Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Direct measurements of laser absorptivity

during metal melt pool formation


associated with powder bed fusion additive
manufacturing processes
Cite as: J. Laser Appl. 30, 032302 (2018); https://doi.org/10.2351/1.5040636
Submitted: 18 May 2018 . Accepted: 18 May 2018 . Published Online: 13 June 2018

Manyalibo Matthews, Johannes Trapp, Gabe Guss, and Alexander Rubenchik

COLLECTIONS

Paper published as part of the special topic on Proceedings of the International Congress of Applications of Lasers &
Electro-Optics (ICALEO® 2017)

ARTICLES YOU MAY BE INTERESTED IN

Laser powder bed fusion additive manufacturing of metals; physics, computational, and
materials challenges
Applied Physics Reviews 2, 041304 (2015); https://doi.org/10.1063/1.4937809

An instrument for in situ time-resolved X-ray imaging and diffraction of laser powder bed
fusion additive manufacturing processes
Review of Scientific Instruments 89, 055101 (2018); https://doi.org/10.1063/1.5017236

Review of selective laser melting: Materials and applications


Applied Physics Reviews 2, 041101 (2015); https://doi.org/10.1063/1.4935926

J. Laser Appl. 30, 032302 (2018); https://doi.org/10.2351/1.5040636 30, 032302

© 2018 Laser Institute of America.


JOURNAL OF LASER APPLICATIONS VOLUME 30, NUMBER 3 AUGUST 2018

Direct measurements of laser absorptivity during metal melt pool formation


associated with powder bed fusion additive manufacturing processes
Manyalibo Matthews,1 Johannes Trapp,2 Gabe Guss,1 and Alexander Rubenchik1
1
Lawrence Livermore National Laboratory, Livermore, California 94550
2
Technische Universität Dresden, Dresden, Germany 01062

(Received 18 May 2018; accepted for publication 18 May 2018; published 13 June 2018)
Direct calorimetric measurements are used to study the effective optical absorptivity at 1070 nm
laser wavelength for bulk and powder-coated discs of industrially relevant metals. Effective absorp-
tivity is plotted as a function of nominal laser power from 30 up to 600 W for scanning velocities
of 100, 500, and 1500 mm/s. The absorptivity versus power curves of the bulk materials typically
shows a reduction in effective absorptivity until the beginning of the formation of a keyhole-type
surface depression that is associated with an increased absorption of the laser light in the growing
keyhole until a saturation value is reached. For powders, an additional plateau of higher absorptivity
can be observed for low laser power, until the curves qualitatively collapse when full melting of the
powder tracks is achieved. It is shown that, under conditions associated with laser powder-bed
fusion additive manufacturing, absorptivity values can vary greatly, and differ from both room
temperature powder layer measurements and liquid metal estimates from the literature. © 2018
Laser Institute of America. https://doi.org/10.2351/1.5040636

Key words: additive manufacturing, 3D printing, selective laser melting, powder bed fusion, optical
absorption, laser keyhole, metal powder, metal liquid absorptivity

I. INTRODUCTION
complex process including a variety of physical effects. The
Metal additive manufacturing through selective laser melt surface interaction with the laser is nonstationary, modu-
melting—also known as selective laser sintering, laser beam lated due to the melt motion which can affect the absorptivity.10
melting, or laser powder-bed fusion additive manufacturing At higher laser intensities, when the recoil momentum is strong
(PBFAM)—produces metal parts via layer by layer deposition enough to create a deep surface depression in the melt pool (the
and selective fusion of said powder layers using a laser beam. keyhole), light interacts with the keyhole walls at steep angles
The gain of design freedom by this process is complicated by and with the ejected vapors. A part of the absorbed energy is
the process complexity. It has been indicated previously1 that ejected with the vapors and a part of the laser light is absorbed
the over 130 parameters could affect the final manufacturing in the vapor plume. An adequate and self-consistent model for-
process, thus establishing a challenging phase space over mulation of these complex effects is extremely difficult and
which process optimization is to be explored. Optimization of remains a challenge to current modeling efforts.
these parameters is required for defect-free part production. In the present work, we demonstrate direct calorimetric
Numerical modeling provides a powerful tool for estimat- measurement to determine the effective value of laser
ing the optimal set of system parameters for specific part geom- absorptivity of metal surfaces and powder layers across a
etries and materials. Today, models describing heat transport wide range of temperatures. Estimated peak temperatures
and hydrodynamics includes the laser melting of the metal range from below the melting point up to above the boiling
powder, the melt pool structure, and dynamics under the effect point. Measurements are achieved by performing single and
of capillary forces and recoil pressure and evaporation.2,3 The multi-track experiments and measuring the resulting tempera-
thermal histories predicted by these models will ultimately ture in the specimen for different scanning velocities, includ-
inform thermodynamic and microstructural models which can ing conditions of the typical PBFAM process. At low laser
be related to final part properties and performance. However, powers, absorptivity values are in agreement with literature
practically in all models describing the laser-matter interaction data at surface absorptivity. At high laser powers, we
of the laser beam with the powder layer surface, the absorption observed a steep absorptivity increase associated with the
of laser light is typically treated as constant, based on either transition to keyhole-mode laser coupling. After taking into
calculated or measured values for polished clean metal plates4,5 account the impact of the scanning velocity, the effects of a
as well as rough6,7 and oxidized surfaces. The relevant absorp- thin powder layer on the effective absorption of the laser
tivity for powders was measured at low homologous tempera- light are demonstrated. At low powers, the absorptivity is
tures8,9 and calculated using a ray tracing approach.8 In significantly higher relative to bare plates with this result
practice, once the temperature of the metallic surface heated by being in agreement with the independent powder absorptivity
the laser reaches the melting point, the laser absorption is a measurements at low temperatures. At higher power, the

1042-346X/2018/30(3)/032302/4/$28.00 032302-1 © 2018 Laser Institute of America


032302-2 J. Laser Appl., Vol. 30, No. 3, August 2018 Matthews et al.

curves for a bare plate and for the powder-coated surface A 600 W Yb-fiber laser was focused to a 60 ± 5 μm 1/e2
qualitatively collapse and display evidence of strong keyhole diameter Gaussian spot and scanned using x–y galvanometer
absorption behavior similar to the case of the bare discs. scanning mirrors. This optical system was designed to closely
approximate that of a typical commercial PBFAM system.
The laser scans created single track or multiple consecutive
II. EXPERIMENTAL DETAILS tracks on the disc or powder surface. The temperature–time
curve was measured for a total time of 60 s with the thermo-
Rolled metal sheets of 316L stainless steel, tungsten
couples operating with 14 Hz. We fit the exponential decay of
(99.95% purity) (both 0.5 mm thick), and aluminum 1100 (0.6
temperature over time for both thermocouples from this time
mm thick) were obtained from Goodfellow Ltd (Huntington,
on and extrapolated the temperature to the time 0 when the
UK). The powder used was commercially available alloy 20ES
laser created the tracks. This procedure was chosen to com-
(316L stainless steel) from Concept Laser (Lichtenfels,
pensate the temperature losses before the temperature equili-
Germany). The mean particle size was 29.9 ± 9 μm for the
bration over the disc (which we attribute mostly to convective
steel powder.
losses as a comparison of the temperature–time curves with a
For the measurements using blank discs, the metal sheets
COMSOL simulation of the process shows). The total time of
were laser cut to discs of 10 mm diameter. The absorptivity of
energy deposition was in the milliseconds range, much
powders was measured using cups as depicted in Fig. 1(c)
shorter than the temperature equilibration time.
that were precision machined from the discs with the rim
With the disc temperature increase known, the effective
height of 100 μm defining the thickness of the powder layer.
absorptivity can be calculated by dividing the energy neces-
The experiments were carried out in a custom-made
sary to uniformly heat the specimen from starting tempera-
build chamber10 with an extensive suite of diagnostic tools
ture T0 to end temperature T1 (EH) by the energy input due
and diagnostic access to the melting pool as well as the
to the laser irradiation (EL):
ability to apply a controlled atmosphere and adjust the flow
rate of the protective argon gas used. The flow rate of argon Ð T1
was set to a fixed value of 2150 and 715 SCCM for the EH T0 mCp (T) dt
Aeff ¼ ¼ , (1)
blank discs and the powder samples, respectively, resulting EL P(l=v)
in a gas pressure slightly above atmospheric pressure.
A key design component of our calorimetric setup is suf- whereby m is the mass of the disc including the powder
ficient thermal isolation of the sample during measurements. layer (if applicable), P is the nominal laser power, v, the
A sample holder custom made from porous alumina (0.14 scanning velocity, and l, the total length of the laser track
W/mK thermal conductivity at 25°C) by Foundry Service & (6 mm for a single track line). The temperature-dependent heat
Supplies, Inc. (Ontario, CA) as shown in Figs. 1(a) and 1(b) capacity at constant pressure was taken from Refs. 11,12.
was designed to minimize the conductive cooling of the
sample well below the convective heat loss. Type K thermo-
couples with blank wires of 76 μm diameter each were spot III. RESULTS and DISCUSSION
welded to the rear side of the discs/cups, one thermocouple We begin with results for the absorptivity measurements
centered, and one about 3 mm away from the center. Thin for flat tungsten discs. The absorptivity dependence on
gold foils of 20 μm thickness were used as an interlayer to power for scanning velocities of 100, 500, and 1500 mm/s is
weld the thermocouples to the aluminum discs. The wires presented in Fig. 2. The curve can be divided into three
were then insulated using alumina tubes. For the experiments parts. At low laser power, the absorptivity slightly decreases
with powder, the powder was spread on the cups by use of a
razor blade with a thickness of about 100 μm defined by the
rim height of the cups.

FIG. 1. Calorimetry setup inside of the process chamber (a); a drawing of


the sample holder made from porous Al2O3 with a mounted metal cup (b), FIG. 2. Effective absorptivity of tungsten discs as a function of laser power
as well as a magnified cross section of the cup filled with powder (c). for scan speeds 100, 500, and 1500 mm/s.
J. Laser Appl., Vol. 30, No. 3, August 2018 Matthews et al. 032302-3

with increasing power. At a certain laser power, the absorp-


tivity curves demonstrate a fast growth to values over two
times that observed at low power and finally saturates at
about 0.7. This onset of absorptivity increase appears to shift
and broaden with increasing laser scan speed.
To understand the physics behind the absorptivity
changes, we cut the tracks perpendicular to the scanning
direction (not shown). We observed two different melt pool
geometries: for low power, a half-spherical shape can be
observed that is a result of thermal conduction of the heat
starting to diffuse into the sample from the laser irradiated
surface. With the growth of this melt pool, the effective
absorptivity slightly decreases. Coinciding with the onset of
growing absorptivity, the melt pool depth increases and a
keyhole is formed. The growth in depth of the keyhole leads
to further increasing absorptivity. After reaching a depth of
about 300 μm, the absorptivity saturates. FIG. 3. Effective absorptivity of aluminum discs as a function of laser
Initially, when the laser power is increased, there is a power for scan speeds 100, 500, and 1500 mm/s.
decrease in measured absorptivity. A melt pool is observed to
form even for the lowest power (30 W) and highest scanning power the absorptivity of the powder is about two times
velocity (1500 mm/s); heat loss due to evaporation of material higher than the surface absorptivity due to the light that is
or the absorption/redirection of laser light in the plume gener- multiply scattering in the powder.13 As laser power increases
ated before the light hits the metallic surface are possible the absorptivity drops, due to the partial powder melting
reasons. The necessary amount of material to compensate making the surface more flat and reducing the scattering
reduced absorptivity for 500 mm/s is about 8 or 1 μm of mater- events of the laser light. With a growing portion of the
ial, respectively. As the inside temperature and the extent of powder being molten with increasing power starting from the
the weld pool on the discs surface is approximately equal for bottom to the top of the picture, the absorptivity decreases.
all scanning velocities when the onset of increased absorptivity As soon as the recoil pressure starts to form a keyhole, the
is reached (viz., the temperature reaches the boiling point), the absorptivity again grows fast due to the multiple reflections
rate of evaporation should be equal for all scanning velocities occurring in the keyhole depression (see inset of Fig. 4). It is
as well. Therefore, the portion of energy lost due to evapor- interesting to note that absorptivity power dependence for
ation of material referred to the energy input must remain the the powder-coated case at high intensities is qualitatively
same for all scanning velocities, as both energy input (power similar to that of the flat plate. However, as reported recently,
times dwell time) and evaporative loss (evaporation rate times the interaction of entrained powder particles that leads to
dwell time) scale linear with scanning velocity. A decrease in material “spatter” further differentiates the laser interaction
absorptivity caused by evaporative losses should, therefore, be between powder layers and bare plates.14
independent on scanning velocity. As this is not the case, those Different from the steel samples, for aluminum and tung-
losses are concluded to be less relevant than absorption/reflec- sten, the difference in the absorptivity of (1) solid and liquid
tion of laser light in the plume. As the plume is more and more
tilted (and interaction therefore reduced), the faster the scan-
ning velocity could allow for a slight increase in absorptivity.
Shown in Fig. 3 is the effective absorptivity of aluminum
discs as a function of laser power for scan speeds 100, 500,
and 1500 mm/s. For scan speeds 100 and 500 mm/s, oxidation
of the surface caused additional heating for laser powers
below ∼200 W, and therefore, this data are not shown.
Consistent with the results for tungsten, a slight decrease in
absorptivity as a function of laser power is observed at low
powers, followed by a steep increase in absorptivity at the
onset of keyhole-mode absorption above ∼200 W. At the
slower speeds (100, 500 mm/s), disc burn-through prevented
higher powers to be accurately measured, leaving only the
absorptivity onset region observable in Fig. 3. Unlike the case
for tungsten, the low power data points appeared slightly
higher than the room temperature literature values for the
material, likely due to the higher reactivity of aluminum and
FIG. 4. Effective absorptivity of bare and powder-coated steel 316L discs as
absorptivity of oxide-containing melt pools.
a function of laser power for a scan speed of 1500 mm/s. A schematic of the
The absorptivity for the bare and powder-coated steel keyhole formation above a critical laser power Pc and corresponding
316L substrate is presented in Fig. 4 for 1500 mm/s. At low increased effecting light absorption is shown in the inset.
032302-4 J. Laser Appl., Vol. 30, No. 3, August 2018 Matthews et al.

metal and (2) an oxide-coated surface compared to a pure ACKNOWLEDGMENT


metallic surface adds to the changes in absorptivity for
This work was performed under the auspices of the U.S.
low power. The observed reduction of effective absorptivity
Department of Energy by Lawrence Livermore National
with increasing power is likely a result of the increasing
Laboratory under Contract No. DE-AC52-07NA27344.
destruction of the oxide layer with its relatively high
absorptivity. 1
I. Yadroitsev, I. Yadroitsava, P. Bertrand, and I. Smurov, “Factor analysis
The onset of increased absorptivity correlates to the of selective laser melting process parameters and geometrical character-
beginning deep penetration regime (the keyhole regime). To istics of synthesized single tracks,” Rap. Prototyp. J. 18, 201–208 (2012).
2
S. A. Khairallah, A. T. Anderson, A. Rubenchik, and W. E. King, “Laser
support this observation, we calculated the peak temperatures
powder-bed fusion additive manufacturing: Physics of complex melt flow
and compared the power necessary to reach the boiling point and formation mechanisms of pores, spatter, and denudation zones,” Acta
which is a necessary prerequisite for the keyhole formation Mater. 108, 36–45 (2016).
3
to the onset of increased absorptivity.15 It was found that the W. King, A. T. Anderson, R. M. Ferencz, N. E. Hodge, C. Kamath, and
S. A. Khairallah, “Overview of modelling and simulation of metal powder
laser power required for boiling was less than that observed bed fusion process at Lawrence Livermore National Laboratory,” Mater.
for keyhole onset in all cases. We believe the difference is Sci. Technol. 31, 957–968 (2015).
due to the excess energy required to deform the meltpool 4
F. Dausinger and J. L. Shen, “Energy coupling efficiency in laser-surface
and increase the surface energy due to capillarity. Further treatment,” ISIJ Int. 33, 925–933 (1993).
5
B. Karlsson and C. G. Ribbing, “Optical-constants and spectral selectivity
work to understand this effect is underway. of stainless-steel and its oxides,” J. Appl. Phys. 53, 6340–6346 (1982).
6
D. Bergstrom, J. Powell, and A. F. H. Kaplan, “The absorption of light by
rough metal surfaces—A three-dimensional ray-tracing analysis,” J. Appl.
Phys. 103 (2008).
IV. CONCLUSIONS 7
A. M. Rubenchik, S. S. Q. Wu, V. K. Kanz, M. M. LeBlanc, W. H.
Lowdermilk, M. D. Rotter, and J. R. Stanley, “Temperature-dependent
The effective optical absorptivity of tungsten, aluminum, 780-nm laser absorption by engineering grade aluminum, titanium, and
and steel discs as a function of laser power (30–600 W) and steel alloy surfaces,” Opt. Eng. 53, 122506 (2014).
8
scan speed (100, 500, and 1500 mm/s) was measured using N. K. Tolochko, T. Laoui, Y. V. Khlopkov, S. E. Mozzharov, V. I. Titov,
and M. B. Ignatiev, “Absorptance of powder materials suitable for laser
microcalorimetry. The absorptivity of steel powder layers as sintering,” Rap. Prototyp. J. 6, 155–160 (2000).
a function of laser power was also measured at a scan speed 9
C. Boley, S. Mitchell, A. Rubenchik, and S. Wu, “Metal powder absorp-
of 1500 mm/s. Low power results agreed mostly with litera- 10
tivity: modeling and experiment,” Appl. Opt. (2016).
ture values of solid or liquid metal absorptivity measure- M. J. Matthews, G. Guss, S. A. Khairallah, A. M. Rubenchik, P. J.
Depond, and W. E. King, “Denudation of metal powder layers in laser
ments. However, at high powers the onset of a vapor recoil powder bed fusion processes,” Acta Mater. 114, 33–42 (2016).
keyhole depression causes a sharp increase in the effective 11
W. F. Gale, T. C. Totemeier (Eds.), Smithells Metals Reference Book,
optical absorptivity, which eventually saturates to values in (Elsevier Butterworth-Heinemann, Oxford, Burlington, 1992).
12
K. C. Mills, Recommended Values of Thermophysical Properties for
the range 0.6–0.8 for all materials. For the case of metal
Selected Commercial Alloys (Woodhead Publishing, Cambridge, U.K., 2002).
powder (steel 316L), an additional transition from powder 13
C. Boley, S. Khairallah, and A. Rubenchik, “Calculation of laser absorption
absorptivity values to liquid and then to keyhole exemplifies by metal powders in additive manufacturing,” Appl. Opt. 54, 2477 (2015).
14
the very strong laser power dependence of energy coupling S. Ly, A. M. Rubenchik, S. A. Khairallah, G. Guss, and M. J. Matthews,
“Metal vapor micro-jet controls material redistribution in laser powder bed
associated with laser power bed fusion additive manufactur- fusion additive manufacturing,” Sci. Rep. 7, 4085 (2017).
ing and may explain the sensitivity of final part quality to 15
T. W. Eagar and N. S. Tsai, “Temperature-fields produced by traveling
laser parameters. distributed heat-sources,” Weld. J. 62, S346–S355 (1983).

You might also like