1. Trang chủ
  2. » Giáo án - Bài giảng

molecular dynamics simulation of heat transfer and phase change during

10 507 0
Tài liệu đã được kiểm tra trùng lặp

Đang tải... (xem toàn văn)

THÔNG TIN TÀI LIỆU

Thông tin cơ bản

Tiêu đề Molecular dynamics simulation of heat transfer and phase change during laser material interaction
Tác giả Xinwei Wang, Xianfan Xu
Trường học Purdue University
Chuyên ngành Mechanical Engineering
Thể loại Bài luận
Năm xuất bản 2001
Thành phố West Lafayette
Định dạng
Số trang 10
Dung lượng 1,62 MB

Các công cụ chuyển đổi và chỉnh sửa cho tài liệu này

Nội dung

Proceedings of NHTC'01 35th National Heat Transfer Conference Anaheim, California, June 10-12, 2001 NHTC2001-20070 MOLECULAR DYNAMICS SIMULATION OF HEAT TRANSFER AND PHASE CHANGE DURIN

Trang 1

Proceedings of NHTC'01 35th National Heat Transfer Conference Anaheim, California, June 10-12, 2001

NHTC2001-20070

MOLECULAR DYNAMICS SIMULATION OF HEAT TRANSFER AND PHASE CHANGE DURING

LASER MATERIAL INTERACTION Xinwei Wang, Xianfan Xu *

School of Mechanical Engineering

Purdue University West Lafayette, IN 47907

* To whom correspondence should be addressed

ABSTRACT

In this work, heat transfer and phase change of an argon

crystal illuminated with a picosecond pulsed laser are

investigated using molecular dynamics simulations The result

reveals no clear interface when phase change occurs, but a

transition region where the crystal structure and the liquid

structure co-exist between the solid and the liquid Superheating

is observed during the melting process The solid-liquid and

liquid-vapor interfaces are found to move with a velocity of

hundreds of meters per second In addition, the vapor is found

to be ejected from the surface with a velocity close to a

thousand meters per second

Keywords: heat transfer, phase change, MD simulation,

laser-material interaction, ablation threshold

NOMENCLATURE

I laser intensity

k thermal conductivity

B

k Boltzmann's constant

m atomic mass

M

P probability for atoms moving with a velocity v

q′′ heat flux applied to the surface of the target for thermal

conductivity calculation

r atomic position

c

r cut off distance

s

r the nearest neighbor distance

t time

T

t preset time constant in velocity scaling

t

T temperature T

δ initial temperature increase for calculating the specific

heat

T

∆ final temperature increase for calculating the specific

heat

v velocity

x coordinate in x direction

y coordinate in y direction

z coordinate in z direction

Greek Symbols

χ velocity scaling factor

ε LJ well depth parameter

φ potential

σ equilibrium separation parameter

ξ current kinetic temperature in velocity scaling

Subscripts

i atomic index

Superscripts

* no-dimensionalized

I INTRODUCTION

In recent years, ultrashort pulsed lasers have been rapidly developed and used in materials processing Due to the extremely short pulse duration, many difficulties exist in experimental investigation of laser material interaction such as measuring the transient surface temperature, the velocity of the solid-liquid interface, and the material ablation rate Ultrashort laser material interaction involves several coupled, non-linear, and non-equilibrium processes inducing an extremely high

Trang 2

heating rate (10 K/s) and a high temperature gradient (10

K/m) The continuum approach of solving the heat transfer

problem becomes questionable under these extreme situations

On the other hand, the molecular dynamics (MD) simulation,

which solves the movement of atoms or molecules directly, is

suitable for investigating the ultrashort laser material interaction

process One aim of this work is to use the MD simulation to

investigate heat transfer occurring in ultrashort laser-material

interaction and to compare the results with those obtained with

the continuum approach

A large amount of work has been dedicated to studying

laser material interaction using MD simulations Due to the

limitation of computer resources, most work was restricted to

systems with a small number of atoms, thus only qualitative

results such as the structural change due to heating were

obtained For instance, using quantum MD simulations,

Shibahara and Kotake studied the interaction between metallic

atoms and the laser beam in a system consisting of 13 atoms or

less [1, 2] Their work was focused on the structural change of

metallic atoms due to laser beam absorption Häkkinen and

Landman [3] studied dynamics of superheating, melting, and

annealing at the Cu surface induced by laser beam irradiation

using the two-step heat transfer model developed by Anisimov

[4] This model describes the laser metal interaction in two

steps including photon energy absorption in electrons and

lattice heating through interaction with electrons A large body

of the MD simulation of laser material interaction was to study

the laser induced ablation in various systems Kotake and

Kuroki [5] studied laser ablation of a small dielectric system

consisting of 4851 atoms Laser beam absorption was simulated

by exciting the potential energy of atoms Applying the same

laser beam absorption approach, Herrmann and Campbell [6]

investigated laser ablation of a silicon crystal containing

approximately 23000 atoms Zhigileit et al [7, 8] studied laser

induced ablation of organic solid using the breathing sphere

model, which simulated laser irradiation by vibrational

excitation of molecules However, because of the arbitrary

properties chosen in the calculation, their calculation results

were qualitative, and were restricted to small systems with tens

of thousands of atoms Ohmura et al [9] attempted to study

laser metal interaction with the MD simulation using the Morse

potential function for metals [10] The Morse potential function

simplified the potential calculation among the lattice and

enabled them to study a larger system with 160,000 atoms Heat

conduction by the electron gas, which dominated heat transfer

in metal, could not be predicted by the Morse potential

function Alternatively, heat conduction was simulated using the

finite difference method based on the thermal conductivity of

metal Laser material interaction in a large system was recently

investigated by Etcheverry and Mesaros [11] In their work, a

crystal argon solid containing about half a million atoms was

simulated For laser induced acoustic waves, a good agreement

between the MD simulation and the standard thermoelastic

calculation was observed

In this work, MD simulations are conducted to study laser argon interaction The system under study has 486,000 atoms, which is large enough to suppress statistical uncertainty Laser heating of argon with different laser fluences is investigated Laser induced heat transfer, melting, evaporation, material ablation are emphasized in this work Phase change relevant parameters, such as the velocity of solid-liquid and liquid-vapor interfaces, ablation rate, and ablation threshold fluence are reported In section II, theories for the MD simulation used in this work are introduced Calculation results are summarized in section III

II THEORY OF MD SIMULATION

Molecular dynamics simulation is a computational method

to investigate the behavior of materials by simulating the atomic motion controlled by a given potential Argon is overwhelmingly explored in MD simulation due to the meaningful physical constants of the widely-accepted Lennard-Jones 12-6 (LJ) potential and the less computation time required than more complicated potentials involving multi-body interaction or electric static force In this calculation, an argon crystal at 50 K is assumed to be illuminated with a spatially uniform laser beam The melting and the boiling temperatures

of argon at one atm are 83.8 K and 87.3 K, respectively, while its critical temperature is 150.87 K The basic problem involves solving Newtonian equations for each atom interacting with its neighbors by means of a pairwise Lennard-Jones force:

=

≠i

j ij

i

dt

r d

2

(1)

where m i and r i are the mass and position of atom i, respectively, F is the interaction force between atoms i and j, ij

which is obtained from the Lennard-Jones potential as

ij ij

F =−∂φ /∂ The Lennard-Jones potential φij is written as

=

6 12

4

ij ij

ij

r r

σ σ

ε

where ε is the LJ well depth parameter, σ is the equilibrium separation parameter, and r ij =r jr i Therefore, the force F ij

can be expressed as

ij ij ij

r r

+

=4ε 12σ1412 6σ86 (2b)

A standard method for solving ordinary differential equations (1) and (2) is the finite difference approach The general idea is to obtain the atomic positions, velocities, etc at time t+ based on the positions, velocities, and other δt

Trang 3

dynamic information at time t The equations are solved on a

step-by-step basis, and the time interval tδ is dependent

somehow on the method applied However, tδ is usually much

smaller than the typical time taken for an atom to travel its own

length Many different algorithms have been developed to solve

Eqs (1) and (2), of which the Verlet algorithm is widely used

due to its numerical stability, convenience, and simplicity [12]

In this calculation, the velocity Verlet algorithm is used, which

is expressed as:

t t t v t r

t

t

r i( +δ)= i()+ ( +δ /2)δ (3a)

ij

ij

t t t

t

F

δ

∂φ

δ ) ( )

=

t m

t t F t t v t

t

v

i

ij

δ

δ δ

δ /2) ( /2) ( )

3

+ +

=

In the calculation, most time is spent on calculating forces

using Eq (3b) When two atoms are far away enough from each

other, the force between them is negligible The distance

between atoms beyond which the interaction force is neglected

is called cutoff distance (potential cutoff), r In this work, c r is c

taken as 2.5σ, which is a cutoff potential widely used in MD

simulations using the LJ potential At this distance, the potential

is only about 1.6% of the well depth In the calculation, the

distance between atoms is first compared with r , and only c

when the distance is less than r , the force is calculated The c

comparison of the atomic distance with r is organized by c

means of the cell structure and the linked list methods [12] In

these methods, the computation domain is divided into many

structural cells with a characteristic size of r To speed up the c

calculation, direct evaluation of the force using Eqs (2) is

avoided by looking up a pre-prepared table for the force in the

range of 2

ij

r from 0.25σ2 to 2

c

r , with an interval of 10-6σ2 Laser energy absorption in the material is simulated by

scaling the velocities of all atoms in each structural cell by an

appropriate factor The amount of energy deposited in each cell

is calculated assuming the laser beam is exponentially absorbed

in the target In order to prevent undesired amplification of

atomic macromotion, the average velocity of atoms in each

layer of structural cells is subtracted before velocity scaling

Non-dimensionalized parameters are used, which are listed

in Table 1 With non-dimensionalization, Eqs (1) and (2)

become

=

t

d

r

2

*

*

2

)

6

* 12

*

*

) (

1 ) (

1

ij ij

ij

r

=

* 8

* 14

*

*

) (

6 ) (

12

ij ij ij

r r

+

The form of Eqs (3a) and (3b) is preserved, while Eq (3c) becomes

*

*

*

*

*

*

*

*

*

*(t 3t /2) v (t t /2) F (t t ) t

v + δ = +δ + ij +δ δ (5) Table 1 Nondimensionalized parameters

Quantity Equation Time t*=t/( m/4εσ) Length r*=r

Mass * / 1

=

=m m m

Velocity v*=v/ 4ε/m Potential φ* φ/4ε

= Force F ij*=F ij/(4ε/σ) Temperature T*=k B T/4ε

Parameters used in the calculation are listed in Table 2 A face-centered cubic (fcc) structure is used to initialize atomic positions The initial atomic velocities are specified randomly from a Gaussian distribution based on the temperature

III CALCULATION RESULTS

The target studied consists of 90 fcc unit cells in x and y directions, and 15 fcc unit cells in the z direction Each unit cell contains 4 atoms, and the system consists of 486,000 atoms In both x and y directions, the computational domain has a size of 48.73 nm In the z direction, the size of the computation domain

is 17.14 nm with the bottom of the target located at 4.51 nm and the top surface (laser irradiated surface ) at 12.63 nm

Trang 4

Table 2 Values of the parameters used in the calculation

Parameter Value

ε , LJ well depth parameter 1.653×10− 21 J

σ , LJ equilibrium separation 0.3406 nm

m, Argon atomic mass 66.3×10− 27 kg

B

k , Boltzmann’s constant 1.38×10− 23 J/K

a, Lattice constant 0.5414 nm

c

r , Cut off distance 0.8515 nm

Size of the sample –x 48.726 nm

Size of the sample –y 48.726 nm

Size of the sample –z 8.121 nm

Time step 25 fs Number of atoms 486000

III.1 Thermal Equilibrium Calculation

The first step in the calculation is to initialize the system so

that it is in thermal equilibrium before laser heating, which is

done by a thermal equilibrium calculation In this calculation,

the target is initially constructed based on the fcc lattice

structure with the (100) surface facing up The nearest neighbor

distance, r s, in the fcc lattice for argon depends on temperature

T, and is calculated using the expression given by Broughton et

al [13],

2

014743 0 054792

0 0964

1

)

 +

 +

=

ε ε

σ

T k T

k T

5 4

3

25057 0 23653

0 083484

 +

+

ε ε

ε

T k T

k T

(7) Initial velocities of atoms are specified randomly from a

Gaussian distribution based on the specified temperature of 50

K using the following formula,

T k v

3

2

1 3

1

2=

=

(8)

where k is the Boltzmann's constant During the equilibrium B

calculation, due to the variation of the atomic positions, the

temperature of the target may change because of the energy

transform between the kinetic and potential energies In order to

allow the target to reach thermal equilibrium at the expected

temperature, velocity scaling is necessary to adjust the

temperature of the target during the early period of

equilibration The velocity scaling approach proposed by

Berendsen et al [14] is applied in this work At each time step, velocities are scaled by a factor

2 / 1

1 







 +

=

ξ

δ

t

t T

(9)

where ξ is the current kinetic temperature, and t is a preset T

time constant, which is taken as 0.4 ps in the simulation This technique forces the system towards the desired temperature at

a rate determined by t , while only slightly perturbing the T

forces on each atom After scaling the velocity for 50 ps, the calculation is continued for another 100 ps to reach thermal equilibrium The final equilibrium temperature of the target is 49.87 K, which is close to the desired temperature of 50 K When the target reaches the thermal equilibrium status, the atomic velocity distribution should follow the Maxwellian distribution

T k mv

B

T k

m v

2 / 3 2

2

2



=

π

π (10)

where P is the probability for an atom moving with a velocity, M

v The velocity distribution based on the simulation results as well as the Maxwell's distribution, are shown in Fig 1, which indicates a good agreement between the two

M D Simulation

M axwell's Distribution

Velocity (m/s)

Figure 1 Comparison of the velocity distribution by the MD simulation with the Maxwellian velocity distribution Figure 2 shows the lattice structure in the x-z plane when the system is in thermal equilibrium For the purpose of illustration, only the atoms in the range of 0< x<12nm and

6 12

0< y< nm are plotted It is seen that atoms are located around their equilibrium positions, and the lattice structure is preserved It is also observed from Fig 2 that at the top and the

Trang 5

bottom surfaces of the target, a few atoms have escaped due to

the free boundary conditions

4

7

10

13

x (nm )

Figure 2 Structure of the target in the x-z plane within the range

of 120< x< nm and 0< y<12.6 nm

III.2 Calculation of Thermophysical Properties

In order to check the validity of the simulation, thermal

physical properties including the specific heat at constant

pressure, the specific heat at constant volume, and the thermal

conductivity are calculated and compared with published data

To calculate the specific heat at constant pressure, the

system is first equilibrated with periodical boundary conditions

in x and y directions, and free boundary conditions in the z

direction, which simulates a target in vacuum A kinetic energy

of 3/2⋅k BδT with δT =8 is added to each atom and the

system is calculated for about 100 ps to reach a new thermal

equilibrium status with a final temperature increase of T∆ The

specific heat is calculated as

) /(

2

/

3 k T m T

c p = ⋅ Bδ ⋅∆ (11)

The specific heat at constant pressure (vacuum) is

calculated to be 787.8 J/kg·K at 51.476 K This value is about

24% higher than the literature data, which is 637.5 J/kg·K [15]

This difference is mainly due to the free boundary conditions of

vacuum used in the MD simulation, while the experimental

results are for samples under atmospheric pressure Under free

boundary conditions, atoms are easier to expand in space when

heated Therefore, more heat is stored in the form of potential

energy and resulting in a larger specific heat

The specific heat at constant volume is calculated in the

similar way as described above except that free boundary

conditions in the z direction are replaced with periodical

boundary conditions in order to keep the volume constant The

specific heat at constant volume is calculated to be 576.0

J/kg·K, which is only 6% higher than the literature value of

543.5 J/kg·K [15] This small difference might be due to the

potential function used in the calculation, which is more suitable for argon in liquid state

The thermal conductivity of argon is calculated as follows The target is first equilibrated with periodical boundary conditions in x and y directions, and free boundary conditions

in the z direction A constant heat flux q′′ is applied to the surface of the target by scaling velocities of atoms in cells on the top surface, and the same amount of heat flux is dissipated from the bottom of the target by scaling velocities of atoms in

cells at the bottom The heat flux q′′ is taken as

8

10 83216

2 × W/m2

, which induces a temperature difference of about 5 K across the target Figure 3 shows the temperature distribution in the target when a heat flux is passing through It

is seen that a linear temperature distribution is established in the

target due to the heat flux The thermal conductivity k is

calculated as

x T

q k

∂ /

′′

The thermal conductivity of the target is calculated to be 0.304 W/m·K, which is about 34% smaller than the experimental value of 0.468 W/m·K This large difference could

be due to the free boundary conditions used in the calculation and possible errors in the potential function Further work is necessary to study the effects of boundary conditions and different potential functions

46 47 48 49 50 51 52 53 54

z (nm )

Figure 3 Temperature distribution in the target subjected to a

constant heat flux

III.3 Laser Material Interaction

In laser material interaction, periodical boundary conditions are assumed on surfaces in x and y directions, and free boundary conditions on surfaces in the z direction The simulation corresponds to the problem of irradiating a block in vacuum The laser beam is uniform in space, and has a temporal Gaussian distribution with a 5 ps FWHM centered at 10 ps The

Trang 6

laser beam energy is absorbed exponentially in the target and

expressed as

τ / )

(z

I

dz

dI

where I is the laser beam intensity, and τ is the characteristic

absorption depth, which is taken as 2.5 nm

Laser Heating

The temperature distribution in the target illuminated with a

laser pulse of 0.03 J/m2

is first calculated and compared with finite difference results With this laser fluence, only a

temperature increase is induced, and no phase change occurs

Figure 4 shows the temperature distribution calculated using the

MD simulation and the finite difference method In MD

simulations, temperature at different locations is calculated as

an ensemble average of a domain with thickness of 2.5σ in the z

direction In the calculation using the finite difference method,

properties of the target obtained with the MD simulation are

used It is observed from Fig 4 that the results obtained from

the MD simulation show proper trends comparing with those by

the finite difference method The difference between them is on

the same order of the statistic uncertainty of the MD simulation

In other words, the continuum approach is still capable of

predicting the heating process induced by a picosecond laser

pulse

Laser Induced Phase Change

In this section, various phenomena accompanying phase

change in an argon target illuminated with a laser pulse of 0.7

J/m2

are investigated The threshold fluence for ablation is also

studied

For argon illuminated with a pulsed laser of 0.7 J/m2

, a series of snapshots of atomic positions at different times is

shown in Fig 5 It is seen that until 10 ps, the lattice structure is

still preserved in the target At about 10 ps, melting starts, and

the lattice structure is destroyed in the melted region and is

replaced by a random atomic distribution After 20 ps, the solid

liquid interface stops moving into the target, and vaporized

atoms are clearly seen Figure 6 shows the distribution of

number density of atoms in space at different times, which

demonstrates the variation of solid structure during laser

heating At the early stage of laser heating, the crystal structure

is preserved in the target, which is seen as the peak number

density of atoms on each lattice layer Due to the increase of the

atomic kinetic energy in laser heating, atoms vibrate more in the

crystal region, causing a lower peak of the number density of

atoms and a wide distribution As laser heating progresses, the

target is melted from its front surface, and the atomic

distribution becomes random Therefore, the number density of

atoms becomes uniform over the melted region However, no

clear interface is observed between the solid and the liquid

Instead, the structure of solid and liquid co-exists within a

certain range between the solid and the liquid, which is shown

as the co-existence of the peak and the high base of the number density of atoms Evaporation happens at the surface of the target, which reduces the number density of atoms significantly

at the location near the liquid surface

48 50 52 54

56

MD simulation Finite Difference t=5 ps

50 52 54

56

MD simulation Finite Difference t=10 ps

50 52 54

56

MD simulation Finite Difference t=15 ps

50 52 54

56

MD simulation Finite Difference t=20 ps

50 52 54

56

MD simulation Finite Difference t=25 ps

50 52 54

56

MD simulation Finite Difference

t=30 ps

z (nm )

Figure 4 Temperature distribution in the target illuminated with

a laser pulse of 0.03 J/m2

In order to find out the rate of melting and evaporation, criteria are needed to determine the solid-liquid and liquid-vapor interfaces For solid argon, the average number density of atoms is 2.52×1028 m-3 with a distribution in space as shown in Fig 6 Owing to the lattice structure, the number density of atoms is higher than the average value around the lattice layer location In this work, if the number density of atoms is higher

Trang 7

than 2.52×1028 m-3, the material is treated as solid At the front

of the melted region, it is seen from Fig 6 that when the number

density of atoms is below 8.42× m27 -3, a relatively sharp

decrease of the number density of atoms happens Therefore,

when the number density of atoms is less than 8.42×1027 m-3,

which is about one third of the number density in solid, the

material is assumed to be vapor Although this criterion for

liquid-vapor interface is not quite rigorous due to the large

transition range from liquid to vapor, further study of the

liquid-vapor interface using radial distribution function shows that the

criterion used here gives a good approximation of the

liquid-vapor interface

0

4

8

12

0

4

8

12

0

4

8

12

0

4

8

12

0

4

8

12

0

4

8

12

z (nm)

Figure 5 Snapshots of atomic positions in argon illuminated

with a laser pulse with a fluence of 0.7 J/m2

Applying these criteria, transient locations of the

solid-liquid and solid-liquid-vapor interfaces, as well as the velocity of

interfaces can be obtained and are shown in Fig 7 It is

observed that melting and evaporation start at 10 ps, while laser

heating starts at around 5 ps It is seen that the solid-liquid

interface moves into the solid owing to the melting of the solid,

and the liquid-vapor interface moves outward as the melted

region expands because liquid is less dense than solid At about

20 ps, both solid-liquid and liquid-vapor interfaces stop moving The velocities of the interfaces are shown in Fig 7b It

is seen that the duration of the interface movement is about 10

ps, which is about the same as the laser pulse width The highest velocity of the liquid-vapor interface is about 200 m/s, close to the equilibrium velocity (233.5 m/s) of the argon atom at the boiling temperature The highest velocity of the solid-liquid interface is about 400 m/s, lower than the sound velocity (1501 m/s) in argon

0.00 0.25 0.50 0.75 1.00 1.25

t=5 ps

0.00 0.25 0.50 0.75 1.00 1.25

t=10 ps

3 )

0.00 0.25 0.50 0.75 1.00 1.25

t=15 ps

0.00 0.25 0.50 0.75 1.00 1.25

t=20 ps

0.00 0.25 0.50 0.75 1.00 1.25

t=25 ps

0.00 0.25 0.50 0.75 1.00 1.25

t=30 ps

z (nm )

Figure 6 Distribution of number density of atoms at different times in argon illuminated with a laser pulse of 0.7 J/m2 The temperature distribution in argon at different times is shown in Fig 8 At 5 ps, laser heating just starts, and the target has a spatially uniform temperature of about 50 K Note that the initial size of the target extends from 4.5 nm to 12.6 nm Melting starts at 10 ps as indicated in Fig 7, and it is clear from Fig 8 that at this moment, the temperature is higher than the melting and the boiling point in the heated region, and is even close to the critical point At 15 ps, a flat region in the temperature distribution is observed around the location of 10

Trang 8

nm, which is the melting interface region The temperature in

this flat region is around 90 K, which is higher than the melting

point, indicating superheating at the melting front

9

10

11

12

13

14

15

16

Solid-liquid Interface Liquid-vapor Interface

Tim e (ps)

(a)

-500

-250

0

250

500

Solid-liquid Interface Liquid-vapor Interface

Tim e (ps)

(b)

Figure 7 (a) Positions (b) velocities of the solid-liquid interface

and the liquid-vapor interface in argon illuminated with a laser

pulse of 0.7 J/m2

An interesting phenomenon is observed at 20 ps, shortly

after melting stops At this moment, a minimum temperature is

observed at 9.5 nm The reason for this temperature drop is not

known yet, and is still under investigation This minimum

temperature disappears gradually due to heat transfer from the

surrounding higher temperature regions It is worth noting that

results of superheating, as well as the lack of a sharp

solid-liquid interface as mentioned previously, could not be predicted

using the continuum approaches without assumptions

40 60 80 100 120 140

t=5 ps t=10 ps t=15 ps t=20 ps t=25 ps t=30 ps

z (nm )

T m

T b

Figure 8 Temperature distribution in argon illuminated with a

laser pulse of 0.7 J/m2

The velocity distribution of vaporized atoms at different times is shown in Fig 9 At 10 ps, melting just starts, and the average velocity of atoms is close to zero except those on the surface, which have high kinetic energy due to the free boundary condition At 15 ps, a higher atomic velocity is observed At the vapor front, the velocity is close to 800 m/s, while at locations near the surface, the vapor velocity is much smaller At 30 ps, non-zero velocities are only observed at locations of 15 nm or further beyond the liquid-vapor interface

as indicated in Fig 7 This shows evaporation from the liquid surface is weak after laser heating stops

-200 0 200 400 600 800

t=5 ps t=10 ps t=15 ps t=20 ps t=25 ps t=30 ps

z (nm )

Figure 9 Spatial distribution of the average velocity in the z direction in argon illuminated with a laser pulse of 0.7 J/m2

Trang 9

1

2

3

M elting Evaporation

Tim e (ps)

(a)

-100

0

100

200

300

400

M elting Evaporation

Tim e (ps)

(b)

Figure 10 (a) Depths of the solid melted and vaporized, and (b)

rate of melting and evaporation in argon illuminated with a laser

pulse of 0.7 J/m2

0

1

2

3

Energy Fluence (J/m 2 )

Figure 11 The ablation depth induced by different laser

fluences in argon

The depth of melting and vaporization, as well as the melting and evaporation rates are shown in Fig 10 It is seen that the melting depth is much larger than the vaporization depth From Fig 10b it is found that melting happens mostly between 10 and 20 ps, while the evaporation process goes on until 25 ps, then reduces to a lower level corresponding to evaporation of liquid in vacuum The depths of ablation induced

by different laser fluences are shown in Fig 11

IV CONCLUSION

In this work, laser material interaction is studied using MD simulations Based on the results, the following conclusions are obtained First, during picosecond laser heating, the heat transfer process predicted using the continuum approach agrees with the result of the MD simulation Second, when melting happens, a transition region of about 1 nm, instead of a clear interface is found between the solid and the liquid During the melting process, the solid-liquid interface moves at almost a constant velocity much lower than the local sound velocity, while the liquid-vapor interface moves with a velocity close to the local equilibrium atomic velocity At the solid-liquid interface, superheating is observed Finally, the laser ablated material is found to move out of the target with a velocity of about a thousand meters per second

ACKNOWLEDGMENTS

Support to this work by the National Science Foundation (CTS-9624890) is gratefully acknowledged

REFERENCES

1 Shibahara, M., and Kotake, S., 1997, "Quantum Molecular Dynamics Study on Light-to-heat Absorption

Mechanism: Two Metallic Atom System," International Journal

of Heat and Mass Transfer, 40, pp 3209-3222

2 Shibahara, M., and Kotake, S., 1998, "Quantum Molecular Dynamics Study of Light-to-heat Absorption

Mechanism in atomic Systems," International Journal of Heat

and Mass Transfer, 41, pp 839-849

3 Häkkinen, H., and Landman, U., 1993, "Superheating,

Melting, and Annealing of Copper Surfaces," Physical Review

Letters, 71, pp 1023-1026

4 Anisimov, S I., Kapeliovich, B L., Perel'man, T L.,

1974, "Electron Emission from Metal Surfaces Exposed to

Ultra-short Laser Pulses," Soviet Physics, JETP, 39, pp

375-377

5 Kotake, S., and Kuroki, M., 1993, "Molecular Dynamics Study of Solid Melting and Vaporization by Laser Irradiation,"

International Journal of Heat and Mass Transfer, 36, pp

2061-2067

6 Herrmann, R F W., and Campbell, E E B., 1998,

"Ultrashort Pulse Laser Ablation of Silicon: an MD Simulation

Study," Applied Physics A, 66, pp 35-42

Trang 10

7 Zhigilei, L V., Kodali, P B S., and Garrison, J., 1997,

"Molecular Dynamics Model for Laser Ablation and Desorption

of Organic Solids," Journal of Physical Chemistry B, 101, pp

2028-2037

8 Zhigilei, L V., Kodali, P B S., and Garrison, J., 1998,

"A Microscopic View of Laser Ablation," Journal of Physical

Chemistry B, 102, pp 2845-2853

9 Ohmura, E., Fukumoto, I., and Miyamoto, I., 1999,

"Modified Molecular Dynamics Simulation on Ultrafast Laser

Ablation of Metal," the International Congress on Applications

of Lasers and Electro-Optics, 1999, pp 219-228

10 Girifalco, L A., and Weizer, V G., 1959, "Application

of the Morse Potential Function to Cubic Metals," Physical

Review, 114, pp 687-690

11 Etcheverry, J I., and Mesaros, M., 1999, "Molecular Dynamics Simulation of the Production of Acoustic Waves by

Pulsed Laser Irradiation," Physical Review B, 60, pp

9430-9434

12 Allen, M P and Tildesley, D J., 1987, Computer

Simulation of Liquids, Clarendon Press, Oxford

13 Broughton, J Q and Gilmer, G H., 1983, "Molecular Dynamics Investigation of the Crystal-fluid Interface I Bulk

Properties," Journal of Chemical Physics, 79, pp 5095-5104

14 Berendsen, H J C, Postma, J P M, van Gunsteren, W F., DiNola, A., and Haak, J R., 1984, "Molecular Dynamics

with Coupling to an External Bath," Journal of Chemical

Physics, 81, pp 3684-90

15 Peterson, O G., Batchelder, D N., and Simmons, R O.,

1966, "Measurements of X-Ray Lattice Constant, Thermal Expansivity, and Isothermal Compressibility of Argon Crystals,"

Physical Review, 150, pp 703-711

Ngày đăng: 06/05/2014, 08:55

TỪ KHÓA LIÊN QUAN