Conference PaperPDF Available

Closed-Loop Simulation Integrating Finite Element Modeling With Feedback Controls in Powder Bed Fusion Additive Manufacturing

Authors:

Abstract and Figures

Powder bed fusion (PBF) additive manufacturing has enabled unmatched agile manufacturing of a wide range of products from engine components to medical implants. While high-fidelity finite element modeling and feedback control have been identified key for predicting and engineering part qualities in PBF, existing results in each realm are developed in opposite computational architectures wildly different in time scale. Integrating both realms, this paper builds a first-instance closed-loop simulation framework by utilizing the output signals retrieved from the finite element model (FEM) to directly update the control signals sent to the model. The proposed closed-loop simulation enables testing the limits of advanced controls in PBF and surveying the parameter space fully to generate more predictable part qualities. Along the course of formulating the framework, we verify the FEM by comparing its results with experimental and analytical solutions and then use the FEM to understand the melt-pool evolution induced by the in-layer thermomechanical interactions. From there, we build a repetitive control algorithm to greatly attenuate variations of the melt pool width.
Content may be subject to copyright.
Proceedings of 2020 International Symposium on Flexible Automation
ISFA 2020
July 5-9, 2020, Chicago, Illinois, U.S.A
[ISFA2020-XXXXX]
CLOSED-LOOP SIMULATION INTEGRATING FINITE ELEMENT MODELING WITH
FEEDBACK CONTROLS IN POWDER BED FUSION ADDITIVE MANUFACTURING
Dan Wang
Dept. of Mechanical Engineering
University of Washington
Seattle, Washington, 98195
Email: daw1230@uw.edu
Xu Chen
Dept. of Mechanical Engineering
University of Washington
Seattle, Washington, 98195
Email: chx@uw.edu
ABSTRACT
Powder bed fusion (PBF) additive manufacturing has en-
abled unmatched agile manufacturing of a wide range of prod-
ucts from engine components to medical implants. While high-
fidelity finite element modeling and feedback control have been
identified key for predicting and engineering part qualities in
PBF, existing results in each realm are developed in opposite
computational architectures wildly different in time scale. Inte-
grating both realms, this paper builds a first-instance closed-loop
simulation framework by utilizing the output signals retrieved
from the finite element model (FEM) to directly update the con-
trol signals sent to the model. The proposed closed-loop simu-
lation enables testing the limits of advanced controls in PBF and
surveying the parameter space fully to generate more predictable
part qualities. Along the course of formulating the framework,
we verify the FEM by comparing its results with experimental
and analytical solutions and then use the FEM to understand the
melt-pool evolution induced by the in-layer thermomechanical
interactions. From there, we build a repetitive control algorithm
to greatly attenuate variations of the melt pool width.
1 Introduction
Additive manufacturing (AM) builds a part directly from its
digital model by joining materials layer by layer, which is dif-
ferent from conventional subtractive machining. In particular,
powder bed fusion (PBF) AM, applying high-precision lasers or
Corresponding author
electron beams as the energy source, has enabled unprecedented
fabrication of complex parts from polymeric and metallic powder
materials. However, broader adoption of the technology remains
challenged by insufficient reliability and in-process variations in-
duced by, for example, uncertain laser-material interactions, en-
vironmental vibrations, powder recycling, imperfect interactions
of mechanical components, and complex thermal histories of ma-
terials [1–3].
Current researches employ finite element modeling and
feedback controls to understand the energy-deposition mecha-
nisms and to regulate the in-process variations in PBF and other
AM technologies such as laser metal deposition (LMD). Partic-
ularly, [4–6] adopt finite element modeling to investigate the ef-
fects of various scan patterns, scan speeds, number of lasers, and
overhanging structures on the thermal fields of the powder bed,
the geometries of the melt pool, and the mechanical properties
of the printed parts. Existing strategies on feedback controls of-
ten implement low-order system models obtained using system
identification techniques [2, 7–9]. A nonlinear memoryless sub-
model [8, 10] and a spatial-domain Hammerstein model [9] have
been built to cover more complicated process dynamics. From
there, PID control [2, 11–13], sliding mode control [10], predic-
tive control [7], and iterative learning control [14] have proved
their efficiencies in improving the dimensional accuracy of the
printed parts in PBF and LMD.
Although finite element models (FEMs) and feedback con-
trols have been identified key for predicting and engineering part
qualities in PBF, existing results in each realm are developed in
separate computational architectures due to their different time
Vt
Vs
W
L
Laser
beam
Powder bed
Unfused
powder
Fused tracks
B
Scanning
direction
Traveling
direction
W
Energy
beam
x
1 2 345678 9 10
Track No.
Track direction
x
T
xodd track
even track
00
Heat
conduction
direction
n-1 n
Track No.
Scan
direction
Heat
conduction
direction
n-2
n-3
Unfused
powder
𝜉
𝑦
A
C
FIGURE 1. Schematic of in-layer sintering process in PBF.
scales. To be more specific, feedback controls are implemented
in real time, while it can take hours or even days for FEMs to
simulate the sintering of a few layers that finishes in seconds
in reality. If we can integrate FEMs with feedback controls di-
rectly in a closed loop, however, we will be able to 1) combine
aforementioned knowledges from each realm, 2) test the limits
of advanced controls in PBF, 3) survey the parameter space fully
to generate more predictable part qualities, and 4) quickly design
controllers and update parameters for novel materials and printer
settings. These benefits are more prominent when the experi-
ments are costly and time-consuming.
In pursuit of the above benefits, this paper builds, in the first
instance to our best knowledge, a closed-loop high-fidelity sim-
ulation framework that leverages modern architectures of finite-
element-modeling tools and the power of data processing and ad-
vanced controls. Specifically, we build a bidirectional communi-
cation so that the output signals (e.g., melt pool width) retrieved
from the FEM can be utilized to directly update the FEM process
parameters (e.g., laser power) in external control toolboxes (e.g.,
MATLAB). Along the course of formulating the framework, we
validate the FEM by comparing its results with experimental and
analytical solutions and furthermore apply the FEM to investi-
gate the periodic in-layer thermal interactions. Under the frame-
work of the closed-loop simulation, we then verify the effective-
ness of the repetitive control (RC) in attenuating the repetitive
variations of the melt pool width.
The remainder of this paper is structured as follows. Sec-
tion 2 builds the main closed-loop simulation framework taking
an FEM and a plug-in RC design for example. Section 3 verifies
the FEM and justifies the existence of the periodic in-layer ther-
mal interactions. Section 4 implement the proposed closed-loop
simulation to evaluate the performance of RC in attenuating the
periodic in-layer disturbances. Section 5 concludes the paper.
2 Proposed high-fidelity closed-loop simulation
A typical part in PBF is built from many thousands of thin
layers. Within each layer (Fig. 1), the energy beam is regulated
to follow trajectories predefined by the part geometry in a slicing
process. After one layer is finished printing, a new thin layer of
powder will be spread on top, and then another cycle begins. This
section frames the main high-fidelity closed-loop simulation. We
first design an FEM to simulate the thermal fields during the PBF
process. After that, a sample RC algorithm is designed and in-
troduced to the closed-loop simulation.
2.1 FEM
We use the COMSOL Multiphysics 5.3a software to build
and refine the FEM of the thermal fields in PBF. The model con-
siders surface convection, surface radiation, conduction, and la-
tent heat of fusion. For brevity and without loss of generality,
the effects of evaporation, fluid flow, and Marangoni force are
neglected. The governing equation for conduction heat flow is
ρcp
dT (x,y,z,t)
dt =·(kT(x,y,z,t)) + qs,(1)
where kis the thermal conductivity, cpthe specific heat capac-
ity, ρthe effective density, tthe time, Tthe temperature, and qs
the rate of local internal energy generated per unit volume [15].
When no confusion would arise in the context, T(x,y,z,t)is ab-
breviated to Tin the remaining of this paper.
2.1.1 Phase change and temperature-dependent
thermal properties We account for the latent heat of fusion
Lfby introducing the effective heat capacity [19]:
cp,e f f (T) =
cp1(T)T0<TTsol
Lf
TmTsol +cp1(Tsol)+cp2(Tm)
2Tsol <T<Tm
cp2(T)TTm
,(2)
where T0is the ambient temperature, Tsol the solidus temperature,
Tmthe melting point, cp1the heat capacity of the powder, and cp2
the heat capacity of the liquid.
For the thermal properties, we adopt k,cp, and ρin [4, 16]
for the solid and liquid materials. For the powder material,
we use the thermal properties generated from the solid mate-
rial by considering the porosity φ[17, 18]: kpowder =ksolid (1
φ)4and ρpowder =ρsolid (1φ), where φis expressed as
φ(T) =
φ0T0<TTsol
φ0
TsolTm(TTm)Tsol <T<Tm
0TTm
with φ0denoting the initial porosity. Here, the heat capacity is
assumed to be the same for the powder and solid materials except
in Tsol <T<Tm[17]. Fig. 2 shows the temperature-dependent
thermal properties used in this paper.
500 1000 1500 2000 2500
Temperature (K)
0
10
20
30
40
50
Thermal conductivity (W/m .K)
500 1000 1500 2000 2500
Temperature (K)
2000
2500
3000
3500
4000
4500
Density (kg/m3)
500 1000 1500 2000 2500
Temperature (K)
0
1
2
3
4
5
6
7
Heat Capacity (J/g .K)
FIGURE 2. Temperature-dependent thermal properties of Ti6Al4V [4, 16–18]. Solid line: solid and liquid materials. Dash-dotted line: powder
material. The two vertical dotted lines respectively indicate Tsol and Tm.
2.1.2 Initial condition, boundary conditions, and
laser beam profile The initial condition is T(x,y,z,0) = T0.
One boundary condition is established by assuming the bottom
(z=h) has no heat loss: kT
z
z=h=0. The other boundary con-
dition considers surface conduction, convection, and radiation:
kT
z
z=0
=Q+hc(TT0) + εσB(T4T4
0),(3)
where Qis the input heat flux, hcthe convection heat transfer
coefficient, εthe emissivity, and σBthe Stefan-Boltzmann con-
stant. Here, we assume Qhas a Gaussian laser beam profile:
Q2q
πR2e2r2
R2, where qis the laser power, Rthe effective laser
beam radius, and rthe radial distance from a certain point to the
center of the laser spot. In Appendix, we list the process param-
eters used in this study unless otherwise specified.
2.1.3 Meshing and scanning schemes The left
plot of Fig. 3 shows the built FEM with a substrate and a thin
layer of powder bed. In this FEM, we use a selective meshing
scheme to balance model accuracy with computation time: a fine
quad-and-swept mesh with a maximum element size of 60 µm
is applied to the central powder bed region that directly inter-
acts with the energy beam, whereas less finer tetrahedral mesh
(3.5mm) and triangular-and-swept mesh (2 mm) are applied to
the substrate and the peripheral powder bed, respectively. The
left plot of Fig. 3 also illustrates the bidirectional scan scheme
used in this study with a hatch spacing (xin Fig. 1) of 60 µm.
The developed FEM will be verified in Section 3.
2.2 Closed-loop simulation framework
We propose here the main closed-loop simulation frame-
work that integrates feedback controls with FEM (e.g., the FEM
in Section 2.1) and enables updating directly the control signals
of the FEM. This closed-loop framework is designed using the
X
Z
Y
Powder bed: F ree triangular
and Swept (2mm)
Substrate: Free
tetrahedral (3.5 mm)
10 mm
5mm
2mm
50 µm
Laser tracks
Powder bed: F ree quad
and Swept (60 µm)
(K)
FIGURE 3. Left: powder bed and substrate with selective meshing
scheme. Right: surface temperature distribution at t=0.14s. The lined
isotherm stands for T=Tm.
Initialization
!"# ", $ # $"
%&'( )( *( !"+ # %
"
% '( )( *( !,# % '( )( *( !-
!,# !-
(.
$ # $&!-+End if !-# !/01
FEM calculation for
!,2!-# !,3 %
4
Calculate melt pool
width 5&!-+from
%&'( )( *( !-+
Apply control
algorithms
(Baseline or RC)
Compare 5
with 51
Get control sign al
$&!6+(laser power)
COMSOL MATLAB
Retriev ing
%&'( )(*( !-+
FIGURE 4. Schematic of proposed closed-loop simulation.
software LiveLink™ for MATLAB and mainly composed of two
parts: FEM developed using COMSOL and feedback control al-
gorithms designed using MATLAB. The key idea of this closed-
loop framework is to use the output signals retrieved from the
FEM to update in MATLAB the control signals sent back to the
model step by step. As a case study, we use melt pool width as
the output signal and laser power as the control signal.
Fig. 4 illustrates the procedures of the proposed closed-loop
simulation. First of all, we initialize the FEM in Section 2.1 by
setting the start time t0as 0, the laser power qas the initial one
q0, and T(x,y,z,t0)as the ambient temperature T0. Note that
the computation time of the FEM is set as one time step from t0
to tf=t0+Ts, and afterwards, MATLAB will call the FEM re-
cursively to finish the whole simulation with a longer time tend.
The design and initialization of the FEM is completed in COM-
C(z)FEM or P(z)
Q(z)
zmˆ
P1(z)zm
d(k)
+
r(k)+e(k)
+
u(k)
+
y(k)
+
ωc(k)
+
+
Plug-in compensator
FIGURE 5. Block diagram of a plug-in RC design.
SOL, while the main file of the closed-loop framework is writ-
ten in MATLAB. When the main file starts running, the com-
mand model.study(’std1’).run first calls COMSOL to compute
the FEM (named std1) for one time step, and then the function
mphinterp retrieves the temperature distribution T(x,y,z,tf)at
t=tffrom COMSOL. Thereafter, the main file calculates the
melt pool width wat t=tffrom T(x,y,z,tf)and, based on the
control algorithms, processes w(tf)and obtains the control sig-
nal q(tf). At the final step, the iterative variables in the FEM are
updated by assigning tfto t0,T(x,y,z,tf)to T(x,y,z,t0), and
q(tf)to the laser power. After this iteration, MATLAB will call
COMSOL again to start a new FEM computation with the up-
dated variables, and then another cycle begins. The closed-loop
simulation will stop when tfreaches to tend .
The proposed closed-loop simulation achieves updating in a
closed loop the control signals of FEM. This simulation frame-
work will benefit and guide experiments by validating before-
hand the effectiveness of the servo designs. Next we will bring a
RC algorithm into the proposed closed-loop simulation.
2.3 Repetitive controller design
RC is designed for tracking/rejecting periodic exogenous
references/disturbances in applications with repetitive tasks [20].
By learning from previous iterations, RC can greatly enhance
current control performance in the structured task space. In digi-
tal RC, an internal model 1/(1zN)is incorporated in the con-
troller, where zis the complex indeterminate in the z-transform.
N=fs/f0is the period of the signal, where fs=1/Tsis the sam-
pling frequency and f0is the fundamental disturbance frequency.
Consider a baseline feedback system consisting a plant P(z)and
a baseline controller C(z)(Fig. 5 with the dotted box removed).
Here, C(z)can be designed by conventional servo algorithms,
such as PID, H, and lead-lag compensation. The signals r(k),
d(k),u(k), and y(k)respectively represent the reference, the in-
put disturbance, the control signal, and the system output. The
sensitivity function S(z) = 1
1+P(z)C(z)is the transfer function from
d(k)to y(k).
We introduce here a plug-in RC design [21] that uses the
internal signals e(k)and u(k)to generate a compensation signal
ωc(k)(Fig. 5). Let mdenote the relative degree of ˆ
P(z), where
ˆ
P(z)is the nominal model of P(z). The transfer function of the
overall controller from e(k)to u(k)is
Call (z) = C(z) + zmˆ
P1(z)Q(z)
1zmQ(z).(4)
The internal model is integrated in Call by designing the
Qfilter as Q(z)=(1αN)zmN/(1αNzN), which gives
1zmQ(z) = (1zN)/(1αNzN), where α[0,1)is
a tuning factor. At the harmonic frequencies ωk=k2πf0Ts
(kZ+), with z=ejωk, we have 1 zN=0, 1 zmQ(z) =
0, Call (z)from (4), and hence the new sensitivity func-
tion S0(z) = 1
1+P(z)Call(z)=0. At the intermediate frequencies
ω6=k2πf0Ts, with z=ejωand αbeing close to 1, Q(z)0,
1zmQ(z)1, Call (z)C(z)from (4), and thereby the origi-
nal loop shape is maintained.
During implementation, zero-phase pairs q0(z1)q0(z)are
attached to Q(z)for robustness against high-frequency plant un-
certainties:
Q(z) = (1αN)zmN
1αNzNq0(z1)q0(z),(5)
where q0(z)=(1+z)n0/2n0and n0Z+. The closed-loop per-
formance S0(z)can be tuned by choosing different αand n0[21].
The plug-in RC and the baseline control can be easily incorpo-
rated into the closed-loop simulation by setting u(k)as q(tf)and
y(k)as w(tf). Under the framework of the closed-loop simula-
tion, we will prove in Section 4 the effectiveness of RC in PBF.
3 Model verification and thermal interactions
In this section, we verify the FEM in Section 2.1 and then
apply it to understand the periodic in-layer thermal cycles.
3.1 Model verification
We compare the melt pool widths obtained from the FEM
first with the experimental results and then with the analytical
solutions. Throughout this paper, melt pool widths are derived
from the temperature distribution (e.g., T(x,y,z,t)in the FEM)
by searching around the position of the laser beam to find the
maximum width of the melt pool geometry bounded by Tm.
We compare in Table 1 the numerical melt pool widths with
the experimental results in [22]. The laser power is fixed to 50 W,
and the scan speed is 100, 200, or 300 mm/s. Overall, the FEM
gives reasonable predictions of the melt pool widths with errors
of 3.61%, 6.41%, and 5.44%, respectively. The main reason
that the numerical melt pool widths are slightly (less than 10 µm)
larger than the experimental results is that evaporation is ignored
TABLE 1. Melt pool widths from FEM and experimental results
[22] with a fixed laser power of 50 W. Difference=FEM-Experiments.
Error=(FEM-Experiments)/FEM.
Scan speed (mm/s) 100 200 300
FEM (µm) 182 152.63 132.56
Experiments (µm) Min/Max 165.71/175.43 140.71/142.85 120.71/125.35
Difference (µm) 6.57 9.78 7.21
Error 3.61% 6.41% 5.44%
X
200 um
Y
Rosent hal solu tion
Z
Y
Sample 1
! = #
$
Sample 2
! = 2#
$
Δ! = #
$
Sample 3
! = 3#
$
FIGURE 6. Melt pool widths from the FEM and analytical solution.
Right and bottom left plots share the same scale and legend.
in the FEM so the overheated material and the heat within are
condensed in the melt pool.
Then we compare the FEM results with the analytical solu-
tions. When a moving point laser source is acting on a thick plate
and the thermal properties of the plate are constant, the analytical
solution of (1) in the steady state is the Rosenthal equation [15]:
T(ξ,y,z)T0=q
2πkr eux(r+ξ)
2κ, where (ξ,y,z)is a coordinate
system attached to the moving source, r=pξ2+y2+z2, and
κ=k/(ρcp). For comparison, the FEM is adapted to accom-
modate the assumptions of the Rosenthal equation, such as con-
stant thermal properties (k=5 W/(m·K), cp=1.1 J/(g·K), and
ρ=4300 kg/m3) and point heat source. Fig. 6 compares the
numerical and analytical solutions. As shown in the right plot
and the bottom left plot, after 27 samples, the numerical melt
pool geometry reaches to the steady state and matches with the
Rosenthal solution (the outline). Also, from the top left plot of
Fig. 6, we can tell that the melt pool widths obtained from the
FEM and the Rosenthal equation match well with each other un-
der different combinations of scan speeds and laser powers.
3.2 Periodic thermal interactions
After having validated the FEM, next we will adopt it to in-
vestigate the periodic in-layer thermal cycles in PBF. Here, we
bidirectionally sinter 10 tracks within one layer (Figs. 1 and 3).
The right plot in Fig. 3 illustrates the simulated surface temper-
ature distribution at t=0.14 s, where the isotherm of T=Tm
indicates the melt pool geometry. From the solid line in the top
plot of Fig. 7, we observe that the melt pool width changes over
time and structurally deviates from the steady-state value 246µm
extracted from the first track. Most importantly, the start of each
track has larger melt pool widths than the rest of the track. This
is because in bidirectional scanning, when the energy beam ap-
proaches the end of one track, the large latent heat does not have
enough time to dissipate out before the next track starts. The re-
sulting increased melt pool widths at the beginning of each track
form a periodic disturbance with a repetitive spectrum in the fre-
quency domain (the solid line in the middle plot of Fig. 7). The
fundamental frequency f0of the disturbance is determined by the
duration of scanning one track t0, that is, f0=1/t0=ux/L, where
uxis the scan speed and Lis the track length. In this example,
f0=100/5=20Hz, and frequency spikes at n f0(nZ+, the set
of positive integers) appear in the fast Fourier transform (FFT) of
the disturbance.
The disturbance periodicity is closely related to the recur-
ring laser scanning trajectories and the repetitive in-layer ther-
momechanical interactions. Besides the bidirectional scan used
in this study, other scan patterns yield similar repetitive distur-
bances (see, e.g., experimental results in [23]). To deal with these
undesired repetitive spectra, we implement the closed-loop simu-
lation by bringing automatic control algorithms [1,21] into finite
element modeling, as will be discussed in Section 4.
4 Results and Analyses
This section employs the proposed closed-loop simulation
to evaluate the performances of the baseline control and RC in
attenuating the variations of the melt pool width (Section 3.2).
First, we identify the plant model of the FEM from the laser
power to the melt pool width as P(s) = 0.001671/(s+1055).
The input signals used for system identification include a pseudo-
random binary sequence (PRBS) signal and multiple sinusoidal
signals (10~300 Hz), with magnitudes of 20 W and add-on DC
components of 60 W. The frequency responses of the measured
and identified systems match well with each other (see Fig. 8).
After that, we design a PI controller as C(s) = Kp+Ki/s
with Kp=9.38 ×105and Ki=1.66 ×109. Under the sam-
pling time Tsof 0.5ms (i.e., fs=2 kHz), the zero-order-hold
equivalents of the plant and controller models respectively are
P(z) = 6.493 ×107/(z0.5901)and C(z)=(9.38z1.08)×
105/(z1). The dashed line in Fig. 9 shows the magnitude re-
sponse of the sensitivity function S(z)in the baseline feedback
loop composed of P(z)and C(z). Such a design provides a band-
FIGURE 7. In-layer thermal disturbance. Top: time-domain. Middle:
frequency-domain (FFT). Bottom: laser power (control signals u(k)in
Fig. 5). The three plots share the same legend. σdenotes the standard
deviation.
10110210 3104
-160
-150
-140
-130
-120
Magnitude (dB)
10110210 3104
Frequency (Hz)
-200
-100
0
100
200
Phase (degree)
Measured system using sinusoidal and PRBS signals
identified system P = 0.001671/(s+1055)
FIGURE 8. Measured and identified system responses.
width at 197Hz, which approximates the limit of 20% of the
Nyquist frequency (1000Hz) and indicates that the PI controller
is well tuned. The closed-loop simulations are designed accord-
ing to Section 2.2 integrating FEM with baseline control and RC,
respectively. Here, in Fig. 5, r(k) = 0, and d(k)comes from the
in-layer thermal interactions. From the frequency-domain results
in Fig. 7, we can tell that the baseline PI control can attenuate
10010110 2103
Frequency (Hz)
-60
-40
-20
0
Magnitude (dB)
Baseline PI control
Repetitive control
FIGURE 9. Magnitude responses of sensitivity functions S(z)in base-
line control and S0(z)in RC.
to some extent the frequency spikes below the closed-loop band-
width but not the other high-frequency spikes. Compared to the
case without control, the baseline feedback loop decreases the
3σvalue of the variations of the melt pool width (y(k)in Fig. 5)
by 21.57%, where σdenotes the standard deviation.
To enhance the disturbance-attenuation performance, we
bring the plug-in RC compensator in Section 2.3 into the closed-
loop simulation. In the Q-filter design in (5), the relative de-
gree mof ˆ
P(z)(=P(z)in this example) is 1, the disturbance pe-
riod N=fs/f0=2000/20 =100, and we choose α=0.99 and
n0=1. With the plug-in RC introduced, high-gain control ef-
forts are generated exactly at 20Hz and its harmonics (see S0(z)
in the solid line of Fig. 9). The bottom plot of Fig. 7 illustrates
the control signals u(k)of the baseline control, the RC, and the
case without control. As shown in the middle plot of Fig. 7,
compared with the baseline control, RC further lowers the peri-
odic frequency spikes especially at high frequencies beyond the
closed-loop bandwidth and decreases the 3σvalue by 35.97%.
Similarly, in the time domain, the increased control efforts of RC
at the harmonic frequencies yield a further-attenuated output y(k)
(the top plot of Fig. 7).
5 Conclusion
In this paper, we first build a finite element model (FEM) to
simulate the temperature response in powder bed fusion (PBF)
additive manufacturing. Then we validate the FEM by compar-
ing the numerical results with the experimental and analytical so-
lutions. Employing the FEM, we justify the existence of the pe-
riodic disturbances in the evolution of the melt pool width. From
there, we develop a first-instance closed-loop simulation frame-
work by integrating FEM with feedback controls (e.g., baseline
PI control and repetitive control) to reduce the in-process vari-
ations and advance the part quality in PBF. Implementing this
closed-loop frameworks, we validate that the repetitive control
algorithm attenuates the periodic disturbances more substantially
by 35.97% compared to the PI control.
Acknowledgment
This material is based upon work supported in part by the
National Science Foundation under Grant No. 1953155.
Appendix: defined parameters of the FEM
Parameters Value Parameters Value
Powder bed size 5mm ×10 mm ×50 µm Material Ti6Al4V
Substrate size 5 mm ×10mm ×2 mm Track length L5 mm
Laser spot diameter 2R220µmm Time step Ts0.5ms
Powder bed absorptance 0.25 Emissivity 0.35
Solidus temperature Tsol 1873K Scan speed ux100mm/s
Latent heat of fusion Lf295kJ/kg Laser power q60 W
T0/Tm293.15 K/1923.15 K φ00.4
hc12.7W/(m2·K) k,cp, and ρFig. 2
References
[1] D. Wang and X. Chen, “A multirate fractional-order repetitive control for
laser-based additive manufacturing,Control Engineering Practice, vol. 77,
pp. 41–51, 2018.
[2] J.-P. Kruth, P. Mercelis, J. Van Vaerenbergh, and T. Craeghs, “Feedback
control of selective laser melting,” in Proceedings of the 3rd international
conference on advanced research in virtual and rapid prototyping, 2007,
pp. 521–527.
[3] V. Seyda, N. Kaufmann, and C. Emmelmann, “Investigation of aging pro-
cesses of ti-6al-4 v powder material in laser melting,Physics Procedia,
vol. 39, pp. 425–431, 2012.
[4] M. Masoomi, S. M. Thompson, and N. Shamsaei, “Laser powder bed fusion
of ti-6al-4v parts: Thermal modeling and mechanical implications, Inter-
national Journal of Machine Tools and Manufacture, vol. 118, pp. 73–90,
2017.
[5] A. Hussein, L. Hao, C. Yan, and R. Everson, “Finite element simulation
of the temperature and stress fields in single layers built without-support
in selective laser melting,Materials & Design (1980-2015), vol. 52, pp.
638–647, 2013.
[6] A. Foroozmehr, M. Badrossamay, E. Foroozmehr et al., “Finite element
simulation of selective laser melting process considering optical penetration
depth of laser in powder bed,Materials & Design, vol. 89, pp. 255–263,
2016.
[7] L. Song and J. Mazumder, “Feedback control of melt pool temperature dur-
ing laser cladding process,” IEEE Transactions on Control Systems Tech-
nology, vol. 19, no. 6, pp. 1349–1356, 2011.
[8] X. Cao and B. Ayalew, “Control-oriented mimo modeling of laser-aided
powder deposition processes,” in American Control Conference (ACC),
2015. IEEE, 2015, pp. 3637–3642.
[9] P. M. Sammons, D. A. Bristow, and R. G. Landers, “Repetitive process
control of laser metal deposition,” in ASME 2014 Dynamic Systems and
Control Conference. American Society of Mechanical Engineers, 2014,
pp. V002T35A004–V002T35A004.
[10] A. Fathi, A. Khajepour, M. Durali, and E. Toyserkani, “Geometry control
of the deposited layer in a nonplanar laser cladding process using a variable
structure controller,Journal of manufacturing science and engineering,
vol. 130, no. 3, p. 031003, 2008.
[11] J. Hofman, B. Pathiraj, J. Van Dijk, D. de Lange, and J. Meijer, “A camera
based feedback control strategy for the laser cladding process,” Journal of
Materials Processing Technology, vol. 212, no. 11, pp. 2455–2462, 2012.
[12] D. Salehi and M. Brandt, “Melt pool temperature control using labview in
nd: Yag laser blown powder cladding process, The international journal
of advanced manufacturing technology, vol. 29, no. 3, pp. 273–278, 2006.
[13] A. Fathi, A. Khajepour, E. Toyserkani, and M. Durali, “Clad height con-
trol in laser solid freeform fabrication using a feedforward pid controller,
The International Journal of Advanced Manufacturing Technology, vol. 35,
no. 3, pp. 280–292, 2007.
[14] L. Tang and R. G. Landers, “Layer-to-layer height control for laser metal
deposition process,” Journal of Manufacturing Science and Engineering,
vol. 133, no. 2, p. 021009, 2011.
[15] E. Kannatey-Asibu Jr, Principles of laser materials processing. John Wi-
ley & Sons, 2009, vol. 4.
[16] A. N. Arce, Thermal modeling and simulation of electron beam melting
for rapid prototyping on Ti6Al4V alloys. North Carolina State University,
2012.
[17] K. Karayagiz, A. Elwany, G. Tapia, B. Franco, L. Johnson, J. Ma, I. Kara-
man, and R. Arróyave, “Numerical and experimental analysis of heat dis-
tribution in the laser powder bed fusion of ti-6al-4v,” IISE Transactions,
vol. 51, no. 2, pp. 136–152, 2019.
[18] J. Yin, H. Zhu, L. Ke, W. Lei, C. Dai, and D. Zuo, “Simulation of tempera-
ture distribution in single metallic powder layer for laser micro-sintering,
Computational Materials Science, vol. 53, no. 1, pp. 333–339, 2012.
[19] I. Yadroitsev, Selective laser melting: Direct manufacturing of 3D-objects
by selective laser melting of metal powders. LAP LAMBERT Academic
Publishing, 09 2009.
[20] T. Inoue, M. Nakano, T. Kubo, S. Matsumoto, and H. Baba, “High accuracy
control of a proton synchrotron magnet power supply,IFAC Proceedings
Volumes, vol. 14, no. 2, pp. 3137–3142, 1981.
[21] X. Chen and M. Tomizuka, “New repetitive control with improved steady-
state performance and accelerated transient,” IEEE Transactions on Control
Systems Technology, vol. 22, no. 2, pp. 664–675, 2014.
[22] I. Yadroitsev, P. Krakhmalev, and I. Yadroitsava, “Selective laser melting
of ti6al4v alloy for biomedical applications: Temperature monitoring and
microstructural evolution,Journal of Alloys and Compounds, vol. 583, pp.
404–409, 2014.
[23] A. J. Dunbar, E. R. Denlinger, M. F. Gouge, T. W. Simpson, and P. Micha-
leris, “Comparisons of laser powder bed fusion additive manufacturing
builds through experimental in situ distortion and temperature measure-
ments,” Additive Manufacturing, vol. 15, pp. 57–65, 2017.
... Linear finite elements were used to approximate solutions to sets of PDE such as convection and beam equations. The study held in [14] presents a simulation framework that uses the output signals retrieved from a FEM-based solution of heat flow equations.This would allow for the testing of different control strategies on an additive manufacturing process in a virtual environment. ...
... where squared "∆" terms vanish. Therefore, our optimal control procedure consists in adapting the perturbation method proposed in section III-D, to the solution of the nonlinear ODE system (10): (1) approximate the solution of the linearized system (10) with the same boundary values as for the nonlinear case, and let the solutions be called (q i 1 , u 1 i ); (2) approximate the solution of the ODE system composed of the perturbed motion equations (14) and the perturbed control equations (15) with vanishing boundary values, and let the solutions be called ...
... for s ← 1, w do w is sufficiently large 5: Insert (q i s , u s i ) into the perturbed motion equations (14) and perturbed control equations (15) 6: ...
Article
Full-text available
We used an optimal control method involving covariant control equations as optimality conditions, to command the actuators of robot manipulators. These form a coupled system of second order nonlinear ordinary differential equations when associated with the robot motion equations. By solving this system, the control action required to take the robot from an initial to a final state is optimized in a prescribed time. However, the target set of equations exhibited stiffness. Therefore, an adequate solution could only be found for short trajectory durations with readily available numerical methods. We examined a time discretization procedure based on cubic and quintic Hermite finite elements which exhibited superconvergence properties for interpolation. This motivated us to develop a time integration algorithm based on Hermite’s technique, where motion and control equations were perturbed to solve the optimal control problem. The optimal motion of a robotic manipulator was simulated using this algorithm. Our method was compared with a commercial differential equations solver on the basis of specific indicators. It outperformed the commercial solver by effectively solving the stiff set of equations for longer trajectory durations, with the cubic elements performing better than the quintic ones in this sense. The convergence analysis of our method confirmed that the quintic elements are more precise at the cost of increased computational burden, but converge at a lower rate than expected. Controlled motion experiments on a robotic manipulator validated our methodology. Trajectories were smoothly tracked and results exposed further methodology improvements.
... The developed FEM has been validated experimentally and analytically in [22] and serves as a simulation platform in this paper. Later on the data generated from the FEM such as melt pool width will be used to identify and verify the accuracies of the proposed models. ...
Preprint
Full-text available
Despite the advantages and emerging applications, broader adoption of powder bed fusion (PBF) additive manufacturing is challenged by insufficient reliability and in-process variations. Finite element modeling and control-oriented modeling have been shown to be effective for predicting and engineering part qualities in PBF. This paper first builds a finite element model (FEM) of the thermal fields to look into the convoluted thermal interactions during the PBF process. Using the FEM data, we identify a novel surrogate system model from the laser power to the melt pool width. Linking a linear model with a memoryless nonlinear sub-model, we develop a physics-based Hammerstein model that captures the complex spatiotemporal thermomechanical dynamics. We verify the accuracy of the Hammerstein model using the FEM and prove that the linearized model is only a representation of the Hammerstein model around the equilibrium point. Along the way, we conduct the stability and robustness analyses and formalize the Hammerstein model to facilitate the subsequent control designs.
... The developed FEM has been validated experimentally and analytically in [22] and serves as a simulation platform in this paper. Later on the data generated from the FEM such as melt pool width will be used to identify and verify the accuracies of the proposed models. ...
Article
Full-text available
Despite the advantages and emerging applications, broader adoption of powder bed fusion (PBF) additive manufacturing is challenged by insufficient reliability and in-process variations. Finite element modeling and control-oriented modeling have been identified fundamental for predicting and engineering part qualities in PBF. This paper first builds a finite element model (FEM) of the thermal fields to look into the convoluted thermal interactions during the PBF process. Using the FEM data, we identify a novel surrogate system model from the laser power to the melt pool width. Linking a linearized model with a memoryless nonlinear submodel, we develop a physics-based Hammerstein model that captures the complex spatiotemporal thermomechanical dynamics. We verify the accuracy of the Hammerstein model using the FEM and prove that the linearized model is only a representation of the Hammerstein model around the equilibrium point. Along the way, we conduct the stability and robustness analyses and formalize the Hammerstein model to facilitate the subsequent control designs.
... The developed FEM has been validated experimentally and analytically in [21] and serves as a simulation platform in this paper. Later on the data generated from the FEM such as melt pool width will be used to identify and verify the accuracies of the proposed models. ...
Preprint
Full-text available
Despite the advantages and emerging applications, broader adoption of powder bed fusion (PBF) additive manufacturing is challenged by insufficient reliability and in-process variations. Finite element modeling and control-oriented modeling have been identified fundamental for predicting and engineering part qualities in PBF. This paper first builds a finite element model (FEM) of the thermal fields to look into the convoluted thermal interactions during the PBF process. Using the FEM data, we identify a novel surrogate system model from the laser power to the melt pool width. Linking a linearized model with a memory-less nonlinear submodel, we develop a physics-based Hammer-stein model that captures the complex spatiotemporal thermomechanical dynamics. We verify the accuracy of the Hammerstein model using the FEM and prove that the linearized model is only a representation of the Hammerstein model around the equilibrium point. Along the way, we conduct the stability and robustness analyses and formalize the Hammerstein model to facilitate the subsequent control designs.
Article
Full-text available
This paper discusses fractional-order repetitive control (RC) to advance the quality of periodic energy deposition in laser-based additive manufacturing (AM). It addresses an intrinsic RC limitation when the exogenous signal frequency cannot divide the sampling frequency of the sensor, e.g., in imaging-based control of fast laser-material interaction in AM. Three RC designs are proposed to address such fractional-order repetitive processes. In particular, a new multirate RC provides superior performance gains by generating high-gain control exactly at the fundamental and harmonic frequencies of exogenous signals. Experimentation on a galvo laser scanner in AM validates effectiveness of the designs.
Article
Full-text available
Laser powder bed fusion (LPBF) of metallic parts is a complex process involving simultaneous interplay between several physical mechanisms such as solidification, heat transfer (convection, conduction, radiation, etc.), and fluid flow. In the present work, a three-dimensional finite element (FE) model is developed for studying the thermal behavior during LPBF of Ti-6 Al-4 V alloy. Two phase transitions are considered in the model: solid-to-liquid and liquid-to-gas. It is demonstrated that metal evaporation has a notable effect on the thermal history evolution during fabrication and should not be overlooked in contrast to the majority of previous research efforts on modeling and simulation of additive manufacturing processes. The model is validated through experimental measurements of different features including the size and morphology of the heat affected zone (HAZ), melt pool size, thermal history. Reasonable agreement with experimental measurements of the HAZ width and depth are obtained with corresponding errors of 3.2% and 10.8%. Qualitative agreement with experimental measurements of the multi-track thermal history is also obtained, with some discrepancies whose sources are discussed in detail. The current work presents one of the first efforts to validate the multi-track thermal history using dual-wavelength pyrometry, as opposed to single track experiments. The effects of selected model parameters and evaporation on the melt pool/HAZ size, geometry and peak predicted temperature during processing, and their sensitivities to these parameters are also discussed. Sensitivity analysis reveals that thermal conductivity of the liquid phase, porosity level of the powder bed, and absorptivity have direct influence on the model predictions, with the influence of the thermal conductivity of the liquid phase being most significant.
Article
Full-text available
Overhanging and floating layers which are introduced during the build in selective laser melting (SLM) process are usually associated with high temperature gradients and thermal stresses. As there is no underlying solid material, less heat is dissipated to the powder bed and the melted layer is free to deform resulting undesired effects such as shrinkage and crack. This study uses three-dimensional finite element simulation to investigate the temperature and stress fields in single 316L stainless steel layers built on the powder bed without support in SLM. A non-linear transient model based on sequentially coupled thermo-mechanical field analysis code was developed in ANSYS parametric design language (APDL). It is found that the predicted length of the melt pool increases at higher scan speed while both width and depth of the melt pool decreases. The cyclic melting and cooling rates in the scanned tracks result high VonMises stresses in the consolidated tracks of the layer.
Article
Full-text available
In repetitive control (RC), the enhanced servo performance at the fundamental frequency and its higher order harmonics is usually followed by undesired error amplifications at other frequencies. In this paper, we discuss a new structural configuration of the internal model in RC, wherein designers have more flexibility in the repetitive loop-shaping design, and the amplification of nonrepetitive errors can be largely reduced. Compared to conventional RC, the proposed scheme is especially advantageous when the repetitive task is subject to large amounts of nonperiodic disturbances. An additional benefit is that the transient response of this plug-in RC can be easily controlled, leading to an accelerated transient with reduced overshoots. Verification of the algorithm is provided by simulation of a benchmark regulation problem in hard disk drives, and by tracking-control experiments on a laboratory testbed of an industrial wafer scanner.
Book
Full-text available
This study concerns the interaction of the powerful laser radiation (0.3-1.3 MW/sq cm) with powder metallic materials. Thermal processes in interaction zones, absorption, reflection, radiation transfer, heat transfer by conduction and convection, evaporation and emission of material are discussed in detail. Numerical modeling of laser interaction with substrate and laser-powder interaction is presented. For commercial metal powders (stainless steel, tool steel, nickel-chromium alloy, cobalt- base and copper alloys, titanium and nickel-titanium alloys) are defined the relations between the basic parameters of selective laser melting and their influence on structure and functional properties of manufactured objects. Based on the results of the present research, net shape functional parts with engineered internal structures, desired porosity, specially structured surface and multi-material objects are fabricated.
Article
A continuum-scale modeling approach is developed and employed with three-dimensional finite element analysis (FEA), for simulating the temperature response of a Ti-6Al-4V, two-layered parallelepiped with dimensions of 10 × 5 x 0.06 mm³ during Laser Powder Bed Fusion (L-PBF), a metals additive manufacturing (AM) method. The model has been validated using experimental melt pool measurements from the literature and also accounts for latent heat of fusion and effective, temperature-dependent transport properties. The discretized temperature, temperature time rate of change (i.e. cooling rate) and temperature gradient are investigated for various scan strategies and number of lasers, i.e. 1, 2 or 4. The thermal response inherent to multi-laser PBF (ML-PBF) is investigated. The number of sub-regional areas of the powder bed dedicated to individual lasers, or ‘islands’, was varied. The average, maximum cooling rate and temperature gradient per layer, as well as the spatial standard deviation, or uniformity, of such metrics, are presented and their implications on microstructure characteristics and mechanical traits of Ti-6Al-4V are discussed. Results demonstrate that increasing the number of lasers will reduce production times, as well as local cooling rates and residual stress magnitudes; however, the anisotropy of the residual stress field and microstructure may increase based on the scan strategy employed. In general, scan strategies that employ reduced track lengths oriented parallel to the part's shortest edge, with islands ‘stacked’ in a unit-row, proved to be most beneficial for L-PBF.
Article
In situ experimental measurements of the laser powder bed fusion build process are completed with the goal gaining insight into the evolution of distortion in the powder bed fusion build process. Utilizing a novel enclosed instrumented system, five experimental builds are performed. Experimental builds compare materials: Ti-6Al-4V and Inconel® 718, differing build geometries, and manufacturing machines: EOS M280 and Renishaw AM250. A combination of in situ measurements of distortion and temperature and post-build measurements of final part geometry are used to compare and contrast the different experiments. Experimental results show that builds completed using Inconel® 718 distort between 50%-80% more relative to Ti-6Al-4V depending on substrate size and build geometry. The experimental build completed on the Renishaw AM250 distorted 10.6% more in the Z direction when compared with the identical build completed on the EOS M280 machine. Comparisons of post-build XY cross-sectional area show a 0.3% contraction from the predefined build geometry for the Renishaw AM250 as compared with the 4.5% contraction for the part built using the EOS M280. Recommendations and future work are also discussed.
Article
The Laser Metal Deposition (LMD) process is an additive manufacturing process in which a laser and a powdered material source are used to build functional metal parts in a layer by layer fashion. While the process is usually modeled by purely temporal dynamic models, the process is more aptly described as a repetitive process with two sets of dynamic processes: one that evolves in position within the layer and one that evolves in part layer. Therefore, to properly control the LMD process, it is advantageous to use a model of the LMD process that captures the dominant two dimensional phenomena and to address the two-dimensionality in process control. Using an identified spatial-domain Hammerstein model of the LMD process, the open loop process stability is examined. Then, a stabilizing controller is designed using error feedback in the layer domain.
Article
A laser metal deposition height control methodology is presented in this paper. The height controller utilizes a particle swarm optimization (PSO) algorithm to estimate model parameters between layers using measured temperature and track height profiles. Using the estimated model, the powder flow rate reference profile, which will produce the desired layer height reference, is then generated using iterative learning control (ILC). The model parameter estimation performance using PSO is evaluated using a four-layer single track deposition, and the powder flow rate reference generation performance using ILC is tested using simulation. The results show that PSO and ILC perform well in estimating model parameters and generating powder flow rate references, respectively. The proposed height control methodology is then tested experimentally for tracking a constant height reference with constant traverse speed and constant laser power. The experimental results indicate that the controller performs well in tracking constant height references in comparison with the widely used fixed process parameter strategy. The application of layer-to-layer height control produces more consistent layer height increment and a more precise track height, which saves machining time and increases powder efficiency.