You are on page 1of 45

Journal Pre-proof

Interface characteristics and precipitation during the


austenite-to-ferrite transformation of a Ti-Mo
microalloyed steel

Ilias Bikmukhametov, Hossein Beladi, Jiangting


Wang, Vahid Tari, Anthony D. Rollett, Peter D.
Hodgson, Ilana Timokhina

PII: S0925-8388(21)03634-3
DOI: https://doi.org/10.1016/j.jallcom.2021.162224
Reference: JALCOM162224

To appear in: Journal of Alloys and Compounds


Received date: 1 May 2021
Revised date: 24 September 2021
Accepted date: 30 September 2021
Please cite this article as: Ilias Bikmukhametov, Hossein Beladi, Jiangting Wang,
Vahid Tari, Anthony D. Rollett, Peter D. Hodgson and Ilana Timokhina,
Interface characteristics and precipitation during the austenite-to-ferrite
transformation of a Ti-Mo microalloyed steel, Journal of Alloys and
Compounds, (2021) doi:https://doi.org/10.1016/j.jallcom.2021.162224
This is a PDF file of an article that has undergone enhancements after acceptance,
such as the addition of a cover page and metadata, and formatting for readability,
but it is not yet the definitive version of record. This version will undergo
additional copyediting, typesetting and review before it is published in its final
form, but we are providing this version to give early visibility of the article.
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.
© 2021 Published by Elsevier.
Interface characteristics and precipitation during the austenite-to-ferrite transformation of a Ti-Mo

microalloyed steel

Ilias Bikmukhametov a, b, Hossein Beladi b, Jiangting Wang b, Vahid Tari c, Anthony D. Rollett c
Peter D. Hodgson b, Ilana Timokhina b
a
Department of Metallurgical & Materials Engineering, The University of Alabama, Tuscaloosa,
AL, United States of America
b
Institute for Frontier Materials, Deakin University, Geelong, Australia
c
Department of Materials Science and Engineering, Carnegie Mellon University, Pittsburgh, PA,
United States of America
Email: ibikmukhametov@ua.edu

f
hossein.beladi@deakin.edu.au

oo
jiangting.wang@deakin.edu.au
vahidtari@gmail.com
rollett@andrew.cmu.edu

pr
peter.hodgson@deakin.edu.au
ilana.timokhina@deakin.edu.au
Abstract e-
The complexity of interphase precipitation in a Fe-0.19C-1.54Mn-0.4Si-0.06Al-0.13Mo-0.06Ti
Pr
(at. %) high strength low-alloy (HSLA) steel at an early stage of the austenite-to-ferrite

transformation was studied by analysing the solute distribution across ferrite-austenite interfaces
al

with Kurdjumov-Sachs (K-S) and non-K-S orientation relationships (OR). Structural


n

characterisation i.e. ferrite/austenite OR and habit plane characteristics was performed by


ur

electron backscatter diffraction (EBSD) and the clustering back-calculation approach, while
Jo

solute distributions i.e. the solute concentration spikes in the interface regions were studied by

atom probe tomography (APT) on the specimens specifically prepared across/near the

ferrite/austenite interfaces.

It was shown for the first time that interphase precipitation is promoted at both types of interface:

(i) a K-S OR and habit plane deviated from ideal (110)α//(111)γ and (ii) a non K-S OR. The key

aspect of interphase precipitation is the distribution of solute atoms across the interface, which is

pronounced Mn, Ti, Mo and C concentration spikes at the interphase boundary. In contrast,

1
interphase precipitates were not formed at the coherent interface with a K-S OR and habit plane

of (110)α//(111). This was correlated with the interfacial condition, where the compositional

ratio of substitutional solute and solvent elements remains almost constant across the interface,

i.e. Mn and C spikes. Interface compositions in this study did not match with local equilibria

(negnigible partitioning local equilibrium and paraequilibrium) limits. In addition, it appeared

that the interfaces with Mn, Ti, Mo and C concentration spikes form ledges leading to randomly

f
redistributed interphase precipitates.

oo
Graphical abstract

pr
e-
Pr
n al
ur

Keywords: Ti-Mo microalloyed steel, Austenite-to-ferrite transformation, Interphase


Jo

precipitation, Kurdjumov-Sachs orientation relationship, Atom probe tomography, Para-

equilibrium, Negligible partitioning local equilibrium, Interface segregation.

1. Introduction

The ferrite/austenite interface characteristics occurring during the austenite-to-ferrite

transformation have a major effect on the resulting micro- and nanostructure of steels, which

subsequently governs their mechanical properties [1]. One of the most prominent examples of

this relationship is the interphase precipitation phenomenon that occurs during the movement of

2
the ferrite/austenite interface in the presence of strong carbide forming elements such as Nb, V,

Ti, Cr, etc. [2], [3], [4], [5], [6], [7], [8], [9], [10], [11], [12], [13], [14], [15]. The resulting

microstructure consists of ferrite grains with parallel rows of finely dispersed precipitates and

clusters [10] presumably formed as a result of the successive interfacial ledge migration [6]. The

clusters have been defined as aggregates of solute atoms that are aligned with the composition

and crystallography governed by the ferrite matrix [11]. This microstructure is recognized to

f
have the optimum strength while maintaining excellent formability [16], [17], [18], [19]. A

oo
significant strengthening effect of 300 MPa from the interphase precipitation was achieved in a

pr
Ti-Mo steel with only ~ 0.09 wt.% Ti [20], where Ti is the main carbide forming element and

Mo appears to increase the coherency of precipitates and, therefore, their aging resistance [21],
e-
[22].
Pr
As previously discussed [6], [23], [24], [25], [26], [27], [28], the main austenite-to-ferrite

transformation parameters that affect interphase precipitation are: (i) the orientation relationship
al

(OR) between ferrite and parent austenite grains [6], [25]; (ii) phase equilibria at the
n

ferrite/austenite interface [26], [27]; (iii) interfacial mobility [23], [28]; and (iv) the solute drag
ur

effect [29]. All of these parameters depend on the steel composition and thermomechanical
Jo

processing [4], [12], [23], [27], [30].

Ferrite normally nucleates in austenite with certain ORs, namely: Bain, Kurdjumov-Sachs (K-S),

Greninger-Troiano (G-T), Pitsch, or Nishiyama-Wassermann (N-W) ORs [31]. However, a

previous study suggested the K-S ((111)γ//(011)α, [1̅01]γ//[1̅1̅1]α) and N-W ((111)γ//(011)α,

[112̅]γ//[01̅1]α) ORs between parent austenite and freshly formed ferrite are dominant [32].

Moreover, ferrite can nucleate in parent austenite with a rational OR (K-S, N-W, etc.) with one

of the austenite grains, but grow into neighbouring austenite grains with an irrational OR [33].

3
Davenport et al. [6] proposed that interphase precipitation can only occur at the K-S oriented

((110)α//(111)) interfaces with low mobility, i.e., semi-coherent interfaces. In contrast, a study of

interphase precipitation in a V-alloyed HSLA steel using combined transmission electron

microscopy (TEM) and the back-calculation approach [34] confirmed the formation of

interphase precipitates in non-K-S (irrational OR) oriented grains, presumably with incoherent

interfaces [25]. The formation of V-rich interphase precipitates, in this case, was attributed to the

f
presence of a V concentration spike at the front of the interfaces between non-K-S related

oo
ferrite/austenite grains, which increased the total driving force for interphase precipitation [26],

pr
[27].

Nucleation and growth of interphase precipitates at the interface require diffusion and depend on
e-
the diffusional flux of microalloying elements across the interface. The diffusional flux of
Pr
microalloying elements is a function of the solute distribution across the ferrite/austenite

interface and could be explained by the mode of phase equilibria at the migrating interface, such
al

as negligible partitioning local equilibrium (NPLE) and para-equilibrium (PE) [35]. In the NPLE
n

model, even though long-range diffusion is not possible, the substitutional atoms can still move
ur

in the vicinity of the interface and build up a solute spike in front of the interface as the interface
Jo

passes through [1], [35]. In contrast, the PE model represents a limiting case where the

substitutional atoms are completely immobile (but not the interstitials), and the compositional

ratio of substitutional solute and solvent elements remains constant across the interface [1], [35],

[36]. However, a transition from the PE to the NPLE condition could occur due to a change in

velocity of the interface during transformation, which is high at the beginning of transformation,

supporting the PE condition, and slows down as transformation proceeds, leading to the NPLE

condition [1].

4
Furthermore, the solute atoms can also segregate directly to the transforming interfaces lowering

their mobility, according to the solute drag and coupled solute drag models [35], [37], [38]. The

solute drag model [37], [38] is based on the dissipation of the Gibbs free energy by substitutional

solute atoms that move by short-range diffusion across the interface and the possible interaction

between substitutional atoms and the migrating interface [37]. In addition, solute species can

mutually alter the activity in the solvent phase and, therefore, change the segregation behaviour

f
at interfaces - this effect is known as the coupled solute-drag effect [38]. The solute drag effect

oo
has been experimentally confirmed by strong segregation of Mn to interfaces with a K-S OR,

pr
which means that the interfaces between grains with the K-S OR also potentially could be

suitable for interphase precipitation [39]. e-


Therefore, the scope of the current study is to address the complex mechanism of interphase
Pr
precipitation formation including structural (i.e., which OR between the parent austenite and

freshly formed ferrite occurs along with its habit plane) and compositional (i.e., which austenite-
al

to-ferrite equilibrium condition is in effect at the interface) effects. Multi-modal characterization


n

comprising electron backscatter diffraction (EBSD) and atom probe tomography (APT) was used
ur

to analyse solute distributions across the ferrite-austenite interfaces and in ferrite grains with K-S
Jo

and non K-S ORs between austenite and ferrite.

2. Experimental

The composition of the laboratory steel was Fe-0.19C-1.54Mn-0.4Si-0.06Al-0.13Mo-0.06Ti

(at.%), equivalent to Fe-0.04C-1.52Mn-0.21Si-0.03Al-0.22Mo-0.05Ti (wt. %). The as-received

slab had an initial thickness of 40 mm and was hot rolled over a temperature range of 1200 -

1100 oC through multi-pass deformation down to a final thickness of 13 mm. The material was

then homogenized at 1200 °C for 4 h under an argon atmosphere. A cylindrical sample 15 mm

5
long and 10 mm in diameter was machined, having the longitudinal axis along the rolling

direction. The sample was then subjected to thermo-mechanical processing (TMP) using a

500 kN thermo-mechanical treatment simulator (Servotest TMTS 500K), which has been

described elsewhere [40]. TMP was designed based on the approach proposed in [26], [27], [39].

This involves an interruption of the austenite-to-ferrite transformation at an early stage, leading

to a ferrite and martensite microstructure, where microstructures with “frozen” interfaces are

f
then further studied by a cross-correlative approach. The cylindrical sample was reheated to

oo
1200 oC for 180 s to dissolve pre-existing precipitates. The samples were then cooled to 650 oC

pr
at 30 K/s and held isothermally for 900 s, followed by water quenching (Fig. 1). The temperature

of 650 oC was chosen based on previous work [5] as the optimum temperature for extensive
e-
precipitate formation during the austenite-to-ferrite transformation. The isothermal section of the
Pr
Mn-C phase diagram as well as the para-equilibrium (PE) limits were calculated using Thermo-

Calc software from the TCFE3 database.


n al
ur
Jo

Fig. 1. Schematic representation of thermomechanical processing schedule.


The specimens for EBSD were initially cut along the axial direction and mechanically polished

with a sequence of 9, 3, 1, and 0.04 µm suspensions. EBSD was conducted using an FEI Quanta

6
3D FEG-SEM operated at 20 kV and 8 nA. The EBSD maps were acquired using a step size of

0.25 µm. The ferrite grains for analysis were selected preferentially having a 100 direction

normal to the sample surface, to have 100 pole clearly seen on APT desorption maps for

reconstruction purposes. The EBSD datasets were post-processed using TSL-OIM Analysis v6.

It is well known that there is no exact orientation relationship between parent austenite and

daughter product/s (i.e., ferrite) upon phase transformation, varying between the K-S and N-W

f
orientation relationships. Owing to the low carbon content in the current steel (i.e., 0.04 wt.% C),

oo
the orientation relationship is expected to be close to the K-S OR [41]. However, the close-

pr
packed habit plane between ferrite and parent austenite may vary from the ideal K-S OR (i.e.,

(110)α//(111)) during the ferrite growth, leading to an irrational interface. Therefore, it is


e-
necessary to examine both the lattice misorientation between parent austenite and ferrite, and
Pr
their habit planes to determine whether the examined ferrite grain matches the K-S OR criteria

with austenite. In the present microstructure, there is no retained austenite to directly determine
al

the closest OR of ferrite with the parent austenite upon transformation. Therefore, we introduced
n

a mathematical approach to indirectly estimate the orientation of the parent austenite. Here, the
ur

transformed region representing a prior austenite grain within the microstructure was initially
Jo

detected through a misorientation angle range of 21-47˚, as it does not meet any misorientations

that result from the ideal K-S OR [41]. The prior austenite grain was verified by plotting its 001

pole figure and comparing with one expected from the K-S orientation relationship. It should be

noted that not all poles can be seen in the 001 pole figure, as the data are related to the two-

dimensional microstructure cross-section and some poles are missing. Afterwards, the

orientations of martensite laths within the defined prior austenite region were used to calculate

the prior austenite orientation, by using the K-S orientation relationship and a back-calculation

7
approach described in [42]. The calculated parent austenite orientation was then used to

theoretically calculate all possible BCC orientations formed through the transformation of a prior

austenite grain-to-ferrite/martensite. This yields 24 orientations in the 001 pole figure, based on

the assumed K-S OR. Afterwards, the orientation of ferrite within the region of interest is

compared with the ones generated through the theoretical calculation, where the lowest value is

considered as the deviation from the K-S orientation relationship for the current ferrite grain. In

f
the current work, any ferrite grain with a deviation of > 5° from the K-S OR was considered

oo
irrational and those with <5° deviation were further analysed using trace analysis approach [43]

pr
to determine the habit planes between ferrite and parent austenite.

The trace analysis is an approximation approach, as resolving the exact habit planes needs
e-
special techniques such as TEM [44] and three-dimensional examination [45]. Here, oriented
Pr
stereographic projections were prepared by overlaying the pole figures of ferrite and adjacent

parent austenite, using their Euler angles. The matching habit planes for a given interphase
al

boundary were determined by choosing the plane poles in the projection with the nearest to
n

coincidence, which were also placed on the locus of possible interface plane normal (i.e., on the
ur

great circle normal to the trace of interface). The stereograms contain only six low-index planes,
Jo

namely {001}, {011}, {111}, {012}, {112}, and {122}. The interface tilt angles were measured

from APT maps, which are then plotted on the pole trace to determine whether the interface

matches the K-S condition or not.

APT specimens from selected interfaces were prepared by the focused ion beam (FIB) in-situ

lift-out procedure using a FEI Quanta 3D FEG-SEM [46]. The APT study was performed on a

CAMECA LEAP-4000HR in the voltage-pulsed mode with 20 % for pulse fraction and 200 kHz

pulse rate at 60 K. IVAS 3.6.12 software was used to reconstruct the APT data. Atom probe

8
datasets were reconstructed using voltage curve by iterating image compression and kf factors

untill plane spacings of 3 crystallographic planes matched their theoretical values [40]. The

concentrations of elements at the ferrite/austenite interfaces were evaluated by plotting one-

dimensional concentration profiles along region of interest (ROI) cylinders with diameters

ranging from 20 to 30 nm. The maximum separation method [47] was used to quantify and

separate precipitates from the matrix. Different minimum numbers of solute atoms in the clusters

f
(Nmin) and threshold distance between solute atoms (dmax) were used for each dataset, depending

oo
on statistics and distribution of solute atoms. The core solute atoms for the cluster analysis were

pr
Ti, Mo and C. Details of the cluster analysis characterisation procedure have been described in

detail elsewhere [48]. According to Dhara [48] C2+ and Ti2+ mass spectrum peak overlap at 24
e-
Da was assigned as Ti. The Fe atoms were considered as part of the matrix, and were not
Pr
included in the composition analysis of precipitates. The size of precipitate or cluster is

represented here as a Guinier radius (rG), which is calculated as


al

𝑟𝐺 =𝑙𝑔 √5⁄3, (1)


n

where lg is the radius of gyration [22]. The radius of gyration lg,which is calculated using the
ur

following equation:
Jo

𝑁𝐶 (𝑟 2
∑𝑖=1 𝐶𝑂𝑀 −𝑟𝑖 )
𝑙𝑔 = √ 𝑁 𝐶 𝑚
, (2)
∑𝑖=1 𝑖

where rcom is the coordinates defining the centre-of-mass of each cluster/precipitate, mi and ri are

the mass and coordinates of each atom in the cluster/precipitate and Nc is the cluster/precipitate

size [49]. The number density of precipitates was calculated by dividing the total number of

precipitates by the dataset volume.

9
3. Results

The microstructure consisted of 5 % of polygonal ferrite and 95 % martensite (Fig. 2). The ferrite

grains appeared to largely nucleate on prior austenite grain boundaries and grow into the adjacent

austenite grain/s. The average prior austenite grain size was 72 ± 10 µm.

f
oo
pr
e-
Pr
al

Fig. 2. Image quality (IQ) and Inverse pole figure (IPF) EBSD maps showing microstructure of
steel after interrupted isothermal treatment. The yellow lines represent 21-47° boundaries; The
n

triangle inset represents the corresponding colour key.


Fig. 3a shows a martensite (α’(γ1)) and a ferrite grain (α1), which appeared to nucleate on the
ur

prior austenite grain boundary and grew into the prior austenite grain γ1. The 001 experimental
Jo

pole figure of the prior austenite region (outlined by white dotted line) clearly revealed that the

orientation of ferrite does not closely match with the distinctive variant clusters expected from

the K-S OR (Fig. 3b). As a result, the ferrite orientation revealed a minimum deviation angle of

17° from the phase transformation products theoretically calculated from the prior austenite

orientation using the K-S OR (Fig. 3c). This suggests that the ferrite had an irrational OR with

the parent austenite.

10
f
oo
pr
e-
Pr
Fig. 3. (a) Inverse pole figure map superimposed by image quality map of ferrite grain α1 and
martensite α’(γ1), which are both products of γ1 prior austenite. The area outlined by dotted line
al

corresponds to the γ1 prior austenite area. The area outlined by dashed line corresponds to the
area of the interface for APT analysis (specimen 1); (b) Experimental 001 pole figure of α1 and
α’(γ1); (c) Calculated (K-S) 001 pole figure of α1 and α’(γ1). Blue triangles correspond to
n

calculated prior austenite orientations. Green squares represent ferrite orientations. Red circles
ur

represent K-S calculated martensite orientations.


The specimen for APT analysis was then extracted from the area, which comprised ferrite
Jo

(α1)/martensite (α’(γ1)) interface, using the in-situ lift-out approach (Fig. 3a). Based on EBSD

study, the α1/γ1 interface was considered as the transforming interface. This procedure allows us

to concurrently study both the solute atoms at the interfaces and interphase precipitation in

ferrite.

The APT atom map superimposed on the TiC iso-concentration from α1/α(γ1) interface (specimen

1 in Fig. 3a) is presented in Fig. 4a. The position of the interface is highlighted by the high

concentration of C, Mn, Ti and Mo atoms in Fig. 4a. The 1D concentration profile across the

11
α1/α’(γ1) interface (Fig. 4c), within the cylinder shown in Fig. 4a, confirmed the presence of Mn,

C and Ti solute spikes at the interface, having maximum values of 1.5 ± 0.1 at.% C, 3.0 ± 0.1

at.% Mn, and 1.0 ± 0.1 at.% Ti, accordingly.

In addition, there was a step feature revealed within the cross section of the 2 at. % Mn iso-

surface with a height of 5 nm (Fig. 4d) dividing the interface into two “terraces”. The 2D Mn

density map (Fig. 4b) constructed from the top terrace of the interface also revealed a profile

f
with a 5 nm kink within the interface plane. The iso-concentration surfaces applied to the APT

oo
reconstruction showed the formation of interphase precipitates/clusters in the ferrite near the

pr
interface (Fig. 4a). The precipitation characteristics were analysed using the standard cluster

analysis procedure, and the average composition of precipitates/clusters was found to be


e-
Ti0.42Mo0.23C0.35 with an average Guinier radius of 1.7 ± 0.5 nm and a number density of 11 x
Pr
1022 m-3.

Fig. 5a shows a polygonal ferrite grain (α2) nucleated on a prior austenite triple junction, largely
al

growing into two adjacent austenite grains (i.e., 2 and 3 in Fig. 5a). The experimental 001 pole
n

figures revealed that the ferrite grain α2 had a near K-S OR with the parent austenite γ2 (rather
ur

than γ3), as it largely matched with the variant clusters expected from the K-S OR (Figs. 5c, d).
Jo

This was also confirmed with the theoretical calculation, as the ferrite orientation had a

minimum deviation angle of 3.7° and 15° from the 2 and 3 parent austenite, respectively,

considering the K-S OR (Figs. 5d, g). This suggests that the ferrite grain α2 had a rational

relationship with the 2 parent austenite and was irrational with the 3 prior austenite; the latter

case also could be deduced from the 21-47° boundary dividing α2 and α’( 3) (Fig. 5a, b).

12
f
oo
pr
e-
Pr
n al
ur
Jo

Fig. 4. APT analysis of the α1/α’(γ1) interface (specimen 1 in Fig. 3a): (a) Ti, Mn, C, Mo atom
maps superimposed with 3 at. % TiC iso-surface; (b) 2D Mn density map from the top terrace of
ledge indicated as dashed ellipse in “a”; (c) 1D concentration profile along arrow direction of
ROI cylinder “I” in “a”. The filled area around the curves represent a standard deviation. Dashed
line indicates total Mn concentration in the sample. (d) Mn iso-surface from cube ROI from
ledge region, indicating ledge profile.

13
f
oo
pr
e-
Pr
n al
ur
Jo

Fig. 5. Microstructural analysis of the α2 ferrite grain: (a) Inverse pole figure map superimposed
on the image quality map of ferrite grain α2; (b) Cross-section of FIB-milled trench during APT
specimen preparation from ferrite grain α2. Numbers indicate site-specific positions of APT
specimens. White dotted indicates prior austenite grain boundary. White dashed line (T 1)
represents trace for α2/α’(γ2) trace analysis. Rainbow triangle represents the colour key; (c)
Experimental 001 pole figure of α2 and α’(γ2); (d) Simulated (K-S) 001 pole figure of α2 and
α’(γ2); Blue squares correspond to calculated prior austenite orientations. Green triangles
represent ferrite orientations. Red dots represent martensite orientations; (e) Trace analysis of the
α2/γ2 interface, where α2 poles are green and γ2 poles are blue; (f) Experimental 001 pole figure
of α2 and α’(γ3); (g) Simulated (K-S) 001 pole figure of α2 and α’(γ3).

14
To further analyse whether the interface between α2 and 2 grains matches the K-S habit plane, a

trace analysis was performed, which revealed that the habit plane of ferrite and austenite is close

to (110)α2//(111) (Fig. 5e). This suggests that α2/ 2 interface could possibly meet the K-S

habit plane criteria (i.e., (110)α//(111)), where deviation angle from the K-S OR was 3.7°.

The APT samples were prepared from the area across the ferrite-austenite interface and the α2

grain interior. A cross-section of FIB-milled lift-out trench across the ferrite grain α2 is shown in

f
Fig. 5b, where the exact locations of the APT specimens are numbered as 2, 3, 4, 5 (Fig. 5b). The

oo
site-specific APT specimens number 2 and 3 originated from the part of the α2 grain with a

pr
rational OR with the adjacent prior austenite grain 2 (Fig. 5a, b), while specimens 4 and 5 were

from the part of the α2 ferrite grain with an irrational OR with the adjacent prior austenite grain
e-
3 (Fig. 5a, b).
Pr
The three-dimensional Ti, Mo, and C atom maps of specimens 2 and 3 (Fig. 6a, b), which

belonged to the rational OR side of the ferrite grain (Fig. 5a, b), did not show any obvious
al

precipitation/clustering, whereas only a single cluster/precipitate was observed in specimen 3


n

(Fig. 6b). The segregation of C along the planar feature in specimen 2 (Fig. 6a) indicates the
ur

position of the α2/α’(γ2) interface on the rational OR side of the ferrite grain. Slight segregation
Jo

of Mn and Ti along the interface in sample 2 was also observed (Fig. 6a). The 1D concentration

profile plotted along the “z” direction of the ROI cylinder across the interface confirmed a

concentration spike of 0.5 ± 0.1 at.% of C and 1.75 ± 0.1 at.% of Mn near the interface (Fig. 6c)

with no apparent Ti peak. As noted before, the C segregation located the α2/α’(γ2) interface and,

from that, the interface tilt angle (deviation of the plane normal from z-axis) could be measured,

which was 63° (Mn map in Fig. 6a). Afterwards, this angle was plotted on the α2/α’(γ2) pole trace

(open triangle), which was

15
f
oo
pr
e-
Pr
n al
ur
Jo

Fig. 6. Site-specific APT analysis of the α2 ferrite grain: (a) Ti, Mo, C and Mn atom maps of
specimen “2” in “Fig. 5a”. The angle on Mn map indicates the deviation of interface normal
from the “z” direction; (b) Ti, Mo, C and Fe atom maps superimposed with 3 at. % TiMoC iso-
surface of specimen “3” in “Fig. 5a”; (c) 1D concentration profile along ROI cylinder “I” in “a”.
The filled area around the curves represent a standard deviation. Dashed line indicates total Mn
concentration in the sample; (d) Fe atom map superimposed with 3 at. % TiMoC iso-surface of
samples “4” and “5” in “Fig.5a”.

16
close to the (110)α//(111) pole overlap (Fig. 5e). The latter suggests that the interface in

specimen 2 is (110)α//(111) or close to this.

Three-dimensional Fe atom maps superimposed with the 3 at. % TiMoC iso-concentration

surface of specimens 4 and 5, which belonged to the irrational OR side of ferrite grain (Fig. 5a,

b), showed the presence of randomly distributed TiMoC precipitates/clusters (4 and 5 in Fig. 6d).

The precipitates/clusters analysis showed a difference between the precipitate characteristics in

f
specimens 4 and 5, i.e., the Guinier radius of precipitates/clusters and number density in

oo
specimen 4 were 1.3 ± 0.4 nm and 36 × 1022 m-3, respectively, whereas the average number of

pr
atoms and number density of precipitates/clusters in specimen 5 were 1.6 ± 0.4 nm and 70 × 1022

m-3, respectively (Fig. 6d). Therefore, the precipitates/clusters, which were closer to the prior
e-
austenite grain boundary (specimen 4) had a roughly factor of two lower number density and
Pr
larger size compared to the precipitates/clusters closer to the interface (specimen 5). However,

the average compositions of precipitates/clusters in both specimens were the same,


al

approximately Ti0.44Mo0.26C0.31.
n

In Figure 7a, two polygonal ferrite grains (i.e. 3 and 4) are present within a given prior
ur

austenite grain, each having a distinct orientation. The experimental 001 pole figure did not
Jo

reveal all possible variant clusters expected from the K-S OR, though the orientations of both

ferrite grains appeared to be close to these clusters (Fig. 7b). The theoretical calculation also

showed that the minimum deviation angle from the K-S OR was 4° and 3° for the α3 and α4,

respectively (Fig. 7c).

Site-specific APT specimens were prepared from the α3 ferrite grain interior (specimen 6) and

α4/α’(γ4) interface (specimens 7 and 8), as shown in Fig. 7a. The trace analysis of interface T2

(from which specimens 7 and 8 were extracted) revealed that the habit plane of this α 4/γ4

17
interface was close to (110)α//(111)γ (Fig. 7d). This suggested that both specimens 7 and 8

potentially have a boundary close to the K-S habit plane (Figs. 7a, d). The three-dimensional Fe

atom map superimposed with a 3 at. % TiMoC iso-surface of specimen 6 (Fig. 7e) shows the

presence of

f
oo
pr
e-
Pr
n al
ur

Fig. 7. Microstructural analysis of the α3 and α4 ferrite grains: (a) Inverse pole figure map
Jo

superimposed with image quality map of ferrite grains α3 and α4. Yellow dotted lines indicate
prior austenite grain boundaries. The triangle inset indicates corresponding colour key. White
dashed line outlines trace of α4/γ4 interface boundary, where 7 and 8 are positions of APT
specimens; (b) Experimental 001 pole figure of the region enclosed by yellow dotted lines in “a”;
(c) Calculated 001 pole figure of the K-S variants originated from γ4 prior austenite grain along
with α3 and α4 ferrite grain orientations; (d) Trace analysis of the α4/γ4 interface, where α4 poles
are green and γ4 poles are blue; (e) Fe atom map superimposed on 3 at. % TiMoC iso-surface of
the specimen “6” in “a”. Black dashed lines indicate interphase precipitation/clustering planes.
interphase precipitates/clusters in the α3 ferrite grain, in the form of parallel sheets with a

separation of 12±2 nm. The three-dimensional Ti, Mo, C and Mn atom map superimposed with 3

at. % TiMoC iso-surface also showed the presence of interphase precipitates/clusters near the

18
α4/α’(γ4) interface in specimen 7 (Fig. 8a). The segregation of C and Mn along the interface

enables us to define the α4/α’(γ4) interface position (Fig. 8c). The 1D concentration profile along

the ROI cylinder “I” showed 1 ± 0.1 at. % C, 3 ± 0.1 at. % Mn, and 0.75 ± 0.05 at. % Ti

concentration

f
oo
pr
e-
Pr
n al
ur
Jo

Fig. 8. APT analysis of the α4/α’(γ4) interface: (a) Ti, Mn, C, Mn atom map superimposed with 3
at. % TiC iso-surface of sample “7” in “Fig. 7a” and corresponding Fe 2D density map along “z”
direction; (b) Ti, Mn, C, Mn atom map superimposed with 3 at. % TiC iso-surface of sample “8”
in “Fig. 7a” and corresponding Fe 2D density map along “z” direction; (c) 1D concentration
profile along arrow direction of ROI cylinder “I” in “a”; (d) 1D concentration profile along
19
arrow direction of ROI cylinder “II” in “b”; The filled area around the curves represent a
standard deviation. Dashed line indicates total Mn concentration in the sample.
spikes (Fig. 8c). Interestingly, APT specimen 8, taken from the same α4/α’(γ4) interface as

specimen 7, did not show any interphase precipitation/clustering (Fig. 8b). In addition, the

concentration spike of Mn near the interface was about 2 at. %, which is lower than in specimen

7. In contrast, the concentration spike of C in specimen 8 was 1.75 at. %, which is higher than in

specimen 7 (1 at. %). 2D Fe density maps, which are often used to reveal the crystallography of

f
APT specimens, confirmed that the upper parts of both specimens belong to the α4 ferrite grain

oo
(Fig. 8a, b). As noted before, the interfaces in specimens 7 and 8 could both meet the K-S habit

pr
plane criteria; however, they have different tilt angles of 64° (Fig. 8a) and 49° (Fig. 8b)

suggesting that only one interface can be (110)α4//(111)γ4. To find which part of the interface is
e-
closer to the (110)α4//(111)γ4 criteria, the interface tilt angles were plotted on the α4/α’(γ4)
Pr
interface pole trace. Figure 7d (open rhombus and triangle are specimen 7 and specimen 8,

respectively, in Fig. 7d). Figure 7d shows that the interface normal (black open triangle) in
al

specimen 8 closely matches the K-S criteria, i.e., (110)α4//(111)γ4 based on the closeness of the
n

110 (blue) and 111 (green) poles, which is not the case for specimen 7 (black open diamond).
ur

4. Discussion
Jo

4.1 The effect of OR on the interphase precipitation/clustering

The effect of the ferrite/austenite OR on the formation of precipitates/clusters was studied for

both type of grains with rational and irrational OR. Grains α1 (Fig. 3, Table 1) and α2 (specimens

4, 5 in Figs. 5a and 6d, Table 1) had an irrational OR with the prior austenite grains. APT

confirmed the formation of interphase precipitates/clusters in both ferrite grains (Figs. 4 and 6d).

Moreover, APT specimens extracted across α1/γ1 interface, which was considered as the

transforming interface, showed C, Mn, Ti and Mo concentration pileups at the interface (Fig. 4,

20
Table 1). It has been reported that interphase precipitates are formed at interfaces with low

mobility, such as a K-S (rational) coherent interface and, in contrast, a non K-S (irrational)

interface is expected to be incoherent with high mobility [6], [28]. However, a non K-S interface

could accommodate solutes leading to the local partitioning and segregation of substitutional

elements at the interface. In this case, the interface mobility could decrease due to the retardation

from the solute partitioning and interphase precipitation/clustering could occur.

f
Table 1. Summary of interface characteristics observed in the current work.

oo
Number OR Interface plane Precipitates Condition C, Mn, Ti, Mo,
of Ferrite from at interface at. % at.% at. % at. %
grain and trace analysis

pr
specimen
α1 (#1) Non irrational Yes Pronounced 1.5 ± 0.1 3 ± 0.1 1 ± 0.1 0.6 ±
K-S e-spike of 0.1
solutes
α2 (#2, 3) K-S {110}α//{111}γ No Negligible 0.6 ± 0.1 1.8 ± 0.1 0.06 0.13
spike of
Pr
sokutes
α2 (#4, 5) Non N/A Yes N/A N/A N/A N/A N/A
K-S
α3 (#6) K-S N/A Yes N/A N/A N/A N/A N/A
α4 (#7) K-S irrational Yes Pronounced 1 ± 0.1 3 ± 0.1 0.75 ± 0.5 ±
al

spike of 0.1 0.1


solutes
α4 (#8)
n

K-S {110}α//{111}γ No Negligible 1.75 ± 0.1 2 0.4 0.13


spike of
ur

solutes
A rational OR was observed in grains α2 (APT specimens 2, 3 in Figs. 5a and 6a, b, Table 1), α3
Jo

(APT specimen 6 in Fig. 7a, e) and α4 (APT specimens 7, 8 in Figs. 7a and 8a, b, Table 1). APT

specimens 2 and 3 from ferrite grain α2 and APT specimen 8 from ferrite grain α4 (Table 1) did

not demonstrate the presence of interphase precipitates/clusters, while APT specimen 6 from

ferrite grain α3 and APT specimen 7 from ferrite grain α4 (Table 1) clearly showed the formation

of interphase precipitates/clusters. Hence, it appears that interphase precipitates were formed at

interfaces between ferrite and austenite grains having both irrational and rational ORs. The

foregoing observation does not support the findings by Miyamoto et al. [25] and Zhang et al.

21
[27] that interphase precipitation is suppressed at the interfaces between ferrite and prior

austenite grains with a K-S OR. However, the trace analysis/APT cross-correlative approach

revealed that the interfaces, which are close to the K-S habit plane criteria i.e. (110)α//(111)γ (α2

(#2, 3) and α4 (#8), Table 1), advanced without interphase precipitate formation. In other words,

the macroscopic rational OR between ferrite and austenite grains can, in some cases, create low-

energy, semi-coherent (110)α//(111)γ facets, where interphase precipitation is suppressed.

f
Therefore, a higher probability of interphase precipitate formation corresponds to ferrite-

oo
austenite interfaces with an irrational OR. This is due to the lower chance to have geometrically

pr
matching (110)α//(111)γ planes between grains holding irrational OR, where interphase

precipitation is suppressed. e-
The importance of the combined effect of both OR and interfacial composition, on the interphase
Pr
precipitation/clustering has been emphasized in [26] and will be discussed below.

4.2 The effect of solutes distribution across the interface on the interphase
al

precipitation/clustering
n

In the current work, the interphase precipitation/clustering was observed to form solely at
ur

interfaces, which had pronounced 1-1.5 ± 0.1 at.% C, 3 ± 0.1 at. % Mn and 0.75-1 ± 0.05 at. %
Jo

Ti concentration spikes (Fig. 4a, c and Fig. 8a, c). In contrast, the interphase

precipitation/clustering reaction was suppressed at the interfaces with negligible solute spikes of

0.5-1.75 ± 0.1 at.% C, 1.75-2 ± 0.1 at.% Mn and 0.05-0.3 ± 0.05 at.% Ti (Fig. 6a, c and Fig. 8b,

d), where the bulk alloy content had 0.19 at.% C, 1.54 at. % Mn and 0.06 at.% Ti. In the previous

studies local equilibria assumptions, namely NPLE and PE, were used to calculate the driving

force for the interphase precipitation [26] and model condition at interfaces during interphase

precipitation [23]. NPLE represents the pronounced spike of interstitial (C) and substitutional

22
(Mn, Mo, and Ti) solutes ahead of austenite/ferrite interface, while PE corresponds to more

pronounced spikes of C and no spikes of substitutional elements ahead of interface [1]. In the

current work it was proved that the interphase precipitates are formed near the interfaces with

prominent spikes of substitutional solutes such as Mn, Mo, Ti along with a spike of C, which is

close to the situation reported for NPLE [23, 26], while the interphase precipitates were not

observed in the situation close to the PE interfacial condition. In order to compare the

f
composition of interfaces obtained during the current study with the theoretical results calculated

oo
by ThermoCalc for NPLE and PE conditions, the 650°C isothermal section of Mn-C phase

pr
diagram was plotted (Fig. 9). The substitutional solute atoms such as Mo and Ti participate in

both the transformation and precipitation events, which can locally alter their concentration at the
e-
interface depending on the state of the precipitation/transformation event. Considering this, we
Pr
made an assumption to use C and Mn as reference elements to identify equilibria conditions at

interfaces, since Mn participates only in transformation (not in precipitation). Figure 9a shows


al

the Mn-C phase diagram with experimental and calculated Mn-C segregation limits to achieve
n

the PE (brown) and NPLE (green) domains during transformation. The compositions of the
ur

concentration spikes of the two groups of interfaces are marked as black triangles (spikes of Mn
Jo

and C, without interphase precipitation) and red squares (spikes of Mn, Ti, Mo and C with

interphase precipitation). The corresponding

23
f
oo
pr
e-
Pr
n al
ur
Jo

Fig. 9. (a) Isothermal section (650 °C) of Mn-C phase diagram for PE and NPLE conditions
calculated using Thermo-Calc. Star “B” is the current alloy composition. The green and brown
areas on the phase diagram are the theoretical NPLE and PE domains, respectively. Red squares
and black triangles denote datasets with NPLE interface condition (with interphase precipitation)
and PE (no interphase precipitation), respectively. D and E are limits for composition at the
NPLE and PE interfaces respectively. Dashed line “A” is the PE tie-line; (b) Schematic Mn
concentration profiles showing calculated NPLE and PE limits for Mn; (c) Schematic C
concentration profiles showing calculated NPLE limit for C; (d) Schematic C concentration
profiles showing calculated PE limit for C.

24
schematic concentration profiles of Mn and C at the interfaces for both NPLE and PE regimes

are given in Fig. 9b and Fig. 9c, d, accordingly. The Mn concentration near the interface for the

PE condition was about 2 ± 0.1 at. %, which is higher than the concentration predicted for the PE

condition, i.e., 1.54 at. % (line A in Fig. 9a, b). During transformation in the PE mode the

concentration of substitutional elements, such as Mn, near the interface is unaffected by the

transformation and it inherits the alloy composition, i.e., 1.54 at. % (line A in Fig. 9a, b)

f
although in this case Mn spikes between 1.75 and 2 ± 0.1 at. % were observed (black triangles in

oo
Fig. 9b). These additional 0.25-0.5 at. % increases in Mn concentration from the predicted PE

pr
limit could mean that the Mn atoms are being segregated directly to the interface in this case, i.e.,

a solute-drag effect of Mn. Concentration spike of Ti (less than 0.4 at. %) was also observed in
e-
both specimens with interfaces, which demonstrated the spikes of Mn and C. Although the Ti
Pr
concentration spike was not observed in the concentration profile of specimen 2 (Fig. 6c), Ti

atoms were seen to decorate the interface in the atom map (specimen 2 in Fig. 6a). However, the
al

presence of Ti spikes at the interfaces with spikes of Mn and C, can be also the consequence of
n

C2+ and Ti2+ mass spectrum peak overlap at 24 Da, which was assigned as Ti according to Dhara
ur

[22]. In the current work, 0.5-0.6 ± 0.1 at. % concentration spikes of Mo were only observed at
Jo

the interfaces with Mn, Ti, Mo and C spikes (Table 1), suggesting that Mo is completely

immobile under this regime conditions and did not participate in solute drag (direct segregation

to the interface) or local equilibrium.

Although the particular interface (specimens 7 and 8 from grain α4 in Figs. 7, 8a, b) reached

neither the calculated NPLE nor PE limits (D and E respectively in Fig. 9), it could be seen that

the C concentration near the PE interface (black triangle 8 in Fig. 9d with 2 at. %) was higher

than near the NPLE interface (red square 7 in Fig. 9c with 1 at. %). This is in agreement with

25
calculations that the C concentration near the interface in the case of PE (4.75 at. % black dot E

in Fig. 9d) is always higher than for NPLE (3.6 at. % black dot D in Fig. 9c) [1].

Therefore, it is suggested that the interphase precipitation/clustering is mainly affected by the

local interfacial condition, i.e. the amount of substitutional solutes in front of and at the interface,

rather than the macroscopic OR between ferrite and prior austenite grains. The local interfacial

conditions could not be described by the PE and NPLE modes of transformation, since they did

f
not meet the predicted limits. Also, due to local trajectory aberrations in the APT (known as the

oo
local magnification effect) [49] complicates this study since it is impossible to establish the true

pr
peak concentration at the spike and the true concentration spike width. In other words, the local

aberration could result in peak width uncertainty, which consequently could lower the peak
e-
concentration and make it deviate from that calculated. Nevertheless, the interphase precipitates
Pr
in the current case are formed only at the interfaces, where pileups of substitutional solute atoms

(Mn, Ti and Mo) were more pronounced.


al

4.3 Effect of local austenite/ferrite interfacial conditions with a rational OR on


n

interphase precipitation/clustering
ur

It is well known that the local interfacial energy varies depending on the type of interface, i.e.,
Jo

coherent, semi-coherent, and incoherent. It is important to highlight that the APT specimens,

which were lifted-out from different areas but within the same interface, displayed different

interfacial conditions, i.e., with Mn, Ti, Mo, C spikes and Mn, C spikes (specimens 7 and 8 in

Figs. 8a, b, respectively). Even though the macroscopic OR between the α4 ferrite and γ4 prior

austenite was near the K-S, rational OR (Fig. 7b) the degree of coherency of interfaces in

specimens 7 and 8 (Fig. 8a, b) should be different. Austenite-to-ferrite transformation involves

participation of the interface network where interfaces are faceted with different degree of

26
coherence. It affects the mobility of the interface leading to different solute distribution and

phase equilibria states at the interface. For example, the local interface in sample 8 was close to

the K-S OR with a habit plane of (110)α//(111)γ (Table 1) and to the PE local interfacial

condition. These conditions suppressed interphase precipitation (Figs 7, 8). In contrast, sample 7,

which was taken across the interface in close proximity to sample 8 (Fig. 7), is also from the

local interface with K-S OR, however, its habit plane deviates from the ideal (110)α//(111)γ K-S

f
conditions leading to conditions with Mn, Ti, Mo and C spikes. It supports the formation of

oo
interphase precipitates (Fig. 8). So, one part of the interface network can be coherent (with low

pr
mobility), while the other part can be incoherent (with high mobility) [50]. This brings

complexity in understanding how the macroscopic OR between the ferrite and austenite grains
e-
can affect the interphase precipitation phenomenon. Therefore, it may have local variation in the
Pr
segregation or pileup of solute atoms at the interface has a more pronounced effect on interphase

precipitation than the macroscopic OR. It appears that the solute spike of substitutional elements
al

provides a sufficient diffusion flux of solutes for interphase precipitation to occur (Figs. 4a, c and
n

8a, c). The solute drag effect [35] should also be considered, as segregation of alloying elements
ur

can significantly lower the interface energy and retard its movement to provide enough time for
Jo

interphase precipitation. Since the interphase precipitation reaction occurs only at interfaces with

solutes concentration spikes, it is believed that interphase precipitation is rather a random event

and related mainly to local coherency of the interfaces.

4.4 The formation mechanism of randomly distributed interphase precipitates

Interphase precipitation is believed to exist in two types according to their distribution: (i) planar

interphase precipitation and (ii) random interphase precipitation [6], [28]. Interphase

precipitation is considered to be planar if the microstructure comprises parallel precipitation

27
sheets with a fixed distance between each other [4]. The parallel sheets of precipitates are

believed to form as a result of successive interface ledge migrations, during which interphase

precipitates nucleate at the ledge terraces, i.e. the ledge mechanism [6]. However, Ricks [28]

observed another mechanism, which leads to the formation of interphase precipitates with

random distribution, where the interface advances by bulging between freshly formed interphase

precipitates, i.e., the bowing mechanism [28], akin to Smith-Zener drag [51] in grain growth. The

f
fact that sheets of planar interphase precipitates can be observed in TEM only if the electron

oo
beam is parallel to the precipitation sheets (otherwise precipitates appear to be randomly

pr
distributed) can lead to a discrepancy in the interpretation of experimental data to explain the

mechanism of interphase precipitation [4]. The strength of the APT technique, in this case, is that
e-
3D reconstructions can be investigated from different angles and the sheets of precipitates are
Pr
easily revealed. Nevertheless, in the current work, the sheets of precipitates (planar interphase

precipitation) were observed only in grain α3 (Fig. 7e), whereas in all other cases, the precipitates
al

were randomly distributed (Fig.4a, 6d, and 8a), therefore considered to be random interphase
n

precipitates.
ur

Even though precipitates were randomly distributed near the interface (specimen 1, Fig.4a), a

stepped structure of the α1/α’(γ1) ferrite/austenite interface decorated by Mn atoms was observed
Jo

(Fig. 4b, d). It is generally accepted for the ledge mechanism that the distance between

precipitation planes is equal to the ledge height [6], which in the current work is ~ 10 nm for this

TMP conditions (Fig. 7c). However, the ledge at the α1/α’(γ1) interface had a height of 5 nm and

was observed at the interface that had Mn, Ti, Mo and C concentration spikes in front of/at the

interface leading to the formation of interphase precipitates with a random distribution (Fig. 4a).

Therefore, the ledge height at the α1/α’(γ1) interface was less than the minimum of the plane

28
spacing of the planar interphase precipitates of 10 nm and of the order of the size of the

interphase precipitates (4 - 5 nm), suggesting that the ledge height decrease (less than 10 nm)

could lead to a change in interphase precipitate distribution from planar to random. At this

condition, interphase precipitates formed by the ledge mechanism appeared as being randomly

distributed, which could suggest that the ledge mechanism operates even in the case of random

interphase precipitation.

f
5. Conclusions

oo
The interphase precipitation phenomenon in a Ti-Mo steel was studied by cross-correlation of

pr
EBSD and APT near the ferrite/austenite interfaces. The most significant findings of the paper

are summarized as follows: e-


1. We refined the generally accepted theory that interphase precipitates form preferentially
Pr
at K-S interfaces with a close-packed habit plane of (110)α//(111)γ and low mobility, i.e.,

rational interfaces. Instead, we found that the rational interfaces had a negligible
al

concentration spikes of Mn (1.75 - 2 at. %) with no interphase precipitation.


n

2. By contrast, the interfaces with a K-S OR but with habit planes that deviate from
ur

(110)α//(111)γ or the incoherent interfaces with non-K-S OR, i.e., irrational interfaces,
Jo

allowed solutes to partition locally with pronounced concentration spikes of substitutional

elements (Mn, Ti and Mo) across the interface. In this case, the interface mobility could

decrease due to retardation from solute partitioning and interphase

precipitation/clustering occurs.

3. Even if the austenite and ferrite have a K-S OR, certain local facets of the interface

between them can have different degrees of coherency and mobility promoting conditions

to develop pronounced Mn, Ti, Mo and C spikes or Mn and C concentration spikes. In

29
this case, interfaces with pronounced Mn, Ti, Mo and C concentration spikes stimulate

the formation of interphase precipitates at the interface, while conditions with Mn and C

concentration spikes retard it.

4. It was suggested that random interphase precipitates can form via the ledge mechanism at

an interface, where diffusion flux of solute elements (Ti and Mo) maintained by

pronounced solute concentration spikes in the interface area.

f
CRediT authorship contribution statement

oo
Ilias Bikmukhametov: conceptualisation, investigation, formal analysis, writing original draft;

pr
Hossein Beladi: formal analysis, writing - review and editing; Jiangting Wang: formal analysis,

writing - review and editing; Vahid Tari: formal analysis, software; Anthony D. Rollett: formal
e-
analysis, software; Peter D. Hodgson: project administration, writing - review and editing; Ilana
Pr
Timokhina: funding acquisition, supervision, writing - review and editing.

Declaration of Competing Interest


al

☒ The authors declare that they have no known competing financial interests or personal
n

relationships that could have appeared to influence the work reported in this paper.
ur

☐The authors declare the following financial interests/personal relationships which may be considered as
potential competing interests:
Jo

30
Acknowledgements

The authors would like to acknowledge the financial support of the Australian Research Council,

Australia via DP150103062 discovery grant. Deakin University's Advanced Characterization

Facility is acknowledged for use of the EBSD, TEM and APT instruments. ADR acknowledges

support from the Office of Naval Research under grant number N00014-18-1-2786.

Declaration of Competing Interest

f
The authors declare that they have no known competing financial interests or personal

oo
relationships that could have appeared to influence the work reported in this paper.

pr
e-
Pr
References

[1] M. Gouné, F. Danoix, J. Ågren, Y. Bréchet, C.R. Hutchinson, M. Militzer, G. Purdy, S. van
der Zwaag, H. Zurob, Overview of the current issues in austenite to ferrite transformation and the
al

role of migrating interfaces therein for low alloyed steels, Mater. Sci. Eng. R 92(0) (2015) 1-38.
[2] G. Miyamoto, R. Hori, B. Poorganji, T. Furuhara, Interphase precipitation of VC and
n

resultant hardening in V-added medium carbon steels, ISIJ Int. 51(10) (2011) 1733-1739.
[3] C.-Y. Chen, M.-H. Liao, Synergistic effects of carbon content and Ti/Mo ratio on
ur

precipitation behavior of HSLA steel: Insights from experiment and critical patent analysis,
Materials & Design 186 (2020) 108361.
Jo

[4] I. Bikmukhametov, H. Beladi, J. Wang, P.D. Hodgson, I. Timokhina, The effect of strain on
interphase precipitation characteristics in a Ti-Mo steel, Acta Materialia 170 (2019) 75-86.
[5] S. Mukherjee, I.B. Timokhina, C. Zhu, S.P. Ringer, P.D. Hodgson, Three-dimensional atom
probe microscopy study of interphase precipitation and nanoclusters in thermomechanically
treated titanium–molybdenum steels, Acta Mater. 61(7) (2013) 2521-2530.
[6] A.T. Davenport, F.G. Berry, R.W.K. Honeycombe, Interphase precipitation in iron alloys,
Metal Sci. 2(1) (1968) 104-106.
[7] H.-W. Yen, P.-Y. Chen, C.-Y. Huang, J.-R. Yang, Interphase precipitation of nanometer-
sized carbides in a titanium–molybdenum-bearing low-carbon steel, Acta Mater. 59(16) (2011)
6264-6274.
[8] M.Y. Chen, M. Gouné, M. Verdier, Y. Bréchet, J.R. Yang, Interphase precipitation in
vanadium-alloyed steels: Strengthening contribution and morphological variability with austenite
to ferrite transformation, Acta Mater. 64(0) (2014) 78-92.

31
[9] J. Wang, P.D. Hodgson, I. Bikmukhametov, M.K. Miller, I.B. Timokhina, Effects of hot-
deformation on grain boundary precipitation and segregation in Ti-Mo microalloyed steels,
Mater. Des. 141 (2018) 48-56.
[10] S. Mukherjee, I.B. Timokhina, C. Zhu, S.P. Ringer, P.D. Hodgson, Clustering and
precipitation processes in a ferritic titanium-molybdenum microalloyed steel, J. Alloys Comp.
690 (2017) 621-632.
[11] J. Wang, M. Weyland, I. Bikmukhametov, M.K. Miller, P.D. Hodgson, I.B. Timokhina,
Transformation from cluster to nano-precipitate in microalloyed ferritic steel, Scr. Mater 160
(2019) 53-57.
[12] Y. Zhang, G. Miyamoto, K. Shinbo, T. Furuhara, Comparative Study of VC, NbC, and TiC
Interphase Precipitation in Microalloyed Low-carbon Steels, Metallurgical and Materials
Transactions A 51(12) (2020) 6149-6158.

f
[13] Z. Xiong, I. Timokhina, E. Pereloma, Clustering, nano-scale precipitation and strengthening

oo
of steels, Prog. Mater. Sci. 118 (2021) 100764.
[14] A. Kostryzhev, C. Killmore, E. Pereloma, Effect of Processing Parameters on Interphase
Precipitation and Mechanical Properties in Novel CrVNb Microalloyed Steel, Metals. 11 (2021).

pr
[15] J. Wang, H. Beladi, L. Pan, F. Fang, J. Hu, L. Kong, P.D. Hodgson, I. Timokhina,
Interphase precipitation hardening of a TiMo microalloyed dual-phase steel produced by
continuous cooling, Mater. Sci. Eng. A. 804 (2021) 140518.
e-
[16] Y. Funakawa, T. Shiozaki, K. Tomita, T. Yamamoto, E. Maeda, Development of high
Strength hot-rolled sheet steel consisting of ferrite and nanometer-sized carbides, ISIJ Int. 44
(2004).
Pr
[17] K. Seto, Y. Funakawa, S. Kaneko, Hot rolled high strength steels for suspension and chassis
parts “NANOHITEN” and “BHT Steel”, JFE tech. rep. (2007) 19-25.
[18] N. Kamikawa, Y. Abe, G. Miyamoto, Y. Funakawa, T. Furuhara, Tensile behavior of Ti,
Mo-added low carbon steels with interphase precipitation, ISIJ int. 54(1) (2014) 212-221.
al

[19] N. Singh, G. Casillas, D. Wexler, C. Killmore, E. Pereloma, Application of advanced


experimental techniques to elucidate the strengthening mechanisms operating in microalloyed
n

ferritic steels with interphase precipitation, Acta Mater. 201 (2020) 386–402.
[20] Y. Funakawa, T. Fujita, K. Yamada, Metallurgical features of NANOHITEN and
ur

application to warm stamping, JFE tech. rep. (2013) 74-79.


[21] Y. Wang, S. Clark, B. Cai, D.A. Venero, K. Yan, M. Gorley, E. Surrey, D. McCartney, S.
Jo

Sridhar, P. Lee, Small-angle neutron scattering reveals the effect of Mo on interphase nano-
precipitation in Ti-Mo micro-alloyed steels, Scripta Materialia 174 (2020) 24-28.
[22] S. Dhara, R.K.W. Marceau, K. Wood, T. Dorin, I.B. Timokhina, P.D. Hodgson,
Precipitation and clustering in a Ti-Mo steel investigated using atom probe tomography and
small-angle neutron scattering, Mater. Sci. Eng. A 718 (2018) 74-86.
[23] R. Lagneborg, S. Zajac, A model for interphase precipitation in V-microalloyed structural
steels, Metall. Mater. Trans. A 32(1) (2001) 39-50.
[24] R. Okamoto, J. Ågren, A model for interphase precipitation based on finite interface solute
drag theory, Acta Mater. 58(14) (2010) 4791-4803.
[25] G. Miyamoto, R. Hori, B. Poorganji, T. Furuhara, Crystallographic analysis of proeutectoid
ferrite/austenite interface and interphase precipitation of vanadium carbide in medium-carbon
steel, Metall. Mater. Trans. A 44(8) (2013) 3436-3443.

32
[26] Y.J. Zhang, G. Miyamoto, K. Shinbo, T. Furuhara, Quantitative measurements of phase
equilibria at migrating α/γ interface and dispersion of VC interphase precipitates: Evaluation of
driving force for interphase precipitation, Acta Mater. 128 (2017) 166-175.
[27] Y.J. Zhang, G. Miyamoto, K. Shinbo, T. Furuhara, T. Ohmura, T. Suzuki, K. Tsuzaki,
Effects of transformation temperature on VC interphase precipitation and resultant hardness in
low-carbon steels, Acta Mater. 84 (2015) 375-384.
[28] R.A. Ricks, P.R. Howell, The formation of discrete precipitate dispersions on mobile
interphase boundaries in iron-base alloys, Acta Metallurgica 31(6) (1983) 853-861.
[29] T. Murakami, H. Hatano, G. Miyamoto, T. Furuhara, Effects of ferrite growth rate on
interphase boundary precipitation in V microalloyed steels, ISIJ int. 52(4) (2012) 616-625.
[30] C.-Y. Chen, C.-C. Chen, J.R. Yang, Synergistic effect of austenitizing temperature and hot
plastic deformation strain on the precipitation behavior in novel HSLA steel, Mater. Sci. Eng. A

f
639 (2015) 145-154.

oo
[31] H. Youliang, S. Godet, J.J. Jonas, Observations of the Gibeon meteorite and the inverse
Greninger-Troiano orientation relationship, J. Appl. Cryst. 39(1) (2006) 72-81.
[32] H. Beladi, G.S. Rohrer, A.D. Rollett, V. Tari, P.D. Hodgson, The distribution of intervariant

pr
crystallographic planes in a lath martensite using five macroscopic parameters, Acta Mater. 63
(2014) 86-98.
[33] R. Honeycombe, R. Mehl, Transformation from austenite in alloy steels, Metallurgical
e-
Transactions A 7(7) (1976) 915-936.
[34] G. Miyamoto, N. Iwata, N. Takayama, T. Furuhara, Mapping the parent austenite
orientation reconstructed from the orientation of martensite by EBSD and its application to
Pr
ausformed martensite, Acta Mater. 58 (2010) 6393-6403.
[35] J. Agren, Y. Brechet, C. Hutchinson, J. Philibert, G. Purdy, Thermodynamics and phase
transformations: the selected works of Mats Hillert, (2006).
[36] A. Hultgren, Isothermal transformation of austenite, Transactions of the American Society
al

for Metals 39 (1947) 915-1005.


[37] M. Hillert, Solute drag, solute trapping and diffusional dissipation of Gibbs energy, Acta
n

Mater. 47(18) (1999) 4481-4505.


[38] H. Aaronson, W. Reynolds, G. Purdy, Coupled-solute drag effects on ferrite formation in
ur

Fe-CX systems, Metallurgical and Materials Transactions A 35(4) (2004) 1187-1210.


[39] F. Danoix, X. Sauvage, D. Huin, L. Germain, M. Gouné, A direct evidence of solute
Jo

interactions with a moving ferrite/austenite interface in a model Fe-C-Mn alloy, Scr. Mater. 121
(2016) 61-65.
[40] I. Bikmukhametov, Investigation of clusters and nano-particles in thermomechanically
processed Ti-Mo steel, Deakin University, 2019.
[41] B. Hutchinson, J. Komenda, G.S. Rohrer, H. Beladi, Heat affected zone microstructures and
their influence on toughness in two microalloyed HSLA steels, Acta Materialia 97 (2015) 380-
391.
[42] V. Tari, A.D. Rollett, H. Beladi, Back calculation of parent austenite orientation using a
clustering approach, J. Appl. Crystall. 46(1) (2013) 210-215.
[43] V. Shterner, I.B. Timokhina, A.D. Rollett, H. Beladi, The role of grain orientation and grain
boundary characteristics in the mechanical twinning formation in a high manganese twinning-
induced plasticity steel, Metallurgical and Materials Transactions A 49(7) (2018) 2597-2611.

33
[44] P.M. Kelly, A. Jostsons, R.G. Blake, The orientation relationship between lath martensite
and austenite in low carbon, low alloy steels, Acta Metallurgica et Materialia 38(6) (1990) 1075-
1081.
[45] H. Beladi, G.S. Rohrer, The relative grain boundary area and energy distributions in a
ferritic steel determined from three-dimensional electron backscatter diffraction maps, Acta
Materialia 61(4) (2013) 1404-1412.
[46] M.K. Miller, K.F. Russell, G.B. Thompson, Strategies for fabricating atom probe specimens
with a dual beam FIB, Ultramicroscopy 102 (4) (2005) 287-298.
[47] M.K. Miller, R.G. Forbes, Atom-probe tomography: the local electrode atom probe,
Springer, 2014.
[48] S. Dhara, R.K.W. Marceau, K. Wood, T. Dorin, I.B. Timokhina, P.D. Hodgson, Atom probe
tomography data analysis procedure for precipitate and cluster identification in a Ti-Mo steel,

f
Data Brief (2018).

oo
[49] B. Gault, M.P. Moody, J.M. Cairney, S.P. Ringer, Atom probe microscopy, Springer , 2012.
[50] T. Furuhara, H.I. Aaronson, On the mechanisms of interphase boundary carbide
precipitation, Scripta Metallurgica 22 (10) (1988) 1635-1637.

pr
[51] C.S. Smith, Grains, phases, and interfaces: an introduction of microstructure, Trans. Metall.
Soc. AIME 175 (1948) 15-51.
e-
Pr

Figure captions
n al
ur
Jo

34
Jo
ur
nal

35
Pr
e-
pr
oo
f
Jo
ur
nal

36
Pr
e-
pr
oo
f
Jo
ur
nal

37
Pr
e-
pr
oo
f
Jo
ur
nal

38
Pr
e-
pr
oo
f
Jo
ur
nal

39
Pr
e-
pr
oo
f
Jo
ur
nal

40
Pr
e-
pr
oo
f
Jo
ur
nal

41
Pr
e-
pr
oo
f
Jo
ur
nal

42
Pr
e-
pr
oo
f
Jo
ur
nal

43
Pr
e-
pr
oo
f
Highlights

 It was shown for the first time that the interphase precipitation is promoted at both interfaces

between the ferrite/austenite grains with K-S and non K-S OR depending on accommodation

of substitutional solutes at the interphase boundary.

 Interphase precipitates are formed at the interfaces with prominent concentration spikes of

Mn, Ti, Mo, C and suppressed at the interfaces with negligible Mn, C concentration spikes.

f

oo
Conditions with negligible Mn, C concentration spikes corresponded to the regions near the

interfaces with habit plane close to ideal K-S

pr
 Austenite-to-ferrite transformation ledge structure first time was captured in the APT
e-
Pr
n al
ur
Jo

44

You might also like