You are on page 1of 11

Optics & Laser Technology 170 (2024) 110313

Contents lists available at ScienceDirect

Optics and Laser Technology


journal homepage: www.elsevier.com/locate/optlastec

Full length article

Numerical and experimental analysis of dynamic process of laser ablation


epoxy resin
Haijun Yu a, b, Zhiyan Zhang a, *, Shusen Zhao a, Hongyang Wang a, b, Han Liang a, b,
Zhiyong Dong a, Xuechun Lin a, b, *
a
Laboratory of All-solid-state Light Sources, Institute of Semiconductors, Chinese Academy of Sciences, Beijing 100083, China
b
College of Materials Science and Opto-Electronic Technology, University of Chinese Academy of Sciences, Beijing 100049 China

A R T I C L E I N F O A B S T R A C T

Keywords: The numerical simulation of polymer laser ablation is essential for applications such as laser propulsion and laser
Laser ablation deposition. Based on the pyrolysis effect of polymers, this paper establishes a dynamics model of polymer laser
Polymer ablation. Numerical solution software FLUENT was used to solve the equations involving mass, momentum, and
Numerical simulation
energy. With user-defined functions (UDF) and dynamic mesh, the deposition of laser energy, the depression of
Dynamics
Pyrolysis
the material surface, and the generation of high-temperature pyrolysis gas plumes were simulated. Epoxy resin
was selected as the experimental material, and its surface changes were observed after exposure to various laser
pulse powers. High-speed shadow imaging techniques were employed to examine epoxy resin’s dynamic laser
ablation process. Compared to experimental results, numerical calculations accurately reflected the dynamic
behaviours of epoxy resin after pulse laser interaction, such as the expansion of pyrolysis gas plumes and the
propagation of shockwaves.

1. Introduction the dynamic characteristics of femtosecond laser ablation of aluminium,


silicon, and glass media[12]. S.L. Johnson et al. and Jifei Ye et al. also
Laser ablation is a process that uses lasers to remove materials. When used time-resolved shadow imaging techniques to study the kinetics of
a high-power laser is focused onto the surface of a material, the laser laser ablation of polymers[13,14]. Numerical studies simulate phe­
energy is absorbed and converted into thermal energy, causing localized nomena like material ablation, shockwaves, and plasmas in laser abla­
heating of the material. At high energy densities, the material’s tem­ tion by solving equations for energy, momentum, and continuity,
perature rises rapidly, exceeding its thermal decomposition or vapou­ focusing mainly on metals and polymers. T. MOŚCICKI et al. theoreti­
rization temperature, and the material begins to evaporate or thermally cally studied the interaction process of a laser beam with metal targets. A
decompose. The laser ablation technique has advantages such as high thermal model of the laser and target interaction was established,
precision, high efficiency, and strong controllability. It has been widely describing the target’s heating and the plasma’s formation and expan­
applied to metals[1–3], semiconductors[4,5], glass[6,7], ceramics[8], sion, and was solved using the Fluent software package[15]. Jianian
polymers[9], and biological tissues[10]. Yang et al. and He Yang et al., respectively, used the commercial finite
Current research on laser ablation mainly involves two aspects: ex­ element software ANSYS to establish a three-dimensional finite element
periments and numerical calculations. Experimental studies primarily model of laser ablation of the resin paint layer, studying the surface
investigate the effects of various laser parameters (pulse width, power, morphology of the material after ablation[16,17].
etc.) on material ablation morphology, as well as using high-speed Laser ablation numerical modelling of polymers differs significantly
cameras, ICCD, etc., to study the kinetics of laser ablation. M.E. Sha­ from metals due to differences in their removal mechanisms. Metal
heen et al. observed how near-infrared (785 nm) picosecond and deep removal by lasers mainly involves melting and vapourization, which are
ultraviolet (193 nm) ArF excimer lasers interact with crystalline silicon, physical processes, while the removal of polymers involves not only
noting differences in surface morphology, ablation rate, and thresholds physical processes but also chemical changes[18–21]. Two research
[11]. Using time-resolved shadow techniques, Nan Zhang et al. studied aspects need to be further developed for numerical studies of polymer

* Corresponding authors at: Laboratory of All-solid-state Light Sources, Institute of Semiconductors, Chinese Academy of Sciences, Beijing 100083, China (X. Lin).
E-mail addresses: zyzhang@semi.ac.cn (Z. Zhang), xclin@semi.ac.cn (X. Lin).

https://doi.org/10.1016/j.optlastec.2023.110313
Received 14 September 2023; Received in revised form 22 October 2023; Accepted 2 November 2023
Available online 10 November 2023
0030-3992/© 2023 Elsevier Ltd. All rights reserved.
H. Yu et al. Optics and Laser Technology 170 (2024) 110313

Table 1 dynamic process of laser ablation, as shown in Fig. 1. The shadow im­
The laser parameters. aging system comprises a high-speed camera (Revealer X213) and a
Parameter Symbol Value supplementary light source (808 nm semiconductor CW laser). The laser
produced by the 808 nm semiconductor laser is expanded and colli­
Wavelength (nm) λ 1064
Average Power (W) P 0–30 mated through a beam expander system. The 808 nm laser is incident
Pulse Width (ns) τFWHM 200 parallel to the surface of the sample and is directly projected into the
Repetition Frequency (Hz) f 30,000 lens of the high-speed camera through a reflective mirror. An 808 nm
Beam Radius (μm) wz ~45 narrow bandpass filter was placed in front of the high-speed camera lens
to filter out stray light generated during the ablation process. When the
laser ablation. Firstly, most numerical models of polymer ablation have pulse laser emitted by the fiber laser illuminates the material surface, the
adopted traditional metal models, which need to be improved. Secondly, material undergoes pyrolysis, producing a large amount of high-
most research focuses on the surface morphology of the material after temperature gas. These high-temperature gases cause changes in the
laser ablation, with fewer studies on kinetic behaviours, such as the refractive index of the air near the material’s surface. Based on this
generation of pyrolysis gases and the expansion of gas plumes during principle, the high-speed camera captures the laser ablation process.
laser ablation. However, research in this area plays an important role in A signal synchronization was established between the laser ablation
applications such as laser propulsion[22] and laser deposition[23]. system and the high-speed camera by the digital delay generator (TDG-
This paper aims to establish a dynamic model of polymer laser IV). When the laser ablation starts, the high-speed camera begins signal
ablation, and to simulate the phenomena of material pyrolysis, plume acquisition simultaneously. The high-speed camera’s pixel resolution,
generation and shock wave propagation. In order to achieve this goal, exposure time, and frame rate are 256 × 176, 500 ns, and 66666 Hz,
the modeling method in reference [24](studied the thermochemical respectively, meaning that the high-speed camera captures an image
ablation of carbon-phenolic jet vane) was referred to. The energy source every 15 μs.
term is added to the polymer internal grid to simulate the process of In addition, the surface morphology of the material was character­
polymer absorbing laser energy. The pyrolysis gas escape process is ized using a 3D scanning microscope (OLS4000).
realized by adding mass, momentum, energy and species source terms to
the first layer flow field grid near the wall. Numerical simulation soft­ 3. Numerical model
ware Ansys Fluent is used as the solver to solve the numerical model.
The User Defined Function (UDF) was employed to realize the loading of 3.1. Model introduction
source terms during the solving process. Additionally, experiments were
conducted to obtain the laser ablation morphology and shadow images, The typical sequence for laser ablation of polymers is as follows:
which were compared with the numerical simulation results. (i) The laser acts on the material surface.
(ii) The material absorbs the laser energy, causing a temperature rise.
2. Experiment (iii) The material undergoes thermal decomposition, generating
gases which then diffuses into the air.
Bisphenol-a epoxy resin was selected as the experimental material To fully simulate the laser ablation process of polymers, it is neces­
for this study. The chosen laser is the YLPN-1–4 × 200–30-M fiber laser sary to consider both the heating of the polymer (occurring in the solid
produced by IPG company. The emitted laser light is reflected by a two- region) and the diffusion process of the decomposition gases (occurring
dimensional galvanometer and subsequently focused onto the material in the fluid region). Therefore, the numerical simulation area must
surface using an f-θ lens. include both solid and fluid regions. The material in the solid region is
The laser parameters are shown in Table 1. the polymer. The materials in the fluid region are various gases gener­
A laser ablation shadow imaging system was set up to observe the ated by the pyrolysis of polymers. An energy source term is added in the

Fig. 1. Laser ablation shadow imaging system.

2
H. Yu et al. Optics and Laser Technology 170 (2024) 110313

solid phase region to simulate the injection of laser energy. Mass source
terms, energy source terms, and specie source terms are added in the
fluid region to simulate the diffusion process of the pyrolysis gases.
During the laser ablation process, ablation craters form on the polymer
surface. This process is simulated using the dynamic mesh function in
FLUENT. All source terms and the dynamic mesh function are defined
and added through the FLUENT User Defined Function (UDF).
Considering the actual situation’s complexity and the numerical
model’s computational efficiency, certain assumptions and simplifica­
tions must be made before establishing the model:
(i)The laser heat source follows a Gaussian distribution.
(ii)All materials are assumed to be homogeneous and isotropic, and
their thermal properties do not change with temperature.
(iii)The various decomposition gases generated by laser ablation and
air are considered compressible ideal gases.
Fig. 2. Schematic diagram of the source term definition in the fluid region.

3.2. Governing equations 3.3. Source term definition

As mentioned earlier, the computational region is divided into fluid The source terms in this article are divided into solid region source
and solid regions. The governing equations for the fluid region are the terms and fluid region source terms.
mass conservation equation, energy conservation equation, and mo­ For the solid region, the energy source term, i.e., the laser heat source
mentum conservation equation. Additionally, species equations are term, is expressed as[17]:
introduced due to the production of various pyrolysis gases. In the solid ( )
region, the energy conservation equation is introduced to address the 2P 2(x2 + y2 )
Sh = (1 − R) • aexp( − az) • 2
• exp − 2
• g(t) (6)
temperature change in the polymer after absorbing laser energy. f πwz τFWHM wz
Solid region energy conservation equation: {
1t ≤ τFWHM
∂ g(t) = (7)
(ρh) = ∇⋅(k∇T) + Sh (1) 0t > τFWHM
∂t
In the equation (1), ρ denotes the density of the material. h signifies In equation (6), R is the reflectivity of the material’s surface, a is the
∫T absorption coefficient of the material for the laser, P is the average laser
the sensible enthalpy of the material, expressed as h = Tref Cp dT. Where
power, f is the repetition frequency of the laser, wz is the radius of the
T represents the temperature of the material and Tref stands for the laser beam spot, and τFWHM is the pulse width of the laser. t is time and ×,
reference temperature, set at Tref = 300 K. k is the conductivity of the y, z are spatial coordinates.
material. Sh is the energy source term within the solid region. In this article, the definition of the source term in the fluid region is
Fluid region mass conservation equation[25]: based on the computational method from the reference [24]. Specif­
∂ρ ically, the first layer of mesh near the wall in the fluid region is treated as
+ ∇⋅(ρν) = Sm (2) the source term zone, and the mass and energy source terms are calcu­
∂t
lated using the real-time temperature from the corresponding solid re­
In equation (2), ρ is the gas density, ν is the velocity vector of the gas,
gion. For clarity, Fig. 2 illustrates a fluid region source term mesh cell
and Sm is the mass source term for the fluid region.
and its corresponding solid region mesh cells. The solid lines represent
Fluid region momentum conservation equation:
the solid region mesh cells, while the dashed lines represent the fluid
∂ region source term zone mesh cells. The source terms of the fluid region
(ρν) + ∇⋅(ρνν) = − ∇p + ∇⋅τ + F (3)
∂t cell are determined by the pyrolytic effects of the corresponding solid
region cells.
In equation (3), p is the static pressure of the gas, τ is the stress tensor
Fluid region mass source term:
of the gas, and F is the momentum source term.
For the solid region cell corresponding to the current source term cell
Fluid region energy conservation equation:
of fluid region, the mass flow rate of the pyrolysis gas is:
( )
∂ ∑ ∫ zv
(ρE) + ∇⋅(ν(ρE + p) ) = ∇⋅ keff ∇T − hi Ji + τ ⋅ν + Sh (4) ṁpol = −
∂ρ
⋅Vdz (8)
∂t i zw ∂t

In equation (4), E represents the total energy per unit mass of the gas. Here, zw and zv represent the starting and ending boundaries of the
keff = kl +kt is the effective thermal conductivity, where kl and kt are the solid region, respectively, and V is the volume of the solid region cell.
laminar and turbulent thermal conductivity, respectively. Ji is the Using the Arrhenius formula to calculate the pyrolysis rate of the
diffusion flux of species i, and hi is the enthalpy of species i. Sh is the polymer material during pyrolysis [24]:
energy source term for the fluid region, representing the heat carried by
∂ρ
the pyrolysis gases. (9)
E0

= − A0 ρ e R0 T
∂t
Species transport equation:
Here, A0 is the pre-exponential factor and E0 is the activation energy.

(ρYi ) + ∇⋅(ρνYi ) = − ∇⋅Ji + Si (5) Combining Equations (8) and (9), and discretizing, we get:
∂t
∑ − E0
In the above equation, Yi represents the mass fraction of species i. Si ṁpol = A0 e R0 Ti
ρi ⋅Vi (10)
is the source term of the species i. Due to the outflow of pyrolysis gases, i

the distribution of the mass fraction of each species in the near-wall grid Here, i is the mesh cell number of the solid region.
cells of the flow field changes, so a species source term needs to be added The final form of the mass source term is:
to the above computational region.

3
H. Yu et al. Optics and Laser Technology 170 (2024) 110313

Fig. 3. Geometry Model and Mesh.

ṁpol
Sm = (11) Table 2
VC
Thermal Properties and Pyrolysis Parameters of Epoxy Resin.
Here, VC represents the volume of the source term zone cell. Parameter Symbol Value
Fluid region energy source term: − 3
Density (kg⋅m ) ρ 1450
The energy source term is the energy generation rate per unit volume
Specific Heat (J⋅kg− 1⋅K− 1) Cp 2500
of the source term cell caused by the diffusion of pyrolysis gas. The Thermal Conductivity (W⋅m− 1⋅K− 1) k 0.3
energy source term generated by species j’s diffusion is: Pre-exponential Factor (s− 1) A0 2.16 × 106
∫ Ti Activation Energy (J⋅mol− 1) E0 8 × 104
∑ E0
Ideal Gas Constant (J⋅mol− 1⋅K− 1) R0 8.314
Shj = [A0 eR0 Ti ρi ⋅Vi ⋅Yj ⋅ Cpj dT]/VC (12) Absorption Coefficient (m− 1) a 8.9 × 104
i Tref
Reflectivity@1064 nm R 0.3
∫ Ti
Here, Yj is the mass fraction of species j in the pyrolysis gas, and Tref

Cpj dT represents the sensible enthalpy of unit mass of species j generated in the z direction in a single time step is:
from the corresponding solid region cell, with the reference temperature ∑ − E0
taken as 300 K. A0 e R0 Ti
ρi ⋅Vi
The final form of the energy source term is: i (16)
Δz = • Δt
∑n ρ • |A|
Sh = j=1
Shj (13)

Fluid region momentum source term: 3.5. Geometry model and mesh
The initial diffusion speed of the pyrolysis gas is 0, so there is no need
to add a momentum source term. The physical model established in this paper is a 2D model with di­
Fluid region species source term: mensions of 2 mm × 1.17 mm, as shown in Fig. 3. The model domain is
The species source term is the mass generation rate of various species divided into a solid region and a fluid region. The solid region is sized at
(CO2, water vapour) in the pyrolysis gas per unit volume: 2 mm × 0.1 mm, while the fluid region measures 2 mm × 1.07 mm. The
first layer of mesh near the wall in the fluid region is set as the source
∑ −
term zone. To utilize the dynamic mesh function in FLUENT, part of the
E0
Sj = A0 e R0 Ti
ρi ⋅Vi ⋅Yj /Vc (14)
i fluid region and the solid region are set with triangular mesh, while
other part are set with quadrilateral mesh. The size of the quadrilateral
mesh is 5 μm × 5 μm.
3.4. Dynamic mesh definition

In order to simulate the material surface retreat caused by laser 3.6. Material properties
ablation, the dynamic mesh function was used in this paper. The ther­
mochemical reaction rate determines the retreat rate ṙ of the material The laser parameters used in both the simulation and the experiment
surface caused by laser ablation: are the same, as shown in Table 1.
The material in the solid region is bisphenol-a epoxy resin. Its ther­
ṁpol
ṙ = (15) mal properties and pyrolysis parameters are shown in Table 2[17,26],
ρ • |A|
where the surface reflectivity is obtained experimentally.
Where ṁpol is the mass flow rate of pyrolysis gas, ρ is the density of Based on previous studies[27], we demonstrated that the primary
the polymer material, and A is the face vector (pointing outside of the gases produced during the pyrolysis of epoxy resin are CO2 and water
mesh cell) of the mesh face that coincides with the wall in the fluid vapour. We only consider these two primary gas components. Hence, the
region. |A| represents the area of the mesh face. Assuming the material fluid region consists of CO2, water vapour, and air, with their thermal
surface moves in the negative z direction, the movement of the grid node properties listed in the Table 3.

4
H. Yu et al. Optics and Laser Technology 170 (2024) 110313

Table 3 exposure, the surface of the epoxy resin exhibits craters. Furthermore, as
Thermal Properties of CO2, Water Vapour, and Air. the laser power escalates, the craters’ diameter and depth increase
CO2 Water Vapour Air progressively. Specifically, for laser powers of 10 W, 20 W, and 30 W,
− 3 the diameters of the craters formed are 63 μm, 89 μm, and 103 μm,
Density (kg⋅m ) Ideal Gas Ideal Gas Ideal Gas
Specific Heat (J⋅kg− 1⋅K− 1) 840 2014 1006 respectively, while the depths are 14 μm, 17 μm, and 18 μm.
Thermal Conductivity 0.0145 0.0261 0.0242 Fig. 5 shows the dynamic shadow images of the laser ablation of the
(W⋅m− 1⋅K− 1) epoxy resin. In this figure, images (a)-(c) illustrate the shadow images of
Viscosity (kg⋅m− 1⋅s− 1) 1.37 × 10− 5
1.34 × 10− 5
1.79 × 10− 5
the material surface post-exposure to a 10 W pulsed laser, with an in­
Molar Mass (kg⋅mol− 1) 44 18 29
terval of 15 μs between each image. For images (d)-(f), the laser power is
20 W, while for (g)-(i), it is 30 W. At the onset of the laser action, plumes
According to reference [28], the volume ratio of CO2 to water vapour and shock waves emerge on the material surface. Both these phenomena
produced during the pyrolysis of epoxy resin is 1:0.65, resulting in mass display a hemispherical shape, and noticeably, the expansion velocity of
fractions of 0.79 and 0.21, respectively. the shockwaves surpasses that of the plumes. Constituting these plumes
are pyrolytic gases like CO2 and water vapour produced due to ablation.
4. Results and discussion At the same time, the shockwaves are generated due to the rapid
expansion of the pyrolytic gases compressing the surrounding air. With
4.1. Experimental results increasing laser power, the radius of the gas plumes also grows. When
the laser power reaches 30 W, the diameter of the plume is approxi­
When the laser power is set to 10 W, 20 W, and 30 W, the surface mately 1 mm. As time progresses, the shadow images undergo alter­
morphology of the laser ablation of the epoxy resin is shown in Fig. 4. All ations. By the second frame, the shockwave has exited the field of view.
these morphologies are the result of a single laser pulse action. Images The gas plume’s expansion in the lateral direction has stopped, and it
(a), (c), and (e) depict the three-dimensional morphology of the material continues to expand upward.
surface after exposure to the 10 W, 20 W, and 30 W lasers, respectively.
Meanwhile, images (b), (d), and (f) show their corresponding two-
dimensional cross-sectional contours. As observed, post-laser

Fig. 4. Surface morphology of epoxy resin after laser ablation (a-b: 10 W, c-d: 20 W, e-f: 30 W).

5
H. Yu et al. Optics and Laser Technology 170 (2024) 110313

Fig. 5. Dynamic shadow images of laser ablation on epoxy resin (a-c: 10 W, d-f: 20 W, g-i: 30 W).

Fig. 6. Temperature contours of laser ablation epoxy resin (a-c: 10 W, d-f: 20 W, g-i: 30 W).

4.2. Simulation results analysis 10 W, and the calculation time of (a), (b), and (c) is 0.05 μs, 1.55 μs, and
3.05 μs, respectively. Initially, at 0.05 μs, the surface of the epoxy resin
4.2.1. Temperature absorbs the laser energy, and the temperature rapidly rises. By 1.55 μs,
Fig. 6 presents the simulated temperature contours of the laser the epoxy resin undergoes thermal decomposition. This causes the sur­
ablation epoxy resin using Fluent. In Fig. 6 (a)-(c), the laser power is face to develop a depression, and a hemispherical high-temperature gas

6
H. Yu et al. Optics and Laser Technology 170 (2024) 110313

Fig. 7. Maximum temperature variation on fluid and solid region (a: fluid region, b: solid region).

Fig. 8. Simulated morphology of ablation craters on the epoxy resin surface (a:10 W, b:20 W, c:30 W).

plume forms in the nearby fluid region. The gas plume is ejected at high peak temperature of 1848℃, while the solid region records a peak
speed, which violently pushes the surrounding stationary air. This action temperature of 1979℃.
creates a series of compression waves that superimpose to form a shock
wave. This shockwave propagates significantly faster than the plume’s 4.2.2. Ablation morphology of epoxy resin
expansion. By 3.05 μs, the shockwave has disappeared from the contour. Fig. 8 shows the ablation morphology on the epoxy resin surface
When compared to 1.55 μs, neither the gas plume nor the resin’s crater obtained through the numerical calculation at 3.05 μs. As the laser
shows noticeable volume changes. This suggests the end of pyrolysis. power increases, the diameter and depth of the ablation craters increase.
Notably, the base of the plume begins to retract, and the gas plume When the laser power is 10 W, 20 W, and 30 W, the diameters of the
shows an upward trend. The sequences for 20 W laser power are ablation craters are 78 μm, 98 μm, and 108 μm, respectively, and their
depicted in figures (d)-(f). For 30 W, they are shown in figures (g)-(i). depths are 12 μm, 16 μm, and 19 μm, respectively. Compared to the
Both showcase a pyrolytic process similar to the 10 W power setting. experimental results, the diameters and depths of the ablation craters
With increasing laser power, the volume of the gas plume formed on the obtained from numerical calculations are slightly off, but they reflect the
resin’s surface enlarges progressively. From figures (c), (f), and (i), it can influence of power on the results. Discrepancy between simulation and
be observed that the diameter of the high-temperature gas plume ob­ experiment mainly arises from the variation in the material’s thermal
tained from the simulation when it stops lateral expanding is slightly properties and the material’s inhomogeneity. During the actual laser
smaller than the experimental results (taken from the first frame of the ablation process, as the temperature of the material continuously in­
high-speed shadow image). This discrepancy is mainly due to the vari­ creases, its thermal properties will change. The material’s absorption
ation in the material’s thermal properties. In the actual ablation process, coefficient for the laser will also vary as the temperature rises. The in­
both the properties of the epoxy resin and the pyrolysis gases change homogeneity of the material can result in uneven edges of the ablation
with the increase in temperature. craters. Moreover, the laser ablation process of epoxy resin involves
Fig. 7 illustrates the peak temperature variations within the fluid and many complex chemical reactions. Specifically, during the pyrolysis
solid regions. The plot reveals that temperatures in both regions rapidly process, residual carbon is produced, which reacts with the air to form
rise to a peak value and then slowly declining. Additionally, the peak CO2. The small amounts of organic gases generated by pyrolysis will also
temperature in the solid region is higher than in the fluid region. As the react with the air, producing CO2 and water vapor. In the simulation, for
laser power escalates, the peak temperatures in both regions increase simplification, only the final products of the reactions were considered.
correspondingly. With a 30 W laser power, the fluid region reaches a From the calculated results of temperature and surface morphology,

7
H. Yu et al. Optics and Laser Technology 170 (2024) 110313

Fig. 9. Pressure contour of laser ablation epoxy resin (a-c: 10 W, d-f: 20 W, g-i: 30 W).

the ablation crater throughout the laser ablation process. Fig. 10 rep­
resents the maximum pressure change in the fluid region during laser
ablation. After the laser action, the pressure rises rapidly, reaching its
peak at around 150 ns and swiftly dropping. As the laser power in­
creases, the peak pressure also grows. When the laser power is 30 W, the
peak pressure reaches 110Mpa.

4.2.4. Velocity
Fig. 11 showcases the velocity vector diagram for epoxy resin’s laser
ablation, derived from numerical simulations. It’s essential to under­
stand that the velocities in this diagram capture only the instantaneous
speeds at given locations, rather than the genuine expansion speeds of
both shock waves and gas plumes. To determine the expansion speeds,
one must compute based on the expansion distance and time.
In contrast to prior temperature and pressure contours, the data re­
veals a 1 mm shockwave expansion distance for a 30 W laser at 1.55 μs.
Simultaneously, the lateral expansion distance for the high-temperature
gas plume is roughly 0.5 mm. This data translates to an average
expansion speed of 645 m/s for the shock wave and 333 m/s for the gas
plume.
The vector diagram further highlights a noticeable vortex on both
Fig. 10. Variation of the maximum pressure in the fluid region. sides at the base of the pyrolysis gas plume. The formation of the vortex
structure is primarily due to the fluid’s viscosity and the frictional
it can be known that this numerical calculation model can accurately resistance between the pyrolytic gas and the wall surface. When a
reflect the dynamic behaviour of pulsed lasers acting on epoxy resin, viscous fluid flows over a wall, its speed decreases as it approaches the
including the formation of ablation craters on the epoxy surface, the wall, due to the combined effects of viscosity and frictional resistance.
generation of pyrolysis gas plumes, and the formation of shock waves. The velocity becomes completely zero at the interface between the fluid
and the wall. This results in a pronounced velocity gradient in the fluid,
4.2.3. Pressure perpendicular to the wall. During the flow of the high-temperature gas
Fig. 9 presents the simulated pressure contour of laser ablation epoxy plume along the wall, this velocity gradient causes the upper and lower
resin. From Fig. 9(b), (e), and (h), it can be observed that after the sections of the plume to gradually separate, forming vortices. The gas
shockwave passes, the pressure suddenly rises and then gradually de­ plume’s instantaneous velocity peak resides close to the ablation crater.
clines. A negative pressure appears at a certain distance after the Meanwhile, Fig. 12 captures the fluid region’s maximum velocity
shockwave, which is similar to the propagation law of explosion shock change. After laser action, the velocity of the fluid region increases
waves in the air [29]. The highest pressure point always remains near rapidly and then decreases gradually. At 30 W laser power, this peak

8
H. Yu et al. Optics and Laser Technology 170 (2024) 110313

Fig. 11. Laser ablation velocity vector diagram of epoxy resin (a-c: 10 W, d-f: 20 W, g-i: 30 W).

5. Conclusions

In this study, a kinetic model for laser ablation of polymers was


established based on the pyrolysis effect of polymers. The model was
solved using numerical computation software FLUENT, obtaining values
for temperature, pressure, velocity, pyrolysis gas concentration, and the
surface morphology of the material. The surface morphology of the
material and shadow images of the laser ablation process was obtained
through experimentation, which was used to validate the numerical
model. The main conclusions are as follows:
(1) The calculation results for the ablation morphology of epoxy
resin show that as the laser power increases, the diameter and depth of
the ablation crater increase. When the laser power is 10 W, 20 W, and
30 W, the diameters of the ablation craters are 78 μm, 98 μm, and
108 μm, respectively, and the depths are 12 μm, 16 μm, and 19 μm,
respectively. Compared with the experimental results, the numerical
calculation effectively reflects the influence of laser power.
(2) The temperature and pressure of the fluid and solid regions
during the laser ablation process were calculated. The temperature and
pressure contours intuitively reflect the generation of high-temperature
Fig. 12. Maximum velocity variation in the fluid region. pyrolysis gases and the propagation of shock waves. As the laser power
increases, the temperature and pressure in the fluid and solid regions
also increase. For a point in the fluid region, after the shockwave passes,
velocity hits 1780 m/s.
the pressure rises sharply and then gradually decreases until a negative
pressure is produced. The propagation rule of the shock wave generated
4.2.5. CO2 concentration
by ablation is similar to that of the explosion shock wave in the air.
The CO2 molar concentration contour is presented to showcase the
(3) The velocity variation of the fluid region during the laser ablation
variation in gas concentration during laser ablation epoxy resin, as
process was calculated. When the laser power is 30 W, the average
shown in Fig. 13. The figure shows that throughout the ablation process,
expansion speed of the shock wave is 645 m/s, while the average
the molar concentration of CO2 gas near the ablation crater is the highest
expansion speed of the high-temperature gas plume is 333 m/s. The
and decays sharply as it diffuses into the air. Fig. 14 displays the
plume gradually separates from the solid wall during its motion. Due the
maximum molar concentration of CO2 in the fluid region during laser
fluid’s viscosity and the frictional resistance between the pyrolytic gas
ablation. The CO2 molar concentration increases with the increasing
and the wall surface, a distinct vortex structure is formed on both sides
incident laser power. When the laser power is 30 W, the peak molar
of the bottom of the gas plume.
concentration of CO2 is 5120 mol⋅m− 3.

9
H. Yu et al. Optics and Laser Technology 170 (2024) 110313

Fig. 13. Molar concentration contour of CO2 during laser ablation (a-c: 10 W, d-f: 20 W, g-i: 30 W).

interests or personal relationships that could have appeared to influence


the work reported in this paper.

Data availability

Data will be made available on request.

Acknowledgements

This work was supported by the National Key R&D Program of China
(2022YFB4601500) and National Natural Science Foundation of China
(U2033211).

References

[1] H.Q. Dou, C.Z. Yao, H. Liu, Y. Wan, R.J. Ding, X.D. Yuan, S.Z. Xu, Femtosecond
laser ablation of Al-Mg alloy in vacuum and air, Appl. Surf. Sci 447 (2018)
388–392.
[2] B. Gakovic, G.D. Tsibidis, E. Skoulas, S.M. Petrovic, B. Vasic, E. Stratakis, Partial
ablation of Ti/Al nano-layer thin film by single femtosecond laser pulse, J. Appl.
Phys 122 (22) (2017).
[3] E. Stratakis, V. Zorba, M. Barberoglou, C. Fotakis, G.A. Shafeev, Femtosecond laser
writing of nanostructures on bulk Al via its ablation in air and liquids, Appl. Surf.
Fig. 14. Maximum molar concentration variation of CO2 in the fluid region. Sci 255 (10) (2009) 5346–5350.
[4] F. Costache, S. Kouteva-Arguirova, J. Reif, Sub-damage-threshold femtosecond
laser ablation from crystalline Si: surface nanostructures and phase transformation,
CRediT authorship contribution statement Appl. Phys. a-Mater 79 (4–6) (2004) 1429–1432.
[5] A. Borowiec, M. MacKenzie, G.C. Weatherly, H.K. Haugen, Femtosecond laser pulse
Haijun Yu: Methodology, Investigation, Validation, Formal analysis, ablation of GaAs and InP: studies utilizing scanning and transmission electron
microscopy, Appl. Phys. a-Mater 77 (3–4) (2003) 411–417.
Writing – original draft. Zhiyan Zhang: Conceptualization, Writing –
[6] S.Z. Xu, C.Z. Yao, H.Q. Dou, W. Liao, X.Y. Li, R.J. Ding, L.J. Zhang, H. Liu, X.
review & editing. Shusen Zhao: Conceptualization, Writing – review & D. Yuan, X.T. Zu, An investigation on 800 nm femtosecond laser ablation of K9
editing. Hongyang Wang: Visualization. Han Liang: Visualization. glass in air and vacuum, Appl. Surf. Sci 406 (2017) 91–98.
Zhiyong Dong: Data curation. Xuechun Lin: Funding acquisition, Su­ [7] J. Kruger, W. Kautek, M. Lenzner, S. Sartania, C. Spielmann, F. Krausz, Laser
micromachining of barium aluminium borosilicate glass with pulse durations
pervision, Project administration. between 20 fs and 3 ps, Appl. Surf. Sci 127 (1998) 892–898.
[8] S.H. Kim, I.B. Sohn, S. Jeong, Ablation characteristics of aluminum oxide and
nitride ceramics during femtosecond laser micromachining, Appl. Surf. Sci 255
Declaration of Competing Interest (24) (2009) 9717–9720.

The authors declare that they have no known competing financial

10
H. Yu et al. Optics and Laser Technology 170 (2024) 110313

[9] Y.Q. Yang, S.F. Wang, Z.Y. Sun, D.D. Dlott, Near-infrared laser ablation of poly [19] E. Kandare, B.K. Kandola, J.E.J. Staggs, Global kinetics of thermal degradation of
tetrafluoroethylene (Teflon) sensitized by nanoenergetic materials, Appl. Phys. Lett flame-retarded epoxy resin formulations, Polym. Degrad. Stabil 92 (10) (2007)
85 (9) (2004) 1493–1495. 1778–1787.
[10] B. Majaron, P. Plestenjak, M. Lukac, Thermo-mechanical laser ablation of soft [20] N.C. Herr, A.E. Gonzales, G.P. Perram, Kinetics, evolving thermal properties, and
biological tissue: modeling the micro-explosions, Appl. Phys. B-Lasers. O 69 (1) surface ignition of carbon fiber reinforced epoxy composite during laser-induced
(1999) 71–80. decomposition, Polym. Degrad. Stabil 152 (2018) 147–161.
[11] M.E. Shaheen, J.E. Gagnon, B.J. Fryer, Studies on laser ablation of silicon using [21] S. Ravi-Kumar, B. Lies, H. Lyu, H.T. Qin, Laser Ablation of Polymers: A Review,
near IR picosecond and deep UV nanosecond lasers, Opt. Laser. Eng 119 (2019) Procedia. Manuf 34 (2019) 316–327.
18–25. [22] H.C. Yu, H.Y. Li, Y. Wang, L.G. Cui, S.Q. Liu, J. Yang, Brief review on pulse laser
[12] N. Zhang, X.N. Zhu, J.J. Yang, X.L. Wang, M.W. Wang, Time-resolved propulsion, Opt. Laser. Technol 100 (2018) 57–74.
shadowgraphs of material ejection in intense femtosecond laser ablation of [23] D.S. Gill, A.A. Anderson, R.W. Eason, T.J. Warburton, D.P. Shepherd, Laser
aluminum, Phys. Rev. Lett 99 (16) (2007). operation of an Nd:Gd3Ga5O12 thin-film optical waveguide fabricated by pulsed
[13] S.L. Johnson, D.M. Bubb, R.F. Haglund, Phase explosion and recoil-induced laser deposition, Appl. Phys. Lett 69 (1) (1996) 10–12.
ejection in resonant-infrared laser ablation of polystyrene, Appl. Phys. a-Mater 96 [24] H. Xue, X. Chen, C. Zhou, Numerical research of thermochemical ablation about
(3) (2009) 627–635. carbon-phenolic jet vane in solid rocket motors, J. Propul. Technol. 37 (10) (2016)
[14] J.F. Ye, X. Jin, H. Chang, Ns-shadowgraphy time resolved plume generation and 1900–1908.
expansion in the laser micro ablation, 2nd International Symposium on Laser [25] ANSYS, Inc., Ansys Fluent Theory Guide, 2022 R2 ed., Canonsburg, 2022.
Interaction with Matter (Limis 2012) 8796 (2013). [26] E.D. McCarthy, B.K. Kandola, G. Edwards, P. Myler, J.F. Yuan, Y.C. Wang,
[15] T. Moscicki, J. Hoffman, Z. Szymanski, Modelling of plasma formation during E. Kandare, Modelling flaming combustion in glass fibre-reinforced composite
nanosecond laser ablation, Arch. Mech 63 (2) (2011) 99–116. laminates, J. Compos. Mater 47 (19) (2013) 2371–2384.
[16] J.N. Yang, J.Z. Zhou, Q. Sun, X.K. Meng, Z.H. Guo, M. Zhu, Digital analysis and [27] H.J. Yu, Z.Y. Zhang, H.Y. Wang, S.S. Zhao, Z.Y. Dong, C. Zou, K.X. Cao, X.C. Lin,
prediction of the topography after pulsed laser paint stripping, J. Manuf. Process Investigation on dynamic behavior of laser paint removal based on high-speed
62 (2021) 685–694. shadow imaging technology, Opt. Laser. Technol 163 (2023).
[17] H. Yang, H.X. Liu, R.X. Gao, X. Liu, X. Yu, F. Song, L.S. Liu, Numerical simulation of [28] M.T. McGurn, P.E. DesJardin, A.B. Dodd, Numerical simulation of expansion and
paint stripping on CFRP by pulsed laser, Opt. Laser. Technol 145 (2022). charring of carbon-epoxy laminates in fire environments, Int. J. Heat. Mass. Tran
[18] R.D. Chippendale, I.O. Golosnoy, P.L. Lewin, Numerical modelling of thermal 55 (1–3) (2012) 272–281.
decomposition processes and associated damage in carbon fibre composites, [29] C.Q. Wu, H. Hao, Modeling of simultaneous ground shock and airblast pressure on
J. Phys. D-Appl. Phys. 47 (38) (2014). nearby structures from surface explosions, Int. J. Impact. Eng. 31 (6) (2005)
699–717.

11

You might also like