Open access

Kinetics and Dynamics of Phase Transformations in Metals Under Action of Ultra-Short High-Power Laser Pulses

Written By

V.I. Mazhukin

Submitted: 14 December 2011 Published: 17 October 2012

DOI: 10.5772/50731

From the Edited Volume

Laser Pulses - Theory, Technology, and Applications

Edited by Igor Peshko

Chapter metrics overview

2,143 Chapter Downloads

View Full Metrics

1. Introduction

Action of super-power and ultra-short laser pulses on highly absorbing condensed media is investigated in the past two decades [1] - [5]. The urgency of this problem is primarily determined by a variety of practical applications of pulsed laser irradiation. In a short period of time scientists have mastered such operations as ablation of elemental materials by femtosecond lasers (100 fs) [6], femtosecond nanostructuring [7], generation of metallic nanoparticles and nanostructures by laser ablation of massive targets by 40 fs and 2ps pulses [8], etc.

The use of super-power G~1012÷1015 W/cm2 and ultra-short τL≈10-12÷10-15 s laser pulses for dimensional processing of materials, such as cutting of various materials by pico-femtosecond laser pulses [9], micro-drilling by a femtosecond laser [10 ], surface etching of metals (Al, Cu, Mo, Ni) and semiconductors (Si) [11], is accompanied by realization of unique physical conditions. In particular, the duration of action becomes comparable with the characteristic times of thermalization and phase transitions in matter. This leads to the need to address complex fundamental problems, including the heating of the material and the kinetics of phase transitions in a strong deviation from local thermodynamic equilibrium.

It should be noted that the majority of laser technologies is associated with the beginning of phase transformations in the material. In particular, the action of pico - femtosecond laser pulses of high intensity on solid targets is one of the ways to create individual particles with unique characteristics, or to form their streams, consisting of a cluster, liquid or solid fragments of the target. The formation of the particle flux in the pulsed laser ablation is observed for a wide range of materials: metals, semiconductors and insulators. The possibility of usage of laser ablation products in practical applications was the impetus for a number of experimental and theoretical studies [12], aimed at studying the conditions and mechanisms of formation of particles of nano - and micro- sizes during the laser exposure.

The accumulation of knowledge of the experimental nature in the first place in this dynamic field leads not only to the variety of practical applications of pulsed laser action, but also to a number of issues of independent physical interest.

The main features of ultrashort action on metals are associated with high speed and voluminous nature of the energy release of the laser pulse. The high rate of heating of a condensed medium is associated with rapid phase transformations of matter, characterized by the transfer of superheated phase boundaries of high-power fluxes of mass and energy. Overheated metastable states at the interface are characterized by temperatures, whose values can be hundreds of degrees higher than the equilibrium values of the melting point or boiling point. Removal of the energy by the flow of matter in conjunction with volume mechanism of energy release of laser radiation contribute to the formation of metastable superheated regions in the volume of solid and liquid phases with near-surface temperature maximum. Calculations [13] showed that the maximum speed of melting front are comparable to the speed of sound ~ (0.5 - 6) km / s, and the phase velocity of the solidification front are (10 - 200) m / s. Accordingly, the maximum superheating / undercooling can reach several thousand / hundred degrees. The achievement of such overheating and overcooling leads to large gradients of the Gibbs energy, which actually determine the driving force for high-speed phase transformations.

Physics of supercooled states in metallic and nonmetallic systems, because of their widespread use in the production of new materials technology, is relatively well-studied [14]. Superheated states received little attention until recently, largely because of the difficulty of experimental investigation and non-obviousness of their application. The situation changed with the advent of femtosecond laser pulses and their application for the production of nanoparticles and nanomaterials [15] - [17].

The purpose of this chapter is the theoretical study of nonequilibrium pulsed laser heating of metals, the kinetics and dynamics of phase transitions in a deep deviation from the local thermodynamic equilibrium.

The main tool for studying the processes initiated by laser pulses of picosecond and femtosecond duration are the methods of mathematical modeling and computational experiment (CE). The possibilities of experimental approaches in this area are very limited due to the large transience of the processes. Computational experiments are preferred in the cases where the natural experiment is not possible, is very difficult or very expensive. The statement of CE is especially convincing in the studies of the kinetics and dynamics of fast processes. For its statement, the computational experiment, which is an important link connecting theory and field experiments, requires the development of appropriate models, determining the properties of all substances studied, the development of computational algorithms and the creation of program codes.

Construction of a theory that covers most of the features of pulsed laser action on materials is very difficult. The methods of mathematical modeling in theoretical constructs have that advantage that you can use the phenomenological and the experimental data.

The description of the kinetics of fast phase transitions of the first kind is carried out in two classes of mathematical models: continuum and atomistic. Continuous models are based on the equations of continuum mechanics, as a rule, are represented as partial differential equations with appropriate boundary conditions and equations of state. Continuum model is used to describe the macro-level processes, heterogeneous kinetics and dynamics of phase transitions of the 1st kind [18]. Atomistic approach is based on a model of molecular dynamics is used to describe the kinetics of homogeneous phase transitions [19].

Advertisement

2. Brief theory of phase transformations of the 1st type

The basis of first-order phase transformations - melting-solidification and evaporation-condensation are two qualitatively different mechanisms: heterogeneous and homogeneous ones. The heterogeneous mechanism is characterized by a sharp interface between the phases (phase front of zero thickness) and determines the dynamics of phase transformations of the 1st type. The homogeneous mechanism of phase transitions that arise usually under the influence of the volume heating or cooling is associated with the processes of volume melting and boiling or spontaneous crystallization and condensation.

Historically, the theory of phase transformations of the 1st type of melting is based on two fundamental sections of classical physics, thermodynamics and kinetics.

2.1. Thermodynamical approach

Thermodynamics is a macroscopic theory [20], which from the energy point of view, considers the properties of macroscopic bodies in equilibrium. This allows to have a great community for the conclusions of thermodynamics. However, thermodynamics does not take into account the internal structure of the considered bodies and some of its conclusions and regulations do not have physical clarity. One way to describe the equilibrium processes in the equilibrium thermodynamics is a theory of thermodynamic potentials.

2.1.1. Phase equilibrium and phase transformations

The properties of a thermodynamic system are determined by thermodynamic parameters. The energy state of a thermodynamic system in equilibrium is uniquely determined by the parameters of the system. There is a unique relationship between the parameters of the system, which is mathematically represented by the state function. From a mathematical point of view, this means that the function has a total differential.

The basis of the method of thermodynamic potentials is just the possibility of introducing the state functions for the equilibrium processes, with total differentials describing the change in the state of a thermodynamic system.

The main identity of thermodynamics of equilibrium processes is usually represented as

TdS=dε+pdVordε=TdSpdVE1

Depending on the choice of the two independent parameters, one can introduce thermodynamic potentials, which differentiation allows determining the other unknown parameters of the state. In general, a thermodynamic potential may be a function of various parameters. In this notation, the internal energy is given as a function of entropy and volume:ε=ε(S,V). It is the function of the state and it has a total differential with respect to its variables. The total differential dε can be used to determine the temperature T and pressurep. However, the usage of entropy S and volume V as two independent variables is inconvenient because they are difficult to control in experiment. It is more convenient to use pressure pand temperature T as two independent variables.

After consecutive transformation of internal energy U first into enthalpy H=H(S,p)

H=ε+pV,E2

which total differential, taking into account the basic identity of thermodynamics has the form

dH=dε+pdV+Vdp=TdS+VdpE3

and then into Gibbs energy G=G(T,p) (Gibbs thermodynamic potential)

G=HTS=ε+pVTSE4

one can obtain the total differential of Gibbs energy, which with account of the basic identity of thermodynamics (1), takes the form

dG=SdT+VdpE5

It is convenient in the fact that the independent variables T and p are easily to modify and to control in experiment.

After reaching the equilibrium state of the system, Gibbs potential takes its minimum value and becomes constant:dG=0. This allows to use the condition of minimum of the Gibbs potential for the description of equilibrium states in which T=const andp=const.

Thermodynamic systems in equilibrium state do not necessarily have to be a homogeneous medium. A system in equilibrium may be composed of several phases, different in their physical and chemical properties, separated by the phase boundaries not changing over time.

The multiphase thermodynamic systems are most simply described, components of which are in equilibrium states, and there is no transfer of matter, energy and momentum through interphase boundaries. In this case, such thermodynamic system is in equilibrium and methods of equilibrium thermodynamics apply to describe it.

Given that during phase transformations, each of the phases is a system with variable mass, the notion of chemical potential μ(p,T) is introduced into the thermodynamic description. It is used to take into account not only exchange of energy but also the exchange of mass (particles). To determine the chemical potential, the term which takes into account the possibility of changing the number of particles in a homogeneous system (the same can be done with the other potentials) is formally introduced into the expression for the thermodynamic Gibbs potential:

dG=SdT+Vdp+μdNE6
dNdetermines the change of the number of particles in the system.

If the thermodynamic potential dG is given as a function of temperature and pressure, then the value of the chemical potential is written as: μ=(GN)p,T

The chemical potential can be written through other thermodynamical functions but in this case it will be written in terms of other state parameters:

μ=(GN)p,T=(UN)V,S=(HN)p,SE7

From the relation (6) it follows, that the chemical potential is a physical quantity that is equal to the value of some thermodynamical potential (with constant certain parameters) that is required to add to the system to change its number of particles by unity.

2.1.2. The conditions of equilibrium for two-phase single-component system

If macroscopic transport does not occur through the interphase boundaries, and the phases themselves are in a state of thermodynamic equilibrium, such thermodynamic system, in spite of its heterogeneity, will be in a state of thermodynamic equilibrium.

For the phase equilibrium in the one-component two-phase system, the following three conditions must be fulfilled:

  1. the condition of thermal equilibrium, that means the equality of temperatures at both sides of the interphase boundary:T1=T2=Teq,

  2. the condition of mechanical equilibrium, consisting of the equality of pressure at the both sides of the interphase boundary:p1=p2=peq,

  3. the condition of the equality of the Gibbs energy per particle, consisting of the requirement of the absence of the macroscopic transfer of molecules (atoms) of this material from one phase to another: μ1(p,T)=μ2(p,T)

In principle, we can use not the Gibbs energy, but any thermodynamic potential, which has a minimum in equilibrium. It is not difficult to show the validity of these conditions. We shall use the total differential for the internal energy, taking into account changes in the number of particles:

dε=TdSpdV+μdNE8

We write this expression for each phase of the closed-loop system:

dε1=T1dS1p1dV1+μ1dN1dε2=T2dS2p2dV2+μ2dN2E9

The closedness of the system automatically gives the following equation:

dε1=dε2,dS1=dS2,dV1=dV2,dN1=dN2E10
.

Consider an equilibrium two-phase system under some simplifying assumptions. Assume that the phases do not change the volume and do not exchange particles, i.e.dV1=dV2=0,dN1=dN2=0. Combining the equations (7) for this case, we obtain the expression:dε1+dε2=dS1(T1T2)=0, which gives the condition of thermal equilibriumT1=T2=Teq. Assuming the constancy of the entropy and the number of particles in phases, i.e. dS1=dS2=0,dN1=dN2=0we obtain the condition of mechanical equilibrium:p1=p2=peq.

Given the conditions of thermal and mechanical equilibrium of (7) we obtain the equality of chemical potentials in the different phasesμ1(p,T)=μ2(p,T). This equation can be solved for the variables T and p and may represent the equilibrium curves of the two phases in the form peq=p(Teq) orTeq=T(peq). If we consider the boundary between liquid and solid, the equilibrium melting curve is obtainedTm=T(pm).

When describing the interface between liquid and gas, the equilibrium vaporization curve is obtainedpsat=p(Tb).

It should be noted that the processes at the interface are static in nature as in the case of equilibrium of different phases and also during phase transitions. There is a constant process of transition of particles from one phase to another at the interface. In equilibrium, these opposing processes compensate each other, and during supply or withdrawal of heat to one of the phases one of these processes begins to dominate leading to a change in the amount of matter in various states of aggregation.

If the components of the thermodynamic system are not in equilibrium with each other, then there are thermodynamic flows through their interface. This will be a process of transformation of matter from one state to another, i.e. phase transformation. Assuming that the occurring processes are quasi-static and the flows are small, one can use the methods of equilibrium thermodynamics to describe such non-equilibrium system. In this case we assume an infinitely small difference between the thermodynamic parameters in different parts of the system.

2.1.3. The driving force of phase transformations

In the process of phase transitions of the 1st type, a number of quantities undergo abrupt changes at the interface, so in the following text in the thermodynamic equations, the sign of the differential will be replace by the corresponding value of the difference dFΔF for the temperature of the phase transitionTeq.

The driving force of phase transitions of the first type is determined by the difference of Gibbs energy (or the magnitude of overheating/overcoolingΔT) for two phases at the interface, defined in (4) and (5) and can be written in two forms

ΔG=(ε1ε2)+peq(V1V2)Teq(S1S2)=Δε+peqΔVTeqΔSE11
ΔG=VΔpSΔTE12

where

Δp=ppeq,ΔT=TTeq,Δε=ε1ε2,ΔV=V1V2,ΔS=S1S2E13

Equilibrium. In equilibrium, ΔG=0and the equality (8) takes the form:

Δε+peqΔVTeqΔS=ΔHTeqΔS=0E14

where the difference of enthalpyΔH=Δε+peqΔV, is known as the equilibrium latent heat of transformation

Leq=ΔH=TeqΔSE15
.

From equation (9) we can obtain the dependence of the equilibrium pressure Peq on the temperatureTeq, that is known as the curve of Clausius-Clapeyron:

dpeqdTeq=LeqTeq1ΔVE16

If one take as ΔV the difference of the volumes of vapor Vvap and condensed Vsol phases, then since Vvap>>Vsol for ideal gas ΔV=VvaporVsolidRTP one obtain the expression

dpeqpeq=LυRdTeqTeq2E17

After integration, we find the temperature dependence of the equilibrium vapor pressure that is widely used for many materials:

peq=pbexp[LυR(1Tb1Teq)]E18

The vapor pressure of the material in equilibrium with solid or liquid phase is called the saturated vapor pressure and with the notationpeq=psat, Teq=Tsuris usually written as

psat=pbexp[LυR(1Tb1Tsur)]E19

where Tsur is the temperature of the surface of the condensed phase, pb,Tbare the equilibrium values of pressure and the boiling point under normal conditions, Lυis the latent heat of evaporation, Ris the gas constant.

As it follows from equation (9), at constant pressure, the difference of Gibbs energy ΔG is linearly proportional to the overheating/ overcooling

ΔG=SΔT=LeqΔTTeqE20

In the future, assuming the difference of energy ΔG to be equal to the rate of the phase transformation, one can obtain, that in the thermodynamic approach, the rate of conversion at constant pressure for small deviations from equilibrium is linearly proportional to the overheating / overcooling ΔT

υKΔT,E21

where K is the constant of proportionality between the normal speed limits and its overcooling. The constant K does not have any clear physical sense and is chosen experimentally for each material. The dependence (14) by its form coincides with the well known relation for the determination of the linear crystal growth rate obtained on the basis of classical molecular-kinetic models in which the constant of proportionality is called the kinetic coefficient [21]. The main application of the relation (14) and its various modification [22, 23] is found in the description of different processes of melting - solidification. Comparison with experiment showed that the equation (14) gives good agreement mainly at small overcoolings [24, 25]. The kinetic coefficient K is the main parameter, characterizing the mobility of the boundary crystal -melt. Despite the great importance of this characteristic, there are only a few experiments to successfully measure the kinetic coefficient in metals and alloys [26]. The main difficulties of experimental determination are associated with the great complexity of measuring the overcooling at the solidification front. Currently, the main approach to the determination of the quantitative evaluation and qualitative understanding of the physics of the processes behind the coefficient K are the methods of molecular dynamics [27] - [30].

2.2. Kinetic approach

The structure particles of matter are in continuous motion and appear in the main provisions of the molecular-kinetic theory in which all processes are considered at the atomic or molecular level, the particles obey the Boltzmann statistics, and the speed of processes is given by

R=R0exp(Q/kBT)E22

where Q is the activation energy, kBis the Boltzmann's constant, kBTis the average thermal energy for one atom, R0is the pre-exponential factor that affects the process rate. The exponential term exp(Q/kBT) is known as Boltzmann's factor, that determines the part of atoms or molecules in the system that have the energy above Q at the temperatureT.

2.2.1. Melt kinetics

For the first time, the conditions of crystal growth from liquids were formulated by Wilson [31], who suggested that the atoms have to overcome the diffusion barrier in order to make the transition from liquid to solid phase. The rate of accession of atoms to the crystal lattice is expressed by the relation analogous to (15)

Rc(T)=aνexp(Q/kBT)E23

where Q is the activation energy for overcoming the diffusion barrier, νis the frequency of attempts of transitions, ais the atom diameter.

The rate of accession of atoms to the crystal was estimated in [31] using the diffusion coefficient of liquidD=a26νDexp(Q/kBT), that allowed to estimate the velocity of the crystallization front as

υs=6DaLmkBTmΔTTsE24

whereΔT=TsTm, Tsis the temperature of the interphase boundary.

Later, Frenkel [32], using the Stokes-Einstein relation D=kBT3πaη between the coefficients of diffusion D and viscosity η expressed the crystal growth rate in terms of viscosity

υs=2kBTπa2ηLmRTmΔTTsE25

The expressions (17), (18) allowed establishing a linear relationship between the rates of crystal growth and overcooling. Wilson-Frenkel theory [31-33] contributed to a better understanding of the microscopic processes associated with the growth of crystals from the melt. A generalization of the obtained results allowed to obtain that, except for very large deviations from equilibrium, where the homogeneous nucleation mechanism can dominate, the process of melting-solidification proceeds heterogeneously. The heterogeneous nucleation mechanism involves the inclusion of the motion of the liquid–solid interface into the consideration. The velocity of this interfaceυs, as a function of the deviation from the equilibrium melting temperature Tm is called the response function of the interface and is the main value characterizing the processes of crystallization and melting. As with the similar equation (14) obtained in the thermodynamic approach, the equations of Wilson-Frenkel showed a good agreement with experiments for very small overcooling of the boundary, i.e. for small deviations from equilibrium.

Theoretical studies of the velocity of the interface are based on some modifications and generalizations of Wilson-Frenkel theory. Their meaning is reduced to taking into account several factors, such as the latent heat of meltingLm, interatomic distancea, efficiency coefficientf, showing the proportion of atoms that remain in the solid phase at the border crossing. In the modifications [24,34,35], it was considered that the crystal has always lower enthalpy than the melt. This is the amount of energy needed for atoms of the crystal to make the transition from crystal to melt. Escape rate of atoms of the crystal followed by the addition to the active points of the liquid, contains this energy difference in the form of the Boltzmann factor exp(Lm/kBTs)

Rm(Ts)=C1exp(Q/kBTs)exp(Lm/kBTs)E26

The rate of the reverse flow of atoms into the crystal from the melt depends only on the diffusion process in liquid

Rc(Ts)=C2exp(Q/kBTs)E27

whereC1, C2are some constants that should be determined.

In the equilibrium pointTs=Tm, the rates Rm(Tm) and Rc(Tm) are equal, which givesC1exp(Q/kBTm)exp(Lm/kBTm)=C2exp(Q/kBTm), andC1=C2exp(Lm/kBTm).

The velocity of the interphase boundary υs is equal to the difference of the rates Rc(T) и Rm(T)

υs(Ts)= Rc(Ts)Rm(Ts)=C2exp(Q/kBTs)C1exp(Q/kBTs)exp(Lm/kBTs)==exp(Q/kBTs)[C2C1exp(Lm/kBTs)]=C2exp(Q/kBTs)[1exp(LmkB(1Tm1Ts))]==C2exp(Q/kBTs)[1exp(LmkB(TsTmTs))]=C2exp(Q/kBTs)[1exp(LmkB(ΔTTs))]E28
ΔT=TsTm, for overheatingΔT>0, for overcoolingΔT<0.

The constant C2 is associated with other physical constants using the following relation:C2=afν. The studies have shown that the resulting equation (21) well predicts the velocity of the melting-solidification front of silicon [36] in a fairly wide temperature range.

The problems of melting-crystallization of monatomic metals in the modes of rapid heating/ cooling, typical of ultra-short laser irradiation, use a different expression for the velocityυs. It is based on the assumption [37, 38], that crystallization of single-atom metals, (that are characterized by high velocities υs50÷100 m/c [35, 39]) is not diffusion-limited, but is limited only by the collision frequency during transition from liquid to crystal surface. The modification of the equation (21) consists of replacement of the diffusion term with the thermal velocityυT=(3kBT/m)1/2. The velocity of interphase boundary υs is written as:

υs(Ts)=afλ(3kBTs/m)1/2[1exp(LmkB(ΔTTs))]=C(Ts)1/2[1exp(LmkB(ΔTTs))]E29

whereC=afλ(3kB/m)1/2, λis the mean free path, mis the atomic mass.

Thus, the kinetic approach allows us to obtain an expression for the response function υs(Ts) without fitting coefficients and suitable for a wide range of overheating/overcooling. The kinetic dependence (22) is asymmetric for the processes of melting and solidification. However, large deviations from equilibrium require additional modification of the kinetic dependences (21) (22), because they do not take into account the dynamic effects associated with the occurrence of high pressures generated by the high velocity of propagation of phase fronts.

2.2.2. Kinetics of vapor

One of the key processes in the zone of laser irradiation is the transition of condensed matter to the gaseous state. Evaporation process is characterized by high power consumption and large increase in the specific volume of the substance.

Investigation of evaporation process began in the 19th century [40, 41] and continues to this day [42 -]. This fact is defined by practical importance and not fully clarified features of non-equilibrium behavior of matter when it evaporates.

2.2.2.1. The simplest model of kinetics of evaporation in vacuum

The thermodynamic relation (9) gives that at a constant temperature the difference between the Gibbs energy between the two phases, one of which is an ideal gas, is linearly proportional to the pressure difference:

ΔG=VΔp=kBTΔppE30

On the basis of the formula (23), it can be assumed that the rate of phase transformation at constant temperature should be linearly proportional to the pressure difference:

υ=υ0ΔppE31

The simplest model for the growth kinetics of vapor phase was developed by Hertz [40] and Knudsen [41] about 100 years ago. Its formulation is qualitatively the same as the thermodynamic model (23), (24)

jm=jm+jm=12πmkB(psatTsurpT)E32

where jm is non-equilibrium flow of atoms at the surface of evaporation, jmis the flux of atoms that collide with the surface under the assumption that the adhesion coefficient is 1. Formally, this flow can be determined using the relation connecting the vapor pressure with the equilibrium particle flux directed to the condensed surface:p=2πmυjm, where υ is the average velocity in one direction, υ=2kBTm, what givesjm=p2πmkBT. Since the nature of the flux jm() remained undetermined and the values of T and correspondingly p unknown, then during the formulation of the boundary conditions in the problem of evaporation into vacuum it is supposed [20, p.281] to use a single-term version of the Hertz-Knudsen formula to determine the evaporation rate:

jmjm+=psat2πmkBTsurE33
,

that takes into account the connection between the pressure of saturated vapor psat and flux of evaporated atoms at the surface temperatureTsur. The equilibrium vapor pressure psatdepends on the temperatureTsur, as it is shown in the equation (12).

However, the representation of the process of surface evaporation in the form of a simple model, which does not take into account the reverse influence of evaporated atoms, does not remove the internal contradictions inherent in the model of Hertz - Knudsen. Let us write the expressions for the fluxes of momentum and ji and energy je of the particles, moving away from the evaporation surface

ρ<Vz2>=ji(+)+ji()=ji=ρυRTυ+ρυu2ρ<V2Vz>=je(+)+je()=je=ρυu(u22+CpTυ)E34
,

It is easy to see that these fluxes, which completely describe this one-dimensional flow, are impossible to be characterized by any temperature. If we equate these fluxes to the corresponding thermodynamic expressions containing the velocity, temperature and density and find these values, we will see that the system of equations has two distinct complex solutions, which have no physical meaning.

The ambiguity of the solution for real values of the thermodynamic parameters is associated with the possibility of discontinuous solutions of the type of shock wave. The complexity of the solution in this case is due to the thermodynamic non-equilibrium of the evaporation flow of Hertz-Knudsen, which can not be described in terms of thermodynamic concepts. When taking into account collisions in the non-equilibrium layer, the evaporative flux is thermalized, but the temperature on the outer side of this layer no longer coincides with the surface temperature.

2.2.2.2. Approximation of the Knudsen layer

Intense surface evaporation is essentially non-equilibrium process. In addition to the thermodynamic equilibrium, this process also has a gas-kinetic non-equilibrium in a thin (Knudsen) layer of vapor, directly adjacent to the interface. Gas-kinetic non-equilibrium is due to the flow of material through the phase boundary. The mass flow increases with the growth of the evaporation rate and, consequently, the degree of non-equilibrium of the process increases. From the physical considerations, the maximum velocity of material flow on the outside of the Knudsen layer is limited to the local speed of sounduusound=(γRTυ)1/2, orM=u/usound1, where u is the gas-dynamic velocity, Mis the Mach number. The maximum deviation from equilibrium is determined by the maximum value of mass flow, which is known to be achieved atM=1.

Under the conditions of phase equilibrium, when the saturated vapor pressurepsat is equal to the external pressure, the flow of vaporized material is balanced by the return flow of particles and the total mass flux through the boundary is zero. The distribution of particle velocities in vapor is in equilibrium and can be described by the Maxwell function with zero average velocity. In cases where the vapor pressure above the surface is less than the saturated vapor pressure, in the system condensed matter-vapor, the directed movement is formed with u>0 and is characterized by non-zero material flow through the phase boundary. The decreasing reverse flow leads to a deviation from the equilibrium in the distribution of the particles. The magnitude of the flux of returning particles decreases with the increase of the rate of evaporation, and the distribution function at the evaporation surface becomes increasingly different from Maxwellian one.

In general, the non-equilibrium distribution function is found by solving the Boltzmann equation in a region with a characteristic size of a few mean free paths. This area is adjacent to the evaporation surface, where the kinetic boundary conditions are set, taking into account the interaction of individual particles with the interface. Similar problem was solved by various methods in many studies, taking into account, in particular, the differences from unity and variability of the coefficient of condensation, which determines the probability of attachment of the particle in its collision with the evaporation surface. (See, for example. [42 - 46]).

Methods of non-equilibrium thermodynamics are used to describe the evaporation process together with other approximate phenomenological approaches [47 -52]. A more general and fundamental approach is to use molecular dynamics method, which was used in [53, 54] to analyze the evaporation process. A recent review on the issue of non-equilibrium boundary conditions at the liquid-vapor boundary is given in [55].

For the equations of continuum mechanics, thin Knudsen layer is a gas-dynamic discontinuity. The knowledge of the relations at this break, connecting the parameters of the condensed medium and the evaporated material, is needed to deal with the full gas-hydrodynamic problem that arises, for example, during the description of laser ablation, taking into account the variability of the Mach number and instability of the evaporation front [56]. The use of kinetic approaches, which explicitly consider the structure of the Knudsen layer, in such cases is difficult because of the emerging problem of significant difference of space-time scales. The solution of these problems is associated with the additional computational difficulties and is not always possible. Therefore, usually another approach is used that allows to determine the matching conditions with certain assumptions about the form of the non-equilibrium distribution function inside the break [57 - 60] without solving the kinetic problem. Approximation of the distribution function in the Knudsen layer was carried out in different models. But to obtain physically reasonable boundary conditions it is necessary to formulate criteria which these conditions must meet. In addition, attention is paid to the peculiarities of the behavior of the fluxes of mass, momentum and energy as the Mach number tends to unity that allows to use the requirement of extremum of total fluxes of massjm, momentumji and energy je as one of the criteria for M=1

"β"- model. In "β" – model [61, 62], a compound distribution function is used to describe one-dimensional non-equilibrium flow of particles on the inner side of the planar Knudsen layer:f=f(+)+f(). Here, the distribution f(+) for particles, flying out of the surface is given as a Maxwell-shaped function f(+)=f(ρsat,Tsur,0) with density of saturated vapor ρsat for the surface temperatureTsur. The distribution f() characterizes the flow of particles returning to the surface and is supposed to be proportional to a “shifted” Maxwell function f()=f(ρυ,Tυ,u) with densityρυ, temperature Tυ and mean velocity u steady flow of vapor on the outside of the Knudsen layer:
f(+)=f(ρsat,Tsur,0),Vz>0f()=f(ρυ,Tυ,u)=β(m2πkBTυ)3/2exp(Vx2+Vy2+(Vz2u)22kBTυ)Vz<0E35

The flows of mass, momentum and energyjk=jk(+)+jk(),k=m,i,e calculated using f(+),f()must be equal to their gas-dynamic valuesjk that are determined by functionf:

ρ<Vz>=jm(+)+jm()=jm=ρυuρ<Vz2>=ji(+)+ji()=ji=ρυRTυ+ρυu2ρ<V2Vz>=je(+)+je()=je=ρυu(u22+CpTυ),E36

where Cp=γγ+1kB is the vapor heat capacity per particle at constant pressure, for single-atom gasγ=53. From the solution of equations (27), it is possible to obtain gas-dynamic conditions on the break, allowing to determine the magnitude ofρυ, Tυи β in terms ofρsat, TsurandM. The calculations showed that total fluxes jm,ji,je have extrema depending on M at M=0.88,1.18,1.22 correspondingly.

Thus, the requirement of the extrema of all flows in the selected point M=1 in the model [61,62] does not met, that can serve as evidence of a bad choice of the distribution functionf().

"εδ"- model. The drawbacks of the “β” – model can be eliminated by selection of a more general form of the distribution function for the reverse flow of particles f() [63]:
f()=βf+αf1α+β=1E37

All 3 flows.. will have extrema atM=1, if the functionf1 is set to be equal tof1=f(ρ1,T1,u1), where the values of ρ1,T1,u1 are written in terms ρsat,Tsur,ρu,Tυ using additional fitting parameters ε and δ

ρ1,T11/2=ερsatTsur1/2+(1ε)ρυTυ1/2,ρ1,T11/2=εTsur1/2+(1ε)Tυ1/2,u1=δM(γRT1)1/2=δ(T1/Tυ)1/2uE38
.

For example, at ε=0.70 and δ=0.32 all three flows jk will have extrema at M=1 with valuesjm=0.853,ji=0.557,je=0.892. It is clear, that this version of selection of fittings coefficients is not the only one.

"α"- model. It is possible to suggest another phenomenological model [63], where strict localization of extrema jk(M=1) is achieved without usage of fitting coefficients for such distribution functionf(), that do not depend on gas-dynamic values. An example of such function is function
f()=α7f0(ρsat,α2Tsur)E39
,

It takes into account the decrease of temperature Tα=α2Tsur of the reverse flow of particles as compared to the surface temperatureTsur. Due to this change ofTα, the ratio of the normalized fluxes je/jm ceases to be a constant and takes the form(1α8)(1α10), which ensures that the correct limiting value is equal to 1.25 in the equilibrium case forα=1. The equation forα, that is obtained from the equity of fluxes (27) has a relatively simple form:

(1α8)(1α10)(1α9)2=π8γ2M2[(γ1)M2+2](γ1)(1+γM2)2E40

The right-hand side of the equation (28) has a maximum atM=1, that determines localization of the extrema of α andjk. The values of jk(M=1) are equal correspondingly tojm=0.85,ji=0.56,je=0.90.

Modified Crout model. The property of localization of the flows jk(M=1) also present in the model suggested by D. Crout [57]. It uses non-equilibrium function of distribution of particles, written in analytical form with temperature that is anisotropic by directions:TL andTT.

f(ρ,TL,TT,u,v,w)={ρ[π3hTm3(2kBTL)(2kBTT)]1/2exp[m((uu0)22kBTL+v22kBTT+w22kBTT)]}E41

where TL is the lateral temperature along x axis, TTis the transversal temperature along y,z axis; u,υ,ωare the components of the velocity vector along corresponding axesx,y,z, u0 is the drift velocity.

The modification of the Crout model [64] consists of explicit introduction of the Mach number into the main relations that allows to obtain:

Tυ=αΤ(M)Tsur,ρυ=αρ(M)ρsat,E42
αΤ(M)=2γM2(m2+0.5)2(1+γM2)2m2t2,αρ(M)=1exp(m2)+π1/2m(1+erf(m))(1+γM2)m2γM2(m2+0.5)2,E43

The value of m is determined from non-linear equation

F(M)(m2 + 0.5)2 - m2(m2+ a+ 1.5)=0, E44

whereF(M)=1+3γM21(γM21)2,.

a=2t20.5π1/2mt1,E45
. t=2mπ1/2+1+erf(m)exp(m2)+π1/2m(1+erf(m)),
erf(m)=2π0mey2dy.E46
For M=1:  Tυ=0,633Tsur,ρυ=0,328ρsat.
pυ=0.208psat,E47
M=0:  Tυ=Tsur,    ρυ=ρsat,   
 pυ=psat.E48

For the numerical solution of the nonlinear equation (30), one can use the Newton's iterative procedure. All fluxes jm,ji,je have extrema at M=1. Calculations using an anisotropic non-equilibrium particle distribution function give the corresponding extreme values of the fluxes:jm=0.84,ji=0.55,je=0.88.

The calculations show that specific choice of the model has relatively little effect on the magnitude of the momentum flux j2, but significantly affects the flux of mass jmand energyje. Fig.1 shows the dependencies of the ratio of the normalized fluxes j¯3/j¯1=j3j1(+)/j1j3(+) on M for all discussed models. From the comparison of the extreme values of the fluxes jk(M=1) and behavior of the curves jm(M),je(M) it follows, that due to bad choice of f(-), the less favorable for the description of evaporation kinetics is "β" – model.

Figure 1.

Dependency of the normalized ratio of energy and mass fluxes j3/j1 on Mach number for different models: 1 – “ε-δ”-model, 2 – Crout model, 3 - “α”-model, 4 – “β”-model (Knight).

Table 1 for all models shows numerical values of ρυ and Tυ (normalized by ρsat and Tsat correspondingly) depending on the Mach number that changes from zero to unity. Comparative analysis of tabular data, as well as the behavior of the curvesjm,je, show a marked difference of the values of ρυ andTυ, obtained using "β" model, from the values, obtained using other models. The values of ρυ turned out to be underestimated, and Tυ overestimated as compared to their real values at the outer side of the Knuden layer.

“β”-model (Knight)“ε-δ”-model“α”-modelCrout model
MTNTNTnTn
0.01.0001.0001.0001.0001.0001.0001.0001.000
0.10.9600.8610.9580.8640.9600.8690.9530.861
0.20.9220.7480.9160.7530.9200.7580.9100.749
0.30.8660.6540.8760.6620.8810.6660.8700.658
0.40.8510.5760.8370.5870.8440.5880.8330.582
0.50.8170.5110.7990.5250.8080.5240.7980.519
0.60.7850.4570.7630.4720.7730.4700.7630.466
0.70.7540.4100.7270.4280.7400.4240.7300.421
0.80.7240.3710.6930.3910.7050.3860.6970.384
0.90.6960.3370.6600.3600.6720.3550.6650.352
1.00.6690.3080.6280.3330.6400.3280.6330.326

Table 1.

.Normalized values ρυ and Tυ depending on the Mach number for different models

The modified Crout model,"εδ"– model,"α"– model fulfill the requirement of the extremum of the flows jk at M=1. The difference of the results that were obtained using these models does not exceed 1.5%. Any of these models can be used to describe the kinetics of the process of non-equilibrium surface evaporation.

2.3. Conclusions

The performed brief analysis of the kinetics of phase transitions is an introduction to the construction of the models that combine mathematical description of kinetics of high-speed phase transformations with dynamics of the macro-processes (heat and mass transfer) under conditions of a strong deviation from local thermodynamic equilibrium that are typical for ultra-short super-power laser action on metals.

Advertisement

3. Thermodynamic and thermo-physical properties of phonon and electron Fermi gas

Determination of physical characteristics of a medium, including equations of state under conditions of local thermodynamic equilibrium can be carried out either experimentally or by means of calculation using distribution functions - a Maxwell-Boltzmann function for ideal gas and ideal plasma and Fermi one for degenerate electron gas, and for the phonon gas – Bose function. In case of violation of the conditions of local thermodynamic equilibrium distribution functions are determined by solving the classical kinetic Boltzmann equation or quantum-kinetic equations. The presence of distribution function is just the required minimum of information that can be used to describe nonequilibrium processes with reasonable accuracy.

The influence of ultrashort high-energy laser on a strongly absorbing media (metals, semiconductors) is in a very short temporal and spatial scales and leads to disturbance of their general local-thermodynamic equilibrium. Irradiated targets in these conditions are presented in the form of two subsystems - electron and phonon each of which is in local thermodynamic equilibrium and are characterized by their temperatures and equations of state. As a consequence, all processes are described in the two-temperature approximation [65], [66]. The target at pico- and femtosecond influence may be heated to very high temperatures and pressures at which the thermal and mechanical properties of matter are not known in general. One of the most important problems for the mathematical modeling is the necessity to determine thermophysical, optical and thermodynamic properties in a wide (tens and hundreds of electronvolts) temperature and frequency ranges for each of the subsystems.

3.1. Electron subsystem

The most important thermophysical and thermodynamic characteristics of the electron Fermi gas within the scope of heat-conducting mechanism of energy transfer are: heat capacityCe, thermal diffusivityχe and thermal conductivityλe. For its determination using fundamental physical quantities, which include the electron mean free pathslee, lephand the characteristic times (frequency) of interaction for two scattering mechanisms: the electron-electronτee and electron-phononτeph.

3.1.1. Fermi-Dirac integral and its approximation

For a quantitative description of electrical, thermodynamic and thermophysical properties of degenerate electron gas with distribution function

fe=1exp(εeμkBTe)+1E49
,

widely used Fermi-Dirac functionsFk+1/2(μ(Te)/Te) expressed in terms of integrals of the form

Fk+1/2(μ(Te)/Te)=0(εekBTe)k+1/2d(εekBTe)exp(εeμkBTe)+1=0Ek+1/2dEexp(EμkBTe)+1E50
,

whereTe, εe- temperature and energy of electron, μ(Te)- chemical potential, μ(0)=εF- Fermi energy.

In the future, the Fermi integrals will be represented as a function of dimensionless energy of the chemical potential:

Fk+1/2(η)=0Ek+1/2dEexp(Eη)+1E51
,

whereE=εe/kBTe, η(Te)=μ(Te)/kBTe- dimensionless energy and chemical potential of electrons.

The integral of form F1/2(η)=0E1/2dEexp(Eη)+1 is used to determine the electron densityNe=0g(E)f(E)dE=0g(E)dEexp(Eη)+1=0F(E)dE, where g(E)=21/2m3/2π23(kBTe)3/2E1/2=g0(kBTe)3/2E1/2 is a density of states, F(E)=g(E)f(E)=g0(kBTe)3/2E1/2exp(Eη)+1is distribution function of the free electron energy,g0=21/2m3/2π23. Then

Ne=g0(kBTe)3/20E1/2dEexp(Eη)+1=g023εF3/2=g0(kBTe)3/2F1/2(η)E52

From the known distribution function of particle energyF(E), using the ratio<x(E)>=0x(E)F(E)dE0F(E)dE the average value of any physical quantity depends on energy can be found. Since the average energy of the electron gas is defined as the ratio of Fermi integrals

<εe>=0EF(E)dE0F(E)dE=(kBTe)0E3/2dEexp(Eη)+10E1/2dEexp(Eη)+1=(kBTe)F3/2(η)F1/2(η)E53
.

Similarly, we can determine the other physical quantities of the electron gas.

The chemical potential μ(Te) depends on the temperature, so the Fermi integrals can be conveniently represented as a function of dimensionless temperatureξ=kBTeεF. In [67] for the integrals of the form (33) has been proposed convenient approximation, which allows to express the integrals Fk+1/2(ξ) through the transcendental gamma-functions Γ(k+1/2) and the dimensionless temperatureξ:

Fk+1/2(ξ)=Aξ3/2[1+(B/ξ)2]k/2E54
,

whereA=23Γ(k+32)Γ(32), B=[A(k+32)]1/kcoefficients expressed in terms of gamma-functions. Equation (36) has correct asymptotics at ξ0 andξ:

Fk+1/2(η)1/ξk+3/2(k+3/2),ξ0Fk+1/2(η)23Γ(k+3/2)Γ(3/2)1ξ3/2ξE55
.

The integral of k=1/2 order can easily be determined from the expression (34):F1/2(ξ)=231ξ3/2.

The most frequently used Fermi integrals that are expressed through (36) have the form:

F1/2(ξ)=43ξ1/2(ξ2+49)1/2, F3/2(ξ)=(ξ2+(25)2)1/2ξ5/2,
F5/2(ξ)=[52ξ3/2+27ξ7/2]E56
.

The maximum error compared with the exact solution [68] for integrals at k=1 and k=1 does not exceed 8%, but increases slightly with increasing of k.

Approximation (36) allows to obtain simple analytical expressions for the physical quantities of the electron gas at arbitrary temperatures.

3.1.2. Equations of state

Using the approximating expressions (36) and (37) equations of state for degenerate electron gas can be written as simple analytical expressions at arbitrary temperatures. Since the average electron energy <εe> and its pressure can be represented as

<εe>=TeF3/2F1/2=32εF(ξ2+0.16)1/2E57
,
p=23Ne<εe>=23NeTeF3/2F1/2=NeεF(ξ2+0.16)1/2E58
,(39)

3.1.3. Electron heat capacity Ce(Te)

The expression for the heat capacity of the electron gas Ce(Te)=T(Neεe) can be obtained from the relations

ηFk+1/2=(k+1/2)Fk1/2, ηT=3TF1/2F1/2,
Ce(Te)=NeF1/2(52F3/292F1/2)E59
.

Using the approximating expressions (37), heat capacity of electron gas can be represented with an error not exceeding 5% as the following function

Ce(Te)32NekB2Te[K][Te2+(3εFπ2)2]12,E60
,

whereNe=zNa, z - the number of valence electrons, Na - the concentration of atoms (ions) of lattice.

The resulting expression gives the classical linear dependence of heat capacity of a degenerate electron gas vs. temperature [69] in low temperature regionTe<<εF, Ce=π22kB2TeNeεFand constant value at Te>>εF equal to heat capacity of gas with Maxwell distributionCe=3/2kBNe. Dependences for copper and aluminum are shown in Fig. 2.

Figure 2.

Temperature dependence of electron heat capacityCe(Te).

3.1.4. The thermal diffusivity of a degenerate electron gasχe.

The thermal diffusivity of electron gas is proportional to the product of the mean free path le and average velocity of the electron<υe>:

χe=13le<υe>E61
,

In metals electron mean free path due to several mechanisms: pair of electron-electron collisions, electron-phonon collisions and scattering by plasmons.

Electron-electron collisions dominate at temperatures comparable to the Fermi energyTe/εF1. Electron-phonon interaction is dominant at low temperaturesTe/εF<<1. The interaction associated with the excitation of plasmons occurs at high temperatures, exceeding the plasma frequency energy Te>ωr (ωr1020eV). Taking into account high temperature region of occurrence and limitation of experimental data about reducing of the mean free path of electrons due to plasmon excitation (it is known only for some metals), electron scattering by plasmons will not be considered.

3.1.4.1. The electron-electron thermal diffusivity χee(Te)

The mean free path of an electron in pair electron-electron collisions lee is determined from the known gas-dynamic formula

lee=1NeσeeE62
,

where σee - scattering cross section with energy Δε transfer for electrons with energiesε1, ε2. The cross section σee is expressed through the transport cross section of the collision of two isolated electrons σtr and Fermi integralsF1/2,F1/2. In turn, the transport cross section of collision of two isolated electrons σtr in a field of screened Coulomb potential U=e2rexp(rd) is expressed through the differential scattering cross section dσ determined in the Born approximation [70, p.560]. The final cross sections will be written in form

σeetr=2π9(49π)4/3z4/3r4rB2[ln(t+1)t(t+1)](ξ2+0.16),σee=σeetr(F1/22F1/2)2=2π9(49π)4/3z4/3r4rB2ξ2[ln(t+1)t(t+1)](ξ2+49)(ξ2+0.16)E63
,

where t=4k2d2=(9π4)4/3z1/3rBr[(ξ2+0.16)(ξ2+49)]1/2, rB=2me2= 0.529 10-8cm – Bohr radius, r=(34πNa)1/3- the average distance between atoms, e - the electron charge, d – field acting radius (Debye)U=e2rexp(rd).

At low temperaturesTe<<εF, the effective cross section σee is small and amounts toσee=σeetrξ2(ξ2+4/9)σeetr(TeεF)2. Maximum of cross section is achieved atTeεF, Fig.3, when the degeneracy is passed and the electron-electron collisions with large energy transfer become possible. At very high temperatures Te>>εF cross section becomes the Coulomb one, Fig.3, and decreases logarithmically. The mean free path of electrons lee is determined by the formula (42):

lee=1Neσee=[Ne2π9(49π)4/3z4/3r4rB2ξ2[ln(t+1)t(t+1)](ξ2+49)(ξ2+0.16)]1==2π2(49π)1/31r21z1/3Nat2ξ2[ln(1+t)t1+t]1E64

Calculations indicate that the mean free path for Al and Cu change in a wide range (~10-2 ÷ 10-7) cm and have a minimum atTeεF.

The average thermal velocity of electron <υe> is expressed through its average energy<εe>:

<υe>=(2m<εe>)1/2=(3εFm(ξ2+0.16)1/2)1/2=
(32)1/2(9π4)1/3z1/3rme(ξ2+0.16)1/4E65
,

Taking into account (44) and (45) electronic thermal diffusivity χee takes the form:

χee=13lee<υe>=132π2(49π)1/31r21z1/3Nat2ξ2[ln(1+t)t1+t]1(32)1/2(9π4)1/3z1/3rme(ξ2+0.16)1/4=(23)1/243πmet2ξ2(ξ2+0.16)1/4[ln(1+t)t1+t]=χ0Φee(ξ)E66
,

where Φee(ξ)=t2ξ2(ξ2+0.16)1/4[ln(1+t)t1+t] is dimensionless function,χ0=2343πme=0.402,[cm2s].

Fig. 4 shows the temperature dependence of χee(Te) for Al and Cu. Temperature dependence of electron-electron thermal diffusivity χee(Te) of both metals has a deep minimum atTeεF. Its value reaches ~ 20÷30 [cm2/s], Fig.4. At removal of degeneracy, when Te>εF the thermal diffusivity increases due to decreasing of the effective cross sectionσee. With further increase of temperature Te>>εF the thermal diffusivity continues to increase and its dependence χee~ξ5/2/lnξ2 coincides with temperature dependence of thermal diffusivity of Maxwell electron plasma. At low temperature Te<<εF dependence of thermal diffusivity is inversely proportional to square of temperature χee~Te2. Strong growth of χee(Te) with decreasing of Te leads to the fact that χee reach (15)105cm2s values which is 3-4 orders higher than the actual electron thermal diffusivity of metals at room temperatures. Thus, the resulting expression (46) is a good approximation only for high temperatures. Under normal conditions (Te~300K) it is necessary to consider the scattering of electrons of metals by phonons to determine thermal diffusivity, this interaction is dominant at low temperatures.

Figure 3.

Temperature dependences of electron-electron scattering cross sectionσee(Te).

3.1.4.2. The electron-phonon thermal diffusivity

The mean free path, defined by electron-phonon interaction is described by the assumption of elastic scattering of conduction electrons of metal on lattice oscillations. To determine it is convenient to use the phenomenological approach [70], in which the crystal is considered as an elastic continuum. Lattice oscillations at the same time considered as a wave of elastic deformations. To simplify the density fluctuations are presented as deviations of each atom (ion) from the average, which square of amplitude is directly proportional to temperatureTph. According to the macroscopic theory of elasticity we can obtain an expression for the mean free path, expressed in terms of macroscopic quantities [71] by expressing the force tending to return the atom (ion) to the equilibrium state through the Young's modulus E:

eph=ErNekBTphE67

Figure 4.

Temperature dependence of electron-electron thermal diffusivityχee(Te).

From the expression (47) it follows that eph is inversely proportional to the lattice temperature Tph in electron-phonon interaction.

During melting of a metal the number of collectivized electrons remains practically unchanged. The modulus of elasticity is only one value (excluding the jump in specific volume, which usually does not exceed 10%) that changes. The melting of most metals is accompanied by decrease in elastic modulus by 2-3 times [72]. This decrease causes a corresponding increase in density fluctuation and, consequently, an abrupt decrease in the mean free patheph:

(eph)k=(ErzNakBTph)kE68
,

where k=s,l subscripts denoting membership in the solid and liquid phases, respectively, r=(34π1Na)1/3=0.6204Na1/3is a distance between the atoms. From below leph is limited by Bohr diameter value. The mean free path of electrons taking into account the scattering by phonons, calculated for aluminum and copper from the relation (48) showed that, compared with the mean free path of the electron-electron scattering values for both metals decreased by several orders of magnitude: for the high-temperature region of 1.5 ÷ 2 orders of magnitude, while at low temperatures for 4 ÷ 5 orders of magnitude.

The thermal diffusivityχeph(Te,Tph), defined by the mechanism in electron-phonon interaction is calculated by the formula:

(χeph(Te,Tph))k=(13eph(Tph)<υe(Te)>)kE69
,

Dependences χeph(Te,Tp) for Al and Cu calculated from (49) at Te=Tph are shown in Fig.5. The calculations show that at room temperature values of the electron thermal diffusivityχeph, for both materials are 3-4 orders of magnitude less than the thermal diffusivity χee and reach values of 102 cm2/s typical for metals under normal conditions. Under equilibrium conditions with temperature Tph increasing, thermal diffusivityχeph, undergoing break at the phase transition decreases rapidly to a value of ~ 1 cm2/s at a temperature of ~ 5 000K. Function χeph(Te,Tp) increase by several orders of magnitude with electron temperature increase.

Figure 5.

Temperature dependences of thermal diffusivity χeph(Te,Tph) for Al and Cu with fixedT=Te=Tph.

3.1.4.3. The resulting thermal diffusivity of electron Fermi gas χe(Te,Tph)

Calculations have shown that taking into account only a pair electron-electron collisions leads to a strong (by several orders of magnitude) overestimation of the thermal diffusivity of the electron gas at low temperaturesTe<<εF. Accounting for electron-phonon collisions gives a more realistic values of χe(Te) at low temperatures.

By averaging mean free paths ee,eph we obtain an expression for resulting thermal diffusivity, at arbitrary temperature

(χe(Te,Tph))k=(χeeχephχee+χeph)k=(13e<υe>)k=<υe>(eeeph3(ee+eph))k=s,E70
,

Fig.6 shows temperature dependences of total electron thermal diffusivity χe(Te,Tph) for Al and Cu in equilibrium Te = Tp.

3.1.5. The thermal conductivity of the electron gas

According to elementary kinetic theory, the thermal conductivity of the gas is

(λe(Te,Tph))k=(13Cee(Te,Tph)<υe>)k=(Ceχe(Te,Tph))k,[WmК]E71

Thus, the thermal conductivity λe can be determined through the heat capacity Сe and averaged thermal diffusivity of electron gasχe(Te,Tph). The temperature dependences of λe for Al and Cu calculated for the equilibrium case when Te=Tph are shown in Fig.6. In accordance with the results, the equilibrium electron thermal conductivity λe(Te,Tph) at temperatures not exceeding the boiling temperature of the equilibrium is practically independent of temperature. In high temperature region Te>1eV thermal conductivity increases rapidly due to the dominance of electron-electron scattering. It is natural that in this region the thermal conductivity of the electron gas depends on electron density and λe increases with increasing of electron concentration. For this reason, electron thermal conductivity of aluminum is higher than the same one of copper at the high-temperature. In low temperature region where electron-phonon interaction is dominated, the ratio is inverse. Because of the greater mean free patheph, electron thermal conductivity of copper is higher than that of aluminum.

Figure 6.

Temperature dependence of total electron thermal conductivityλe(Te,Tph).

3.2. Phonon gas

3.2.1. Heat capacity of phonons

The main consequence of the existence of lattice oscillations is the possibility of its thermal excitation, which is appeared as a contribution to the heat capacity of solid.

Taking into account the process of melting, heat capacity of phonon gas can be written as [69], [73]:

Cph,k={3kBNa,k,k=s,l,T>TD12π45NakB(TphTD)3,T<TD,[Jm3K]E72
TDis a Debye temperature.

3.2.2. The thermal diffusivity of phonon gas χph

Phonons are considered as a gas of particles. From elementary kinetic theory, thermal diffusivity of a gas is given by

χph=13phυsoundE73

where ph - phonon mean free path. It is assumed that the phonons move with velocity of soundυsound.

The mean free path of phonons ph is determined from the description of thermal motion in a solid by means of notions of the phonon gas. The interaction between the phonons can be characterized by some effective cross section which is proportional to the mean square of thermal expansion of the body or the mean square of density fluctuation ρ [69]:Δ2=NakBTpβ, where β is a coefficient of compressibility. Taking into accountβ=(ρυsound2)1:

Δ2=NakBTphβ=NakBTphρυsound2=kBTphMυsound2E74

Assigning to phonons radius equal to thermal oscillations amplitude, we can taken into account scattering of sound waves and determine the mean free path ph

ph=rΔ2γ2=Mυs2kBTphrγ2,E75

where M - the mass of the atom, γ- Griuneyzen constant.

Accounting for the expressions for the mean free path ph (54) and the velocity of sound υsound [73] expressed through the Fermi velocity

υsound=(m3M)1/2υF=(m3M)1/2(9π4z)1/3m1r,M=M0A,E76

thermal diffusivity of phonons can be written as:

χph=13phυs=M0AυS33kBTphrγ2=(34π)1/3π2(3m)3/23(M0A)1/2zγ2Na2/3kBTphE77

Taking into account the melting process:

(χph)k=(2.831×1013zA1/2γ2Na2/3Tph[K])k[cm2s],k=s,lE78

Figure 7.

Temperature dependences of phonon thermal diffusivityχph(Tph).

3.2.3. Thermal conductivity of phonon gas λph

The thermal conductivity of phonon gas can be expressed through its heat capacity Cph and thermal conductivity. Taking into account expressions for Cph and χph in the high-temperature region Tph>TD thermal conductivity takes the form:

(λph)к=Cphχph=(3π2(3m)3/23M01/2zA1/2γ2Na5/3Tph[K])k=(1.17×1035zA1/2γ2Naк5/3Tph[K])кE79
,

Graphic dependences of λph for Al and Cu are represented in Fig. 8.

Figure 8.

Temperature dependence of the phonon thermal conductivity.

3.2.4. Equilibrium heat capacity and thermal conductivity of metal

The temperature dependences of the total heat capacity and thermal conductivity of a metal can be represented as corresponding sums, which consist of two parts – electron (40), (51) and phonon (52), (55):

C(T)=Ce(T)+Cph(T),λ(T)=λe(T)+λph(T)E80
,

Temperature dependences of λ(T) and C(T) are shown in Figs 9, 10. For comparison, reference data from [72], [74] - [77] are shown by markers.

According to the equilibrium theory of metals [69], [73] in the temperature range T0 < T < Tm heat transport by electrons is dominated. As predicted by the Wiedemann-Franz law, equilibrium electron thermal conductivity is practically independent of temperature. Its contribution to the total thermal conductivity of the metal with temperature increasing, is much greater than the contribution of the phonon component λе > λph. The phonon part of thermal conductivity is inversely proportional to the temperature, and its contribution is noticeable only at low temperatures. As a result, the resulting thermal conductivity of solid-state becomes linearly decreasing temperature dependence. At the phase transition solid - liquid thermal conductivity of both metals decreases abruptly in 2 ÷ 3 times and decreases slowly with increasing temperature.

At very low temperatures, much smaller than Debye temperature T<TD, these relations become invalid, which is associated with the change of interaction mechanisms. Since the thermal conductivity of metal bodies at low temperatures is determined by the transfer of thermal energy by sound waves, rather than electrons. The heat capacity and the number of phonons in this case are proportional to ~ T3, and the electrical resistance, defined by the scattering of electrons by phonons is inversely proportional to absolute temperature.

Figure 9.

Temperature dependences of total thermal conductivityλ(T).

Calculations showed that at temperatures close to the plasma T≥1 eV, the thermal conductivity of metals increases rapidly, and the thermal conductivity of aluminum, due to three times exceeded electron density becomes higher than the thermal conductivity of copper. However, this situation can have meaning only in the case of strongly nonequilibrium with hot electrons and cold lattice. Under equilibrium conditions, the use of calculated data should be restricted to the region of the critical point, in which neighborhood, as known, all physical properties of metals vary sharply.

Comparison of the obtained theoretical data with reference data was carried out in a relatively narrow temperature range, since the physical properties of most metals were measured in the temperature range from several tens to several thousand degrees, and usually not exceeding the boiling temperature of the equilibrium. In the range 300 ÷ 1500 K, the comparison has shown a complete qualitative agreement of results with a good quantitative correlation, Figs.9, 10.

3.3. Electron-phonon interaction

During laser action on metals, entire pulse energy is transmitted directly to the electrons. The result is a highly nonequilibrium region in the solid with hot electrons and cold lattice. Cooling of electron subsystem realized by two mechanisms: energy transfer by electron thermal conductivity, leveling the gradients of electron temperature and electron-phonon interaction, which leads to heating of the lattice.

Figure 10.

Temperature dependences of total heat capacityC(T).

3.3.1. The average frequency of electron-phonon collisions νeph.

Per time unit, the electron undergoes an average of ~ 1/2 τ collisions with the emission of phonons, and roughly the same, but slightly smaller, the number of collisions with the absorption of phonon. Estimate the time of inelastic collisions by using the following equation [70]

12τeph=1NeυeσephE81
.

Hence, the average frequency of electron-phonon collisions with energy transfer of electrons to crystal lattice can be represented as

νeph=2Ne<υe>σeph=2zNa<υe>σephE82

where σeph is scattering cross section of electrons by phonons.

The cross section for electron-phonon scattering, calculated under the assumption that the product of the wave number of electron and atom displacements from equilibrium [78] has the form

σeph=4π(49π)4/3mkBTph2(rrB)2r4z4/3ξ[ln(t1+1)t1(t1+1)](ξ2+0.16)1/2(ξ2+49)1/2E83

The cross section has a maximum atTeεF. At Te<<εF due to the Pauli principle, the cross section is proportional to temperatureTe. After removing the degeneracyTe>>εF, cross section of electron-phonon collisions decreases with Te increasing. The linear lattice temperature Tph dependence of the cross section of electron scattering in metals reflects the fact that the effective radius of the atoms can be identified with the amplitude of its thermal fluctuations.

Taking into account expressions for σeph (59) and <υe> (45), the average rate of energy transfer can be written as:

νeph=(32)1/283π(rrB)2kBTphΦeph(ξ){Φeph(ξ)=ξ[ln(t1+1)t1(t1+1)](ξ2+0.16)1/4(ξ2+49)1/2при   ξ<1Φeph(ξ)=ξ[ln(t1+1)t1(t1+1)](ξ2+0.16)3/4(ξ2+49)1/2приξ1}E84

The dependences of the mean frequency of transmission electron energy to phonons from the temperature Te at Tph=300 K and Tph= 1000 for aluminum and copper are shown in Fig.11.

Figure 11.

Temperature dependences of frequency of energy transfer of electrons to phonons at 300 K and = 1000 K.

3.3.2. Energy transfer

The mechanism of energy transfer to the lattice due to electron-phonon collisions was discussed in several papers [79-81]. The collision of an electron with the lattice phonon is absorbed or is born, for which law of conservation of energy and momentum works:

Δε=mυ12/2mυ22/2=±Ωp=2mυFsinφ2ΩυsoundE85

where Ω - the phonon frequency, φ- the angle between the initial and final electron momentum mυ1 andmυ2, υsound- velocity of sound, p- the electron momentum.

From the conservation laws (61) it follows that energy of phonon, excited by an electron with a momentum p equal toΔε=Ω=2pυsoundsinφ2.

At a spherically symmetric scattering the energy transferred by one electron per unit time, averaged byφ, equals [81]:

<Δε>=23ΩDkBTphpυsoundνeph[1132(ΩDpυsound)3], where ΩD - the Debye energy.

Taking into account velocity of soundυsound, expressed in terms of the Fermi velocity [73], <Δε>in its final form is written:

<Δε>=23pFυsound2kDkBTph[1132(kDυsoundpFυsound)3]νephkBTph==(9π2)1/3(32)1/243zmM(rrB)23r4×[1(124z)]Φeph(ξ),E86

3.3.3. The electron-fonon coupling factor

The average energy transferred to phonons by electrons per volume unit per time unit is equal to the product of the average energy transferred by a single electron <Δε> and concentration of electrons, which may participate in the transfer of energy to phonons

dεdt={<Δε>kB(TeTph)εFNeTe<εF<Δε>z(Te)NeTeεFE87

or

dεdt=<Δε>kB(TeTph)εFNe=g(ξ)(TeTph)E88
,

where g(ξ) is electron-fonon coupling factor, equal to

g(ξ)=ΔεkBNeεF=(32)1/2(89π)1/383kBrB2M0z2ANa[1(124z)]Φeph(ξ),[Wm3K].E89

The temperature dependence of the electron-fonon coupling factor g(Te) is shown in Fig.12. The results calculated for Al in [82] are shown to compare by markers.

Figure 12.

Temperature dependences of electron-fonon coupling factorg(Te).

3.4. Conclusions

Important thermodynamic and thermophysical properties of a degenerate electron and phonon gas: equation of state, heat capacity, the energy exchange between subsystems in an arbitrary temperature range are determined using fundamental physical quantities - the mean free paths and times (frequency) of the electron - electron and electron-phonon collisions.

Using Fermi integrals technique and its subsequent approximations have provided the temperature dependence of all characteristics in the form of simple analytical expressions.

Advertisement

4. Formulation and solution of the problem of laser action on metals using the method of dynamic adaptation

Action of intense laser radiation on metals may be accompanied by heating, evaporation, plasma formation. The dynamics of the processes in the condensed media and in the flow of vapor and gas environment surrounding the evaporating surface depend on several parameters of the laser pulse: the level of the absorbed intensity, wavelength, duration and spatial and temporal distribution of the energy, as well as optical, thermal and hydrodynamic characteristics of the condensed and gaseous media. It should be noted that the process of interaction is qualitatively different depending on where the target is placed: in the vacuum or the gaseous medium.

As it was already mentioned, the main features of ultra-short super-power action on metals are associated with the high speed and voluminous nature of the laser pulse energy release. The high rate of energy input leads to a strong deviation of the system from the local thermodynamic equilibrium (LTE). The basis of the mathematical description of the pulsed laser heating of metals is a phenomenological two-temperature model (TTM) of parabolic type proposed in the 50s by the authors of [65], [66]. The application of TTM involves the description of the situations with small deviations from LTE. Ultrafast heating of metal targets by super-power laser pulses causes a strong deviation from the LTE and requires appropriate adjustments of the mathematical model. In the first place, the kinetic non-equilibrium phase transitions and powerful dynamic effects associated with the movement of the phase fronts must be taken into account. The consideration of these processes requires the explicit description of the kinetics of the phase transitions and the formulation of the conservation laws at the phase fronts, which are hydrodynamic breaks.

4.1. Mathematical description of processes in the irradiation zone

Laser radiation propagates from the right to the left and strikes the surface of the metal target. Then the radiation is partially absorbed and partially reflected. The absorbed energy is consumed for heating, phase transformations and the generation of shock waves in the solid phase. Fig.13 shows spatial location of phases, moving interphase boundariesΓs(t), Γυ(t)and shock wave in solidΓsh,s(t).

Figure 13.

Spatial phase configuration.

Formulation of the problem was carried out under the following assumptions and limitations:

  • The mechanisms of volume of melting and evaporation are not included into the consideration.

  • The melting and evaporation fronts appear at the irradiated surface during its noticeable overheating ΔTs=TsurTm andΔTυ=TsurTb(psat), the overheated metastable states act sustainably during the consideration.

The mathematical description and modeling of pico- and femtosecond laser action on metal target in vacuum was performed within the framework of a two-temperature and spatially one-dimensional multi-front nonequilibrium hydrodynamic Stefan problem, written for the two phases - solid and liquid.

4.1.1. System of equations

Action of high-power laser pulses on metal targets initiate in them a number of interrelated processes with nonlinear behavior. The description of their behavior is performed using a system of hydrodynamic equations, two energy balance equations for electron and phonon subsystems, and the transport equation of the laser radiation. The system of equations can be written in the domain with three moving boundariesΓsh,s(t),Γs(t) ,Γkυ(t).

(ρt+(ρu)x=0(ρu)t+(ρu2)x+px=0(ρeεe)t+(ρeuεe)x=(pux+Wex+g(Te)(TeTph)+Gx)(ρεph)t+(ρuεph)x=(pux+Wphxg(Te)(TeTph))Gx+α(Te)G=0,ρe=zmMρ,P=Pe(ρe,Te)+Pph(ρ,Tph),εe=Ce(Te)Te,εph=Cph(Tph)Tph)k=s,t>0,Γs<x<Γsh,s(t)Γsh,s(t)<x<Γs(t)Γs(t)<x<Γv(t)We=λ(Te,Tph)Tex,Wph=λ(Tph)Tphx,P(ρ,T)=P(ρe,Te)+P(ρ,Tph),E90
,

the designations: ρ,u,ε,T,pare the density, gas-dynamic velocity, internal energy, temperature and pressure correspondingly, α(Te), R(Te)are the coefficient of volume absorption of the laser radiation and surface reflectivity, G is the density of the laser radiation, Ce(Te), Cph(Tph), λe(Te,Tph), λph(Tph)are electron and phonon heat capacity and heat conductivity, g(Te)electron-phonon coupling factor. Indexess,,υ, designate solid, liquid, vapor phases, e,phdesignate electron and phonon components.

4.1.2. Edge conditions

We use edge conditions in the form of initial conditions at t=0 and boundary conditions atx=Γs,Γsh,s(t),Γs(t),Γv(t).

Initial conditions

t=0:u(0,x)=0,p=0,ρ=ρ0,Te=Tph=T0=293KE91

Boundary conditions

x=Γs: the condition of zero mass and heat flow is used as boundary conditions at the left (fixed) boundary:x=Γs:
ρsus=0,WT=0;E92
x=Γs(t): the model of surface melting-crystallization is used as boundary conditions at the moving interphase boundaryΓs(t).

The model of surface melting-crystallization is a non-equilibrium kinetic version of the Stefan problem [18], [83,84], formulated for the conditions of significant deviation from LTE and consisting of three conservation laws: mass, momentum and energy, written in the stationary (laboratory) coordinate system:

x=Γs(t):
ρs(usυs)=ρ(uυs)E93
Ps+ρs(usυs)2=P+ρ(uυs)2,E94
(λphTphx)s(λphTphx)=ρsLmneυs,E95

These conservation laws are accompanied by the kinetic dependence of the interphase front velocity υs(ΔTs)on the overheating of the solid surface (22).

υs(Ts)=afλ(3kBTs/m)1/2[1exp(LmkB(ΔTTs))]E96
,

Additional account of hydrodynamic effects was carried out using the expressions for the curve of equilibrium meltingTm(Ps), temperature dependence of the equilibrium latent melting heat Lmeq(Tm(P)) and non-equilibrium latent melting heat Lmne

Lmne=Lmeq(Tm(P))+ΔCpsΔTs+ρs+ρρsρ(usu)22,
Tm=Tm(P)=Tm,0+kP,E97

whereΔCps=(CpCps),

ΔTs=(TsTm(P))E98
.

The electron component is assumed continuous with respect to the electron density Neand temperature Teduring transition through the phase front:

(λeTex)s=(λeTex),Te,s=Te,E99
.x=Γkυ(t): the model of surface evaporation is used as boundary conditions at the moving interphase boundaryΓkυ(t). The process of surface evaporation within the approximation of the Knudsen layer is described by three conservation laws and three additional parameters at the outer side of the Knudsen layer (temperatureTυ, density ρυ and velocityuυ). In the general case, two of these parameters (usually Tυ andρυ) are determined using certain approximation relations, e.g. (30), while the third one (usually the Mach numberM=u/uc) is found from the solution of the gas-dynamic equations.

Three conservation laws are written in the stationary (laboratory) coordinate system and have the form:

x=Γkυ(t):ρk(ukυkυ)=ρυ(uυυkυ),E100
pk+ρk(ukυkυ)2=pυ+ρυ(uυυkυ)2E101
,
(λphTphx)k(λυTυx)υ=ρk(ukυkυ)Lυne,E102
ρυ=αρ(M)ρsat,Tυ=αT(M)Tsur,E103
psat(Tsur)=pbexp[LυneRTb(ΔTsurTsur)],ρsat=psat(Tsur)RTsur,ΔTsur=TsurTb(psat(Tsur))Lυne=Lυeq(Tsur)+Cpυ(TυTsur)+ρk+ρυρkρυ(ukuυ)22,E104

The boundary conditions for the electron component and the laser radiation are written as:

λeTex=σTe4,E105
G(t)=(1Rk(Te))G0exp((tτL)2)E106
.σis the Stefan-Boltzmann constant.x=Γsh,s(t): three conservation laws are written at the moving front of the shock wave, so-called Rankine-Hugoniot relations, that have the following form in the laboratory coordinate system:
jsh,sm=ρ1(u1υsh,s)=ρ0(u0υsh,s)E107
,
jsh,si=p1+ρ1(u1υsh,s)2=p0+ρ0(u0υsh,s)2E108
,
jsh,se=WT,1+jsh,sm[ε1+(u1υsh,s)22]=WT,0+jsh,sm[ε1+(u0υsh,s)22]E109

The index “0” signifies the values at the background side, “1” relates to the shock wave side.

4.2. Numerical algorithm and finite difference schemes

Mathematical feature of Stefan problems is the lack of explicit expressions for the interconnected quantities at the interphase boundaries: temperatureTs, Tsurand velocities of the phase fronts υs andυkυ. So the problem (65) – (81) will be nonlinear even for constant values of thermo-physical and optical properties. For low velocities of phase transformations (υs,υkυ<<υsound), the processes behave in a quasi-equilibrium way, and their description can be performed within the framework of equilibrium models. For example, in the problem of melting-crystallization, a phenomenological condition of the temperature equity Ts=Ts=T=Tmis used instead of the kinetic condition (71), and the differential Stefan condition is completely omitted. The influence of the phase transition is taken into account using a singular adding of the latent heat Lmeq to the heat capacity (equation of state) in the point of the phase transition, so-called enthalpy method [85 - 87]. To solve the problems in this statement, the methods of the type of “pass-through” or uniform schemes have found wide application [85], where the velocity and location of the phase transition are not determined explicitly.

Fast phase transitions (υs,υkυ<<υsound), that are typical for powerful pulsed laser action, occur under the conditions of high non-equilibrium caused by the powerful flow of material [88] over the interphase boundary. Using the smoothing procedure of the enthalpy reduces the class of solutions of the phase transformations problems in the material; in particular, it excludes from consideration the effects of overheating and overcooling of the condensed matter [89, 90]. High-speed phase transitions require explicit tracking of the phase boundaries [91]. Typically, the velocity of phase boundaries is determined numerically. In computational respect, the presence of moving boundaries leads to a significant complication of the numerical solution [92].

4.2.1. Method of dynamic adaptation

Finite-difference method of dynamic adaptation [93], [94],was used to numerically solve the discussed problem. This method allows performing explicit tracking of any number of interphase boundaries and shock waves [95].

The method of dynamic adaptation is based on the procedure of transition to an arbitrary nonstationary coordinate system. Its usage allows to formulate the problem of grid generation and adaptation on a differential level, i.e. part of differential equations in the obtained mathematical model describes physical processes, and other part describes the behavior of the grid nodes [96].

4.3. Arbitrary nonstationary coordinate system

The transition to an arbitrary nonstationary coordinate system is performed using an automatic coordinate transformation using the sought solution. Formally, it is possible to use a variable change of the common form x=f(q,τ), t=τ to perform a transition from the physical space Ωx,t with Euler variables (x,t) to some computational space with arbitrary nonstationary coordinate system Ωq,τ and variables(q,τ). It is assumed that this transformation corresponds to a univalent nondegenerate reverse transformationq=ϕ(x,t),    τ=t. During the transition between coordinate systems, the partial derivatives of the dependent variables are connected using the following expressions:

t=τ+Qq,x=1ψq,2x2=1ψq1ψqE110
,

where ψ=x/q is the Jacobian of the reverse transformation, Qis the transformation function that should be determined.

In the arbitrary coordinate system, moving with velocityQ, wherexτ=Q, orψτ=Qq, the differential model (65) in the new variables (q,τ) takes the form:

ψτ=Qq,ψ=xqE111
τ(ψρ)+q(ρ(u+Q))=0E112
τ(ψρu)+q(p+ρu(u+Q))=0E113
τ(ψρeεe)+q(ρeεe(u+Q))=(puq+Weq+g(Te)(TeTph)+Gq)E114
(ψρεph)τ+q(ρεph(u+Q))=(puq+Wphqg(Te)(TeTph))E115
Gq+ψα(Te)G=0E116
p=pe(ρe,Te)+pph(ρ,Tph),εe=Ce(Te)Te,εph=Cph(Tph)Tph,W=λ(T)ψTq,E117

Thus, usage of an arbitrary nonstationary coordinate system is accompanied by the transformation of the initial differential model (65) into the extended model (83) – (88), in which the equation (83) is the equation of the inverse transform. Equations (84) – (88) describes the physical processes. The type, characteristics and kind of boundary conditions for the inverse transformation equations depend on the type conversion functionQ. Because of this equation, the initial (66) and boundary conditions (67) – (81) are changed accordingly.

Thus, the unknown functions in the computational space are not only the functions of the physical fields but also the coordinates of the grid nodes. The equation of the reverse transformation is used after the determination of the function Q to construct the adaptive grid in the physical space. Its differential analogue describes the dynamics of the grid nodes, and the function Q performs controlled movement of the nodes in agreement with the dynamics of the sought solution. In the computational space, the nodes of the grid and all discontinuous and interphase boundaries are steady. The value of the function Q at the boundaries of the domain are determined from the boundary conditions.

Selection of the functionQ. The function Q that is in agreement with the sought solution can be determined from the quasi-steady principle [96], which states that we should search for such nonstationary coordinate system where all processes occur in a steady way. Since the energy balance equation includes all hydrodynamic variables, to determine the functionQ, we can use only two equations (86), (87), assuming that εeτ=εphτ= 0.Then the function Q will have the form

Q=u2puq+(+Weq+Wphq+Gq)(ρeεeq+ρεphq+re)E118

whereWe=λeΨTeq,

Wph=λphΨTphqE119

Differential schemes

The differential model (83) – (88) was approximated by a family of conservative differential schemes obtained using integration-interpolation method [97]. Computational grids (ωmj)s and (ωmj)are introduced in the computational space Ωq,τin each subdomain with integer qi and half-integer qi+1/2 nodes for the spatial variable q and step Δτj for the variableτ.

(ωmj)k={ (qm,τj),(qm+1/2,τj);qm+1=qm+h,qm+1/2=qm+0.5h,τj+1=τj+Δτj,i=0,..,N1,j=0,1,...}k=s,E120

The flow variablesWe,, Wph,, G, and functionsu,Q,x, correspond to the integer nodes during approximation, while functionsεe, εph, Te,Tph, ρ,ψ correspond to half-integers ones(qm+1/2,τj).

The family of conservative differential schemes has the form:

ψm1/2j+1ψm1/2jΔτj=Qmσ1Qm1σ1hψm1/2j+1ρm1/2j+1ψm1/2jρm1/2jΔτj=ρmσ2(Qmσ1umσ3)ρm1σ2(Qm1σ1um1σ3)humj+1(ψm1/2j+1ρm1/2j+1+ψm+1/2j+1ρm+1/2j+1)umj(ψm1/2jρm1/2j+ψm+1/2jρm+1/2j)2Δτj=pm1/2σ4pm+1/2σ4h++ρmσ2(Qmσ1umσ3) um+1σ3+ρm+1σ2(Qm+1σ1umσ3) umσ32hρmσ2(Qmσ1umσ3) um1σ3+ρm1σ2(Qm1σ1um1σ3) umσ32hE121
(ψm1/2j+1ρe,m1/2j+1εe,m1/2j+1)(ψm1/2jρe,m1/2jεe,m1/2j)Δτj={[ρe,mσ2εe,mσ5(Qmσ1umσ3)ρe,m1σ2εe,m1σ5(Qm1σ1um1σ3)]h++pm1/2σ4(um1σ3umσ3)h+2[λmσ6Tm+1/2σ7Tm1/2σ7ψm+1/2σ8+ψm1/2σ8λm1σ6Tm1/2σ7Tm3/2σ7ψm1/2σ8+ψm3/2σ8]h+[g(Te)σ9(TeTph)σ10]m1/2++(Gmσ11Gm1σ11)h+(We,mσ12We,m1σ12)h},E122
(ψm1/2j+1ρm1/2j+1εph,m1/2j+1)(ψm1/2jρm1/2jεph,m1/2j)Δτj={[ρmσ2εph,mσ13(Qmσ1umσ3)ρm1σ2εph,m1σ13(Qm1σ1um1σ3)]h+pm1/2σ4(um1σ3umσ3)h++2[λph,mσ14Tph,m+1/2σ15Tph,m1/2σ15ψm+1/2σ8+ψm1/2σ8λph,m1σ14Tph,m1/2σ15Tph,m3/2σ15ψm1/2σ8+ψm3/2σ8]h++[g(Te)σ9(TeTph)σ10]m1/2+(We,mσ16We,m1σ16)h},E123
Herefσr=σrfj+1+(1σr)fj, and σr=σ1,σ2,...□are the weight coefficients, determining the degree of implicitness of the differential schemes.

To solve the obtained system of nonlinear partial-differential equations we use a computational algorithm with enclosed iterative cycles consisting of one external and two internal cycles [98]. Each of the internal cycles uses the Newton iterative procedure.

Advertisement

5. Modeling results

We consider two modes of pulse with pico - and femtosecond laser fluence with a wavelength λL=0.8µm on two metal target of aluminum (Al) and copper (Cu). The influence on each of the targets was carried out with the same energy densityF, respectively F=0.7Jcm2 for Al and F=2.0Jcm2 for Cu.

Temperature dependences of the reflectivity R(Te) and volumetric absorption coefficient α(Te) were calculated according to [99, 100]. Fig. 14 shows the time profiles of the incident and the absorbed laser pulse with a Gaussian profile τL=1012s and τL=1015s.

Figure 14.

Time profile of pico- and femtosecond pulses on the surface of Al.

Figure 15.

Time dependence of the surface temperature of Al.

The incident laser radiation is absorbed in the electron component, causing it to warm up fast. The main feature of the nonequilibrium heating is that it occurs under conditions of intense competition between the rapid release of the laser energy and the factors limiting the heat - strongly varying electron thermal conductivity coefficient λe(Te,Tph) and electron-fonon coupling factorg(Te).

At picosecond influenceτL=1012, the maximum of electron temperature of the surface Te,max1.8104K is achieved at the descending branch of the laser pulse at the momenttτL, Fig. 15, and the values of the coefficients vary in the range λe,max/λAleq10 andgmax/gAleq70. The maximum value of the lattice temperature is achieved after the pulse t5τL and amountsTph,max7.5103K. The maximum gap between the temperatures reaches the valueΔT=TeTph104K. Temperature equalization TeTph occurs within a time ~ 10 ps.

Femtosecond influence τL=1015s differs from picosecond that the same energy is released during the 3 orders of magnitude less than that causes more of a deviation from local thermodynamic equilibrium. The values of the coefficients λe(Te,Tph) and g(Te) limiting the heating of electronic components, and also increase λe,max/λAleq27 andgmax/gAleq200, but not as much asG(t). Time profile of the electron temperature of the surface Te,surf(t) significantly shifted relative to laser pulse time profileG(t). The maximum value of the electron temperature is achieved at the end of the pulse t2τL and isTe,max4.1104K, Fig. 15. A second difference of a femtosecond influence, despite the rapid energy exchange between the subsystems is a slow heating of the lattice, which differs little from the effects of heating at picosecond influence. The maximum value of the lattice temperature is achieved at the time t5 ps and isTph,max7.2103K. The maximum gap between the temperatures reaches a valueΔT=TeTph3.3104K. Temperature equalization TeTph occurs within the time ~ 9 ps.

The high heating rate determines high rate of phase transformations. Fig. 16 shows the time dependences of the melting front propagation velocity υsl(t) for two modes of influence. Since the rate of heating of the lattice by pico - and femtosecond pulses differ slightly, then the rate υsl is almost the same. The only difference is in the initial stage of the process. At the picosecond influence, the birth of the melting front starts at the front of the laser pulset0.1ps, while at the femtosecond with a considerable delay after the pulse,t30fs. The maximum values of υslmax(t)3.75 km/s are achieved in the time intervalt2÷2.5ps. Such a high rate of propagation of the front leads to a pressure jump at the melting surface of the solid phase ps=0.16Mbar for τL=1012 s (ps=0.14Mbarfor τL=1015 s) and the formation of a shock wave moving ahead of the front of melting, Fig. 17.

Appearance of high dynamic pressure causes an increase in the equilibrium melting temperature, the curve Tm(ps) shown in Fig. 18. Its maximum value reachesTm(ps)1900K. Due to the dynamic pressure of about one order of magnitude as compared to the equilibrium value, increases the value of non-equilibrium heat of meltingLmne(ps).

Comparison of the Ts(t) and Tm(ps(t)) curves indicates a significant overheating of surface of the phase boundaryTs(t)>Tm(ps(t)). The degree of nonequilibrium of melting process is conveniently characterized by the response functionΔTs(υsl), which is shown in Fig.19.

Figure 16.

The time profile of the velocity of melting of aluminum.

Figure 17.

The spatial profiles of temperature of aluminum.

Figure 18.

The time profiles of the equilibrium melting temperature and temperature of the melting front of Al.

Figure 19.

Response functions ΔTs(υsl) for Al.

At low overheating (supercooling), the function ΔTs increases linearly with growthυsl. In the range of υsl ~ 1 km/s, at which the shock wave is formed in the solid phase, ΔTsfunction has a smooth maximum indicating the change of the mechanism of phase transformation. The thermal conductivity mechanism is dominant at low velocities υsl1km/s will be displaced by hydrodynamic at large valuesυslυsound.

Fig. 20 shows fragment of dependence of υsl(t)>0 which characterizes the hardening process which takes place in the nanosecond time range. Maximum velocity of the solidification front does not exceed 45 m/s, respectively, the maximum undercooling is10÷15K. High velocity of propagation of melting fronts are associated with strong cross-flows of matter and energy across the phase boundary.

Figure 20.

The time profiles of the velocity of solidification of aluminum.

In aggregate with volumetric nature of the energy transfer from the electron gas to the lattice it leads to the formation of highly superheated metastable state, characterized by the appearance of the surface temperature maximum ΔTph,max=TphTm(ps)4.5103K in the solid phase, Fig. 17. In the liquid phase is also formed near-surface temperature maximum, Fig.20, but since the maximum velocity of evaporation front υυ,max40m/s is two orders of magnitude lower than the velocityυsl, magnitude and its role are insignificant.

Action of ultrashort laser pulses on metal target with different thermo-physical properties – copper – has several differences and many in common with the action of the aluminum target. The main difference of the non-equilibrium heating and non-equilibrium phase transformation in the condensed media is in the first place connected with the rate of energy exchange between the electron and phonon subsystems. The coefficient of energy exchange g(Te) in copper is by about 1.5 orders lower than that of aluminum, so that during laser energy release, the action of one of the main factors that limit the heating of the electron subsystem turns out to be significantly lower. As a result, the maximum values of the electron temperature of the surface are several times higher than that for aluminum: Te,max4.104Kfor picosecond pulse Te,max1.2105K for femtosecond pulse, fig. 21.

Heating and melting of the lattice take place in the picosecond range tm+0.1 ps, fig.22. For copper, the rate of energy transfer to the lattice is significantly lower, so that the velocity of the melting front turns out to be several times lower than that for aluminum υs,max(t)1.3km/s, fig.23. The maximum value of the pressure in the solid phase is also lower,ps,max=0.14Mbar. Despite the fact that the velocity υs for copper is significantly lower than that for Al, the value of overheating of the interphase surface turns out to be significantly higher due to decrease of the hydrodynamic effects:ΔTs,max350K, fig. 24. The overheating of the sub-surface region of the solid phase reaches the value of4103K.

The solidification of the liquid phase takes place in the nanosecond rage – the same as for aluminum. The lifetime of the melt reaches the value of 15÷20ns, fig.25, the maximum velocity of the crystallization front - υs35m/s. The maximum value of the overcooling of the front ~ 22 K. The overcooling of the volume of the solid reaches the value of 800 K.

Figure 21.

Time dependence of the surface temperature of Cu.

Figure 22.

The spatial profiles of temperature of copper.

Figure 23.

The time profile of the velocity of melting of copper.

Figure 24.

Response functions ΔTs(υsl) for Cu.

Figure 25.

The time profiles of the velocity of solidification of copper.

Modern experimental methods of the investigation of the processes of melting- crystallization of the metal systems determine the velocities of the solid-melt front to be within the range of 10÷100m/s with corresponding overcooling of liquid within the range of10÷400K. Thus, the data from the modeling of the metastable overcooled states in the liquid phase of metals is in relatively good agreement with the experimental measurements. It is not possible to perform the same comparison with the overcooled states in the solid phase because of the lack of reliable experimental data which is caused by the complexity of statement of such experiments.

Advertisement

6. Conclusion

The detailed analysis of the physical processes of the non-equilibrium heating and the fast phase transformations in the metal systems under the action of ultra-short laser pulses was performed in this chapter. A mathematical model was built based on the brief analysis of the thermodynamical and kinetic approaches for the description of the phase transformations of the first order. This model is a non-equilibrium hydrodynamic version of the Stefan problem and describes the kinetics of high-speed heterogeneous phase transformations: melting, crystallization, evaporation. The model is supplemented by the explicit description of the moving phase fronts for correct reflection of the dynamics of the phase transitions.

The methods of mathematical modeling were used to investigate the regimes of pico- and femtosecond action on metals - Al, Cu. The kinetics and dynamics of phase transformations are analyzed. The results of the modeling confirmed the skewness of the high-speed processes of melting – solidification. In particular, the maximum velocity of melting reaches the value of several kilometers per second, while the maximum velocity of solidification does not exceed hundred of meters per second. The temperature dependence υs(Ts) does not have a sharp bend in the point of change of the direction of the phase transition.

Special attention is paid to the overheated and overcooled metastable states that are caused by the fast phase transitions. The deviation from the local thermodynamical equilibrium is noticed both in the bulk and at the interphase boundary. The response function Ts(υs) has a linearly increasing character for low values of velocity υs in the range of[0÷±50]m/s. As the velocity υs rises more thanυs>100m/s, a maximum is formed in the response function for both metals. The appearance of the maximum means that the dominating mechanism of the phase transition changes: the mechanism of heat conductivity is changed by the hydrodynamic one. The maximum value of overheating of the interphase boundary reaches the value of 200 ÷ 400 K, the one for overcooling is by one order lower than that value. The overheating and overcooling of the bulk are significantly higher - 4103÷5103K for overheating and 6102÷8102K for overcooling. The results for different materials are compared between each other and with experimental data.

These results do not pretend to be an exhaustive description of the variety of fast phase transitions, because they used the assumption of the thermodynamic stability of overheated / overcooled metastable states during all times considered. Refusal of such assumptions will require further research to include the kinetics of homogeneous phase transformations into the consideration.

Acknowledgement

This work was supported by the RFBR (projects nos. 10-07-00246-a, 12-07- 00436-a, 11-01-12086-ofi-m).

References

  1. 1. FujimotoJ. G.LiuJ. M.IppenE. P.BloembergenN.1984Femtosecond Laser Interaction with Metallic Tungsten and Nonequilibrium Electron and Lattice Temperatures. Physical Review Letters. 53: 19/18371840
  2. 2. ChichkovB. N.MommaC.NolteS.vonAlvensleben. F.TunnermannA.1996Femtosecond, picosecond and nanosecond laser ablation of solids. Appl. Phys. A 63109115
  3. 3. ShirkM. D.MolianP. A.1998A review of ultrashort pulsed laser ablation of materials. Journal of Laser Applications. 10: 1/1828
  4. 4. Von derLinde. D.Sokolovski-TintenK.2000The physical mechanisms of short-pulse laser ablation. Appl. Surf. Sci. 154-155: 1-10.
  5. 5. 2007(200Part.UltrafastInteractions.InClaude. R.Phippseditor.OpticalSciences.Springer Series in Laser ablation and its applications. Santa Fe, New Mexico: Springer. 129121280
  6. 6. GonzalesP.BernathR.DuncanJ.OlmsteadT.RichardsonM.2004Femtosecond ablation scaling for different materials. Proc. SPIE. 5458265272
  7. 7. ChichkovB. N.KorteF.KochJ.NolteS.OstendorfA.2002Femtosecond laser ablation and nanostructuring. Proc. SPIE. 47601927
  8. 8. HamadS.KrishnaPodagatlapalli. G.SreedharS.TewariS. P.RaoS. V.2012Femtosecond and picosecond ablation of aluminum for synthesis of nanoparticles and nanostructures and their optical characterization. Proc. SPIE 8245: 82450L.
  9. 9. OstendorfA.KamlageG.KlugU.KorteF.ChichkovB. N.2005Femtosecond versus picosecond laser ablation. Proc. SPIE 571318
  10. 10. KamlageG.BauerT.OstendorfA.ChichkovB. N.2003Deep drilling of metals by femtosecond laser pulses. Applied Physics A: Materials Science & Processing. 77307310
  11. 11. LopezJ.KlingR.TorresR.LidolffA.DelaigueM.RicaudS.HönningerC.MottayE.2012Comparison of picosecond and femtosecond laser ablation for surface engraving of metals and semiconductors. Proc. SPIE 8243: 82430O.
  12. 12. BulgakovA. V.BulgakovaN. M.BurakovI. M.2009Nanosized Material Synthesis by Action of High- Power Energy Fluxes on Matter. Novosibirsk: Institute Thermophysics SB RAS. 461 p.
  13. 13. MazhukinV. I.LobokM. G.ChichkovB. N.2009Modeling of fast phase transitions dynamics in metal target irradiated by pico and femto second pulsed laser. Applied Surface Science. 25551125115
  14. 14. HerlachD. M.GalenkoP. K.Holland-MorizD.2005Metastable materials from undercooled metallic melts. Amsterdam: Elsevier. 485 p.
  15. 15. EliezerS.EliazN.GrossmanE.FisherD.GouzmanI.HenisZ.PeckeS.HorovitzY.FraenkelM.MamanS.LereahY.2004Synthesis of nanoparticles with femtosecond laser pulses. Phys. Rev. B 69: 144119/16
  16. 16. Mene´ndez-Manjo´nA.BarcikowskiS.ShafeevG. A.MazhukinV. I.ChichkovB. N.2010Influence of beam intensity profile on the aerodynamic particle size distributions generated by femtosecond laser ablation. Laser and Particle Beams. 284552
  17. 17. StratakisE.BarberoglouM.FotakisC.ViauG.GarciaC.ShafeevG. A.2009Generation of Al nanoparticles via ablation of bulk Al in liquids with short laser pulses. Opt. Express. 171265012659
  18. 18. MazhukinV. I.SamarskiiA. A.1994Mathematical Modeling in the Technology of Laser Treatments of Materials. Review. Surveys on Mathematics for Industry. 4: 2/85149
  19. 19. IvanovD. S.ZhigileiL. V.2003Combined atomistic-continuum modeling of short-pulse laser melting and disintegration of metal films. Phys. Rev. B. 68: 064114-1- 064114-22.
  20. 20. Landau.L. D.LifshicE. M.1976Teoreticheskaja fizika, t. V. Statisticheskaja fizika. Chast’ 1. M.: Nauka. 616 p.
  21. 21. ChristianJ. W.1965The theory of transformations in metals and alloys: an advanced textbook in physical metallurgy. Oxford: Pergamon Press. 975 p.
  22. 22. KirkpatrickR. J.1975Crystal Growth from the Melt. Review. American Mineralogist A: 60798614
  23. 23. AstaM.BeckermannC.KarmaA.KurzW.NapolitanoR.PlappM.PurdyG.RappazM.TrivediR.2009Solidification microstructures and solid-state parallels: Recent developments, future directions. Acta Materialia: 57: 146/941971
  24. 24. JacksonK. A.2004Kinetic Processes:Crystal Growth, Diffusion, and Phase Transitions in Materials. Weinheim: Wiley-VCH Verlag GmbH & Co. KGaA. 409 p.
  25. 25. HoytJ. J.AstaM.2002Atomistic computation of liquid diffusivity, solid-liquid interfacial free energy, and kinetic coefficient in Au and Ag. Phys. Rev. B. 65: 21/214106-1- 214106-11.
  26. 26. RodwayG. H.HuntJ. D.1991Thermoelectric investigation of solidification of lead I. Pure lead. Journal of Crystal Growth. 11223
  27. 27. HoytJ. J.AstaM.KarmaA.2002Atomistic Simulation Methods for Computing the Kinetic Coefficient in Solid-Liquid Systems. Interface Science. 1023
  28. 28. ButaD.AstaM.HoytJ. J.2007Kinetic coefficient of steps at the Si(111) crystal-melt interface from molecular dynamics simulations. Journal of Chemical Physics. 127: 7/074703074713
  29. 29. AminiM.LairdB. B.2006Kinetic Coefficient for Hard-Sphere Crystal Growth from the Melt. Physical Review Letters. 97: 216102-1- 216102-4.
  30. 30. MonkJ.YangY.MendelevM. I.AstaM.HoytJ. J.SunD. Y.2010Determination of the crystal-melt interface kinetic coefficient from molecular dynamics simulations. Modelling and Simulation in Materials Science and Engineering. 18: 015004-1- 015004-18.
  31. 31. Wilson H.A.1900On the velocity of solidification and viscosity of supercooled liquids. Philos. Mag. 50238250
  32. 32. FrenkelJa. I.1932Note on the relation between the speed of crystallization and viscosity. Phys. Z. Sowjet Union. 1498499
  33. 33. Frenkel’Ja. I.1975Kineticheskaja teorija zhidkostei. L.: Nauka. 592 p.
  34. 34. JacksonK. A.ChalmersB.1956Kinetics of solidification. Can. J. Phys. 34473490
  35. 35. ChalmersB.1964Principles of Solidification., New York: John Wiley &Sons. 129 p.
  36. 36. KludeM. D.RayJ. R.1989Velocity versus Temperature Relation for Solidification and Melting of Silicon: A Molecular-Dynamics Study. Phys. Rev. B. 39: 3/17381746
  37. 37. Tymczak C.J. Ray J.R.1990Interface response function for a model of sodium: A molecular dynamics study, J. Chem. Phys. 92: 12/75207530
  38. 38. Tymczak C.J. Ray J.R.1990Asymmetric crystallization and melting kinetics in sodium: A molecular-dynamics study. Phys. Rev. Lett. 641112781281
  39. 39. MacDonald. C. A.MalvezziA. M.SpaepenF.1989Picosecond time-resolved measurements of crystallization in noble metals. J. Appl. Phys. 65: 1/129137
  40. 40. HertzH.1882Uber die Verdunstung der Flussigkeiten, insbesondere des Quecksilbers, im luftleeren Raume. Ann. Phys. und Chemie. 17177190
  41. 41. KnudsenM.1915Die Maximale Verdampfungs-geschwindigkeit Quecksilbers. Ann. Phys. und Chemie. 47697705
  42. 42. SiewertC. E.ThomasJ. R.1981Strong evaporation into a half space. J. of Applied Mathem. and Physics( ZAMP). 32421433
  43. 43. Siewert C.E.2003Heat transfer and evaporation/condensation problems based on the linearized Boltzmann equation. European Journal of Mechanics B/Fluids. 22391408
  44. 44. FrezzottiA.2007A numerical investigation of the steady evaporation of a polyatomic gas. European Journal of Mechanics B/Fluids. 2693104
  45. 45. TitarevV. A.ShakhovE. M.2001Teplootdacha i isparenie s ploskoi poverkhnosti v poluprostranstvo pri vnezapnom povyshenii temperatury tela. Mekhanika zhidkosti i gaza. 1/141153
  46. 46. AbramovA. A.ButkovskyA. V.2009Effect of the Mode of Gas-Surface Interaction on Intensive Monatomic Gas Evaporation. Fluid Dynamics. 44: 1/8087
  47. 47. BondM.StruchtrupH.2004Mean evaporation and condensation coefficients based on energy dependent condensation probability. Phys. Rev. E. 70: 061605-1- 20.
  48. 48. FordI. J.Lee-LT.2001Entropy production and destruction in models of material evaporation. J. Phys. D: Appl. Phys. 34413417
  49. 49. RoseJ. W.2000Accurate approximate equations for intensive sub-sonic evaporation. Int. Journal of Heat and Mass Transfer.. 4338693875
  50. 50. BedeauxD.KjelstrupS.RubiJ. M.2003Nonequilibrium translational effects in evaporation and condensation. J. Chemical Physics. 119: 17/91639170
  51. 51. FrezzottiA.GibelliL.LorenzaniS.2005Mean field kinetic theory description of evaporation of a fluid into vacuum. Phys. Fluids. 17: 012102-1- 012102-12.
  52. 52. RahimiP.WardCh. A.2005Kinetics of Evaporation: Statistical Rate Theory Approach Int.J. of Thermodynamics. 8: 1/114
  53. 53. MelandR.FrezzottiA.YtrehusT.HafskjoldB.2004Nonequilibrium molecular-dynamics simulation of net evaporation and net condensation, and evaluation of the gas-kinetic boundary condition at the interphase. Phys. Fluids. 16: 2/223243
  54. 54. IshiyamaT.YanoT.FujikawaS.2005Kinetic boundary condition at a vapor-liquid interface. Phys. Rev. Lett. 95: 0847504.
  55. 55. FrezzottiA.2011Boundary conditions at the vapor-liquid interface. Phys. Fluids. 23: 030609-(1-9).
  56. 56. KartashovI. N.SamokhinA. A.SmurovI.Yu2005Boundary conditions and evaporation front instabilities. J. Phys. D: Appl. Phys. 3837033714
  57. 57. CroutD.1936An application of kinetic theory to the problem of evaporation and sublimation of monatomic gases. J. Math. Physics. 15154
  58. 58. LabuntsovD. A.KryukovA. P.1979Analysis of Intensive Evaporation and Condensation. Inter. J. Heat and Mass Transfer. 22: 7/9891002
  59. 59. MazhukinV. I.SamokhinA. A.1987Matematicheskoe modelirovanie fazovykh perekhodov i obrazovanija plazmy pri deistvii lazernogo izluchenija na pogloshhajushhie sredy. In: Samarskii A.A., Kurdjumov S.P., Mazhukin V.I., editors. Nelineinye differencial’nye uravnenija matematicheskoi fiziki. Moskva: Nauka. 191244
  60. 60. Samokhin A.A.1990Effect of laser radiation on absorbing condensed matter. In: Prokhorov A.M. Series editor. Fedorov V.B. editor. Proceedings of the Institute of General Physics Academy of Sciences of the USSR. New-York: Nova Science Publishers. 13: 203 p.
  61. 61. Anisimov S.I.1968Ob isparenii metalla, pogloshhajushhego lazernoe izluchenie. ZhEhTF. 54: 1/339342
  62. 62. Khight C.J.1979Theoretical Modeling of Rapid Surfact Vaporation with Back Pressure. AIAA J. 17: 5/8186
  63. 63. MazhukinV. I.PrudkovskiiP. A.SamokhinA. A.1993O gazodinamicheskikh granichnykh uslovijakh na fronte isparenija. Matematicheskoe modelirovanie. 5: 6/39
  64. 64. MazhukinV. I.SamokhinA. A.2012Boundary conditions for gas-dynamical modeling of evaporation processes. Mathematica Montisnigri. XXIV: 817
  65. 65. LifshicI. M.KaganovM. I.TanatarovL. V.1956Relaksacija mezhdu ehlektronami i kristallicheskojj reshetkojj. ZhEhTF. 31: 2(8)/232237
  66. 66. LifshicI. M.KaganovM. I.TanatarovL. V.1959K teorii radiacionnykh izmenenijj v metallakh. Atomnaja ehnergija. 6/391402
  67. 67. MartynenkoJu. V.JavlinskiiJu. N.1983Okhlazhdenie ehlektronnogo gaza metalla pri vysokojj temperature. DAN SSSR. 270: 1/8891
  68. 68. BeerA. C.ChaseM. N.ChoquardP. F.1955Extension of McDougall-Stoner tables of the Fermi-Dirac functions. Helvetica Physica Acta. XXVII: 5/529542
  69. 69. KittelC.1978Introduction to Solid State Physics, 6h Edition. New York: J. Wiley & Sons. 789 p.
  70. 70. LandauL. D.LifshicE. M.1963Teoreticheskaja fizika. Kvantovaja mekhanika. Nereljativistskaja teorija. T.III. Moskva: Fizmatgiz. 702 p.
  71. 71. Frenkel’Ja. I.1958Vvedenie v teoriju metallov. Moskva: Fizmatlit. 368 p.
  72. 72. TouloukianY. S.PowellR. W.HoC. Y.KlemensP. G.1970Thermophysical Properties of Matter 1Thermal Conductivity: Metallic Elements and Alloys, New York, Washington: IFI/Plenum Data Corp.
  73. 73. Ziman J.M1974Principles of the theory of solid. Gambridge: University Press. 472 p.
  74. 74. CheynetB.DuboisJ. D.MilesiM.1996Données thermodynamiques des éléments chimiques. In : Techniques de L’Ingenieur, Materiaux metalliques. M153Forum M64. 122
  75. 75. Aluminum: Properties and physical metallurgy1984In: John E. Hatch editor. American society for metals. Ohio: Metals Park. 421 p.
  76. 76. Fizicheskie velichiny. Spravochnik1991Grigor’eva I.S., Mejjlikhova E.Z. editors. Moskva: Ehnergoatomizdat. 1232 p.
  77. 77. IidaT.GuthrieR. I. L.1988The Physical Properties of Liquid Metals. Oxford: Clarendon Press. 288 p.
  78. 78. PinesD.1965Ehlementarnye vozbuzhdenija v tverdykh telakh. Moskva: Mir. 383 p.
  79. 79. Qiu T.Q., Tien C.L1994Femtosecond Laser Heating of Multi-Layer Metals. I. Analysis. Int. J. Heat Mass Transfer. 37: 17/27892797
  80. 80. Ginzburg V.L., Shabanskii V.L1955Kineticheskaja temperatura ehlektronov v metallakh i anomal’naja ehlektronnaja ehmissija. Dokl. AN SSSR. 100: 3/445448
  81. 81. Martynenko Ju.V., Javlinskii Ju.N1987Vozbuzhdenie ehlektronov metalla oskolkom delenija. Atomnaja ehnergija. 62: 2/8083
  82. 82. LinZ.ZhigileiL. V.CelliV.2008Electron-phonon coupling and electron heat capacity of metals under conditions of strong electron-phonon nonequilibrium. Phys. Rev. B. 77: 7/075133-1- 075133-17.
  83. 83. MazhukinV. I.MazhukinA. V.DeminM. M.ShapranovA. V.2011Ehffekty neravnovesnosti pri vozdejjstvii impul’snogo lazernogo izluchenija na metally. Opticheskiy zhurnal. 78: 8/2937
  84. 84. MazhukinV. I.MazhukinA. V.DeminM. M.ShapranovA. V.2011Mathematical modeling of short and ultrashort laser action on metals. In: Sudarshan T.S., Beyer Eckhard, Berger Lutz-Michael editors. Surface Modification Technologies (STM 24). Dresden: Fraunhofer Institute for Material and Beam Technology IWS. 24201208
  85. 85. Samarskii A.A., Moiseenko B.D1965Ehkonomichnaja skhema skvoznogo scheta dlja mnogomernojj zadachi Stefana. ZhVM i MF. 5: 5/816827
  86. 86. Meyer G.H1973Multidimensional Stefan Problems. SIAM J. Numer. Anal. 10: 3/522538
  87. 87. White B.R.E1983A modified finite difference scheme for the Stefan problem. Mathem. of Comput. 41: 164/337347
  88. 88. Klimontovich Ju.L1982Statisticheskaja fizika. Moskva: Nauka. 608 p.
  89. 89. MazhukinV. I.SmurovI.FlamantG.1995Overheated Metastable States in Pulsed Laser Action on Ceramics. J. Applied Physics. 78: 2/12591270
  90. 90. MazhukinV. I.SmurovI.DupuyC.JeandelD.1994Simulation of Laser Induced Melting and Evaporation Processes in Superconducting. J. Numerical Heat Transfer Part A. 26587600
  91. 91. MazhukinV. I.LobokM. G.ChichkovB. N.2009Modeling of fast phase transitions dynamics in metal target irradiated by pico and femto second pulsed laser. Applied Surface Science. 255/51125115
  92. 92. MazhukinV. I.SamarskiiA. A.ChujjkoM. M.1999Metod dinamicheskojj adaptacii dlja chislennogo reshenija nestacionarnykh mnogomernykh zadach Stefana. Doklady RAN. 368: 3/307310
  93. 93. Mazhukin A.V., Mazhukin V.I2007Dinamicheskaja adaptacija v parabolicheskikh uravnenijakh. Zhurnal vychislitel’nojj matematiki i matematicheskojj fiziki. 47: 11/19111934Mazhukin A.V., Mazhukin V.I (2007) Dynamic Adaptation for Parabolic Equations. Computational Mathematics and Mathematical Physics. 47: 11/1833- 1855.
  94. 94. Breslavsky P.V., Mazhukin V.I2008Metod dinamicheskojj adaptacii v zadachakh gazovojj dinamiki s nelinejjnojj teploprovodnost’ju. Zhurnal vychislitel’nojj matematiki i matematicheskojj fiziki. 48: 11/20672080Breslavskii P. V. and Mazhukin V. I (2008) Dynamic Adaptation Method in Gasdynamic Simulations with Nonlinear Heat Conduction. Computational Mathematics and Mathematical Physics. 48: 11/2102-2115.
  95. 95. Koroleva O.N., Mazhukin V.I2006Matematicheskoe modelirovanie lazernogo plavlenija i isparenija mnogoslojjnykh materialov. Zhurnal vychislitel’nojj matematiki i matematicheskojj fiziki. 46: 5/910925Koroleva O.N., Mazhukin V.I (2006) Mathematical Simulation of Laser Induced Melting and Evaporation of Multilayer Materials. Computational Mathematics and Mathematical Physics. 46: 5/848- 862.
  96. 96. Mazhukin V.I., Samarskii A.A., Kastel’janos Oplando, Shapranov A.V1993Metod dinamicheskojj adaptacii dlja nestacionarnykh zadach’s bol’shimi gradientami. Matematicheskoe modelipovanie. 5: 4/3256
  97. 97. Samarskii A.A1989Teorija raznostnykh skhem. Moskva: Nauka. 616 p.
  98. 98. Samarskii A.A., Popov Ju.P1992Raznostnye metody reshenija zadach gazovojj dinamiki. Moskva: Nauka. 421pp.
  99. 99. MazhukinV. I.MazhukinA. V.KorolevaO. N.2009Optical properties of electron Fermi-gas of metals at arbitrary temperature and frequency. Laser Physics. 19: 5/11791186
  100. 100. Mazhukin A.V., Koroleva O.N2010Raschjot opticheskikh kharakteristik aljuminija. Matematicheskoe modelirovanie. 22: 5/1528

Written By

V.I. Mazhukin

Submitted: 14 December 2011 Published: 17 October 2012