\magnification=\magstep1 \input amstex \documentstyle{amsppt} \baselineskip=20pt \parskip=6pt \hsize=6.5truein \vsize=9truein \NoBlackBoxes \define\bb{\bold{B}} \define\pp{\Bbb{P}} \define\hh{\Bbb{H}} \centerline{\bf On Moments of Negative Eigenvalues for the Pauli Operator} \bigskip \centerline{Zhongwei Shen*\footnote{ Research supported in part by the AMS Centennial Research Fellowship, the MSRI at Berkeley, California, and the NSF grant DMS-9596266.}} \centerline{Department of Mathematics} \centerline{University of Kentucky} \centerline{Lexington, KY 40506} \centerline{E-mail: shenz\@ms.uky.edu} \centerline{Tel: (606) 257-3231} \centerline{Fax: (606) 257-4078} \bigskip \bigskip \centerline{\bf Dedicated to the Memory of Professor Ruilin Long} \bigskip \bigskip \noindent{\bf Abstract.}\ \ This paper concerns the three-dimensional Pauli operator $$ \Bbb{P}=\big( \bold{\sigma} \cdot (\bold{p}-\bold{A}(x))\big)^2 +V(x) $$ with a non-homogeneous magnetic field $\bold{B}=\text{curl\,}\bold{A}$. The following Lieb-Thirring type inequality for the moment of negative eigenvalues is established: $$ \sum_{\lambda_j<0} |\lambda_j| \le C_1 \int_{\Bbb{R}^3}|V(x)|_-^{5/2}dx + C_2\int_{\Bbb{R}^3} [b_p(x)]^{3/2} |V(x)|_-dx $$ where $p>3/2$ and $b_p(x)$ is the $L^p$ average of $|\bb|$ over certain cube centered at $x$ with a side length scaling like $|\bb|^{-1/2}$. We also show that, if $\bb$ has a constant direction, $$ \sum_{\lambda_j<0} |\lambda_j|^\gamma \le C_{1,\gamma} \int_{\Bbb{R}^3}|V(x)|_-^{\gamma+3/2}dx + C_{2,\gamma} \int_{\Bbb{R}^3} b_p(x) |V(x)|_-^{\gamma +1/2}dx $$ where $\gamma>1/2$ and $p>1$. \vfill\eject \centerline{\bf 1. Introduction} Consider the Pauli operator $$ \Bbb{P}=\Bbb{P}_0 +V(x) =\big(\bold{\sigma}\cdot(\bold{p}-\bold{A}(x))\big)^2 +V(x), \tag 1.1 $$ acting on $L^2(\Bbb{R}^3,\Bbb{C}^2)$. Here $\bold{\sigma}=(\sigma_1,\sigma_2,\sigma_3)$ denotes the vector of Pauli matrices, $\bold{p}=-i\nabla$, and $\bold{A}(x)=(A_1(x),A_2(x),A_3(x))\in L^2_{loc}(\Bbb{R}^3,\Bbb{R}^3)$ is a vector potential. We shall use $\bold{B}=\text{curl\,}\bold{A}$ to denote the magnetic field generated by the potential $\bold{A}$. The goal of this paper is to establish the Lieb-Thirring type estimates for the moments of negative eigenvalues of $\Bbb{P}$: $$ \Cal{M}_\gamma=\sum_{\lambda_j<0}|\lambda_j|^\gamma,\ \ \ \ \ \ \gamma >0. \tag 1.2 $$ We remark that in quantum mechanics $\Bbb{P}$ is used to describe the motion of a charged spin-1/2 particle in an electromagnetic field. Lieb-Thirring type estimates are an important tool in the study of stability of matter and in semi-classical analysis \cite{13,10,11,12,2,4,5}. In the case of Schr\"odinger operator $-\Delta +V(x)$ in $\Bbb{R}^n$, the classical Lieb-Thirring inequality \cite{13} states that $$ \Cal{M}_\gamma \le L_{\gamma,n}\int_{\Bbb{R}^n} |V(x)|_-^{\gamma +n/2}dx \tag 1.3 $$ where $\gamma >0$, $n\ge 2$, and $|V|_-=\max (-V,0)$ denotes the negative part of $V$ (in the case $\gamma=0$ and $n\ge 3$, (1.3) is the well-known Rosenblum-Lieb-Cwickel estimate \cite{15}). For the Pauli operator $\Bbb{P}$ in $\Bbb{R}^3$ with a constant field, it was shown in \cite{11} that $$ \Cal{M}_\gamma \le C_{1,\gamma}\int_{\Bbb{R}^3} |V(x)|_-^{\gamma+3/2}dx + C_{2,\gamma}\, |\bold{B}|\int_{\Bbb{R}^3} |V(x)|_-^{\gamma+1/2}dx \tag 1.4 $$ for $\gamma >1/2$. Subsequently, L.~Erd\"os \cite{3} initiated the study of Lieb-Thirring estimates for Pauli operators with non-homogeneous fields. He observed that the direct extension of (1.4) to the case of the non-constant fields: $$ \Cal{M}_\gamma \le C_{1,\gamma}\int_{\Bbb{R}^3} |V(x)|_-^{\gamma+3/2}dx + C_{2,\gamma} \int_{\Bbb{R}^3}|\bold{B}(x)|\, |V(x)|_-^{\gamma+1/2}dx \tag 1.5 $$ as well as its consequence (by H\"older's inequality) $$ \Cal{M}_\gamma \le \widetilde{C}_{1,\gamma}\int_{\Bbb{R}^3} |V(x)|_-^{\gamma+3/2}dx + \widetilde{C}_{2,\gamma} \int_{\Bbb{R}^3}|\bold{B}(x)|^{3/2} |V(x)|_-^{\gamma}dx \tag 1.6 $$ is false without substantial regularity conditions on $\bb$. Moreover he conjectured that it would be necessary to replace $|\bold{B}|$ by some screened version of $|\bold{B}|$. Motivated by Erd\"os' observation, A.~Sobolev \cite{19,20} obtained estimates in the forms of (1.5) and (1.6), but with $|\bold{B}|$ replaced by a so-called effective (scalar) magnetic field $b(x)$: $$ \sum_{\lambda_j<0} |\lambda_j| \le C_1\int_{\Bbb{R}^3} |V(x)|_-^{5/2}dx +C_2\int_{\Bbb{R}^3} b(x)^{3/2} |V(x)|_-dx.\tag 1.7 $$ Roughly speaking, $b(x)$ is a slow varying function which dominates $|\bold{B}(x)|$ pointwise. We remark that the two-dimensional Pauli operator as well as the three-dimensional case with a constant direction field was also investigated in \cite{3,19,20}. Recently, L.~Bugliaro, C.~Fefferman, J.~Fr\"ohlich, G.~Graf, and J.~Stubbe \cite{2} established (1.7) with a $b(x)$ whose energy is comparable to that of $|\bb|$ : \ $\| b\|_{L^2}\approx \|\bb\|_{L^2}$. This in particular implies the following estimate of E.~Lieb, M.~Loss, and J.~Solovej \cite{10}: $$ \Cal{M}_1 \le C_1\int_{\Bbb{R}^3}|V(x)|_-^{5/2}dx + C_2\left(\int_{\Bbb{R}^3}|\bold{B}(x)|^2dx\right)^{3/4} \left(\int_{\Bbb{R}^3}|V(x)|_-^4dx\right)^{1/4}. \tag 1.8 $$ In this paper we further extend the results in \cite {3,19,20,2}. The main novelty of our work is that the effective field $b(x)$ we found is much simple and more natural than the previous ones in \cite{20,2}. Indeed, $b(x)=b_p(x)$ is defined to be the $L^p$ average of $|\bb|$ over a suitable cube centered at $x$ with a side length scaling like $|\bold{B}|^{-1/2}$. We believe that this choice of the effective field is optimal. To state our main results, we first define the basic length scale: $$ \ell_p(x) =\sup\left\{ \ell>0:\ \ell^2\left(\frac{1}{\ell^3}\int_{Q(x,\ell)} |\bb(y)|^pdy\right)^{1/p}\le 1\right\}\tag 1.9 $$ where $Q(x,\ell)$ denotes the cube centered at $x$ with side length $\ell$. It is easy to see that if $|\bb|\in L^p_{loc}(\Bbb{R}^3)$ for some $p>3/2$, then $0<\ell_p(x)<\infty$ unless $|\bb|\equiv 0$. Our effective field is now given by $$ b_p(x)\equiv\frac{1}{\{ \ell_p(x)\}^2} =\left\{\frac{1}{\ell^3_p(x)}\int_{Q(x,\ell_p(x))} |\bb(y)|^pdy\right\}^{1/p}, \tag 1.10 $$ i.e., $b_p(x)$ is the $L^p$ average of $|\bb|$ over the cube centered at $x$ with side length $\ell_p(x)$. In particular, we have $\|b_p\|_{L^q}\le C\|\bold{B}\|_{L^q}$ for any $q\ge p$. \proclaim{\bf Proposition 1.1} Let $p>3/2$. Then, for any $q\ge p$, there exists a constant $C_{p,q}>0$ such that $$ \int_{\Bbb{R}^3} |b_p(x)|^q dx \le C_{p,q}\int_{\Bbb{R}^3} |\bb(x)|^q dx. \tag 1.11 $$ \endproclaim The following are the main results of the paper. \proclaim{\bf Theorem 1.1} Let $p>3/2$ and $\gamma\ge 1$. Then there exist constants $C_1(\gamma,p) >0$ and $C_2 (\gamma, p)>0$ such that $$ \sum_{\lambda_j<0} |\lambda_j|^\gamma \le C_1(\gamma, p)\int_{\Bbb{R}^3}|V(x)|_-^{\gamma +3/2}dx + C_2(\gamma, p)\int_{\Bbb{R}^3} \left[ b_p(x)\right]^{3/2} |V(x)|_-^\gamma dx. $$ \endproclaim We have a stronger estimate for magnetic fields with constant directions. \proclaim{\bf Theorem 1.2} Let $p>1$ and $\gamma>1/2$. Suppose that $\bb$ has a constant direction. Then there exist constants $C_3(\gamma,p) >0$ and $C_4 (\gamma, p)>0$ such that $$ \sum_{\lambda_j<0} |\lambda_j|^\gamma \le C_3(\gamma, p)\int_{\Bbb{R}^3}|V(x)|_-^{\gamma +3/2}dx + C_4(\gamma, p)\int_{\Bbb{R}^3} b_p(x) |V(x)|_-^{\gamma+1/2} dx. $$ \endproclaim A few remarks are in order. \remark{\bf Remark 1.1} The definition of $b_p(x)$ is motivated by the auxiliary function $m(x,|\bb|)$, which is instrumental in the study of eigenvalue problems for the magnetic Schr\"odinger operator $-(\bold{p}-\bold{A}(x))^2 +V(x)$ with certain degenerate potentials \cite{16,17}. We should point out that, although the definition (1.9) is not rotation invariant, the basic length scales $\ell_p(x)$, hence $b_p(x)$, are equivalent in different coordinates systems. One can certainly use balls centered at $x$ instead of cubes in (1.9). We also remark that, if the components of $\bold{B}$ are polynomials, then $$ b_p(x)\approx \sum_{|\alpha|\le k} |\nabla^\alpha \bold{B}(x)|^\frac{2}{|\alpha|+2}\tag 1.12 $$ where $k$ is the degree of $\bb$. \endremark \remark{\bf Remark 1.2} Suppose that $\bold{B}=\bold{B}_1+\bold{B}_2$ where $|\bold{B}_1|\in L^\infty(\Bbb{R}^3)$ and $|\bold{B}_2|\in L^p (\Bbb{R}^3)$ for some $p>3/2$. Let $\ell =\ell_p(x)$, then $$ \ell^2\left(\frac{1}{\ell^3}\int_{Q(x,\ell)}|\bold{B}(y)|^pdy\right)^{1/p}=1. $$ It follows that $$ \ell^2 \|\bold{B}_1\|_{L^\infty} +\ell^{2-3/p}\|\bold{B}_2\|_{L^p}\ge 1. $$ A simple computation yields that $$ b_p(x) =\frac{1}{\ell^2} \le C\big\{ \|\bold{B}_1\|_{L^\infty} +\|\bold{B}_2\|_{L^p}^\frac{2p}{2p-3}\big\}. \tag 1.13 $$ This, together with Theorem 1.1 ($p=2$), gives the estimate in \cite{3, Theorem 2.4, p.635}. Same argument also shows that Theorem 2.1 in \cite{3, p.632} follows from Theorem 1.2. \endremark \remark{\bf Remark 1.3} To compare our results with that in \cite{2}, we note that the effective field in \cite {2} is defined by $\widetilde{b}(x)=1/[r(x)]^2$ where $$ \frac{1}{r(x)} =\int_{\Bbb{R}^3} \varphi\left(\frac{y-x}{r(x)}\right) |\bb(y)|^2dy\tag 1.14 $$ and $\varphi(z)=(1+\frac{1}{2}z^2)^{-2}$. Let $r=r(x)$. It is easy to see that $$ \int_{Q(x,r)}|\bb(y)|^2dy\le \frac{C}{r}. $$ Thus, if $\delta$ is small, we have $$ (\delta r)^2\left\{\frac{1}{(\delta r)^3} \int_{Q(x,\delta r)} |\bb(y)|^2dy\right\}^{1/2} \le C\delta^{1/2}<1. $$ By definition, $\ell_2(x) \ge \delta r$. Hence, $$ b_2(x)=\frac{1}{\{ \ell_2(x)\}^2} \le \frac{1}{(\delta r)^2} =\frac{\widetilde{b}(x)}{\delta ^2}. $$ Thus Theorem 1.1 extends the results in \cite {2}, with an effective field much easier to compute. One may deduce the estimates in \cite{20} from Theorems 1.1 and 1.2 with $p=\infty$ by a similar argument. \endremark \remark{\bf Remark 1.4} In Theorems 1.1 and 1.2, we have implicitly assumed that $\Bbb{P}_0+V$ is a self-adjoint operator. In fact, if the right hand side of the estimate in Theorem 1.1 (or 1.2) is finite, then $\Bbb{P}_0+V$ has a unique self-adjoint realization on $L^2(\Bbb{R}^3,\Bbb{C}^2)$ associated with its quadratic form. Indeed, under the assumption $\bold{A}\in L^2_{loc}(\Bbb{R}^3, \Bbb{R}^3)$, we may define $\Bbb{P}_0$ as the self-adjoint operator associated with the closed quadratic form $\| \Bbb{D}\psi\|_{L^2}^2$ where $\Bbb{D}=\sigma\cdot (\bold{p}-\bold{A})$, i.e., $\Bbb{P}_0 =\Bbb{D}^*\Bbb{D}$. Let $V_j=V\chi_{\{ |V|\le j\} }$. Since $V_j$ is bounded, the operator $\Bbb{P}_0-\frac{1}{\varepsilon}|V_j|$ is self-adjoint with a form core Domain$(\Bbb{D})$ for any $\varepsilon\in (0,1)$. We now apply the variation principle and Theorem 1.1 to $\Bbb{P}_0-\frac{1}{\varepsilon}|V_j|$. We obtain $$ \aligned \int_{\Bbb{R}^3} |\sigma\cdot (\bold{p} & -\bold{A})\psi|^2dx -\frac{1}{\varepsilon}\int_{\Bbb{R}^3} |V_j|\, |\psi|^2dx\\ &\ge \lambda_1(\varepsilon) \int_{\Bbb{R}^3} |\psi|^2dx\ge -C(\delta, \varepsilon, V, \bold{B}) \int_{\Bbb{R}^3}|\psi|^2dx \endaligned $$ for any $\psi\in\text{Domain} (\Bbb{D})$. Thus, $$ \int_{\Bbb{R}^3} |V_j|\, |\psi|^2dx \le \varepsilon\int_{\Bbb{R}^3} |\sigma\cdot (\bold{p} -\bold{A})\psi|^2dx +\varepsilon\, C(\delta, \varepsilon, V, \bold{B}) \int_{\Bbb{R}^3}|\psi|^2dx. $$ Let $j\to \infty$. By Fatou's Lemma, we have $$ \int_{\Bbb{R}^3} |V|\, |\psi|^2dx \le \varepsilon\int_{\Bbb{R}^3} |\sigma\cdot (\bold{p} -\bold{A})\psi|^2dx +\varepsilon\, C(\delta, \varepsilon, V, \bold{B}) \int_{\Bbb{R}^3}|\psi|^2dx $$ for $\psi\in \text{Domain} (\Bbb{D})$. It then follows from the KLMN Theorem \cite{14, p.167} that $\Bbb{P}_0+V$ can be extended to the unique self-adjoint operator associated with its quadratic form. Furthermore, Domain$(\Bbb{D})$ is a form core for $\Bbb{P}_0+V$. \endremark To prove Theorem 1.1, we will compare the Pauli operator $\pp_0$ with the magnetic Schr\"odinger operator $\hh_0 =(\bold{p}-\bold{A})^2$ in an appropriate scale, as in \cite{2,8,19,20}. This is possible since $\pp_0=\hh_0-\sigma\cdot\bb$. To this end, our first step is to localize the operators to cubes $Q$ over which the $L^p$ average of $|\bb|$ is small compare to $\ell(Q)^{-2}$ (the localization error in the kinetic energy), by a Calder\'on--Zygmund decomposition. More precisely, we divide $\Bbb{R}^3$ into a grid of disjoint cubes $\{ Q_j\}$ where each $Q_j$ is a maximal dyadic cube such that $$ \left\{ \int_{12Q_j}|\bb(x)|^pdx\right\}^{1/p} \le \frac{\varepsilon}{[\ell (Q_j)]^{2-\frac{3}{p}}}. \tag 1.15 $$ Here $\ell(Q_j)=\ell_j$ is the side length of $Q_j$, and $12Q_j$ denotes the cube which has the same center as $Q_j$ and side length $12 \ell(Q_j)$. It can be proved that $\ell_j\approx \ell_k$ if $4Q_j\cap 4Q_k \neq \emptyset$. Using this property, we construct a partition of unity for $\Bbb{R}^3$: $\sum_j \phi_j^2(x)\equiv 1 $, with $\phi_j\in C_0^\infty (2Q_j,\Bbb{R})$. Next we apply the Birman-Schwinger principle which reduces the problem to the estimate of singular values of $|V|^{1/2}(\pp_0+\lambda)^{-1/2}$ for $\lambda>0$. We then use the resolvent identity to compare $\phi_j (\pp_0 +\lambda)^{-1}$ with $(\pp_0+\Phi +\frac{1}{\ell_j^2} +\lambda)^{-1}\phi_j$ where $\Phi(x)=\sum_j \frac{1}{\ell_j^2}\, \phi_j^2(x) \approx b_p(x) $. We will show that, if $\varepsilon$ in (1.15) is sufficiently small, then $$ \hh_0\le C\left\{ \pp_0+\Phi\right\} \tag 1.16 $$ (Theorem 3.1). With (1.16) at our disposal, we are able to estimate the contribution from $(\pp_0+\Phi+\frac{1}{\ell_j^2} +\lambda)^{-1}\phi_j$. This leads to the first term in Theorem 1.1. Finally, to deal with the error term $ \phi_j (\pp_0 +\lambda)^{-1}- (\pp_0+\Phi+\frac{1}{\ell_j^2}+ \lambda)^{-1}\phi_j $ which can be written as $$ (\pp_0+\Phi+\frac{1}{\ell_j^2}+ \lambda)^{-1} \big\{ \big(\Phi+\frac{1}{\ell_j^2}\big) \phi_j+[\pp_0,\phi_j]\big\} (\pp_0 +\lambda)^{-1}, $$ we use the resolvent identity again to compare $\psi_j (\pp_0+\Phi +\frac{1}{\ell_j^2})^{-1}$ with $(\hh_0+\frac{1}{\ell_j^2})^{-1}\psi_j$ where $\psi_j$ is a bump function such that $\psi_j\phi_j\equiv \phi_j$. The desired result follows from certain regularity estimates for the operator $\psi_j(\pp_0+\Phi +\frac{1}{\ell_j^2})^{-1}$ (see Lemma 4.2). A similar approach is used in the case of constant direction fields. Assuming $\bb=(0,0,B(x_1,x_2))$ without the loss of generality, we construct a partition of unity for $\Bbb{R}^2$ associated with $B$. The inequality $$ \| (-\partial_{x_3}^2 +\lambda)^{1/2} (\pp_0 +\lambda)^{-1/2}\|_{L^2\to L^2} \le 1,\ \ \ \ \ \ \ \ \ \ \lambda>0 \tag 1.17 $$ is used to exploit the fact that $\pp_0$ commutes with $\partial_{x_3}$. The paper is organized as follows. In Section 2 we collect some basic facts that will be used concerning the norm in the Neumann-Schatten classes. Section 2 also contains the proof of Proposition 1.1. The partition of unity will be constructed in Section 3. Section 4 is devoted to the proof of Theorem 1.1. Finally we study the case of constant direction fields in Section 5 where Theorem 1.2 is proved. Throughout this paper we will use $\|\psi\|_{L^p}$ to denote the $L^p$ norm of the function $\psi$. $\| T\|_{L^p\to L^q}$ will denote the operator norm, while $\| T\|_p$ is reserved for the norm in the Neumann-Schatten classes (see (2.1)). Finally we will use $C$ to denote constants, which are not necessarily the same at each occurrence, which may depend on $p$. \bigskip\medskip \centerline{\bf 2. Some Preliminaries} Most materials in this section on the Neumann-Schatten classes can be found in \cite {18}. We include them here for the reader's convenience. The proof of Proposition 1.1 will be given at the end of the section. Let $T$ be a compact operator on $L^2(\Bbb{R}^3,\Bbb{C}^2)$. We will use $n(s,T)$ to denote the number of singular values $\{ s_n(T)\}$ (counting multiplicity) of $T$ greater than $s$ where $s>0$. For $p\ge 1$, let $$ \| T\|_p^p\equiv\sum_n |s_n(T)|^p=p\int_0^\infty s^{p-1}n(s,T)ds. \tag 2.1 $$ The functional $\|T\|_p$ defines a norm on the Neumann-Schatten class $\Cal{S}_p$ which consists of compact operators $T$ with $\|T\|_p<\infty$. The following facts will be useful to us. $$ \align n(s^2, T^*T)&=n(s,T), \tag 2.2\\ \|T\|_p^2&=\|T^* T\|_{p/2},\tag 2.3\\ n(s,T)&\le \frac{\| T\|_p^p}{s^p}.\tag 2.4 \endalign $$ We will also need $$ \align &n(s_1+s_2,T_1+T_2)\le n(s_1,T_1)+n(s_2+T_2), \tag 2.5\\ &\|T_1 T_2\|_p\le \|T_1\|_p \|T_2\|_{L^2\to L^2},\tag 2.6\\ & \|T_1 T_2\|_p\le \| T_1\|_r \|T_2\|_s \ \ \ \ \ \ \ \ \text {if } \ \ \ \frac1p =\frac1r +\frac 1s,\ \ r, \, s\ge 1\tag 2.7 \endalign $$ where $\|T\|_{L^2\to L^2}$ denotes the operator norm of $T$. The proof of the following lemma may be found in \cite{18}. \proclaim{\bf Lemma 2.1} If $f,\ g\in L^p(\Bbb{R}^3)$ with $2\le p<\infty$. Then $$ \| f(x)g(-i\nabla)\|_p\le C_p \| f\|_{L^p} \| g\|_{L^p}. $$ \endproclaim It follows from Lemma 2.1 that, if $p\ge 2$, $\alpha > \frac{3}{2p}$, and $\lambda >0$, $$ \| V (-\Delta +\lambda )^{-\alpha}\|_p \le C_{p, \alpha} \lambda^{\frac{3}{2p}-\alpha}\| V\|_{L^p}. \tag 2.8 $$ Also, by Lemma 2.1, if $\alpha >1/2$ and $0 < \lambda_1\le \lambda_2$, $$ \| V(-\partial _{x_3}^2+\lambda_1)^{-1/2} (-\Delta+\lambda_2)^{-\alpha}\|_2 \le C_\alpha \lambda_1^{-\frac14}\lambda_2^{-\alpha +\frac12}\|V\|_{L^2}. \tag 2.9 $$ Thus, by the diamagnetic inequality \cite{9}, $$ |\big( \hh_0+\lambda\big)^{-\alpha}\psi| \le (-\Delta +\lambda)^{-\alpha}|\psi|, \tag 2.10 $$ where $\hh_0 =(\bold{p}-\bold{A}(x))^2$, and Theorem 2.13 in \cite{18, p.36}, we have $$ \align &\|V( \hh_0+\lambda)^{-\alpha}\|_2 \le C_\alpha \lambda^{\frac34-\alpha} \|V\|_{L^2} \ \ \ \ \text{ for } \alpha >3/4, \tag 2.11\\ & \|V( \hh_0+\lambda)^{-\alpha}\|_4 \le C_\alpha \lambda^{\frac38-\alpha} \|V\|_{L^4} \ \ \ \ \text{ for } \alpha >3/8, \tag 2.12 \endalign $$ and for $\alpha>1/2$ and $0<\lambda_1\le \lambda_2$, $$ \| V(-\partial _{x_3}^2+\lambda_1)^{-1/2} (\hh_0+\lambda_2)^{-\alpha}\|_2 \le C_\alpha \lambda_1^{-\frac14}\lambda_2^{-\alpha +\frac12}\|V\|_{L^2}. \tag 2.13 $$ The following two lemmas concern the operator norm of $V(-\Delta +\lambda)^{-\alpha}$ on $L^2(\Bbb{R}^3, \Bbb{C}^2)$. \proclaim{\bf Lemma 2.2} Let $\alpha >0$ and $\lambda >0$. Then $$ \align & \| V (-\Delta +\lambda)^{-\alpha}\|_{L^2\to L^2} \le C_\alpha \| V\|_{L^\frac{3}{2\alpha}} \ \ \ \ \ \ \ \ \ \text{ if } \ \ \alpha <3/4,\tag 2.14\\ & \| V (-\Delta +\lambda)^{-\alpha}\|_{L^2\to L^2} \le C_\alpha \lambda^{\frac{3}{4}-\alpha} \| V\|_{L^2}\ \ \ \ \text { if } \ \ \alpha >3/4.\tag 2.15 \endalign $$ \endproclaim \demo{\bf Proof} Note that, if $\alpha <3/4$, $$ \|V(-\Delta+\lambda)^{-\alpha}\psi\|_{L^2} \le \|V\|_{L^{p_1}}\| (-\Delta +\lambda )^{-\alpha}\psi \|_{L^{p_2}} \le C_\alpha \|V\|_{L^{p_1}}\|\psi\|_{L^2} $$ where $\frac{1}{p_1}+\frac{1}{p_2}=\frac12$, $\frac{1}{p_2}=\frac12 -\frac{2\alpha}{3}$, and we have used H\"older's inequality and the well-known fractional integral estimate \cite{21}. (2.14) is proved since $p_1=3/(2\alpha)$. Now suppose $\alpha >3/4$. Using the Fourier transform, we have $$ \aligned \|(-\Delta& +\lambda)^{-\alpha}\psi\|_{L^\infty} \le C\int_{\Bbb{R}^3} (|\xi|^2 +\lambda)^{-\alpha} |\hat{\psi}(\xi)|d\xi\\ &\le C \|\psi\|_{L^2} \left\{\int_{\Bbb{R}^3} (|\xi|^2 +\lambda)^{-2\alpha} d\xi\right\}^{1/2} \le C_\alpha \lambda^{\frac34 -\alpha}\|\psi\|_{L^2}. \endaligned \tag 2.16 $$ It follows that $$ \|V(-\Delta+\lambda)^{-\alpha}\psi \|_{L^2} \le \|V\|_{L^2} \| (-\Delta +\lambda)^{-\alpha}\psi \|_{L^\infty} \le C_\alpha \lambda^{\frac34 -\alpha} \|V\|_{L^2} \|\psi\|_{L^2}. $$ \enddemo \proclaim{\bf Lemma 2.3} Let $0<\alpha <1/2$ and $\lambda >0$. Suppose that $V$ depends only on the first two variables. Then $$ \| V (-\Delta +\lambda )^{-\alpha}\|_{L^2\to L^2} \le C_\alpha \| V\|_{L^{1/\alpha}(\Bbb{R}^2)}. $$ \endproclaim \demo{\bf Proof} By the fractional integral estimates \cite{21}, if $\psi\in C^\infty (\Bbb{R}^3,\Bbb{C}^2)$ and decays at $\infty$, $$ \|\psi\|_{L^q(\Bbb{R}^2)} \le C_\alpha \| (-\Delta_2 +\lambda )^\alpha\psi\|_{L^2(\Bbb{R}^2)} $$ where $\frac1q=\frac12 -\alpha$ and $\Delta_2 =\partial_{x_1}^2 +\partial_{x_2}^2$ denotes the Laplacian in $\Bbb{R}^2$. It then follows from H\"older's inequality that $$ \align \|V\psi\|_{L^2(\Bbb{R}^2)} &\le \|V\|_{L^p(\Bbb{R}^2)} \|\psi\|_{L^q(\Bbb{R}^2)}\\ &\le C_\alpha \|V\|_{L^p(\Bbb{R}^2)} \| (-\Delta_2 +\lambda )^\alpha\psi\|_{L^2(\Bbb{R}^2)} \endalign $$ where $\frac1p +\frac1q =\frac12$. Hence $$ \align \int_{\Bbb{R}^3} |V|^2 |\psi|^2dx &\le C_\alpha^2 \|V\|_{L^p(\Bbb{R}^2)}^2 \int_{\Bbb{R}^3} |(-\Delta_2 +\lambda )^\alpha\psi|^2dx\\ &\le C_\alpha^2 \|V\|_{L^p(\Bbb{R}^2)}^2 \int_{\Bbb{R}^3} |(-\Delta +\lambda )^\alpha\psi|^2dx \endalign $$ where the second inequality can be verified through the Fourier transform. Thus $$ \| V\psi \|_{L^2} \le C_\alpha \|V\|_{L^p(\Bbb{R}^2)} \|(-\Delta +\lambda )^\alpha\psi\|_{L^2}. $$ The lemma now follows easily since $\frac1p =\frac12 -\frac1q=\alpha$. \enddemo We end this section with the \demo{\bf Proof of Proposition 1.1} It follows from (1.10) that $b_p(x)\le \{ M(|\bb|^p)(x)\}^{1/p}$ where $M$ is the Hardy-Littlewood maximal operator. This gives the desired estimate in the case $q>p$ since $M$ is bounded on $L^s$ for $s>1$ \cite{21}. For the case $q=p$, we appeal to an argument in \cite{2}. Let $\widetilde{b}_p(x)=1/[r_p(x)]^2$ where $r_p(x)$ is defined by $$ \frac{1}{[r_p(x)]^{2p-3}} =\int_{\Bbb{R}^3} \varphi \left(\frac{x-y}{r_p(x)}\right)|\bb(y)|^pdy \tag 2.17 $$ and $\varphi(z)=[1+\frac12 |z|^2]^{-2}$. Since $b_p(x)\le C\, \widetilde{b}_p(x)$ (see Remark 1.4 for $p=2$), it suffices to show that $$ \int_{\Bbb{R}^3} | \widetilde{b}_p(x)|^pdx \le C\int_{\Bbb{R}^3} |\bb(x)|^pdx. \tag 2.18 $$ Using (2.17) and Fubini's Theorem, one may reduce (2.18) to $$ \int_{\Bbb{R}^3} \varphi \left(\frac{x-y}{r_p(x)}\right) \frac{dx}{[r_p(x)]^3} \le C<\infty. \tag 2.19 $$ The proof of (2.19) relies on the following estimate $$ r_p(y)g_-\left(\frac{|x-y|}{r_p(y)}\right) \le r_p(x) \le r_p(y)g_+\left(\frac{|x-y|}{r_p(y)}\right) \tag 2.20 $$ where $g_+(s)$ and $g_-(s)$ are solutions of the equation $$ s^2=2(t^{p+\frac12}-1)(1-t^{\frac32-p}) \tag 2.21 $$ on $(1,\infty)$ and $(0,1)$ respectively. Since $p>3/2$, it is easy to see that $g_+(s)$, $g_-(s)$ are well defined. We omit the proof of (2.20), which is similar to that of Lemma 2 in \cite{2, p.569}. It is not hard to see that $g_+(s)\approx s^{\frac{4}{2p+1}}$ if $s\ge 1/2$, and $g_-(s)\ge c_0>0$ if $s<1/2$. This, together with (2.20), implies that $$ \alignedat2 &r_p(x)\le C\, [r_p(y)]^{1-\frac{4}{2p+1}} |x-y|^{\frac{4}{2p+1}} \ \ & \text{ if }\ \ |x-y|\ge \frac12 r_p(y),\\ &r_p(x)\ge c \, r_p(y) & \text{ if }\ \ |x-y|<\frac12 r_p(y). \endalignedat $$ It then follows that $$ \aligned &\int_{\Bbb{R}^3} \varphi \left(\frac{x-y}{r_p(x)}\right) \frac{dx}{[r_p(x)]^3} \le \int_{\Bbb{R}^3} \frac{r_p(x)\, dx} {\{ [r_p(x)]^2 +\frac12 |x-y|^2\}^2}\\ &\le 4 \int_{|x-y|\ge \frac12 r_p(y)} \frac{r_p(x)\, dx} {|x-y|^4} + \int_{|x-y|< \frac12 r_p(y)} \frac{dx}{[r_p(x)]^3}\\ &\le C \, \big[r_p(y)\big]^{\frac{2p-3}{2p+1}} \int_{|x-y|\ge \frac12 r_p(y)} \frac{ dx} {|x-y|^{3+\frac{2p-3}{2p+1}}} +\frac{C}{[r_p(y)]^3} \int_{|x-y|< \frac12 r_p(y)} dx\\ &\le C. \endaligned $$ The proof is finished. \enddemo \bigskip \centerline{\bf 3. A Partition of Unity} Throughout this section we fix $p>3/2$. We will assume that $|\bb|\in L^p_{loc}(\Bbb{R}^3)$ and $|\bb|\not\equiv 0$. Let $\varepsilon\in (0,1)$ be a small constant to be determined later. Let $\Cal{A}$ be the set of all dyadic cubes in $\Bbb{R}^3$ such that $$ \left(\int_{12 Q}|\bb(x)|^pdx\right)^{1/p}\le \frac{\varepsilon}{[\ell(Q)]^{2-\frac{3}{p}}} \tag 3.1 $$ where $\ell(Q)$ denotes the side length of $Q$. We say that $Q$ is a maximal element of $\Cal{A}$ if $Q\in\Cal{A}$ and $Q$ is not properly contained in any other cube in $\Cal{A}$. Let $\Cal{B}$ denote the set of all maximal elements of $\Cal{A}$. Clearly, by definition, the interiors of the cubes in $\Cal{B}$ are disjoint. \proclaim{\bf Lemma 3.1} Let $\Cal{B}=\left\{ Q_j,\ j=1,2,\dots\right\}$. Then, $$ \Bbb{R}^3=\bigcup_j Q_j. $$ \endproclaim \demo{\bf Proof} It suffices to show that any point in $\Bbb{R}^3$ belongs to some cube $Q_j$ in $\Cal{B}$. To this end, fix $x\in \Bbb{R}^3$, let $\{ Q_{\alpha_k}\}_{k=-\infty}^\infty$ be an increasing sequence of dyadic cubes such that $x\in Q_{\alpha_k}$ for all $k$, $\ell(Q_{\alpha_k})\to 0$ as $k\to -\infty$, and $\ell(Q_{\alpha_k})\to\infty$ as $k\to\infty$. Note that, as $\ell(Q)\to 0$, the l.h.s. of (3.1) goes to $0$ while the r.h.s. goes to $\infty$. This implies that $Q_{\alpha_k}\in\Cal{A}$ if $\ell(Q_{\alpha_k})$ is small. Also, as $\ell(Q_{\alpha_k})\to\infty$, the l.h.s. of (3.1) goes to $\|\bb\|_{L^p} >0$ and the r.h.s. goes to 0. Thus, $Q_{\alpha_k}\notin\Cal{A}$ if $\ell(Q_{\alpha_k})$ is large. Hence there must be a maximal element $Q_j$ in $\Cal{A}$ (which may not be in $\{ Q_{\alpha_k}\}$) such that $x\in Q_j$. \enddemo \proclaim{\bf Lemma 3.2} Let $Q_j$, $Q_k \in \Cal{B}$. Suppose that $4Q_j\cap 4Q_k\neq\emptyset$. Then $$ \frac{1}{2}\ell(Q_k) \le \ell(Q_j) \le 2\ell(Q_k). $$ \endproclaim \demo{\bf Proof} We adapt an argument found in \cite {8}. Suppose $\ell(Q_j)>2\ell (Q_k)$. Since $\ell(Q_j)$ and $\ell(Q_k)$ are powers of $2$, we have $ \ell(Q_j)\ge 4\ell(Q_k). $ Let $Q_k^+$ be the dyadic parent of $Q_k$, i.e., $Q_k$ can be obtained by bisecting $Q_k^+$. Since $Q_k$ is a maximal element in $\Cal{A}$, $Q_k^+\notin \Cal{A}$. It follows that $$ \left( \int_{12Q_k^+}|\bb|^pdx\right)^{1/p} >\varepsilon \left[\frac{1}{\ell(Q_k^+)}\right]^{2-\frac{3}{p}} =\varepsilon \left[\frac{1}{ 2\ell(Q_k)}\right]^{2-\frac{3}{p}} \ge \varepsilon \left[\frac{2}{\ell(Q_j)}\right]^{2-\frac{3}{p}} \tag 3.2 $$ where we have used $p>3/2$ in the last inequality. We claim that $$ 12Q_k^+\subset 12 Q_j. \tag 3.3 $$ This, together with (3.2), would imply that $$ \left(\int_{12Q_j}|\bb|^pdx\right)^{1/p} \ge \left(\int_{12Q_k^+}|\bb|^pdx\right)^{1/p} > \varepsilon \left[\frac{2}{\ell(Q_j)}\right]^{2-\frac{3}{p}} $$ which contradicts with the assumption that $Q_j\in \Cal{A}$. To see (3.3), let $x_j$ and $x_k$ be the centers of $Q_j$ and $Q_k$ respectively. Since $4Q_j\cap 4Q_k\neq \emptyset$, there exists $z$ such that $$ |z-x_j|_*\le 2\ell(Q_j)\ \ \ \text{ and } \ \ \ |z-x_k|_*\le 2\ell(Q_k) $$ where $|\cdot|_*$ is the norm in $\Bbb{R}^3$ defined by $$ |x|_*=|(a,b,c)|_*=\max \{ |a|,|b|,|c|\}. $$ Thus, $$ |x_j-x_k|_*\le 2\ell(Q_j)+2\ell(Q_k). $$ Now suppose $y\in 12 Q_k^+$. Then $|y-x_k|_* \le |y-x_k^+|_*+ |x_k^+-x_k|_* < 13 \ell(Q_k)$ where $x_k^+\in Q_k$ is the center of $Q_k^+$. It follows that $$ |y-x_j|_*\le |y-x_k|_*+|x_k-x_j|_* < 15\ell(Q_k) +2\ell(Q_j) < 6\ell(Q_j). $$ Hence, $y\in 12 Q_j$. (3.3) is then proved. \enddemo \remark{\bf Remark 3.1} It follows easily from Lemma 3.2 that $\{ 4Q_j: Q_j\in \Cal{B}\}$ has the finite intersection property: $\sum_j \chi_{4Q_j}\le C$ where $C$ is an absolute constant. \endremark We are now in a position to construct the partition of unity associated with the field strength $|\bb|$. \proclaim{\bf Lemma 3.3} There exists a sequence of functions $\{ \phi_j\}$ such that (i)\ \ $\phi_j\in C_0^\infty(2Q_j)$ and $0\le \phi_j\le 1$, (ii)\ \ $|\nabla^\alpha \phi_j(x)|\le c_\alpha/\ell_j^{|\alpha|}$ where $\ell_j=\ell(Q_j)$, (iii)\ \ $\sum_j\phi_j^2\equiv 1$ in $\Bbb{R}^3$. \endproclaim \demo{\bf Proof} Choose $\eta\in C_0^\infty(Q(0,2))$ such that $\eta\equiv1$ in $Q(0,1)$ and $0\le \eta \le 1$. Let $\Cal{B}=\{ Q_j\}_{j=1}^\infty$ and $Q_j=Q(x_j,\ell_j)$. We define $$ \eta_j(x)=\eta\left(\frac{x-x_j}{\ell_j}\right). $$ and $\varphi (x)=\sum_j \eta_j^2(x)$. Clearly, by Lemma 3.1 and Remark 3.1, $1\le \varphi \le C$ where $C$ is an absolute constant, and $\varphi\in C^\infty (\Bbb{R}^3)$. Finally let $\phi_j(x)=\eta_j(x)/\sqrt{\varphi (x)}$. It is easy to check that $\phi_j$ satisfies (i)-(iii). We omit the details. \enddemo Recall that $\ell_j=\ell(Q_j)$ and $Q_j$ is a maximal cube. Define $$ \Phi=\sum_j\frac{1}{\ell_j^2}\, \phi_j^2.\tag 3.4 $$ We will show in Section 4 that $\Phi(x)\approx b_p(x)$ (see Lemma 4.5). \proclaim{\bf Theorem 3.1} There exist constants $C>0$ and $\varepsilon_0>0$ such that, if $0<\varepsilon <\varepsilon_0$, we have $$ \hh_0\le C\big\{ \pp_0+\Phi\big\}. $$ \endproclaim \demo{\bf Proof} We first show that, if $\varepsilon$ in (3.1) is small, then $$ \int_{\Bbb{R}^3} |\, |\bb|^{1/2}\phi_j\psi|^2dx \le C\int_{\Bbb{R}^3} |\sigma\cdot (\bold{p}-\bold{A})(\phi_j\psi)|^2dx\tag 3.5 $$ for any $\psi\in C^\infty_0(\Bbb{R}^3,\Bbb{C}^2)$. To this end, we note that, by H\"older's inequality $$ \left\{\int_{2Q_j}|\bb|^{3/2}dx\right\}^{2/3} \le |2Q_j|^{\frac23-\frac1p} \left\{\int_{2Q_j}|\bb|^p dx\right\}^{1/p} \le 4\varepsilon.\tag 3.6 $$ It then follows from the Sobolev embedding and the inequality $|\nabla |\psi|\,|\le |(\bold{p}-\bold{A})\psi|$ that $$ \aligned \bigg\{ \int_{\Bbb{R}^3} |\, &|\bb|^{1/2}\phi_j\psi|^2dx\bigg\}^{1/2} \le \left\{\int_{2Q_j}|\bb|^{3/2}dx\right\}^{1/3} \left\{\int_{\Bbb{R}^3}|\phi_j\psi|^6dx\right\}^{1/6}\\ &\le C\varepsilon^{1/2} \left\{\int_{\Bbb{R}^3}|\nabla|\phi_j\psi|\,|^2dx\right\}^{1/2}\\ &\le C\varepsilon^{1/2} \left\{\int_{\Bbb{R}^3}|(\bold{p}-\bold{A})(\phi_j\psi)|^2 dx\right\}^{1/2}\\ &\le C\varepsilon^{1/2} \left\{\int_{\Bbb{R}^3}|\sigma\cdot (\bold{p}-\bold{A})(\phi_j\psi)|^2 dx\right\}^{1/2} +C\varepsilon^{1/2} \left\{ \int_{\Bbb{R}^3} |\, |\bb|^{1/2}\phi_j\psi|^2dx\right\}^{1/2} \endaligned $$ where we have used the fact that $\pp_0 =\hh_0 -\sigma\cdot \bb$. This gives (3.5) if $C\varepsilon^{1/2}<1/2$. Next we use Lemma 3.3 and (3.5) to obtain $$ \aligned &\int_{\Bbb{R}^3} |(\bold{p}-\bold{A})\psi|^2dx =\sum_j\int_{\Bbb{R}^3} |\phi_j(\bold{p}-\bold{A})\psi|^2dx\\ & \le 2\sum_j \int_{\Bbb{R}^3} |(\bold{p}-\bold{A})(\phi_j\psi)|^2dx +2\sum_j \int_{\Bbb{R}^3}|\nabla\phi_j|^2|\psi|^2dx\\ &\le 2\sum_j \int_{\Bbb{R}^3} |\sigma\cdot (\bold{p}-\bold{A})(\phi_j\psi)|^2dx +2\sum_j\int_{\Bbb{R}^3} |\, |\bb|^{1/2}\phi_j\psi|^2dx + 2\sum_j \int_{\Bbb{R}^3}|\nabla\phi_j|^2|\psi|^2dx\\ &\le C\sum_j \int_{\Bbb{R}^3} |\sigma\cdot (\bold{p}-\bold{A})(\phi_j\psi)|^2dx +2\sum_j \int_{\Bbb{R}^3}|\nabla\phi_j|^2|\psi|^2dx\\ &\le C\int_{\Bbb{R}^3} |\sigma\cdot (\bold{p}-\bold{A})\psi|^2dx +C\sum_j \frac{1}{\ell_j^2} \int_{2Q_j}|\psi|^2dx. \endaligned $$ Clearly, the theorem follows if we have $$ \sum_j\frac{1}{\ell_j^2}\, \chi_{2Q_j}\le C\, \Phi. \tag 3.7 $$ We claim that (3.7) is an easy consequence of Lemma 3.2. Indeed, if $x\in Q_k$, the l.h.s. of (3.7) is bounded by $$ \sum_{2Q_j\cap Q_k\neq\emptyset} \frac{1}{\ell_j^2} \le \frac{C}{\ell_k^2}\cdot \# \{ Q_j: \ 2Q_j\cap Q_k\neq\emptyset\} \le\frac{C}{\ell_k^2}\le C\, \Phi(x). $$ The proof is now complete. \enddemo It follows easily from Theorem 3.1 that, if $0<\lambda_1\le\lambda_2$, $$ \| (\hh_0+\lambda_1)^{1/2} (\pp_0+\Phi +\lambda_2)^{-1/2}\|_{L^2\to L^2} \le C. \tag 3.8 $$ The following lemma will be used in the next section. \proclaim{\bf Lemma 3.4} There exists a constant $C>0$ such that, for $\lambda>0$, $$ \| (\pp_0+\Phi+\frac{1}{\ell_j^2}+\lambda)^{-1/2}\phi_j (\pp_0+\lambda)^{1/2}\|_{L^2\to L^2}\le C. $$ \endproclaim \demo{\bf Proof} By duality, it suffices to show that $$ \|(\pp_0+\lambda)^{1/2}\phi_j (\pp_0+\Phi+\frac{1}{\ell_j^2}+\lambda)^{-1/2}\|_{L^2\to L^2} \le C. \tag 3.9 $$ To this end, we note that, for any $\psi\in C^\infty_0(\Bbb{R}^3,\Bbb{C}^2)$, $$ \aligned &\int_{\Bbb{R}^3} |(\pp_0+\lambda)^{1/2}\phi_j (\pp_0+\Phi+\frac{1}{\ell_j^2}+\lambda)^{-1/2}\psi|^2dx\\ & = \int_{\Bbb{R}^3}|\sigma\cdot (\bold{p}-\bold{A})\phi_j (\pp_0+\Phi+\frac{1}{\ell_j^2}+\lambda)^{-1/2}\psi|^2dx +\lambda \int_{\Bbb{R}^3} |\phi_j(\pp_0+\Phi+\frac{1}{\ell_j^2}+\lambda)^{-1/2}\psi|^2dx\\ & \le 4\int_{\Bbb{R}^3}|\psi|^2 +2\int_{\Bbb{R}^3} |\nabla \phi_j|^2 |(\pp_0+\Phi+\frac{1}{\ell_j^2}+\lambda)^{-1/2}\psi|^2dx\\ & \le C\int_{\Bbb{R}^3} |\psi|^2dx \endaligned $$ where we have used $|\nabla \phi_j|\le C/\ell_j$ and $$ \align &\| \sigma \cdot (\bold{p}-\bold{A})(\pp_0+\Phi+\lambda)^{-1/2}\|_{L^2 \to L^2}\le 1,\tag 3.10\\ &\|(\pp_0+\Phi+\lambda)^{-1/2}\|_{L^2 \to L^2}\le \frac{1}{\sqrt{\lambda }}. \tag 3.11 \endalign $$ The proof is finished. \enddemo \bigskip \centerline{\bf 4. The Proof of Theorem 1.1} For $\lambda >0$, we denote by $N(\lambda, \Bbb{P}_0+V)$ the number of eigenvalues (counting multiplicity) of $\Bbb{P}_0+V$ smaller than $-\lambda$. Since $V\ge -\lambda -|V+\lambda|_-$, we have $$ N(2\lambda, \Bbb{P}_0+V)\le N(\lambda, \Bbb{P}_0-|V+\lambda|_-). $$ It then follows from the Birman-Schwinger principle and (2.2) that $$ N(2\lambda, \Bbb{P}_0+V) \le n(1,Y(\Bbb{P}_0+\lambda)^{-1}Y ) = n(1,Y(\pp_0+\lambda)^{-1/2}) \tag 4.1 $$ where $Y=|V+\lambda|^{1/2}_-$. As in \cite {19,20}, our approach will be based on the following resolvent identity: $$ \varphi H_1^{-1} =H_2^{-1}\varphi + H_2^{-1} \big\{ (H_2-H_1)\varphi +[H_1,\varphi]\big\} H_1^{-1} \tag 4.2 $$ where $\varphi\in C^\infty_0(\Bbb{R}^3,\Bbb{R})$, $H_1,\, H_2$ are two operators, and $[H_1,\varphi]=H_1\varphi-\varphi H_1$ denotes the commutator between $H_1$ and $\varphi$. Using (4.2), we may write $$ \aligned Y(\pp_0+\lambda)^{-1/2}&=\sum_j Y\phi_j(\pp_0+\Phi +\frac{1}{\ell_j^2}+\lambda)^{-1} \phi_j(\pp_0+\lambda)^{1/2}\\ &\ \ \ \ \ \ +\sum_j Y\phi_j(\pp_0+\Phi +\frac{1}{\ell_j^2}+\lambda)^{-1} (\Phi +\frac{1}{\ell_j^2})\phi_j (\pp_0+\lambda)^{-1/2}\\ &\ \ \ \ \ \ +\sum_jY\phi_j(\pp_0+\Phi +\frac{1}{\ell_j^2}+\lambda)^{-1} [\pp_0,\phi_j](\pp_0+\lambda)^{-1/2}\\ &=I_1 +I_2 +I_3 \endaligned $$ where $\Phi$ is the function defined by (3.4). Thus, by (2.5), $$ n(1, Y(\pp_0+\lambda)^{-1/2}) \le n(1/3, I_1)+n(1/3,I_2)+n(1/3, I_3). \tag 4.3 $$ We begin with the estimate of $n(1/3, I_1)$. \proclaim{\bf Lemma 4.1} There exists a constant $C>0$ such that, for any $\lambda>0$, $$ n(1/3, I_1)\le \frac{C}{\sqrt{\lambda}}\int_{\Bbb{R}^3}|Y|^4 dx. $$ \endproclaim \demo{\bf Proof} First observe that, by (2.4), $$ n(1/3, I_1)\le C\|\sum_j Y\phi_j(\pp_0+\Phi +\frac{1}{\ell_j^2}+\lambda)^{-1} \phi_j(\pp_0+\lambda)^{1/2}\|_4^4. \tag 4.4 $$ Let $$ I_{1j}=Y\phi_j(\pp_0+\Phi +\frac{1}{\ell_j^2}+\lambda)^{-1} \phi_j(\pp_0+\lambda)^{1/2}. \tag 4.5 $$ It follows from (2.6) and Lemma 3.4 that $$ \aligned \|I_{1j}\|_4 &\le \|Y\phi_j(\pp_0+\Phi +\frac{1}{\ell_j^2}+\lambda)^{-1/2} \|_4 \|(\pp_0+\Phi +\frac{1}{\ell_j^2}+\lambda)^{-1/2} \phi_j(\pp_0+\lambda)^{1/2}\|_{L^2\to L^2}\\ &\le C\|Y\phi_j(\pp_0+\Phi +\frac{1}{\ell_j^2}+\lambda)^{-1/2}\|_4\\ &\le C\|Y\phi_j (\hh_0+\lambda)^{-1/2}\|_4 \| (\hh_0+\lambda)^{1/2} (\pp_0+\Phi +\frac{1}{\ell_j^2}+\lambda)^{-1/2}\|_{L^2\to L^2}\\ &\le C\|Y\phi_j (\hh_0+\lambda)^{-1/2}\|_4 \endaligned $$ where we also used (3.8) in the last inequality. We now apply (2.12) with $\alpha =1/2$ to obtain $$ \|I_{1j}\|_4 \le C\lambda^{-1/8}\|Y\phi_j\|_{L^4}. \tag 4.6 $$ Next we note that, by (2.3), $$ \|\sum_jI_{1j}\|_4^4 =\|\sum_j\sum_k\sum_m\sum_n I_{1j} I_{1k}^* I_{1m}I_{1n}^*\|_1 \le \sum_j\sum_k\sum_m\sum_n \|I_{1j} I_{1k}^* I_{1m}I_{1n}^*\|_1. $$ Since $\phi_j$ is supported in $2Q_j$ and $\phi_j (\pp_0+\lambda)^{1/2} (\pp_0+\lambda)^{1/2}\phi_k = \phi_j (\pp_0+\lambda)\phi_k=0$ unless $2Q_j\cap 2Q_k \neq \emptyset$, it is not very hard to see that $$ I_{1j} I_{1k}^* I_{1m}I_{1n}^*=0 $$ unless $$ 2Q_j\cap2Q_k\neq\emptyset,\ \ 2Q_k\cap 2Q_m\neq\emptyset,\ \ \text {and } \ \ 2Q_m\cap 2Q_n\neq \emptyset . \tag 4.7 $$ On the other hand, by Lemma 3.2, if we fix any one of indices $j,k,m,n$, the number of remaining indices which satisfy (4.7) is bounded by an absolute constant. Thus, by (2.7), $$ \aligned n(1/3,I_1) &\le C\|\sum_jI_{1j}\|_4^4 \le C\sum \|I_{1j}\|_4 \|I_{1k}\|_4\|I_{1m}\|_4\|I_{1n}\|_4\\ &\le C\sum \big\{ \|I_{1j}\|_4^4+ \|I_{1k}\|_4^4 +\|I_{1m}\|_4^4 +\|I_{1n}\|_4^4\big\}\\ &\le C\sum_j\|I_{1j}\|^4_4 \endaligned $$ where the second and third sums are over all $(j,k,m,n)$ satisfying (4.7). The lemma now follows from (4.6). \enddemo Using $|Y|\le |V|^{1/2}\chi_{\{ x\in\Bbb{R}^3: V(x) <-\lambda\} }$, we may deduce easily from Lemma 4.1 that $$ \int_0^\infty n(1/3, I_1)\, d\lambda \le C\int_{\Bbb{R}^3} |V(x)|^{5/2}dx. \tag 4.8 $$ Let $\psi_j\in C^\infty_0(3Q_j, \Bbb{R})$ such that $0\le \psi_j\le 1$, $\phi_j\psi_j\equiv \phi_j$ and $ |\nabla^\alpha \psi_j |\le C_\alpha/\ell_j^{|\alpha|}$. We will need the following lemma in the estimates of $n(1/3, I_2)$ and $n(1/3,I_3)$. \proclaim{\bf Lemma 4.2} There exist constants $C>0$ and $\varepsilon_0>0$ such that, if $0<\varepsilon<\varepsilon_0$, $$ \align &\| \, [\pp_0,\psi_j] (\pp_0+\Phi +\frac{1}{\ell_j^2})^{-1} \|_{L^2\to L^2} \le C, \tag 4.9 \\ &\| \psi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}\|_{L^2\to L^\infty} \le C \ell_j^{\frac12}, \tag 4.10 \\ & \| (\hh_0 +\frac{1}{\ell_j^2})^{-\frac14 +\delta} (\sigma\cdot \bb)\psi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1} \|_{L^2\to L^2} \le C_\delta \, \ell_j^{\frac12 -2\delta} \tag 4.11 \endalign $$ where $0<\delta <\min\big(1/4, (p-(3/2))/2\big)$. \endproclaim \demo{\bf Proof} To show (4.9), we note that $$ [\pp_0,\psi_j]= \big[\Bbb{D}^*,[\Bbb{D},\psi_j]\big] +[\Bbb{D}^*,\psi_j]\Bbb{D} +[\Bbb{D},\psi_j]\Bbb{D}^* \tag 4.12 $$ where $\pp_0=\Bbb{D}^*\Bbb{D}$ and $\Bbb{D}=\sigma\cdot (\bold{p}-\bold{A})$ ($\pp_0$ is defined as the operator associated with the quadratic form $\|\Bbb{D}\psi\|_{L^2}^2$. We do not need that $\Bbb{D}$ is self-adjoint). Since $$ \align & |\, [\Bbb{D}^*,\psi_j]\psi | +|\, [\Bbb{D},\psi_j]\psi |\le 2|\nabla \psi_j|\, |\psi| \le \frac{C}{\ell_j}|\psi|, \tag 4.13 \\ & |\, \big[\Bbb{D}^*,[\Bbb{D},\psi_j]\big]\psi|\le |\nabla^2\psi_j|\, |\psi | \le \frac{C}{\ell_j^2}|\psi|, \tag 4.14 \endalign $$ we have $$ \|\, [\pp_0,\psi_j] (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}\psi\|_{L^2} \le C\| \psi\|_{L^2}. \tag 4.15 $$ where we also used (3.10)--(3.11), $\Bbb{D}^* =\Bbb{D}$ on the domain of $\Bbb{D}$, and $\|(\pp_0+\Phi +\lambda)^{-1}\|_{L^2\to L^2} \le \lambda^{-1}$. To prove (4.10), we use the resolvent identity (4.2) and $\pp_0 =\hh_0 -\sigma\cdot \bb$ to write $$ \aligned &\psi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}= (\hh_0+\frac{1}{\ell_j^2})^{-1}\psi_j\\ &\ \ \ \ \ \ +(\hh_0+\frac{1}{\ell_j^2})^{-1} \big\{ (\sigma\cdot\bb -\Phi)\psi_j +[\pp_0,\psi_j]\big\} (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}. \endaligned \tag 4.16 $$ Hence, by the diamagnetic inequality (2.10) and (2.16), we get $$ \aligned &|\psi_j(\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}\psi|\\ &\le (-\Delta +\frac{1}{\ell_j^2})^{-1} |\psi_j\psi| +(-\Delta +\frac{1}{\ell_j^2})^{-1}|\bb| \psi_j |(\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}\psi|\\ & \ \ \ \ \ \ \ +(-\Delta +\frac{1}{\ell_j^2})^{-1} \big| (-\Phi\psi_j +[\pp_0, \psi_j]) (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}\psi\big|\\ & \le C\ell_j^{\frac12}\|\psi\|_{L^2} +C\ell_j^{2\delta} \| (-\Delta +\frac{1}{\ell_j^2})^{\delta-\frac14} |\bb|\psi_j |(\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}\psi|\, \|_{L^2} \endaligned \tag 4.17 $$ where $0<\delta <\min\big(1/4, (p-(3/2))/2\big)$ and we also used (4.15) in the second inequality. To deal with the second term in r.h.s. of (4.17), we apply (2.14) to obtain $$ \aligned &\| (-\Delta +\frac{1}{\ell_j^2})^{\delta-\frac14} |\bb|\psi_j |(\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}\psi|\, \|_{L^2}\\ &\le \|(-\Delta +\frac{1}{\ell_j^2})^{\delta-\frac14} |\bb|^{-\delta +\frac14}\chi_{3Q_j}\|_{L^2\to L^2} \|\, |\bb|^{\delta +\frac34}\psi_j |(\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}\psi|\, \|_{L^2}\\ &\le C \| \, |\bb|^{-\delta +\frac14}\|_{L^{\frac{6}{1-4\delta}}(3Q_j)} \|\, |\bb|^{\delta +\frac34}\|_{L^2(3Q_j)} \|\psi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}\psi \|_{L^\infty}\\ &\le C\ell_j^{2-2\delta-\frac{3}{p}} \left\{ \int_{3Q_j} |\bb|^pdx\right\}^{1/p} \| \psi_j(\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}\psi \|_{L^\infty}\\ &\le C\ell_j^{-2\delta}\varepsilon \|\psi_j(\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}\psi\|_{L^\infty} \endaligned \tag 4.18 $$ where we also used H\"older's inequality, $(3/2)+2\delta 3/2$, we may deduce that the function is in $L^q$ for any $q<\infty$. It then follows that it is in $L^\infty$ by the same argument. Finally (4.11) follows from (4.18) and (4.10). \enddemo We now are ready to estimate $n(1/3, I_2)$ and $n(1/3, I_3)$. \proclaim {\bf Lemma 4.3} There exists a constant $C>0$ such that $$ \int_0^\infty n(1/3,I_2)\, d\lambda \le C\int_{\Bbb{R}^3}|V(x)|^{5/2}dx +C\int_{\Bbb{R}^3} \Phi(x)^{3/2} |V(x)|\, dx. $$ \endproclaim \demo{\bf Proof} We first note that, since $\| (\pp_0+\lambda)^{-1/2}\|_{L^2\to L^2} \le 1/\sqrt{\lambda}$, $$ \align n(1/3,I_2) &=n\big (1/3,\sum_j Y\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2} +\lambda)^{-1} (\Phi+\frac{1}{\ell_j^2}) \phi_j (\pp_0 +\lambda)^{-1/2}\big)\\ &\le n\big(\sqrt{\lambda}/3, \sum_j Y\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2} +\lambda)^{-1} (\Phi+\frac{1}{\ell_j^2}) \phi_j\big)\\ &\le n\big(\sqrt{\lambda}/6, \sum_j Y\phi_j \big\{ (\pp_0 +\Phi +\frac{1}{\ell_j^2} +\lambda)^{-1} -(\pp_0 +\Phi +\frac{1}{\ell_j^2} )^{-1}\big\} (\Phi+\frac{1}{\ell_j^2}) \phi_j\big)\\ &\ \ \ \ \ \ + n\big(\sqrt{\lambda}/6, \sum_j Y\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1} (\Phi+\frac{1}{\ell_j^2}) \phi_j\big)\\ & =K_1+K_2. \endalign $$ Using $$ (\pp_0 +\Phi +\frac{1}{\ell_j^2} +\lambda)^{-1} -(\pp_0 +\Phi +\frac{1}{\ell_j^2} )^{-1} =-\lambda (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1} (\pp_0 +\Phi +\frac{1}{\ell_j^2} +\lambda)^{-1} $$ and (2.4), we have $$ \aligned K_1 &=n\big(\sqrt{\lambda}/6, \sum_j Y\phi_j \lambda (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1} (\pp_0 +\Phi +\frac{1}{\ell_j^2} +\lambda)^{-1} (\Phi+\frac{1}{\ell_j^2}) \phi_j\big)\\ &\le C{\lambda^2} \| \sum_j Y\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1} (\pp_0 +\Phi +\frac{1}{\ell_j^2} +\lambda)^{-1} (\Phi+\frac{1}{\ell_j^2}) \phi_j\|_4^4\\ &\le C\lambda^2 \sum_j \| Y\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1} (\pp_0 +\Phi +\frac{1}{\ell_j^2} +\lambda)^{-1} (\Phi+\frac{1}{\ell_j^2}) \phi_j\|_4^4\\ &\le C\lambda^2 \sum_j \| Y\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1} (\pp_0 +\Phi +\frac{1}{\ell_j^2} +\lambda)^{-1} \phi_j\|_4^4\cdot\frac{1}{\ell_j^8} \endaligned $$ where we have used the finite intersection property of the supports of $\phi_j$ as in the proof of Lemma 4.1. We now use (2.6) and the $(L^2, L^2)$ bound of $(\pp_0 +\Phi +\frac{1}{\ell_j^2} +\lambda)^{-\alpha}$ ($\alpha =1, 1/2$) to obtain $$ \align \| Y\phi_j (\pp_0 &+\Phi +\frac{1}{\ell_j^2})^{-1} (\pp_0 +\Phi +\frac{1}{\ell_j^2} +\lambda)^{-1} \phi_j\|_4\\ &\le \| Y\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}\|_4 \cdot\frac{1}{\lambda +\frac{1}{\ell_j^2}}\\ &\le \| Y\phi_j (\pp_0 +\Phi+\frac{1}{\ell_j^2})^{-1/2}\|_4 \cdot\frac{\ell_j}{\lambda +\frac{1}{\ell_j^2}}\\ &\le \|Y\phi_j (\hh_0 +\frac{1}{\ell_j^2})^{-1/2}\|_4 \| (\hh_0 +\frac{1}{\ell_j^2})^{1/2} (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1/2}\|_{L^2\to L^2} \cdot\frac{\ell_j}{\lambda +\frac{1}{\ell_j^2}}\\ &\le C\|Y\phi_j (\hh_0 +\frac{1}{\ell_j^2})^{-1/2}\|_4 \cdot\frac{\ell_j}{\lambda +\frac{1}{\ell_j^2}}\\ &\le C\|Y\phi_j\|_{L^4} \cdot\frac{\ell_j^2}{\lambda^{5/8}} \endalign $$ where we also used (3.8) and (2.12) in the last two inequality. It then follows that $$ K_1\le \frac{C}{\sqrt{\lambda}} \sum_j \int_{\Bbb{R}^3} |Y\phi_j|^4dx \le \frac{C}{\sqrt{\lambda}}\int_{\{x\in \Bbb{R}^3:V(x)< -\lambda\} } |V|^2dx. \tag 4.20 $$ It remains to estimate $K_2$. Since $Y\le |V|^{1/2}$, we have $$ K_2 \le n\big(\sqrt{\lambda}/6, \sum_j |V|^{1/2}\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1} (\Phi+\frac{1}{\ell_j^2})\phi_j\big). $$ It follows that $$ \aligned \int_0^\infty K_2\, d\lambda &\le C\| \sum_j |V|^{1/2}\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1} (\Phi+\frac{1}{\ell_j^2})\phi_j\|_2^2\\ &\le C\sum_j \|\, |V|^{1/2}\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1} \phi_j\|_2^2\cdot\frac{1}{\ell_j^4}. \endaligned \tag 4.21 $$ Using $\phi_j\psi_j\equiv \phi_j$, (4.16), and (4.9), we have $$ \align &\|\, |V|^{1/2}\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1} \phi_j\|_2\\ & \le \|\, |V|^{1/2}\phi_j (\hh_0+\frac{1}{\ell_j^2})^{-1} \phi_j\|_2 +\|\, |V|^{1/2}\phi_j (\hh_0+\frac{1}{\ell_j^2})^{-1} (\sigma\cdot \bb)\psi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}\phi_j\|_2\\ & +\|\, |V|^{1/2}\phi_j (\hh_0+\frac{1}{\ell_j^2})^{-1} (-\Phi\psi_j +[\pp_0,\psi_j]) (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}\phi_j\|_2\\ &\le C \|\, |V|^{1/2}\phi_j (\hh_0+\frac{1}{\ell_j^2})^{-1}\|_2\\ &\ \ \ \ + \|\, |V|^{1/2}\phi_j (\hh_0+\frac{1}{\ell_j^2})^{-\frac34 -\delta} \|_2 \|(\hh_0+\frac{1}{\ell_j^2})^{-\frac14+\delta} (\sigma\cdot \bb)\psi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2})^{-1}\phi_j\|_{L^2\to L^2}\\ &\le C \|\, |V|^{1/2}\phi_j (\hh_0+\frac{1}{\ell_j^2})^{-1}\|_2 +C\ell_j^{\frac12-2\delta} \|\, |V|^{1/2}\phi_j (\hh_0+\frac{1}{\ell_j^2})^{-\frac34 -\delta} \|_2\\ &\le C\ell_j^{\frac12}\|\, |V|^{1/2}\phi_j\|_{L^2} \endalign $$ where we also used (4.11) in the third and (2.11) in the last inequality. Inserting the estimate above into the r.h.s. of (4.21), we obtain $$ \int_0^\infty K_2 \, d\lambda \le C\sum_j \frac{1}{\ell_j^3} \int_{\Bbb{R}^3} |V|\, \phi_j^2\, dx \le C\int_{\Bbb{R}^3} \Phi(x)^{3/2} |V(x)|dx. $$ This, together with (4.20), gives the desired estimate. \enddemo \proclaim{\bf Lemma 4.4} There exists a constant $C>0$ such that $$ \int_0^\infty n(1/3, I_3)\, d\lambda \le C\int_{\Bbb{R}^3} |V(x)|^{5/2}dx +\int_{\Bbb{R}^3} \Phi(x)^{3/2} |V(x)|dx. $$ \endproclaim \demo{\bf Proof} It follows from (2.5) and (4.12) that $$ \aligned &n(1/3, I_3) =n\big(1/3, \sum_j Y\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2}+\lambda)^{-1} [\pp_0,\phi_j](\pp_0+\lambda)^{-1/2}\big)\\ &\le n\big(1/6, \sum_j Y\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2}+\lambda)^{-1} [\Bbb{D}^*, [\Bbb{D},\phi_j]](\pp_0 +\lambda)^{-1/2}\big)\\ &\ \ \ \ \ + n\big(1/6, 2\sum_j Y\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2}+\lambda)^{-1} \left\{ [\Bbb{D}^*, \phi_j] \Bbb{D} +[\Bbb{D}, \phi_j] \Bbb{D}^*\right\} (\pp_0 +\lambda)^{-1/2}\big) \endaligned \tag 4.22 $$ where $\Bbb{D}=\sigma\cdot (\bold{p}-\bold{A})$. Since $|\big[\Bbb{D}^*, [\Bbb{D},\phi_j]\big]\psi|\le |\nabla^2 \phi_j|\, |\psi|$, the first term on the r.h.s. of (4.22) can be treated exactly as $n(1/3, I_2)$. With $\| \Bbb{D} (\pp_0 +\lambda )^{-1/2}\|_{L^2\to L^2}\le 1$ and $\| \Bbb{D}^* (\pp_0 +\lambda )^{-1/2}\|_{L^2\to L^2}\le 1$, we may use the same argument as in the proof of Lemma 4.1 to bound the second term by $$ \aligned C\| \sum_j Y\phi_j (\pp_0 +\Phi & +\frac{1}{\ell_j^2}+\lambda)^{-1} [\Bbb{D}, \phi_j]\|_4^4 +C\| \sum_j Y\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2}+\lambda)^{-1} [\Bbb{D}^*, \phi_j]\|_4^4\\ &\le C\sum_j \|Y\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2}+\lambda)^{-1}\|_4^4 \cdot\frac{1}{\ell_j^4}\\ & \le C\sum_j \|Y\phi_j (\pp_0 +\Phi +\frac{1}{\ell_j^2}+\lambda)^{-1/2}\|_4^4\\ &\le C\sum_j \|Y\phi_j (\hh_0+\lambda)^{-1/2}\|_4^4\\ &\le\frac{C}{\sqrt{\lambda}} \int_{\{ x\in \Bbb{R}^3:V(x)< -\lambda\} } |V|^2dx. \endaligned $$ The lemma then follows by integration. \enddemo We need one more lemma before we carry out the proof of Theorem 1.1. \proclaim{\bf Lemma 4.5} There exist two constants $C_1>0$, $C_2>0$ depending on $\varepsilon$ and $p$, such that, for every $x\in \Bbb{R}^3$, $$ C_1 \Phi (x)\le b_p(x)\le C_2 \Phi(x). $$ \endproclaim \demo{\bf Proof} Suppose $x\in Q_j$ for some $j$. Then $\Phi (x)\approx 1/\ell_j^2$. Let $\delta\in (0,1)$ and $\ell_j=\ell(Q_j)$. Since $Q(x,\delta \ell_j)\subset 12 Q_j$, we have $$ \align (\delta \ell_j)^2 \bigg(\frac{1}{(\delta\ell_j)^3} \int_{Q(x,\delta \ell_j)}& |\bb(y)|^pdy\bigg)^{1/p} = (\delta \ell_j)^{2-\frac{3}{p}} \left(\int_{Q(x,\delta \ell_j)} |\bb(y)|^pdy\right)^{1/p}\\ &\le (\delta \ell_j)^{2-\frac{3}{p}} \left(\int_{12 Q_j} |\bb(y)|^pdy\right)^{1/p}\\ &\le (\delta \ell_j)^{2-\frac{3}{p}} \cdot \frac{\varepsilon}{\ell_j^{2-\frac{3}{p}}} =\delta^{2-\frac{3}{p}} \varepsilon <1. \endalign $$ Hence, by the definition of $\ell_p(x)$ (see (1.9)), we have $\ell_p(x)\ge \ell_j$. It follows that $$ b_p(x)=\frac{1}{\{ \ell_p(x)\}^2 } \le\frac{1}{\ell_j^2}\le C_2 \Phi (x). $$ To show $b_p(x)\ge C_1 \Phi(x)$, we note that, if $N$ is large, $12 Q_j^+\subset Q(x,N\ell_j)$ where $Q_j^+$ is the dyadic parent of $Q_j$. Hence $$ \align (N\ell_j)^2 \bigg(\frac{1}{(N\ell_j)^3} \int_{Q(x,N\ell_j)}& |\bb(y)|^pdy\bigg)^{1/p} = (N\ell_j)^{2-\frac{3}{p}} \left( \int_{Q(x,N\ell_j)} |\bb(y)|^pdy\right)^{1/p}\\ &\ge (N\ell_j)^{2-\frac{3}{p}} \left( \int_{12 Q_j^+} |\bb(y)|^pdy\right)^{1/p}\\ &> (N\ell_j)^{2-\frac{3}{p}}\cdot \frac{\varepsilon} {(2\ell_j)^{2-\frac{3}{p}}} =\varepsilon \left(\frac{N}{2}\right)^{2-\frac{3}{p}} >1 \endalign $$ if $N$ is large enough. Again, by definition, this implies that $\ell_p(x)\le N\ell_j$. Therefore $$ b_p(x)=\frac{1}{\{\ell_p(x)\}^2 } \ge \frac{1}{N^2\ell_j^2} \ge C_1 \Phi (x). $$ \enddemo Finally we are in a position to give the \demo{\bf Proof of Theorem 1.1} It suffices to prove the theorem for $\gamma=1$ (see \cite{20}). We may also assume, without the loss of generality, that $V\le 0$ a.e.. By means of (4.1) and (4.3), we have $$ \align \Cal{M}_1&=\sum_{\lambda_j<0} |\lambda_j| =2\int_0^\infty N(2\lambda, \pp_0+V)\, d\lambda\\ &\le 2\int_0^\infty n(1, Y(\pp_0+\lambda)^{-1/2})\, d\lambda\\ &\le 2\sum_{j=1}^3 \int_0^\infty n(1/3, I_j)\, d \lambda. \endalign $$ The theorem now follows from (4.8) and Lemmas 4.3--4.5. \enddemo \bigskip \centerline{\bf 5. Fields with Constant Directions} In this section we study the special case where the magnetic field $\bb$ has a constant direction. The goal is to prove Theorem 1.2 stated in the Introduction. Without the loss of generality, we may assume that $\bold{A}=(A_1(x_1,x_2), A_2(x_1,x_2),0)$ and $\bb=(0,0,B)$ where $B=\frac{\partial A_2}{\partial x_1} -\frac{\partial A_1}{\partial x_2}$. With this assumption, we have the following identity: $$ \int_{\Bbb{R}^3} |\sigma\cdot (\bold{p}-\bold{A})\psi|^2dx =\int_{\Bbb{R}^3} | [\sigma_1(p_1-A_1)+\sigma_2(p_2-A_2)]\psi|^2dx +\int_{\Bbb{R}^3} |\frac{\partial \psi}{\partial x_3}|^2dx. $$ This in particular implies that, for $\lambda>0$, $$ \| (-\partial_{x_3}^2 +\lambda)^{1/2} (\pp+\lambda)^{-1/2}\|_{L^2\to L^2}\le 1. \tag 5.1 $$ We shall use $x^\prime =(x_1,x_2)$, $y^\prime=(y_1,y_2)$ to denote points in $\Bbb{R}^2$. Our approach to the case of constant direction fields is similar to that in Section 4 for arbitrary fields. We begin by observing that, if $\bb=(0,0,B(x^\prime))$, the basic length scale $\ell_p(x)$ defined by (1.9) is reduced to $$ \ell_p(x^\prime) =\sup \left\{ \ell>0:\ \ \ell^2 \left\{ \frac{1}{\ell^2}\int_{S(x^\prime,\ell)} |B(y^\prime)|^p dy^\prime\right\}^{1/p}\le 1 \right\} \tag 5.2 $$ where $S(x^\prime,\ell)$ denotes the square in $\Bbb{R}^2$ centered at $x^\prime$ with side length $\ell$. We will assume that $B\not\equiv 0$ and $B\in L^p_{loc}(\Bbb{R}^2)$ for some $p>1$. With this we have $0<\ell_p(x^\prime)<\infty$ for any $x^\prime\in \Bbb{R}^2$. We now sketch the construction of the partition of unity for $\Bbb{R}^2$ associated with $B$, which is parallel to that in Section 3. First we write $$ \Bbb{R}^2 =\bigcup_j S_j \tag 5.3 $$ where $\{ S_j\}_{j=1}^\infty$ are maximal elements in the set of all dyadic squares $S$ in $\Bbb{R}^2$ such that $$ \left\{ \int_{12 S} |B(x^\prime)|^pd x^\prime\right\}^{1/p} \le \frac{\varepsilon} {[\ell(S)]^{2-\frac{2}{p}}} \tag 5.4 $$ where $\varepsilon\in (0,1)$ is a constant to be determined later. Next we use the same argument as in the proof of Lemma 3.2 to show that, $$ \frac12\ell(S_k) \le \ell(S_j) \le 2\ell(S_k), \ \ \ \ \text{ if } \ \ 4S_j\cap 4S_k\neq\emptyset. \tag 5.5 $$ It follows from (5.5) that there exists a sequence of functions $\{ \varphi_j\}$ such that $$ \align &\varphi_j\in C_0^\infty (2S_j, \Bbb{R})\ \ \ \text{ and }\ \ 0\le \varphi_j\le 1, \tag 5.6\\ & |\nabla^\alpha \varphi_j|\le C_\alpha/\ell_j^{|\alpha|} \ \ \ \text{ where }\ \ \ell_j=\ell(S_j), \tag 5.7\\ & \sum_j\varphi_j^2\equiv 1\ \ \ \text{ in } \ \Bbb{R}^2. \tag 5.8 \endalign $$ Let $$ \Psi =\sum_j\frac{1}{\ell_j^2}\, \varphi_j^2. \tag 5.9 $$ \proclaim{\bf Theorem 5.1} There exist constants $C>0$ and $\varepsilon_0>0$ such that, if $\varepsilon$ in (5.4) is less than $\varepsilon_0$, then $$ \hh_0\le C\left\{ \pp_0 +\Psi\right\}. $$ \endproclaim \demo{\bf Proof} We will only show that, if $\varepsilon$ in (5.4) is small, then $$ \int_{\Bbb{R}^3} |\, |B|^{1/2} \varphi_j\psi|^2dx \le C\int_{\Bbb{R}^3} |\sigma\cdot (\bold{p}-\bold{A})(\varphi_j\psi)|^2dx \tag 5.10 $$ for any $\psi\in C^\infty_0(\Bbb{R}^3,\Bbb{C}^2)$. The rest of the proof is exactly the same as that of Theorem 3.1. To show (5.10), we need the following Sobolev inequality $$ \left(\int_S |f|^qdx^\prime\right)^{1/q} \le C_q |S|^{1/q} \left( \int_S |\nabla_2 f|^2dx^\prime\right)^{1/2} \tag 5.11 $$ for $f\in C_0^\infty (S,\Bbb{R})$ where $S$ is a square, $\nabla_2 =(\partial_{x_1},\partial_{x_2})$, and $10$ such that, for any $\lambda>0$, $$ n(1/3, J_1) \le \frac{C}{\sqrt{\lambda}} \int_{\{ x\in \Bbb{R}^3: V(x)<-\lambda\} } |V(x)|^2dx. $$ \endproclaim The proof of Lemma 5.1, which uses (5.12)--(5.13) and (2.12), is similar to that of Lemma 4.1. We leave it to the reader. Let $\xi_j\in C_0^\infty (3S_j,\Bbb{R})$ such that $0\le \xi_j\le 1$, $\xi_j\varphi_j\equiv \varphi_j$, and $|\nabla^\alpha\xi_j|\le C_\alpha/\ell_j^{|\alpha|}$. The following lemma is needed to estimate $n(1/3,J_2)$ and $n(1/3,J_3)$. \proclaim{\bf Lemma 5.2} Let $0<\delta \le \frac12 (1-\frac1p)$. There exists a constant $C_\delta>0$ such that, for any $\lambda >0$, $$ \| (\hh_0 +\frac{1}{\ell_j^2} +\lambda)^{-\frac12 +\delta} (\sigma\cdot \bb)\xi_j (\pp_0+\Psi +\frac{1}{\ell_j^2}+\lambda)^{-1} \|_{L^2\to L^2} \le C_\delta \, \ell_j^{1-2\delta}. $$ \endproclaim \demo{\bf Proof} Note that, by (5.12), $$ \aligned \| (\hh_0 +&\frac{1}{\ell_j^2} +\lambda)^{-\frac12 +\delta} (\sigma\cdot \bb)\xi_j (\pp_0+\Psi +\frac{1}{\ell_j^2}+\lambda)^{-1} \|_{L^2\to L^2}\\ &\le \| (\hh_0 +\frac{1}{\ell_j^2} +\lambda)^{-\frac12 +\delta} (\sigma\cdot \bb)\xi_j (\pp_0+\Psi +\frac{1}{\ell_j^2}+\lambda)^{-1/2} \|_{L^2\to L^2}\cdot \ell_j\\ &\le \| (\hh_0 +\frac{1}{\ell_j^2} +\lambda)^{-\frac12 +\delta} (\sigma\cdot \bb)\xi_j (\hh_0+\frac{1}{\ell_j^2}+\lambda)^{-1/2} \|_{L^2\to L^2}\\ &\ \ \ \ \ \ \ \ \ \ \ \ \ \cdot \|(\hh_0+\frac{1}{\ell_j^2}+\lambda)^{1/2} (\pp_0+\Psi +\frac{1}{\ell_j^2}+\lambda)^{-1/2} \|_{L^2\to L^2}\cdot \ell_j\\ &\le C\ell_j \| (\hh_0 +\frac{1}{\ell_j^2} +\lambda)^{-\frac12 +\delta} (\sigma\cdot \bb)\xi_j (\hh_0+\frac{1}{\ell_j^2}+\lambda)^{-\frac12} \|_{L^2\to L^2}\\ &\le C\ell_j^{1+2\delta} \| (\hh_0 +\frac{1}{\ell_j^2} +\lambda)^{-\frac12 +\delta} (\sigma\cdot \bb)\xi_j (\hh_0+\frac{1}{\ell_j^2}+\lambda)^{-\frac12 +\delta} \|_{L^2\to L^2}. \endaligned $$ Thus, by the diamagnetic inequality (2.10) and duality, it suffices to show that $$ \|\, |B|^{1/2}\chi_{3S_j} (-\Delta +\frac{1}{\ell_j^2} +\lambda )^{-\frac12 +\delta} \|_{L^2\to L^2} \le C_\delta\, \ell_j^{-2\delta}. \tag 5.15 $$ We claim that (5.15) follows from Lemma 2.3. Indeed, by Lemma 2.3, the l.h.s. of (5.15) is bounded by $$ C_\delta \, \|\, |B|^{1/2}\|_{L^{\frac{2}{1-2\delta}}(3S_j)} \le C_\delta \ell_j^{-2\delta} $$ where we used $p\ge 1/(1-2\delta)$, H\"older's inequality, and (5.4) in the inequality. The proof is complete. \enddemo We are now ready to deal with $n(1/3,J_2)$ and $n(1/3,J_3)$. \proclaim{\bf Lemma 5.3} There exists a constant $C>0$ such that $$ n(1/3, J_2) \le \frac{C}{\sqrt{\lambda}} \int_{\{ x\in\Bbb{R}^3: V(x)<-\lambda\} } \Psi (x) |V(x)|dx. $$ \endproclaim \demo{\bf Proof} First we note that, by (2.4), (2.6), and (5.1), $$ \aligned n(1/3, J_2) &=n\big(1/3, \sum_j Y\varphi_j (\pp_0 +\Psi +\frac{1}{\ell_j^2} +\lambda)^{-1} (\Psi +\frac{1}{\ell_j^2})\varphi_j (\pp_0+\lambda)^{-1/2}\big)\\ &\le C \| \sum_j Y\varphi_j (\pp_0 +\Psi +\frac{1}{\ell_j^2} +\lambda)^{-1} (\Psi +\frac{1}{\ell_j^2})\varphi_j (\pp_0+\lambda)^{-1/2} \|_2^2\\ &\le C \| \sum_j Y\varphi_j(-\partial_{x_3}^2+\lambda)^{-1/2} (\pp_0 +\Psi +\frac{1}{\ell_j^2} +\lambda)^{-1} (\Psi +\frac{1}{\ell_j^2})\varphi_j \|_2^2\\ &\ \ \ \ \ \ \ \ \cdot \| (-\partial_{x_3}^2+\lambda)^{1/2} (\pp_0+\lambda)^{-1/2}\|^2_{L^2\to L^2}\\ &\le C \|\sum_j Y\varphi_j(-\partial_{x_3}^2+\lambda)^{-1/2} (\pp_0 +\Psi +\frac{1}{\ell_j^2} +\lambda)^{-1} (\Psi +\frac{1}{\ell_j^2})\varphi_j \|_2^2\\ &\le C \sum_j \|Y\varphi_j(-\partial_{x_3}^2+\lambda)^{-1/2} (\pp_0 +\Psi +\frac{1}{\ell_j^2} +\lambda)^{-1} (\Psi +\frac{1}{\ell_j^2})\varphi_j \|_2^2\\ &\le C \sum_j \|Y\varphi_j(-\partial_{x_3}^2+\lambda)^{-1/2} (\pp_0 +\Psi +\frac{1}{\ell_j^2} +\lambda)^{-1}\varphi_j \|_2^2\cdot\frac{1}{\ell_j^4} \endaligned $$ where we used the finite intersection property of supp$\varphi_j$ in the fourth inequality. The fact that $\pp_0+\Psi$ commutes with $\partial_{x_3}$ is essential in the estimate above. Next we use the resolvent identity (4.2) to obtain $$ \aligned &\|Y\varphi_j(-\partial_{x_3}^2+\lambda)^{-1/2} (\pp_0 +\Psi +\frac{1}{\ell_j^2} +\lambda)^{-1}\varphi_j \|_2\\ &= \|Y\varphi_j(-\partial_{x_3}^2+\lambda)^{-1/2} \xi_j(\pp_0 +\Psi +\frac{1}{\ell_j^2} +\lambda)^{-1}\varphi_j \|_2\\ &\le \|Y\varphi_j(-\partial_{x_3}^2+\lambda)^{-1/2} (\hh_0+\frac{1}{\ell_j^2} +\lambda)^{-1}\varphi_j \|_2\\ &\ \ \ \ \ +\|Y\varphi_j(-\partial_{x_3}^2+\lambda)^{-1/2} (\hh_0+\frac{1}{\ell_j^2} +\lambda)^{-1}(\sigma\cdot \bb)\xi_j (\pp_0 +\Psi +\frac{1}{\ell_j^2} +\lambda)^{-1}\varphi_j \|_2\\ &\ \ \ \ \ +\|Y\varphi_j(-\partial_{x_3}^2+\lambda)^{-1/2} (\hh_0+\frac{1}{\ell_j^2} +\lambda)^{-1} \left\{ -\Psi\xi_j +[\pp_0,\xi_j]\right\} (\pp_0 +\Psi +\frac{1}{\ell_j^2} +\lambda)^{-1}\varphi_j \|_2\\ &=K_1+K_2+K_3. \endaligned $$ By (2.13), $$ K_1\le C\lambda^{-1/4} (\frac{1}{\ell_j^2}+\lambda)^{-1/2} \|Y\varphi_j\|_{L^2} \le C\lambda^{-1/4}\ell_j \|Y\varphi_j\|_{L^2}. \tag 5.16 $$ To estimate $K_2$, we use (2.6), Lemma 5.2, and (2.13) to obtain $$ \aligned K_2&\le \|Y\varphi_j (-\partial_{x_3}^2+\lambda)^{-1/2} (\hh_0+\frac{1}{\ell_j^2}+\lambda)^{-\frac12 -\delta}\|_2\\ &\ \ \ \ \ \ \ \ \cdot \| (\hh_0+\frac{1}{\ell_j^2}+\lambda)^{-\frac12 +\delta} (\sigma\cdot\bb )\xi_j (\pp_0+\Psi +\frac{1}{\ell_j^2}+\lambda)^{-1}\|_{L^2\to L^2}\\ &\le C\, \|Y\varphi_j (-\partial_{x_3}^2+\lambda)^{-1/2} (\hh_0+\frac{1}{\ell_j^2}+\lambda)^{-\frac12 -\delta}\|_2 \cdot \ell_j^{1-2\delta}\\ &\le C\, \lambda^{-1/4} (\frac{1}{\ell_j^2}+\lambda)^{-\delta} \|Y\varphi_j\|_{L^2}\cdot \ell_j^{1-2\delta}\\ &\le C\, \lambda^{-1/4}\ell_j \|Y\varphi_j\|_{L^2} \endaligned \tag 5.17 $$ where $\delta=\frac12 (1-\frac1p)$. To bound $K_3$, we observe that $$ \| \left\{ -\Psi\xi_j +[\pp_0,\xi_j]\right\} (\pp_0+\Psi +\frac{1}{\ell_j^2} +\lambda)^{-1} \|_{L^2\to L^2}\le C $$ using the same argument as in the proof of (4.9). It follows that $$ \aligned K_3 &\le \| Y\varphi_j (-\partial_{x_3}^2 +\lambda)^{-1/2} (\hh_0+\frac{1}{\ell_j^2}+\lambda)^{-1}\|_2\\ &\ \ \ \ \ \ \ \ \ \ \cdot \| \left\{ -\Psi\xi_j +[\pp_0,\xi_j]\right\} (\pp_0+\Psi +\frac{1}{\ell_j^2} +\lambda)^{-1} \|_{L^2\to L^2}\\ &\le C\lambda^{-1/4}\ell_j \|Y\varphi_j\|_{L^2}. \endaligned \tag 5.18 $$ Finally, we add (5.16), (5.17) and (5.18) to conclude that $$ \|Y\varphi_j (-\partial_{x_3}^2+\lambda)^{-1/2} (\pp_0+\Psi +\frac{1}{\ell_j^2} +\lambda)^{-1} \varphi_j \|_2 \le K_1+K_2 +K_3 \le C\lambda^{-1/4}\ell_j \|Y\varphi_j\|_{L^2}. $$ It follows that $$ \aligned n(1/3,J_2) &\le C\sum_j \|Y\varphi_j (-\partial_{x_3}^2+\lambda)^{-1/2} (\pp_0+\Psi +\frac{1}{\ell_j^2} +\lambda)^{-1} \varphi_j \|_2^2\cdot\frac{1}{\ell_j^4}\\ &\le\frac{C}{\sqrt{\lambda}} \sum_j \frac{1}{\ell_j^2} \int_{\Bbb{R}^3} |Y\varphi_j|^2dx\\ &\le \frac{C}{\sqrt{\lambda}} \int_{\{ x\in\Bbb{R}^3: V(x)< -\lambda\} } \Psi(x) |V(x)|dx \endaligned $$ since $|Y|\le |V|^{1/2}\chi_{\{ x\in\Bbb{R}^3: V(x)< -\lambda\} }$. The proof is finished. \enddemo \proclaim{\bf Lemma 5.4} There exists a constant $C>0$ such that $$ n(1/3,J_3) \le\frac{C}{\sqrt{\lambda}} \int_{\{ x\in\Bbb{R}^3: V(x)< -\lambda\} } |V(x)|^2dx + \frac{C}{\sqrt{\lambda}} \int_{\{ x\in\Bbb{R}^3: V(x)< -\lambda\} } \Psi(x) |V(x)|dx. $$ \endproclaim \demo{\bf Proof} Using $[\pp_0,\varphi_j]=\big[\Bbb{D}^*,[\Bbb{D},\varphi_j]\big] +[\Bbb{D},\varphi_j]\Bbb{D}^* +[\Bbb{D}^*,\varphi_j]\Bbb{D} $, we have $$ \aligned J_3&= \sum_jY\varphi_j (\pp_0+\Psi +\frac{1}{\ell_j^2}+\lambda)^{-1} [\pp_0,\varphi_j] (\pp_0 +\lambda)^{-1/2}\\ &= \sum_jY\varphi_j (\pp_0+\Psi +\frac{1}{\ell_j^2}+\lambda)^{-1} \big[\Bbb{D}^*,[\Bbb{D},\varphi_j]\big] (\pp_0 +\lambda)^{-1/2}\\ &\ \ \ \ +\sum_jY\varphi_j (\pp_0+\Psi +\frac{1}{\ell_j^2}+\lambda)^{-1} \left\{ [\Bbb{D},\varphi_j] \Bbb{D}^* + [\Bbb{D}^*,\varphi_j] \Bbb{D} \right\} (\pp_0 +\lambda)^{-1/2}\\ &=J_{31} +J_{32}. \endaligned $$ Thus, $$ n(1/3,J_3)\le n(1/6, J_{31}) +n(1/6, J_{32}). $$ Since $| [\Bbb{D}^*,[\Bbb{D},\varphi_j]]\psi|\le |\nabla^2 \varphi_j|\, |\psi|$, we may use the same argument as in the proof of Lemma 5.3 to show that $$ n(1/6, J_{31}) \le\frac{C}{\sqrt{\lambda}} \int_{\{ x\in\Bbb{R}^3: V(x)< -\lambda\} } \Psi (x) |V(x)|dx. $$ On the other hand, using $\| \Bbb{D} (\pp_0+\lambda)^{-1/2} \|_{L^2\to L^2}\le 1 $, $\| \Bbb{D}^* (\pp_0+\lambda)^{-1/2} \|_{L^2\to L^2}\le 1 $, and (5.12), we may bound $n(1/6, J_{32})$ by $$ \aligned &C\| \sum_jY\varphi_j (\pp_0+\Psi +\frac{1}{\ell_j^2}+\lambda)^{-1} \left\{ [\Bbb{D}^*,\varphi_j] \Bbb{D} + [\Bbb{D},\varphi_j] \Bbb{D}^*\right\} (\pp_0 +\lambda)^{-1/2}\|_4^4\\ &\le C\| \sum_jY\varphi_j (\pp_0+\Psi +\frac{1}{\ell_j^2}+\lambda)^{-1} [\Bbb{D}^*,\varphi_j]\|_4^4 + C\| \sum_jY\varphi_j (\pp_0+\Psi +\frac{1}{\ell_j^2}+\lambda)^{-1} [\Bbb{D},\varphi_j]\|_4^4\\ &\le C \sum_j\|Y\varphi_j (\pp_0+\Psi +\frac{1}{\ell_j^2}+\lambda)^{-1} [\Bbb{D}^*,\varphi_j]\|_4^4 + C \sum_j\|Y\varphi_j (\pp_0+\Psi +\frac{1}{\ell_j^2}+\lambda)^{-1} [\Bbb{D},\varphi_j]\|_4^4\\ &\le C \sum_j\|Y\varphi_j (\pp_0+\Psi +\frac{1}{\ell_j^2}+\lambda)^{-1/2}\|_4^4 \cdot\frac{1}{(\frac{1}{\ell_j^2}+\lambda)^2} \cdot \frac{1}{\ell_j^4}\\ &\le C \sum_j\|Y\varphi_j (\hh_0+\frac{1}{\ell_j^2}+\lambda)^{-1/2}\|_4^4 \| (\hh_0+\frac{1}{\ell_j^2}+\lambda)^{1/2} (\pp_0+\Psi +\frac{1}{\ell_j^2}+\lambda)^{-1/2}\|_{L^2\to L^2}^4\\ &\le C \cdot \sum_j\|Y\varphi_j (\hh_0+\frac{1}{\ell_j^2}+\lambda)^{-1/2}\|_4^4\\ &\le \frac{C}{\sqrt{\lambda}} \int_{\{ x\in\Bbb{R}^3: V(x)< -\lambda\} } |V(x)|^2dx \endaligned $$ where (2.12) is used in the last inequality. The proof is complete. Finally we are in a position to give the \demo{\bf Proof of Theorem 1.2} It follows from (5.14), Lemmas 5.1, 5.3, and 5.4 that $$ \aligned N(2\lambda, \pp_0 +V) &\le n(1/3, J_1) +n(1/3, J_2) +n(1/3, J_3)\\ &\le \frac{C}{\sqrt{\lambda}} \int_{\{ x\in\Bbb{R}^3: V(x)< -\lambda\} } |V(x)|^2dx +\frac{C}{\sqrt{\lambda}} \int_{\{ x\in\Bbb{R}^3: V(x)< -\lambda\} } \Psi (x)|V(x)|dx. \endaligned $$ Thus, if $\gamma>1/2$, $$ \aligned \Cal{M}_\gamma &=\sum_{\lambda_j<0} |\lambda|^\gamma =\gamma \int_0^\infty \lambda^{\gamma -1} N(\lambda,\pp_0+V)\, d\lambda\\ &= \gamma 2^\gamma \int_0^\infty \lambda^{\gamma -1} N(2\lambda,\pp_0+V)\, d\lambda\\ &\le C_\gamma \int_{\Bbb{R}^3} |V(x)|_-^{\gamma+3/2}dx +C_\gamma \int_0^\infty \Psi(x) |V(x)|_-^{\gamma+1/2}dx. \endaligned $$ Noting that $\Psi(x)\approx b_p(x)$ for $p>1$ by an argument similar to that in the proof of Lemma 4.5, we are done. \enddemo \Refs \widestnumber\key{20} \ref\key 1 \by J.~Avron, I.~Herbst, and B.~Simon \paper Schr\"odinger Operators with Magnetic Fields. I. General Interaction \jour Duke Math.~J.\vol 45(4)\yr 1978\pages 847--883 \endref \ref \key 2 \by L.~Bugliaro, C.~Fefferman, J.~Fr\"ohlich, G.~Graf, and J.~Stubbe \paper A Lieb-Thirring bound for a magnetic Pauli Hamiltonian \jour Commun.~Math.~Phys. \vol 187 \yr 1997 \pages 567-582 \endref \ref \key 3 \by L.~Erd\"os \paper Magnetic Lieb-Thirring inequalities \jour Commun.~Math.~Phys \vol 170 \yr 1995 \pages 629-668 \endref \ref \key 4 \by L.~Erd\"os and J.~Solovej \paper Semiclassical eigenvalue estimates for the Pauli operator with strong nonhomogeneous magnetic fields. I. Non-asymptotic Lieb-Thirring type estimates \jour Preprint \endref \ref \key 5 \bysame \paper Semiclassical eigenvalue estimates for the Pauli operator with strong nonhomogeneous magnetic fields. II. Leading order asymptotic estimates \jour Commun.~Math.~Phys. \vol 188 \pages 599-656 \yr 1997 \endref \ref \key 6 \by C. Fefferman \paper The uncertainty principle \jour Bull. Amer. Math. Soc. \vol 9(2) \yr 1983 \pages 129-206 \endref \ref \key 7 \bysame \paper Stability of Coulomb systems in a magnetic field \jour Proc.~Natl.~Acad.~Sci. \vol 92 \yr 1995 \pages 5006-5007 \endref \ref \key 8 \bysame \paper On electrons and nuclei in a magnetic field \jour Adv.~Math. \vol 124 \yr 1996 \pages 100-153 \endref \ref\key 9 \by H.~Leinfelder and C.~Simader \paper Schr\"odinger Operators with Singular Magnetic Vector Potentials \jour Math.~Z.\vol 176\yr 1981\pages 1--19 \endref \ref \key 10 \by E. Lieb, M. Loss, and J. Solovej \paper Stability of matter in magnetic fields \jour Phys. Rev. Lett. \vol 75\pages 985-989 \yr 1995 \endref \ref \key 11 \by E.~Lieb, J.~Solovej, and J.~Yngvason \paper Asymptotics of heavy atoms in high magnetic fields: II. Semi-classical regions \jour Commun.~Math.~Phys \vol 161 \pages 77-124 \yr 1994 \endref \ref \key 12 \bysame \paper Ground states of large quantum dots in magnetic fields \jour Phys. Rev. B 51 \yr 1995\pages 10646-10665 \endref \ref \key 13 \by E.~Lieb and W.~Thirring \paper Bound for the kinetic energy of fermions which proves the stability of matter \jour Phys.~Rev.~Lett. \vol 35 \pages 687-689 \endref \ref \key 14 \by M.~Reed and B.~Simon \book Methods of Modern Mathematical Physics, II \publ Academic Press, New York \yr 1975 \endref \ref \key 15 \bysame \book Methods of Modern Mathematical Physics, IV \publ Academic Press, New York \yr 1978 \endref \ref \key 16 \by Z.~Shen \paper Eigenvalue asympotics and exponential decay of eigenfunctions for Schr\"odinger operators with magnetic fields \jour Trans.~Amer.~Math.~Soc. \vol 348 \yr 1996 \pages 4465-4488 \endref \ref \key 17 \bysame \paper On the number of negative eigenvalues for a Schr\"odinger operator \jour Commun. Math. Phys. \vol 182 \yr 1996 \pages 637-660 \endref \ref \key 18 \by B.~Simon \book Trace ideals and their applications \publ Cambridge Univ.~Press \yr 1979 \endref \ref \key 19 \by A.~Sobolev \paper On the Lieb-Thirring estimates for the Pauli operator \jour Duke Math.~J. \vol 82\yr 1996 \pages 607-635 \endref \ref \key 20 \bysame \paper Lieb-Thirring inequalities for the Pauli operator in three dimensions \jour Preprint \endref \ref\key 21 \by E.~Stein \book Singular Integrals and Differentiability Properties of Functions \publ Princeton Univ. Press\yr 1970 \endref \endRefs \end