You are on page 1of 28

11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics

Laser Surface Treatment


Laser surface treatment allows the refinement of microstructures,
dissolution of inclusions and precipitates and formation of non-
equilibrium supersaturated solid solutions, quasi-crystalline and
amorphous materials [2], microstructural changes often leading to
improved corrosion resistance.
From: Laser Surface Modification of Alloys for Corrosion and Erosion Resistance,
2012

Related terms:

Phase Change, Nitrogen, Evaporation, Residual Stress, Corrosion Resistance,


Surface Melting, Surface Roughness, Alloying

Characterization and modification of technical ceramics


through laser surface engineering
P. Shukla, J. Lawrence, in Laser Surface Engineering, 2015

5.1 Introduction
Laser surface treatment of materials is an important technique because it offers a
possibility to enhance various properties such as the surface strength, hardness,
roughness, coefficient of friction, chemical resistance, and corrosion of various
materials. Such improvements to a material surface are not only ideal for
applications when wear rate and shear stresses are high but could also be used for
maintaining or elongating the component’s functional life by means of covering
over the microcracks in surfaces, such as those of technical ceramic-based
components. In addition, aesthetics can also be improved using laser surface
treatment (for ceramics in particular) by creating a modified surface layer.
Research has advanced in the field of laser machining of ceramics [1–5], surface
treatment [6–16], laser cutting [4,15,17–25], and drilling of ceramics [26–29]. But
the effects of laser processing of various advanced ceramics are still unknown and
are not fully reported in most of the published literature. The unknown aspects, in
particular, are the effects during the laser's interaction with oxide and nitride
technical grade ceramics. The interaction of the high brightness, and fairly novel,
near-infrared (NIR) wavelength fiber laser with such technical ceramics is still not
understood from the viewpoint of a compositional, microstructural, internal
phases, and mechanical aspects. In this chapter, laser surface treatment process
parameters are described for the ceramic material systems used herein, followed by
an analysis of the surface finish, material removal, chemical composition, K1c and
the distribution of surface temperature. This should lead to a better understanding
of the laser beam-ceramic interactions for future applications, such as laser-
assisted joining, because joining technical ceramics-to-ceramics would be useful as
an alternative process that could give an opportunity for new applications. On

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tre… 1/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics
account of this, the interaction of the laser beam with selected ceramics is
presented to understand how the laser surface treatment of ceramics can be
improved as a process. Furthermore, it is generally difficult to characterize
technical ceramics and, particularly if they are laser processed, because the laser-
engineered surface becomes amorphous and glassy. This, in turn, does not allow
them to be characterized by methods such as X-ray diffraction, which is a
conventional tool for the determination of the composition, residual stress, and
phase transformation. Having said that, this chapter also indicates the alternative
characterization methods that can be used to analyze the laser engineered technical
grade ceramics.
The ZrO2 and the Si3N4 technical ceramics were selected as they both have
different mechanical and thermal properties. This, in turn, will create different
effects after applying laser surface treatment. Furthermore, both the ZrO2 and
Si3N4 technical ceramics are frequently used for many applications within the
engineering, medical, and power generation sectors, so undertaking this research
contributes to achieving further diversity of current and future applications for the
two ceramics employed.

Read full chapter


URL: https://www.sciencedirect.com/science/article/pii/B9781782420743000052

Corrosion Protective Coatings for Ti and Ti Alloys Used


for Biomedical Implants
Liana Maria Muresan, in Intelligent Coatings for Corrosion Control, 2015

17.4 Laser Oxidation


Laser surface treatment is a thermal process that has an advantage over
conventional furnace treatment. It is based on the heating caused by the light
adsorption of the surface layer and the cooling ensured by the high conductivity of
the material.31 The adsorbed laser energy results in a thin surface layer with
desired properties while the bulk of the material is unaffected.
By using laser treatments followed by AFM investigations, it was proved that the
morphological modifications of Ti-based alloys play a key role in cell adherence and
proliferation by influencing cell-material interactions.32
Porous Ti samples were fabricated by using laser power in an attempt to improve
the osteoconductivity and reduce the problems associated with stress shielding.5
NiTi were treated with a pulsed Nd:YAG laser in air, by using selected parameters
aiming to obtain uniform oxide films of controlled thickness.7 The corrosion
resistance of the so-treated NiTi samples increased about 15 times, while the
surface Ni/Ti ratio was reduced. The last feature is important as the high Ni
content in NiTi constitutes a potential threat to its safe use in vivo.

Read full chapter


URL: https://www.sciencedirect.com/science/article/pii/B9780124114678000179

Laser Machining and Surface Treatment

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tre… 2/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics
A.Z. Sahin, ... S.S. Akhtar, in Comprehensive Materials Processing, 2014

9.15.4 Results and Discussion


Laser surface treatment of alumina surface is carried out. The microstructural
changes and stress levels in the laser-treated region are examined for two power
levels. The thermal efficiency analysis is introduced to assess the influence of laser
power levels on the first and second law efficiencies of the laser treatment process
(Figure 1).
Temporal variation of temperature predicted from the simulations and obtained
from the thermocouple data is shown in Figure 2. It can be observed that the
temperature predicted agrees well with the thermocouple data. However, the small
discrepancies between both results are small, which is within the experimental
error.

Figure 2. Comparison of temporal variation of surface temperature predicted from


the simulations and obtained from the experiment.

Figure 3 shows temperature distribution along the x-axis for different cooling
periods and for two laser power intensities (power source A = 0.22 × 109 W m−2 and
power source B = 0.25 × 109 W m−2). It should be noted that the cooling cycle starts
at t = 0.05 s immediately after the laser beam power is ceased off. In this case, the
location of the laser beam is at x = 0.005 m away from the initial starting point
along the x-axis (laser scanning axis). Temperature attains its maximum at a
location of x = 0.005 m where the laser intensity is ceased. Temperature decay
along 0 ≤ x ≤ 0.005 m is gradual while it is sharp for x > 0.0005 m. The gradual
decay of temperature is associated with the heating of this region during laser
scanning prior to ceasing the laser power. Moreover, the sharp decay of
temperature in the frontal region of the laser spot (x > 0.005 m) results in a high
temperature gradient. This is particularly true for high laser power intensity. The
superheating of the liquid phase in the irradiated region is evident at t = 0.05 s
after the initiation of laser scanning. In this case, temperature in the irradiated
region exceeds the melting temperature of alumina. This situation is visible in the
temperature curve, at which the slope changes across the melting temperature. As
the cooling period progresses, temperature decays rapidly in the region where the
laser power is ceased, i.e., at x = 0.005 m. The rapid decay of temperature in this

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tre… 3/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics
region is related to the convective and radiative cooling of the surface and
conduction loss from the irradiated region to the solid bulk of the substrate
material. Moreover, the rate of temperature decay is more pronounced at high
laser power level since the surface temperature attains high value at the onset of
cooling.

Figure 3. Temperature distribution along the x-axis for two laser power intensities
(power source A = 0.22 × 109 W m−2 and power source B = 0.25 × 109 W m−2) and
for different cooling periods. The cooling period starts at t = 0.05 s after the laser
treatment is initiated.

Figure 4 shows von Mises stress along the x-axis for different cooling periods and
for two laser output powers. von Mises stress attains low values in the region of
high temperature during the early cooling periods, which is true for both laser
output powers. This occurs because of the elastic modulus of alumina, which
reduces considerably with increasing temperature (Table 2). In this case, low elastic
modulus at high temperatures results in low stress levels. Moreover, von Mises
stress attains high values in the frontal region of the laser-irradiated spot (x > 0.005
m). This is attributed to the attainment of high temperature gradient in this region.
Moreover, von Mises stress attains high values in the initially heated region (x ≤
0.005 m), which is associated with the development of high temperature gradient
during the laser scanning. As the cooling period progresses, temperature reduces
significantly (Figure 3) and von Mises stress becomes the residual stress. The
residual stress remains high along the x-axis, which is particularly true for high
laser output power intensity (0.25 × 109 W m−2). This is because of the
development of high temperature gradient during laser scanning along the x-axis,
i.e., high intensity beam results in high temperature gradient in the laser-
irradiated region. The maximum value of the residual stress is in the order of 3
GPa, which is considerably high and may cause cracks at the laser-treated surface.

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tre… 4/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics

Figure 4. von Mises stress distribution along the x-axis for two laser power
intensities (power source A = 0.22 × 109 W m−2 and power source B = 0.25 × 109 W
m−2) and for different cooling periods. The cooling period starts at t = 0.05 s after
the laser treatment is initiated.

Figure 5 shows the first law efficiency with the laser output power for different laser
scanning velocities. Increasing laser power reduces the first law efficiency almost
linearly, which is particularly true for high laser scanning velocities. In this case, the
slope of the first law efficiency decay becomes larger for high laser scanning
velocities. This is attributed to the energy deposition through absorption and
energy used for the laser treatment process. It should be noted that increasing
scanning speed reduces the rate of laser energy deposition per unit length at the
surface during the scanning process. In addition, the irradiated energy absorbed in
the surface region results in a temperature increase in this region. However, the
useful treated depth for the laser heating is limited with the depth of melting.
Consequently, when evaluating the first law efficiency, the useful energy is limited
to the energy required for melting in the surface region. Hence heat conduction
from the melted zone to the solid bulk reduces the first law efficiency. In addition,
convective and radiative losses from the surface also lower the first law efficiency.
Therefore, increasing laser power intensity enhances the conduction, convection,
and radiation losses while lowering the first law efficiency. Similarly reducing the
scanning speed lowers the first law efficiency, since the rate of laser energy
deposited via absorption increases at low scanning speeds. Although lowering laser
output power enhances the first law efficiency, further reduction in the laser power
ceases the melting process; in this case, laser treatment via melting replaces the
solid phase heating of the surface.

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tre… 5/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics

Figure 5. The first law efficiency with laser power for various values of laser
scanning speed.

Figure 6 shows the second law efficiency with the laser output power for different
laser scanning speeds. The second law efficiency reduces almost linearly with
increasing laser power, which is true for all laser scanning speeds used in the
calculations. This behavior of the second law efficiency is similar to that
corresponding to the first law efficiency. This is because of the energy losses from
the melted zone via conduction, convection, and radiation. Moreover, the second
law efficiency attains lower values than the first law efficiency. This is associated
with the rate of entropy generation during the laser treatment process, which
reduces the exergy required for melting in the surface region.

Figure 6. The second law efficiency with laser power for various values of laser
scanning speed.

Figure 7 shows optical photographs of the laser-irradiated surface of alumina at


two laser power levels. It can be observed that increasing laser output power results
in the wide laser scanning tracks in the surface region. This is associated with the
high melting rates at the irradiated surface. Moreover, the microcrack formation is
less intensive at low laser power intensity compared to the high power intensity.
Figure 8 shows SEM micrographs of the laser-treated region at two laser power
levels. It is evident that the depth of the laser-treated region is slightly larger at the
high laser power level than that corresponding to the low laser power level. The

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tre… 6/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics
formation of longitudinal cracks is evident, which contributes to the crack networks
in the surface region. Moreover, fine grains are observed in the surface region
because of the high cooling rates and the formation of AlN in the surface region,
which is evident from XRD diffractograms (Figure 9). The formation of the
microcracks at the surface is also associated with the nitride species formation in
the surface vicinity. Since nitrogen is used as an assisting gas during the laser
treatment process, aluminum nitride is formed at the surface. It should be noted
that AlN results in slight volume shrinkage in the surface region because of its
lower density as compared to Al2O3 (13). This enhances the microstress levels at
the grain boundaries, which contribute to the crack formation in the surface
region. As the depth below the surface increases, larger and elongated grains with
randomly distributed voids are observed. This is attributed to the relatively lower
cooling rates in this region as compared to the surface. The depth of the heat-
affected zone is narrow for both laser power levels. This is associated with the low
thermal diffusivity of alumina, in which case, the heat diffusion from the melt
region to the solid bulk remains low.

Figure 7. Optical photographs of laser-treated surfaces at two power intensities.

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tre… 7/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics

Figure 8. SEM micrographs of a cross section of a laser-treated surface at two


power intensities.

Figure 9. XRD diffractogram for laser-treated and as-received alumina surfaces.

Read full chapter


URL: https://www.sciencedirect.com/science/article/pii/B978008096532100916X

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tre… 8/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics

Structures, properties and development trends of laser-


surface-treated hot-work steels, light metal alloys and
polycrystalline silicon
L.A. Dobrzański, ... A. Drygała, in Laser Surface Engineering, 2015

1.3 Laser treatment of light metal casting alloys


The laser surface treatment of light metal alloys enables the production of a surface
layer with a thickness of under a millimeter to several millimeters and with special
functional properties: high hardness and wear resistance, while maintaining the
properties of the substrate material [14–19]. Laser power and the type of the
powder cladded into the surface conditions the structure and properties of the light
metal casting alloy’s surface layer. The material, following laser treatment made
with the HPDL, exhibits different properties from those made with other high-
power lasers; especially, it features a more homogenous remelting area and smaller
surface roughness [14–19]. The advantages stem from the unique properties of the
HPDL, that is, high efficiency of approximately 30-50% (2.0 kW laser consumes
7.5 kW together with a laser cooling system), a very high radiation absorption
coefficient, as well as a linear shape of a laser radiation beam. HPDLs are also
relatively low-priced, highly durable (over 10,000 h), do not need extra maintenance
apart from optic system cleaning, and are easy to use and mobile [14–19].
The structure of the solidified material after laser treatment is characterized by a
zonal construction with diversified morphology related to the crystallization of
magnesium alloys (Figures 1.13 and 1.14) [14–24]. Multiple changes in crystal
growth direction have been observed for these areas. In the area located on the
boundary between the solid and liquid phase, minor dendrites occur on the main
axes oriented along the heat disposal directions (Figures 1.15 and 1.16) [14,15,19].

Figure 1.13. Border between the RZ and the HAZ of the surface layer of
MCMgAl12Zn1 alloy after titanium carbide powder cladding, the laser power of
1.6 kW, alloying speed of 0.75 m/min; SEM.

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tre… 9/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics

Figure 1.14. MCMgAl12Zn1 alloy remelting border after cladding TiC powder, laser
power of 1.6 kW, alloying speed of 0.75 m/min; SEM.

Figure 1.15. Central zone of the MCMgAl3Zn1 alloy surface layer after cladding
with TiC particles; scan rate of 0.75 m/min, laser power of 1.2 kW.

Figure 1.16. Structure of the interface between the laser-melted zone, HAZ, and
the substrate of the MCMgAl9Zn1 alloy after laser treatment with SC particles; scan
rate of 0.75 m/min, laser power of 2.0 kW.

The results obtained from the microstructure investigation performed using a SEM
(ZEISS Supra 25) with a magnification of up to 500 times reveal the presence of the
SiC used (Figure 1.12) and WC ceramic powder (Figure 1.13). For microstructure
evaluation, the backscattered electrons detection method was used, with an
accelerating voltage of 20 kV. Based on these investigations, the distribution of the
powder particles in the surface layer of the AlSi7Cu and AlSi9Cu aluminum casting
alloys was presented. It was also found that in the laser-treated surface layer, there

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 10/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics
were no pores or cracks in the coating produced, nor did any defects or failures
occur in this layer.
Occasionally, discontinuity of the layer can be seen as a product of the heat transfer
process and may be neutralized by properly adjusted powder quality and powder
feed rate. It is also possible, on the basis of the cross-section micrograph, to
evaluate the thickness of the powder feed depth, which is ca. 120 μm in the case of
SiC (Figure 1.17) and ca. 1.2 mm in the case of WC powder (Figure 1.18).

Figure 1.17. Surface layer of the AlSi9Cu alloy after laser alloying with SiC particles;
scan rate of 0.25 m/min, laser power of 2.0 kW, powder feed rate of 8.0 g/min.

Figure 1.18. Surface layer of the AlSi7Cu alloy after laser alloying with WC particles;
scan rate of 0.25 m/min, laser power of 2.0 kW, powder feed rate of 3.0 g/min.

In the case of the WC powder, the particles are located mainly on the bottom of the
RZ. It was also found that the examined layers consisted of three subzones—the
RZ, the heat influence zone, and the substrate material. Further investigations will
be needed to reveal the morphology and nature of these zones that occur after
alloying with different process parameters and different ceramic powders [14,15].
As a result of the metallographic observations, it was confirmed that the structure
of the composite layers produced is free of defects, with distinct grain refinement
containing evenly distributed dispersion particles of TiC, WC, SiC carbide, or Al2O3
oxide applied, which was also confirmed by EDS (Figure 1.19) or electron graphic
tests (Figure 1.20).

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 11/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics

Figure 1.19. Structure of the thin foil of casting magnesium alloy MCMgAl12Zn1
after laser treatment with TiC (TEM), laser power: 1.6 kW: (a) light field; (b) dark
https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 12/28
11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics
field from the [220] reflex TiC; (c) diffraction pattern from the area as in (a); (d)
solution of the diffraction pattern from (c).

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 13/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 14/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics

Figure 1.20. Surface layer structure of the magnesium alloy MCMgAl6Zn1


subjected to laser treatment using titanium carbide: (a) the image obtained using
secondary electrons and the surface distribution of elements, (b) Mg, (c) Al, (d) Zn,
(e) Ti, (f ) C; scanning speed of 0.75 m/min, laser power of 1.6 kW; SEM.

The alloys with laser-cladded particles of vanadium carbide (their share in the
remelting zone being slight) are the exception to the rule. A strong circulation of
the liquid metal took place during laser cladding. After the laser bundle remelting,
rapid solidification of the initially liquid material took place. The thickness of the
laser-formed surface layer is of vital importance in the determination of the
material properties, period of use, and final application of the material obtained
[14–19].
Three zones occur in the surface layer of cast magnesium alloys: a zone rich with
unsolved particles cladded on the surface of magnesium alloys, a RZ, and a HAZ.
Both RZ and HAZ (depending on the concentration of aluminum in the
magnesium matrix, laser power applied, and ceramic powder) are of different
thickness and shape (Figures 1.13–1.18). It was proven that along with the growth
of the power applied, the area of occurrence of both RZ and HAZ increased (Figure
1.21), as well as the face of weld changes, which is also confirmed by the reference
studies. The MCMgA112Zn1 alloys are characterized with the largest thickness of
the surface layer, 3.59 mm, processed with a laser power of 2.0 kW, with silicon
carbide cladded into their surface [14,15].

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 15/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics

Figure 1.21. Influence of laser power on thickness of the RZ, HAZ, and the surface
layer of casting magnesium alloys after cladding.

The power of the laser within the range of 1.2-2.0 kW provides the possibility to
obtain flat regular remelting welds with a highly smooth surface (Figures 1.22 and
1.23).

Figure 1.22. View of the MCMgAl9Zn1 casting magnesium alloy face of weld after
laser treatment with WC powder, scan rate: 0.5 m/min, laser power: 2.0 kW.

Figure 1.23. View of the MCMgAl9Zn1 casting magnesium alloy face of weld after
laser treatment with SiC powder, scan rate: 0.5 m/min, laser power: 2.2 kW.

An intensive heating of the surface causes the creation of the uneven areas and
hollows in the surface layer of the Mg-Al-Zn and Al-Si-Cu cladded using carbide

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 16/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics
particles. The formation of the melted material in the remelting lake depends on
parameters such as the type of substrate, laser power, cladding rate, and the laser
powder. Some of the alloy and ceramic parts embedded in the remelting zone are
evaporated at high temperature occurring during laser treatment, therefore the
characteristic hollows appear on the remelting surface. It was also found that,
disregarding the ceramic powder used, in the laser power range from 1.2 to
2.0 kW, the porosity of the composite layers obtained increases, in comparison to
that of the raw casting surfaces of magnesium alloys.

Read full chapter


URL: https://www.sciencedirect.com/science/article/pii/B9781782420743000015

Laser nitriding and carburization of materials


D. Höche, ... P. Schaaf, in Laser Surface Engineering, 2015

2.4.4 Nitrocarburization of materials


This type of laser surface treatment was carried out by Shapochkin and Smirnov
[128] but in a step-by-step way. The alloying effect is based on a classical furnace
treatment followed by laser irradiation leading to a high resistance to contact
fatigue. All-encompassing, a comprehensive number of studies in terms of pure
laser-based processes does not exist, hence future possibilities should be fathomed
out. New studies on AISI H12 tool steel [129] applying a nitrogen gas flow might
be a base for extending application-oriented research.

Read full chapter


URL: https://www.sciencedirect.com/science/article/pii/B9781782420743000027

Laser surface treatment of biomedical alloys


R. Vilar, A. Almeida, in Laser Surface Modification of Biomaterials, 2016

2.3 Microstructure and phase transformations in laser-treated


materials
The properties of a material are ultimately determined by its composition and
microstructure. It is of utmost importance to understand the microstructure
formation mechanisms in liquid phase laser coating methods to be able to choose
processing parameters that optimize the properties of the coating. Laser coating
processes involve solidification and cooling rates typically in the ranges 0.5–
50 mm s−1 and 5 × 10−2–105 K s−1. At these high cooling rates solid-state
transformations are suppressed and the final microstructure and properties of the
laser-treated layer are those that result from solidification. However, in some
materials, such as carbon and alloy steels and titanium alloys, martensitic
transformations may occur as the material cools down to room temperature.
Sometimes diffusive solid-state transformations occur in regions where the
resolidified material is reheated due to laser track overlapping [43,56,57], but their
effect is extremely localized.
2.3.1 Solidification

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 17/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics
In liquid phase laser surface treatment processes the liquid is in contact with a
crystalline substrate. Heat is extracted to the bulk, creating a positive temperature
gradient at the solid–liquid (S/L) interface. Solidification starts at the bottom of the
melt pool, usually by epitaxial growth on the crystalline substrate, without
nucleation being required. The solidification front progresses towards the surface
as the laser beam moves away from its initial position. If undercooling is neglected,
the S/L interface coincides with the liquidus and solidus isotherms, which mark the
beginning and end of solidification according to the equilibrium phase diagram.
Under stationary conditions, ie, at some distance from the beginning and from the
end of the tracks, the shape and size of the melt pool are approximately constant
and the S/L interface moves in the laser beam displacement direction, remaining
parallel to it. Under these conditions growth is columnar because the liquid, which
is permanently irradiated by the laser beam, is overheated, ie, at temperatures
higher than the solidification temperature.
The solidification rate (R) depends on the laser scanning speed (V), melt pool
shape, and the crystal preferential growth directions. R and V are related by (Eq.
[2.2]) [58]:
[2.2]

where θ and φ are the angles between R and V vectors and between the normal to
the interface (the temperature gradient direction) and the nearest preferential
growth direction, if any (for example, the <100 > directions in the dendritic growth
of cubic crystals), respectively. If no preferential growth direction exists φ  =  0.
Since cos θ  =  0 at the bottom of the melt pool, R = 0. R increases rapidly as
solidification proceeds, up to a value lower than the scanning speed (cos θ < 1) near
the surface. The temperature gradient (G) is highest at the bottom of the melt pool
and decreases as the solidification front approaches the surface. If R and G are
known, two other important solidification parameters can be calculated: the G/R
ratio, which controls the S/L interface morphology at solidification rates lower than
the limit of absolute stability [59], and the cooling rate, given by (Eq. [2.3]):
[2.3]

G/R presents an infinite value at the bottom of the melt pool and decreases as
solidification proceeds, while the cooling rate increases as the S/L interface
approaches the surface. In general, for the values of the G/R ratio prevalent in laser
deposition, the S/L interface is plane at the bottom of the melt pool due to the
large value of this ratio in this region, and evolves to a cellular and eventually a
dendritic morphology because G/R decreases as the interface moves towards the
surface, leading to featureless, cellular and dendritic solidification microstructures
(Fig. 2.5). The solidification parameters vary rapidly near the bottom of the melt
pool, but transformation rates slow down rapidly and these parameters remain
approximately constant during most of the solidification process. As a result, the
plane front solidification and the cellular microstructure regions are usually very
narrow (sometimes even difficult to detect) and the microstructure of the
resolidified layer is almost completely dendritic, as shown in Fig. 2.5. Since the
solidification parameters are approximately constant for a given set of processing
parameters the morphology and characteristic dimensions of the dendritic
structure are approximately uniform. In the later stages the liquid may become
sufficiently undercooled for equiaxed growth to occur, in particular if there are
particles in suspension in the liquid that facilitate heterogeneous nucleation of the
equiaxed crystals [60].

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 18/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics

Figure 2.5. Transition between plane front to solidification, cellular solidification


and dendritic solidification near the coating/substrate interface. The coating
material is an NiCrAlY high-temperature material.

For single-phase materials, the strength and hardness of resolidified materials is


often inversely proportional to the dendrites size. This size is usually characterized
by the secondary dendrite arm spacing (λ2), which varies with the cooling rate. The
secondary dendrite arm spacing varies with the cooling rate according to equation
(Eq. [2.4]) [61]:
[2.4]

with c and n constants. In general, the exponent n is in the range 0.33–0.5. This
equation shows that the dendrite characteristic dimensions decrease with
increasing scanning speed, potentially leading to increased hardness and strength.
It also allows estimation of the cooling rate from secondary dendrite arm spacing
measurements, because c and n are similar for a wide range of alloys, and their
values are known for many systems.
Another important factor affecting the coating properties, in particular, their
corrosion resistance is chemical inhomogeneity. The composition of the solid
resulting from a solidification process is usually different from the composition of
the liquid and varies during solidification. The ratio of the concentrations of solute
in the solid and the average concentration of the solute in the liquid is
characterized by the effective partition coefficient (k). In the absence of solid-state
diffusion, the distribution of solute in a solid solution resulting from unidirectional
solidification can be calculated using Scheil’s equation (Eq. [2.5]) [62].
[2.5]

where Cs is the solute concentration in the solid, k the partition coefficient, C0 the
initial solute concentration in the liquid and fs the fraction of solid. In the case of
dendritic solidification a model such as the KGT (Kurz–Giovanola–Trivedi) model
[63] can be applied to analyze the solidification process. The effective partition
coefficient, which characterizes the redistribution of solute between the solid and
the liquid, depends on the solidification parameters according to the equation (Eq.
[2.6]):
[2.6]

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 19/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics
where k0 is the equilibrium partition coefficient, R the solidification rate, D the
solute diffusion coefficient in the liquid and δ the thickness of the stationary
boundary layer, which depends on the velocity of the fluid convective motion. The
effective partition coefficient varies within the range k0 to 1.
For a specific alloy composition, the solidification microstructure depends
essentially on the local solidification parameters (solidification rate, R, and
temperature gradient at the S/L interface, G), which, in turn, depend on the heat
and mass transfer in the system. The solidification parameters cannot be easily
measured, but they can be estimated using approximate analytical solutions to the
heat conduction equation [64,65] or numerical simulations based on finite
differences or finite element methods [34,38,40].
By using a suitable microstructure selection criterion and the response functions of
all phases and interface morphologies liable to form in the system, Kurz et al. [66–
69] calculated microstructure selection maps that allow prediction of the
solidification microstructure of simple metallic alloys when the solidification
parameters R and G are known. These principles were applied by several authors to
predict the solidification structure of several engineering alloys, such as medium
carbon steels [70], Al-Si alloys [71] and austenitic stainless steels [72–74]. These
predictions are a useful tool for optimization of the materials’ properties.
2.3.2 Solid-state transformations
At the high cooling rates prevailing in laser surface treatments, diffusive solid-state
transformations are usually suppressed, but in some alloys, such as α/β titanium
alloys and steels, martensitic transformations may occur. In a transformation of
this type the proportion of martensite formed does not depend on the cooling rate
but on the undercooling below the martensite start temperature. For many alloys it
can be estimated by the Koistinen and Marburger equation (Eq. [2.7]) [75]:
[2.7]

where T is temperature, Ms the martensite start temperature and α a constant that


depends on the material, typically of the order of 0.011. The martensite start
temperature is a function of the material’s chemical composition and there are a
number of empirical equations relating Ms with the chemical composition. When
laser-treated tracks are overlapped, the underlying material is reheated and
diffusive transformations may occur. The extent of a diffusional solid-state
transformation during an isothermal heat cycle can be calculated from isothermal
kinetics information, using the concept of kinetic strength of the thermal cycle (I)
[64] defined by (Eq. [2.8]):
[2.8]

where Q is the activation energy for the transformation concerned and R the gas
constant. Costa et al. [43,56,57] and Crespo et al. [76–78] developed a
thermokinetic finite element model that allows calculation of the distribution of
microstructures, properties and internal stresses in laser-deposited materials,
taking into consideration the diffusional and martensitic solid-state
transformations that occur during the laser treatment process.

Read full chapter


URL: https://www.sciencedirect.com/science/article/pii/B9780081008836000022

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 20/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics

Laser surface modification of nickel–titanium (NiTi)


alloy biomaterials to improve biocompatibility and
corrosion resistance
K.W. Ng, H.C. Man, in Laser Surface Modification of Alloys for Corrosion and
Erosion Resistance, 2012

5.1.3 Importance of laser surface treatment


Among many different types of surface treatment, laser surface treatment has been
shown to be a viable means for modifying the surface properties of materials [14–
20]. Lasers have been employed for surface modification because of their flexibility,
high precision and intense beam power [21]. The electromagnetic radiation of a
laser beam is absorbed within the surface layer for metals. The laser energy can be
deposited precisely at the point where it is needed. The bulk properties of NiTi
alloy will not be affected after laser surface treatment [22]. The treated surface layer
has a strong metallurgical bond with the substrate. The controlled thermal
penetration and non-contact feature of laser minimize the need for the post-
machining processes. Laser surface treatment is a clean and fast process for surface
engineering. The technique can be applied to bone fixation plates or implants with
relatively large surface area by overlapping the melt tracks.

Read full chapter


URL: https://www.sciencedirect.com/science/article/pii/B9780857090157500051

Laser surface modification of steel and cast iron for


corrosion resistance
R. Vilar, in Laser Surface Modification of Alloys for Corrosion and Erosion
Resistance, 2012

1.3.2 Solid-state transformations


At the high cooling rates prevailing in laser surface treatment, solid-state diffusive
transformations are usually suppressed, but in steels and cast irons austenite may
transform into martensite by a martensitic (diffusionless) transformation, which is
sufficiently fast to occur even in these conditions. In a typical martensitic
transformation the proportion of martensite formed (ym) does not depend
noticeably on the cooling rate but on the undercooling below the martensite start
temperature. It can be given, for example, by the Koistinen and Marburger
equation (Eq. 1.6) [43]:
[1.6]

where T is temperature, Ms the martensite start temperature and α a constant that


depends on the material and is typically of the order of 0.011.
When laser-treated tracks are overlapped to cover extensive areas, considerable
reheating of the previously deposited material occurs and diffusive transformation
may be induced, in particular martensite tempering. Knowing the thermal cycles at
each point, the phase transformations that may occur can be identified and their
extent calculated using a finite element heat transfer model [23]. The expectable
transformations in a typical martensitic stainless steel (AISI 420 – Fe-0.33%C–
13.5%Cr) are represented in Fig. 1.1. The microstructure after solidification
https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 21/28
11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics
consists essentially of primary austenite dendrites [17]. A proportion of this
austenite given by the Koistinen and Marburger equation transforms into
martensite during cooling to room temperature. The transformations occurring
during cooling in this first thermal cycle are indicated as Cooling in Fig. 1.1. During
overlapping, the previously deposited material undergoes a new thermal cycle,
which may originate martensite tempering, a complex sequence of precipitation
reactions that, eventually, leads to the decomposition of martensite into ferrite and
carbides [18]. If the temperature exceeds Ac1 temperature, austenitisation occurs.
These transformations are indicated as Reheating in Fig. 1.1. The transformations
observed during the corresponding cooling period depend on the transformations
that occurred during heating and are indicated as 2nd cooling in the diagram. If no
austenite is formed during heating, no significant phase transformations occur, but
if the material was austenitised (Ac1 < T < Ac3), austenite will partially or totally
transform into martensite. Reheating may also cause destabilisation of retained
austenite and its transformation into martensite [18].

1.1. Diagram indicating the phase transformation occurring during the repeated
thermal cycles generated due to single layer overlapping in an AISI 420 martensitic
stainless steel. M – martensite; α – ferrite; γ – austenite; L – liquid phase [23].

Read full chapter


URL: https://www.sciencedirect.com/science/article/pii/B9780857090157500014

Advances in Additive Manufacturing and Tooling


R. Vilar, in Comprehensive Materials Processing, 2014

10.07.5.2 Solid-State Transformations


At the high cooling rates prevailing in laser surface treatment, diffusive solid-state
transformations are usually suppressed, but in some alloys, such as α/β titanium
alloys, steels, and cast irons martensitic transformations may occur. In a
transformation of this type the proportion of martensite formed does not depend
noticeably on the cooling rate but on the undercooling below the martensite start
temperature. It is given by Koistinen and Marburger equation (eqn [22]) (122):
[22]

where T is the temperature, Ms the martensite start temperature, and α a constant


that depends on the material, but is typically of the order of 0.011. The martensite
start temperature is a function of the materials chemical composition and there are

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 22/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics
a number of equations relating Ms with the chemical composition (123). In general,
laser-melted steels present a much larger proportion of retained austenite than
conventionally treated materials (113,124,125). This is partially explained by the
prevalence of nonequilibrium γ solidification over equilibrium δ solidification at
high solidification rates, as previously mentioned, but in order to retain austenite
at room temperature the martensitic transformation must be suppressed. Several
factors other than the chemical composition affect the martensite start
temperature (Ms) and, hence, the proportion of austenite retained at any
temperature lower than Ms. Particularly relevant in the context of laser deposition
is the fact that laser melting dissolves entirely the carbides initially existing in the
microstructure, leading to a much larger contents of alloying elements dissolved in
austenite and, consequently, to lower Ms temperatures as compared with
conventionally treated steels (126). Residual stresses (127) and a large density of
crystallographic defects such as dislocations and grain boundaries also contribute
to reduce Ms (128). On the basis of the existing experimental evidence, Colaço and
Vilar (113) concluded that the refinement of austenite dendrites due to rapid
solidification (see eqn [19]) is the most probable explanation for the large
proportions of retained austenite observed in high alloy steels treated at high laser
scanning speeds.
When laser-treated tracks are overlapped, the underlying material is reheated and
diffusive transformations may occur. In this context, tempering of steels is
particularly important because these materials are frequently processed by laser
deposition and tempering affects considerably their properties. In laser-treated
steels tempering reactions may be different from those observed in conventionally
treated steels, due to their larger proportions of retained austenite (113,129).
Laser-deposited tool steels undergo secondary hardening (130), but their
secondary hardening temperature is usually higher than that after conventional
quenching (114,129,131), while the maximum secondary hardness may be either
higher or lower. For example, the secondary hardening temperature of laser-
treated DIN X42Cr13 martensitic stainless steel is 600 °C and the maximum
hardness 620 HV, as compared with 500 °C and 570 HV for the same steel in the
quenched condition (129). The microstructure after laser surface melting, which
consists of about 40% lath-type martensite, 59% retained austenite and 1% M23C6
carbide, does not change noticeably during 1 h isochronous tempering treatments
up to 500 °C but at higher temperatures martensite decomposes into a mixture of
ferrite and M7C3 and M23C6 carbides, which precipitate within the martensite laths
and at the lath boundaries. Above 575 °C M7C3 carbide precipitates within retained
austenite as well, leading to destabilization of this phase. Destabilized austenite
decomposes into a mixture of α+ carbides if the material remains at high
temperature or transforms into martensite if the material is cooled down below Ms
temperature. At temperatures higher than 600 °C recrystallization occurs. A similar
behavior is observed for AISI 440C steel (Fe–1Cr–18Cr) (130).
The extent of a diffusional solid-state transformation during an anisothermal heat
cycle can be calculated from isothermal kinetics information, using the concept of
kinetic strength of the thermal cycle (I) (99). The kinetic strength is defined by (eqn
[23]):


[23]

where Q is the activation energy for the transformation concerned and R, the gas
constant.
Ashby and Easterling (99) proposed a simplified approximated equation to calculate
I (eqn [24]):
[24]

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 23/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics
where Tmax is the peak temperature in the thermal cycle and τ is the cycle thermal
constant. For their calculation, see Ref. (99). The kinetics of phase transformations
during anisothermal thermal cycles, such as those existing in laser powder
deposition, was analyzed by Mittemeijer (132).
An anisothermal kinetics model was used by Costa et al. (133) for modeling the
diffusional phase transformations occurring during laser deposition of steels.
Knowing the thermal cycles at each point of a part, for example, by finite element
calculations, and the phase transformations that occur in the material in the
temperature range, the extent of these transformations can be calculated using
anisothermal transformation kinetics (134). The anticipated transformations in a
typical martensitic stainless steel (AISI 420: Fe–0.33%C–13.5%Cr) are represented
in the diagram of Figure 19. The microstructure after solidification consists
essentially of primary austenite dendrites (112). A proportion of this austenite given
by the Koistinen and Marburger equation (eqn [22]) transforms into martensite
during cooling down to room temperature. The transformations occurring during
the first thermal cycle are indicated as Cooling in Figure 19. During overlapping,
the previously deposited material undergoes a new thermal cycle, which may
originate tempering, a complex sequence of precipitation reactions that, eventually,
leads to the decomposition of martensite into ferrite and carbides (129). If the
temperature exceeds Ac1 temperature, austenitization occurs. These
transformations are indicated as Reheating in Figure 19. The transformations
observed during the corresponding cooling period depend on the transformations
that occurred during heating and are indicated as 2nd cooling in Figure 19. If no
austenite is formed during heating, no significant phase transformations occur, but
if the material was austenitized (Ac1 < T < Ac3), austenite will partially or totally
transform into martensite.

Figure 19. Diagram indicating the phase transformations occurring during the
repeated thermal cycles generated due to single layer overlapping in an AISI 420
martensitic stainless steel. M – martensite; α – ferrite; γ – austenite; L – liquid
phase.
Adapted from Costa, L.; Vilar, R.; Reti, T. Rapid Tooling by Laser Powder Deposition: Process
Simulation Using Finite Element Analysis. Acta Mater. 2005, 53, 3987–3999.

Ti alloys also undergo solid-state transformations during laser powder deposition.


Pure Ti and Ti alloys with very low alloying elements concentration (usually called α
Ti alloys) solidify at 1670 °C into β phase, with BCC structure. In equilibrium
conditions, β transforms at 882 °C into HCP α-phase by a diffusion-controlled
allotropic transformation and this phase remains stable down to room
temperature. At high cooling rates, this diffusional transformation is replaced by a
martensitic type transformation leading to the formation of massive (lath or packet)
martensite (α′). The corresponding transformation is sometimes designated by
https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 24/28
11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics
massive transformation. The transformation product presents a platelike
morphology, as shown in the micrograph of Figure 20.

Figure 20. SEM micrograph of the transverse cross-section of a CP Grade 2 Ti wall


showing α′ platelets.
Adapted from Meacock, C. Laser Powder Microdeposition of Biomedical Alloys. Ph.D. Thesis,
Technical University of Lisbon, 2010.

Similar transformations occur in α + β Ti alloys, such as Ti–6Al–4V. These alloys


solidify as β-phase. If the cooling rate to room temperature is lower than 410 °C
s−1 β transforms by a diffusion-controlled process into the HCP α-phase. α
presents, in general, a Widmanstätten morphology for the cooling rates normally
existing in laser deposition (135) (Figure 21(a)). For higher cooling rates, the
transformation is martensitic and leads to the formation of acicular α′ (Figure 21(b))
(135). In both cases, a Burger's crystallographic orientation relation exists between
the newly formed α phase and the parent β phase (136):

Figure 21. Micrographs showing (a) Widmanstätten α structure and (b) acicular α′
martensite, in the transverse cross-section of a Ti–6Al–4V wall.
Adapted from Meacock, C. Laser Powder Microdeposition of Biomedical Alloys. Ph.D. Thesis,
Technical University of Lisbon, 2010.

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 25/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics

In addition to the aforementioned phase transformations, which occur during


quenching, metastable α′ martensite formed in Ti–6Al–4V decomposes upon
heating into equilibrium α and β phases with a Johnson–Mehl–Avrami kinetics.
This tempering transformation may be induced by reheating due to track
overlapping, but its kinetics in Ti6Al4V is so slow that it is unlikely that tempering
plays a significant role in determining the properties of parts of this material
produced by laser deposition. The phase transformations that may occur in Ti–6Al–
4V alloy during deposition are summarized in Figure 22.

Figure 22. Diagram indicating the phase transformations occurring during the
repeated thermal cycles generated in laser deposition of Ti–6Al–4V alloy, ignoring
tempering.
Adapted from Crespo and Vilar.Reproduced from Crespo, A.; Vilar, R. Finite Element Analysis of the
Rapid Manufacturing of Ti-6Al-4V Parts by Laser Powder Deposition. Scr. Mater. 2010, 63, 140–
143.

Read full chapter


URL: https://www.sciencedirect.com/science/article/pii/B9780080965321010050

Laser surface remelting to improve the erosion–


corrosion resistance of nickel–chromium–aluminium–
yttrium (NiCrAlY) plasma spray coatings

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 26/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics
B. Singh Sidhu, in Laser Surface Modification of Alloys for Corrosion and Erosion
Resistance, 2012

10.3 Applications of laser remelted coatings to combat erosion and


corrosion
The research work published on the effect of laser surface treatment on the high-
temperature behaviour of protective coatings or alloys used in these coatings can
be classified into two categories depending on the main aspect of modification
involved: one deals with the structural modification of coatings and the other with
consolidation of coatings by the elimination of porosity.
However, the presence of trapped oxygen in atmospheric plasma sprayed coatings
causes unfavourable oxide inclusions during the laser melting process. These oxide
inclusions may develop into sites of oxidation attack. Atmospheric plasma sprayed
coating seems therefore not to be the best candidate for laser treatment. For these
reasons laser surface melting was applied to low-pressure plasma sprayed coatings.
These coatings have no residual porosity and the laser treatment affects only the
phase composition and distribution within the coating.

Read full chapter


URL: https://www.sciencedirect.com/science/article/pii/B9780857090157500105

Recommended publications:

Applied Surface Science


Journal

Colloids and Surfaces B: Biointerfaces


Journal

Polymer Degradation and Stability


Journal

Journal of Molecular Structure


Journal

Browse Journals & Books

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 27/28


11/16/2020 Laser Surface Treatment - an overview | ScienceDirect Topics

Copyright © 2020 Elsevier B.V. or its licensors or contributors.


ScienceDirect ® is a registered trademark of Elsevier B.V.

https://www.sciencedirect.com/topics/chemistry/laser-surface-treatment#:~:text=Laser surface treatment is a,advantage over conventional furnace tr… 28/28

You might also like