r}c_s \\ {}&\quad = \frac{1}{(2\pi )^{d/2}} \langle c,Mc \rangle _{{\mathbb {C}}^k} \end{aligned}$$

with the matrix \(M= (\widehat{W}(\omega _r-\omega _s))_{r,s=1,\ldots ,k}\). If this matrix is negative definite, then \(\langle c,Mc \rangle _{{\mathbb {C}}^k} <0\) for all \(c\ne 0\), thus \(T(p)+W\), hence also \(T(p)+V\), will be strictly negative on the subspace \(N_{k,n}=\text {span}(\varphi _{r,n}, r=1,\ldots ,k)\). For large n the functions \(\varphi _{r,n}\) do not overlap in Fourier-space, thus \(\dim N_{k,n}=k\) for all large enough n. This gives the existence of at least k strictly negative eigenvalues of \(T(p)+V\) by the usual variational arguments, see [4, 8, 53].

To prove part a, we simply note that if \(V\le 0\) and \(V\ne 0\), then for \(B_R\) a centered ball of radius R

$$\begin{aligned} W^m_R:=-\min (V_-, m) \mathbf {1}_{B_{R}} \end{aligned}$$

is integrable for all \(m,R>0\) and \(V\le W^m_R\le 0\). Since \(V\not =0\), one has \(\int W^m_R\, {\mathrm{d}}x <0\) for large enough mR. Using the variational principle again, we can assume \(V=W\in L^1\), \(W\le 0\), and \(W\not =0\), without loss of generality.

The matrix M above will be negative definite. For \(|c|^2=\sum _{r=1}^k |c_r|^2=1\), we have

$$\begin{aligned} \langle c,Mc \rangle _{{\mathbb {C}}^k} = \frac{1}{(2\pi )^{d/2}} \int _{{\mathbb {R}}^d} W(x)\, \left| \sum _{r=1}^k c_r e^{-ix\omega _r}\right| ^2\, \mathrm {d}x <0 \end{aligned}$$

since \(x\mapsto \sum _{r=1}^k c_r e^{-ix\omega _r}\) is real-analytic, thus not zero on any open set of positive Lebesgue measure and \(W\le 0\), \(W\ne 0\). Moreover, let \(S^{k-1}\) be the unit sphere in \({\mathbb {C}}^k\) and note that the map \(S^{k-1}\ni c\mapsto \langle c,Mc\rangle _{{\mathbb {C}}^k}\) is continuous and M is a hermitian matrix. Thus by the above, we see that the largest eigenvalue of M is negative. So M is negative definite and by part b, we conclude that \(T(p)+V\) has at least k strictly negative eigenvalues.

To prove part c, we note that it is enough to show that \(T(p)+W\) has at least k negative eigenvalues. Let M be the \(k\times k\) matrix as above and let \(a=(a_1, \ldots ,a_k)^t\) be the normalized eigenvector corresponding to the eigenvalue zero, which we assume to be non-degenerate. Let \(U_{a}\) be the \(k-1\) dimensional orthogonal complement of the vector a in \({\mathbb {C}}^k\). Furthermore, we set

$$\begin{aligned} N_{a,n}:=\left\{ \sum _{r=1}^k c_r\varphi _{r,n}:\, c=(c_1,\ldots ,c_k)^t\in U_{a} \right\} , \end{aligned}$$

where the functions \(\varphi _{r,n}\) are defined just below Eq. (5.2). Note that \(N_{a,n}\) is a \(k-1\)-dimensional subspace of \(L^2({\mathbb {R}}^d)\) for large enough n since then \(\varphi _{r,m}\), \(r=1,\ldots ,k\), have disjoint support in Fourier space. Define, for some positive \(\alpha \),

$$\begin{aligned} \widetilde{\varphi }:=\sum _{r=1}^k a_r \varphi _{r,n} + \alpha \phi \end{aligned}$$

and \(L_{k,\alpha ,n}:=\text {span}\{ N_{a,n}, \widetilde{\varphi }\}\). That is, any vector in \({L_{k,\alpha ,n}}\) can be written as

$$\begin{aligned} \psi _{n,\alpha }= \psi _{n,\alpha }(\gamma ,c) = \sum _{r=1}^k c_r\varphi _{r,n} +\gamma \left( \sum _{r=1}^k a_r \varphi _{r,n} + \alpha \phi \right) = \sum \tilde{c}_r \varphi _{r,n} + \gamma \alpha \phi \end{aligned}$$
(5.3)

were we set \(\tilde{c}= \tilde{c}(\gamma ,c)= \gamma a+c\in {\mathbb {C}}^k\). Since \(U_a\) is the orthogonal complement of a in \(C^k\), the map \(\tilde{c}:{\mathbb {C}}\times U_a\rightarrow C^k\) is a bijection.

Our goal is to show that the dimension of \(L_{k,n}\) is k and \(T(p)+W\) is negative on \(L_{k,\alpha ,n}{\setminus }\{0\}\), for large enough n and a suitable choice of \(\phi \) and \(\alpha \in {\mathbb {R}}\). Writing \(\psi _{n,\alpha }= (c+\gamma a)\varphi _{\cdot ,n}+\gamma \alpha \phi \), we have

$$\begin{aligned} \begin{aligned} \langle \psi _{n,\alpha }, (T(p)+W) \psi _{n,\alpha }\rangle&= \langle (c+\gamma a)\varphi _{\cdot ,n}, (T(p)+W)(c+\gamma a)\varphi _{\cdot ,n} \rangle \\&\quad +2\, \mathrm {Re}\langle \gamma \alpha \phi , (T(p)+W)(c+\gamma a)\varphi _{\cdot ,n}\rangle \\&\quad + |\gamma \alpha |^2 \langle \phi ,(T(p) +W)\phi \rangle . \end{aligned} \end{aligned}$$
(5.4)

Since the \(\varphi _{r,n}\), \(r=1,\ldots ,k\), have disjoint supports in Fourier space when \(n\in {\mathbb {N}}\) is large and, as before, \(\langle \varphi _{r,n}, T(p)\varphi _{r,n}\rangle = o_n(1)\rightarrow 0\) as \(n\rightarrow \infty \), we have

$$\begin{aligned} \langle (c+\gamma a)\varphi _{\cdot ,n}, T(p)(c+\gamma a)\varphi _{\cdot ,n} \rangle = o_n(1)|c+\gamma a|^2= o_n(1) (|c|^2+|\gamma |^2) \end{aligned}$$

where we also used that since a is normalized in \(C^k\) and \(c\perp a\), we have \(|c+\gamma a|^2= |c|^2+|\gamma |^2\). Also

$$\begin{aligned} \langle (c+\gamma a)\varphi _{\cdot ,n}, W(c+\gamma a)\varphi _{\cdot ,n} \rangle = \langle (c+\gamma a), (M+\Delta M_n)(c+\gamma a) \rangle _{C^k} , \end{aligned}$$

where \(M= (\widehat{W}(\omega _r-\omega _s))_{r,s=1,\ldots ,k}\) is the \(k\times k\) matrix as before, but now it has a single zero eigenvalue since \(Ma=0\) and \(\langle c, Mc\rangle _{C^k}\le -\lambda _1|c|^2 \) for some \(\lambda _1>0\) and all \(c\perp a\). Moreover, \(\Delta M_n\) is a \(k\times k\) matrix which converges to zero as \(n\rightarrow \infty \), that is, \(\langle \tilde{c},\Delta M_n \tilde{c}\rangle = o_n(1)|\tilde{c}|^2\) for all \(\tilde{c}\in C^k\). Using \(M a=0\) and \(c\perp a\), we get

$$\begin{aligned} \langle (c+\gamma a), (M+\Delta M_n)(c+\gamma a) \rangle _{C^k}&= \langle c, M c \rangle _{C^k} +\langle (c+\gamma a), \Delta M_n(c+\gamma a) \rangle _{C^k} \\&\le -\lambda _1 |c|^2 +o_n(1)(|c|^2+|\gamma |^2) . \end{aligned}$$

For the part due to the kinetic energy in the cross term we use the Cauchy Schwarz inequality to bound it as

$$\begin{aligned}&\left| \langle \gamma \alpha \phi , T(p)(c+\gamma a)\varphi _{\cdot ,n}\rangle \right| \\&\quad \le |\gamma \alpha | \langle \phi , T(p) \phi \rangle ^{1/2} \, \langle (c+\gamma a)\varphi _{\cdot ,n}, T(p)(c+\gamma a)\varphi _{\cdot ,n}\rangle ^{1/2} \\&\quad = o_n(1)\alpha |\gamma |(|c|^2+ |\gamma |^2)^{1/2}\, \le o_n(1)\alpha (|c|^2+ |\gamma |^2). \end{aligned}$$

The part due to the potential in the cross term is bounded as

$$\begin{aligned} \mathrm {Re}\langle \gamma \alpha \phi , W(c+\gamma a)\varphi _{\cdot ,n}\rangle&= \alpha \mathrm {Re}( \overline{\gamma } \langle \phi , W c\varphi _{\cdot ,n}\rangle + |\gamma |^2 \alpha \mathrm {Re}\langle \phi , W a\varphi _{\cdot ,n}\rangle \\&\le \alpha |\gamma | \Vert |W|^{1/2}\phi \Vert \Vert |W|^{1/2} c\varphi _{\cdot ,n} \Vert + |\gamma |^2 \alpha \mathrm {Re}\langle \phi , W a\varphi _{\cdot ,n}\rangle \\&\le C\alpha |\gamma ||c| + |\gamma |^2\alpha \mathrm {Re}\langle \phi , W a\varphi _{\cdot ,n}\rangle , \end{aligned}$$

where we also used

$$\begin{aligned} \Vert |W|^{1/2}\phi \Vert ^2 = \langle \phi , |W|\phi \rangle \le C \ \text {and}\ \Vert |W|^{1/2} c\varphi _{\cdot ,n} \Vert ^2 = \langle c\varphi _{\cdot ,n}, |W|c\varphi _{\cdot ,n} \rangle \le C|c|^2 \end{aligned}$$

for some large enough constant C.

Plugging all of this into (5.4) and collecting terms we get the upper bound

$$\begin{aligned} \begin{aligned} \langle \psi _{n,\alpha },&(T(p)+W) \psi _{n,\alpha }\rangle \\&\le o_n(1)(|c|^2+|\gamma |^2) -\lambda _1 |c|^2 + \alpha \big ( \mathrm {Re}\langle \phi , W a\varphi _{\cdot ,n}\rangle + C\alpha \big )|\gamma |^2 + C\alpha |\gamma ||c| . \end{aligned} \end{aligned}$$
(5.5)

for some large enough constant C, any \(\psi _{n,\alpha }\in L_{k,n,\alpha }\) in the form given by (5.3), and \(\alpha \ge 0\).

Note that

$$\begin{aligned} \langle \phi , W a\varphi _{\cdot ,n}\rangle&= (2\pi )^{-d/2}\int \overline{\phi (x)} W(x) \sum _{r=1}^k a_r\varphi _{r,n}\, {\mathrm{d}}x \rightarrow (2\pi )^{-d/2}\\ {}&\quad \times \int \overline{\phi (x)} W(x) \sum _{r=1}^k a_r e^{i\omega _r\cdot x} \, {\mathrm{d}}x \end{aligned}$$

so with the choice \(\phi (x)= \zeta (x) \sum _{r=1}^k a_r e^{i\omega _r\cdot x} \) for some real-valued function \(\zeta \)

$$\begin{aligned} \langle \phi , W a\varphi _{\cdot ,n}\rangle&\rightarrow (2\pi )^{-d/2}\int \zeta (x) W(x) \left| \sum _{r=1}^k a_r e^{i\omega _r\cdot x}\right| ^2 \, {\mathrm{d}}x . \end{aligned}$$

Of course,

$$\begin{aligned} (2\pi )^{-d/2}\int W(x) \left| \sum _{r=1}^k a_re^{i\omega _r\cdot x}\right| ^2 \, {\mathrm{d}}x =\langle a,Ma\rangle _{C^k} = 0 \end{aligned}$$

but, since \( \Big |\sum _{r=1}^k a_r e^{i\omega _r\cdot x}\Big |\) can vanish only on a set of measure zero and W is not identically zero, there must exist a real-valued Schwartz function \(\zeta _0\) with compact support such that \( \int \zeta _0(x) W(x) \Big |\sum _{r=1}^k a_r e^{i\omega _r\cdot x}\Big |^2 \, {\mathrm{d}}x <0 \). We cut this Schwartz function in Fourier space with a spherically symmetric and smooth cut-off function, to get a real-valued Schwartz function \(\zeta \) which has compact support in Fourier space and for which, by making the cut-off in Fourier space large enough, one has

$$\begin{aligned} \int \zeta (x) W(x) \left| \sum _{r=1}^k a_r e^{i\omega _r\cdot x}\right| ^2 \, {\mathrm{d}}x <0. \end{aligned}$$

With this choice we then have

$$\begin{aligned} \mathrm {Re}\langle \phi , W a\varphi _{\cdot ,n}\rangle \le -\rho \end{aligned}$$

for some \(\rho >0\) and all large enough n. Thus (5.5) implies

$$\begin{aligned} \begin{aligned}&\langle \psi _{n,\alpha }, (T(p)+W) \psi _{n,\alpha }\rangle \\ {}&\quad \le o_n(1)(|c|^2+|\gamma |^2) -\lambda _1 |c|^2 - \alpha \big ( \rho - C\alpha \big ) |\gamma |^2 +C\alpha |\gamma ||c|\\ {}&\quad \le o_n(1)(|c|^2+|\gamma |^2) -\lambda _1 |c|^2 - \alpha \big ( \rho - C\alpha \big ) |\gamma |^2 + C\alpha ^{1/2}|c|^2 + C\alpha ^{3/2}|\gamma |^2\\ {}&\quad = -\big ( o_n(1) + C\alpha ^{1/2} -\lambda _1\big )|c|^2 +\alpha \big ( C(\alpha +\alpha ^{1/2}) -\rho \big )|\gamma |^2 \\ {}&\quad \le -\frac{\lambda _1}{2}|c|^2-\frac{\alpha \rho }{2}|\gamma |^2 <0 , \end{aligned} \end{aligned}$$
(5.6)

for all large enough n and all small enough and positive \(\alpha \) and all \(c+\gamma a\ne 0\). Thus the quadratic form \(T(p)+W\) is negative definite on the space \(L_{k,n}\) and the usual min–max variational arguments show that \(T(p)+V\) has no less than \(\dim (L_{k,\alpha ,n})\) negative eigenvalues.

It remains to show that \(L_{k,n}\) has dimension k. Assume that

$$\begin{aligned} \psi _{\alpha ,n}(c,\gamma )= (c+\gamma a )\varphi _{\cdot ,n} + \gamma \alpha \phi =0 \end{aligned}$$
(5.7)

and \(\alpha \ne 0\). Since in the construction above, we chose \(\zeta _0\) to have compact support, its Fourier transform \(\widehat{\zeta }_0\) will not have compact support. The function \(\widehat{\varphi }_{r,n}\) has support in a small ball, depending on how large n is, around each center \(\omega _r\), \(r=1,\ldots ,k\). So for large enough n the supports of \(\widehat{\varphi }_{r,n}\) are pairwise disjoint and, moreover, since \(\widehat{\zeta }_0\) does not have compact support, we can choose the cut-off in Fourier space so large that the support of \(\widehat{\zeta }\) is not contained in the union of the supports of the \(\widehat{\varphi }_{r,n}\). But then (5.7) immediately implies that \(\gamma =0\). Once this is the case, the linear independence of the \(\varphi _{r,n}, r=1,\ldots ,k\) for large n shows that also \(c=0\).

Thus for fixed \(a\in C^k{\setminus }\{0\}\), the map \(C^k\ni (c+\gamma a)\mapsto \psi _{\alpha ,n}(c,\gamma )\in L_{k,n,\alpha }\), with \(c\perp a\), is a bijection. Hence \(L_{k,n,\alpha }\) is k-dimensional for all large enough n. This finishes the proof. \(\square \)

Now we come to the

Proof of Corollary 3.6

A simple calculation shows that the assumption of Theorem 3.4 is fulfilled at k distinct points \(\omega _1,\ldots ,\omega _k\). So Theorem 3.4 applies. \(\square \)

For the proof of Corollary 3.7 the following Lemma is helpful, which gives the so-called nearest point projection parametrization of a suitable open neighborhood of \(\Sigma \).

Lemma 5.3

Let \(\Sigma \) be a \(\mathcal {C}^2\) submanifold in \({\mathbb {R}}^d\) of codimension \(1\le n\le d-1\). Then for each point \(\omega \in \Sigma \), there exists a neighborhood \(\mathcal {O}\) of \(\omega \) in \({\mathbb {R}}^d\) and neighborhoods \(\mathcal {U}_1\) in \({\mathbb {R}}^{d-n}\) and \(\mathcal {U}_2\) in \({\mathbb {R}}^{n}\) both containing zero and a \(\mathcal {C}^1\) diffeomorphism \(\Psi : \mathcal {U}_1 \times \mathcal {U}_2 \rightarrow \mathcal {O}\), such that

$$\begin{aligned} \Psi (0, 0) = \omega \text { and } \Psi (y, 0) \in \Sigma \text { for all } y\in \mathcal {U}_1 . \end{aligned}$$

Moreover,

$$\begin{aligned} {\text {dist}}(\Psi (y, t),\Sigma ) = |t|. \end{aligned}$$

This type of result seems to be well-known to geometers, at least in the analytic category, see for example [54]. However, we could not find a reference which assumes only that \(\Sigma \) is a \(\mathcal {C}^2\) manifold. So for the convenience of the reader, and ours, we give the proof of this Lemma in Appendix C. We now come to the

Proof of Corollary 3.7

Assume that \(\Sigma \) has codimension \(1\le m\le d-1\). Pick a point \(\omega \in \Sigma \) and let \(\mathcal {O}\), \(\mathcal {U}_1\), \(\mathcal {U}_2\) be the neighborhoods and \(\psi \) the \(\mathcal {C}^1\) diffeomorphism from Lemma 5.3. Since \(\mathcal {O}\) is open there exists \(\delta _0>0\) such that \(B_\delta (\omega )\subset \mathcal {O}\) for all \(0<\delta \le \delta _0\). Fix such a \(\delta \) and choose \(A_1\subset \mathcal {U}_1\) and \(A_2\subset \mathcal {U}_2\) both centered closed balls in \({\mathbb {R}}^{d-m}\), respectively, \({\mathbb {R}}^m\), with

$$\begin{aligned} \psi (A_1\times A_2)\subset B_\delta (\omega ). \end{aligned}$$

We use \(\psi \) to change coordinates in the calculation of a lower bound for \(\int _{B_\delta (\omega )}T(\eta )^{-1}\, \mathrm {d}\eta \). Parametrize \(\eta \) as \(\eta =\psi (y,t)\), then the change of variables formula gives

$$\begin{aligned} \int _{B_\delta (\omega )}T(\eta )^{-1}\, \mathrm {d}\eta&\ge \int _{\psi (A_1\times A_2)}T(\eta )^{-1}\, \mathrm {d}\eta \\&= \iint _{A_1\times A_2} T(\psi (y,t)) \, |\det (D(\psi (y,t)))|\,\mathrm {d}y \mathrm {d}t\\& > rsim \iint _{A_1\times A_2} |t|^{-\gamma } \,\mathrm {d}y \mathrm {d}t \sim \int _{A_1} \Big ( \int _0^{{{\,\mathrm{diam}\,}}(A_2)} r^{-\gamma +d-1}\, \mathrm {d}r \Big )\, \mathrm {d}y\\&= \infty \int _{A_1}\, \mathrm {d}y = \infty , \end{aligned}$$

where in the second inequality we used the assumption on the symbol T, and the fact that \(D\psi \) is continuous, so \(|\det (D\psi (y,t))| > rsim 1\) on the compact set \(A_1\times A_2\). In the last steps we simply used \(\gamma \ge m\). Together with the \(k=1\) case of part c of Theorem 3.4 this shows that \(T(p)+V\) has at least one strictly negative eigenvalue if V is relatively form compact with respect to T(p), \(V\in L^1({\mathbb {R}}^d)\), and \(\int V\, {\mathrm{d}}x \le 0\).

Of course, we can pick arbitrarily many distinct points \(\omega _l\in \Sigma \) and then the above shows that for arbitrarily many distinct points \(\omega _l\in \Sigma \) one has

$$\begin{aligned} \int _{B_\delta (\omega _l)}T(\eta )^{-1}\, \mathrm {d}\eta =\infty \end{aligned}$$

for all small enough \(\delta \), hence by monotonicity also for all \(\delta >0\). Thus if \(V\ne 0\) and \(V\le 0\), the assumption of part a of Theorem 3.4 is fulfilled for any \(k\in {\mathbb {N}}\) and so \(T(p)+V\) has infinitely many strictly negative bound states in this case. \(\square \)

Finally, we will prove Theorem 1.1, by reducing it to the \(k=1\) case of Theorem 3.4.c. For this, the following Lemma is useful.

Lemma 5.4

Assume that \(T:{\mathbb {R}}^d\rightarrow [0,\infty )\) is measurable and that there exists a compact set \(M\subset {\mathbb {R}}^d\) such that (1.1) holds. Then there exists a point \(\omega \in M\) such that T has a thick zero set near \(\omega \).

Proof

By assumption we know that there exist a compact subset \(M\subset {\mathbb {R}}^d\) with

$$\begin{aligned} \int _{M_\delta } T(\eta )^{-1}\, \mathrm{d}\eta = \infty \end{aligned}$$

for all \(\delta >0\), where \(M_\delta \) is the closed \(\delta \)-neighborhood \( M_\delta = \{\eta \in {\mathbb {R}}^d:\, {\text {dist}}(\eta ,M)\le \delta \}. \)

Assume, by contradiction, that for every \(\omega \in M\) there exists \(\delta _\omega >0\) with

$$\begin{aligned} \int _{B_{\delta _\omega }(\omega )} T(\eta )^{-1}\, \mathrm{d}\eta <\infty . \end{aligned}$$

We clearly have

$$\begin{aligned} M\subset \bigcup _{\omega \in M} B_{\delta _\omega }(\omega ) \end{aligned}$$

and by compactness of M there exist a finite subcover, i.e., \(N\in {\mathbb {N}}\) and points \(\omega _l\in M \), \(l=1,\ldots ,N\), such that

$$\begin{aligned} \mathcal {O}:=\bigcup _{l=1}^N B_{\delta _{\omega _l}}(\omega _l) \supset M . \end{aligned}$$

Clearly

$$\begin{aligned} \int _{\mathcal {O}} T(\eta )^{-1} \, \mathrm{d}\eta <\infty \end{aligned}$$
(5.8)

by construction of \(\mathcal {O}\). Since M is compact and contained in the open set \(\mathcal {O}\), it has a strictly positive distance from the closed set \(\mathcal {O}^c\). Thus there exists \(\delta >0\) such that \(M_\delta \subset \mathcal {O}\), but then with (5.8) we arrive at the contradiction

$$\begin{aligned} \infty >\int _{\mathcal {O}} T(\eta )^{-1} \, \mathrm{d}\eta \ge \int _{M_\delta } T(\eta )^{-1} \, \mathrm{d}\eta = \infty . \end{aligned}$$

Hence there exists \(\omega \in M\) for which (3.1) holds. \(\square \)

Now we can give the short

Proof of Theorem 1.1

From the assumption of the Theorem and Lemma 5.4 we have that there exists a point \(\omega \in {\mathbb {R}}^d\) such that T has a thick zero set near \(\omega \) and hence the \(k=1\) case of Theorem 3.4.c applies. \(\square \)

6 Quantitative Bounds

6.1 Proof of Theorem 1.3: Quantitative Bound

Our approach is inspired by Cwikel’s proof of the Cwikel–Lieb–Rozenblum inequality. We will give a slight modification of Cwikel’s proof, which enables us to reduce his constant by a factor of two, see Lemma 6.1. For different modifications of Cwikel’s proof see [57, 59] and, in particular, [61]. Since \(T(p)+V\ge T(p)-V_-\), in the sense of quadratic forms, the variational principle shows

$$\begin{aligned} N(T(p)+V)\le N(T(p)-V_-) \end{aligned}$$

where N(A) denotes the number of negative eigenvalues of an operator A. Thus it is enough to bound the number of strictly negative eigenvalues of \(T(p)-U\), where \(U\ge 0\).

Since U is relatively form compact with respect to T(p), the operator \(\sqrt{U}(T(p)+E)^{-1/2}\) is compact for all \(E>0\), see [56, Lemma 6.28]. Let A be a compact operator with singular values \(s_j(A)\), \(j\in {\mathbb {N}}\), and let

$$\begin{aligned} n(A;1):=\#\{j\in {\mathbb {N}}:\, s_j(A)\ge 1\} \end{aligned}$$

be the number of singular values of A greater or equal to one. Furthermore, for \(E>0\) let \(N(T(p)-U,-E)\) be the number of eigenvalues of \(T(p)-U\) which are less or equal to \(-E\). The Birman–Schwinger principle, [53, Theorem 7.9.4], shows

$$\begin{aligned} N(T(p)-U,-E) = n(K_E;1) \end{aligned}$$
(6.1)

with the so-called Birman–Schwinger operator \(K_E= \sqrt{U}(T(p)+E)^{-1}\sqrt{U}\), which is also a compact operator for any \(E>0\). Factorizing \(K_E= A_E A_E^*\) with \(A= f(x)g_E(p)\), where we introduced the multiplication operator \(f=\sqrt{U}\) and the Fourier-multiplier \(g_E(p)= (T(p)+E)^{-1/2}\), the Birman–Schwinger principle shows

$$\begin{aligned} N(T(p)-U,-E) = n(A_E;1) \end{aligned}$$

since the singular values of \(A_E\) are just the square roots of the positive eigenvalues of \(K_E\). Since \(N(T(p)-U)=\lim _{E\rightarrow 0+}N(T(p)-U,-E)\), we have to control \(n(A_E;1)\) for small \(E>0\). For convenience, we will write g for \(g_E\) below. Following Cwikel, we decompose f and g as

$$\begin{aligned} f&=\sum _{n\in {\mathbb {Z}}} f_n \quad \text {and } g= \sum _{n\in {\mathbb {Z}}} g_n , \\ \text { where }f_n&:=f\mathbf {1}_{\{\alpha r^{n-1}<f\le \alpha r^n\}} \text { and } g_n:=g\mathbf {1}_{\{r^{n-1}<g\le r^n\}} \end{aligned}$$

for some \(\alpha >0\) and \(r>1\) and introduce the operators

$$\begin{aligned} B_{\alpha ,r} :=\sum _{k+l\le 1} f_k(x) g_l(p), \quad H_{\alpha ,r} :=\sum _{k+l\ge 2} f_k(x) g_l(p) . \end{aligned}$$

We have the bounds \(\square \)

Lemma 6.1

For any \(\alpha >0\) and \(r>1\) and any functions \(f,g\ge 0\) the operator \(B_\alpha \) is bounded and its operator norm is bounded by

$$\begin{aligned} \Vert B_{\alpha ,r}\Vert&\le \alpha \left( \frac{r^4}{r^2-1} \right) ^{1/2} . \end{aligned}$$

Moreover, define \(\widetilde{G}_\alpha (u)\) for \(u,\alpha >0\) by

$$\begin{aligned} \widetilde{G}_\alpha (u):=u^2\int \limits _{ug(\eta )>\alpha } g(\eta )^{2}\,\frac{\mathrm {d}\eta }{(2\pi )^d}. \end{aligned}$$

If for \(\alpha >0\) we have \(\int _{{\mathbb {R}}^d} G_\alpha (f(x))\, \mathrm {d}x<\infty \), then \(H_{\alpha ,r}\) is a Hilbert–Schmidt operator for all \( r>1\) and its Hilbert–Schmidt norm is bounded by

$$\begin{aligned} \Vert H_{\alpha ,r}\Vert _{HS}^2 \le \int _{{\mathbb {R}}^d} \widetilde{G}_\alpha (f(x))\, {\mathrm{d}}x . \end{aligned}$$

Remarks 6.2

  1. (i)

    If g is ‘locally’ \(L^2\) in the sense that \(g\mathbf {1}_{\{g>\alpha \}}\in L^2({\mathbb {R}}^d)\) for any \(\alpha >0\), then \(\widetilde{G}_\alpha (u)<\infty \) for all \(u,\alpha >0\) and \(\lim _{u\rightarrow 0} \widetilde{G}_\alpha (u)=0\) for any \(\alpha >0\).

  2. (ii)

    Note that the right hand side of the bound on the operator norm of \(B_{\alpha ,r}\) is minimized by the choice \(r=\sqrt{2}\) and the bound for the Hilbert–Schmidt norm of \(H_{\alpha ,r}\) is independent of \(r>1\). This improves the constant from Cwikel’s original proof by a factor of two. We will use the choice \(r=\sqrt{2}\) later.

Before we give the proof of the lemma, we state and prove an immediate consequence.

Corollary 6.3

If \(\int _{{\mathbb {R}}^d} \widetilde{G}_\alpha (f(x))\, \mathrm {d}x<\infty \) for all \(\alpha >0\), then the operator f(x)g(p) is compact.

Proof

By Lemma 6.1 we have \(f(x)g(p)= B_\alpha +H_\alpha \), where \(B_\alpha \) is bounded and \(H_\alpha \) is a Hilbert–Schmidt, in particular, a compact operator. Since the operator norm of \(B_\alpha \) is bounded by \(\Vert B_\alpha \Vert \le 2\alpha \), where we chose \(r=2\) for convenience, we see that f(x)g(p) is the norm limit, as \(\alpha \rightarrow 0\), of the compact operators \(H_\alpha \), so it must be compact. \(\square \)

Proof of Lemma 6.1

The proof of the bound for the operator norm of \(B_{\alpha ,r}\) follows Cwikel’s ideas closely, with the difference that we defined \(B_{\alpha ,r}\) slightly differentlyFootnote 7 than Cwikel in [11]. We give the short proof for the convenience of the reader and in order to implement a little trick, which allows to improve on Cwikel’s constant by a factor of two:

Let \(\Psi ,\Phi \in L^2({\mathbb {R}}^d)\), \(\widetilde{f}_k:=\alpha ^{-1} r^{-k} f_k \) and \(\widetilde{g}_l:=r^{-l} g_l\) . Then

$$\begin{aligned} \langle \Psi , B_{\alpha ,r}\Phi \rangle&= \sum _{k+l\le 1} \langle f_k(x)\psi , g_l(p)\Phi \rangle =\alpha \sum _{k+l\le 1} r^{k+l} \langle \widetilde{f}_k\Psi , \widetilde{g}_l\widehat{\Phi } \rangle \\&= \alpha \sum _{k\in {\mathbb {Z}}} \Big \langle \widetilde{f}_k\Psi , \sum _{n\le 1}r^n\,\widetilde{g}_{n-k}\widehat{\Phi } \Big \rangle \end{aligned}$$

Thus

$$\begin{aligned} |\langle \Psi , B_{\alpha ,r}\Phi \rangle |&\alpha \le \sum _{k\in {\mathbb {Z}}} \Big |\Big \langle \widetilde{f}_k\Psi , \sum _{n\le 1}r^n\,\widetilde{g}_{n-k}\widehat{\Phi } \Big \rangle \Big | \le \alpha \sum _{k\in {\mathbb {Z}}} \big \Vert \widetilde{f}_k\Psi \big \Vert \Big \Vert \sum _{n\le 1}r^n\,\widetilde{g}_{n-k}\widehat{\Phi } \Big \Vert \\&\le \alpha \left( \sum _{k\in {\mathbb {Z}}} \Vert (\widetilde{f}_k\Psi \Vert ^2 \right) ^{1/2}\, \left( \sum _{n\le 1} \Big \Vert \sum _{n\le 1}r^n\, \widetilde{g}_{n-k}\widehat{\Phi } \Big \Vert ^2 \right) ^{1/2} . \end{aligned}$$

Moreover, \(0\le \widetilde{f}_k\le 1\) and they have disjoint supports, so \(\sum _{k\in {\mathbb {Z}}}(\widetilde{f}_k)^2\le 1 \) pointwise, hence

$$\begin{aligned} \sum _{k+l\in {\mathbb {Z}}} \Vert \widetilde{f}_k\psi \Vert ^2 = \Big \langle \Psi , \sum _{k\in {\mathbb {Z}}}(\widetilde{f}_k)^2 \Psi \Big \rangle \le \langle \Psi , \mathbf {1}_{\{f>0\}} \Psi \rangle \le \Vert \Psi \Vert ^2. \end{aligned}$$

For the second term, we note

$$\begin{aligned} \Big \Vert \sum _{n\le 1} r^{n}\, \widetilde{g}_{n-k}\widehat{\Phi } \Big \Vert ^2&= \sum _{n_1,n_2\le 1} r^{n_1+n_2} \langle \widehat{\Phi }, \widetilde{g}_{n_1-k}\widetilde{g}_{n_2-k}\widehat{\Phi } \rangle = \sum _{n\le 1} r^{2n} \langle \widehat{\Phi }, (\widetilde{g}_{n-k})^2\widehat{\Phi } \end{aligned}$$

since \(\widetilde{g}_{n_1-k}\) and \(\widetilde{g}_{n_2-k}\) have disjoint supports when \(n_1\ne n_2\). In addition, we also have \(\sum _{k\in {\mathbb {Z}}} (\widetilde{g}_{n-k})^2 \le 1\) for any \(n\in {\mathbb {Z}}\), hence

$$\begin{aligned} \sum _{k\in {\mathbb {Z}}} \Big \Vert \sum _{n\le 1} r^{n}\, \widetilde{g}_{n-k}\widehat{\Phi } \Big \Vert ^2&= \sum _{n\le 1} r^{2n} \Big \langle \widehat{\Phi }, \sum _{k\in {\mathbb {Z}}}(\widetilde{g}_{n-k})^2\widehat{\Phi } \le \sum _{n\le 1} r^{2n} \Vert \widehat{\Phi }\Vert ^2 = \frac{r^4}{r^2-1} \Vert \widehat{\Phi }\Vert ^2 \end{aligned}$$

follows. So we get the bound

$$\begin{aligned} \Vert B_{\alpha ,r}\Vert \le \alpha \left( \frac{r^4}{r^2-1} \right) ^{1/2} \end{aligned}$$

for the operator norm of \(B_{\alpha ,r}\).

The bound of the Hilbert–Schmidt norm of \(H_\alpha \) is a simple calculation. It is convenient to consider the operator \(\widetilde{H}_{\alpha }= f(x)\mathcal {F}^{-1}g(\eta )\), since \(H_\alpha ^*H_\alpha \) is unitarily equivalent to \(\widetilde{H}_{\alpha }^*\widetilde{H}_{\alpha }\), so their Hilbert–Schmidt norms are the same. The advantage is that one can easily read off the kernel of \(\widetilde{H}_{\alpha }\), for which we have the bound

$$\begin{aligned} |\widetilde{H}_{t}(x,\eta )|&\le (2\pi )^{-d/2} \sum _{k+l\ge 2} f_k(x) g_l(\eta ) = (2\pi )^{-d/2}\! \sum _{k+l\ge 2} f_k(x) g_l(\eta )\mathbf {1}_{\{f(x)g(\eta )>\alpha \}} \\&\le (2\pi )^{-d/2}f(x) g(\eta )\mathbf {1}_{\{f(x)g(\eta )>\alpha \}}, \end{aligned}$$

since the supports of \(f_k\), \(g_l\), respectively, are pairwise disjoint and for \((x,\eta )\) in the support of \(f_k g_l\) we have \( f(x) g(\eta ) = f_k(x)g_l(\eta )> tr^{k+l-2}\ge t \) by construction of \(f_k\) and \(g_l\) and since \(k+l\ge 2\) in the above sum. Thus with Tonelli’s theorem one sees

$$\begin{aligned} \Vert H_\alpha \Vert _{HS}^2&= \Vert \widetilde{H}_\alpha \Vert _{HS}^2 = \iint |\widetilde{H}_\alpha (x,\eta )|^2\, \mathrm {d}x\mathrm {d}\eta \\&\le (2\pi )^{-d}\iint \limits _{f(x)g(\eta )>\alpha } f(x)^2g(\eta )^2\, \mathrm {d}\eta \mathrm {d}x \\&= \int f(x)^2 \int \limits _{f(x)g(\eta )>\alpha }g(\eta )^2\,\frac{\mathrm {d}\eta \mathrm {d}x}{(2\pi )^d} =\int _{{\mathbb {R}}^d} \widetilde{G}_\alpha (f(x))\, \mathrm {d}x \end{aligned}$$

with

$$\begin{aligned} \widetilde{G}_\alpha (u)= u^2\int \limits _{ug(\eta )>\alpha } g(\eta )^2\,\frac{\mathrm {d}\eta }{(2\pi )^d}. \end{aligned}$$

as claimed. \(\square \)

Now we come to the

Proof of Theorem 1.3

The usual arguments, see [56, Lemma 6.26] or [53, Theorem 7.8.3], show that the essential spectrum does not change, \(\sigma _{\mathrm{ess}}(T(p)+V)=\sigma _{\mathrm{ess}}(T(p))\), when V is a relatively form compact perturbation of T(p). That \(\sigma _{\mathrm{ess}}(T(p))=\sigma (T(p))\subset [0,\infty )\) is clear, since T(p) is a Fourier multiplier with a positive symbol T.

It remains to prove the bound (1.3): As already discussed in the beginning of this section, setting \(f=V_-^{1/2}\) and \(g=g_E=(T+E)^{-1/2}\), the Birman Schwinger principle and the variational theorem yield

$$\begin{aligned} \begin{aligned} N(T(p)+V, -E)&= n(\sqrt{V_-}(T(p)+E)^{-1/2};1) = \#\{n:\, s_n(f(x)g(p))\ge 1\} \\&\le \sum _{n\in {\mathbb {N}}} \frac{(s_n(f(x)g(p))-\mu )_+^2}{(1-\mu )^2} = \sum _{n\in {\mathbb {N}}} \frac{(s_n(B_\alpha +H_\alpha )-\mu )_+^2}{(1-\mu )^2} \end{aligned} \end{aligned}$$
(6.2)

for any \(0\le \mu <1\), where the inequality follows from the simple bound \((s-\mu )_+^2/(1-\mu )^2\ge 1\) for all \(s\ge 1\) and where we split \( f(x)g(p) = B_\alpha +H_\alpha \), with the optimal choice of \(r=2\).

Ky–Fan’s inequality for the singular values and the first part of Lemma 6.1 gives

$$\begin{aligned} s_n(B_\alpha +H_\alpha )\le s_1(B_\alpha )+s_n(H_\alpha )= \Vert B_\alpha \Vert +s_n(H_\alpha )\le 2\alpha +s_n(H_\alpha ) . \end{aligned}$$

So choosing \(\mu =2\alpha \) in (6.2), we arrive at the bound

$$\begin{aligned} N(T(p)+V, -E)\le (1-2\alpha )^{-2} \Vert H_\alpha \Vert _{HS}^2 \le (1-2\alpha )^{-2} \int _{{\mathbb {R}}^d} \widetilde{G}_\alpha (f(x))\, {\mathrm{d}}x , \end{aligned}$$

for all \(0<\alpha <1/2\). Since \(g=g_E=(T+E)^{-1/2}\) and \(f=\sqrt{V_-}\) a straightforward calculation and a simple monotonicity argument shows

$$\begin{aligned} \widetilde{G}_\alpha (u)&= u^2\int _{T+E<u^2/\alpha ^2} \frac{1}{T(\eta )+E}\frac{\mathrm {d}\eta }{(2\pi )^d}\\&\le u^2\int _{T<u^2/\alpha ^2} \frac{1}{T(\eta )}\frac{\mathrm {d}\eta }{(2\pi )^d} = \alpha ^2 G\left( \frac{u^2}{\alpha ^2}\right) \end{aligned}$$

with G from (1.2). So, since \(f(x)=\sqrt{V_-(x)}\), we have

$$\begin{aligned} N(T(p)+V,-E)\le \frac{\alpha ^2}{(1-2\alpha )^2} \int _{{\mathbb {R}}^d} G(V_-(x)/\alpha ^2)\, \mathrm {d}x \end{aligned}$$
(6.3)

and letting \(E\rightarrow 0\) finishes the proof. \(\square \)

6.2 Proof of Theorem 1.5: Dichotomy

We start with

Lemma 6.4

Under the conditions of Theorem 1.5 we have

$$\begin{aligned} \int _{Z_\delta }\frac{1}{T}\, \mathrm {d}\eta =\infty \text { for some }\delta>0&\Longrightarrow \int _{T<u} \frac{1}{T}\, \mathrm {d}\eta =\infty \text { for all }u>0 \\&\Longrightarrow \int _{Z_\delta } \frac{1}{T}\, \mathrm {d}\eta =\infty \text { for all }\delta >0 . \end{aligned}$$

Remark 6.5

Lemma 6.4 clearly shows

$$\begin{aligned} \int _{Z_\delta }\frac{1}{T}\, \mathrm {d}\eta =\infty \text { for some }\delta>0&\Longleftrightarrow \int _{Z_\delta }\frac{1}{T}\, \mathrm {d}\eta =\infty \text { for all }\delta >0 \end{aligned}$$

and

$$\begin{aligned} \int _{Z_\delta }\frac{1}{T}\, \mathrm {d}\eta<\infty \text { for some }\delta>0&\Longleftrightarrow \int _{T<u}\frac{1}{T}\, \mathrm {d}\eta =\infty \text { for some } u>0 , \end{aligned}$$

which explains Remark 1.6.iii.

Proof of Lemma 6.4

Note the simple identity

$$\begin{aligned} \int _{T<u} \frac{1}{T}\, \mathrm {d}\eta= & {} \int _{\{T<u\}\cap Z_\delta } \frac{1}{T} \, \mathrm {d}\eta + \int _{\{T< u\}\cap Z_\delta ^c} \frac{1}{T} \, \mathrm {d}\eta \nonumber \\= & {} \int _{Z_\delta } \frac{1}{T} \, \mathrm {d}\eta - \int _{\{T\ge u\}\cap Z_\delta } \frac{1}{T} \, \mathrm {d}\eta + \int _{\{T< u\}\cap Z_\delta ^c} \frac{1}{T} \, \mathrm {d}\eta \end{aligned}$$
(6.4)

where \(\int _{\{T<u\}\cap Z_\delta ^c} \frac{1}{T} \, \mathrm {d}\eta <\infty \) for all \(\delta >0\) and all small enough \(u>0\), depending on \(\delta \), because of (1.7). Also \(\int _{\{T\ge u\}\cap Z_\delta ^c} \frac{1}{T} \, \mathrm {d}\eta \le |Z_\delta |/u<\infty \) by assumption. So the left hand side of (6.4) is infinite for all small enough \(u>0\) if for some \(\delta >0\) we have \(\int _{Z_\delta }\frac{1}{T}\, \mathrm {d}\eta =\infty \). But by monotonicity, then also \(\int _{T<u} \frac{1}{T}\, \mathrm {d}\eta =\infty \) for all \(u>0\), which proves the first implication in Lemma 6.4.

On the other hand, once \(\int _{T<u} \frac{1}{T}\, \mathrm {d}\eta =\infty \) for all \(u>0\), one sees from (6.4) that \(\int _{ Z_\delta } \frac{1}{T} \, \mathrm {d}\eta =\infty \) for any \(\delta >0\), since the last two terms in (6.4) are finite for all small enough \(u>0\). \(\square \)

Now we come to the

Proof of Theorem 1.5

For part (a) we note that by Lemma 6.4 one has

$$\begin{aligned} \int _{Z_\delta }\frac{1}{T}\, \mathrm {d}\eta =\infty \text { for some }\delta>0&\Longleftrightarrow \int _{Z_\delta }\frac{1}{T}\, \mathrm {d}\eta =\infty \text { for all }\delta >0 \end{aligned}$$

so one can use Theorem 1.1 to see that one weakly coupled bound states exist once (1.8) holds.

On the other hand, assume that (1.8) fails. Then, again by Lemma 6.4, we have \(\int _{T<u} \frac{1}{T}\, \mathrm {d}\eta <\infty \) for all small enough \(u>0\). Thus G(u) defined in (1.2) is finite for all small enough \( u>0\) and \(\lim _{u\rightarrow 0+}G(u)=0\). Then a simple argument, see Remark 1.4.i, yields a strictly negative potential V such that \(T(p)+V\) has no negative spectrum. Thus condition (1.8) is equivalent to having weakly coupled bound states.

For part (b) we simply note that Lemma 6.4 shows that \(\int _{Z_\delta }\frac{1}{T}\mathrm {d}\eta <\infty \) for some \(\delta >0\) implies \(\int _{T<0}\frac{1}{T}\mathrm {d}\eta <\infty \) for all small enough \(u>0\). Then Theorem 1.3 shows that a quantitative bound on the number of strictly negative eigenvalues of \(T(p)+V\) in the form (1.3) holds.

Conversely, assume that (1.9) fails. Then Theorem 1.1 applies. Thus weakly coupled bound states always exist for any non-trivial attractive potentials, hence no quantitative bound on the number of strictly negative bound states can exist. \(\square \)