Download Supplementary information of Accelerating Ferroic Ageing

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Chemical thermodynamics wikipedia , lookup

Transcript
Supplementary information of
Accelerating Ferroic Ageing Dynamics upon Cooling
Junyan Zhang1,2,6, Yunwei Mao1,2,7, Dong Wang1,2, Ju Li1,2,3,4* Yunzhi Wang1,2,5*
Frontier Institute of Science and Technology, Xi’an Jiaotong University, Xi’an 710049, China
State Key Laboratory for Mechanical Behavior of Materials, Xi’an Jiaotong University, Xi’an 710049,
China
1
2
3
Department of Nuclear Science and Engineering, Massachusetts Institute of Technology, Cambridge, MA
02139, USA
4
Department of Materials Science and Engineering, Massachusetts Institute of Technology, Cambridge,
MA 02139, USA
5
Department of Materials Science and Engineering, The Ohio State University, 2041 College Road,
Columbus, OH 43210, USA
Present address: 6 Department of Chemical Engineering, Northeastern University, Boston, MA 02115, USA
7
Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139,
USA
*Corresponding authors: [email protected]; [email protected]
S1. Phase field simulation details
A generic cubictetragonal martensitic transformation in 2D is considered. The two
crystallographically equivalent and hence energetically degenerate Bain paths generate
two deformation variants, which can be described by two non-conserved order parameter
fields, η1 (𝑟) , and η2 (𝑟) , respectively, in the Landau free energy, where r is the
coordinate vector. Any given microstructural state in the system can be described by local
values of these two order parameter fields, e.g., (η1 , η2 ) = (0,0) represents the austenite
and (η1 , η2 ) = (1,0), or (0,1) represents one of the two deformation variants of the
martensitic phase, respectively.
ε0𝑖𝑗 (𝑟⃑) = ε01
⃑)12 + ε02
⃑)22
𝑖𝑗 η(𝑟
𝑖𝑗 η(𝑟
where the stress free transformation strains of the two deformation variants are ε01
𝑖𝑗 =
1 0
1̅ 0). The total energy of the system consists of two parts: the
02
ε0 (
)
and
ε
=
ε
(
0
𝑖𝑗
1 1̅
1 1
chemical free energy Fchem and the elastic strain energy Eel . The former one can be
1
1
expressed in the form of:Fchem = ∫ d2 r [2 β(∇η1 )2 + 2 β(∇η2 )2 + f(η1 , η2 )], where the
first two terms in the square blanket are the non-local gradient energy terms with β being
the gradient energy coefficient and the last term is the “local” chemical free energy
described with 6-order Landau Polynomial(1):
1
1
1
f(η1 , η2 ) = A10 (η1 2 + η2 2 ) − A2 (η1 4 + η2 4 ) + A3 (η1 2 + η2 2 )2
2
4
4
1
+ A4 (η1 2 + η2 2 )3
6
(T
),
(A10 = A00
−
T
A
to
A4 are the expansion coefficients. In our simulation, A00
2
0
1
1 =
0.05,T0 = 220, A2=30, A3=19, A4=10). The second part of the total energy, the elastic
strain energy, is given by Khachaturyan (1, 2):
2
2
1
00
3 2
2
Eel = Cijkl ∑ ∑ ε00
ij (p)εij (q) ∫ d rηp (r, t)ηq (r, t)
2
2
p=1 q=1
2
1
d2 k
𝐤
∗
− ∑∑∫
Bpq ( ) × {η2p (𝐫, t)} {η2q (r, t)}
2
𝐤
𝐤
(2π)
2
k
p=1 q=1
If the system is doped with point defects, whose role can be generalized as the global
transition temperature effect (GTTE) and local field effect (LFE)(3), the total free energy
only needs to be revised in the “local” chemical free energy term as:
1
1
1
1
f(η1 , η2 ) = A1 (η1 2 + η2 2 ) − A2 (η1 4 + η2 4 ) + A3 (η1 2 + η2 2 )2 + A4 (η1 2 + η2 2 )3
2
4
4
6
1
1
2
−ηloc (1) [A10 η1 − (A2 − A3 )η1 3 + A3 η2 2 η1 + A4 η1 5 + A4 η2 4 η1 + A4 η2 2 η1 3 ]
3
5
3
1
1
2
3
2
5
4
−ηloc (2) [A10 η2 − (A2 − A3 )η2 + A3 η1 η2 + A4 η2 + A4 η1 η2 + A4 η1 2 η2 3 ]
3
5
3
where A1 = A10 + b ∙ c̅(b is defect strength and c̅ is average defect concentration of the
whole system) in the firstterm describes the LTTE, and ηloc(1) and ηloc(2) in the last two
terms describe the LFE, which are local random fields created by the point defects. One
possible physical origin of the LFE is lattice distortion caused by doped point defects,
resulting in symmetry breaking of the Landau free energy(4). The stochastic timedependent kinetic equation is used to describe the temporal evolution of the system
microstructure:
ηp (r, t + Δt) = ηp (r, t) + (dηp (r, t)/dt)Δt
,
where
dηp (r,t)
dt
=
−M (δF/δηp (r, t)) + ξp (r, t), p = 1,2. ξp (r, t) is thermal fluctuations, which is produced
by a random number generator. The kinetic equation is solved numerically in a
computational cell consisting of 256 × 256 grid sites with periodical boundary conditions
applied in both dimensions and macroscopic strain at the boundary is kept zero.
The length scale in our simulation can be estimated by the formula:
ΔF
ΔF∗
= (γ ×
d𝑥)/(γ∗ × d𝑥 ∗ ) , where ΔF is the free energy between austenite and martensite in
experiments, γis the interfacial energy for coherent face in experiments, and dx is the
character length scale in experiment. The physical quantities with star *, ΔF ∗, γ∗ and d𝑥 ∗ ,
are the corresponding quantities in simulation. Then l0=dx/dx* is the length unit in
simulation. Here we assume the interfacial energy for coherent face to be 0.05J/m2and the
chemical free energy ΔF for martensitic system is about −1.85 × 106J/m3. Therefore,
l0~0.9nm, meaning one grid length in simulation equals to about 0.9nm in experiment.
S2. Structural analysis of the ferroelastic systems in cooling process
Figure S1 analyzes the structural changes upon cooling for systems with different defect
concentrations from 0.02 to 0.40 (dimensionless), equivalent to 0.00222 to 0.04444
(at.%) in dimensional concentration unit. The dimensionless defect concentration, c, can
be converted to mole fraction, x, by x=c*a02/(l02) (1), l0 (0.9nm) is the length scale and a0
is the lattice parameter (~ 0.3nm for NiTi system). A step cooling process was used in our
simulations. We run the system under each temperature for enough time steps in order to
reach the equilibrium state (volume fraction change is less than 0.05%, approximately for
20,000 steps in our simulations); then decrease the temperature with ΔT (1K) and repeat
the procedure until the lowest temperature considered. The results are shown in Fig. S1.
Fig. S1 Structural analysis of ferroelastic systems under different defect
concentrations from 0.02 to 0.40. (a) depicts martensitic volume fraction change
along with temperature. The defect concentration of each line is indicated by the
color bar. (b) is the derivative of martensitic volume fraction with respect to
temperature when c=0.02, 0.14 and 0.40. Part of (b) is enlarged in the inserted
figure. (c) is the domain number change along with temperature. (d) shows the
maximum domain number (MaxN, the blue curve), final domain number (FinaN,
the green curve), and the critical temperature corresponding to MaxN (Tm, the red
curve) with respect to defect concentration. (a)-(c) share the same color bar.
S3. DSC results of the systems
Furthermore, we calculated the DSC curves of three typical systems (c=0.02, 0.14
and 0.40) in the cooling process. The results are shown in Fig. S2. We can see that
for typical martensitic transition, there is a sharp peak indicating the first order
martensitic transition. For intermediate concentraion, there are two peaks, one for
austenite to amorphous transition at TP-A and the other for amorphous to
martensitic transition at TA-M, whereas for a typical amorphization process, there
is only one diffuse peak at the transitional temperature.
Fig. S2 DSC curves of ferroelastic systems with three typical defect
concentrations 0.02, 0.14 and 0.40
1.
2.
3.
4.
Semenovskaya S & A.G. K (1997) Coherent structural transformations in random
crystalline systems. Acta Materialia 45(10):18.
Khachaturyan AG (1983) Theory of structural transformations in solids.
Levanyuk AP & Sigov AS (1988) Defects and structural phase transitions.
Wang D, Wang Y, Zhang Z, & Ren X (2010) Modeling Abnormal Strain States in
Ferroelastic Systems: The Role of Point Defects. Physical Review Letters 105:4.