Download as pdf or txt
Download as pdf or txt
You are on page 1of 38

Accepted Manuscript

Title: Laser Direct Deposition of AISI H13 Tool Steel Powder


with Numerical Modeling of Solid Phase Transformation,
Hardness, and Residual Stresses

Authors: Neil S. Bailey, Christopher Katinas, Yung C. Shin

PII: S0924-0136(17)30155-3
DOI: http://dx.doi.org/doi:10.1016/j.jmatprotec.2017.04.020
Reference: PROTEC 15198

To appear in: Journal of Materials Processing Technology

Received date: 23-9-2016


Revised date: 5-4-2017
Accepted date: 25-4-2017

Please cite this article as: Bailey, Neil S., Katinas, Christopher, Shin, Yung C., Laser
Direct Deposition of AISI H13 Tool Steel Powder with Numerical Modeling of Solid
Phase Transformation, Hardness, and Residual Stresses.Journal of Materials Processing
Technology http://dx.doi.org/10.1016/j.jmatprotec.2017.04.020

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Laser Direct Deposition of AISI H13 Tool Steel Powder with Numerical Modeling of

Solid Phase Transformation, Hardness, and Residual Stresses

Neil S. Bailey, Christopher Katinas, Yung C. Shin

Center for Laser-based Manufacturing, Purdue University, West Lafayette Indiana,


47907, USA

ABSTRACT

This study is concerned with the systematic study on the laser direct deposition of AISI

H13 tool steel powder on H13 substrates to determine and investigate the hardness and

residual stresses resulting from the laser deposition process. In this paper, numerical

models are presented to predict three-dimensional solid phase transformation, material

hardness, and residual stresses produced by the laser direct deposition process with multi-

track and multi-layer capabilities. The simulated laser deposition track is based on the

predicted temperature and geometry from a free-surface-tracking laser deposition model.

The predictive modeling results are validated by experiments in terms of tempering

patterns, hardness field measurements, and XRD residual stress measurements. Predicted

results show the variation of stresses and hardness into the depth and along the surface of

the work piece provide a reasonable agreement with measured values. It was found that

the heat-affected zone with high phase fractions of martensite result in strong

compressive stresses, but the tempering effect due to multi-track laser deposition can

somewhat alieviate the compressive stress. It was also found that the ultimate tensile

strength of laser deposited H13 steel is 15% to 30% higher than published values for

commercial H13.

1
Keywords: laser deposition, predictive modeling, phase transformation, residual stress,

H13 tool steel

1. Introduction

In high temperature applications, H13 tool steel has excellent material properties, as

described by Roberts et al. (1998), such as high toughness, high stability in heat

treatment, and high resistance to thermal fatigue cracking. Because of these properties,

H13 is the standard material used in industry for toolings in hot work applications such as

die casting, forging, and extrusion. However, despite being the best material for these

applications, H13 tooling parts will still fail after many cycles. Common failures include

corrosion, pitting, thermal fatigue cracking, and wear. The failure often occurs at a small

location on the part, rendering the entire part unusable, and replacing the entire part can

be costly.

Traditional tooling repair methods are labor intensive and involve manually

removing the damaged area (via grinding, for example), rebuilding the damaged area via

manual welding, and finishing the rebuilt geometry with traditional machining.

According to Thompson (1999), rebuilding the damaged area via manual welding is

usually the limiting factor in determining if a failed tooling part can be repaired, and even

if a part can be repaired, Tušek and Ivančič (2004) cite that the side-effects of manual

welding (such as porosity, distortion, and thermal stress) often give the tool a much

shorter production life than a new part.

Tooling repair and remanufacture via laser direct deposition (LDD) has many

advantages over traditional methods. Besides the obvious advantage of a much more

precise and versatile addition of material, according to Majumdar and Manna (2012) and

2
Jhavar et al. (2013), the LDD process itself can improve the material properties and

extend the life of the tool. Pinkerton et al. (2008) studied the component repair of deep

cracks or defects using LDD of H13 steel. They machined grooves into a substrate to

simulate deep cracks and demonstrated the efficacy of using LDD to repair the defect.

Telasang et al. (2014) studied laser cladding of H13 on an H13 substrate to determine the

effectiveness of the laser cladding process in repairing/refurbishing large H13 steel

components. Using pulsed and continuous wave techniques, they found that post-

cladding laser hardening or conventional tempering increased the hardness and the

magnitude of the compressive residual stress near the surface of the clad. Remanufacture

of parts that have had a design change that requires adding material are also good

candidates for the LDD process, although this hasn’t seen much attention in the literature.

Numerical prediction of residual stresses in LDD must be built on a solid

foundation of temperature and geometry numerical prediction. Various methods of

numerical analysis have been used by various research groups to predict ideal process

parameters and resulting geometry of the repaired or remanufactured tooling. Numerical

models have been developed so as to predict the temperature, melt flow, powder flow,

and resulting geometry during laser cladding and laser deposition. Pinkerton, (2015)

reviewed the state of the art in modeling of LDD. Many analytical models exist, which

define the melt pool surface as half ellipses or circles. Fathi et al. (2006) presented a

model based on a moving heat source and an energy balance that predicts the depth of the

melt pool during LDD. Lalas et al. (2007) built a model to predict LDD geometry based

on the surface tension between the melt pool and the substrate. El Cheikh et al. (2012)

used experimental data to build a model that predicts deposition track geometry based on

3
a disk shape. Although computationally efficient, these models fail to capture much of

the physics that occur during laser deposition. He et al. (2009) developed a numerical

model to study the temperature, fluid flow, and species transport and used it to study two-

track laser cladding of H13 powder onto an H13 substrate. Ibarra-Medina et al. (2011)

included the inflow of powder particles in their model and used the volume of fluid

method to track the free surface. Manvatkar et al. (2015) built a simplified model that

predicts the temperature and approximate geometry during LDD and used it to analyze

the deposition of a thin wall (single-track, multi-layer) of H13.

Fewer numerical models have used the data from thermal models to predict

residual stresses that are caused by the extreme thermal cycles inherent in LDD. Using

ANSYS, Labudovic et al. (2003) developed a finite element model that calculates

transient temperature profiles and residual stresses, but does not predict the detailed

profile of the laser track. They did not consider phase transformational strains in their

residual stress prediction since they simulated MONEL 400 powder deposited onto a

steel substrate. Using COMSOL, Alimardani et al., (2007) created a similar FE model

and predicted the geometry and residual stresses during LDD of 304 stainless steel.

Kamara et al. (2011) used ANSYS to predict temperature and resulting residual stresses

in a thin wall of Waspaloy deposited onto an Inconel substrate. Similarly, Wang et al.

(2008) used SYSWELD for modeling residual stresses in thin walls of deposited stainless

steel.

For certain materials, residual stresses cannot be accurately predicted without

predicting phase transformation during LDD. For steels, strong tensile and compressive

stresses can develop in the material due to transformations between solid phases such as

4
ferrite, pearlite, austenite and martensite as shown by Bailey et al. (2009).

Santhanakrishnan et al. (2011) used a finite element model to predict temperature during

high power direct diode laser cladding of H13, and then used the temperature to predict

the phase transformation kinetics and the hardness in the clad material. Brückner et al.

(2007) presented a finite element model for laser cladding to predict residual stresses and

they also accounted for strains caused by phase transformation in steel. However, instead

of simulating a free surface, their clad geometry was generated from a sheared ellipsoid

passing over to an elliptical cylinder at the solidification front. Farahmand and Kovacevic

(2014) have developed a model that predicts residual stresses in multi-track cladding with

a high power diode laser, including phase transformational stresses in steels. Again, the

predicted track geometry is simplified and not based on a free-surface tracking model.

To make modeling easier, all of these residual stress finite element models use a

simplified computational domain where the deposited track profile is assumed to have a

known profile based on a geometric shape such as a square, ellipse, or circle. This

simplification allows for the deactivation and reactivation of finite elements to be based

on a mathematical function (i.e. the simplified melt pool geometry traveling at the laser

travel speed) rather than the physical track profile as predicted by a deposition model’s

free-surface tracking. In order to more accurately predict the residual stress in a LDD

process, the deactivation and reactivation of finite elements ought to be based on an

accurate prediction of melting and solidification rather than a simple function. This will

allow for prediction in more general cases of laser deposition of complex geometry.

The purpose of the work presented in this paper is to use experimentally-

determined operating parameters for laser direct deposition of H13 tool steel to repair and

5
remanufacture two industrial tooling parts and analyze the resulting microstructure and

mechanical properties of the deposited material. Then a finite element model is presented

that predicts the solid phase transformations and the residual stresses developed during

multi-track and multi-layer laser direct deposition based on the actual geometry of the

laser track as predicted by a deposition model with free-surface tracking. The phase

transformation portion of the model is validated via hardness measurements of the

deposited material and the residual stress portion of the model is validated via residual

stress XRD measurements.

2. Experimental setup and operating parameters

The LDD system used in this work is an Optomec LENS 750, which uses a 500 W

fiber laser with a focused beam diameter of 740 µm and four nozzles to deliver the

deposition powder carried by argon gas to the focal plane of the laser. Particle size of the

H13 tool steel powder ranged between 50 µm and 150 µm. The laser and powder feeder

nozzles move vertically on the Z axis and are focused on an XY table. The system is

enclosed in an environmental chamber charged with argon gas with an oxygen level

below 20 ppm.

The determination of the optimal LDD parameters of H13 powder is not the focus

of this work, but the experimentally-determined optimal parameters for this deposition

system are given in Table 1. These parameters are used in all experiments and

simulations throughout this work.

3. Modeling

6
A numerical model has been developed to predict the phase transformation

kinetics and residual stresses during the LDD process. The model uses a previously

developed LDD model presented by Wen and Shin (2011, 2010) to provide temperature

and surface profile data to the residual stress model. A flowchart for the comprehensive

model is shown in Figure 1, and a detailed description of the model is presented here.

As can be seen from Figure 1, the LDD model provides temperature and free-surface

profile information to the residual stress model, which utilizes commercial finite element

software ABAQUS and user subroutines to calculate the stress produced during laser

deposition. The temperature and free surface information are used to determine the

deactivation and reactivation of each element in the finite element model. The pseudo-

steady temperature data is interpolated onto the finite element mesh and the local solid

phase of the material is calculated based on the temperature data according to the phase

transformation kinetic model. The thermal strain and the phase transformational strain is

then calculated and provided to the finite element solver and ABAQUS solves for the

prediction of resultant material stresses.

3.1 Laser deposition model and validation

Simulation of the deposition process was performed using a three dimensional

numerical model developed by Wen and Shin (2011, 2010). The computational domain

consisted of a 12mm x 5mm x 12mm substrate composed of H13 tool steel to which H13

powder was being deposited via the LDD process with a total of 343,728 elements

arranged in a structured, non-uniform mesh. Thermal and physical properties of the H13

tool steel are provided in Table 2. The substrate was assumed to be initially at 300K, and

7
convective boundary conditions were set at all faces of the steel substrate using a constant

heat transfer coefficient of 10 W/m2-K with an ambient temperature of 300K.

A series of experiments was performed using an Optomec LENS 750 LDD

system in order to validate the simulation by comparing measured and predicted track

geometry, heat affected zone (HAZ), and molten pool depth and width for both single and

multiple track depositions. The deposition parameters used in the simulations correspond

to those used for experimentation and measurement as listed in Table 1.

The simulations were validated against the experimental data, first for a single

track case, then for a multi-track case. Figure 2 provides an isometric view of the

deposition of the 1st track showing the temperature isocontours upon completion of the

single track simulation. In Figure 3, overlays of the predicted track profile, molten pool,

and HAZ geometries are shown on an image of a cross section from a single-track

experiment, comparing the predicted and measured geometries. Based on the shape and

location of each of the contours shown in Figure 2 and Figure 3, it can be seen that the

deposition process was reasonably well captured in the simulation model for a single

track simulation.

Simulation of the LDD process involves a high computational cost. Since the goal

of this simulation was to acquire the geometry and temperature profile of bulk deposition,

the model must calculate the multi-track case. From experimentation it was determined

that the cross-sectional profiles of the first two tracks in a multi-track deposition sample

were unique, but the third and all subsequent track cross-sectional profiles were similar.

8
Therefore, three subsequently deposited tracks provided an average steady-steady profile

which essentially remained unchanged upon sufficient additional deposited tracks.

Similar trends were been observed by Mazumder et al. (2012). Thus, to minimize the

amount of simulation time, a solidified multi-track cross-sectional profile was captured

based on a two-track experimental sample, as shown in Figure 4, and was used as the

initial surface condition for the simulation of the third track deposition process. The

profile of the two-track surface was created via manual selection of the interface points at

10 µm intervals along the cross section of the two-track sample, and was compared to

additional two-track experiments to ensure a reasonable and non-extreme track profile

was being selected as the basis for multi-track work.

Simulation for the multi-track case was allowed to run until a pseudo-steady

profile developed with respect to the direction of travel. This required approximately 4.5

mm of travel in the positive x-direction. Figure 5 shows an isometric view of the

simulated three-track deposition process with temperature isocontours at the end of

simulation. In Figure 6, overlays of the predicted track profile and molten pool are shown

on an image of a cross section from a multi-track experiment, comparing predicted and

measured geometries with a reasonable agreement.

b. Phase transformation kinetics

In material processing of steels, phase transformation kinetics have a strong

influence on the behavior of the steel. Skvarenina and Shin (2006) developed a 2D

9
kinetic model that simulates the solid phase transformations of hypoeutectoid steels

during the heating. Lakhkar et al. (2008) modified this model to include tempering cycles

and Bailey et al. (2009) expanded the model to 3D and further modified it to include the

cooling cycle. A sufficient description of this model and it’s adaptation for use in multi-

track and multi-layer residual stress prediction during LDD is presented here.

Phase transformation kinetics during the LDD processes can be divided into three

parts: heating, cooling, and tempering. As the laser spot approaches a point of interest,

the temperature increases rapidly. When the local temperature passes the eutectoid

temperature (727°C), the pearlite colonies, consisting of alternating plates of high-carbon

iron carbide and low carbon eutectoid α-ferrite, quickly transform into γ-austenite

because the carbon in the iron carbide can quickly diffuse into the close eutectoid α-

ferrite plates. As the temperature continues to increase, the carbon from the austenite

colonies diffuses into the proeutectoid ferrite colonies, transforming them to austenite.

Once the temperature passes the austenitization temperature (850°C for H13 steel), all of

the ferrite will have transformed to austenite. The transformation from ferrite to austenite

between the eutectic and austenitization temperatures is governed by carbon diffusion. In

slow heating processes, it is important to characterize carbon diffusion in order to

simulate the austenitization of the steel. However, the LDD process results in such rapid

heating to temperatures high above melting that the carbon diffusion may be neglected.

Therefore, as temperatures pass the eutectic temperature, all pearlite transforms to

austenite. Between the eutectic temperature and the austenitization temperature, the

fraction between ferrite and austenite is determined by the lever rule as defined by

10
Callister and Rethwisch (2007). Once the temperature passes the austenitization

temperature, the element is transformed to 100% austenite.

The modeling of the cooling cycle is based on Sheil’s Additivity Rule, the

Johnson-Mehl-Avrami (JMA) model as used by Kang and Im (2005), and the Koistinen-

Marburger (KM) equation as used by Woodard et al. (1999). The details of this model are

presented by Bailey et al. (2009). Since the cooling rates are very rapid, as predicted by

the LDD model for the simulation presented above, the cooling model predicts that all of

the austenite transforms to martensite.

The tempering model, as presented by Lakhkar et al. (2008), has been expanded

and adapted to the residual stress model. Ferrite and pearlite phases are stable below the

eutectoid temperature, while martensite is easily tempered at temperatures as low as

100°C. In single-track laser processing, tempering of martensite is not an issue because,

once the martensite is formed, there is no opportunity for the martensite of the single

track to be tempered by the heat of a second track. With multi-track and multi-layer

LDD, the heat from neighboring laser tracks can raise the local temperature high enough

to temper the martensite, but not high enough to transform to austenite. When this occurs,

a certain fraction of the martensite is tempered. This phenomenon must be accurately

modeled in order to predict the resultant phase and ultimately the resultant residual

stresses.

In this model, if an element has a phase fraction of martensite greater than zero,

and the temperature of that element reaches a maximum of temperature greater than

100°C but less than the eutectoid temperature 727°C, then the martensite may transform

to the following tempering phases: ε-carbide and ferrite are formed between 100°C and

11
250°C, and cementite and ferrite are formed between 250°C and 727°C. The amount of

martensite that is tempered, , is calculated by the JMA equation:

= − Eq. 1

where is an empirically derived constant with a value of 0.109. The term must be

solved for non-isothermal conditions and is given by the differential equation,

= + Eq. 2

where = 1.97x10 kJ/mol is the activation energy, = 5.11x10 is a constant, R is

the universal gas constant, and temperature T is in Kelvin.

For temperatures between 100°C and 250°C, the fractions of -carbide ( ) and

ferrite ( ) that form from the fraction of martensite that was tempered ( ) are calculated

as

= = − Eq. 3

where is the carbon concentration of H13, set at 0.4% for this simulation, and =

8.55% and = 8.22% are the carbon concentrations of the phases -carbide and ferrite,

respectively. Similarly, for temperatures between 250°C and 727°C, the fractions of

cementite ( ) and ferrite ( ) are calculated as

= = − Eq. 4

12
where the carbon concentration of cementite is = 6.7%. At the end of the simulation,

the hardness, H, of the resultant multi-layer, multi-track laser deposited material is

calculated by

=∑ Eq. 5

where are the hardness values of the various phases and are the phase fractions.

c. Residual stress model

Many simultaneously occurring physical phenomena affect the mechanical

response of the material, and in order to predict the resultant residual stress state, the

relevant mechanical strains must be accurately modeled. The residual stress model used

in this work is presented by Bailey et al. (2009)and has been expanded to include the

effects of melting, multiple laser tracks and multiple layers, and laser tempering effects.

The relevant thermal ∆ and phase ∆ strain increments can be added to the elastic

∆ and plastic ∆ strain increments according to Equation 6.

∆ =∆ +∆ +∆ +∆ Eq. 6

It is assumed that elastic and plastic strains are independent of material phase, but are
strongly dependent on temperature. To model the plastic strain, a Johnson-Cook
hardening model was used with coefficients reported by Shatla et al. (2001) and shown
here in
Table 3.

The elastic strain is dependent on the material’s Young’s modulus and Poisson’s ratio,

and both are functions of temperature. According to the High Performance Alloy

13
Database by CINDAS (Benedyk, 2008), the Poisson’s ratio of H13 steel at room

temperature is 0.28, but increases with temperature in a nonlinear fashion, approaching

0.3 at 800 K and 0.33 at 1000 K. Philip and McCaffrey (1997) reported the Young’s

modulus of H13 as a function of temperature up to 800 K, while Bohler-Uddeholm

Corporation (2016) reports the Young’s modulus of their commercial H13 steel at higher

temperatures up to 1300 K. Both the Poisson’s ratio and the Young’s modulus of H13

used in this study are shown in Figure 7 as functions of temperature.

Anisotropic thermal strains exist in a workpiece wherever thermal gradients exist.

The thermal strain increment is defined by the temperature increment, ∆ , and the multi-

component thermal expansion coefficient, , according to Equations 7 and 8.

∆ = (∆ ) Eq. 7

=∑ Eq. 8

Here, is the phase fraction of phase x and is the thermal expansion coefficient of

phase x. When a given volume in a material transforms from one solid phase to another,

the crystal structure can change. For example, austenite changing to martensite will

transform from a face-centered cubic structure to a body-centered tetragonal structure,

thereby expanding the volume. There are two parts to the phase transformational strain

given in Equation 9 as first reported by Oddy et al. (1989) and subsequently used by Das

et al. (1992) and Ghosh and Choi, (2005). First, the strain increment due to this

volumetric expansion (or contraction as the case may be) is given by the first term on the

right-hand side of Equation 9. Second, the transformation-induced plasticity strain is the

second term on the right-hand side of Equation 9.

14
∆ ∆
∆ = ∆ + ( − − ∆ )∆ Eq. 9

The percent volume change, ∆ / , due to phase transformation from one phase to
another is listed in

15
Table 4, along with the thermal expansion coefficients for each phase. The incremental

change in phase fraction is defined as ∆ . is the deviatoric stress tensor, is the

amount of phase already changed during the entire cycle, and is the yield stress of the

weaker phase, which is usually austenite during these transformations.

In order to predict residual stress in LDD, the numerical model must take melting,

solidification, and material addition into account. In order to simulate melting and

solidification, this model utilizes the ABAQUS command *MODEL CHANGE. This

command deactivates and reactivates elements according to element sets given for each

step in ABAQUS. This command must be implemented in the input file and therefore the

temperature and the level-set data must already be known.

A program was written, which imports the temperature and level-set fields from

the deposition model and interpolates these fields onto the ABAQUS mesh. Then, during

a given time increment, each element that melts during this time increment is added to

one list and each element that solidifies during this time increment is added to another

list. Both melting and solidification lists for each time increment are written to a file and

copied into the ABAQUS input file as element sets: two sets for each step in ABAQUS.

During the first step of a simulation, the top layer of elements that will represent the

deposition layer will be deactivated so that they may be reactivated to represent the

additional material added during LDD. The time duration of a step is the given time

increment mentioned above, which, for the simulation presented here, is set to 0.004

seconds.

16
Thus, during each step, two sets of elements that represent the regions of melting

and solidification become deactivated and reactivated, respectively. The stress existing in

the recently deactivated elements is lost and the stress in the recently reactivated elements

is initialized to zero. In this way, the melt pool cannot hold stress and the mechanical

response of the deposited material is accurately simulated.

4. Implementation and results

In order to validate the residual stress model presented here, a multi-track, multi-

layer finite element simulation was run in ABAQUS to predict the residual stresses that

form during LDD of H13 tool steel powder onto a 50 mm x 50 mm x 5 mm H13

substrate. The physical samples consist of two layers of multiple tracks and were built

using the LDD operating parameters listed in Table 1. Two samples were built: one was

used to analyze the resultant solid phase transformation and hardness, and the other one

was used to measure the resultant residual stress.

The simulation consists of 2 layers of 8 tracks each. The width and length of the

square 8-track domain are 2.1 mm (the first and last tracks of each layer are partial

tracks). In the fine-mesh area, the width of each element is 50 µm and the height of each

element is 25 µm. The simulated time duration for each track is exactly 1 second. At

14.85 mm/s, the laser travels 14.85 mm for each track. Since the length of the domain is

only 2.1 mm, the laser spot spends most of the simulation time outside the computational

domain. But the temperature field far from the laser spot still has a strong influence on

the thermal strains and phase strains. Since the temperature influence from subsequent

tracks can temper the martensite produced by previous tracks, the simulation time for

17
each layer was extended to 15 seconds, 7 seconds after the 8 th track has finished, to allow

for tempering from the tracks that occur past the computational domain. The orientation

of the tracks in the 2nd layer is perpendicular to those of the 1st layer. The total

computational cost for both layers was 9 days using 8 processors at 2.1 GHz. An image

of the predicted temperature field during the 3rd track is shown in Figure 8.

The elements shown in Figure 8 are the currently active elements for the current

ABAQUS step. During each step as the laser advances, a set of elements near the front of

the melt pool will be deactivated due to melting and a set of elements near the back of the

melt pool, including elements of the deposition layer, will be reactivated due to

solidification. After the first layer has been simulated, the resulting residual stress state is

imported into commercial software Tecplot where it is rotated 90° (because the 2nd

layer’s deposition direction will be orthogonal to that of the 1st) and shifted down a

distance equivalent to the deposition layer thickness (to create room in the computational

domain for the 2nd layer). The rotated and translated residual stress state is then imported

back into ABAQUS (via user subroutine SIGINI, as shown in Figure 1) as the initial

stress state in the sample before the 2nd layer simulation is begun. For this simulation, the

top of the deposition’s 1st layer is smooth enough to be assumed flat. Indeed, part of the

determination criteria for the optimal operating parameters listed in Table 1 was the

smoothness of the deposited surface.

Phase transformation is calculated during each time increment, including heating,

cooling, and tempering as discussed above. An image of the phase fraction of martensite

during the 3 rd track of the 1st layer is shown in Figure 9. From Figure 9, one can see that

18
during deposition of the 1st layer, the bulk of the manufactured component substrate has

no martensite. After a laser track has passed, the material will cool to martensite (as seen

in the left front side), some of which is transformed back to austenite as the next track’s

laser passes.

Images of the phase fraction of martensite during the 3 rd and 7 th tracks of the 2nd

layer are shown in Figure 10. From Figure 10, one can see that during deposition of the

2nd layer, the material deposited by the 1st layer has a varying amount of martensite. As

each track passes a previous track, the heat from the current track will temper (to some

degree) the previous track and previous layer. This effect can clearly be seen in the

bottom image of Figure 10. The previous tracks near the left of the image have a lower

martensite phase fraction than the track that has most recently passed. The material of the

1st layer, below the recently deposited tracks, has an even lower phase fraction of

martensite because it has also been tempered by the 2nd layer.

In comparison to the predicted tempering pattern, a micrograph image is shown in

Figure 11 from a cross-section of a two-layer multi-track sample. This sample was made

with the operating parameters given in Table 1, with two deposition layers: the 1stlayer

orientation running parallel to the image and the 2nd layer (top layer) orientation running

perpendicular to the image. Two distinctive patterns can also be seen in Figure 11,

labeled the melt boundary and the temper boundary. The melt boundary is clearly the

edge of the melt pool for each individual deposition track, while the temper boundary

shows an isocontour of the tempering experienced by previous tracks resulting from the

19
heat of following tracks. The last track at the right of Figure 11 is 100% martensite

because there were no following tracks to cause tempering. Indeed, it can be clearly seen

that the tempering pattern cannot be seen in the last track. The tempering pattern

predicted by the solid phase transformational model is similar to the pattern shown in the

experimental sample.

Hardness measurements from a single-layer multi-track sample were used to

validate the kinetic model. At the end of the simulation of the 1st layer, the material

hardness was calculated for each element in the finite element mesh according to

Equation 5. The predicted hardness field is shown as a 3D surface plot in Figure 12 as a

function of both distance parallel to the surface and depth from the top surface of the

deposited material. Notice that the dark peaks in Figure 12 represent the hardest material

near the top surface of the sample in the deposited material and they clearly show a

pattern of hardness repeating in the parallel direction. This repeating pattern represents

the tempering that is experienced by each laser track. The width between each laser track

is 300 µm, which is the same width of the hardness pattern shown in Figure 12. As depth

into the sample increases, the hardness value decreases sharply.

The sample was sectioned and polished in preparation for hardness measurements, and

then the sample was indented with a uniform grid of 7 rows and 13 columns (with grid

spacing of 100 µm), as shown in Figure 13. The automatic measurements were made

with a Vickers indenter, a load of 200 gram force, and a load time of 13 seconds. To

show repeatability, the same measurement grid was performed at three separate locations

along the sample cross-section.

20
Hardness values predicted by the kinetic model are compared with the measured

values in Figure 14. Each hardness comparison graph shows a line of measured and

predicted values at given depths of 100 µm, 300 µm, and 500 µm below the sample’s

surface. The measured values shown in these graphs are the average values of the three

separate measurements and the error bars show the maximum and minimum values of the

three measurements at each location. The predicted values match the magnitude and the

general trend of the horizontally-varying patterns of the measured values.

A section view of the predicted transverse stress field after the simulation of the

2nd layer is shown in Figure 15 (transverse stress component is in the direction

perpendicular to the laser travel direction of the top layer and parallel to the surface).

Two lines are shown on Figure 15: the surface line and the depth line. To validate

the residual stress model, the same transverse stress component was measured using the

X-ray diffraction technique on two different LDD samples (two samples to show

repeatability) at 5 and 7 locations, respectively, along the depth line of the two-layer

samples. The transverse stresses of the samples were measured at the surface where the

surface line and the depth line intersect; then the samples were electropolished (so that no

additional stress was introduced) to subsequent depths along the depth line. The

transverse stress was measured at each depth. The predicted transverse stress values

along the depth line are compared to measured transverse stress values in Figure 16. The

predicted transverse stress values along the surface line are compared to the measured

average residual stress value at the surface in Figure 17.

The predicted stress values match reasonably well with measured values. From

Figure 16 and Figure 17, it is clear that a compressive stress exists near the surface of the

21
sample that increases with depth to a value of -400 MPa. As depth increases, the residual

stress becomes tensile up to a value of +400 MPa. This general trend of compressive

stress followed by tensile stress as depth increases is primarily due to thermal strains

experienced in the sample. As shown in Figure 17, the compressive stress varies between

-50 MPa and -200 MPa along the surface, and is supported by the upper and lower

bounds of the two measured values. The two XRD measurements performed on the

surface of the sample resulted in values of -107 MPa and -94 MPa with error bounds of

±7 MPa and ±45 MPa, respectively, as indicated by the upper and lower bounds for each

measurement in Figure 17. The horizontal variation in residual stress could not be

validated by the XRD method because the smallest beam size available in the XRD

apparatus (1 mm) was larger than the length scale of the predicted variation. Therefore,

the measured residual stress values on the surface are an average of the residual stress

within the spot covered by the XRD beam.

As noted in

22
Table 4, the transformation from austenite to martensite is accompanied by a

1.03% increase in volume. This explains the strong compressive stresses within the first

0.5 mm of the surface of the sample where phase transformation to martensite occurred

throughout the heat-affected zone. Additionally, the tempering process is accompanied by

a smaller decrease in volume of 0.81% or 0.71%, which helps explain the weaker

compressive stresses closer to the surface of the sample where stronger tempering occurs,

and that the periodic variation of stress along the surface is due to the periodic tempering

of the martensite from the multiple tracks of the 2nd layer.

In order to test the strength of laser-deposited H13 tool steel, three test specimens

were built via LDD with the same operating parameters as those listed in Table 1. Each

sample is a cylinder with a diameter of 12.7 mm and a height of at least 90 mm, and was

machined on a lathe to meet the ASTM E8 tensile testing standard (ASTM, 2001). The

reduced section for each sample is 6.35 mm diameter and 31.75 mm long. The samples

were tested in an MTS 793 tensile testing machine at a rate of 5.0 mm/min. The

stress/strain curve of each sample is shown in Figure 18 .

According to the test data, the ultimate tensile strength (UTS) of the three test

samples are 2035 MPa, 2100 MPa, and 2081 MPa, with an average UTS of 2072 MPa.

Published UTS values for commercial H13 hot-work die steel (AZoM Materials, 2013;

Bohler Uddeholm, 2013; Ellwood Specialty Steels, 2016) vary between 1600—1800

MPa. Laser-deposited H13 tool steel, therefore, shows a UTS value between 15% and

30% stronger than commercial H13 tool steel.

5. Conclusion

23
Optimal operating parameters were experimentally determined for multi-track and

multi-layer laser direct deposition (LDD) of H13 tool steel in order to characterize the

laser-deposited material and to validate a predictive residual stress model. A finite

element model was built and utilized to predict the phase transformation kinetics and

residual stresses of multi-track multi-layer LDD. Predicted tempering patterns and

predicted variation of hardness across deposition tracks matched well with experimental

images and measured values. For a two-layer simulation, predicted strong compressive

stresses of -400 MPa at a depth of 0.5 mm and strong tensile stress of +400 MPa at a

depth of 1.0 mm were confirmed by measurement using XRD. Compressive stresses near

the surface are primarily caused by phase transformational strains when high phase

fractions of martensite result from the thermal processing of laser deposition. Small

variations in compressive stresses along the top surface of the deposited tracks, most

likely due to the tempering of the martensite from the heat of additional laser tracks, were

also predicted and verified by micrograph. The ultimate tensile strength of laser-

deposited H13 was shown to be 2030 MPa, a value of 15% to 30% stronger than

commercial H13 tool steel.

24
References

Alimardani, M., Toyserkani, E., Huissoon, J.P.,;1; 2007. A 3D dynamic numerical


approach for temperature and thermal stress distributions in multilayer laser solid
freeform fabrication process. Opt. Lasers Eng. 45, 1115–1130.
ASTM, E.,;1; 2001. Standard test methods for tension testing of metallic materials. Annu.
B. ASTM Stand. ASTM.

AZoM Materials,;1; 2013. H13 Tool Steel [WWW Document]. URL


http://www.azom.com/article.aspx?ArticleID=9107

Bailey, N.S., Tan, W., Shin, Y.C.,;1; 2009. Predictive modeling and experimental results
for residual stresses in laser hardening of AISI 4140 steel by a high power diode laser.
Surf. Coatings Technol. 203, 2003–2012.

Benedyk, J.C.,;1; 2008. H-13, High Performance Alloys Database.

Bohler-Uddeholm Corporation,;1; H13 Tool Steel data sheet [WWW Document], 2016.
URL http://www.bucorp.com/media/H13_data_sheet_09032013.pdf

Bohler Uddeholm,;1; 2013. Bohler-Uddeholm H13 Tool Steel.

Brückner, F., Lepski, D., Beyer, E.,;1; 2007. Modeling the influence of process
parameters and additional heat sources on residual stresses in laser cladding. J. Therm.
spray Technol. 16, 355–373.

Callister, W.D., Rethwisch, D.G.,;1; 2007. Materials science and engineering: an


introduction. Wiley New York.

Das, S., Upadhya, G., Chandra, U.,;1; 1992. Prediction of Macro- and Micro-Residual
Stress in Quenching Using Phase Transformation Kinetics. Quenching Distortion Control
July, 229–234.

El Cheikh, H., Courant, B., Hascoët, J.-Y., Guillén, R.,;1; 2012. Prediction and analytical
description of the single laser track geometry in direct laser fabrication from process
parameters and energy balance reasoning. J. Mater. Process. Technol. 212, 1832–1839.

Ellwood Specialty Steels,;1; 2016. Exell H13 Premium [WWW Document]. URL
http://ess.elwd.com/product/exell-h13-premium/

Farahmand, P., Kovacevic, R.,;1; 2014. An experimental--numerical investigation of heat


distribution and stress field in single-and multi-track laser cladding by a high-power
direct diode laser. Opt. Laser Technol. 63, 154–168.

25
Fathi, A., Toyserkani, E., Khajepour, A., Durali, M.,;1; 2006. Prediction of melt pool
depth and dilution in laser powder deposition. J. Phys. D. Appl. Phys. 39, 2613.

Ghosh, S., Choi, J.,;1; 2005. Three-dimensional transient finite element analysis for
residual stresses in the laser aided direct metal/material deposition process. J. Laser Appl.
17, 144–158.

He, X., Fuerschbach, P.W., DebRoy, T.,;1; 2003. Heat transfer and fluid flow during
laser spot welding of 304 stainless steel. J. Phys. D. Appl. Phys. 36, 1388.

He, X., Yu, G., Mazumder, J.,;1; 2009. Temperature and composition profile during
double-track laser cladding of H13 tool steel. J. Phys. D. Appl. Phys. 43, 15502.

Ibarra-Medina, J., Vogel, M., Pinkerton, A.J.,;1; 2011. A CFD model of laser cladding:
From deposition head to melt pool dynamics. Proc. 30th Int. Congr. Appl. Lasers Electro-
optics 378–386.

Jhavar, S., Paul, C.P., Jain, N.K.,;1; 2013. Causes of failure and repairing options for dies
and molds: A review. Eng. Fail. Anal. 34, 519–535.

Jin, L.,;1; 2001. Simulation of Quenching and Temperaing of Steels. Purdue University.

Jung, M., Lee, S.-J., Lee, Y.-K.,;1; 2009. Microstructural and dilatational changes during
tempering and tempering kinetics in martensitic medium-carbon steel. Metall. Mater.
Trans. A 40, 551–559.

Kamara, a. M., Marimuthu, S., Li, L.,;1; 2011. A Numerical Investigation into Residual
Stress Characteristics in Laser Deposited Multiple Layer Waspaloy Parts. J. Manuf. Sci.
Eng. 133, 31013.

Kang, S.-H., Im, Y.-T.,;1; 2005. Three-dimensional finite-element analysis of the


quenching process of plain-carbon steel with phase transformation. Metall. Mater. Trans.
A 36, 2315–2325. doi:10.1007/s11661-005-0104-5

Labudovic, M., Hu, D., Kovacevic, R.,;1; 2003. A three dimensional model for direct
laser metal powder deposition and rapid prototyping. J. Mater. Sci. 38, 35–49.

Lakhkar, R.S., Shin, Y.C., Krane, M.J.M.,;1; 2008. Predictive modeling of multi-track
laser hardening of AISI 4140 steel. Mater. Sci. Eng. A 480, 209–217.

Lalas, C., Tsirbas, K., Salonitis, K., Chryssolouris, G.,;1; 2007. An analytical model of
the laser clad geometry. Int. J. Adv. Manuf. Technol. 32, 34–41.

Lin, Y., McHugh, K.M., Zhou, Y., Lavernia, E.J.,;1; 2007. Modeling the spray forming
of H13 steel tooling. Metall. Mater. Trans. A 38, 1632–1637.

26
Majumdar, J.D., Manna, I.,;1; 2012. Laser-assisted fabrication of materials. Springer
Science & Business Media.

Manvatkar, V.D., Gokhale,;1; a. a., Reddy, G.J., Savitha, U., De, A., 2015. Investigation
on laser engineered net shaping of multilayered structures in H13 tool steel. J. Laser
Appl. 27, 32010. doi:10.2351/1.4921493

Mazumder, J., Conde, O., Vilar, R., Steen, W.,;1; 2012. Laser processing: surface
treatment and film deposition. Springer Science & Business Media.

Oddy, A.S., Goldak, J.A., McDill, J.M.J.,;1; 1989. Transformation effects in the 3D finite
element analysis of welds, in: Proceedings of the Second International Conference on
Trends in Welding Research. pp. 15–19.

Philip, T. V, McCaffrey, T.J.,;1; 1997. Ultrahigh strength steels. Mech. Eng. YORK
BASEL-MARCEL DEKKER- 149–162.

Pinkerton, A.J.,;1; 2015. Advances in the modeling of laser direct metal deposition. J.
Laser Appl. 27, S15001. doi:10.2351/1.4815992

Pinkerton, A.J., Wang, W., Li, L.,;1; 2008. Component repair using laser direct metal
deposition 222, 827–836. doi:10.1243/09544054JEM1008

Roberts, G., Krauss, G., Kennedy, R.,;1; 1998. Tool Steels, 5th ed. ASM International,
Materials Park, OH.

Santhanakrishnan, S., Kong, F., Kovacevic, R.,;1; 2011. An experimentally based


thermo-kinetic hardening model for high power direct diode laser cladding. J. Mater.
Process. Technol. 211, 1247–1259.

Shatla, M., Kerk, C., Altan, T.,;1; 2001. Process modeling in machining. Part I:
determination of flow stress data. Int. J. Mach. Tools Manuf. 41, 1511–1534.

Skvarenina, S., Shin, Y.C.,;1; 2006. Predictive modeling and experimental results for
laser hardening of AISI 1536 steel with complex geometric features by a high power
diode laser. Surf. Coatings Technol. 201, 2256–2269.

Telasang, G., Dutta Majumdar, J., Padmanabham, G., Tak, M., Manna, I.,;1; 2014. Effect
of laser parameters on microstructure and hardness of laser clad and tempered AISI H13
tool steel. Surf. Coatings Technol. 258, 1108–1118. doi:10.1016/j.surfcoat.2014.07.023

Thompson, S.,;1; 1999. Handbook of Mould, Tool and Die Repair Welding, Handbook of
Mould, Tool and Die Repair Welding. Elsevier. doi:10.1533/9781855738324.37

Tušek, J., Ivančič, R.,;1; 2004. Computer-aided analysis of repair welding of stamping
tools. Zeitschrift f{ü}r Met. 95, 8–13.

27
Wang, L., Felicelli, S.D., Pratt, P.,;1; 2008. Residual stresses in LENS-deposited AISI
410 stainless steel plates. Mater. Sci. Eng. A 496, 234–241.

Wen, S., Shin, Y.C.,;1; 2011. Comprehensive predictive modeling and parametric
analysis of multitrack direct laser deposition processes. J. Laser Appl. 23, 22003.

Wen, S., Shin, Y.C.,;1; 2010. Modeling of transport phenomena during the coaxial laser
direct deposition process. J. Appl. Phys. 108.

Woodard, P.R., Chandrasekar, S., Yang, H.T.Y.,;1; 1999. Analysis of temperature and
microstructure in the quenching of steel cylinders. Metall. Mater. Trans. B 815–822.

Zhao, J.Z., Mesplont, C., Cooman, B.C. De,;1; 2001. Kinetics of Phase Transformations
in Steels: A New Method for Analysing Dilatometric Results. ISIJ Int. 41, 492–497.

Zhao, J.Z., Mesplont, C., De Cooman, B.C.,;1; 2002. Quantitative analysis of the
dilatation during an isothermal decomposition of austenites. Mater. Sci. Eng. A 332, 110–
116.

28
Figure 1: Computational model flow chart.

Figure 2: Isocontours at the conclusion of simulation for the 1st track deposited on H13
substrate.

29
Figure 3: Molten pool (dotted black), heat affected zone (dashed black), and deposited
track geometry (solid red line) for single track deposition simulation compared with
experiment.

Figure 4: Two-track cross-sectional profile showing the extracted track geometry for the
initial seeding of the three-track simulation.

30
Figure 5: Results of the three-track simulation – pseudo-steady-state surface during
deposition.

Figure 6: Molten pool (dashed black line) and deposited track geometry (solid red line)
for the three-track deposition simulation compared with experiment.

31
Figure 7: Young's modulus and Poisson's ratio from Bohler-Uddeholm Corporation

(2016) and Philip and McCaffrey (1997) as functions of temperature.

32
Figure 8: Temperature field during 3rd track.

Figure 9: Phase fraction of martensite during the 3rd track of the 1 st layer.

33
Figure 10: Phase fraction of martensite during the 3rd (top) and 7 th (bottom) tracks of the
2nd layer.

34
Figure 11: Experimental result showing tempering pattern similar to the predicted
tempering pattern in Figure 10.

Figure 12: Predicted hardness field.

35
Figure 13: Cross-section of a single-layer LDD sample showing the hardness
measurement grid.

36

You might also like