Published online by Cambridge University Press: 06 March 2006
We study the details of a gradual change in electron properties from those of a nearly-free-electron (NFE) metal to those of a strongly-coupled plasma, in ultrashort pulse energy deposition in solid metal targets. Time scales shorter than those of a target surface layer expansion are considered. Both the case of an optical laser (visible or near infrared wavelengths range) and of a free electron laser (vacuum ultraviolet or X-ray) are treated. The mechanisms responsible for the change in electron behavior are isochoric melting, lattice charge disordering, and electron mean free path reduction. We find that the transition from metal to plasma usually occurs via an intermediate stage of a charge-disordered solid (solid plasma), in which ions are at their lattice sites but the ionization stages of individual ions differ due to ionization from localized bound states. Charge disordered state formation is very rapid (typically, few femtoseconds or few tens of femtoseconds). Pathway to charge-disordered state differs in simple metals and in noble metals. Probabilities are derived for electron impact ionization and 3-body recombination of a bound ionic state in solid-density medium, applicable both in metal and in plasma regime. An evolution of energy coupling between electron and ion subsystems, from metallic electron-phonon (e-ph) to plasma electron-ion (e-i) coupling, is considered. Substantial increase in coupling parameter is expected as a result of charge disorder.
Interaction of an ultrashort laser pulse with metal surface leads to a rapid increase in electron temperature, whereas the thermalization between electron and ion subsystems occurs on a substantially longer time-scale (Anisimov et al., 1967, 1974). At laser intensities of 1012 to 1015 W/cm2, for pulse duration of a few tens of femtoseconds, electron temperature Te typically reaches 1 to 100 eV, while ions remain essentially cold in the duration of the laser pulse. In that situation, a significant redistribution occurs of electrons by their energy states. The ensuing modification of electron charge distribution and lattice potential may facilitate ultrafast melting, or even result in a nonthermal melting of the irradiated material. Ultrafast and/or nonthermal melting facilitated by depopulation of covalent-bonding orbitals was observed in several semiconductors (Shank et al., 1983; Tom et al., 1988; Saeta et al., 1991; Sokolowski-Tinten et al., 1995, 1998; Siders et al., 1999; Cavalleri et al., 2001; Lindenberg et al., 2005), C (Reitze et al., 1992), and Ga (Uteza et al., 2004), see review by Bennemann (2004). Ultrafast melting phenomenon was extensively studied theoretically (Stampfli & Bennemann, 1992; Rethfeld et al., 2002, 2004), and reproduced in molecular dynamics simulations (Silvestrelli et al., 1996; Gambirasio et al., 2000; Jeschke et al., 2001; Dumitrica & Allen, 2002). Softening of phonon modes was observed in Te (Hunsche et al., 1995) and Bi (DeCamp et al., 2001). Ultrafast thermal melting of Te was observed in (Ashitkov et al., 2002). Significant reflectivity increase on a subpicosecond time-scale, incompatible with Drude model and interpreted as melting, was detected in W (Wang & Downer, 1992). For self-absorption measurements or short-delay pump-probe absorption measurements, caution must be exercised to discriminate between suppression of interband absorption in a solid metal and an ultrafast melting of the metal (Fisher et al., 2002).
Transient increase in Al reflectivity at 800 nm pumping/probing wavelength was observed by Guo et al. (2000) in pump-probe experiments. Difference in target damage effect from 800 nm and 400 nm wavelength pulse was also reported ibidem. An increase in Al reflectivity at 800 nm wavelength, and absence of such increase at 400 nm wavelength, were observed by Fisher et al. (2002) in self-absorption experiments. Two alternative explanations for the experimental findings were provided in these works. Guo et al. (2000) suggested a nonthermal melting of Al, associated with pumping of a bonding-to-antibonding-state interband transition, as an explanation for their experimental results. This hypothesis was further investigated by Youn et al. (2004). Fisher et al. (2002) have, on the other hand, suggested an electron impact broadening of the interband absorption peak, without a loss of crystalline order, as an explanation for their experimental results. The fact that the increase in Al reflectivity at 800 nm was observed by Fisher et al. (2002) to occur on a time-scale of about 50 fs or shorter, at modest laser intensities, agrees with the impact broadening hypothesis, but can hardly be explained by a nonthermal melting. The controversy was eventually laid to rest by the experimental findings of Siwick et al. (2003) that rule out the nonthermal melting hypothesis of Guo et al. (2000). In Siwick et al. (2003) the melting of Al heated by a laser pulse of 800 nm wavelength was found to occur thermally, on a picosecond time-scale.
As the electron temperature increases, the electron-electron (e-e) and electron-ion (e-i) energy and momentum relaxation times characteristic of a periodic-lattice electron behavior must be replaced continuously by the relaxation times characteristic of a nonideal-plasma electron behavior. Although that transition was expected in theory (Lee & More, 1984; Mott, 1990) and observed in experiment (Milchberg et al., 1988; Ng et al., 1994; Price et al., 1995), the details of mechanism, conditions and time-scale of metal-to-plasma electron transition are not known. In the present work we attempt to fill this gap.
It is generally accepted that the exceedingly large probabilities for elastic electron collisions at Te ∼ TF (where TF is Fermi temperature), at solid metal densities, result in electron mean free path (MFP) with respect to longitudinal momentum relaxation being of the same order as the typical interionic distance a. The electron longitudinal momentum relaxation rate ν is then approaching ue /a, where ue is the characteristic electron velocity. This is the resistivity saturation (also referred to as minimal MFP, minimal conductivity, or Ioffe-Regel) regime (see Chapter 1.6.3 in Mott, 1990). However, very little is known on the details of this transition: how the e-ph interaction is replaced by e-i interaction, how the Umklapp electron-electron scattering dies out, and so on. Also, the details of the transition from (a relatively low) e-ph energy exchange rate to (a relatively high) e-i energy exchange rate, as well as the actual mechanism(s), Te values, and time-scales on which this transition occurs in various metals remained largely unknown until recently or until now. Attempts to evaluate theoretically the e-i energy exchange rate in a solid target with Te ∼ 1 − 100 eV and Ti < 1 eV have lead, under different approximations, to results varying by well over an order of magnitude for the same conditions (Dharma-wardana & Perrot, 1998, 2001; Hazak et al., 2001; Gericke et al., 2002). The key to understanding the e-i energy exchange physics, under the conditions considered is, again, the knowledge of the details of mechanism, conditions, and time-scale of the gradual transition in the conductivity electron properties from nearly-free-electron (NFE)-like to strongly-coupled-plasma-like. As we explain below, this gradual transition involves usually not only the conductivity electrons, but rather the core-state electrons as well.
This work is structured as follows. In Section 2 we consider the mechanisms for lattice periodicity breakdown and for gradual transition from metal-like to plasma-like electron behavior (electron metal-to-plasma transition, eMPT) in metal targets subjected to an ultrashort radiative energy deposition. In Sections 3 to 5, we study in detail the core-state electron impact ionization and three-body (Auger) recombination processes which, as we show, dominate the eMPT in most metals. In Sections 6 to 9, we demonstrate the important differences in eMPT mechanisms between various metallic elements. Three distinct categories of normal metals are found with respect to detailed mechanisms of eMPT under ultrashort laser irradiation. Metals with relatively deep-lying weakest-bound core states (Li, Be, Na, Mg, Al, K, Ca) form the first category. “Relatively” here is defined by comparison with the Fermi energy value in the metal under consideration. Metals with relatively shallow-lying weakest-bound core states (Zn, Cd, Sn, Pb) form the second category. Noble metals Cu, Ag, Au, in which d-band states localize with increasing electron temperature (Fisher et al., 2004, 2005a), form the third category. Transitional metals and semimetals are not considered in this work, and will be considered separately. Energy exchange rate between electron and ion subsystems is discussed in Section 7.
We use cgs units in all expressions. In the text, values of particle energies and temperatures are given for convenience in units of electron-volt rather than, respectively, erg or Kelvin.
Fundamental differences exist between electronic properties of a metal and electronic properties of a plasma, at the same ion number density and temperature. Electron transport properties and electron-ion energy exchange rate in metal are governed by lattice periodicity effects on (quasi-) electron states, so that the energy and momentum exchange occurs between collective perturbations of conductivity electron distribution (quasielectrons), and ion distribution (phonons). In plasma, on the other hand, energy and momentum exchange can be described as occurring between individual electrons and ions, in pair collisions, with an environment providing merely a screening background. The key to the transition from quasiparticle to individual-particle picture lies in the relation of the quasiparticle decay rate (in energy units) to the quasiparticle energy. For the quasiparticle picture to be valid, the former must be much smaller than the latter (Abrikosov, 1972). In other words, periodic-lattice electron behavior breaks down and eMPT occurs when (quasi-)electron total collision frequency, in energy units, is no longer much smaller than the electron Fermi energy.
There are five mechanisms by which the periodic-lattice behavior of conductivity electrons can break down in femtosecond laser-matter interaction. The mechanisms are as follows.
(1) Ultrafast isochoric melting. In metals irradiated by an ultrashort laser pulse the melting may occur either with an expansion of the free surface of the target (heterogeneous melting), or, provided the overheating of the solid is substantial, isochorically (homogeneous melting), see (Rethfeld et al., 2004). The mechanism we consider in this paragraph is the isochoric, homogeneous melting; lattice disordering mechanisms involving the free surface are outlined in paragraph (5) below. In NFE metals, homogeneous melting occurs as a result of energy transfer from electrons (they are heated by laser radiation) to lattice ions. As mentioned in the Introduction, in non-NFE metals and in semiconductors the thermal redistribution of electrons over energy levels changes the spatial charge distribution in the lattice and leads to weakening of covalent bonding. The weakened bonding facilitates ultrafast thermal melting. The spatial charge redistribution can induce non-thermal melting in not closely packed lattices. For NFE metals, on the other hand, conductivity electron spatial distribution is much closer to uniform even at low temperatures. Then, a shift of an ion from its equilibrium lattice location occurs on the time scale of one or few phonon periods, 102-103 fs, after isochoric melting ion temperature has been reached. It is important to stress that melting alone does not produce an eMPT. It is well known that a room-pressure isobaric equilibrium melting of a metal does not destroy a short-range lattice order, and thus produces only a modest increase in ν across the melting point. Relation ν << ue /a certainly remains valid. Laser-induced ultrafast isochoric melting occurs at higher ion temperatures and may have a stronger effect on the short-range order, eliciting stronger increase in ν. Still, an eMPT can not be produced by melting alone; the joint action of some (or all) of the mechanisms listed below is required.
(2) Charge disorder (Fisher et al., 2004, 2005a). Ionization of localized core states (in simple metals) or localization of d-band states (in noble metals) with increasing Te produce ions in a variety of ionization stages occupying individual lattice sites. Ions remain positioned at their lattice sites (at least for some time, estimated to be of the order of 100 fs), but their charge states vary from site to site in a random fashion. Lattice translational symmetry is thus lost even though the geometric ion ordering persists. This is a solid plasma, a charge-disordered solid state. At comparable abundances of two or more different ionization stages, the loss of short-range order is complete. Individual impact ionization events occur on a time-scale of the knocked-out electron evacuation from its parent ion potential well, that is, a/ue ∼ 0.1 fs. However, charge disorder emerges only when the abundance of the next ionization stage becomes significant ([gsim ]1%). It may take ∼1 to 100 fs, depending on ionization rate, after Te reaches sufficiently high values (1 to 20 eV). The details of this mechanism are considered in the Sections 3 to 9.
(3) Reduction in conductivity electron MFP due to the increasing collision rate. A typical linear size L of an electron wave-packet envelope can not significantly exceed the MFP. The MFP, and thus L as well, can become as small as a, thereby destroying Bloch state structure. For ν this phenomenon produces resistivity saturation. The values of ν and L are determined by Te and, at Te [lsim ] 1 eV, also by an ion temperature Ti (Fisher et al., 2002). In absence of charge disorder, resistivity saturation occurs usually as Te approaches Fermi temperature (Milchberg et al., 1988; Price et al., 1995; Fisher et al., 2002). Time scale for electron MFP reduction is given by the inverse of the electron total collision frequency, which is sub-femtosecond in the vicinity of resistivity saturation regime. Time scale on which Te changes is significantly longer in most cases, and determines therefore the resistivity saturation onset time.
The three aforementioned mechanisms can operate equally well in the metal bulk and on the metal surface, as they involve neither ion density change nor deviation from a local charge neutrality. There exist another two mechanisms that can affect the metallic electron properties of the target. Those two mechanisms involve explicitly the free target surface.
(4) The local breach of charge neutrality due to the escape of fast electrons from the target surface into the void, and the associated production of strong electric and magnetic fields (Gamaly et al., 2002). The effect of the strong surface fields on the electron states and transport inside the target, in the laser intensities range considered here (1012 − 1015 W/cm2), remains largely uninvestigated as far as we know.
(5) An anisotropic expansion of the surface target layers normally to the free surface of the target (Eliezer, 2002, 2005; Eidmann et al., 2000). This expansion is driven by the thermal pressure of conductivity electrons heated by the laser radiation absorption to temperatures of a few electron-volts or few tens of electron-volts. Estimates in Fisher et al. (2002) show that the typical time-scales for this effect, under our conditions, are in order of
, where δ is skin layer thickness, and Mi is ion mass.
In this work we only concentrate on the “bulk,” isochoric mechanisms (1)–(3). As we said in the Introduction, ultrafast melting mechanism (1) is important in semiconductors. Charge disorder mechanism (2) and electron MFP reduction mechanism (3) are dominant in simple metals. Furthermore, we show that there are two alternative pathways to charge disorder, one realized by ionization of the core states and the other realized by localization of the d-band states of noble metals. After the initial isochoric stage, which lasts several tens or hundreds of femtoseconds, expansion of the heated layers into the void can no longer be neglected, and mechanism (5) becomes important in most systems.
Charge disorder in normal metals is produced by ionization of localized core electron states. First we consider metals with core states lying at least several eV below the bottom of the conductivity band. Noble metals (Cu, Ag, Au), in which the d-band energy lies above the bottom of the conductivity band but below the Fermi level, or metals with core states or d-band located close to the bottom of the conductivity band (Zn, Cd, Sn, Pb), are considered in Sections 8 and 9. The charge disordering mechanism is more complicated in those metals.
Both in metals and in plasma the ionization of a bound electron state can occur either by a free electron impact or by photon absorption. At the laser intensities relevant to this work (below 1015 W/cm2) multiphoton ionization of core states can be neglected; therefore, for photon energies below the photoionization threshold (optical laser), the only ionization channel is the electron impact ionization (EII) of core states by sufficiently energetic conductivity electrons. For a vacuum-ultraviolet free electron laser (see, e.g., Lee et al., 2002; Ke Lan & Meyer-ter-Vehn, 2004; Alesini et al., 2004), or any other source of high-intensity femtosecond/attosecond radiation pulses with photon energies above the core state photoionization threshold in the irradiated sample, both EII and photoionization channels must be accounted for.
Finally, in a solid matter bombarded by an intense energetic ion beam (either laser or accelerator produced, see e.g. Malka & Fritzler, 2004; Hoffmann et al., 2005), dynamics of isochoric transition from an ordered solid to a high-energy-density matter is of significant interest. There, on sufficiently short time scales, both charge disordering due to interaction with projectile ions and charge disordering due to EII by hot electrons are expected.
Here we restrict ourselves to consideration of only the shallowest-lying (weakest-bound) core states of any ion, as core hole abundances even lower than one per ion are already sufficient for the eMPT. Core states considered have binding energies in the range of 15 to 100 eV, depending on the chemical element. Deeper-lying core states are not affected under present conditions. Zero of energy is chosen at the bottom of conductivity band, so that the core states have negative energy (i.e., positive binding energy Eb), and conductivity electron states have positive energy. An EII event occurs when a sufficiently energetic conductivity electron collides with a core-state electron, with energy transfer in the collision exceeding the core state binding energy. Final states of both electrons belong then in conductivity band, and an ionization degree of the parent ion increases by unity.
As the gradual eMPT is what we are ultimately interested in, we must develop a unified description for EII events (their cross-section, probability, and rate) in metal and in plasma. To the best of our knowledge, this was never carried out. The weakest-bound core state in metal is equivalent to an optical electron state of a plasma ion, and the NFE conductivity-band states in metal are equivalent to free electron states in plasma. The core state binding energy in metal is therefore measured relative to the bottom of the conductivity band. Due to density-dependent screening effects, the binding energy for the same ionic state in metal and in dilute plasma differ (see, for example, Johansson & Martensson, 1980). The EII cross section must be scaled to the correct binding energy (Salzmann, 1998). The appropriate expression for the EII cross-section σEII(E), where E is the incident electron energy, is presented at the end of this Section. For now, it is sufficient to state that σEII(E) is a smooth finite non-negative function of E, and
This presumes the absence of excitons in metal under conditions considered here.
To evaluate the probability (per second) P of ionization of a core state in a single ion, we use
where Ne is the number density of conductivity electrons, u(E) is the incident electron velocity, and the average 〈〉 is over the conductivity electron energy distribution. Π is a Pauli blocking factor accounting for reduction in P due to the significant probability of the electron final state(s) being occupied. The conductivity band is assumed parabolic in the extended zone scheme,
with k denoting the (quasi)electron momentum, and with the (quasi)electron mass equal to a free electron mass. Then,
The averaging in Eq. (2) is over the energy distribution g(E) f (E) of the incident electron,
Here
is the conductivity electron density of states (in units of [1 / erg cm3]) in a NFE band with dispersion relation (3), and
is the Fermi-Dirac distribution function, where kB is Boltzmann constant, and μ is electron chemical potential. The values of μ(Te) are determined implicitly by the identity
The EII event is only possible when the final states are vacant for both electrons involved. In contrast to a classical plasma, in nonclassical plasma or in metal the probability for the final states to be vacant may be low. Therefore, in the average (in right-hand side of Eq. (2)) a Pauli blocking factor 0 < Π(E − Eb) < 1 is included, and the expression for EII rate becomes
We estimate the blocking factor Π(E − Eb) in the following way. The total energy of the two electrons in the final state is E − Eb > 0. The probability to find the final state vacant decreases nearly exponentially as the energy of one of the final states decreases. Thus, the most favorable scattering event with respect to both final states being vacant is the scattering in which final energy of both electrons is approximately equal, each one being given then by (E − Eb)/2. Hence, the factor Π(E − Eb) may be approximated by
It approaches unity, of course, as E increases. It also approaches unity for any E at large Te due to a drop in μ(Te). In a low-Te limit, kBTe << μ, Eq. (10) is straightforward to derive; this is done by assuming any realistic form of differential EII cross-section dσ/dε (where 0 ≤ ε ≤ E − Eb is the energy of ejected electron) and integrating over ε. A prefactor close to unity appears then in the low-Te limit of the right-hand side of (10). The prefactor depends on E, Eb, μ, as well as on the details of the functional form of dσ/dε. Most important, however, it does not depend on Te. We neglect the said prefactor altogether.
We note that Ne does not appear explicitly in Eq. (9). Instead, it enters implicitly through the value of μ(Te), which appears in Eq. (7) for f (E). In classical plasmas f (E) << 1 for all E ≥ 0. In that limit we find Π(E − Eb) ≈ 1 for all E ≥ Eb, and f (E) ≈ exp[(μ − E)/kBTe]. Using the nondegenerate free-electron asymptote
one finds f (E) ∼ NeTe−3/2 exp(−E/kBTe), hence the familiar explicit factor Ne in EII probability in classical plasma.
In the above model for P, the use of a constant value for Eb is only justified as long as the depletion of the core state is not too large (say, no more than 10% vacancies abundance in the core state; for example, no less than 5.4 electrons on average per ion in the 2p state of Al). The abundance of holes (vacancies) in a core state is defined here as a ratio of the number of vacant electron states to the total number of electron states. At larger vacancy abundances, the reduction in nucleus screening for the remaining core electrons becomes important and Eb increases. This effect can be accounted for in equilibrium, by the following extension of the model. The equilibrium core hole abundance h(Te) is evaluated by including a narrow core state in the density of states, Eq. (6), and recalculating the chemical potential from Eq. (8) with Ne now including the core electrons as well; then h = 1 − f (−Eb). Values of Eb(Te) may be found using INFERNO model (Liberman, 1979, 1982).
For a laser pulse not much longer than the typical equilibration times, on the other hand, the equilibrium number of vacancies in the core state is not reached, and the core states are thus overpopulated for the given Te (i.e., actual vacancy abundance is lower than h). In the nonequilibrium case it is possible to solve rate equations for h, and to use Eb(h) as above; such a procedure, however, is beyond the scope of the present work. The hole population equilibration time is estimated in Section 6.
The following point must be clarified. Since the core state is localized, not all ions have the same number of core electrons. Consequently, the notion of Eb(Te) is meaningful only in the average sense, as explained in the Appendix of Liberman (1979). Simply put, ions with K = 1,…,q electrons in the core state have different binding energies EbK of the core state. The smaller is K, the larger is EbK. The average Eb(Te) increases with Te following the shift in the ionization composition of the matter, so that Eb(Te) is not far from EbK for the dominant K value at any given Te. Due to the dependence of free-electron screening on free electron density and temperature, all EbK are also functions of Te. However, the Te dependence of EbK is significantly weaker than the Te dependence of the average Eb(Te). As we already said, core-hole ion (K = q − 1) abundance ∼1% in metal is sufficient to produce the charge disorder and the ensuing eMPT. Thus, for normal metals below we assume Eb = Ebq, that is, binding energy is that of an electron in a fully occupied core state, unless explicitly stated otherwise. For noble metals the situation with the d-band energy is different, see Section 8.
Let us now turn to the determination of an appropriate expression for σEII(E). In a single ion, the cross-section for a direct EII event may be approximated by Lotz formula,
see e.g., Pal'chikov and Shevelko (1995). Here a0 is Bohr radius, q is the degeneracy of the core state, and Ry = 13.6 eV is the binding energy of a ground state of an isolated hydrogen atom. Lotz formula was obtained semi-empirically to fit the EII cross-sections measured for multiply charged ions. It was later found to be applicable to majority of direct EII cross sections in dilute ideal plasmas (including EII of weakly charged ions and EII from inner shells) (Griem, 1997). However, Lotz formula in its original form is inapplicable to either metals or dense plasmas. It does not account for the screening of a momentum transfer between electrons in a dense conducting medium. It also does not account for the difference in low-energy conductivity electron wave functions between a condensed matter and a dilute plasma. The latter problem is especially acute for low projectile energy values, E ≈ Eb. The E/Eb >> 1 nonrelativistic asymptotic (Bethe) form of Eq. (11) is applicable to an atom in dense medium as well as to an atom in dilute plasma. The screening effect of the medium becomes important either in relativistic limit (Fermi, 1940; Sternheimer & Peierls, 1971; Inokuti & Smith, 1982) which is not considered here, or for
(see, for example, Nozieres & Pines (1959) and Murillo & Weisheit (1998)). In most metals, the plasma frequency ωpe obeys ħωpe [lsim ] Eb. Thus, both the screening and the continuum wavefunction modification by the dense medium necessitate the amendment of Eq. (11) at the low (near-threshold) projectile energies, E ∼ Eb. Under present conditions, the σEII values in medium tend to be lower than for an individual ion. We introduce, therefore, an empirical correction factor (1 − Eb /E)Y into the EII cross-section,
Here Y > 0 is a fitting constant depending on the chemical element (and possibly on a lattice structure of metal, for elements in which a number of metallic phases exist). The value of Y can be found from the room-temperature experimental values of the core hole lifetimes, as explained in the next Section. Note that in the present model Y is independent of Te. By adopting Eq. (12) we therefore neglect the effect on EII cross-section stemming from changes in electron eigenstates and electron screening with an increasing electron temperature at a given ion density. This is, of course, an oversimplification; however, as Te increases, the dominant contribution to EII comes from larger E values, and the factor (1 − Eb /E)Y becomes less important. We stress that Y is a function of ion density, and may be assumed constant only as long as ions remain motionless (as in the present case). In expanding plasmas Y is expected to decrease as Ni decreases, and to approach zero eventually when Ni is low enough for the density effects on EII cross-section to be negligible.
In this work we concentrate on the ionization in metal targets with Te no larger than, say, 100 eV. This roughly corresponds to laser intensities below 1015 W/cm2, for a pulse duration of a few tens of femtoseconds. Rapid ionization of target material at higher laser intensities is considered, for example, by Kemp et al., (2004).
For completeness, we note that the EII contribution to stopping of a hot electron in a cold metal, and a related problem of an Auger lifetime of a core vacancy in metal at room conditions, have been studied (Almbladh et al., 1989; Schoene et al., 1999; Campillo et al., 2000; Knorren et al., 2001). The results in these publications are obtained by means of detailed band structure calculations, and are inapplicable to the systems with Te comparable with Fermi temperature or higher.
The process inverse to EII is the three-body recombination or, synonymously, an Auger process involving two electrons in a conductivity band and a core state hole. The three-body recombination probability and rate are found from the detailed balance requirements for the equilibrium conditions. We consider the simple case of relatively low abundance of holes, such that only K = q and K = q − 1 species are present in the system. That is, abundances of species with K < q − 1 are negligible. This is the case relevant to eMPT. The equilibrium core hole abundance h = 1 − f (−Eb). There are q core states per ion, and each of those states can be occupied or vacant. Thus, the probability to have a hole in the core state of a given ion equals qh, provided the probability to have more than one core-state hole per ion is negligible. Of course, this approximation is only accurate for qh << 1. The probability to have no holes in the core state is then 1 − qh. Rate of EII (in units of events per second per cm3) is given by
where Nq is the number density of ions with no core-state holes (i.e., with q electrons in the core state per ion). In equilibrium,
where Ni is the total number density of lattice ions. Likewise, three-body recombination rate W3 is expressed via the three-body recombination probability P3 by
where Nq−1 is the number density of ions with one core hole per ion. The physical meaning of P3 is the inverse mean lifetime, 〈τh〉−1 ≡ P3, of a hole in a core state of a lattice ion. For equilibrium conditions, Nq−1 is given by
In equilibrium between the core state population and the next ionization stage ground state population, W3 = W. Thus, for qh << 1,
The three-body recombination occurs in an interaction of two conductivity electrons in presence of an ion with a core-state vacancy. One of the conductivity electrons makes a transition into the vacant core state and the other one takes the excess energy. In the limit kBTe << min(Eb,EF), where EF ≡ kBTF ≡ μ(Te = 0), one should expect P3 to become independent of Te, since the energy distribution of conductivity electrons becomes nearly independent of Te at kBTe << EF. It is easy to show that Eq. (17) yields indeed the correct limiting behavior: limTe→0 P3 = const. The prefactor in Eq. (17) becomes
To find the limit limTe→0 P, we must consider two distinct cases: EF < Eb and EF ≥ Eb. Let us first consider the case EF < Eb. Using Eq. (9) for P and Eq. (7) for f, we find
In the case under consideration, unity next to exp((E − μ)/kBTe) is negligible since kBTe << EF < Eb, E ≥ Eb, and lim μ(Te → 0) = EF. Thus,
independent of Te. For the case EF ≥ Eb, which corresponds to a relatively weakly bound (shallow-lying) core state, one finds in a similar way
independent of Te, too. Thus, for both cases limTe→0 P3 can be described by a single expression
The value of limTe→0 P3 depends on Ne both via EF ∼ Ne2/3 and via Y. The dependence of P3 on Ne in this degenerate-conductivity-electrons limit is therefore far more complicated than the trivial P3 ∼ Ne2 dependence encountered in an ideal classical plasma limit.
Eqs. (18) and (21) show that, despite the Pauli blocking, for kBTe << EF the dominant contribution to EII probability P comes from slow projectile electrons, Eb < E ≤ Eb + 2EF. This underlines the importance of the empirical correction factor (1 − Eb /E)Y introduced in the previous Section.
We have calculated P(Te) and P3(Te) for solid-density Li, Be, Na, Mg, Al, K, and Ca. For these metals the ratio Eb /EF > 4. Parameters used in the calculations are presented in Table 1. Elements are listed in the order of increasing atomic number. Values of Ne include conductivity electrons only. Values of the binding energy with respect to Fermi level, Eb + EF, are taken from NIST data tables http://physics.nist.gov/PhysRefData/FFast/Text/Cover.html. Values of EF are obtained from table values for Ne using Eq. (8) at Te = 0. Values of room-temperature hole lifetimes 〈τh〉 were found in literature, as follows. For Li, we adopted the experimental value of 〈τh〉 = ħ/Γ ≈ 20 fs as reported by Citrin et al. (1979), where Γ is the lifetime broadening of the core hole state. For Be, no experimental data were found, so we used the value of 〈τh〉 ≈ 16 fs obtained by band structure calculations (Almbladh et al., 1989). For Na and Mg we adopted the 〈τh〉 experimental values of ≈70 fs and ≈30 fs, respectively (Citrin et al., 1979; Neddermeyer, 1976). For Al, we adopted the experimental value of 〈τh〉 ≈ 20 fs reported by Theis and Horn (1993), see also, Almbladh et al. (1989) and Citrin et al. (1979). For K, experimental 〈τh〉 ≈ 44 fs (First et al., 1988). All the above 〈τh〉 data are typically accurate to within a factor of 2. Datum for K is accurate to within 15%. For Ca no data were found either experimental or theoretical, so we used a typical value 〈τh〉 ≈ 20 fs; this is merely an order-of-magnitude “guesstimate.” Table 1 lists also the fitted Y values for room-density metallic state of Li, Be, Na, Mg, Al, K, and Ca. The Y values were determined by equating the limTe→0 P3 value, Eq. (21), to 〈τh〉−1 for each of the elements. We note the ostensible regularity in Y: the latter seems to grow nearly monotonously with the increasing atomic number.
Parameters of weakest-bound core states for several metals
Calculation results for EII probability P in the aforementioned metals are presented in Table 2. Data are only given for P > 106 s−1. Figure 1 shows P(Te) for the same metals. Figure 2 shows calculation results for P3(Te) in the same metals, up to qh = 0.1. One can see that P3 decreases with increasing temperature. It is important to note that constant Eb and Ne were assumed in the above calculations, which corresponds to core-ionized ion abundance much smaller than unity. As we said, under equilibrium conditions the present values of P(Te) and P3(Te) are only valid for Te such that the equilibrium core-ionized ion abundance qh(Te) << 1. For very short pulses, as long as the hole abundance is substantially lower than in equilibrium, present values of P(Te) can be used at higher Te as well (provided the conductivity electrons are thermalized between themselves, i.e., the conductivity electron distribution function can be approximated by Eq. (7)).
Calculated probabilities of core state electron impact ionization for several metals
Calculated probabilities of core state electron impact ionization for several metals.
Calculated probabilities of core vacancy 3-body recombination for several metals.
Rate with which an equilibrium value h of the core hole abundance is approached, for a given electron temperature Te, is easy to evaluate. Assuming, as above, that only K = q and K = q − 1 species are present in the system, for Nq and Nq−1 the following rate equations are readily written down:
with the initial condition Nq(t = 0) = Nq0, Nq−1(t = 0) = Nq−10 = Ni − Nq0. The solution is
where
and
The hole abundance relaxes to its equilibrium value h at the rate Ω which is given by
Ω values vary between 1013 and 1015 s−1 for the metals considered in this Section.
For a short-range lattice order to be destroyed by charge disordering, the fraction Nq−1 /Ni of core-ionized ions in the medium must be in excess of, say, 0.1/n ∼ 0.01. Here n is the number of near neighbors of a given ion in the lattice, typically between 4 and 12. To attain these values of Nq−1 /Ni by means of ultrashort optical laser energy deposition one needs, first of all, to reach Te such that qh(Te) [gsim ] 0.01. Furthermore, to attain these values of Nq−1 /Ni on a 100-fs scale or faster, one needs the EII rate P > 1011 s−1, see Eq. (22). Temperatures at which qh equals 0.001, 0.01, and 0.1 are listed in Table 3. For readers' convenience, Fermi energies of the elements are listed again in Table 3. By looking at Tables 2 and 3, one can immediately see that the charge disorder emerges at Te ≈ (0.5 − 2)TF. That is, the same temperature range as for the minimal conductivity regime. For that reason, it is hard (if at all possible) to detect charge disorder by self-absorption or pump-probe measurements of optical properties of the metals considered in this section. This confirms the conclusion drawn by Fisher et al. (2002) that for 50-fs laser pulse irradiation of Al targets at intensities below 1015 W/cm2, at 1.5 eV and 3 eV photon energies, the core state ionization does not affect appreciably the laser pulse absorption coefficient.
Temperature values corresponding to specified equilibrium core-ionized ion abundances qh
Charge disorder can also be produced, temporarily, by a hard ultraviolet or X–ray laser irradiation of a metal sample. To produce an observable charge disorder effect in a relatively cool metal, photons with energy close to the photoionization threshold Eb + EF are required. Tunability of a free electron laser presents a great advantage in this case. The optimal laser energy deposition should be such that the Nq−1 /Ni value produced is in the range 0.01 − 0.1. Lower core hole populations are harder to observe, while higher core hole populations yield, upon recombination to equilibrium hole abundances, Te values that are too high (close to Fermi temperature). At Te ≈ TF or higher, the ν values, as we said, are close to minimal conductivity value anyway, and the charge disorder effect is not manifested in optical properties. Core hole population thermalization is expected to occur, under the conditions relevant to X–ray induced charge disorder, at a rate of (1 − 5) × 1013 s−1, that is, on a time scale of tens of femtoseconds, see Figure 2.
It is important to stress that the charge disorder being hard to detect optically in metals listed in Tables 1, 2, and 3 does not imply that charge disorder is hard to detect at all in these metals. Charge disorder produces a significant increase in electron-ion energy coupling (Fisher, 2005b). Indeed, a solid-density medium with partially degenerate conductivity electrons and identical ions in a short-range order may have an e-i energy coupling coefficient significantly lower than a strongly-coupled plasma under the same conditions. The plasma lacks a short-range order by virtue of its ionization composition, and is characterized by Braginsky e-i coupling value (Braginskii, 1965); originally derived by Kogan (1959). Plasma e-i coupling value is limited by the maximal possible coupling constraint (Fisher, 2005b) due to a limit on energy transfer in a single e-i collision. (The maximal possible coupling value is evaluated by simply assuming a large-angle elastic scattering of electrons on ions, with a MFP equal to the interionic distance). As the plasma e-i energy coupling constant is typically an order of magnitude larger than e-i (or, more accurately, e-ph) energy coupling constant of a simple metal, one would expect a notable shortening in ultrafast melting times of a charge-disordered metal in comparison to a model in which the charge disorder is disregarded. A comparison for Cu between melting times with and without an account for the charge disorder is given in Figure 1 of (Fisher, 2005b). Charge disorder in Cu is discussed in the next Section.
The occurrence of charge disorder was first predicted by Fisher et al. (2004, 2005a, 2005b) for noble metals under ultrashort pulse laser irradiation. In noble metals (Cu, Ag, Au) the charge disorder is brought along by a mechanism that is somewhat different from that in the normal metals described above. In normal metals, charge disorder is brought along by core state vacancies production, the core state being localized throughout the Te range considered. In noble metals, on the other hand, charge disorder is brought along by d-band state localization which occurs at Te values of a few electron-volts. At room temperature, d-band hole is mobile, similar to a conductivity-band hole, rather than localized in a single ion potential well. The criterion for mobile versus localized holes is according to the relationship between (A) hole tunneling rate out of the resonance and (B) the hole inverse lifetime with respect to inelastic processes. At room temperature, in d-bands of noble metals the typical hole inelastic lifetimes (dominated by three-body recombination) are 5 − 50 fs, see for example, Zhukov et al. (2003). We estimate the room-temperature tunneling time to be of order of 1 fs, so the holes are not localized. As Te increases, binding energy of the d-band increases with it, and the d-hole tunneling rate between individual ion sites drops. This leads to localization of the d-holes at individual lattice sites. In the localization Te domain, Te of a few eV, the d-band hole concentration is already large. It exceeds one hole per ion at Te [gsim ] 3 eV, see Table 4. Thus, the charge disorder follows directly from d-band localization. In other words, in noble metals charge disorder is brought along by a collective process (band structure evolution with increasing Te) rather than by inelastic pair collisions (EII events) as in normal metals. As the result, in noble metals the parameters governing the eMPT are the energy E(Te) and the lifetime width of the d-band, while in normal metals the parameter governing eMPT is the core-state vacancy abundance. Moreover, during a femtosecond laser irradiation of normal metals with large Eb, the core state population is not necessarily in thermal equilibrium with the conductivity-band electrons. In that case, the core-state vacancy abundance depends on time explicitly (and not via Te), and is determined by the ionization rate.
Equilibrium number of electrons per ion in d-bands of noble metals, as a function of electron temperature at solid density, evaluated using INFERNO model
The metals Zn, Cd, Sn, and Pb represent an intermediate case between noble metals on one hand and normal metals on the other. The d-bands of Zn and Cd are located close to the bottom of the conductivity band, and are localized in the above sense even at room temperature. In the case of Zn and Cd Eq. (12) is inapplicable. In Sn (white tin) and in Pb, the fully occupied d-bands are located several electron-volts below the bottom of the conductivity band, Eb ≈ EF ≈ 10 eV. For Sn and Pb, Eq. (12) can be applied, provided an amenable Y calibration is carried out. Such a procedure is, however, largely unnecessary. Indeed, due to the relative proximity in energy between the d-band and the conductivity band one can assume that, at least on a time scale of 10 fs or longer, the d-band hole population is in thermal equilibrium with the conductivity electrons. INFERNO model calculation results for average fraction of core-ionized lattice ions, Nq−1 /Ni = qh, in 4d band of Sn and in 5d band of Pb are given in Table 5. One can see that the charge disorder onset is expected at Te ≈ 3 eV. At Te = 5 eV the loss of short-range order is complete. These temperatures are by a factor of 2 − 3 lower than the Fermi temperatures of Sn and Pb.
Equilibrium fraction of core-ionized ions in Sn and Pb at solid densities
INFERNO model calculation results for average equilibrium fraction of core-ionized lattice ions qh in 3d band of Zn and in 4d band of Cd are given in Table 6. The core-ionized ion abundance in metallic Zn and Cd becomes significant at Te as low as 1 − 2 eV. A complete charge disorder is expected at Te ≥ 2 eV. For comparison, Fermi energy of Zn is 9.4 eV. Zn thus seems to be the best candidate for charge disorder effect observation using optical lasers. The Fermi energy of Cd is 7.5 eV, so Cd should also exhibit clearly the optical properties characteristic of charge disorder (resistivity saturation) at Te ≥ 2 eV.
Equilibrium fraction of core-ionized ions in Zn and Cd at solid densities
Transition from metallic to plasma electron transport properties, eMPT, has been studied in metals undergoing ultrashort pulse electromagnetic energy deposition. The mechanisms contributing to eMPT have been discussed, and the role of charge disorder phenomenon in the eMPT has been determined. Approximate analytic expressions for cross-section, probability, and rate of EII of core states, applicable to both metals and plasmas, have been derived. Probability of three-body recombination was determined through detailed balance. Calculations were carried out for a number of normal metals.
It was shown that the charge disorder occurs at room densities in all metals considered. Under equilibrium conditions for core population, in Li, Be, Na, Mg, Al, K, and Ca the charge disorder occurs at temperatures close to Fermi temperature, when the electron momentum relaxation rate is already close to resistivity saturation limit. The resistivity saturation in those metals is reached, with rising electron temperature, due to an increasing electron elastic collision rate. For Zn, Cd, Sn, and Pb the charge disorder occurs at temperatures a few times lower than the Fermi temperature, with charge disorder leading to resistivity saturation. For noble metals, resistivity saturation is also reached through charge disorder. In the case of noble metals, in contrast to simple metals, charge disorder occurs due to d-band sinking and localization, rather than due to core state ionization. It was concluded that effect of charge disorder on visible or near-infrared femtosecond laser absorption coefficient should be pronounced in Zn, Cd, Sn, Pb, Cu, Ag, Au, and probably can not be detected in Li, Be, Na, Mg, Al, K, and Ca. Charge disorder produced directly by energetic photon ionization of core states was also considered. Core hole lifetimes (with respect to recombination from conductivity band) were found to be 10 − 100 fs for Te = 0 − 20 eV and Eb = 15 − 100 eV.
Differences in experimentally determined optical properties of Cu and Al near the resistivity saturation domain were noted by Price et al. (1995), and the shift of Cu d-band was proposed as an explanation. Lack of detectable effect of core ionization on optical properties of Al was determined both experimentally and theoretically by Fisher et al. (2002). In both works, the absorption of visible or near-infrared femtosecond laser pulses was used both to heat the target and to probe its electron transport properties. Unfortunately, we are not aware of any detailed experiment on the femtosecond laser absorption in Zn, Cd, Sn, or Pb that would allow to test the predictions of the present work.
In noble metals, as we said, the charge disorder is expected to increase the value of electron-ion energy coupling coefficient g and thus to reduce melting time. The charge disorder at sub-Fermi temperatures also affects optical properties of the noble metals, although that effect may be somewhat harder to single out. Pump-probe results for Cu were reported in several works. In the pioneering study of Elsayed-Ali et al. (1987) Fermi level smearing effect with increasing Te was demonstrated, for Te ≤ 0.2 eV. In a recent study (Sandhu et al., 2005) detailed pump-probe results for Cu and Al are presented for pump intensity of 3 × 1015 W/cm2, showing a transition to plasma regime. This intensity produces peak Te several times higher than the Fermi temperature. For the present purpose (to study the specifics of the eMPT, that is), this temperature is too high. Experiments at 10–100 times lower intensities are needed. In Wang et al. (1994) an increase in g was indeed observed by means of a two-pulse thermionic emission, for Au at modest laser intensities (Te of about 1 eV or even lower). This result is suggestive of a charge disorder effect, however, the temperature seems too low. Thus, the experimental evidence for charge disorder in the noble metals remains inconclusive so far.
Several experiments can be carried out to test the predictions made in this work. First of all, pump-probe or self-absorption ultrashort laser experiments can be carried out on Zn targets in visible or near infrared range. Excessive electron momentum relaxation rate at Te > 1 eV, and resistivity saturation at Te
2 eV are predicted as a result of charge disorder. Similar experiments on Cd are also possible, as explained in the previous Section. For Sn and Pb, the charge disorder is predicted to cause resistivity saturation at Te
3 eV. Experiments with probe wavelength near parallel-band interband absorption peak (the peak being produced by conductivity band splitting at Brillouin zone boundary (Ashcroft & Sturm, 1971)) should be particularly conclusive, since the peak is destroyed by the loss of short-range order in a charge disordered solid. Note that, by contrast, an interband absorption peak produced by electron excitation from a core state or from a d-band to the Fermi surface is not likely to be destroyed by charge disorder. It is subjected, however, to an increased lifetime broadening as a result of an increased electron collision rate; it is also broadened by Fermi level smearing.
Another possibility is to detect the charge disorder effect in ultrafast melting experiments. The X-ray probe experiments (see Siders et al., 1999; von der Linde et al., 2001; Lindenberg et al., 2005) or electron-beam probe experiments (Siwick et al., 2003) allow to directly probe the ion spatial arrangement and to resolve the ultrafast homogeneous or heterogeneous melting dynamics. Those are especially relevant for noble metals where a significant difference in melting times with and without charge disorder is predicted; and for normal metals where, at Te close to Fermi temperature, various theoretical predictions for electron-ion coupling coefficient value differ substantially.
The authors thankfully acknowledge discussions with Z. Zinamon, S.I. Anisimov, N.A. Inogamov, and S.N. Gordienko.
Parameters of weakest-bound core states for several metals
Calculated probabilities of core state electron impact ionization for several metals
Calculated probabilities of core state electron impact ionization for several metals.
Calculated probabilities of core vacancy 3-body recombination for several metals.
Temperature values corresponding to specified equilibrium core-ionized ion abundances qh
Equilibrium number of electrons per ion in d-bands of noble metals, as a function of electron temperature at solid density, evaluated using INFERNO model
Equilibrium fraction of core-ionized ions in Sn and Pb at solid densities
Equilibrium fraction of core-ionized ions in Zn and Cd at solid densities