1: %\documentclass[prl, showpacs, twocolumn, floatfix]{revtex4}
2: \documentclass[pre, showpacs, floatfix, preprint, endfloats*]{revtex4}
3: %\documentclass{article}
4:
5: \usepackage{graphicx}
6: \usepackage{amsmath, amsfonts, amssymb, bm}
7: \usepackage[]{psfrag}
8:
9: \psfragscanoff
10: \setlength{\parindent}{0pt}
11:
12: \begin{document}
13:
14: \title{Positronium in intense laser fields}
15: \author{Bj\"orn \surname{Henrich}}
16: \email{henrich@physik.uni-freiburg.de}
17: \author{Karen Z. \surname{Hatsagortsyan}}
18: \email{khats@physik.uni-freiburg.de}
19: \author{Christoph H. \surname{Keitel}}
20: \email{keitel@uni-freiburg.de}
21: \affiliation{Theoretische Quantendynamik, Physikalisches Institut,
22: Universit\"at Freiburg, Hermann-Herder Str. 3, D-79104 Freiburg, Germany}
23:
24:
25: \date{\today}
26:
27: \begin{abstract}
28: The dynamics and radiation of positronium is investigated in intense laser fields. %, with emphasis placed on
29: %peculiaries due to the equal masses and opposite charges of its two constituents.
30: Our two-body quantum mechanical treatment displays the tunneling, free-evolution and recollision
31: dynamics of electron and positron both in the oscillating laser electric and laser magnetic field
32: components. In spite of significant momentum transfer of the numerous incoming laser photons,
33: recollisions of both particles are shown to occur automatically after tunneling ionization, along
34: with substantial x-ray and gamma-ray emission during recombination and annihilation processes.
35: \end{abstract}
36:
37: \pacs{42.50.Hz, 42.65.Ky, 36.10.Dr, 78.70.Bj}
38:
39: \maketitle
40:
41: The highly nonlinear interaction of gaseous atoms with super intense laser pulses has been demonstrated
42: to give rise to the emission of coherent high-harmonic generation (HHG) up to the x-ray regime, with numerous
43: applications in high-resolution spectroscopy and diagnostics (see e.g. \cite{hhg}).
44: Since the kinetic energy and consequently the frequency of the emitted radiation of a particle increases
45: with rising laser intensity, considerable interest has been directed towards understanding the complex
46: relativistic quantum dynamics of atoms and ions in ultra intense laser pulses \cite{Protopapas}.
47: Once electrons reach velocities nonnegligible to that of light, however, the magnetically induced Lorentz
48: force or in other words the momentum transfer of the numerous incoming photons induce a separation of
49: electrons and ionic core in the laser propagation direction, resulting in a strong reduction
50: of high harmonic yields \cite{Protopapas,Reiss,Walser}. While highly charged ions and crystals were
51: studied towards a reduction of this effect \cite{cry}, there is still a clear lack for an efficient system, where
52: radiation pressure does not induce substantial ionization in the laser propagation direction and thus reduce coherent
53: high frequency generation. On a different front there is also a quest for physics beyond atomic and classical
54: plasma physics in ultra intense lasers, such as nuclear reactions \cite{ditmire} and QED effects \cite{qed}.
55:
56: In this letter the dynamics and high harmonic generation of positronium is investigated in intense laser fields.
57: The two-body system is shown to display unique properties: While tunnel-ionization of electron and positron may
58: occur almost oppositely in the laser polarisation direction, both particles sense the identical drift in the
59: laser propagation direction due to their equal magnitudes of mass and charge (see Fig. \ref{diagram}).
60: Periodic electron-positron recollisions are shown to occur automatically head-on in spite of the influence of
61: the Lorentz force. %, along with substantial high harmonic generation in the x-ray regime.
62: In addition to substantial coherent x-ray generation during periodic electron-positron recombinations we predict
63: gamma radiation in the less likely events of laser-enhanced annihilations of both particles.
64:
65: \begin{figure}[b]
66: %\vspace{-2.0cm}
67: %\psfrag{B}{${\bf H}$}
68: %\psfrag{E}{${\bf E}$}
69: %\psfrag{k}{${\bf k_L}$}
70: %\psfrag{p}{$p$}
71: %\psfrag{e}{$e$}
72: \includegraphics[width=12.8cm]{diagram.eps}
73: \caption{\label{diagram}Schematic diagram displaying positronium dynamics in an
74: intense laser field. The bound system depicted by the density of its
75: wave function may ionize in the laser field. Once free, both electron $e$ and
76: positron $p$ could be described as classical particles. Their trajectories are
77: shown by the solid lines. The electric field ${\bf E}$ accelerates both particles in
78: opposite directions while the Lorentz force due to the magnetic field ${\bf H}$
79: leads to an identical drift in the propagation direction ${\bf k_L}$. Without an initial
80: center-of-mass motion these trajectories are symmetric and thus both particles
81: overlap periodically giving rise to possible recombinations and annihilations and thus
82: high-frequency light emission.
83: }
84: \end{figure}
85:
86:
87: Positronium consists of an electron and a positron and is known to be unstable.
88: While ortho-positronium annihilates into three photons with a lifetime of
89: $1.4\cdot 10^{-7}$s, para-positronium does so with two photons and a lifetime of
90: $1.25\cdot 10^{-10}$s \cite{Landau}.
91: The presence of a strong laser field may induce substantial reductions \cite{Mittleman}
92: or enhancements \cite{Rivlin,Ritus} of the annihilation process into gamma photons.
93: However, even for the shorter lifetime of para-positronium it is sufficient for the interaction
94: with many cycles of a femto second laser pulse with interaction lengths not exceeding
95: much the centimeter range \cite{Rivlin}.
96: %For ortho-positronium, the lifetime shall not impose any serious restrictions.
97:
98: The laser field shall be described by the vector potential $\mathbf{A}(\mathbf{x},t)$ propagating in the z direction
99: and linearly polarized along the x direction. Being interested in the tunneling regime with
100: moderately intense laser field strengths we restrict ourselves to Schr\"odinger dynamics in two
101: dimensions, though beyond the dipole approximation. The quantum dynamics of positronium in the laser field,
102: taking fully into account of the Lorentz force due to the laser magnetic field component,
103: is thus governed by the following Hamiltonian (in atomic units as throughout the article):
104: \begin{eqnarray}
105: H=\frac{\left(-i\nabla_{\mathbf{x}_e}-\frac{\mathbf{A}(t-z_e/c)}{c}\right)^2}{2}
106: +\frac{\left(-i\nabla_{\mathbf{x}_p}+\frac{\mathbf{A}(t-z_p/c)}{c}\right)^2}{2}
107: \nonumber \\
108: -\frac{1}{|\mathbf{x}_e-\mathbf{x}_p|} \hspace{2.82cm}
109: \label{e1}
110: \end{eqnarray}
111: where, $\mathbf{x}_i$ and $-i\nabla_{\mathbf{x}_i}$, ($i\in\{e,p\}$),
112: are the operators of the coordinates and momenta of the electron and positron, respectively.
113: Further we introduce relative $\mathbf{r}=(x,y,z)=\mathbf{x}_e-\mathbf{x}_p $ and center of mass
114: $\mathbf{R}=(X,Y,Z)=\frac{\mathbf{x}_e+\mathbf{x}_p}{2}$
115: coordinates. In accordance with the Ehrenfest theorem, the center of mass transversal canonical momenta
116: $-i\partial_X$, $-i\partial_Y$ as well as $i\left(\partial_t+c\partial_Z\right)$ turn out to be
117: conserved quantities, i.e. commute with $H$. The eigenvalues are defined to be $P_x, P_y, g$, respectively,
118: and $\cal E$ is a separation constant that can be understood as the energy of the system before the interaction.
119: Therefore, we consider the ansatz
120: \begin{equation}\Phi(\mathbf{r},\mathbf{R},\tau)=\exp\left(i\left(P_xX+P_yY-\frac{g}{c}Z-\cal
121: E\tau\right)\right)
122: \phi(\mathbf{r},\tau) \label{e3} \end{equation}
123: singling out eigenfunctions of the conserved quantities in the wave function and introducing
124: the running time $\tau=t-\frac{Z}{c}$.
125: Employing a $1/c$ expansion of the vector potential as function of the coordinates
126: of the relative motion of the particles, we obtain the following equation for $\phi$:
127: \begin{eqnarray}i\left(1-B_z\right)\partial_{\tau}\phi(\mathbf{r},\tau)=
128: \left\{\left(\frac{1}{i}\nabla_{\mathbf
129: r}-\frac{\mathbf{A}(\tau)}{c}\right)^2+\frac{\mathbf{P}^2}{4}\right.\nonumber\\
130: \left.+
131: \frac{P_x\dot{A}(\tau)z}{2c^2} -\frac{1}{r}-{\cal E}-\frac{1}{4c^2}\partial^2_{\tau}
132: \right\}\phi(\mathbf{r},\tau). \label{e4} \end{eqnarray}
133:
134: Here, the center of mass velocity is introduced via $\mathbf{V}=\mathbf{P}/2$ with
135: $\mathbf{B}=\mathbf{V}/c$ and $\mathbf{P}=\left(P_x,P_y,({\cal E} -g)/c\right)$.
136: The applied expansion takes into account of the leading multipole operators
137: describing the magnetically induced relative motion of electron and positron in the laser propagation
138: direction. Since $P_x$ is a conserved quantity,
139: this operator describes an oscillation of the relative coordinate in the
140: propagation direction while for atoms the analogous term would lead to a drift.
141: For consistency, the higher-order term for the center-of-mass motion $-\frac{1}{4c^2}\partial^2_{\tau}$
142: in Eq. (\ref{e4}) will be neglected in the following.
143:
144: We proceed by carrying out a transformation to the length gauge:
145: $\exp(-i\chi(\mathbf{r},\tau))\phi=\Psi$ with
146: $\chi(\mathbf{r},\tau)=\mathbf{A}(\tau)\cdot\mathbf{r}/c.$
147: In this gauge we note the full velocity dependence of the transition matrix
148: elements to be discussed later.
149: The resulting time dependent Schr\"odinger equation reads
150: \begin{eqnarray} & i\left(1-B_z\right)\partial_{\tau}\Psi(\mathbf{r},\tau)=
151: \bigg\{-\nabla^2_{\mathbf{r}} + I_p & \bigg.\nonumber\\
152: \bigg. & -E(\tau)\left(1-B_z\right)x-B_xE(\tau)z -\frac{1}{r}
153: %& \bigg.\nonumber\\
154: %\bigg. & -\frac{1}{4c^2}\Big(\partial_{\tau}-iE(\tau)x \Big)^2
155: \bigg\}\Psi(\mathbf{r},\tau) &
156: \label{e5} \end{eqnarray}
157:
158: with electric field $\mathbf{E}(\tau)=-\frac{1}{c}\partial_{\tau}\mathbf{A}$
159: and positronium ionization potential $I_p$.
160: Further we employ the reasonable assumptions proposed earlier by Lewenstein and
161: co-workers \cite{Lewenstein1} for atomic HHG in the dipole approximation, being:
162: a) The contribution of all bound states except the ground state is neglected;
163: b) The depletion of the ground state is neglected;
164: c) In the continuum the electron is treated as a free particle in the laser field.
165: Then the time dependent wave function can be written as
166: \begin{equation}\left|\Psi(\tau)\right\rangle=
167: \left|0\right\rangle+\int_{-\infty}^{\infty} d^3\tilde{p}b(\tilde{\mathbf{p}},\tau)\left|\tilde{\mathbf{p}}\right\rangle
168: \label{e6} \end{equation}
169: Here $\left|0\right\rangle$ is the ground state
170: and $b(\tilde{\mathbf{p}},\tau)$ denotes the amplitude of the corresponding continuum state
171: $\left|\tilde{\mathbf{p}}\right\rangle$.
172: Integration of Eq. (\ref{e5}) with the ansatz (\ref{e6})
173: %and neglecting all $1/c^2$ and higher order contributions
174: yields for the ionization amplitude $b$:
175: \begin{eqnarray}b(\tilde{\mathbf{p}},\tau)&=&\frac{-i}{1-B_z}\int_{-\infty}^{\tau}
176: \left\langle\mathbf{p}-\mathbf{A}(\tau')/c
177: - \mathbf{\Lambda}(\tau')\right|H_I(\tau')\left|0\right\rangle \nonumber\\
178: & &\times
179: \exp\left(-i S(\mathbf{p},\tau,\tau')\right) d\tau', \label{e7} \end{eqnarray}
180: where the quasiclassical action is
181: $S(\mathbf{p},\tau,\tau')=\int_{\tau'}^{\tau}d\tau''\left\{\left(\mathbf{p}-\mathbf{A}(\tau'')/c
182: -\mathbf{\Lambda}(\tau'')\right)^2+I_p\right\}/(1-B_z)$.
183: Here, the new variable
184: $\mathbf{p}=\tilde{\mathbf{p}}+\mathbf{A}(\tau)/c+\mathbf{\Lambda}(\tau)$ is
185: introduced with $\mathbf{\Lambda}(\tau)=\Lambda(\tau)\mathbf{e}_z=
186: B_xA(\tau)/\left[c\left(1-B_z\right)\right]\mathbf{e}_z$
187: and $H_I(\tau)=-E(\tau)x(1-B_z)-B_xE(\tau)z$.
188: %We note that $b$ still contains small higher order terms in $1/c$
189: %for the sake of compactness of the expression which however will be dropped in
190: %the final result to be derived.
191: In solving the equation for $b$ we have neglected all terms
192: involving continuum-continuum transitions of no interest for the
193: high-frequency part of harmonic generation as in \cite{Lewenstein1}.
194:
195: We calculate the dipole moment $\mathbf{x}_d(\tau)=\left\langle\Psi(\tau)\right|\mathbf{r}
196: \left|\Psi(\tau)\right\rangle$ of which the Fourier transform $\mathbf{x}_d([\omega-\mathbf{k \cdot V}]/[1-B_z])$
197: yields the Doppler shifted harmonic spectrum, where $\omega$ and ${\bf k}$ denote the emitted radiation frequency
198: and wave vector, respectively.
199: In the tunneling regime the integrals can be obtained using the saddle point method,
200: $\nabla_{\mathbf{p}}S\left(\mathbf{p},\tau,\tau'\right)=0$. Sustaining all momenta up to first order in $1/c$ yields:
201: $p_x(\tau,\tau')=\int_{\tau'}^{\tau}d\tau''A(\tau'')/{c(\tau-\tau')},p_y=0, p_z(\tau,\tau') = B_xp_x(\tau,\tau').$
202: The matrix elements can be evaluated analytically employing the relation \cite{Bethe}
203: $\mathbf{d}(\mathbf{p})=(d_x,d_y,d_z)=\left\langle\mathbf{p}\right|\mathbf{r}
204: \left|0\right\rangle=4 i\left(I_p\right)^{5/4} \mathbf{p}/(\pi \left(\mathbf{p}^2+I_p \right)^3)$
205: and performing the integration with respect to $\mathbf{p}$ gives
206: \begin{eqnarray}x_d(\tau)&=&\frac{1}{\sqrt{i}}
207: \int_{-\infty}^{\tau}d\tau'\left(\frac{\pi}{\tau-\tau'}\right)^{3/2}\nonumber\\
208: & & \times d_x^{\ast}\left(\mathbf{p}(\tau,\tau')-\mathbf{A}(\tau)/c - \mathbf{\Lambda}(\tau) \right)\nonumber\\
209: & & \times E(\tau')d_x\left(\mathbf{p}(\tau,\tau')-\mathbf{A}(\tau')/c - \mathbf{\Lambda}(\tau') \right)\nonumber\\
210: & & \times \exp\left(-iS(\mathbf{p}(\tau,\tau'),\tau,\tau')\right)+c.c.,
211: \label{e8a}\end{eqnarray}
212: with classical action reading $S(\mathbf{p}(\tau,\tau'),\tau,\tau')=\int_{\tau'}^{\tau}d\tau''
213: \big[\left( \mathbf{p}(\tau,\tau')-\mathbf{A}(\tau'')/c - \mathbf{\Lambda}(\tau'') \right)^2+I_p\big]/
214: \left(1-B_z\right)$.
215: Taking into account the pole in the $\tau'$ integration finally results in
216: \begin{eqnarray}x_d(\tau) & = &
217: \frac{1}{\sqrt{i}}\sum_{\tau_0}^{}\left(\frac{\pi}{\tau-\tau_0}\right)^{3/2}\frac{I_p^{3/4}}{2E(\tau_0)}\nonumber\\
218: & & \times d_x^*\Big(p_x(\tau,\tau_0)-\frac{A(\tau)}{c}\Big)\nonumber\\
219: & & \times \exp\Big\{-\frac{2}{3}\frac{I_p^{3/2}}{E(\tau_0)}(1+B_z)\Big\}\nonumber\\
220: & & \times \exp \big\{-i S\big(\mathbf{p}(\tau,\tau_0),\tau,\tau_0\big) \big\}+c.c..
221: \label{e8}\end{eqnarray}
222: %with $S(\mathbf{p}(\tau,\tau_0),\tau,\tau_0)=\int_{\tau_0}^{\tau}d\tau''
223: %\Big\{\big[\left(p_x(\tau,\tau_0)-A(\tau'')/c\right)^2+I_p\big](1-B_z)
224: %+\left(p_z(\tau,\tau_0)-\Lambda(\tau'')\right)^2\Big\}$.
225: The birth times $\tau_0$ are determined, for given $\tau$, by the condition $p_x(\tau,\tau_0)=A(\tau_0)/c$.
226: The leading terms in $1/c$ have been sustained only in the phases because they play a far
227: less substantial role in the prefactor being identical to the dipole case.
228:
229: The dipole moment in Eq. (\ref{e8}) displays essential differences in comparison with atomic HHG
230: which shows especially in the real exponential factor in Eq. (\ref{e8}):
231: $C:=\exp[- 2 I_p^{3/2} (1+B_z)/(3 E(\tau_0))]$.
232: We stress that in the atomic HHG case \cite{Walser}
233: there is an additional contribution exchanging $I_p$ by $I_p^a + c^{2}\xi ^{4}/2$
234: due to the drift in the laser propagation direction which can strongly reduce the amplitude
235: and thus the efficiency of HHG. Here $\xi=\sqrt{2} A /c^2$ and $I_p^a$ denotes the atomic ionization
236: potential. The expression $C$, without the factor $1+B_z$, coincides with
237: the corresponding result within the dipole approximation for atoms.
238: This means that for positronium, the magnetically-induced drift in the laser propagation direction has no
239: diminishing impact on HHG.
240: %The factor $1+B_z$ in the exponent is due to the Lorentz transformation
241: %of the electric field in the center-of-mass system of reference.
242:
243: \begin{figure}[b]
244: %\psfrag{xx}[c]{}
245: %\psfrag{x}[c]{harmonic order}
246: %\psfrag{y}[c]{$log|\omega^2 x_d(\omega)|$(arb.u.)}
247: %\psfrag{a}[c]{a)}
248: %\psfrag{b}[c]{b)}
249: %\psfrag{y1}[c]{$10^{-3}$}
250: %\includegraphics[width=5.5cm]{henrichfig2a.eps}
251: \includegraphics[width=11.0cm]{spectra.eps}
252: \caption{\label{spektrum1}High harmonic spectra of positronium in the laser propagation direction
253: in a laser pulse with vector potential $A=194.0$a.u. and angular frequency $\omega=0.0091$a.u.
254: (corresponding to a laser intensity of 8.5 $\times 10^{12}$ W/cm${^2}$
255: and the second harmonic of a CO$_2$ laser with wavelenth 10 $\mu m$) and pulse shape parameters $c_1=1.0$ and $c_2=5.0$.
256: The symbols $c_1$ and $c_2$ denote the number of cycles of ${\rm sin}^2$ turn-on/turn-off phases and of the intermediate cycles at full amplitude, respectively. The initial positronium velocity is
257: for a) $V_x=V_z=0.0001$a.u. and for b) $V_x=V_z=13.7$a.u. corresponding to 10$\%$ of the speed of
258: light.
259: }
260: \end{figure}
261:
262:
263: The figures \ref{spektrum1} display harmonic spectra in the
264: laser propagation direction following the numerical integration of the dipole via Eq. (\ref{e8a}).
265: In agreement with the numerical results, the maximal harmonic frequency $\omega_{max}$ is shown to critically
266: depend on both the center-of-mass velocity and the observation direction via the Doppler shift with
267: $\omega_{max}(1-\mathbf{k \cdot V}/|\mathbf{k}|) = n_{max} \omega_L (1-B_z) = 3.17 U_p + I_p$ for
268: maximal harmonic number $n_{max}$ and $U_p=c^{2}\xi ^{2}/2$.
269:
270:
271:
272: In order to quantify the advantage of positronium for HHG in comparison to conventional HHG in atomic gas jets,
273: we consider an example. In Lithium vapor being irradiated by a laser field at $\lambda =0.75 \mu $m wavelength
274: and intensity $I=7\cdot 10^{16}$W/cm$^{2}$ ($\xi=0.17$), HHG up to the 10 keV photon energy
275: range can be achieved in the tunneling regime for Li$^{2+}$ with $I_{p}^a=122.88$ eV.
276: Magnetic field effects in this case, however, are problematic as the extent of the electron drift during
277: the laser period exceeds the atomic size: $\lambda \xi ^{2}/16\pi \gg a_B$, where $a_{B}$ is the Bohr radius.
278: This substantially reduces the HHG efficiency, because of the earlier discussed $c^{2}\xi^4$ term being absent
279: in positronium. Therefore, at $\xi =0.12$ (e.g. $\lambda \approx 50 \mu $m and
280: $I=8\cdot 10^{12}$W/cm$^{2}$), the HHG cut-off frequency for positronium is equal to that of Li$^{2+}$ at the
281: above mentioned parameters, however with a yield 15 orders of magnitude larger.
282: Thus, the low densities of positronium may be compensated for by the absence of drift-reduced recollisions.
283: %It means that more efficient HHG with positronium can be realized in much low densities avoiding as well
284: %the phase mismatching problem caused by free electrons (positrons).
285:
286: Finally we discuss the possibility of $\gamma $-ray emission with positronium via
287: laser induced annihilation. In a super strong laser field with $\xi \gtrsim (mc^{2}/\hbar \omega )^{1/3} \approx 10^{2}$,
288: annihilation can occur via emission of $n \gtrsim 10^{6}$ photons \cite{Ritus} along with one $\gamma $-quantum (Fig. 3a).
289: %but the dynamics of positronium in such high intensity fields is out of the scope of the present letter.
290: In more moderate laser intensities in the tunneling regime with $\xi \leq 1$ of interest here,
291: ortho-positronium, may annihilate into two $\gamma $-quanta (with induced emission of an odd number of laser photons, Fig. 3b)
292: or spontaneously into three photons (Fig. 3c)).
293:
294: We proceed with an order-of-magnitude estimation of $\gamma $-ray emission via strong laser
295: assisted annihilation of ortho-positronium as depicted in Fig. 3b in comparison to spontaneous 3-photon annihilation (Fig. 3c).
296: %The laser-induced annihilation can be described by a 2-vortex diagram with electron-positron states being dressed by the laser field.
297: The 2-vortex diagram in Fig. 3b, at first with bare particles, can be evaluated as product of the Thomsonian
298: cross-section $\sigma \approx\pi r_{0}^{2}$ with $r_{0}$ being the classical electron radius \cite{Landau} and the particle flux
299: to yield the annihilation probability $W\approx \sigma \cdot \rho c$, with electron and positron density $\rho =a^{-3}_B$.
300: The annihilation process is considered in a strong laser field here, with laser frequency being low in comparison
301: to the characteristic frequency $\omega_c=c^2$. In this situation the probability for bare particles may be multiplied by
302: a factor $F_s$ which takes into account of $s$ laser photons being emitted during the annihilation process (analogous to
303: Low theorem \cite{Low}).
304: The laser-induced differential probability of $\gamma $-ray emission at frequency $\omega_c$
305: being accompanied by the emission of $s$ laser photons can then be written as:
306: $dW_{ind}\approx \pi r_{0}^{2}\rho c(d\omega /\Delta \omega _{D})(d\Omega /4\pi )\sum_s F_s$.
307: Here we assumed that the $\gamma$-emission is distributed within a $4\pi $ solid angle $\Omega$ and that the Doppler
308: width $\Delta \omega _{D}$ describes the leading spectral broadening. Further
309: $F_s=A_{0}^{2}(s\alpha \beta )+\xi ^{2}(2u-1)[A_{1}^{2}(s\alpha \beta )-A_{0}(s\alpha \beta )A_{2}(s\alpha \beta )]$ \cite{Ritus}
310: is represented via generalized Bessel functions
311: $A_{0}(s\alpha \beta ), A_{1}(s\alpha \beta ), A_{2}(s\alpha \beta )$, $\alpha =k^{'}A/(kq)c$, $\beta =-kk^{'}A^{2}/(8c^2(kq)^{2})$,
312: $u=(kk^{'})^{2}/(4(kq)(kq^{'}))$, where $q_{\mu }=p_{\mu }-k_{\mu }A^{2}/(4kp)c^2$, $q^{'}_{\mu }=p^{'}_{\mu }-k_{\mu }A^{2}/(4kp)c^2$
313: for $\{\mu\}\in \{0,1,2,3\}$ and
314: $k, k^{'}, p, p^{'}$ denote the laser photon, $\gamma $-quanta, electron and positron 4-momenta, respectively. $A$ is the
315: 4-vector-potential. In a weak laser field with $\xi \ll 1$ our formula for $dW_{ind}$ matches the known result in \cite{Rivlin}.
316: With a similar heuristic argumentation we may estimate the spontaneous 3-photon annihilation probability
317: without laser field at $\omega_c$ depicted in Fig. 3c, to give $dW_{sp}\approx \alpha \pi r_{0}^{2}\rho c(d\omega /\omega_c)(d\Omega /4\pi )$
318: with fine structure constant $\alpha=1/c$. This result for Fig. 3c , however, is well-known from the literature \cite{Landau}.
319:
320:
321: \begin{figure}[t]
322: %\psfrag{x}[c]{Harmonic order}
323: %\psfrag{y}[c]{$|\omega^2 x_d(\omega)|$ (arb.u.)}
324: \includegraphics[width=14cm]{feynman-graphen.eps}
325: \caption{\label{feynman}
326: Feynman diagrams displaying $\gamma$-photon emission via electron-positron annihilation by
327: a) multiphoton annihilation in super strong laser fields at $\xi \gtrsim 100$ with emission of
328: $n\gtrsim 10^{6}$ laser photons along with one $\gamma $-quantum;
329: b) annihilation in a laser field at $\xi \ll 100$ with emission of laser photons along with two
330: $\gamma $-quanta;
331: c) annihilation of ortho-positronium with emission of three photons without laser field.
332: Bold lines correspond to the electron (positron) Volkov states, i.e. in the presence of the laser field,
333: dashed lines to the emitted photons.}
334: \end{figure}
335:
336:
337: %For comparison we estimate also the spontaneous 3-$\gamma $-quanta annihilation probability.
338: %It is described by a 3-vortex diagram. 2-vortex diagram brings the Thomsonian cross-section and an additional photon
339: %emission does an additional factor $\alpha $. Assuming also that emission spans the spectral region up to photon energy of $mc^{2}$,
340: %the differential probability of $\gamma $-ray emission during the spontaneous
341: %annihilation by order of magnitude can be represented as
342: %$dW_{sp}\simeq \alpha \pi r_{0}^{2}\rho c(d\omega /mc^{2})(d\Omega /4\pi )$, that agrees with the known result \cite{Landau}
343: %supporting our heuristic approach.
344:
345: For moderately intense laser fields with $\xi \ll 1$, the main contribution to $dW_{ind}$ arises from
346: $F_1\approx \xi ^{2}$, and at $\xi \approx 1$, for particular $s$ there are arguments with
347: $F_s\approx 1$. In this latter case, the induced emission $dW_{ind}$ can exceed the spontaneous background $dW_{sp}$
348: up to a factor of $(\omega_c/\Delta \omega _{D})/\alpha \approx 10^{6}$ for a Doppler linewidth
349: $\Delta \omega _{D}/\omega_c\approx \sqrt(k_{B}T/\omega_c)\approx 10^{-4}$ with Boltzmann constant $k_{B}$
350: and temperature $T$. For para-positronium, however, a similar analysis shows that the intensity of
351: $\gamma $-radiation via laser-induced annihilation can not exceed that via spontaneous annihilation.
352: %To have an idea of absolute
353: %magnitude of the $\gamma $-radiation due to laser induced annihilation we calculate the maximal attainable brilliance.
354: Regarding the $\gamma $-ray yield via laser-induced annihilation, a brilliance at photon energy $0.5$ MeV
355: of $B\simeq 10^{11}$ $photons/(s\cdot mm^{2}mrad^{2}(0.1\% bandwidth))$ can be estimated for
356: ortho-positronium density of $10^{10}$cm$^{-3}$ and an interaction length of $1$ cm.
357: This appears clearly detectable though moderate as compared to large-scale $\gamma $-ray
358: sources in the same spectral region
359: %, see e.g. the undulator
360: %radiation brilliance at the same frequency of $10^{19}$ $photon/(s\cdot mm^{2}mrad^{2}(0.1\% bandwidth))$ in
361: \cite{Gizzi}.
362:
363: Concluding, positronium was shown to provide various promising dynamical features in strong laser fields such as
364: energetic head-on recollisions of its constituents, with substantial coherent x-ray generation and $\gamma $-ray emission.
365:
366: This work has been supported by the German science foundation (Nachwuchsgruppe within SFB276) and an Alexander-von-Humboldt
367: fellowship for KZH.
368: We benefited from help on numerical issues by D.Diehl and A.Staudt, and especially from a seminal discussion
369: with D. Habs which initiated this project.
370:
371:
372: \begin{thebibliography}{99}
373: %\bibitem{hhg}
374: %A. Rundquist, C.G. Durfee, Z.H. Chang, C. Herne, S. Backus, M. M. Murnane,
375: %H. C. Kapteyn, Science {\bf 280}, 1412 (1998).
376: \bibitem{hhg}
377: %Extreme ultraviolet interferometry measurements with high-order harmonics
378: D. Descamps et al., %, Lynga C, Norin J, L'Huillier A, Wahlstrom CG, Hergott JF, Merdji H, Salieres P, Bellini M, Hansch TW
379: Opt. Lett. {\bf 25}, 135 (2000);
380: J. F. Hergott et al., % Salieres P, Merdji H, et al.
381: % XUV interferometry using high-order harmonics: Application to plasma diagnostics
382: Las. Part. Beams {\bf 19}, 35 (2001);
383: M. Bauer et al.,
384: %C.Lei, K. Read, R. Tobey, J. Gland, M.M. Murnane, and H.C. Kapteyn,
385: Phys. Rev. Lett. {\bf 87}, 025501 (2001).
386: \bibitem{Protopapas}
387: M. Protopapas et al., Rep. Prog. Phys. {\bf60}, 389 (1997);
388: C. J. Joachain et al., Adv. At. Mol. Phys. {\bf 42}, 225 (2000);
389: C. H. Keitel, Contemp. Phys. {\bf 42}, 353 (2001);
390: A. Maquet and R. Grobe, J. Mod. Opt. {\bf 49}, 2001 (2002).
391: \bibitem{Reiss} U.W. Rathe et al., J. Phys. B {\bf 30}, L531 (1997);
392: H. R. Reiss, Phys. Rev. A {\bf 63}, 013409 (2000);
393: M. J. Kylstra et al., Phys. Rev. Lett. {\bf 85}, 1835 (2000);
394: M. Dammasch et al., Phys. Rev. A {\bf 64}, 061402 (2001);
395: J. R. V\'azquez de Aldana et al., Phys. Rev. A {\bf 64}, 013411 (2001);
396: C. C. Chiril\v{a} et al., Phys. Rev. A{\bf 66}, 063411 (2002).
397: \bibitem{Walser} M. W. Walser et al., Phys. Rev. Lett. {\bf 85}, 5082 (2000).
398: \bibitem{cry} K. Z. Hatsagortsyan and C. H. Keitel, Phys. Rev. Lett. {\bf 86}, 2277 (2001);
399: D. B. Milo\v{s}evi\'{c}, S. X. Hu, and W. Becker, Phys. Rev. A {\bf 63}, 011403(R) (2001);
400: C. H. Keitel and S. X. Hu, Apl. Phys. Lett. {\bf 80}, 541 (2002).
401: \bibitem{ditmire} G. Pretzler et al., Phys. Rev. E {\bf 58}, 1165 (1998);
402: T. Ditmire et al., Nature {\bf 398}, 489 (1999); K. W. D. Ledingham et al.,
403: Phys. Rev. Lett. {\bf 84}, 1459 (2000); N. Izumi et al., Phys. Rev. E {\bf 65}, 036413 (2002);
404: G. Grillon et al., Phys. Rev. Lett. {\bf 89}, 065005 (2002).
405: \bibitem{qed} C. Bula et al. Phys. Rev. Lett. {\bf 65}, 3116 (1996);
406: C. H. Keitel et al., J. Phys. B {\bf 31}, L75 (1998); C. Bamber et al., Phys. Rev. D {\bf 60}, 092004 (1999).
407: \bibitem{Landau} V. B. Berestetskii, E. M. Lifshitz, L. P. Pitaevskii,
408: {\it Relativistic Quantum Theory} (Pergamon Press, New York, 1971).
409: \bibitem{Mittleman} M. M. Mittleman, Phys. Rev. A {\bf 33}, 2840 (1986).
410: \bibitem{Rivlin} L. A. Rivlin, Quantum Electron. {\bf 6}, 1313 (1976).
411: \bibitem{Ritus} H. R. Reiss, J. Math. Phys. {\bf 3}, 59 (1962);
412: A. I. Nikishov and V. I. Ritus, Sov. Phys. JETP {\bf 19}, 529 (1964).
413: %\bibitem{Lewenstein2} M. Lewenstein et al., Phys. Rev. A {\bf 51}, 1495(1995).
414: \bibitem{Lewenstein1} M. Lewenstein et al., Phys. Rev. A {\bf 47}, 2117(1994).
415: \bibitem{Bethe} H. A. Bethe and E. E. Salpeter, {\it Quantum Mechanics of One and
416: Two Electron Atoms} (Academic, New York, 1957).
417: \bibitem{Low} F. E. Low, Phys. Rev. {\bf 110}, 974 (1958).
418: \bibitem{Gizzi} D. Guiletti and L.A. Gizzi, Riv. Nuovo Cim. {\bf 21}, 1 (1998).
419: \end{thebibliography}
420:
421:
422: \end{document}
423:
424:
425:
426: