+ All Categories
Home > Documents > Nonlinear diffusion, bonding, and mechanics of the...

Nonlinear diffusion, bonding, and mechanics of the...

Date post: 08-Jul-2020
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
7
1464 | Phys. Chem. Chem. Phys., 2019, 21, 1464--1470 This journal is © the Owner Societies 2019 Cite this: Phys. Chem. Chem. Phys., 2019, 21, 1464 Nonlinear diffusion, bonding, and mechanics of the interface between austenitic steel and ironQin Qin,* a Wei He, a Lu Xie, a Junchao Deng, a Xuehui Zhu a and Qing Peng * b We investigate the atomic diffusivity and mechanics of the interface between bulk austenitic steel (fcc structure) and bcc iron at various temperatures and strain rates using molecular dynamics simulations. We adopt the system of Fe 74 Cr 16 Ni 10 corresponding to 316L steel as a representative model of austenitic steels, denoted as FeCrNi. We find that the compressive strength of the FeCrNi/Fe system decreases by 44.3% and the corresponding strain decreases by 7.2% when the temperature increases from 1500 K to 1800 K. The temperature enhances nonlinearly the diffusion of interfacial atoms and improves the cohesion of FeCrNi/Fe by forming a thicker diffusion layer, of which the thickness increases by 56.3% when the temperature increases from 1600 K to 1700 K, and by nearly 48% when the temperature increases from 1700 K to 1800 K. However, the thickness of the diffusion layer decreases by 33.3% when the compressive strain rate increases from 1 10 9 s 1 to 4 10 9 s 1 . Our study sheds light on the atomistic mechanism of the interfaces of bimetals and might be helpful in optimizing the process of the fabrication of bimetal composites. Introduction Bimetallic composites are widely used in aerospace, machinery, chemical, power and electronics industries due to their superior properties including strength, corrosion resistance, electron con- ductivity, and thermal conductivity. The interface of bimetallic composites has also become one of the important research areas of metal composites because the microstructure and strength of the interface of bimetallic composites are key factors influencing the composite’s quality and properties. Both experimental and numerical methods are used to investigate bimetallic interfaces. For example, the experimental method was adopted to observe and analyze the microstructure of the interface. 1,2 For numerical modeling, a binding model or cohesive model was adopted to simulate the interface. 3 Despite these efforts, a clear cohesion mechanism of the different metals is still lacking. The prediction of the bonding strength of the interface is desirable. An atomistic model with two metals forming an interface is indispensable to study the cohesion mechanism of the interface. Despite extensive studies on both single crystals 4–13 and bi-crystals, 14–21 the aforementioned simulations are carried out on a pure metal phase, i.e. a single element in one structure. Few studies have been reported on two kinds of alloys at high temperature with multiple elements within one alloy. To the authors’ best knowledge, no attempt has been made to simulate the interfaces, especially the cohesion mechanism, of austenitic steels and ferritic steels, due to the complexity of the problems. We are particularly interested in the common commercial 316L steel (austenitic) and Q345R carbon steel (ferritic) owing to their balanced mechanical properties, corrosion resistance, and attrac- tive price. However, it is a great challenge to perform the accurate modeling of their interfaces. Part of the reason might be the fact that various kinds of elements (the details are shown in Table S1 in the ESI) are present in both 316L and Q345R, presenting a great challenge in atomically modeling 316L and Q345R in the same simulation system. Creatively and for the first time, we adopt the system of Fe 74 Cr 16 Ni 10 corresponding to 316L steel as a representative model of austenitic steels, denoted as FeCrNi. The other components in 316L stainless steel are ignored because they have trivial effect on the mechanical properties owing to their low concentration. We model the Q345R steel as pure Fe because Fe is the predominant element with over 98 at%. The mechanical properties and diffusion behaviors, and the effect of temperature and compression strain rate on the compression mechanism and interfacial diffusion have been comprehensively investigated using a FeCrNi/Fe model. Methodologies Atomic model of the FeCrNi/Fe interface The preparation for the initial configurations of the atomic FeCrNi/Fe model consists of three steps. First, pure Fe with a a School of Mechanical Engineering, University of Science and Technology Beijing, Beijing 100083, China. E-mail: [email protected] b Nuclear Engineering and Radiological Sciences, University of Michigan, Ann Arbor, MI 48109, USA. E-mail: [email protected] Electronic supplementary information (ESI) available. See DOI: 10.1039/ c8cp07123c Received 18th November 2018, Accepted 12th December 2018 DOI: 10.1039/c8cp07123c rsc.li/pccp PCCP PAPER
Transcript
Page 1: Nonlinear diffusion, bonding, and mechanics of the ...qpeng.org/publication/pdf/Peng_2019_pccp_steel_interface.pdf · Nonlinear diffusion, bonding, and mechanics of ... bi-crystals,14–21

1464 | Phys. Chem. Chem. Phys., 2019, 21, 1464--1470 This journal is© the Owner Societies 2019

Cite this:Phys.Chem.Chem.Phys.,

2019, 21, 1464

Nonlinear diffusion, bonding, and mechanics ofthe interface between austenitic steel and iron†

Qin Qin,*a Wei He,a Lu Xie, a Junchao Deng,a Xuehui Zhua and Qing Peng *b

We investigate the atomic diffusivity and mechanics of the interface between bulk austenitic steel (fcc structure)

and bcc iron at various temperatures and strain rates using molecular dynamics simulations. We adopt the

system of Fe74Cr16Ni10 corresponding to 316L steel as a representative model of austenitic steels, denoted as

FeCrNi. We find that the compressive strength of the FeCrNi/Fe system decreases by 44.3% and the

corresponding strain decreases by 7.2% when the temperature increases from 1500 K to 1800 K. The

temperature enhances nonlinearly the diffusion of interfacial atoms and improves the cohesion of FeCrNi/Fe

by forming a thicker diffusion layer, of which the thickness increases by 56.3% when the temperature

increases from 1600 K to 1700 K, and by nearly 48% when the temperature increases from 1700 K to 1800 K.

However, the thickness of the diffusion layer decreases by 33.3% when the compressive strain rate increases

from 1 � 109 s�1 to 4 � 109 s�1. Our study sheds light on the atomistic mechanism of the interfaces of

bimetals and might be helpful in optimizing the process of the fabrication of bimetal composites.

Introduction

Bimetallic composites are widely used in aerospace, machinery,chemical, power and electronics industries due to their superiorproperties including strength, corrosion resistance, electron con-ductivity, and thermal conductivity. The interface of bimetalliccomposites has also become one of the important research areasof metal composites because the microstructure and strength ofthe interface of bimetallic composites are key factors influencingthe composite’s quality and properties. Both experimental andnumerical methods are used to investigate bimetallic interfaces.For example, the experimental method was adopted to observeand analyze the microstructure of the interface.1,2 For numericalmodeling, a binding model or cohesive model was adopted tosimulate the interface.3 Despite these efforts, a clear cohesionmechanism of the different metals is still lacking. The predictionof the bonding strength of the interface is desirable. An atomisticmodel with two metals forming an interface is indispensable tostudy the cohesion mechanism of the interface.

Despite extensive studies on both single crystals4–13 andbi-crystals,14–21 the aforementioned simulations are carried outon a pure metal phase, i.e. a single element in one structure. Fewstudies have been reported on two kinds of alloys at high

temperature with multiple elements within one alloy. To theauthors’ best knowledge, no attempt has been made to simulatethe interfaces, especially the cohesion mechanism, of austeniticsteels and ferritic steels, due to the complexity of the problems.We are particularly interested in the common commercial 316Lsteel (austenitic) and Q345R carbon steel (ferritic) owing to theirbalanced mechanical properties, corrosion resistance, and attrac-tive price. However, it is a great challenge to perform the accuratemodeling of their interfaces. Part of the reason might be the factthat various kinds of elements (the details are shown in Table S1in the ESI†) are present in both 316L and Q345R, presenting agreat challenge in atomically modeling 316L and Q345R in thesame simulation system. Creatively and for the first time, weadopt the system of Fe74Cr16Ni10 corresponding to 316L steel as arepresentative model of austenitic steels, denoted as FeCrNi. Theother components in 316L stainless steel are ignored because theyhave trivial effect on the mechanical properties owing to their lowconcentration. We model the Q345R steel as pure Fe because Fe isthe predominant element with over 98 at%. The mechanicalproperties and diffusion behaviors, and the effect of temperatureand compression strain rate on the compression mechanism andinterfacial diffusion have been comprehensively investigatedusing a FeCrNi/Fe model.

MethodologiesAtomic model of the FeCrNi/Fe interface

The preparation for the initial configurations of the atomicFeCrNi/Fe model consists of three steps. First, pure Fe with a

a School of Mechanical Engineering, University of Science and Technology Beijing,

Beijing 100083, China. E-mail: [email protected] Nuclear Engineering and Radiological Sciences, University of Michigan, Ann Arbor,

MI 48109, USA. E-mail: [email protected]

† Electronic supplementary information (ESI) available. See DOI: 10.1039/c8cp07123c

Received 18th November 2018,Accepted 12th December 2018

DOI: 10.1039/c8cp07123c

rsc.li/pccp

PCCP

PAPER

Page 2: Nonlinear diffusion, bonding, and mechanics of the ...qpeng.org/publication/pdf/Peng_2019_pccp_steel_interface.pdf · Nonlinear diffusion, bonding, and mechanics of ... bi-crystals,14–21

This journal is© the Owner Societies 2019 Phys. Chem. Chem. Phys., 2019, 21, 1464--1470 | 1465

BCC crystal structure is built. Second, the atoms of FeCrNi arerandomly positioned at the FCC crystal structure accordingto the concentration proportion. Finally, the two separatedmodels, Fe and FeCrNi, are placed together inside a simulationbox with dimensions of 13.5(X) � 13.5(Y) � 27(Z) nm3. Thecontact surfaces of Fe and FeCrNi are both on (0 0 1) planes,as shown in Fig. 1. There are 355 023 atoms in this model, with172 773 atoms in Fe and 182 250 atoms in FeCrNi.

Molecular dynamics simulations

The molecular dynamics (MD) method has been well establishedover the last several decades and becomes an indispensable toolfor various investigations.13 The first-principles calculation basedempirical embedded atom method (EAM) potential was adoptedin MD simulations throughout this study.22,23 The total EAMenergy E can be expressed as a multi-body potential energyfunction in the following form

E ¼ 1

2

XN

i;j¼1jai

fa;b rij� �þXN

i¼1Fti rið Þ (1)

where f(rij) is the short-range pair energy which is defined as afunction of the interatomic distance rij between atoms i and j.F(ri) is the embedding energy which is dependent on the localelectron density, ri. Here N represents the total number ofatoms in the system, and a and b denote element types ofatoms i and j. And the local electron density around atom i,contributed by its neighbours, is given as

ri ¼XN

j¼1jai

jtjrij� �

(2)

where j denotes the electron density function of the consideredelement. Thus, for the FeCrNi ternary system, twelve functions

need to be defined: jFe, jCr, jNi, FFe, FCr, FNi, VFeFe, VCrCr, VNiNi,VFeCr, VFeNi, and VCrNi. The parametrization and the fittedparameters are from these literature studies.22,23

To make sure that this potential can afford the high-temperature study in this report, this potential has been usedto measure the melting points of Fe and Fe74Cr16Ni10, respectively.As detailed in the ESI,† the melting point was 1868 K for Fe and2123 K for Fe74Cr16Ni10. Obviously, this potential can afford thehigh-temperature (1500 K–1800 K) study in this report.

Uniaxial compression loading

There are four regions in our model. As shown in Fig. 1, fixedlayers contain three layers of atoms at the top of Fe and thebottom of FeCrNi, respectively. The initial thermal velocities ofatoms follow the Maxwell distribution. Periodic boundary con-ditions are implemented in the x and y directions. The velocity-Verlet algorithm24 is applied to integrate the motion equationfor the atoms in the structure with a constant time step of1 � 10�15 s. In order to ensure the interface combination athigh temperature, before compression loading, the system isrelaxed sufficiently to obtain an equilibrium state at 1500 Kwith the isobaric–isothermal (NPT) ensemble. After relaxation,the system is uniaxially compressed at a constant strain rate ofloading along the z direction of about 3.0 � 109 s�1.

In order to study how the temperature and strain rate affectthe compression process and interfacial diffusion, eight sets ofsimulations have been carried out. As shown in Table 1, there arefour models uniaxially compressed at a constant strain rate of3.0 � 109 s�1 with four temperatures (1500 K, 1600 K, 1700 K and1800 K), and another four models uniaxially compressed at 1500 Kwith four strain rates (1.0 � 109 s�1, 1.5 � 109 s�1, 3.0 � 109 s�1

and 4.0 � 109 s�1). All molecular dynamics (MD) simulations areperformed by using LAMMPS.25 OVITO26 visualization softwarewas used to visualize the atomic structures, generate simulationsnapshots, and carry out dislocation analysis (DXA).27

Results and discussionStress–strain curve under uniaxial compression

The uniaxial compression process for FeCrNi and Fe has beendiscussed in detail at a temperature of 1500 K with a constantstrain rate of 3 � 109 s�1. The compressive stress–strain curvein this process is shown in Fig. 2. Coherency strains naturallydevelop in both layers because of the lattice mismatch28,29

between FeCrNi and Fe when creating the structure with acoherent interface. Therefore, the stress at the initial stage ofthis process fluctuates around zero. Then the stress from pointa to point c shows a linear increase when the strain increases.The maximum stress is 4.74 GPa when the strain at point c

Fig. 1 Initialization configuration of the FeCrNi/Fe model (the top layer isFe and the bottom layer is FeCrNi; : Fe atom in the Fe, : Fe atom in theFeCrNi, : Cr atom in the FeCrNi, : Ni atom in the FeCrNi). For theremainder of this paper, the Fe atoms in the Fe are referred to as cFe(Fe in carbon steel), and the Fe atoms in the FeCrNi are referred to as sFe(Fe in stainless steel).

Table 1 Compression conditions of different models

Models 1 2 3 4 5 6 7 8

Temperature (K) 1500 1600 1700 1800 1500 1500 1500 1500strain rate (109 s�1) 3 3 3 3 1 1.5 3 4

Paper PCCP

Page 3: Nonlinear diffusion, bonding, and mechanics of the ...qpeng.org/publication/pdf/Peng_2019_pccp_steel_interface.pdf · Nonlinear diffusion, bonding, and mechanics of ... bi-crystals,14–21

1466 | Phys. Chem. Chem. Phys., 2019, 21, 1464--1470 This journal is© the Owner Societies 2019

reaches 0.16. After the apex, the stress value from point c topoint d decreases rapidly. Then the values of compressive stressdecrease from point d to point f until the strain reaches 0.3 at f.

We examine the atomic structures at the different stagesduring the compression process. As shown in Fig. 3(a–c), noapparent interfacial diffusion takes place with strain up to 0.08.However, the atoms in FeCrNi and Fe have diffused graduallyinto their opposite side in Fig. 3(d–f). In addition, the structurenear the interface begins to exhibit an amorphous structure asthe strain increases beyond 0.19.

In order to understand the compression mechanism ofFeCrNi/Fe, the dislocation evolution is examined based onthe dislocation extraction analysis. The dislocation structures

at various strain stages during the compression process areshown in Fig. 3(g–l). The dislocation multiplication anddislocation tangle are evident in Fig. 3(g and h). These phenomenaindicate that the total dislocation length increases until the strainreaches 0.11. However, the total dislocation length decreases whenthe strain goes beyond, Fig. 3(i) to Fig. 3(l). As shown in Fig. 2,the total dislocation length has a positive correlation with thestress–strain curve. This could be understood as follows.The dislocation multiplication and dislocation tangle resultedin high-density dislocations and created new barriers to dis-location slip. Therefore, a greater compressive stress is requiredfor further plastic deformation. Then the total dislocationdecreases in Fig. 3(i–l) because the aggravating amorphizationoccurs in both FeCrNi and Fe with the strain increasing at hightemperature. In addition, the dislocation mainly occurs in theFe layer from Fig. 3. According to Fig. 3(g), it can be concludedthat the plastic deformation begins before the strain 0.08 in Fewhile FeCrNi is still in the elastic deformation stage. Thus, thestress–strain curve presents a linear increase with increasingstrain in the initial deformation stage (e o 0.16).

The temperature has a great influence on the compressionprocess of the FeCrNi/Fe system. The stress–strain curves andtotal dislocation length variation at temperatures of 1500 K,1600 K, 1700 K and 1800 K are illustrated in Fig. 4(a). Thesnapshots of interfacial diffusion at various temperatures areillustrated in Fig. S1 in the ESI.† A higher compression tem-perature results in a lower compressive stress. The stress–straincurves at each temperature show a similar trend. The fluctuat-ing stress gradually increases during the elastic stage anddecreases rapidly after the stress reaches its maximum valuewhen the strain increases. During the compression process ofFeCrNi and Fe at high temperature, the compressive strength of

Fig. 2 Stress–strain curve (red line) and total dislocation length (blue line)at 1500 K, where points a to f in the stress–strain curve correspond to thestrain values of (a) 0.08, (b) 0.11, (c) 0.16, (d) 0.19, (e) 0.25 and (f) 0.30,respectively.

Fig. 3 Snapshots of the FeCrNi/Fe model (a–f) and total dislocation length evolution (g–l) at various stages of uniaxial compression. The strains are 0.08,0.11, 0.16, 0.19, 0.25 and 0.30, respectively. Dislocation segments were extracted by DXA, colored by the magnitude of the Burgers vector. The green linerepresents a Shockley partial dislocation, and the red line represents other dislocations. The total dislocation length is 3067 Å, 10 078 Å, 2748 Å, 2583 Å,2413 Å, and 2960 Å for g–l, respectively.

PCCP Paper

Page 4: Nonlinear diffusion, bonding, and mechanics of the ...qpeng.org/publication/pdf/Peng_2019_pccp_steel_interface.pdf · Nonlinear diffusion, bonding, and mechanics of ... bi-crystals,14–21

This journal is© the Owner Societies 2019 Phys. Chem. Chem. Phys., 2019, 21, 1464--1470 | 1467

the FeCrNi/Fe system decreases by 44.3% and the corres-ponding strain decreases by 7.2% when the temperatureincreases from 1500 K to 1800 K. The maximum stress of theFeCrNi/Fe system reduced, and the maximum stress pointfollowed in advance at the same time when the temperatureincreased from 1600 K to 1800 K (Fig. 4(a)). For example, themaximum stress value reduces by 17.6% and the correspondingstrain reduces by 0.91% as the temperature increases from1600 K to 1700 K. Upon increasing the temperature from 1600 Kto 1800 K, the maximum stress value decreases by 33.5% andthe corresponding strain increases by about 10.1%. The tem-perature enhances plasticity and promotes the combination ofFeCrNi and Fe. A higher compression temperature results in alower total dislocation length in Fig. 4(c). The higher degreeof amorphization at a higher temperature leads to a decreaseof the total dislocation length.

Strain rate is an important factor regarding the dynamicalbehavior of a system. The effect of strain rate on the compres-sion process has also been examined. The stress–strain curvesobtained at four different compression strain rates at a tem-perature of 1500 K are shown in Fig. 4(b), which are 1 � 109 s�1,1.5 � 109 s�1, 3 � 109 s�1 and 4 � 109 s�1. The snapshots ofinterfacial diffusion at various strain rates are illustrated inFig. S3 in the ESI.† The strain–strain curves at different strainrates are in coincidence at the initial deformation stage.The maximum stress value is about 4.7 GPa and occurs nearthe strain of 0.16 when the strain rate is 1 � 109 s�1, 1.5 �109 s�1, 3 � 109 s�1 and 4 � 109 s�1. Then the stress–strain curvedecreases rapidly when the strain exceeds the maximum stresspoint. However, the stress–strain curve decreases slightly slowly ata higher strain rate. Finally, the stress value decreases until thestrain reaches 0.3. As shown in Fig. 4(d), a higher compressionstrain rate results in a lower total dislocation length.

Diffusion across interfaces

We observed an atomic diffusion near the interface betweenFeCrNi and Fe during the compression stage. The diffusion

layer specifies that the atomic concentrations of both FeCrNi andFe along the compression direction (Z axis) are more than 5%.15,20

The thickness of the diffusion layer as a function of time at 1500 Kis shown in Fig. 5(a). The thickness of the diffusion layer graduallyincreased to 6.5 Å. Fig. 5(b) shows the mean square displacement(MSD) as a function of time for sFe, Cr, Ni and cFe, which clearlymanifested that FeCrNi was more diffusive. By using the Einsteinrelation,30 it was possible to find the diffusivity of a particularspecies from its time vs. MSD plot. The diffusivity of sFe wasfound to be 6.7178� 10�10 m2 s�1. The diffusivity of Cr was foundto be 6.9745 � 10�10 m2 s�1. The diffusivity of Ni was found to be6.5144 � 10�10 m2 s�1. The diffusivity of cFe was found to be5.1133 � 10�10 m2 s�1. Obviously, the diffusivity of Fe in carbonsteel was smaller than that of Fe, Cr and Ni in stainless steel.Therefore, the main phenomenon that occurred was the diffusionof 316L into Q345R. Meanwhile, the diffusivity of Cr was higherthan that of Ni in stainless steel. Therefore, the diffusion of Cr wasfaster than that of Ni and was more likely to invade first when316L diffused into Q345R because the concentration gradient ofCr on both sides of the interface was higher than that of Ni, asshown in Fig. 5(c). In the region of 316L, the ratio of sFe atoms, Cratoms and Ni atoms was 74 : 16 : 10, while in the diffusion layer,the ratio of sFe atoms, Cr atoms and Ni atoms was 73.5 : 16.3 : 10.2.Therefore, Cr, Ni and Fe in 316L diffused through the originalinterface to Q345R when FeCrNi diffused into Fe. And the amountof diffusion of Cr atoms was higher than that of Ni atoms, whichwas consistent with the conclusion obtained by MSD. Goodagreement was obtained in comparison with experimental results,which were detailed in the ESI.†

The thicknesses of the diffusion layer at various tempera-tures are shown in Fig. 6a, which manifests that temperature

Fig. 4 Stress–strain curves and total dislocation length variation atdifferent temperatures and stain rates, (a and b) exhibit stress–straincurves; (c and d) exhibit total dislocation length variation.

Fig. 5 The thickness of the diffusion layer (a) and MSD (b) as a function oftime at 1500 K with a constant strain rate of 3 � 109 s�1; (c) atomicconcentration distribution of position Z at 1 � 10�10 s (snapshots of theFeCrNi/Fe model at 0 s, 3.3 � 10�11 s, 6.6 � 10�11 s and 1 � 10�10 s are alsoshown, in which the diffusion layer is marked in grey).

Paper PCCP

Page 5: Nonlinear diffusion, bonding, and mechanics of the ...qpeng.org/publication/pdf/Peng_2019_pccp_steel_interface.pdf · Nonlinear diffusion, bonding, and mechanics of ... bi-crystals,14–21

1468 | Phys. Chem. Chem. Phys., 2019, 21, 1464--1470 This journal is© the Owner Societies 2019

plays an important role in the interface atomic diffusion. Andthe atomic concentration distribution of position Z at varioustemperatures is illustrated in Fig. S2 in the ESI.† The thicknessof the diffusion layer increases exponentially with increasingtemperature, where the thickness of the diffusion layer is 6.5 Å,8 Å, 12.5 Å and 18.5 Å when the temperature is 1500 K, 1600 K,1700 K and 1800 K, respectively. For the temperature rangefrom 1500 K to 1600 K, the thickness of the diffusion layerincreases from 6.5 to 8 Å, with an increment of 1.5 Å. Incontrast, when the temperature increases from 1600 K to1700 K, the thickness of the diffusion layer abruptly increasesto 12.5 Å, with an increment of 4.5 Å, which is 3 times that ofthe previous rate. Regarding the percentage of growth, thethickness of the diffusion layer increases by 23.1% when thetemperature increases from 1500 K to 1600 K and by nearly56.3% when the temperature increases from 1600 K to 1700 K.Our results agree with a previous study,15 and show that atomscan diffuse beyond the interface to the other side only when thetemperature reaches 0.6–0.8Tm (Tm is the melting temperatureof the material, and Tm is about 2000 K here). The main reasonis that the atomic thermal motions become intense, resultingin a significant diffusion when the temperature is close to themelting point of the material. Therefore, the thickness of thediffusion layer becomes larger with temperature. This is directevidence that the growth of the layer is diffusion controlledwhich is greatly influenced by the temperature. As a result, athicker diffusion layer and better interface combination are

more efficiently obtained when the composite temperatureis higher.

Besides the temperature effect, we have investigated thestrain rate effect on the atomic diffusion. Similarly, the thicknessof the diffusion layer at various strain rates can be obtained(see Fig. 6b) by analyzing the concentration of FeCrNi and Fe inthe Z direction at various strain rates (see Fig. S4 in the ESI†). Asshown in Fig. 6b, the thickness of the diffusion layer decreasesexponentially with increasing strain rate, which are 9 Å, 7.5 Å,6.5 Å and 6 Å at 1 � 109 s�1, 1.5 � 109 s�1, 3 � 109 s�1 and4 � 109 s�1, respectively. Obviously, the thickness of thediffusion layer decreases as the strain rate increases, whichdecreases by 33.3% when the strain rate increases from 1� 109 s�1

to 4 � 109 s�1. It can be explained that the compression time isinversely proportional to the strain rate on the premise of reachingthe same strain 0.3. Therefore, it is necessary to select theappropriate compressive strain rate to obtain a thicker diffu-sion layer for better cohesion of the bimetal composite. We canconclude that the optimized condition is a hot compressionprocess with a strain rate of 3 � 109 s�1 when taking bothcomposite quality and simulation time into consideration.

Atomistic structure of the bimetal interface

The interface is defined as the region with the largest thicknessof the diffusion layer. The atomic spatial distributions andstructures can be characterized by the radius distributionfunction (RDF). The influence of temperature and strain rateon the degree of interface amorphization can be obtained by

Fig. 6 Diffusion as a function of temperature (a) and strain rate (b). Redline stands for the thicknesses of the diffusion layer that can be obtained byextracting the coordinate information of the atoms from the LAMMPSdump files; blue line is for the exponential fitting of the thicknesses.

Fig. 7 Radius distribution functions on the interface at various tempera-tures (a) and strain rates (b).

PCCP Paper

Page 6: Nonlinear diffusion, bonding, and mechanics of the ...qpeng.org/publication/pdf/Peng_2019_pccp_steel_interface.pdf · Nonlinear diffusion, bonding, and mechanics of ... bi-crystals,14–21

This journal is© the Owner Societies 2019 Phys. Chem. Chem. Phys., 2019, 21, 1464--1470 | 1469

comparing the width and height of the peaks of the RDF curves.A general rule is that a broader and lower peak implies a higherdegree of amorphization. We have examined the RDF of theinterface as shown in Fig. 7.

All peaks of the RDF curves in Fig. 7a decrease withincreasing temperature. The trend is that a first narrower peakappears at 2.61 Å and splits the second peak, which shows awell crystalline structure in the interface region. Then the peaksgrow broader and lower, which indicates that an amorphousstructure has formed at the interface. The split of the secondpeak is also strong evidence of the amorphous structure ofthe interface,31 consistent with the previous results and thevisualization in Fig. 3 and 5. We have calculated the averagepositions of the four nearest peaks in the RDF of FeCrNi, Fe andthe interface region from Fig. 7a, as summarized in Table 2.The peaks of the interface region are between FeCrNi and Fe,and the first two peaks are closer to FeCrNi. These resultsindicate that the structure of the interface region undergoes aphase change from the crystalline state to an amorphous statewhen the strain increases.

Our further investigation of the strain rate effect on the RDFof the interface is shown in Fig. 7b, where the RDF is insensitiveto the strain rates in our strain rate range of 1 � 109 s�1 to4 � 109 s�1. The first peak of the RDF decreased with increasingstrain rate and the other peaks were broad and faded awayquickly. All these phenomena indicated that the structure ofthe interface region changed from the crystalline state to anamorphous state when the strain rate increased. The simulatedRDF was basically insensitive to the strain applied at the ratesused in this work. In order to clarify the reason, one moresimulation at 1500 K with a strain rate of 3.0 � 108 s�1 has beenexamined. Obviously, the first peak of the RDF at 3.0 � 108 s�1

was the highest. The simulated RDF was basically insensitive tothe strain applied at the rates used in this work because of theultra-high rate in applying strain.

To gain more insights into the influence of temperature andstrain rate on the interfacial structure, we present the atomicconcentration of the interface region in Fig. 8. The atomicconcentration of sFe, Cr, and Ni slightly increases with increasingtemperature, and the concentration of cFe slightly decreases. Thismainly results from the fact that the increase of temperaturepromotes atomic diffusion between FeCrNi and Fe, whichindicates that more and more FeCrNi diffuses into the Fe layer.From Fig. 8b, the concentration of Cr and cFe slightly increaseswith increasing strain rate and the concentration of Ni and sFeslightly decreases. It can be explained by the fact that there is notenough time for atoms to diffuse across the interface and thediffusion of the Fe layer is more dominant than that of FeCrNi at ahigher strain rate.

Conclusions

The diffusion and mechanics of the interface between austeniticsteel (FeCrNi) and bcc iron in a hot compression process havebeen investigated via molecular dynamics simulations. The effectsof compression temperature and strain rate on the compressionmechanism and interfacial diffusion were examined. A hightemperature favors the diffusion of interfacial atoms and improvesthe combination of FeCrNi/Fe to form a thicker diffusion layer.The thickness of the diffusion layer increases by 23.1% when thetemperature increases from 1500 K to 1600 K and by nearly 56.3%when the temperature increases from 1600 K to 1700 K. Thethickness of the diffusion layer is affected by the strain rate, witha decrease of 33.3% when the compressive strain rate increasesfrom 1 � 109 s�1 to 4 � 109 s�1. When the temperature and strainrate increase, the structure of the interface region changes fromthe crystalline state to an amorphous state.

The stress–strain curve at the initial stage of this process islinear until the stress reaches the maximum stress. After that, thestress decreases rapidly with fluctuations. During the compressionprocess of FeCrNi and Fe at high temperature, the compressivestrength of the FeCrNi/Fe system decreases by 44.3% and thecorresponding strain decreases by 7.2% when the temperatureincrease from 1500 K to 1800 K. However, the mechanical proper-ties are not sensitive to the strain rate in our study. Our study

Table 2 The average positions of the four nearest peaks in the RDF of316L, Q345R and the interface at various temperatures

First peak (Å) Second peak (Å) Third peak (Å) Fourth peak (Å)

FeCrNi 2.63 4.78 6.92 9.37Fe 2.62 4.83 6.92 9.35Interface 2.61 4.81 6.98 9.41

Fig. 8 Concentration variation at the interface (red) and concentration ofthe interface (blue) at various temperatures (a) and strain rates (b).

Paper PCCP

Page 7: Nonlinear diffusion, bonding, and mechanics of the ...qpeng.org/publication/pdf/Peng_2019_pccp_steel_interface.pdf · Nonlinear diffusion, bonding, and mechanics of ... bi-crystals,14–21

1470 | Phys. Chem. Chem. Phys., 2019, 21, 1464--1470 This journal is© the Owner Societies 2019

sheds light at the atomic level on the mechanical strength anddiffusivity of the interface in the bimetal during the hotcompression process which might be beneficial to the optimi-zation of the fabrication of bimetal composites.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The authors would like to thank the financial support to thisstudy from the National Natural Science Foundation China(No. 51375041, 21703007 and 51475039) and the FundamentalResearch Funds for the Central Universities (FRF-GF-17-B18and FRF-TP-16-044A1).

Notes and references

1 R. Jiangwei, L. Yajiang and F. Tao, Mater. Lett., 2002, 56,647–652.

2 H. Yang, M. Gu, W. Jiang and G. Zhang, Interface micro-structure and reaction in Gr/Al metal matrix composites, 1996.

3 M. Elices, G. V. Guinea, J. Gomez and J. Planas, Eng. Fract. Mech.,2002, 69, 137–163.

4 R. Komanduri, N. Chandrasekaran and L. M. Raff, Int. J. Mech.Sci., 2001, 43, 2237–2260.

5 P. M. Derlet and H. Van Swygenhoven, Scr. Mater., 2002, 47,719–724.

6 J. Schiøtz and K. W. Jacobsen, Science, 2003, 301, 1357–1359.7 B. Eduardo and M. Wang, Science, 2005, 5742.8 C. Brandl, P. Derlet and H. van Swygenhoven, Strain rates in

molecular dynamics simulations of nanocrystalline metals, 2009.9 D. E. Spearot, M. A. Tschopp, K. I. Jacob and D. L. McDowell,

Acta Mater., 2007, 55, 705–714.10 M. A. Tschopp and D. L. McDowell, J. Mech. Phys. Solids,

2008, 56, 1806–1830.11 P. Heino, H. Hakkinen and K. Kaski, Molecular-dynamics

study of mechanical properties of copper, 2007.12 Q. Peng, X. Zhang and G. Lu, Structure, mechanical and

thermodynamic stability of vacancy clusters in Cu, 2009.

13 Q. Peng, F. Meng, Y. Yang, C. Lu, H. Deng, L. Wang, S. Deand F. Gao, Nat. Commun., 2018, 9, 4880.

14 J. Jimenez-Saez, A. M. C. Perez-Martın, J. Domınguez-Vazquez and J. J. Jimenez-Rodrıguez, Ion beam mixingcalculation of multilayer samples, 1996.

15 S. Chen, F. Ke, M. Zhou and Y. Bai, Acta Mater., 2007, 55,3169–3175.

16 S.-G. Lee, S.-P. Kim, K.-R. Lee and Y.-C. Chung, J. Magn.Magn. Mater., 2005, 286, 394–398.

17 K. Gall, M. F. Horstemeyer, M. Van Schilfgaarde and M. I.Baskes, J. Mech. Phys. Solids, 2000, 48, 2183–2212.

18 M. Weissmann, R. Ramırez and M. Kiwi, Phys. Rev. B:Condens. Matter Mater. Phys., 1992, 46, 2577–2583.

19 S. D. Chen, A. K. Soh and F. J. Ke, Scr. Mater., 2005, 52,1135–1140.

20 S. Y. Chen, Z. Wu, K. Liu, X. Li, N. Luo and G. Lu, Atomicdiffusion behavior in Cu–Al explosive welding process, 2013.

21 H. Liu, F. J. Ke, H. Pan and M. Zhou, Molecular dynamicssimulation of the diffusion bonding and tensile behavior of aCu–Al interface, 2007.

22 G. Bonny, D. Terentyev, R. C. Pasianot, S. Ponce andA. Bakaev, Interatomic potential to study plasticity in stainlesssteels: The FeNiCr model alloy, 2011.

23 G. Bonny, N. Castin and D. Terentyev, Interatomic potentialfor studying ageing under irradiation in stainless steels: TheFeNiCr model alloy, 2013.

24 L. Verlet, Computer experiments on classical fluids. i Thermo-dinamical properties of Lennard-Jones molecules, 1967.

25 S. Plimpton, J. Comput. Phys., 1995, 117, 1–19.26 A. Stukowski, Visualization and Analysis of Atomistic Simulation

Data with OVITO - the Open Visualization Tool, 2010.27 A. Stukowski and K. Albe, Extracting Dislocations and non-

Dislocation Crystal Defects from Atomistic Simulation Data,2010.

28 R. Hoagland, T. Mitchell, J. Hirth and H. Kung, Philos. Mag.A, 2002, 82, 643–664.

29 I. N. Mastorakos, H. M. Zbib and D. F. Bahr, Appl. Phys. Lett.,2009, 94, 173114.

30 M. P. Allen and D. J. Tildesley, Computer simulation ofliquids, Oxford University Press, 2017.

31 X. J. Liu, Y. Xu, X. Hui, Z. P. Lu, F. Li, G. L. Chen, J. Lu andC. T. Liu, Phys. Rev. Lett., 2010, 105, 155501.

PCCP Paper


Recommended