Next Article in Journal
Probing Electronic Strain Generation by Separated Electron-Hole Pairs Using Time-Resolved X-ray Scattering
Previous Article in Journal
Immune-Related Circulating miR-125b-5p and miR-99a-5p Reveal a High Recurrence Risk Group of Pancreatic Cancer Patients after Tumor Resection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sequences of Sub-Microsecond Laser Pulses for Material Processing: Modeling of Coupled Gas Dynamics and Heat Transfer

by
Andrey V. Gusarov
1,2,* and
Oleg B. Kovalev
2
1
Moscow State University of Technology STANKIN, Vadkovsky Per. 3a, 127055 Moscow, Russia
2
Khristianovich Institute of Theoretical and Applied Mechanics SB RAS, Institutskaya Str. 4/1, 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Appl. Sci. 2019, 9(22), 4785; https://doi.org/10.3390/app9224785
Submission received: 13 September 2019 / Revised: 3 November 2019 / Accepted: 5 November 2019 / Published: 8 November 2019
(This article belongs to the Section Applied Physics General)

Abstract

:

Featured Application

Material processing by pulsed lasers can suffer from plasma ignition, which reduces efficiency. This work compares the conditions at single- and multipulse laser processing and proposes a modeling tool to optimize the sequences of pulses.

Abstract

Multipulse laser processing of materials is promising because of the additional possibilities to control the thickness of the treated and the heat-affected zones and the energy efficiency. To study the physics of mutual interaction of pulses at high repetition rate, a model is proposed where heat transfer in the target and gas-dynamics of vapor and ambient gas are coupled by the gas-dynamic boundary conditions of evaporation/condensation. Numerical calculations are accomplished for a substrate of an austenitic steel subjected to a 300 ns single pulse of CO2 laser and a sequence of the similar pulses with lower intensity and 10 μs inter-pulse separation assuring approximately the same thermal impact on the target. It is revealed that the pulses of the sequence interact due to heat accumulation in the target but they cannot interact through the gas phase. Evaporation is considerably more intensive at the single-pulse processing. The vapor is slightly ionized and absorbs the infrared laser radiation by inverse bremsstrahlung. The estimated absorption coefficient and the optical thickness of the vapor domain are considerably greater for the single-pulse regime. The absorption initiates optical breakdown and the ignition of plasma shielding the target from laser radiation. The multipulse laser processing can be applied to avoid plasma ignition.

1. Introduction

Pulsed laser processing of materials has some advantages over continuum-wave processing because it can offer thinner processed and heat-affected zones and higher surface temperatures. The serious drawback of laser pulses with durations ranging from nanoseconds to microseconds is the ignition of plasma. The plasma shields the treated surface from the laser beam. Thus, the efficiency of laser treatment decreases with the pulse energy. High repetition-rate sequences of pulses is a well-known solution to increase the efficiency [1,2,3,4].
The computational model developed by Vora et al. [2] to understand the influence of multiple millisecond laser pulses on the surface finish is in agreement with experimentally observed surface roughness [2]. Ranjbar et al. [3] numerically simulated the dynamics of vapor plumes induced by nanosecond multi-pulse laser irradiation in the burst mode. They showed that the multi-pulse mode can considerably decrease the plasma ignition threshold and increase the efficiency of ablation. Cai et al. [4] modeled laser processing by the sequences of sub-millisecond pulses. Their model provided a theoretical guidance for optimizing the process parameters of laser ablation.
The complex of physical phenomena related to the interaction of a laser pulse with material includes absorption of laser radiation, heat transfer in the irradiated target followed by melting and evaporation, vapor expansion and interaction with the ambient gas, and laser beam absorption in the vapor, which can result in optical breakdown and plasma ignition. They were studied experimentally [5,6,7,8,9] and analyzed theoretically [9,10,11,12,13,14,15,16,17,18]. The above works indicate that the evaporation process is very important in laser processing. First, it is responsible for laser ablation. Conventional surface evaporation and phase explosion (volumetric evaporation) are the principal mechanisms for nanosecond pulses [19]. It seems that the phase explosion is more important for higher energies and shorter pulses while the surface evaporation becomes the principal ablation mechanism for longer pulses [2]. Due to high latent heat, evaporation can be the principal mechanism of heat loss. Thus, it influences heat transfer in the target. Finally, low ionization potential and high temperature of metal vapor facilitate optical breakdown [16], which drastically changes the mode of laser/material interaction. The surface temperature is the principal parameter influencing the rate of evaporation. This is why one can decouple gas dynamics and heat transfer in the target [19] under certain conditions. Generally, evaporation from the target surface and back condensation from the gas phase are competing processes [20]. Thus, the net evaporation rate depends also on the gas-phase parameters.
Gas dynamics of vapor and ambient gas coupled with heat transfer in the target was calculated by Bird’s direct Monte Carlo simulation (DSMC) [12,13,18]. If the mean free path in the gas phase is small, the numerical solution of Euler’s gas-dynamic equations is useful [10,11]. A strongly non-equilibrium Knudsen layer is formed in the gas phase near the target surface [20,21,22,23,24,25,26,27]. Rigorous boundary conditions at the Knudsen layer should be taken into account [15]. Gas-dynamic modeling of laser ablation [9,10,11,12,13,14,15,16,17,18] shows formation of a shock wave in the ambient gas and a contact temperature jump between the vapor and the ambient gas. The results of these calculations describe plume expansion and shock wave propagation at nanosecond pulsed laser ablation in ambient gas, which were observed experimentally [5,7]. Net back condensation of the vapor is expected at later stages of the gas-dynamic flow when the laser pulse is over [15].
The DSMC of laser evaporation by the sequences of nanosecond pulses [3] indicated that the second and the following pulses generate shock waves not only in the ambient gas but also in the vapor formed by the previous pulses. The pulses of the sequence are interdependent because of accumulation of heat in the target and because the gas parameters near the evaporating surface depend on the previous pulses [3]. If the inter-pulse separation time increases, the ablation depth decreases while the ignition of laser plasma becomes less probable [3]. The time of gas-phase relaxation after a laser pulse is shorter than that of target relaxation. Therefore, at high inter-pulse separation, the only mechanism of pulse interaction is the accumulation of heat in the target. For example, at modeling the sequences of millisecond pulses with inter-pulse separation of the order of 100 ms [2], the mutual influence of pulses through the gas phase was reasonably neglected.
Cai et al. [4] studied surface heat treatment by laser pulses of 1064 nm wavelength, 210 ns duration and the inter-pulse separation of the order of 10 µs. A rigorous model of conductive heat transfer in the target was proposed [4]. The dynamics of vapor and ambient gas was not considered in detail while simple models for evaporation and vapor ionization were used to take into account the shielding of the target by laser plasma. The model was validated by comparison with experimental results. The models with rigorous gas dynamics are known [3,15] but they were applied to nanosecond laser pulses.
A strong interdependence between heat transfer in the target and gas dynamics of vapor and ambient gas at laser–matter interaction was clearly realized back in the 1990s. The first models of heat transfer coupled with gas dynamics indicated the importance of the boundary conditions at the interface between the target and the gas phase [10,11]. Under the conditions of strong overheating and intensive evaporation, the near-boundary region of the gas phase (Knudsen layer) is highly non-equilibrium. Thus, the key issue is not gas dynamics but kinetic processes in the gas phase. In the first models, the kinetic description of the gas phase was used in the framework of DSMC [12,13]. Such an approach has been successfully applied to nanosecond laser ablation up to now [3]. However, the DSMC computation becomes extremely time consuming at small Knudsen numbers. This is why it is not applied in the recent models for sub-microsecond [4] and millisecond [3] laser pulses.
The present work implements a rigorous but less time-consuming computation model where the interface region of the Knudsen layer is considered explicitly [15]. It concerns modeling the sequences of CO2-laser pulses with 10.7 μm wavelength, the duration of 300 ns, and inter-pulse separation of 10 μs, which are promising for high-efficiency surface heat treatment [28]. Estimation of the ionization degree [27] and especially modeling the optical breakdown [4,16] require strong assumptions. To obtain conclusions that are more general, we restrict our consideration to qualitative effects of laser–plasma interaction. The calculation results obtained are rigorous before plasma ignition only. Note that both experiments and modeling [4,16] indicate that plasma ignition instantly interrupts the laser beam. It should be avoided at laser processing. The goals of the present study are to estimate the threshold of evaporation, the parameters of the vapor important for plasma ignition, and the physical mechanisms governing the mutual influence of the pulses in the multipulse sequences. The fraction of laser energy absorbed in the gas phase depends on its optical thickness. This is why the probability of plasma ignition is assessed by the optical thickness.

2. Model

In this model, heat transfer in the target and the dynamics of the ambient gas and the vapor are coupled through the boundary conditions at the interface. Only the central region of the laser spot is considered where the laser energy flux is supposed to be uniform. The thermal and the gas-dynamic responses are also uniform over this domain. Thus, the problem is essentially one-dimensional, at least at the initial stage of laser–material interaction. The position of the evaporating surface is not fixed, its recession velocity us being consistent with gas-dynamic and temperature fields. Recession velocity due to laser evaporation is much smaller than the characteristic gas-dynamic velocities [15]. Therefore, it is neglected in the gas-dynamic problem considered in half-space z > 0 of the frame where the positive direction of axis Z is the external normal to the target surface approximated by plane z = 0 and the origin is fixed at the surface (see Figure 1). By contrast, the recession can influence heat transfer in the target [15] considered in half-space z < 0.
The mathematical model used in this work is the thermal model of laser ablation [14,15], which was initially applied for single nanosecond pulses. It consists of the heat transfer equation in the target and the gas-dynamic equations for vapor and ambient gas. The heat transfer and gas dynamics are coupled by the gas-dynamic boundary conditions of evaporation/condensation [26].

2.1. Governing Equations

Heat conduction is supposed to be the principal mechanism of heat transfer in the target and is described by the heat equation in enthalpy form [14,15].
h t + u s h z = z λ T z
where h is the enthalpy per unit volume, t the time, λ the thermal conductivity, and T the temperature. The second term in the left hand side of Equation (1) is responsible for advection in the frame-moving relative medium. In the case of net evaporation, the interface recesses, the medium of the target moves in the positive direction of axis Z, and us > 0. In the case of net condensation, the interface rises, the medium of the target moves in the negative direction of axis Z, and us < 0. The volumetric specific heats in solid cs and liquid cl phases are approximately taken as constants. Then, the thermal equation of state [14],
T = h / c s T m T m + ( h c s T m q m ) / c l ,   ,   ,   h c s T m c s T m < h < c s T m + q m h c s T m + q m
relates the enthalpy and the temperature, where Tm is the melting point and qm the latent heat of melting per unit volume.
Gas-dynamic equations of continuity (3) and (4), momentum (5) and energy (6) [15],
n v t + ( n v u ) z = 0
n g t + ( n g u ) z = 0
( ρ u ) t + ( ρ u 2 + p ) z = 0
( E + ρ u 2 / 2 ) t + [ u ( E + ρ u 2 / 2 + p ) ] z = 0
are applied at z > 0, where nv and ng are the number densities of vapor and ambient gas, respectively, u the gas velocity, ρ the gas density, p the pressure and E the internal energy per unit volume. In the gas mixture [15],
ρ = mvnv + mgng,
p = (nv + ng)kT,
E = n v γ v 1 + n g γ g 1 k T
where mv and mg are the molecular masses and γv and γg the adiabatic exponents of vapor and ambient gas, respectively.
Mass balance at the interface, z = 0, is [15]
ntus = nvu,
where nt is the number density in the target. If laser energy is absorbed at the target surface with the heat effect per unit time per unit area equal to absorbed laser energy flux Ka, the energy balance at z = 0 is [15]
K a + u s h ( T s ) λ T z z = 0 0 = u n v U + u ( E + ρ u 2 / 2 + p )
where Ts is the temperature of the target surface and U the binding energy of target atoms at T = 0. Note that gas temperature at the interface T(0) is necessary to calculate the right hand side of Equation (11). Temperature T(0) is generally different from Ts because of a temperature jump across the Knudsen layer formed near the interface at strong evaporation or condensation. Binding energy U is the principal part of the latent heat of evaporation qv. In addition to Equations (10) and (11), two boundary conditions at z = 0 are necessary at evaporation and only one boundary condition is applied at condensation [26].
Gas-dynamic boundary conditions of evaporation/condensation relate the parameters of target surface with the parameters of vapor formed at the outer edge of the Knudsen layer. The target parameters are its surface temperature Ts and saturated vapor pressure ps at Ts. The saturated vapor pressure is estimated by integrating the Clausius–Clapeyron equation in assumption that the latent heat of evaporation qv is independent of temperature [14]
p s = p a exp q v k T b q v k T s ,
where Tb is the boiling temperature at atmospheric pressure pa. The thickness of the Knudsen layer at strong evaporation/condensation is about few mean free paths [26]. It is neglected in this work. Therefore, gas parameters, such as temperature T, pressure p and Mach number M are taken at z = 0. For example, the Mach number is defined as [15]
M = u ( 0 ) γ v k T ( 0 ) m v .
At evaporation M > 0 while at condensation M < 0. Boundary conditions of strong evaporation/condensation were obtained by gas-kinetic calculations of the Knudsen layer [20,21,22,23,24,25,26]. Analytical boundary conditions used in this work approximate the above cited rigorous calculations under the conditions typical for laser processing [15]. Evaporation and condensation are distinguished by comparing gas pressure p and saturated vapor pressure ps. At evaporation, p < ps, the two boundary conditions are [15]:
M = 1 ( p / p s ) 0.275 1 ( p 0 ) 0.275 ,
T T s = T 0 ( p 0 ) 0.125 + ( 1 T 0 ) ( p / p s ) 0.125 1 ( p 0 ) 0.125 ,
where p0 = 0.20742 is the pressure ratio and T0 = 0.6437 the temperature ratio at evaporation with M = 1. Note that M > 1 is impossible at evaporation [20,21,22,23,24,25,26]. At condensation, p > ps, the boundary condition is [15]
M = 1 ( p s / p ) α 1 ( I 0 ) α ,
where parameters α and I0 are functions of temperature ratio T/Ts specified in [15].

2.2. Computations

Partial differential Equations (1), (3)–(6) with boundary conditions at the interface given by Equations (10), (11), (14)–(16) are solved in computation domains shown in Figure 1 by the numerical method of finite volumes on uniform grids. The parameters of undisturbed ambient gas are imposed on the top boundary of the gas-phase domain. Ambient temperature Ta is the boundary condition at the bottom of the target domain. When the shock wave approaches to the top of the gas-phase domain or the heat-affected zone approaches to the bottom of the target domain, the corresponding computation domain is enlarged, the grid is rearranged, and the discrete fields from the previous grid are interpolated to obtain new discrete fields on the new grid. Thus, the computation domains are always greater than the disturbed zones, and the influence of the external boundaries is excluded. Differential Equations (1), (3)–(6) are approximated by second order in space- and first order in time-explicit finite difference equations. A second-order Godunov scheme [29] is used for gas-dynamic Equations (3)–(6). Central differences are applied to the spatial derivatives of heat Equation (1).
The present work uses original software for the thermal model of laser ablation, which was developed earlier [15]. The details of the numerical method and the calculation algorithm are given elsewhere [15].

3. Results

Laser treatment of austenitic steel X18H10T by pulsed CO2 laser in an air atmosphere is modeled. The parameters of the target and the ambient gas are listed in Table 1 and Table 2, respectively. Binding energy U from Equation (11) is approximated by latent heat of evaporation qv. The temporal profile of the laser pulse is specified by a triangle with the increasing front of τf = 50 ns and the decreasing back of τb = 250 ns (see Figure 2). The maximum of intensity Imax is a variable parameter. It is attained at 50 ns after the pulse beginning. To exclude the uncertainty due to the absorptance of the target, the absorbed part of laser energy Ka is specified as the model parameter. In the energy balance, Equation (11), it is explicitly defined as the following function of time:
K a = I max t τ f , 0 < t < τ f I max τ f + τ b t τ b , τ f < t < τ f + τ b ,
The pulse parameters are listed in Table 3.
Figure 2, Figure 3 and Figure 4 show the effect of a single pulse with the maximum absorbed intensity Imax = 15 MW/cm2. Figure 2 shows the temporal profile of the laser pulse specified by the function of absorbed laser power Ka versus time t. Figure 3 shows temperature fields in the target calculated for this pulse. Surface temperature attains the boiling point Tb and becomes greater than Tb. Therefore, an intensive vapor flow is formed along with a compressible flow of ambient air. Figure 4 shows gas-dynamic fields. Notice a shock front separating undisturbed and disturbed domains of air and a contact jump between air and vapor. The shock front is very close to the boundary of the computation domain for t = 0.2, 0.3, and 1 μs. Those are the instances just before the enlargement of the domain and the rearrangement of the computation grid. The shock front never touch the right boundary. At t = 0.1 and 0.2 μs (see Figure 4), gas-dynamic velocity u is positive at the interface z = 0 indicating net evaporation but at the end of the laser pulse at t = 0.3 μs, u(0) becomes negative indicating net condensation. Figure 2 shows the dynamics of the surface temperature Ts and the heat loss to evaporation calculated from the right hand side of Equation (11). One can see that the evaporation loss is much smaller than the absorbed laser energy for this pulse and the surface temperature can be considerably greater than the boiling point Tb = 3200 K.
A single pulse with smaller intensity Imax = 10 MW/cm2 cannot heat the target up to the boiling point (see the dotted line in Figure 5). However, the heat of the pulse is accumulated in the surface layer of the target. Figure 6 shows that at the moment of t = 10 μs after the pulse, the surface temperature is still about 600 K. The second pulse of the sequence starting at t = 10 μs, interacts with the already heated target. Therefore, it results in higher surface temperature (dashed line in Figure 5). The surface temperature becomes greater than the boiling point and evaporation starts. At the beginning of the third pulse at t = 20 μs, the surface temperature attains 800 K (see Figure 6a), the temperature after the third pulse is higher (full line in Figure 5), and evaporation should be more intensive. Figure 6b shows the evolution of the temperature field in the target after the third pulse and Figure 7 presents the gas-dynamic fields. The gas-dynamic fields are qualitatively similar to the fields shown in Figure 4 for a single pulse with Imax = 15 MW/cm2. The maximum vapor pressure in Figure 4 is about 5 atm. while it is around 2 atm. in Figure 7. Therefore, one can conclude that evaporation is less intensive at Imax = 10 MW/cm2, even after the third pulse.
Evaporation/condensation at the target surface is characterized by such parameters of the Knudsen layer as pressure ratio p/ps, temperature ratio T/Ts, and Mach number M related by the boundary conditions of evaporation/condensation, Equations (14)–(16). Figure 8 shows the evolution of these parameters at the studied regimes of laser processing. The effect of laser radiation on the target is characterized by the evolution of the temperature field shown in Figure 3 and Figure 6 along with the melt depth in the target shown in Figure 9.

4. Discussion

The modeling indicates that at inter-pulse separation 10 μs, the heat transferred to the target by a laser pulse is not completely dissipated when the next pulse of the sequence starts. Thus, thermal energy is accumulated in a near-surface layer of the target. It is illustrated by Figure 5 where the temperature curve of each pulse is above the temperature curve of the previous pulse. The intensity of evaporation by the pulses of a sequence increases because of increasing surface temperature.
In the modeled regimes of single and multipulse laser processing, the vapor velocity near the target surface becomes negative even before the termination of the 300 ns laser pulse (see Figure 4 and Figure 7). Thus, back vapor condensation on the target surface starts. According to the model, all the evaporated material condenses back. The period of time necessary for full back condensation depends on the parameters of laser processing. The drop of the absolute value of Mach number |M| down to zero in Figure 8 indicates the instant of complete back condensation. For single pulse with Imax = 15 MW/cm2 (see Figure 8b), it is t = 1900 ns. At Imax = 10 MW/cm2, the duration of back condensation is considerably less. For the second pulse (see Figure 8c), t = 315 ns. For the third pulse (see Figure 8d), t = 415 ns. In any case, back condensation finishes well before the beginning of the next laser pulse at t = 10 μs. This is why the laser radiation of the next pulse cannot interact with the vapor formed by the previous pulse.
Figure 8 shows that Mach number M approaches its theoretical limit for evaporation M = 1 [20,21,22] but does not attain it. This is because of a relatively low laser intensity. Calculations by the similar model at higher intensities [15] indicated that M = 1 can be attained. At evaporation, temperature ratio T/Ts and pressure ratio p/ps are functions of M given by Equations (14) and (15). Note that the temperature ratio attains its minimum when evaporation is the most intensive (see Figure 8). At the stage of condensation, M tends to −1 (see Figure 8). Note that supersonic condensation with |M| > 1 is theoretically possible [24,25]. However, it is not observed in the present work as well as in the previous one [15]. For the calculated examples, temperature ratio at condensation is in the interval between 0.4 and 1 (see Figure 8). This means that the vapor is colder than the target. Such an unusual ratio is explained by adiabatic expansion of vapor where its temperature rapidly decreases as a power function of the density. The temperature of the target surface is controlled by heat conduction and decreases more slowly. Even if the vapor is colder, its pressure is still greater than the saturated vapor pressure at the temperature of the target surface. This is why condensation continues. Once the Mach number of the condensing vapor flow attains −1, the value of M is around −1 till the complete condensation of the vapor (see Figure 8). Similar behavior of the condensing vapor flow was observed at modeling for much shorter nanosecond pulses [15].
Figure 9 shows the calculated dynamics of the melting/solidification front. The maximum melt depth estimates the thickness of the surface layer with the modified structure. This value appears to be approximately the same for the single laser pulse with Imax = 15 MW/cm2 and for the sequence of three weaker pulses with Imax = 10 MW/cm2 (see Figure 9). It is slightly greater than one micron. However, the parameters of vapor are quite different. Table 4 compares the parameters at 200 ns after the beginning of the single pulse with Imax = 15 MW/cm2 (t = 0.2 μs in Figure 4) and the third pulse of the sequence with Imax = 10 MW/cm2 (t = 20.2 μs in Figure 7). In both cases, the vapor temperature is about 3000 K. The vapor pressure at 15 MW/cm2 is four times the pressure at 10 MW/cm2. The thickness of the vapor domain at 15 MW/cm2 is four times its thickness at 10 MW/cm2. In summary, at the miltipulse laser processing, vapor is considerably less dense and occupies a considerably thinner domain.
Metallic vapor is expected to be slightly ionized at the temperature of 3000 K. The source of charges is thermoelectron and thermoion emission from the target surface [27]. Modeling [27] revealed the Knudsen layer and the sheath fused together at the interface between an evaporating surface and slightly ionized vapor. It was shown that the electronic temperature is around the surface temperature and the ionization degree is around the equilibrium one at the electronic temperature [27]. The last row of Table 4 shows the calculated surface temperature estimating the electronic temperature of the vapor. One can see that the electronic temperature is greater for the single pulse with Imax = 15 MW/cm2. According to the Saha Equation [31], the ionization degree sharply grows with the electronic temperature. Therefore, the ionization degree can be considerably greater at 15 MW/cm2. The principal mechanism of the absorption of infrared radiation in ionized gas is inverse bremsstrahlung [31]. The absorption coefficient is proportional to the concentration of electrons [31].
The probability of optical breakdown in the vapor can be assessed from the optical thickness, which is roughly the vapor layer thickness (optical path) multiplied by the absorption coefficient of the vapor. Compare, for example, the distributions of vapor density at the end of the single laser pulse (see Figure 4, t = 0.3 μs) and at the end of the third pulse in the sequence (see Figure 7, t = 20.3 μs). The thickness of the vapor domain is about 140 μm in the former case and only 30 μm in the latter case. Thus, the optical path through vapor considerably reduces at the multipulse laser processing. One can see from these figures that the vapor density is around the density of the ambient gas in the former case and approximately a quarter of the ambient gas density in the latter case. The electron temperature and the ionization degree in the vapor are lower at the multipulse laser processing (see above). The electron density is the vapor density multiplied by the ionization degree. Therefore, the electron density along with the absorption coefficient considerably reduces at the multipulse laser processing. Finally, the optical thickness of the vapor considerably reduces too. The absorption initiates optical breakdown and the ignition of plasma shielding the target from laser radiation. Thus, the calculations indicate that at the miltipulse laser processing, the vapor parameters are considerably less favorable for plasma ignition while the thermal impact on the target is essentially the same. This is why the multipulse processing can be applied to avoid plasma ignition.
The present work proposes a modeling tool, which gives tendencies and can be a guideline in experiments. To the best knowledge of the authors, the present work is the first where gas-kinetic processes are taken into account and analyzed for sub-microsecond laser pulses. The two examples considered show the contribution of different physical processes. Based on the achieved understanding of the competing processes, one can qualitatively estimate the influence of the laser parameters. For the given laser setup, only two parameters are variable: the pulse energy and the repetition rate. The modeling indicates that increasing the pulse energy sharply increases the optical thickness of the vapor, which results in optical breakdown. This is why increasing the pulse energy is not desirable. On the other hand, increasing the repetition rate above the 100 kHz considered here (corresponds to the 10 μs inter-pulse separation) will increase heat accumulation in the target at a much lower value of the optical thickness. Thus, the thermal impact on the target will increase while the probability of plasma ignition will be still low. Therefore, increasing the repetition rate above 100 kHz seems to be promising in order to increase the efficiency of the laser processing.

5. Conclusions

  • Heat transfer in the target and gas dynamics of vapor and ambient gas coupled by gas-dynamic boundary conditions of evaporation/condensation are calculated for a single laser pulse with the maximum absorbed power density Imax = 15 MW/cm2 and a multipulse laser processing with Imax = 10 MW/cm2.
  • The calculations indicate that the thermal impact on the target is essentially the same for the single pulse and the multipulse sequences. However, the vapor parameters are very different. Generally, evaporation is considerably more intensive at the single-pulse processing.
  • Thermal impact on the target and evaporation become more intensive with an increase in the number of pulses in the sequence because of heat accumulation in the target.
  • In the considered regimes, vapor formed during a laser pulse condenses back on the target surface during the time interval between consecutive pulses. Therefore, the pulses of the sequence can not interact through the gas phase.
  • The modeling indicates that the absorption coefficient and the total optical thickness are considerably greater in single-pulse laser processing. The absorption initiates optical breakdown and the ignition of plasma shielding the target from laser radiation. Therefore, multipulse laser processing can be applied to avoid plasma ignition.

Author Contributions

Conceptualization, O.B.K.; methodology, A.V.G.

Funding

This research was funded by Russian Science Fundation grant number 18-19-00430.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

cspecific heat
Einternal energy per unit volume
henthalpy per unit volume
Imaxmaximum laser energy flux
Klaser energy flux
kBoltzmann constant
MMach number
mmolecular mass
nnumber density
ppressure
qlatent heat
Ttemperature
ttime
Ubinding energy
uvelocity
zcoordinate
Greek Symbols
γadiabatic exponent
λthermal conductivity
ρmass density
τtime interval
Subscripts
aambient, absorption
bboiling, back
ffront
ggas
lliquid
mmelting
ssolid, surface
vvapor

References

  1. Dahotre, N.B.; Harimkar, S.P. Laser Fabrication and Machining of Materials; Springer: New York, NY, USA, 2008. [Google Scholar]
  2. Vora, H.D.; Santhanakrishnan, S.; Harimkar, S.P.; Boetcher, S.K.S.; Dahotre, N.B. One-dimensional multipulse laser machining of structural alumina: Evolution of surface topography. Int. J. Adv. Manuf. Technol. 2013, 68, 69–83. [Google Scholar] [CrossRef]
  3. Ranjbar, O.A.; Lin, Z.; Volkov, A.N. One-dimensional kinetic simulations of plume expansion induced by multipulse laser irradiation in the burst mode at 266 nm wavelength. Vacuum 2018, 157, 361–375. [Google Scholar] [CrossRef]
  4. Cai, S.; Chen, G.; Zhou, C. Research and application of surface heat treatment for multipulse laser ablation of materials. Appl. Surf. Sci. 2015, 355, 461–472. [Google Scholar] [CrossRef]
  5. Wen, S.; Mao, X.; Greif, R.; Russo, R.E. Laser ablation induced vapor plume expansion into a background gas. II. Experimental analysis. J. Appl. Phys. 2007, 101, 023115. [Google Scholar] [CrossRef]
  6. Li, X.; Wang, Z.; Mao, X.; Russo, R.E. Spatially and temporally resolved spectral emission of laser-induced plasmas confined by cylindrical cavities. J. Anal. At. Spectrom. 2014, 29, 2127–2135. [Google Scholar] [CrossRef]
  7. Harilal, S.S.; Bindhu, C.V.; Tillack, M.S.; Najmabadi, F.; Gaeris, A.C. Internal structure and expansion dynamics of laser ablation plumes into ambient gases. J. Appl. Phys. 2003, 93, 2380. [Google Scholar] [CrossRef]
  8. Wood, R.F.; Leboeuf, J.N.; Geohegan, D.B.; Puretzky, A.A.; Chen, K.R. Dynamics of plume propagation and splitting during pulsed-laser ablation of Si in He and Ar. Phys. Rev. B 1998, 58, 1533–1543. [Google Scholar] [CrossRef]
  9. Bulgakov, A.V.; Bulgakova, N.M. Gas-dynamic effects of the interaction between a pulsed laser-ablation plume and the ambient gas: Analogy with an underexpanded jet. J. Phys. D 1998, 31, 693–703. [Google Scholar] [CrossRef]
  10. Ho, J.R.; Grigoropoulos, C.P.; Humphrey, J.A.C. Computational study of heat transfer and gas dynamics in the pulsed laser evaporation of metals. J. Appl. Phys. 1995, 78, 4696–4709. [Google Scholar] [CrossRef]
  11. Ho, J.R.; Grigoropoulos, C.P.; Humphrey, J.A.C. Gas dynamics and radiation heat transfer in the vapor plume produced by pulsed laser irradiation of aluminum. J. Appl. Phys. 1996, 79, 7205–7215. [Google Scholar] [CrossRef]
  12. Itina, T.E.; Katassonov, A.A.; Marine, W.; Autric, M. Numerical study of the role of a background gas and system geometry in pulsed laser deposition. J. Appl. Phys. 1998, 83, 6050–6054. [Google Scholar] [CrossRef] [Green Version]
  13. Itina, T.E.; Hermann, J.; Delaporte, P.; Sentis, M. Laser-generated plasma plume expansion: Combined continuous-microscopic modeling. Phys. Rev. E 2002, 66, 066406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Gusarov, A.V.; Gnedovets, A.G.; Smurov, I. Gas dynamics of laser ablation: Influence of ambient atmosphere. J. Appl. Phys. 2000, 88, 4352–4364. [Google Scholar] [CrossRef]
  15. Gusarov, A.V.; Smurov, I. Thermal model of nanosecond pulsed laser ablation: Analysis of energy and mass transfer. J. Appl. Phys. 2005, 97, 014307. [Google Scholar] [CrossRef]
  16. Mazhukin, V.I.; Nossov, V.V.; Smurov, I. Modeling of plasma-controlled evaporation and surface condensation of Al induced by 1.06 and 0.248 μm laser radiations. J. Appl. Phys. 2007, 101, 024922. [Google Scholar] [CrossRef]
  17. Chen, Z.; Bleiner, D.; Bogaerts, A. Effect of ambient pressure on laser ablation and plume expansion dynamics: A numerical simulation. J. Appl. Phys. 2006, 99, 063304. [Google Scholar] [CrossRef] [Green Version]
  18. Palya, A.; Ranjbar, O.A.; Lin, Z.; Volkov, A.N. Effect of the background gas pressure on the effectiveness of laser-induced material removal from deep cavities in irradiated targets. Appl. Phys. A 2018, 124, 32. [Google Scholar] [CrossRef]
  19. Wang, Y.; Shen, N.; Befekadu, G.K.; Pasiliao, C.L. Modeling pulsed laser ablation of aluminum with finite element analysis considering material moving front. Int. J. Heat Mass Transf. 2017, 113, 1246–1253. [Google Scholar] [CrossRef]
  20. Anisimov, S.I. Vaporization of metal absorbing laser radiation. Sov. Phys. JETP 1968, 27, 182. [Google Scholar]
  21. Ytrehus, T. Theory and experiments on gas kinetics in evaporation. In Rarefied Gas Dynamics; Potter, J.L., Ed.; AIAA: New York, NY, USA, 1977; pp. 1197–1212. [Google Scholar]
  22. Knight, C.J. Theoretical modeling of rapid surface vaporization with back pressure. AIAA J. 1979, 17, 519–523. [Google Scholar] [CrossRef]
  23. Sone, Y.; Sugimoto, H. Strong evaporation from a plane condensed phase. In Adiabatic Waves in Liquid-Vapor Systems; Meier, G.E.A., Thompson, P.A., Eds.; Springer: Berlin, Germany, 1990; pp. 293–304. [Google Scholar]
  24. Aoki, K.; Sone, Y.; Yamada, T. Numerical analysis of gas flows condensing on its plane condensed phase on the basis of kinetic theory. Phys. Fluids A 1990, 2, 1867–1878. [Google Scholar] [CrossRef]
  25. Sone, Y.; Takata, S.; Golse, F. Notes on the boundary conditions for fluid-dynamic equations on the interface of a gas and its condensed phase. Phys. Fluids 2001, 13, 324. [Google Scholar] [CrossRef]
  26. Gusarov, A.V.; Smurov, I. Gas-dynamic boundary conditions of evaporation and condensation: Numerical analysis of the Knudsen layer. Phys. Fluids 2002, 14, 4242–4255. [Google Scholar] [CrossRef]
  27. Gusarov, A.V.; Aoki, K. Ionization degree for strong evaporation of metals. Phys. Plasmas 2005, 12, 083503. [Google Scholar] [CrossRef] [Green Version]
  28. Bagayev, S.N.; Grachev, G.N.; Ponomarenko, A.G.; Smirnov, A.L.; Demin, V.N.; Okotrub, A.V.; Baklanov, A.M.; Onischuk, A.A. A new method of laser-plasma synthesis of nanomaterials. First results and prospects. Proc. SPIE 6732. In Proceedings of the International Conference on Lasers, Applications and Technologies 2007: Laser-assisted Micro-and Nanotechnologies, Minsk, Belarus, 28 June 2007; pp. 62–71. [Google Scholar] [CrossRef]
  29. Harten, A.; Engquist, B.; Osher, S.; Chakravarthy, S.R. Uniformly high order accurate essentially non-oscillatory schemes, III. J. Comput. Phys. 1997, 131, 3–47. [Google Scholar] [CrossRef]
  30. Gray, D.E. American Institute of Physics Handbook; McGraw-Hill: New York, NY, USA, 1972. [Google Scholar]
  31. Zel’dovich, Y.B.; Raiser, Y.P. Physics of Shock Waves and High-Temperature Hydrodynamic Phenomena, 2nd ed.; Academic Press: New York, NY, USA, 1967. [Google Scholar]
Figure 1. Interaction between the laser beam, the target, and the gas phase.
Figure 1. Interaction between the laser beam, the target, and the gas phase.
Applsci 09 04785 g001
Figure 2. Temporal profiles of the absorbed laser power, the surface temperature, and the evaporation loss for a single pulse with Imax = 15 MW/cm2.
Figure 2. Temporal profiles of the absorbed laser power, the surface temperature, and the evaporation loss for a single pulse with Imax = 15 MW/cm2.
Applsci 09 04785 g002
Figure 3. Dynamics of the temperature field in the target for a single pulse with Imax = 15 MW/cm2.
Figure 3. Dynamics of the temperature field in the target for a single pulse with Imax = 15 MW/cm2.
Applsci 09 04785 g003
Figure 4. Evolution of the gas-dynamic fields at single-pulse laser processing with Imax = 15 MW/cm2. The arrow near each curve points to the corresponding axis. The axis title contains the name of the physical quantity. Flow velocity (dashed curves) and mass densities of vapor (red curves) and gas (blue curves): (a,c,e,g). Temperature (black curves) and pressure (blue curves): (b,d,f,h). Time t: 0.1 μs (a,b); 0.2 μs (c,d); 0.3 μs (e,f); 1 μs (g,h).
Figure 4. Evolution of the gas-dynamic fields at single-pulse laser processing with Imax = 15 MW/cm2. The arrow near each curve points to the corresponding axis. The axis title contains the name of the physical quantity. Flow velocity (dashed curves) and mass densities of vapor (red curves) and gas (blue curves): (a,c,e,g). Temperature (black curves) and pressure (blue curves): (b,d,f,h). Time t: 0.1 μs (a,b); 0.2 μs (c,d); 0.3 μs (e,f); 1 μs (g,h).
Applsci 09 04785 g004
Figure 5. Increase of the surface temperature due to heat accumulation in the target at laser processing by the sequence of pulses with Imax = 10 MW/cm2 and the inter-pulse separation 10 μs.
Figure 5. Increase of the surface temperature due to heat accumulation in the target at laser processing by the sequence of pulses with Imax = 10 MW/cm2 and the inter-pulse separation 10 μs.
Applsci 09 04785 g005
Figure 6. Dynamics of the temperature field in the target for a sequence of pulses with Imax = 10 MW/cm2, and inter-pulse separation 10 μs: (а) moment t = 10 μs corresponds to the beginning of the second pulse, the third pulse starts at t = 20 μs, and the fourth pulse starts at t = 30 μs; (b) more details for the third pulse.
Figure 6. Dynamics of the temperature field in the target for a sequence of pulses with Imax = 10 MW/cm2, and inter-pulse separation 10 μs: (а) moment t = 10 μs corresponds to the beginning of the second pulse, the third pulse starts at t = 20 μs, and the fourth pulse starts at t = 30 μs; (b) more details for the third pulse.
Applsci 09 04785 g006
Figure 7. Evolution of the gas-dynamic fields after the third pulse of the sequence with Imax = 10 MW/cm2 and inter-pulse separation 10 μs. The arrow near each curve points to the corresponding axis. The axis title contains the name of the physical quantity. Flow velocity (dashed curves) and mass densities of vapor (red curves) and gas (blue curves): (a,c,e,g). Temperature (black curves) and pressure (blue curves): (b,d), (f,h). Time t: 20.15 μs (a,b); 20.2 μs (c,d); 20.3 μs (e,f); 21 μs (g,h).
Figure 7. Evolution of the gas-dynamic fields after the third pulse of the sequence with Imax = 10 MW/cm2 and inter-pulse separation 10 μs. The arrow near each curve points to the corresponding axis. The axis title contains the name of the physical quantity. Flow velocity (dashed curves) and mass densities of vapor (red curves) and gas (blue curves): (a,c,e,g). Temperature (black curves) and pressure (blue curves): (b,d), (f,h). Time t: 20.15 μs (a,b); 20.2 μs (c,d); 20.3 μs (e,f); 21 μs (g,h).
Applsci 09 04785 g007
Figure 8. Parameters of the Knudsen layer at evaporation/condensation at different regimes of laser processing: single pulse processing (a,b); multipulse processing: (c,d). The arrow near each curve points to the corresponding axis: temperature ratio (black curves); Mach number (red curves); inverse pressure ratio (dashed blue curves).
Figure 8. Parameters of the Knudsen layer at evaporation/condensation at different regimes of laser processing: single pulse processing (a,b); multipulse processing: (c,d). The arrow near each curve points to the corresponding axis: temperature ratio (black curves); Mach number (red curves); inverse pressure ratio (dashed blue curves).
Applsci 09 04785 g008
Figure 9. Melt depth in the target at single pulse laser processing with Imax = 15 MW/cm2 and multipulse processing with Imax = 10 MW/cm2 and inter-pulse separation 10 μs.
Figure 9. Melt depth in the target at single pulse laser processing with Imax = 15 MW/cm2 and multipulse processing with Imax = 10 MW/cm2 and inter-pulse separation 10 μs.
Applsci 09 04785 g009
Table 1. Properties of steel Х18Н10Т accepted for modeling [30].
Table 1. Properties of steel Х18Н10Т accepted for modeling [30].
Melting point, Tm, (K)1750
Boiling point, Tb (K)3200
Latent heat of melting, qm (GJ/m3)2.18
Latent heat of evaporation, qv (GJ/m3)49.7
Specific heat in solid phase, cs (MJ/(m3 K))4.25
Specific heat in liquid phase, cl (MJ/(m3 K))5.95
Thermal conductivity, λ (W/(m K))20
Density, ρ (kg/m3)7890
Atomic mass, mv, (a.m.u.)μ55.85
Adiabatic exponent of vapor, μv5/3
Table 2. Properties of air accepted for modeling.
Table 2. Properties of air accepted for modeling.
Ambient pressure, pa (atm.)1
Ambient temperature, Ta (K)298
Density at the ambient conditions, ρ0 (kg/m3)1.185 [30]
Atomic mass, mg, (a.m.u.)29 [30]
Adiabatic exponent, μg7/5 [30]
Table 3. Parameters of the laser pulse.
Table 3. Parameters of the laser pulse.
Duration (ns)300
Inter-pulse separation (μs)10
Absorbed intensity at the maximum, Imax (MW/cm2)10 … 15
Table 4. Calculated parameters of vapor.
Table 4. Calculated parameters of vapor.
Maximum absorbed intensity, Imax (MW/cm2)1510
Time t, (μs)0.220.2
Position of the contact surface, z (μm)7518
Vapor pressure, p (atm.)~8~2
Vapor temperature, T (K)~3000~3000
Target surface temperature, (K)42003400

Share and Cite

MDPI and ACS Style

Gusarov, A.V.; Kovalev, O.B. Sequences of Sub-Microsecond Laser Pulses for Material Processing: Modeling of Coupled Gas Dynamics and Heat Transfer. Appl. Sci. 2019, 9, 4785. https://doi.org/10.3390/app9224785

AMA Style

Gusarov AV, Kovalev OB. Sequences of Sub-Microsecond Laser Pulses for Material Processing: Modeling of Coupled Gas Dynamics and Heat Transfer. Applied Sciences. 2019; 9(22):4785. https://doi.org/10.3390/app9224785

Chicago/Turabian Style

Gusarov, Andrey V., and Oleg B. Kovalev. 2019. "Sequences of Sub-Microsecond Laser Pulses for Material Processing: Modeling of Coupled Gas Dynamics and Heat Transfer" Applied Sciences 9, no. 22: 4785. https://doi.org/10.3390/app9224785

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop