\documentclass[reqno]{amsart} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2004(2004), No. 145, pp. 1--10.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu (login: ftp)} \thanks{\copyright 2004 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2004/145\hfil Estimates for the mixed derivatives] {Estimates for the mixed derivatives of the Green functions on homogeneous manifolds of negative curvature} \author[R. Urban\hfil EJDE-2004/145\hfilneg] {Roman Urban} \address{Roman Urban \hfill\break Institute of Mathematics, University of Wroclaw, Pl. Grunwaldzki 2/4, 50-384 Wroclaw, Poland} \email{urban@math.uni.wroc.pl} \date{} \thanks{Submitted October 17, 2003. Published December 7, 2004.} \thanks{Partly supported by grant 1PO3A01826 from KBN, and by contract HPRN-CT-2001-\hfill\break\indent 00273-HARP from RTN Harmonic Analysis and Related Problems} \subjclass[2000]{22E25, 43A85, 53C30} \keywords{Green function; second-order differential operators; $NA$ groups; \hfill\break\indent Bessel process; evolutions on nilpotent Lie groups} \begin{abstract} We consider the Green functions for second-order left-invariant differential operators on homogeneous manifolds of negative curvature, being a semi-direct product of a nilpotent Lie group $N$ and $A=\mathbb{R}^+$. We obtain estimates for mixed derivatives of the Green functions both in the coercive and non-coercive case. The current paper completes the previous results obtained by the author in a series of papers \cite{amuc,pota,ejde,CM}. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{proposition}[theorem]{Proposition} \section{Introduction} Let $M$ be a connected and simply connected homogeneous manifold of negative curvature. Such a manifold is a solvable Lie group $S=NA$, a semi-direct product of a nilpotent Lie group $N$ and an Abelian group $A=\mathbb{R}^+$. Moreover, for an $H$ belonging to the Lie algebra $\mathfrak{a}$ of $A$, the real parts of the eigenvalues of $\mathop{\rm Ad}_{\exp H}|_{\mathfrak{n}}$, where $\mathfrak{n}$ is the Lie algebra of $N$, are all greater than 0. Conversely, every such a group equipped with a suitable left-invariant metric becomes a homogeneous Riemannian manifold with negative curvature (see \cite{H}). On $S$ we consider a second-order left-invariant operator \begin{equation*} \mathcal{L} =\sum_{j=0}^mY_j^2+Y. \end{equation*} We assume that $Y_0,Y_1,\dots,Y_m$ generate the Lie algebra $\mathfrak{s}$ of $S$ and $Y\in\mathfrak{s}$. We can always make $Y_0,\dots, Y_m$ linearly independent and moreover, we can choose $Y_0,Y_1,\dots,Y_m$ so that $Y_1(e),\dots, Y_m(e)$ belong to $\mathfrak{n}$ (write $\mathcal{L}$ as $\sum_{i,j=0}^{\mathop{\rm dim}\mathfrak{n}}\alpha_{i,j}E_iE_j +\sum_{j=0}^{\mathop{\rm dim}\mathfrak{n}}\beta_jE_j$, $E_0\in\mathfrak{a}$, $\{E_j\}$ is a basis of $\mathfrak{n}$, $\alpha_{i,j},\beta_j\in\mathbb{R}$ and then rewrite $\mathcal{L}$ as a sum of squares). Let $\pi:S\to A=S\slash N$ be the canonical homomorphism. Then the image of $\mathcal{L}$ under $\pi$ is a second-order left-invariant operator on $\mathbb{R}^+$, \begin{equation*} \pi(\mathcal{L})=(a\partial_a)^2-\gamma a\partial_a, \end{equation*} where $\gamma=\gamma_\mathcal{L}\in\mathbb{R}$. We say that a second order differential operator $\mathcal{L}$ on a Riemannian manifold is \textit{non-coercive} (\textit{coercive} resp.) if there is no $\varepsilon>0$ such that $\mathcal{L}+\varepsilon\text{Id}$ admits the Green function (if such an $\varepsilon$ exists resp.). It is worth noting that our definition of coercivity is slightly different than that used e.g. in \cite{A}. Namely, for us, $\mathcal{L}$ is coercive if it is weakly coercive in Ancona's terminology. There is a relation between the notion of coercivity property in the sense used in the theory of partial differential equations (i.e., that an appropriate bilinear form is coercive, \cite{Stamp}) and weak coercivity. For this the reader is referred to \cite{A}. In this paper we shall study both coercive and non-coercive operators. In this case $\mathcal{L}$ can be written as \begin{equation}\label{operator} \mathcal{L}=\mathcal{L}_\gamma=\sum_j\Phi_a(X_j)^2+\Phi_a(X)+a^2\partial_a^2 +(1-\gamma)\partial_a, \end{equation} where $\gamma=\gamma_{\mathcal{L}}\in\mathbb{R}$, $X,X_1,\dots,X_m$ are left-invariant vector fields on $N$, moreover, $X_1,\dots,X_m$ are linearly independent and generate $\mathfrak{n}$, $$\Phi_a={\mathop{\rm Ad}}_{\exp(\log a)Y_0}=e^{(\log a) \mathop{\rm ad}_{Y_0}}=e^{(\log a) D},$$ where $D=\mathop{\rm ad}_{Y_0}$ is a derivation of the Lie algebra $\mathfrak{n}$ of the Lie group $N$ such that the real parts $d_j$ of the eigenvalues $\lambda_j$ of $D$ are positive. By multiplying $\mathcal{L}_\gamma$ by a constant, i.e., changing $Y_0$, we can make $d_j$ arbitrarily large (see \cite{DHU}). Let $\mathcal{G}_\gamma(xa,yb)$ be the {\it Green function} for $\mathcal{L}_\gamma$. $\mathcal{G}_\gamma$ is (uniquely) defined by two conditions: \begin{itemize} \item [(i)] $\mathcal{L}_\gamma\mathcal{G}_\gamma(\cdot,yb)=-\delta_{yb}$ as distributions (functions are identified with distributions via the right Haar measure), \item [(ii)] for every $yb\in S$, $\mathcal{G}_\gamma(\cdot,yb)$ is a potential for $\mathcal{L}_\gamma$, i.e, is a positive superharmonic function such that its largest harmonic minorant is equal to zero (cf. \cite{CC}). \end{itemize} Let \begin{equation}\label{gf} \mathcal G_\gamma(x,a):=\mathcal G_\gamma(e,xa), \end{equation} where $e$ is the identity element of the group $S$. Since $\mathcal{L}_\gamma$ is left-invariant it is easily seen that \begin{equation*} \mathcal G_\gamma(xa,yb)=\mathcal G_\gamma(e,yb(xa)^{-1})=\mathcal G_\gamma(yb(xa)^{-1}). \end{equation*} In this article we call $\mathcal G_\gamma(x,a)$ defined in \eqref{gf} the Green function for $\mathcal L_\gamma$. The main goal of this paper is to give estimates for derivatives of the Green function \eqref{gf} for $\mathcal{L}_\gamma$. To illustrate the general set up, before we proceed further, we would like to give a simple and explicit example of the operator $\mathcal L$ in coordinates. Consider $S=\mathbb{R}^n\times\mathbb{R}^+$. Let $d_1,\dots,d_n$ be positive constants. For every $a>0$, define $\Phi_a(\partial_{x_j})=a^{d_j}\partial_{x_j}$. Then $\Phi_a$ on $\mathbb{R}^n$ becomes $\Phi_a(x)=\Phi_a(x_1,\dots,x_n)=(a^{d_1}x_1,\dots,a^{d_n}x_n)$ and we get on $\mathbb{R}^n$ a structure of the homogeneous group with the homogeneous dimension $Q=\sum d_j$ (see \cite{FS}). The multiplication law in $S$ is given by the formula $(x,a)\cdot(y,b)=(x+\Phi_a(y),ab)$. In this example the operator \eqref{operator} is $\mathcal{L}=\sum_j a^{2d_j}\partial_{x_j}^2+a^2\partial_a^2+(1-\gamma)a\partial_a$. The Green function for $\mathcal{L}$ is $\mathcal{G}((x,a),(y,b))=\int_0^\infty p_t(x,a;y,b)dt$, where $p_t(x,a;y,b)$ is the heat diffusion kernel on $S$, such that $u(t,y,b):=p_t(x,a;y,b)$ is the minimal solution of $\mathcal{L}u=\partial_t u$, $u(0,y,b)=\delta_{(x,a)}(y,b)$ and $\delta_{(x,a)}(\cdot)$ stands for Dirac's delta. Let us go back to the general setting. We are going to prove (or at least to sketch the proof of) the following estimates. Let $\gamma\geq 0$. For every neighborhood $\mathcal U$ of the identity $e$ of $NA$ there is a constant $C=C(\gamma)$ such that we have \begin{equation}\label{ge} |\partial_a^k\mathcal{X}^I\mathcal{G}_{-\gamma}(x,a)|\leq \begin{cases} C(|x|+a)^{-\|I\|-Q-\gamma}a^{-k}&\\ \times (1+|\log(|x|+a)^{-1}|)^{\|I\|_0} &\text{ for $(x,a)\in(\mathcal Q\cup\mathcal U)^c$,}\\[3pt] Ca^{-k}&\text{ for $(x,a)\in \mathcal Q\setminus\mathcal U$} \end{cases} \end{equation} and \begin{equation}\label{ge1} |\partial_a^k\mathcal{X}^I\mathcal{G}_{\gamma}(x,a)|\leq \begin{cases} C(|x|+a)^{-\|I\|-Q-\gamma}a^{\gamma-k}&\\ \times (1+|\log(|x|+a)^{-1}|)^{\|I\|_0} &\text{ for $(x,a)\in(\mathcal Q\cup \mathcal U)^c$,}\\[3pt] Ca^{\gamma-k}&\text{ for $(x,a)\in \mathcal Q\setminus \mathcal U,$} \end{cases} \end{equation} where $|\cdot|$ stands for a ``homogeneous norm" on $N$, $\mathcal Q=\{|x|\leq 1, a\leq 1\}$, $\|I\|$ is a suitably defined length of the multi-index $I$ and $\|I\|_0$ is a certain number depending on $I$ and the nilpotent part of the derivation $D$. In particular, $\|I\|_0$ is equal to 0 if the action of $A=\mathbb{R}^+$ on $N$, given by $\Phi_a$, is diagonal or, if $I=0$. $\mathcal{X}_1,\dots,\mathcal{X}_n$ is an appropriately chosen basis of $\mathfrak{n}$. For the precise definitions of all the notions that have appeared here see Sect. \ref{preliminaries}. It should be said that the estimate for the Green function itself (i.e., $I$=0) with $\gamma>0$, also from below, was proved by E. Damek in \cite{D} and then by the author for $\gamma=0$ in \cite{CM} but at that time it was impossible to prove analogous estimate for derivatives. The reason was that we did not have sufficient estimates for the derivatives of the transition probabilities of the evolution on $N$ generated by an appropriate operator which appears as the ``horizontal" component of the diffusion on $N\times\mathbb{R}^+$ generated by $a^{-2}\mathcal L_{-\gamma}$ (cf. \cite{DHU}). These estimates have been obtained by the author in \cite{MathZ} and eventually led up to the estimates for derivatives of the Green functions in the non-coercive case, i.e., $\gamma=\gamma_{\mathcal{L}}=0$ (see \cite{pota} for derivatives with respect to $N$ and $A$-variables separately and \cite {ejde} for the mixed derivatives which required a slightly different approach). Next, in \cite{amuc} the results from \cite{pota} have been used to get estimates for derivatives in the coercive case. This note completes the previous works of the author in that we provide a proof of the estimates which is valid for both the coercive and non-coercive cases. The proofs of \eqref{ge} and \eqref{ge1} require both analytic and probabilistic techniques. Some of them have been introduced in \cite{DHZ,DHU} and \cite{CM}. The structure of the paper is as follows. In Sect. \ref{preliminaries} we set the notation and give all necessary definitions. In particular, we recall a definition of the Bessel process which appears as the``vertical" component of the diffusion generated by $a^{-2}\mathcal{L}_{-\gamma}$ on $N\times\mathbb{R}^+$ as well as the notion of the evolution on $N$ generated by an appropriate operator which appears as the``horizontal" component of the diffusion on $N\times\mathbb{R}^+$ mentioned in the above (cf. \cite{DHU, pota}). Moreover, we state Theorem \ref{r} which is the main tool in the proof of Theorem \ref{main}. Finally, in Sect. \ref{proof of theorem} we state the estimates \eqref{ge}, \eqref{ge1} precisely (see Theorem \ref{main}) and we give a sketch of its proofs. \section{Preliminaries.}\label{preliminaries} \subsection*{$NA$ groups} Good reference for this topic are \cite{DHZ, DHU} and \cite{FS}. Let $N$ be a connected and simply connected nilpotent Lie group. Let $D$ be a derivation of the Lie algebra $\mathfrak{n}$ of $N$. For every $a\in\mathbb{R}^+$ we define an automorphism $\Phi_a$ of $\mathfrak{n}$ by the formula \begin{equation*} \Phi_a=e^{(\log a)D}. \end{equation*} Writing $x=\exp X$ we put \begin{equation*} \Phi_a(x):=\exp\Phi_a(X). \end{equation*} Let $\mathfrak{n}^\mathbb{C}$ be the complexification of $\mathfrak{n}$. Define \begin{equation*} \mathfrak{n}_\lambda^\mathbb{C}=\{X\in\mathfrak{n}^\mathbb{C}:\exists k>0\text{ such that }(D-\lambda I)^k=0\}. \end{equation*} Then \begin{equation}\label{gradation} \mathfrak{n}=\bigoplus_{\text{Im}\lambda\geq 0}V_\lambda, \end{equation} where \begin{equation*} V_\lambda= \begin{cases} V_{\bar\lambda}=(\mathfrak{n}_\lambda^\mathbb{C}\oplus\mathfrak{n}_{\bar\lambda}^\mathbb{C})\cap\mathfrak{n}&\text{if Im$\lambda\not=0$,}\\ \mathfrak{n}_\lambda^\mathbb{C}\cap\mathfrak{n}&\text{if Im$\lambda=0.$} \end{cases} \end{equation*} We assume that the real parts $d_j$ of the eigenvalues $\lambda_j$ of the matrix $D$ are strictly greater than 0. We define the number \begin{equation}\label{qdef} Q=\sum_j\text{Re }\lambda_j=\sum_j d_j \end{equation} and we refer to this as a \textit{``homogeneous dimension"} of $N$. In this paper $D=\mathop{\rm ad}_{Y_0}$ (see Introduction). Under the assumption on positivity of $d_j$, \eqref{gradation} is a gradation of $\mathfrak{n}$. We consider a group $S$ which is a \textit{semi-direct} product of $N$ and the multiplicative group $A=\mathbb{R}^+= \{\exp tY_0:t\in\mathbb{R}\}:$ \begin{equation*} S=NA=\{xa:x\in N,a\in A\} \end{equation*} with multiplication given by the formula \begin{equation*} (xa)(yb)=(x\Phi_a(y)\ ab). \end{equation*} On $N$ we define a \textit{``homogeneous norm"}, $|\cdot|$ (cf. \cite{DHZ,DHU}) as follows. Let $(\cdot,\cdot)$ be a fixed inner product in $\mathfrak{n}$. We define a new inner product \begin{equation}\label{product} \langle X,Y\rangle =\int_0^1\Big(\Phi_a(X),\Phi_a(Y)\Big)\frac{da}{a} \end{equation} and the corresponding norm \begin{equation*} \|X\|=\langle X,X\rangle^{1\slash 2}. \end{equation*} We put \begin{equation*} |X|=\left(\inf\{a>0:\|\Phi_a(X)\|\geq 1\}\right)^{-1}. \end{equation*} One can easily show that for every $Y\ne 0$ there exists precisely one $a>0$ such that $Y=\Phi_a(X)$ with $|X|=1$. Then we have $|Y|=a$. Finally, we define the homogeneous norm on $N$. For $x=\exp X$ we put \begin{equation*} |x|:=|X|. \end{equation*} Notice that if the action of $A=\mathbb{R}^+$ on $N$ (given by $\Phi_a$) is diagonal the norm we have just defined is the usual homogeneous norm on $N$ and the number $Q$ in \eqref{qdef} is simply the homogeneous dimension of $N$ (see \cite{FS}). Having all that in mind we define appropriate derivatives (see also \cite{DHZ}). We fix an inner product \eqref{product} in $\mathfrak{n}$ so that $V_{\lambda_j}$, $j=1,\dots,k$ are mutually orthogonal and an orthonormal basis $\mathcal{X}_1,\dots,\mathcal{X}_n$ of $\mathfrak{n}$. The enveloping algebra $\mathfrak{U}(\mathfrak{n})$ of $\mathfrak{n}$ is identified with the polynomials in $\mathcal{X}_1,\dots,\mathcal{X}_n$. In $\mathfrak{U}(\mathfrak{n})$ we define $\langle \mathcal{X}_1\otimes\dots\otimes\mathcal{X}_r,\mathcal{Y}_1\otimes\dots\otimes\mathcal{Y}_r\rangle = \prod_{j=1}^r\langle \mathcal{X}_j,\mathcal{Y}_j\rangle$. Let $V_j^r$ be the symmetric tensor product of $r$ copies of $V_{\lambda_j}$. For $I=(i_1,\dots,i_k)\in(\mathbb{N}\cup\{0\})^k$ let \begin{equation*} \mathcal{X}^I=\mathcal{X}_1^{(i_1)}\dots\mathcal{X}_k^{(i_k)},\text{ where }\mathcal{X}_j^{(i_j)}\in V_j^{i_j}. \end{equation*} Then for $\mathcal{X}\in V_{\lambda_j},$ \begin{equation*} \|\Phi_a(\mathcal{X})\|\leq c\exp(d_j\log a+D_j\log(1+|\log a|)), \end{equation*} where $d_j=\text{Re}\lambda_j$ and $D_j=\dim V_{\lambda_j}-1$, and so \begin{equation}\label{den} \|\Phi_a(\mathcal{X}^I)\|\leq\exp\Big(\sum_{j=1}^ki_j(d_j\log a+D_j\log(1+|\log a|))\Big)\prod_{j=1}^k\|\mathcal{X}_j^{(i_j)}\| \end{equation} \subsection*{Bessel process} Let $\sigma(t)$ denote the Bessel process with a parameter $\alpha\geq 0$ (cf. \cite{RY}), i.e., a continuous Markov process with the state space $[0,+\infty)$ generated by $\partial_a^2+\frac{2\alpha+1}{a}\partial_a$. The transition function with respect to the measure $y^{2\alpha+1}dy$ is given (cf. \cite{Borodin, RY}) by: \begin{equation*} p_t(x,y)=\begin{cases} \frac{1}{2t}\exp\big(\frac{-x^2-y^2}{4t}\big) I_\alpha\left( \frac{xy}{2t}\right) \frac{1}{(xy)^\alpha} &\text{for $x,y>0,$}\\ \frac{1}{2^\alpha (2t)^{\alpha+1}\Gamma(\alpha+1)} \exp\big(\frac{-y^2}{4t}\big) &\text{for $x=0$, $y>0,$} \end{cases} \end{equation*} where \begin{equation*} I_\alpha(x)= \sum_{k=0}^\infty\frac{(x\slash 2)^{2k+\alpha}}{k!\Gamma(k+\alpha+1)} \end{equation*} is \textit{the Bessel function} (see \cite{L}). Therefore for $x\geq 0$ and a measurable set $B\subset(0,\infty)$: \begin{equation*} \mathbf{P}_x(\sigma(t)\in B)=\int_Bp_t(x,y)y^{2\alpha+1}dy. \end{equation*} If $\sigma(t)$ is the Bessel process with a parameter $\alpha$ starting from $x$, i.e. $\sigma(0)=x$, then we will write that $\sigma(t)\in\mathop{\rm BESS}_x(\alpha)$ or simply $\sigma(t)\in\mathop{\rm BESS}(\alpha)$ if the starting point is not important or is clear from the context. Properties of the Bessel process are very well known and their proofs are rather standard. They can be found e.g. in \cite{RY, DHU, pams, Phd}. However, in our paper we will not explicitly make use of any particular property of the Bessel process. \subsection*{Evolutions} Let $X,X_1,\dots, X_m$ be as in \eqref{operator}. Let $\sigma: [0,\infty)\longrightarrow [0,\infty)$ be a continuous function such that $\sigma(t)>0$ for every $t>0$. We consider the family of evolutions operators $L_{\sigma(t)}-\partial_t$, where \begin{equation}\label{operatorsigma} L_{\sigma(t)}=\sigma(t)^{-2}\Big(\sum_j\Phi_{\sigma(t)}(X_j)^2+ \Phi_{\sigma(t)}(X)\Big). \end{equation} For the main result of the paper we are mainly interested in the operator \eqref{operatorsigma} with $\sigma(t)$ being a trajectory of an appropriate Bessel process. Since we assume $X_1,\dots,X_m$ being linearly independent, we select $X_{m+1},\dots, X_n$ so that $X_1,\dots ,X_n$ form a basis of $\mathfrak{n}$. For a multi-index $I=(i_1,\dots,i_n)$, $i_j\in\mathbb{Z}^+$ and the basis $X_1,\dots,X_n$ of the Lie algebra $\mathfrak{n}$ of $N$ we write: $X^I=X_1^{i_1}\dots X_n^{i_n}$ and $|I|=i_1+\dots+i_n$. For $k=0,1,\dots,\infty$ we define: \begin{equation*} C^k=\{f:X^If\in C(N),\text{ for }|I|r>s$. The existence of the family $U^\sigma(s,t)$ follows from \cite{T}. For $\alpha\geq 0$, on a direct product $G=N\times\mathbb{R}^+$ we consider the following operator $$ \mathbf{L}_\alpha=a^{-2}\sum_j \Phi_a(X_j)^2+a^{-2}\Phi_a(X)+\partial_a^2+\frac{2\alpha+1}{a}\partial_a.$$ For $f\in C^\infty_c(G)$, we define on $G$ the following function \begin{equation}\label{rozw} u(t,x,a):=\mathbf{E}_a U^\sigma (0,t)f(x,\sigma(t)), \end{equation} where the expectation is taken with respect to the distribution of the Bessel process $\sigma(t)$ starting from $a$. The following theorem, which gives the formula for the solution of the heat equation for $L_\alpha$ in terms of the evolution on $N$ driven by the Bessel process, is one of the main tool in the proof of our results. \begin{theorem}\label{r} Let $u=u(t,x,a)$ be a function on $G$ defined by \eqref{rozw}. Then \begin{equation}\label{3.2} \mathbf{L}_\alpha u=\partial_tu,\ \ u(0,x,a)=f(x,a). \end{equation} Moreover, there exists the only one bounded from below solution $u$ of \eqref{3.2}. \end{theorem} \begin{proof} For the proof of the first part of Theorem \ref{r} see \cite{DHU}. The uniqueness of the bounded from below solution $u$ follows by some kind of the maximum principle which is a modification of Theorem. 3.1.1 in \cite{SV}. The proof for a diagonal action given in \cite{Phd} can be easily generalized. \end{proof} \section{The main result and its proof.}\label{proof of theorem} In this section we obtain pointwise estimates for derivatives of the Green function \eqref{gf}. For a positive $\delta<1\slash 2$ define \begin{align*} T_\delta=&\{(x,a)\in N\times\mathbb{R}^+:1-\delta0. \end{equation*} It is not difficult to check that although the operator $\mathbf{L}_\alpha$ is not left-invariant it has some homogeneity with respect to the family of dilations introduced above: \begin{equation*} \mathbf{L}_\alpha(f\circ D_t)=t^2\mathbf{L}_\alpha f\circ D_t. \end{equation*} This implies that \begin{equation}\label{homogeneity} G_\alpha(x,a;y,b)=t^{-Q-2\alpha}G_\alpha(D_{t^{-1}}(x,a);D_{t^{-1}}(y,b)). \end{equation} It turns out (see (1.17) in \cite{DHU}) that \begin{equation}\label{g-g} \mathcal G_{-\gamma}(x,a)=G_{\gamma\slash 2}(e,1;x,a)=G^*_{\gamma\slash 2}(x,a;e,1), \end{equation} where $G^*_{\alpha}$ is the Green function for the operator \begin{equation*} \mathbf{L}^*_\alpha=a^{-2}\sum_j\Phi_a(X_j)^2-a^{-2}\Phi_a(X)+\partial_a^2+\frac{2\alpha+1}{a}\partial_a \end{equation*} formally conjugate to $\mathbf{L}_\alpha$ with respect to the measure $a^{2\alpha+1}d ad x$. Moreover, \begin{equation*} G^*_\alpha(x,a;e,1)=\lim_{\eta\to 0}\int_0^\infty\mathbf{E}_1p^\sigma(t,0)(x)m_\alpha(I_{a,\eta})^{-1}1_{I_{a,\eta}}(\sigma_t)d t, \end{equation*} where \begin{equation*} m_\alpha(I)=\int_Ia^{2\alpha+1}da \end{equation*} and the expectation is taken with respect to the distribution of the Bessel process with the parameter $\alpha$ starting from 1, i.e., $\mathop{\rm BESS}_1(\alpha)$ on the space $C((0,\infty),(0,\infty))$, $p^\sigma(t,0)$ is the transition function of the evolution generated by the operator \eqref{operatorsigma} and $I_{a,\eta}=[a-\eta,a+\eta]$. Since $\mathcal L_{-\gamma}(\cdot)=a^{-\gamma}\mathcal L_\gamma(a^\gamma\cdot)$ it follows that \begin{equation}\label{conj} \mathcal G_\gamma(xa,yb)=a^{\gamma}\mathcal G_{-\gamma}(xa,yb)b^{-\gamma} \end{equation} and therefore, by \eqref{g-g} and \eqref{conj}, \begin{equation*} \mathcal G_\gamma(x,a)=G^*_{\gamma\slash 2}(x,a;e,1)a^{\gamma}. \end{equation*} Before we go to the proof of our main result we note the following important proposition which gives estimates on the set $\mathcal Q\setminus T_\delta$ of some functional of the evolution $p^\sigma$ which plays the crucial role in the proof of Theorem \ref{main}. \begin{proposition}\label{mainproposition} i) For every $1>\delta>0$ and for every multi-index $I$ such that $|I|>0$ there exists a constant $C>0$ such that for every $(x,a)\in(\mathcal{Q}\setminus T_\delta)\cap\{(x,a)\in N\times\mathbb{R}^+:a\leq 1-\delta\}$ and for every $0\leq l\leq k-1$, \begin{equation*} \sup_{0<\eta<\delta\slash 2}|\int_0^\infty\mathbf{E}_1 X^Ip^\sigma(t,0)(x)\partial_a^lm_\alpha(I_{a,\eta})^{-1}\partial_a^{k-l}1_{I_{a,\eta}}(\sigma_t)dt|\leq Ca^{-k}. \end{equation*} ii) For every $0<\chi_0\leq 1,$ $00$ there exists a constant $C>0$ such that for every $\chi\leq\chi_0,$ for every $(x,a)\in\{0\delta>1\slash 2$ and for every multi-index $I$ such that $|I|>0$ there exists a constant $C>0$ such that for every $\chi\leq 1\slash 2-\delta,$ for every $(x,a)\in\{(1-\delta)\slash 2\leq a\leq 1\slash 2\}$ and for every $0\leq l\leq k-1$, \begin{equation*} \sup_{0<\eta<\delta\slash 4}|\int_0^\infty\mathbf{E}_\chi X^Ip^\sigma(t,0)(x)\partial_a^lm_\alpha(I_{a,\eta})^{-1}\partial_a^{k-l}1_{I_{a,\eta}}(\sigma_t)dt|\leq Ca^{-k}. \end{equation*} \end{proposition} \begin{proof}[Sketch of the proof] The case $\alpha=0$ has been proved in \cite{ejde}. (See the proof of Proposition 3.1 in \cite{ejde}. We take the opportunity to say that the formulation of Proposition 3.1 in \cite{ejde} is wrong and that it should have been stated exactly as above with $\alpha=0.$) The generalization from $\alpha=0$ to an arbitrary $\alpha>0$ is (almost) straightforward. One only needs to imitate the proof of Proposition 3.1 in \cite{ejde}. \end{proof} After this preparatory facts we are ready to give \begin{proof}[Sketch of the proof of Theorem \ref{main}] Let $0<\delta<1\slash 2$, $k\in\mathbb{Z}^+$ and a multi-index $I$ be fixed. \noindent \textbf{Case 1.} We consider the set \begin{equation*} S_1=\mathcal{Q}\setminus T_\delta. \end{equation*} By Proposition \ref{mainproposition} and Theorem \ref{r} it follows that there exists a positive constant $C$ such that \begin{equation}\label{c1} |\partial_a^k\mathcal X^IG^*_{\gamma\slash 2}(x,a;e,1)|\leq Ca^{-k} \end{equation} for every $(x,a)\in \tilde S_1:=S_1\cap\{(x,a)\in N\times\mathbb{R}^+:a\leq 1-\delta\}$. But $S_1\setminus\text{Int}\tilde S_1$ is a compact set and $G^*_{\gamma\slash 2}$ is a continuous function so we get \eqref{c1} on $S_1$. Therefore, on $S_1$ we have that \begin{equation*} |\partial_a^k\mathcal X^I\mathcal G_{-\gamma}(x,a)|=|\partial_a^k\mathcal X^IG^*_{\gamma\slash 2}(x,a;e,1)|\leq Ca^{-k} \end{equation*} and \begin{equation*} |\partial_a^k\mathcal X^I\mathcal G_{\gamma}(x,a)|=|\partial_a^k\mathcal X^IG^*_{\gamma\slash 2}(x,a;e,1)a^{\gamma}|\leq Ca^{\gamma-k}. \end{equation*} \noindent \textbf{Case 2.} Outside the set $S_1$ we imitate the proof of Theorem 3.1 in \cite{amuc}, where derivative with respect to $\mathcal{X}$ have been considered. The homogeneity \eqref{homogeneity} of the Green function plus \eqref{den} play the crucial role. \end{proof} \begin{thebibliography}{10} \bibitem{A} A.~Ancona. \newblock Negatively curved manifolds, elliptic operators, and the {M}artin boundary. \newblock {\em Ann. of Math.}, 125: 495--536, 1987. \bibitem{CC} C.~Constantinescu and A.~Cornea. \newblock {\em Potential theory on harmonic spaces}. \newblock Springer-Verlag, New York, 1972. \newblock With a preface by H. Bauer, Die Grundlehren der mathematischen Wissenschaften, Band 158. \bibitem{Borodin} A.N. Borodin and P.~Salminen. \newblock {\em Handbook of {B}rownian motion - facts and formulae}. \newblock Birkh{\"{a}}user Verlag, 1996. \bibitem{D} E.~Damek. \newblock Fundamental solutions of differential operators on homogeneous manifolds of negative curvature and related {R}iesz transforms. \newblock {\em Coll. Math}, 73(2): 229--249, 1997. \bibitem{DHU} E.~Damek, A.~Hulanicki, and R.~Urban. \newblock Martin boundary for homogeneous {R}iemannian manifolds of negative curvature at the bottom of the spectrum. \newblock {\em Rev. Mat. Iberoamericana}, 17(2): 257--293, 2001. \bibitem{DHZ} E.~Damek, A.~Hulanicki, and J.~Zienkiewicz. \newblock Estimates for the {P}oisson kernels and their derivatives on rank one {$NA$} groups. \newblock {\em Studia Math.}, 126(2): 115--148, 1997. \bibitem{FS} G. B. Folland and E.~Stein. \newblock {\em Hardy spaces on homogeneous groups}. \newblock Princeton Univ. Press, 1982. \bibitem{H} E.~Heintze. \newblock On homogeneous manifolds of negative curvature. \newblock {\em Math. Ann.}, 211: 23--34, 1974. \bibitem{Stamp} D.~Kinderlehrer and G.~Stampacchia. \newblock {\em An introduction to variational inequalities and their applications}. \newblock Academic Press, 1980. \bibitem{L} N. N. Lebedev. \newblock {\em Special functions and their applications}. \newblock Dover Publications, Inc., 1972. \bibitem{RY} D.~Revuz and M.~Yor. \newblock {\em Continuous martingales and {B}rownian motion}. \newblock Springer-{V}erlag, 1991. \bibitem{T} H.~Tanabe. \newblock {\em Equations of evolutions}. \newblock London, Pitman, 1979. \bibitem{SV} D.~Stroock and S.R. Varadhan. \newblock {\em Multidimensional difussion processes}. \newblock Springer, 1979. \bibitem{amuc} R.~Urban. \newblock Estimates for the derivatives of the {G}reen functions on homogeneous manifolds of negative curvature. \newblock {\em Acta Math. Univ. Comenianae (N.S.)}, 73(1): 43--53, 2004. \bibitem{pota} R.~Urban. \newblock Estimates for derivatives of the {G}reen functions for the noncoercive differential operators on homogeneous manifolds of negative curvature. \newblock {\em Potential {A}nalysis}, 19(4): 317--339, 2003. \bibitem{ejde} R.~Urban. \newblock Estimates for derivatives of the {G}reen functions for the noncoercive differential operators on homogeneous manifolds of negative curvature, {II}. \newblock {\em Electron. J. Differential Equations}, 2003(86): 1--8, 2003. \bibitem{Phd} R.~Urban. \newblock {\em Estimates for the {P}oisson kernels on {$NA$} groups. A probabilistic method}. \newblock PhD thesis, Wroclaw {U}niversity, 1999. \bibitem{pams} R.~Urban. \newblock Estimates for the {P}oisson kernels on homogeneous manifolds of negative curvature and the boundary {H}arnack inequality in the noncoercive case. \newblock {\em Probab. Math. Statist.}, 21(1): 213--229, 2001. \bibitem{CM} R.~Urban. \newblock Noncoercive differential operators on homogeneous manifolds of negative curvature and their {G}reen functions. \newblock {\em Colloq. Math.}, 88(1): 121--134, 2001. \bibitem{MathZ} R.~Urban. \newblock Estimates for derivatives of the {P}oisson kernels on homogeneous manifolds of negative curvature. \newblock {\em Math. Z.}, 240(4): 745--765, 2002. \end{thebibliography} \end{document}