Open Access
6 September 2022 Multiscale model for ultrashort pulsed parallel laser structuring—part II. The macroscale model
Author Affiliations +
Abstract

The increasing pixel density in displays demands high quality in the production of fine metal masks (FMMs). The production process of FMMs boils down to structuring tiny holes in thin metal sheets or foils. The manufacturing requirements of FMMs are high precision in terms of the hole geometry to let enough light escape from each diode and high productivity to produce the required amount. To achieve both objectives, high power ultrashort pulse (USP) lasers can be utilized. Because USP lasers fall short of the productivity requirements, they are combined with multibeam scanners. During production, the multibeam scanners deposit a lot of heat in the metal foil, which can ultimately yield temperature-induced distortions. To understand and finally avoid such distortions, a process simulation is sought. In a preceding study, the structuring of a single hole (the microscale) was investigated, but due to the large differences in the time and spatial scales involved, it was not feasible to simulate the production of the whole part (the macroscale). Within this treatise, a multiscale approach that takes into account the necessary information from the microscale to describe temperature-induced distortions on the macroscale is described. This approach targets laser ablation processes with pulse durations ranging from picoseconds up to nanoseconds provided the ablation is not melt-driven. First, a representative volume element (RVE) is generated from the results of the microscale model. Then, this RVE is utilized in the thermo-elastic structural mechanics simulation on the macroscale. The multiscale model is validated numerically against a hole-resolved computation, which shows good agreement. Naturally, the simulation is highly dependent on the microscale model, which in turn depends on the material properties. To handle material changes well, an experimental calibration has to be performed. This calibration is not part of this treatise, but it will be described in a future publication. In addition to the calibration process, the validation with experiments will be conducted in future research. Additionally, the authors envision the automation of the whole process, resulting in a first-time-right approach for the development of FMMs. Finally, the procedure might be extended to the requirements of other filtration purposes.

1.

Introduction

In today’s consumer electronics, displays with ever-increasing pixel density are required. Within the production of such displays, fine metal masks (FMMs) with the highest quality have to be fabricated. Conventionally, the holes are etched into the metal foil; however, this process is reaching its limit as the defects accumulate for pixel densities around 600 PPI.1 A technology that allows for such quality is high power ultrashort pulsed lasers.2 The main reason for resorting to USP lasers is their ability to insert a lot of energy into the material without too much heating, which would destroy such thin foils.2 To meet the productivity requirements, the laser can be combined with a multibeam scanner.3 During the production of FMMs, it is important to avoid distortions or discoloration due to heat accumulation.4 Therefore, process parameters and processing strategies that avoid such problems are sought.

Within this treatise, an extension to a simulation that computes the structuring of a single borehole, which has been explained in a prior publication,5 is described. This time-consuming computation of a single borehole shape is used as the basis for a multiscale model with which the simulation of larger workpieces becomes feasible. The multiscale model consists of two scales.610 First, the heat conduction and deformations are computed for a single borehole, the microscale. Second, the information on the microscale is collected in a representative volume element (RVE) with which the heat conduction and deformations of a whole workpiece, the macroscale, can be computed. The microscale model assumes pulse durations from femtoseconds up to nanoseconds as long as the ablation is not melt-driven.11

The unit cell, which is the computational domain in the microscale simulation, is the direct result of the simulation described in the prior publication.5 Both the heat conduction and the thermo-elastic structural mechanics problems are discretized using a Bubnov–Galerkin method.1214 Averaging the results over the unit cell domain yields the material properties of the RVE.

On the macroscale, the computational domain is an FMM. It consists of two regions: one for a hole and one for the solid material. It is essential to use the RVE in the region that accounts for a hole and the ordinary material properties of the solid everywhere else. Again, a Bubnov–Galerkin scheme is employed to compute the temperature field and the distortions of the FMM.1214

This treatise presents the multiscale approach using a simplified hole geometry. The homogenization of both models, the heat-conduction and the thermo-elasticity models, is performed using an asymptotic expansion in the scale variable as described in Ref. 6. To account for multibeam scanners, the two-scale model can be adapted easily. This novel approach is capable of simulating the temperature field and elastic distortions occurring during the structuring of thin metal foils. Plastic strains and melt formation are not considered. The multiscale simulation computes accurate results in a fraction of the time compared with a fully-resolved simulation.

2.

Multiscale Model

In a prior publication, the ablation of a structure with a single beam was explained.5 This treatise extends the results to a multiscale model with which a periodic ablation pattern can be simulated for larger domains in a reasonable time.

The multiscale approach follows closely the works of Fish [8, Chapter I.3] or [6, Chapter 2.2]. To further reduce the computational demands, a residual-free method that allows for computing the microstructure off-line and in advance of the macroscale simulation was developed.

2.1.

Geometric Set-Up and Notation

Hereafter, it is assumed that two length scales that differ in size enough to be separated exist, i.e., 0<ζfc1. Variables referring to a particular length scale are denoted with a subscript xf for fine and xc for coarse. Furthermore, the notation for functions that account for information on all scales uses a preceding superscript fζ and functions that are valid on the I’th scale use fI. If not indicated otherwise, the variable x lives on the coarse scale, whereas the variable yx/ζ lives on the fine scale. The computational domain is denoted as Ω with the boundary Ω. To account for different physical interactions with the environment, it is necessary to split this boundary into a Neumann part ΓN and a Dirichlet part ΓD, such that Ω=ΓNΓD and ΓNΓD=0 (c.f. Fig. 1). Analogously, the domain and boundaries for the unit cell are defined, except the symbol Θ is used. Whenever a distinction between Dirichlet and Neumann Boundary has to be made, ΓD and ΓN, respectively, are used.

Fig. 1

(a) An artificial unit cell geometry is shown. In reality, the hole shape depends on the ablation strategy and the laser beam properties. (b) Displays a foil created using the unit cell in (a), although the hole sizes are emphasized.

OE_61_9_095103_f001.png

The reason the artificial unit cell geometry [c.f. Fig. 1(a)] is used throughout this treatise is twofold. First, the computation of a realistic hole is already described in a prior publication,5 and the geometries can be easily exchanged. Second, on the large scale, the difference between the artificial unit cell and a realistic unit cell is negligible.

In addition to the geometric notation, the Einstein summation convention is employed, i.e., repeated indices are summed. The derivative of a function f(xi) with respect to xi is sometimes abbreviated with xif to emphasize vector field operators such as the divergence. Averaged quantities are denoted with an overline, e.g., x¯, vectors and matrices are highlighted in a boldface x and uppercase boldface X, respectively. Finally, the subscript f(i,xj) is used to denote the symmetrized gradient tensor entry, i.e.

Eq. (1)

f(i,xj)12(fixj+fjxi).

More complex geometries can be created using a level set function Φ(x):RnR. Such a level set function is the key result in the companion publication5 wherein Φ is the solution that describes the ablation surface of a laser structuring process. A unit cell domain can be defined using this function on voxels by removing all elements that lie completely inside the ablated domain, i.e., for which Φ(x)<0 holds. This can be seen in Fig. 2. The second result of the simulation described in Ref. 5 is the heat distribution after the ablation process is done. This heat distribution can be used as a volume source in Sec. 2.2.

Fig. 2

A level set function can be used to mark elements that lie inside an ablated domain [c.f. Fig. 2(a)]. The final unit cell after removal of marked elements is shown in Fig. 2(b).

OE_61_9_095103_f002.png

2.2.

Two-Scale Heat Conduction

The heat conduction equation for a material with specific density ρ:ΩζR, specific heat capacity cp:ΩζR, and thermal conductivity λ:ΩζR heated by a heat source Q:ΩζR reads

Eq. (2)

ρcpTζt+qiζxi=Qζin  Ωζqiζ=λTζxiin  Ωζqiζniζ=qNζon  ΓNζTζ=TDζon  ΓDζ.
Herein, ni, refers to the i’th entry of the outward pointing unit normal n, e.g. n=(n1,n2,n3)T in three dimensions. The solution to this problem is the temperature field T:ΩζR. In general, the functions depend on both the coarse and fine scales, but for the density and the specific heat capacity a dependence on the fine scale variable only is assumed, i.e., ρρ(y) and cpcp(y).

Now, an ansatz function for the temperature is defined by means of an asymptotic expansion15 using the scale ζ, which reads

Eq. (3)

Tζ(x)T(x,y)=ζ0T0(x,y)+ζ1T1(x,y)+O(ζ2)+.
Note that terms of order 2 and above have been omitted as they do not contribute significantly to the temperature due to the condition ζ1. Fourier’s law in Eq. (2) relates the temperature gradient to the heat flux; therefore, the spatial derivative of Eq. (3) is taken. Rearranging the terms in accordance to the order of ζ yields

Eq. (4)

ζTxi(x)=Txi(x,y)=ζ10T(x,y)+ζ0(0Txi(x,y)+1Tyi(x,y))+ζ11Txi(x,y)+O(ζ2)+.
Again, higher-order terms are omitted in the two-scale formulation. Multiplying this equation with the negative thermal conductivity and a comparison of the coefficients results in

Eq. (5)

λTxi(x,y)=ζ1λT0(x,y)ζ0λ(T0xi(x,y)+T1yi(x,y))ζ1λ1Txi(x,y)O(ζ2)qi(x)=ζ1q1i(x,y)+ζ0q0i(x,y)+ζ1q1i(x,y)+O(ζ2)+.
The divergence of the heat flux is needed, too. Taking the derivative with respect to the i’th space component and sorting in ascending orders of ζ yields

Eq. (6)

qixi=ζ21qiyi+ζ1(1qixi+0qiyi)+ζ0(0qixi+1qiyi)+ζ1(1qixi+2qiyi)+O(ζ2)+.
Here, the functions’ arguments are dropped for brevity. Inserting the ansatz functions—Eq. (3) for the temperature and Eq. (5) for the heat flux—in the energy balance in Eq. (2) gives

Eq. (7)

ρcp(T0+ζT1)t+ζ21qiyi+ζ1(1qixi+0qiyi)+ζ0(0qixi+1qiyi)+O(ζ)+=Qζ.
Note the truncation at O(ζ). The terms are now rearranged and collected in orders of ζ. More rigorously, Eq. (7) is multiplied by ζ2 first, then by ζ1, which, after taking the limit ζ0+, yields a system of three equations:

Eq. (8)

O(ζ2):  1qiyi=0,

Eq. (9)

O(ζ1):  1qixi+0qiyi=0,

Eq. (10)

O(ζ0):  ρcp(T0+ζT1)t+0qixi+1qiyi=Qζ.
Henceforth, it is assumed that the heat source Qζ and the boundary conditions qNζ and TDζ in Eq. (2) depend on the coarse scale only. The system still needs closure.

Multiplying Eq. (8) with T0 and integrating over the unit cell gives

Eq. (11)

ΘT01qiyidy=0.
which, after applying the product rule of the divergence operator and the divergence theorem, reads

Eq. (12)

ΘT10qinidyΘ0Tyiqi1dy=0.
Because periodic boundary conditions are employed on the unit cell, the temperature and heat flux on opposite points of the boundary are equal, but the normal points in exactly opposite directions, and therefore, the boundary integral vanishes. Inserting the corresponding coefficient from the ansatz in Fourier’s law Eq. (5) gives

Eq. (13)

Θ0Tyiλ0Tyidy=0.
From λ>0 follows 0Tyi=0, which implies an independence on the fine scale, i.e., T0T0(x).

Through Eq. (5), this also holds for the corresponding part of the heat flux

Eq. (14)

qi1=λT0yi=0.

Now, Eq. (9) is considered. Inserting Eq. (14) results in

Eq. (15)

0qiyi=0.
Again, the expression for the corresponding part of the heat flux from Eq. (5) can be inserted, which yields

Eq. (16)

yi(λ(0Txi+1Tyi))=0.
Using the ansatz T1(x,y)=Hj(y)xjT0(x) and factoring out xjT0(x) gives

Eq. (17)

yi(λ(δij+Hj(y)yi))=0,
with the temperature influence function Hj:ΩζRC0(Θ) and the usual Dirac delta δij. Boundary conditions depend on the coarse scale only, and the fact that T0 already accounts for all information on the boundary, the fine scale problem reads

Find Hj(y), such that

Eq. (18)

yi(λ(δij+Hj(y)yi))=0,in  ΘHj(y)=0on  Θ.
With the current information, the temperature ansatz Eq. (3) is given as

Eq. (19)

Tζ(x)=T(x,y)T0(x)+ζ1Hj(y)0T(x)xj.
On the coarse scale, the temperature is defined to be the multiscale temperature averaged over the unit cell:

Eq. (20)

Tc(x)=1vol(Θ)ΘTζ(x)dy1vol(Θ)ΘT0(x)+ζ1Hj(y)0T(x)xjdy=T0(x)vol(Θ)Θdy+ζ1xjT0(x)vol(Θ)ΘHj(y)dy,
and thus, Tc=T0, if and only if ΘHj(y)dy=0.

Inserting the corresponding relation of the heat flux ansatz Eq. (5) into Eq. (17) yields

Eq. (21)

qi0=Λij0TxjΛijλ(δij+Hjyi),
in which Λij is termed the heat flux influence function.

Finally, the zeroth-order terms of ζ Eq. (10) are averaged over the unit cell, resulting in

Eq. (22)

1vol(Θ)Θρcp0Tt+0qixidy+1vol(Θ)Θ1qiyidy1vol(Θ)ΘQζdy=0.
Rearranging the first integral and applying the divergence theorem to the second gives

Eq. (23)

0Tt1vol(Θ)Θρcpdy+xi(1vol(Θ)Θqi0dy)+1vol(Θ)Θqi1nidy1vol(Θ)ΘQζdy=0,
wherein the boundary integral vanishes due to qi1=λxi T1=λHjxj T0 and Hj=0 for yΘ, resulting in

Eq. (24)

0Tt1vol(Θ)Θρcpdy+xi(1vol(Θ)Θqi0dy)1vol(Θ)ΘQζdy=0.
Henceforth, averaged quantities are denoted with a horizontal line above them, e.g., ρcp¯1vol(Θ)Θρcpdy. With this notation in place, the coarse scale heat conduction reads

Eq. (25)

0Ttρcp¯+q0i¯xiQζ¯=0.

In summary, two problems have to be solved: one on the fine scale and one on the coarse scale. The fine scale problem reads

Find Hj(y), such that

Eq. (26)

yi(λ(δij+Hjyi))=0,in  ΘHj=0,on  Θ.

From the solution, the heat flux influence function Λij can be computed using Eq. (21). The coarse scale problem is then defined as

Find T0(x), such that

Eq. (27)

ρcp¯0Ttxi(Λij¯0Txj)=Q¯,in  Ω×[0,tmax]T0=T¯0,in  Ω×{0}Λij¯0Txjni=qN,on  Ω.
The overall temperature function Tζ can then be constructed as

Eq. (28)

Tζ(x,y)T0+ζHj(y)0T(x)xj.

2.3.

Thermo-Elastic Deformations

To define the multiscale structural-mechanics equations, which describe the deformation of a solid under volume and thermal loading conditions, a few functions have to be introduced. For an n-dimensional domain, the displacement is denoted u:ΩζRn. Note that it does not depend on time. The body forces acting on the work piece are written as b:ΩζRn. The notations for strains and stresses resulting from the loads read ϵ:ΩζRn×n and σ:ΩζRn×n, respectively. They are related through the stiffness tensor CRn×n×n×n. Using these definitions, the static equilibrium of stresses in index notation is given as

Eq. (29)

σijζxj+biζ=0,in  Ωζσijζ=Cijklζϵklelζ,in  Ωζσijζnj=t¯iζ,on  ΓNζuiζ=u¯iζ,on  ΓDζ.
Herein, the total ϵtot strain is the sum of the elastic ϵel, plastic ϵpl, and thermal ϵth strains:

Eq. (30)

ϵkltotζ=ϵklelζ+ϵklthζ,in  Ωζϵkltotζ=12(ukζxl+ulζxk),in  Ωζϵklthζ=αΔTζδkl,in  Ωζ.
The stiffness tensor obeys a symmetry condition

Eq. (31)

Cijklζ=Cjiklζ=Cijlkζ=Cklijζ,
and is positive in the sense that

Eq. (32)

  c>0:  Cijklζnijnklcnijnij,  nij=nji.

Analogously to the two-scale heat conduction in Sec. 2.2, the asymptotic ansatz for the displacement is defined to be

Eq. (33)

uiζ(x)ui(x,y)=ui0(x,y)+ζ1ui(x,y)+ζ2u2i(x,y)+O(ζ3)+.
From Eq. (29), it is obvious that the derivative is required, too. Taking the derivative with respect to xj, applying the chain rule, and sorting in ascending order of ζ yields

Eq. (34)

uixj(x,y)=ζ10uiyj(x,y)+ζ0(0uixj(x,y)+1uiyj(x,y))+ζ1(1uixj(x,y)+2uiyj(x,y))+ζ2(2uixj(x,y)+3uiyj(x,y))+O(ζ3)+=ζ10uiyj(x,y)+s=02ζs(suixj(x,y)+s+1uiyj(x,y))+O(ζ3)+.
Inserting into the strain–displacement relation in Eq. (30) gives

Eq. (35)

ϵijel(x,y)=ζ112(0uiyj+0ujyi)+s=02ζs12(suixj+suixj+s+1uiyj+s+1uiyj)+O(ζ3)+=ζ11ϵijel+s=02ζsϵijels+O(ζ3)+.
Assuming the stiffness tensor is independent of the coarse scale, i.e., Cijkl(x,y)Cijkl(y), and inserting the ansatz for the elastic strain Eq. (35) into the stress–strain relation Eq. (29) results in the two-scale ansatz for the stress

Eq. (36)

σij(x,y)=Cijkl(y)(ζ11εijel+s=02ζsεijels)+O(ζ3)+=ζ11σij+s=02ζsσijs+O(ζ3)+.
Again, the derivative of the stress with respect to space is needed as it appears in Eq. (29). Thus, inserting the derivative, which reads

Eq. (37)

σijxj(x,y)=ζ21σijyj+ζ1(1σijxj+0σijyj)+ζ0(0σijxj+1σijyj)+O(ζ),
in the stress balance and sorting in ascending order of ζ yields

Eq. (38)

ζ21σijyj+ζ1(1σijxj+0σijyj)+ζ0(0σijxj+1σijyj+biζ)+O(ζ)=0.
Analogously to the derivation of the two-scale heat conduction problem, the leading order terms of the stress balance are obtained by multiplication with ζ2 and ζ, respectively, and taking the limit ζ0:

Eq. (39)

O(ζ2):  1σijyj=0,

Eq. (40)

O(ζ1):  1σijxj+0σijyj=0,

Eq. (41)

O(ζ0):  0σijxj+1σijyj+biζ=0.
It is, hereafter, assumed that the body force depends on the fine scale only biζ(x)bi(y), and the boundary conditions in Eq. (29) are coarse scale functions t¯iζ(x)t¯i(x) and u¯iζ(x)u¯i(x). For closure, the moments are computed for each order. Therefore, multiplying Eq. (39) by ui0 and integrating over the unit cell gives

Eq. (42)

Θui01σijyjdy=0.
Using the product rule of divergence and the divergence theorem, as well as noting that the resulting boundary integral vanishes due to periodicity of both integrands and opposing normals, results in

Eq. (43)

Θ0uiyjσij1dy=0.
Inserting the definitions of σ1ij and ϵ1ijel from Eqs. (36) and (35), respectively, yields

Eq. (44)

Θ0uiyjCijkl12(0ukxl+0ulxk)dy=0.
From the symmetry and positiveness of the stiffness tensor C, it can be inferred that

Eq. (45)

0uixj=0,
and, with Eqs. (36) and (35),

Eq. (46)

σij1=0.

Inserting Eq. (46) into Eq. (40) gives

Eq. (47)

σ0ijyj=0,
which, with the definition of σij0 from Eq. (36) and ϵij0 from Eq. (35), results in

Eq. (48)

yj(Cijkl(12(0ukxl+0ulxk)+12(1ukxl+1ulxk)))=0.
This shows a direct dependence between the displacement of the zeroth and first scale. Hence, the separation of variables ansatz

Eq. (49)

u1iHimn(y)12(0umxn+0unxm),
with the first-order displacement influence function H, is employed. It is assumed that the influence function is symmetric Himn=Himn, locally periodic, and continuous HC0(Ω). Inserting the ansatz in Eq. (48) yields

Eq. (50)

yj(Cijkl(u(k,xl)0+H(k,yl)0mnu(m,xn)))=0,
and after factoring out u(m,xn)0

Eq. (51)

yj(Cijkl(Iklmn+H(k,yl)mn))u(m,xn)0=0,
with the definition Iklmn12(δkmδln+δlmδkn). Because u0(m,xn) is arbitrary, it is required that

Eq. (52)

yj(Cijkl(Iklmn+H(k,yl)mn))=0.
To have a well-defined problem in the sense of Hadamard, an additional condition is needed. In the literature,68 two conditions are reported:

  • 1. Hkmn(y)=0,  on  Θvert, and

  • 2. ΘHkmn(y)dy=0,  in  Θ.

Herein, Θvert denotes the vertices of the unit cell boundary Θ. As always, each has advantages over the other. Although condition 1 is simpler to implement, condition 2 associates ui0(x) with the average displacement uic through

Eq. (53)

uc(x):=1vol(Θ)Θuζ(x,y)dy.
With the fine-scale displacement ansatz, the overall displacement function reads

Eq. (54)

ui(x,y)=ui0(x)+ζmnHi(y)u(m,xn)0(x)+O(ζ2),
which can be inserted into Eq. (53)

Eq. (55)

uic(x)1vol(Θ)Θui0(x)+ζmnHi(y)u0(m,xn)(x)+O(ζ2)dy=u(m,xn)0(x)1vol(Θ)Θ1dy=u0(x).
Within this treatise, condition 2 has been applied. From this, the leading order strain is given as

Eq. (56)

ϵijtot0=u(i,xj)0+u(i,xj)1=(Iijkl+H(k,yl)mn(y))u(m,xn)0=Emnklu(m,xn)0=:ϵkltotf(x,y),
with the strain influence function EmnklIijkl+H(k,yl)mn(y) and the fine-scale total strain ϵtotf. Averaging the fine-scale total strain over the unit cell yields the coarse-scale total strain

Eq. (57)

ϵmntotc(x)1vol(Θ)Θϵmntotf(x,y)dy=u(m,xn)0.
Herein, the integral vanishes due to the divergence identity, the divergence theorem, and the local periodicity of the displacement influence function. With this, the coarse scale total strain is given as

Eq. (58)

ϵkltotf(x,y)Emnkl(y)ϵmntotc(x).
The leading stress is computed according to

Eq. (59)

σij0(x,y)Σijmn(y)ϵmntotc(x)=:σijf(x,y),
with the stress influence function ΣijmnCijklEklmn.

Finally, the highest order, i.e., Eq. (41), averaged over the unit cell reads

Eq. (60)

1vol(Θ)Θ0σijyj+1uijyj+biζdy=01vol(Θ)Θ0σijyj+biζdy=0.
Herein, the second term vanishes after applying the divergence theorem and due to local periodicity. Inserting the stress–strain and strain–displacement relations for the respective scale yields

Eq. (61)

1vol(Θ)ΘCijklEklmn(y)dyxju0(m,xn)(x))+1vol(Θ)Θbiζdy=0Ccijmnxj(ϵmntotc)+bic=0cσijxj+bic=0,
with the according definitions of Ccijmn and σcij.

In summary, on the fine scale, the problem reads

Find Himn(y), such that

Eq. (62)

yj(Cijkl(H(k,yl)mn+Iklmn))=0,in  ΘΘHimn(y)dy=0,in  ΘHimn(y)=Himn(y+l),on  Θ,
where l is chosen, such that y+l is the point on the boundary opposite of y. The problem on the coarse scale is

Find uci(x), such that

Eq. (63)

σijcxj+bic=0,in  Ωσijc=Cijmncϵmntotc,in  Ωuic=u¯ic,on  ΓDσijcnj=t¯ic,on  ΓN.
Finally, the overall displacements can be recovered with

Eq. (64)

ui(x,y)=ui0(x)+ζmnHk(y)u(m,xn)0(x).

3.

Weak Formulation

To solve the aforementioned systems numerically, a Galerkin method is applied in the space dimension. The time discretization is performed using the Euler method. First, the two-scale heat conduction is explained, and second, the thermo-elasticity problem is discussed.

3.1.

Two-Scale Heat Conduction

Remembering the equation for the temperature influence function, i.e., Eq. (26), and choosing trial and test functions from the Hilbert space ϕ,φH01(Ω), containing functions vanishing on the boundary, the temperature influence function can be approximated within this space to read

Eq. (65)

Hj(y)H^jkϕk(y),k=1,,K.
Multiplying Eq. (26) by arbitrary test functions and integrating over the unit cell yields

Eq. (66)

Θyi(λ(δij+H^jkϕkyi))φldy=0,  φl,l=1,,L.
Constant terms can be factored out and brought to the right side. Using the product rule for the divergence operator and applying the divergence theorem gives

Eq. (67)

Θλyi(H^jkϕk)φlyidy=Θyi(λδijφl)dy.
Note that the boundary integral vanishes due to vanishing test functions. With the definition of a system matrix ARL×K and a thermal load vector bjRL, such that

Eq. (68)

AlkΘλϕkyiφlyidy,blΘyi(λδij)φldy.
The problem can be written in matrix form, i.e.

Eq. (69)

AH^j=bj.
Solving Eq. (69) for the coefficient vector H^jRK yields the approximate solution through Eq. (65), which in turn gives the heat flux influence function

Eq. (70)

Λij=λ(δij+H^jkϕkyi).

On the macroscale, the discretization in space is performed analogously. Again, the same notation is used to refer to trial and test functions. Note, however, that the functions are defined on the macroscopic space this time, e.g., ϕ=ϕ(x). Representing the temperature in the Hilbert space H01(Ω) with time-dependent coefficients T^=T^(t)

Eq. (71)

T0T^k(t)ϕk(x),k=1,,K,
and multiplying Eq. (27) by test functions φl(x) yields

Eq. (72)

Ωρcp¯T^k(t)tϕk(x)φl(x)T^k(t)xi(Λij¯ϕk(x)xj)φl(x)dx=ΩQ¯φl(x)dx,  φl,l=1,,L.
Henceforth, the arguments of test and trial functions are dropped for a terser notation. The second term can be rewritten using the divergence theorem and integrating the Neumann boundary condition from Eq. (27) and, therefore, it can be brought to the right side, too:

Eq. (73)

Ωρcp¯T^k(t)tϕkφldx=ΩQ¯φldx+ΓNqNφldx.
Now, the time domain is discretized using an explicit Euler scheme. Henceforth, the time step is denoted with a superscript n and the time step with Δt. With this notation in place, the temperature coefficients at the next time step T^kn+1 can be computed as

Eq. (74)

Ωρcp¯T^kn+1ϕkφldx=Ωρcp¯T^knϕkφldx+Δt(ΩQ¯nφldx+ΓNqNnφldx).
With the definition of the system matrix ARL×K and the thermal load vector bnRL, i.e.

Eq. (75)

AlkΩρcp¯ϕkφldx,blnΩρcp¯T^knϕkφldx+Δt(ΩQ¯nφldx+ΓNqNnφldx),
the matrix form of Eq. (74) reads

Eq. (76)

AT^n+1=bn.

3.2.

Thermo-Elastic Deformations

In analogy to Sec. 3.1, the weak formulation is now derived for the thermo-elastic deformations. For the fine-scale problem, the test and trial functions are chosen from the same Hilbert space H01(Θ) of functions vanishing on the boundary. With the approximation

Eq. (77)

HmnH^qqmnϕ,q=1,,Q,
for the stress influence function, and after Eq. (62) is multiplied by test functions φp and integrated over the unit cell, the problem reads

Eq. (78)

Θyj(Cijkl(H^qqmnϕ(k,yl)+Iklmn))φpdy=0,  φp,p=1,,P.
Applying the product identity for the divergence operator followed by the divergence theorem and taking into account the test functions vanishing on the boundary gives

Eq. (79)

ΘCijkl(H^qqmnϕ(k,yl)+Iklmn)φpyjdy=0.
The known terms can be written on the right side, which reads

Eq. (80)

ΘCijklH^qqmnϕ(k,yl)φpyjdy=ΘCijklIklmnφpyjdy.
To turn this expression into matrix form, the system matrix ARP×Q and load vector bRP are defined. In coefficient notation, they read

Eq. (81)

ApqiΘCijklqϕ(k,yl)φpyjdy,bimnΘCijklIklmnφpyjdy.
With this, the matrix form

Eq. (82)

AiH^mn=bmni,
can be solved for the coefficient vector H^mnRQ.

The solution to the discretized fine-scale problem is used to define the homogenized coarse-scale material properties. First, the displacement influence function can be reconstructed using

Eq. (83)

Himn(y)H^qmnϕqi(y).
Then, the strain influence function is given as

Eq. (84)

Eklmn(y)Iklmn+H(k,yl)mn(y),H(k,yl)mn(y)=H^qmn12(ϕqkyl(y)+ϕqlyk(y)).
With this, and the stress influence function

Eq. (85)

Σijmn(y)Cijkl(y)Eklmn(y),
the coarse-scale stress tensor can finally be computed to

Eq. (86)

Cijklc1vol(Θ)ΘΣijmn(y)dy.

Now, the coarse-scale stress equilibrium problem Eq. (63) is considered. Multiplying with suitable test functions φpH01(Ω)={fH1|f|ΓD=0} and integrating over the domain Ω yields

Eq. (87)

Ωcσijxjφpdx=Ωbicφpdx,  φp,p=1,,P.
Using, again, the product rule of the divergence operator and the divergence theorem, after inserting the traction defined on the Neumann boundary, gives

Eq. (88)

Ωσijcφpxjdx=Ωbicφpdx+ΓNt¯icφpdx.

Inserting the stress–strain and in turn the strain–displacement relations yields

Eq. (89)

ΩCijklc(u^qqcϕ(k,xl)+αΔTcδkl)φpxjdx=Ωbicφpdx+ΓNt¯icφpdx.
Again, constant terms are brought to the right, i.e.

Eq. (90)

ΩCijklcu^qqcϕ(k,xl)φpxjdx=ΩbicφpCijklcαΔTcδklφpxjdx+ΓNt¯icφpdx.
Factoring out the coefficients yields

Eq. (91)

ΩCijklqcϕ(k,xl)φpxjdxu^qc=ΩbicφpCijklcαΔTcδklφpxjdx+ΓNt¯icφpdx,
which can be brought into the matrix form

Eq. (92)

Aiu^c=bic,RP×QAipqΩCijklcϕq(k,xl)φpxjdx,RPbipΩbicφpCijklcαΔTcδklφpxjdx+ΓNt¯icφpdx.

4.

From a Single Beam to Multibeam Patches

To scale up production, the laser beam can be split into multiple beams, each structuring the same cavity in parallel. This scenario can also be accounted for in the described model. The trick is to replace the periodic unit cell domain to account for a multibeam patch. Therefore, instead of the unit cell shown in Fig. 1(a), the unit cell consists of multiple holes ablated from a cuboid (c.f. Fig. 3). Naturally, the volume heat source, which is included in the two-scale heat conduction problem in Sec. 2.2, has to account for the multiple beams as well. In addition, the two-scale formulation is the same.

Fig. 3

A unit cell accounting for multiple holes produced by a multibeam parallel ablation process.

OE_61_9_095103_f003.png

5.

Numerical Experiments

To validate the two-scale model, the structuring of a foil is simulated with the multiscale approach and compared with a hole-resolved simulation.

The geometric setup is similar to the foil shown in Fig. 1(b). The foil’s thickness is 2×105  m. The holes are assumed to be rectangular (c.f. Fig. 1) with a size of 4×105  m by 4×105  m on the top and a size of 2×105  m by 2×105  m on the bottom of the foil. There are 14×14 patches in the middle of the foil with 8×8 holes per patch. The holes are distributed on a patch with a hole pitch size of 5.7×105  m. This setup models a multibeam scanner unit with eight times eight laser beams (c.f. Table 1 for the beam parameters).

Table 1

Laser beam.

DescriptionVariable
TypeGaussian beam
PolarizationCircular
Wavelength1×106  m
Beam radius90×106  m
Power100 W
Focal length0.0 m
Rayleigh length500 m

To structure the hole foil, a hatch strategy is needed. The scanner is set to move in the y-direction first and then in the x-direction with a speed of 0.05  m/s. For the hole-resolved simulation, each hole is discretized with a mesh of 2×2×1 elements in the x, y, and z directions, respectively. On the other hand, in the multiscale simulation, the discretization is patch resolved; therefore, each patch corresponds to 1 element.

Beneath the foil, which is made of INVAR36 (c.f. Ref. 16 for the material parameters), an air layer and a soda-lime float glass layer, with a thickness of 5×106  m each, are added to dissipate heat and avoid thermally induced plastic distortions which cannot be handled by the described model. The material parameters for air and glass are shown in Tables 2 and 3, respectively.

Table 2

Material parameters for air.

DescriptionVariableValue
Poisson ratioν0.156
Young’s modulusE0.1 MPa
Thermal expansion coefficientα4.6×1016  1/K
Thermal conductivityλ0.026 W/(m K)
Specific heat capacitycp1000 J/kg K
Densityρ1  kg/m3

Table 3

Material parameters for soda-lime float FL glass.

DescriptionVariableValue
Poisson ratioν0.156
Young’s modulusE9.1×104  MPa
Thermal expansion coefficientα4.6×106  1/K
Thermal conductivityλ0.76 W/(m K)
Specific heat capacitycp800 J/kg K
Densityρ2530  kg/m3

It remains to define the initial conditions, the boundary conditions, and the sources. The load vectors for the microscale heat conduction problem read b0=(1,0,0)T, b1=(0,1,0)T, and b2=(0,0,1)T. For the mechanical microscale problem, the fine scale stiffness tensor has to be defined. In Voigt notation, i.e., after compressing the symmetric coefficients, the stiffness tensor can be represented as a symmetric matrix, which in turn can be stored as a vector with the entries C=(C11,C22,C33,C12,C23,C13). The six load cases read C6=(1,0,0,0,0,0), C7=(0,1,0,0,0,0), C8=(0,0,1,0,0,0), C9=(0,0,0,1,0,0), C10=(0,0,0,0,1,0), and C11=(0,0,0,0,0,1). For the heat conduction problem, the initial condition is ambient temperature, i.e., TDTa=298.15  K. The Neumann boundary is chosen to be isolating on all boundaries except for the bottom one, on which a heat transition condition is set, i.e., qN10(TTa). Finally, the source is given by QAPh/(Velneph) with the absorption coefficient A=0.4, the hole power Ph=0.0163416  W, the volume of the hole Velneph, and the number of elements per hole neph. For the mechanical problem, there is neither an initial condition nor a source. Instead, two boundary conditions have to be provided. The foil is clamped u(x)=(0,0,0)T at x=(1000,1000,0)T. Additionally, it is constrained in the x-direction at x=(1000,1000,0)T and in the y-direction at x=(1000,1000,0)T. There are no traction forces applied to the surface; therefore, t¯ic=0.

To solve the time dependent system Eq. (82), the time step is chosen to yield a Courant–Friedrichs–Lewy number of C=1000 and a total of 400 time steps is simulated. Then, the system is solved using the GMRES17 algorithm and the ILU018 preconditioner from the PETSc project.19 The algorithm stops when either a tolerance of 1×106 is reached or the number of iterations surpasses 100. The structural mechanics problem Eq. (92) is solved using the iterative Newton SOR method20 with a relaxation factor of 1.5.

6.

Results

The solutions to the microscale problem were computed once per load case. This information was then fed into the patch-resolved simulation. Therefore, there are no comparisons to the hole-resolved simulation. It can be seen that the temperature influence function depends highly on the fine scale thermal conductivity λ, which enters Eq. (26) on the right side (c.f. the left and middle pictures compared with the right one in Fig. 4).

Fig. 4

The temperature influence function Hj for the three load cases (1,0,0)T, (0,1,0)T, and (0,0,1)T is shown in Figs. 4(a)4(c), respectively.

OE_61_9_095103_f004.png

Figure 5 shows the displacement influence function for the first three load cases. Again, it can be seen that the geometry and the stiffness tensor play the dominant role as the displacement influence function in the right most plot is not as affected as in the other two cases.

Fig. 5

The magnitude of the displacement influence function Hmn for the first three (out of six) load cases, i.e., (1,0,0,0,0,0)T, (0,1,0,0,0,0)T, and (0,0,1,0,0,0)T is shown in Figs. 5(a)5(c), respectively.

OE_61_9_095103_f005.png

In Fig. 6, it can be seen that the temperature field is highest in the middle of the structured area. Even though the ablation area is rectangular, there is an almost radial symmetric temperature dissipation. Note, however, that the midpoint of this radial symmetric field is slightly shifted toward the right and the top of the structured area. This is an effect of the hatch strategy, which drives the laser from the lower left to the upper right corner. Because the heat cannot dissipate into the holes, the cooling rate is highest at the corners, high at the border of the pattern, and lowest at the midpoint. Hence, it is a radial symmetric temperature field.

Fig. 6

The temperature of the hole-resolved simulation is displayed in Fig. 6(a) and the multiscale simulation in Fig. 6(b).

OE_61_9_095103_f006.png

The displacements in Fig. 7 are higher in the positive x or y direction, respectively. As the deformation is solely driven by the temperature differences, this is again an impact of the hatching strategy. The elastic deformations in the left or bottom edge are already reducing to the initial configuration, whereas the deformation at the upper and right edges have just occurred. The somewhat lower displacements at the corner are a direct result of the clamping boundary conditions, which act on the bottom side of the foil.

Fig. 7

The displacement of the hole-resolved simulation in the x- and y-directions [c.f. Figs. 7(a) and 7(b)] and the multiscale simulation in the x- and y-directions [c.f. Figs. 7(c) and 7(d)].

OE_61_9_095103_f007.png

The von Mises stresses in Fig. 8 are lower at the corners and the edges of the foil and highest in the middle. This actually visualizes the clamping conditions and explains the reduced deformations toward the corners. Because the process is temperature based, it is obvious that the von Mises stress show a similar pattern as the temperature field.

Fig. 8

The von Mises stress of the hole-resolved simulation is displayed in Fig. 8(a), and the multiscale simulation is in Fig. 8(b).

OE_61_9_095103_f008.png

Finally, it is observed that the results of both the hole-resolved and the patch-resolved simulations coincide. The total runtime for the hole-resolved simulation was 36 067 s. The patch-resolved simulation took 421 s to compute the RVEs and 23 s to simulate the foil using the RVEs. In total, it is an improvement of over 98%, but because the RVEs only need to be computed once per material and hole shape, the improvement for further runs is even better at over 99%.

7.

Conclusion and Outlook

To develop the necessary process understanding to allow for a first-time-right production of FMMs, it is essential to have a fast simulation that aids in the search for better processing strategies. Therefore, a multiscale approach was developed within this treatise. This mathematical model consists of a two-scale heat conduction problem and a two-scale thermo-elasticity problem. In total, four tasks needed to be solved, two for the heat conduction and two for the structural mechanics. The discretization in space was performed using a Bubnov-Galerkin method, and the time discretization of the heat conduction problem was realized with the Euler method.

The implementation of the multiscale model is capable of reproducing the results simulated with a microscale model in a fraction of the time. The total achieved improvement in runtime is over 98% with an absolute runtime of a few seconds for the shown experiment.

The multiscale simulation enables scientists and researchers to explore and evaluate different scanning strategies with respect to the temperature-induced distortions in the work piece. Additionally, the processing strategies for filters or FMMs can be found faster compared with experiments, which might yield a quicker production ramp-up.

The current model does not account for plastic deformations. However, the yield stress is a good indicator for the occurrence of plastic deformations and can therefore be used to prevent situations in which the model fails to work. In future research, the authors plan to validate the model against physical experiments. In addition, the automation of the chain of calibrations and simulations is desired. Finally, the simulation might be used as a basis for an optimization algorithm, which selects the best scanning strategy for structuring thin foils.

Acknowledgments

Funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy—EXC-2023 Internet of Production—390621612. In addition, the authors acknowledge the publication of a prior version of this treatise in the Proceedings of the SPIE Photonics WEST 2022 conference.21 The authors have no relevant financial interests in this manuscript and no other potential conflicts of interest to disclose.

References

1. 

K. Kim et al., “FMM materials and manufacturing process: review of the technical issues,” SID Symp. Dig. Tech. Pap., 49 (1), 1011 –1013 (2018). https://doi.org/10.1002/sdtp.12230 DTPSDS 0097-966X Google Scholar

2. 

C. Kim et al., “Fine metal mask material and manufacturing process for high-resolution active-matrix organic light-emitting diode displays,” J. Soc. Inf. Disp., 28 (8), 668 –679 (2020). https://doi.org/10.1002/jsid.901 JSIDE8 0734-1768 Google Scholar

3. 

O. Hofmann et al., “Highly dynamic positioning of individual laser beams in a multi-beam system for laser surface processing,” Proc. CIRP, 94 812 –816 (2020). https://doi.org/10.1016/j.procir.2020.09.123 Google Scholar

4. 

S. Eifel, Effizienz- und Qualitätssteigerung bei der Lasermikrobearbeitung mit UKP-Lasern durch neue optische Systemtechnik, 1. aufl ed.Ergebnisse aus der Lasertechnik, Apprimus(2015). Google Scholar

5. 

C. Heinigk et al., “A multi-scale model for ultra short pulsed parallel laser structuring—Part I. The micro-scale model,” J. Laser Micro Nanoeng., 16 (2), 144 –149 (2021). https://doi.org/10.2961/jlmn.2021.02.2011 Google Scholar

6. 

J. Fish, Practical Multiscaling, John Wiley & Sons Inc.(2013). Google Scholar

7. 

J. Fish and R. Fan, “Mathematical homogenization of nonperiodic heterogeneous media subjected to large deformation transient loading,” Int. J. Numer. Methods Eng., 76 (7), 1044 –1064 (2008). https://doi.org/10.1002/nme.2355 IJNMBH 0029-5981 Google Scholar

8. 

Multiscale Methods: Bridging the Scales in Science and Engineering, Oxford University Press(2010). Google Scholar

9. 

E. Weinan, Principles of Multiscale Modeling, Cambridge University Press(2011). Google Scholar

10. 

W. E. B. Engquist et al., “Heterogeneous multiscale methods: a review,” Commun. Comput. Phys., 2 (3), 367 –450 (2007). https://doi.org/10.48550/arXiv.physics/0205048 1815-2406 Google Scholar

11. 

B. Neuenschwander, B. Jaeggi and M. Schmid, “From fs to Sub-ns: dependence of the material removal rate on the pulse duration for metals,” Phys. Proc., 41 794 –801 (2013). https://doi.org/10.1016/j.phpro.2013.03.150 PPHRCK 1875-3892 Google Scholar

12. 

T.-R. Hsu, Finite Element Method in Thermomechanics, Springer Netherlands(2012). Google Scholar

13. 

A. F. Bower, Applied Mechanics of Solids, CRC Press(2010). Google Scholar

14. 

O. C. Zienkiewicz, R. L. Taylor and D. Fox, The Finite Element Method for Solid and Structural Mechanics, 7th ed.Elsevier/Butterworth-Heinemann(2014). Google Scholar

15. 

I. Temizer, “On the asymptotic expansion treatment of two-scale finite thermoelasticity,” Int. J. Eng. Sci., 53 74 –84 (2012). https://doi.org/10.1016/j.ijengsci.2012.01.003 IJESAN 0020-7225 Google Scholar

16. 

G. Hausch and H. Warlimont, “Single crystalline elastic constants of ferromagnetic face centered cubic Fe-Ni invar alloys,” Acta Metall., 21 (4), 401 –414 (1973). https://doi.org/10.1016/0001-6160(73)90197-1 AMETAR 0001-6160 Google Scholar

17. 

Y. Saad and M. H. Schultz, “GMRES: a generalized minimal residual algorithm for solving nonsymmetric linear systems,” SIAM J. Sci. Stat. Comput., 7 (3), 856 –869 (1986). https://doi.org/10.1137/0907058 SIJCD4 0196-5204 Google Scholar

18. 

Y. Saad, Iterative Methods for Sparse Linear Systems, PWS Publ. Co(1996). Google Scholar

19. 

S. Balay et al., PETSc/TAO Users Manual, 7. Argonne National Laboratory, Illinois (2022). Google Scholar

20. 

D. M. Young and W. Rheinboldt, Iterative Solution of Large Linear Systems, Elsevier Science(2014). Google Scholar

21. 

C. Heinigk et al., “A multi-scale model for ultra short pulsed parallel laser structuring—Part II. The macro-scale model,” Proc. SPIE, 11994 119940B (2022). https://doi.org/10.1117/12.2608151 PSISDG 0277-786X Google Scholar

Biographies of the authors are not available.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Christian Heinigk, T. Barthels, W. Schulz, and M. Nießen "Multiscale model for ultrashort pulsed parallel laser structuring—part II. The macroscale model," Optical Engineering 61(9), 095103 (6 September 2022). https://doi.org/10.1117/1.OE.61.9.095103
Received: 4 March 2022; Accepted: 18 August 2022; Published: 6 September 2022
Advertisement
Advertisement
KEYWORDS
Pulsed laser operation

Optical engineering

Computer simulations

Electroluminescence

Metals

Heat flux

Laser ablation

Back to Top