You are on page 1of 11

Available online at www.sciencedirect.

com

ScienceDirect
Acta Materialia 77 (2014) 17–27
www.elsevier.com/locate/actamat

Atomic and nanoscale chemical and structural changes in quenched


and tempered 4340 steel
A.J. Clarke a,⇑, M.K. Miller b, R.D. Field a,c, D.R. Coughlin a, P.J. Gibbs a, K.D. Clarke a,
D.J. Alexander a, K.A. Powers b, P.A. Papin a, G. Krauss c
a
Los Alamos National Laboratory, Materials Science and Technology Division, PO Box 1663, Los Alamos, NM 87545, USA
b
Oak Ridge National Laboratory, Center for Nanophase Materials Science, PO Box 2008, Oak Ridge, TN 37831, USA
c
Colorado School of Mines, 1500 Illinois St., Golden, CO 80401, USA

Received 26 February 2014; received in revised form 14 May 2014; accepted 19 May 2014

Abstract

Atom probe tomography and transmission electron microscopy (TEM) have been used to determine the location and distribution of
carbon and alloying elements associated with the complex structural changes that occur at the atomic and nanoscale in 4340 steel after
quenching to martensite and tempering at 325, 450 or 575 °C. Tempering at 325 °C resulted in carbide formation without partitioning of
chromium, manganese, molybdenum, aluminum, nickel or phosphorus, but with early-stage silicon rejection from the carbide. TEM
verified the presence of cementite and the Bagaryatsky orientation relationship with the tempered martensite matrix and detected com-
plex precipitate structures. Tempering at 450 or 575 °C developed concentrations of all alloying elements at ferrite–cementite interfaces:
chromium, manganese and molybdenum partitioned into the cementite, and silicon, aluminum, nickel and phosphorus were clearly
rejected from the cementite. These results provide direct evidence for staged cementite growth, where early-stage growth likely occurs
under paraequilibrium conditions, followed by initial silicon redistribution and subsequent alloying element redistribution during
late-stage growth. Tempering at 575 °C induced spheroidization of the cementite, loss of the Bagaryatsky orientation relationship,
and phosphorus concentrations at Cottrell atmospheres within the cementite and at ferrite–cementite interfaces, correlating with early
observations of the retardation of spheroidization by phosphorus.
Ó 2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Steel; Tempering; Martensite; Atom probe tomography (APT); Transmission electron microscopy (TEM)

1. Introduction resistance to softening during tempering. Although some


alloy carbides may be retained, austenitizing is designed to
Low-alloy, medium-carbon steels are austenitized, dissolve all carbon and other alloying elements into the
quenched to martensite and tempered at various tempera- face-centered cubic crystal structure of the austenite. As a
tures to provide beneficial combinations of strength, tough- result, the martensite, formed by the diffusionless shear trans-
ness and fracture resistance. These steels are alloyed with formation of the austenite, is supersaturated with carbon and
substitutional alloying elements, such as chromium (Cr), other alloying elements compared to their equilibrium con-
manganese (Mn), molybdenum (Mo) and nickel (Ni), to centrations in ferrite-cementite microstructures. These super-
provide good hardenability, i.e. depth of hardening, and saturations are relieved by the precipitation of a variety of
carbide types and morphologies during tempering [1].
It is well known that the alloying elements in steels
⇑ Corresponding author. Tel.: +1 505 665 3467. retard softening during tempering, and Grange et al. [2]
E-mail address: aclarke@lanl.gov (A.J. Clarke). have quantitatively documented the effects of various

http://dx.doi.org/10.1016/j.actamat.2014.05.032
1359-6454/Ó 2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
18 A.J. Clarke et al. / Acta Materialia 77 (2014) 17–27

elements in steels in retaining the hardness of as-quenched APT was performed using a local electrode atom probe
structures as a function of element content and tempering (Cameca Instruments Inc. LEAPÒ 4000X HR). Sections
temperature. However, the mechanisms of alloying element were machined from quenched and tempered steel plate
redistribution as the saturation in martensite is relieved are and mechanically ground to a thickness of 0.25 mm.
only now beginning to be characterized by virtue of recent Abrasive wire-cutting, followed by mechanical grinding,
developments in atom probe tomography (APT). was then used to fabricate specimen blanks with a square
At low tempering temperatures, carbon is mobile, but cross-section, nominally 0.25  0.25 mm.
the diffusivity of the substitutional elements is too low to Specimen blanks were electropolished into sharp needles
significantly influence the microstructure and substructure by a two-stage double-layer and micropolishing technique
of the martensite [1,3]. Therefore, carbon plays a dominant [12]. The electrolytes used were 25% perchloric acid
role in the chemical and structural changes that occur, (70%) in glacial acetic acid for rough polishing and 2%
permitting the achievement of high strengths and moderate perchloric acid in 2-butoxyethanol for final polishing [12].
toughness after tempering below 250 °C [1,3–5]. At Focused ion beam (FIB) annular milling was then
tempering temperatures above 250 °C, the replacement of performed in an FEI Nova 200 instrument to sharpen each
transition carbides and low-carbon martensite by cementite needle for APT. For the LEAP analyses, the specimen
and ferrite occurs [1,6–9]. Higher tempering temperatures temperature was 50 K and the specimens were run in
(>425 °C) provide sufficient thermal energy for substitu- voltage mode. A pulse fraction of 0.2 and pulse repetition
tional elements to be mobile and influence structural rate of 200 kHz were used.
coarsening, resulting in microstructures that exhibit TEM specimens were prepared by two different meth-
combinations of high toughness and moderate strengths ods: electropolishing and FIB milling. The electropolishing
[1,4,5]. solution used was 95% glacial acetic acid with 5% per-
Babu et al. reported early-stage cementite growth under chloric acid. Polishing was performed at room temperature
paraequilibrium conditions (i.e. dependent only upon car- in a Fishione automatic polisher at a potential of 25–30 V.
bon diffusion) in an Fe–C–Si–Mn steel after tempering at The FIB foil preparation was performed in an FEI DB235
lower temperatures, whereas higher tempering tempera- dual-beam FIB equipped with an Omniprobe nanomanip-
tures and longer times resulted in Si and Mn partitioning ulator for foil extraction. TEM analysis was performed in
[10]. Zhu et al. have shown no partitioning of Cr, Mo, an FEI TF30 Tecnai instrument operating at 300 kV.
Mn or Ni with the initiation of Si partitioning after temper-
ing 4340 steel at 300, 350 or 400 °C for 1 h [11]. Tempering 3. Results
at 400 °C for 10 h resulted in significant Cr and Mn parti-
tioning within the cementite and the rejection of Si and Ni 3.1. As-quenched martensitic microstructure
[11]. The purpose of this study was to systematically eval-
uate the redistribution of interstitial and substitutional ele- Oil quenching of the 4340 steel produced a lath martens-
ments and the subtle structural changes that occur in 4340 itic microstructure. Martensite crystals contain dislocation
steel at the atomic and nanoscale after quenching and densities as high as 1012 cm2, resulting from the lattice
tempering between 325 and 575 °C for 2 h using APT invariant deformation component of the diffusionless mar-
and transmission electron microscopy (TEM). tensitic phase transformation [13]. Fine twins were also
observed in some laths, as shown in the bright-field TEM
2. Experimental procedure images in Fig. 1.
Nanoscale, plate-like precipitates were also revealed by
Aircraft-quality 4340 steel was used for this study; the TEM, as shown in Fig. 2a. Selected-area diffraction
composition of this steel in wt.% and at.% is provided in (Fig. 2b) suggests that these precipitates are e0 (also referred
Table 1. Nominally 6.35 mm thick plate was austenitized to as g) transition carbide [9,14,15]. Three reflections, labeled
under vacuum for 30 min at 845 °C, oil quenched, and tem- as A, B and C, are highlighted in Fig. 2b. Position A corre-
pered in air at 325 or 450 °C or under vacuum at 575 °C for sponds to a d-spacing of 0.20 nm, consistent with f0 2 1ge0
2 h. A sample was also encapsulated in quartz under ultra- planes. Position B corresponds to superlattice reflections
high-purity argon, austenitized for 30 min at 845 °C, and from the ordered e0 carbide and position C results from dou-
then ice-water quenched for microstructural comparison ble diffraction between f0 2 1ge0 carbide and f1 1 0ga martens-
with the oil-quenched condition. ite reflections. The pattern is consistent with two variants of
the following orientation relationship [15]:

Table 1
4340 steel composition.
C Mn Si Ni Cr Mo Cu P S
wt.% 0.42 0.78 0.26 1.78 0.83 0.24 0.03 0.010 0.003
at.% 1.92 0.78 0.51 1.67 0.88 0.14 0.03 0.018 0.005
A.J. Clarke et al. / Acta Materialia 77 (2014) 17–27 19

Fig. 1. (a) Bright-field TEM image of lath martensite produced by oil quenching 4340 steel, containing twins highlighted by the arrow in (b). The inset
selected-area diffraction pattern in (b) is indexed as a h1 1 0i beam direction containing {1 1 2} twins and is properly rotated with respect to the image to
show the {1 1 2} habit plane. (c) Simulated electron diffraction pattern. The filled circles correspond to primary a reflections, the open circles correspond to
twin reflections of the filled circles, and the “x” symbols correspond to double-diffraction reflections.

Fig. 2. (a) Bright-field TEM image of nanoscale carbides (examples indicated by the arrows) in oil-quenched 4340 and (b) selected-area diffraction using
the ½0 0 1a zone axis. Data are consistent with that of Taylor et al. [15] for the e0 carbide with an orientation relationship of ½1 0 0e0 k½0 0 1a and
ð0 0 1Þe0 kð1 1 0Þa . Position A corresponds to f0 2 1ge0 reflections, position B corresponds tof0 1 1ge0 superlattice reflections, and position C corresponds to
double diffraction between f0 2 1ge0 and f1 1 0ga reflections. ð0 2 0Þe0 and ð0 0 2Þe0 reflections should be present near f1 1 0ga , but are presumed to be too weak
to be visible in close proximity to the much stronger matrix reflections.
20 A.J. Clarke et al. / Acta Materialia 77 (2014) 17–27

½1 0 0e0 k½0 0 1a other substitutional elements, has been observed and
ð0 0 1Þe0 kð1 1 0Þa shown to correlate with the more rapid diffusion of Si com-
pared to that of Cr, Mn and Ni at temperatures between
From a tempering study of Fe–C–Ni steel, Taylor et al. 300 and 400 °C [11].
reported initial f0 2 3ga “spinodal” habit planes that TEM micrographs and selected-area diffraction and
migrated toward f0 2 1ga for large precipitates [14,15]. microdiffraction results from the martensitic matrix and
These habit plane normals are indicated in Fig. 2a for ref- carbide are shown in Fig. 5. The carbide contains internal
erence. The large precipitates in the central region of structure with laths ranging from 10 to 20 nm (Fig. 5a),
Fig. 2a have e0 carbide habit planes near ð0 2 0Þa , suggesting along with finer features ranging from 1 to 2 nm
that these precipitates are well beyond any early stage (Fig. 5b). Diffraction data from these features are
“spinodal” clustering. consistent with the presence of Fe3C (cementite) with the
APT revealed ultra-fine-scale carbon clustering in the Bagaryatsky orientation relationship [20–22] (from
oil-quenched martensitic microstructure, which is high- Fig. 5c–f) for the tempered martensite and cementite:
lighted by the 5 at.% carbon isoconcentration surfaces in
Fig. 3a. Similar clustering, although qualitatively finer ½0 1 1 k½0 0 1
a Fe3 C
in scale, produced after ice-water quenching is shown in ð2 1 1Þa kð0 1 0ÞFe3 C
Fig. 3b. The features indicated by the arrows in Fig. 3a
and b display maximum carbon levels approaching 12–14 at.%, Double diffraction accounts for the (weaker) reflections
as shown in the proximity histograms. Carbon atom in  16 f1 1 2ga positions. Complementary selected-area dif-
redistribution during the quench contributes to the high fraction and microdiffraction results from other precipi-
strength of as-quenched 4340 steel [16]. Such rearrange- tates sometimes contained reflections only in  13 f1 1 2ga
ment has been attributed to carbon segregation to disloca- positions (i.e. reflections at 16 f1 1 2ga positions are missing),
tions and the formation of Cottrell atmospheres [17–19], which may indicate cementite-like carbide and/or twin-
with carbon levels ranging from 6 to 8 at.% [18,19]. The related structure. Simulated electron diffraction patterns
maximum carbon levels associated with the ultra-fine-scale using the Isaichev orientation relationship [21,23] reveal
1
carbon clusters in Fig. 3 exceed the range typically 3
f1 1 2ga reflections, but additional reflections are also pre-
associated with Cottrell atmospheres at dislocations; these dicted that are not experimentally observed. Nagakura
levels are also lower than those anticipated for transition et al. have reported that the transition from chi carbide
carbides or cementite. (Fe5C2) to Fe3C occurs by microsyntactic growth, or car-
bide intergrowth, requiring only iron atom displacements
3.2. Martensitic microstructure tempered at 325 °C and carbon diffusion [1,24,25].

APT revealed the formation of plate-like features with 3.3. Martensitic microstructure tempered at 450 °C
carbon levels above the equilibrium amount of 25 at.%
anticipated for Fe3C-cementite after 325 °C tempering for APT of a carbide after tempering at 450 °C for 2 h is
2 h, as shown in Fig. 4. There is no evidence of Cr, Mn shown in Fig. 6a. A proximity histogram with a maximum
or Mo partitioning into the carbide (Fig. 4b), and the Ni, carbon level of 25 at.% is provided in Fig. 6b, suggesting
Al and P levels are similar in the carbide and matrix that the carbide is cementite. Strong evidence of Cr, Mn
(Fig. 4c). There is apparent Si rejection from the carbide and Mo partitioning to the carbide is also highlighted. Less
(Fig. 4c). Similar rejection of Si from cementite at low clear trends with respect to the carbide exist for Si, Al, Ni
tempering temperatures, with no detectable diffusion of and P in the proximity histogram in Fig. 6c, but based

Fig. 3. (a and b) Ultra-fine-scale carbon clusters in 4340 after oil or ice-water quenching, respectively, highlighted by 5 at.% C isoconcentration surfaces,
and proximity histograms showing the carbon levels associated with the features indicated by the arrows. APT results.
A.J. Clarke et al. / Acta Materialia 77 (2014) 17–27 21

Fig. 4. (a) C atom map and a 10 at.% C isoconcentration surface highlighting the carbide–matrix interface indicated by the arrow after oil quenching and
tempering 4340 steel at 325 °C for 2 h and (b and c) proximity histograms showing the C, Cr, Mn, Mo, Si, Al, Ni and P levels associated with the carbide
and martensitic matrix. Error bars are provided every fifth data point in the proximity histograms. APT results.

upon the rejection of Si from the carbide after tempering at achieved. Atom maps for Si, Al, Ni and P are shown in
325 °C, it is speculated that Si is also rejected from the Fig. 7e–h, respectively. A proximity histogram is shown
carbide at 450 °C. in Fig. 7j, highlighting the behaviors of these elements with
Carbides formed during high-temperature tempering are respect to the cementite and matrix. The Si map in Fig. 7e
intra- or inter-lath or at prior austenite grain boundaries. and the proximity histogram in Fig. 7j clearly show that Si
Inter-lath carbides, typical of retained austenite decompo- is rejected from the cementite and partitions to the ferrite
sition to ferrite and cementite, were observed after temper- matrix, but small Si clusters remain within the cementite,
ing at 450 °C. Complex carbides, similar to those observed perhaps trapped at internal structure or defects. The
after tempering at 325 °C (e.g. Fig. 5) with selected-area behaviors of Al, Ni and P are less obvious, but closer exam-
diffraction and microdiffraction reflections at  16 f1 1 2ga ination of the proximity histogram and carbide–matrix
or  13 f1 1 2ga positions, were also observed after temper- interface in Fig. 7j reveals that these elements behave
ing at 450 °C. similarly to Si. Clustering of these elements within the
cementite, often at the same location(s) as the Si, is also
3.4. Martensitic microstructure tempered at 575 °C qualitatively observed in the element maps. P appears to
not only be segregated near the carbide–matrix interface
APT results after tempering at 575 °C for 2 h are shown in Fig. 7h, but is also segregated within the cementite.
in Fig. 7. A plate-like carbide is defined by 10 at.% isocon- From three-dimensional visualization of this data set at dif-
centration surfaces in Fig. 7a. These carbon isoconcentra- ferent orientations, the lower arrowed P-enriched region is
tion surfaces are also shown in Fig. 7b–g to highlight the linear in morphology, and therefore, most likely a disloca-
location of the carbide within the analyzed volume and tion. The extent of P segregation around the line is consis-
the complex element partitioning associated with the car- tent with that of a dislocation. Such features have been
bide and ferrite (matrix). Atom maps are provided for attributed to APT observations of Cottrell atmosphere for-
Cr, Mn and Mo in Fig. 7b–d, respectively. A complemen- mation at dislocations [17], but in the case of the cementite
tary proximity histogram for these elements with respect to may also have been associated with an internal boundary in
the upper isoconcentration surface is provided in Fig. 7i. the cementite, as shown in Fig. 5.
The maximum carbon content is slightly higher than, but Bright-field TEM images of the ferrite (matrix) and
close to, that anticipated for equilibrium cementite coarse carbides after tempering at 575 °C for 2 h are shown
(25 at.%). Cr, Mn and Mo clearly partition to the in Fig. 8a and b. Diffraction from four different carbides in
cementite, with elevated levels within the cementite adja- three different ferrite grains matched orthorhombic cement-
cent to the carbide–matrix interface. Equilibrium partition- ite, with lattice parameters of a = 5.090 Å, b = 6.748 Å and
ing of these elements within the cementite has not yet been c = 4.523 Å. A selected-area diffraction pattern for the
22 A.J. Clarke et al. / Acta Materialia 77 (2014) 17–27

Fig. 5. Cementite with internal structure (arrows) and selected-area diffraction for (a) the ½113 a zone axis and (b) the ½0 1 1a zone axis after oil quenching
and tempering 4340 steel at 325 °C for 2 h. (c) Selected-area diffraction for the ½0 1 1a zone axis and (d) a simulated electron diffraction pattern for the
matrix (filled circles) and cementite (open circles) and microdiffraction for (e) the matrix and (f) cementite used to determine the Bagaryatsky orientation
relationship. Bright-field TEM images and electron diffraction.

cementite and ferrite is shown in Fig. 8c, and a simulated Bagaryatsky, did not produce the observed rotation in
electron diffraction pattern is shown in Fig. 8d. The Bagar- the experimental pattern. Low dislocation density is also
yatsky orientation relationship was determined. Many observed in the ferrite, suggesting that 575 °C tempering
cementite particles still have a f1 1 2ga habit plane, but is sufficient to result in significant recovery of the
some particles appear to be spheroidizing, resulting in dislocation substructure produced by the martensitic
slight rotation of the cementite with respect to the ferrite, transformation.
as shown in Fig. 7c. This rotation suggests a loss of precip-
itate coherency with the matrix and rotation away from the 4. Discussion
Bagaryatsky orientation relationship. Simulated electron
diffraction patterns based on the Isaichev orientation rela- Medium-carbon, alloyed steels are typically tempered
tionship, which can be considered as a small rotation from after quenching to achieve the desired hardness, strength,
A.J. Clarke et al. / Acta Materialia 77 (2014) 17–27 23

Fig. 6. (a) C atom map and a 5 at.% C isoconcentration surface highlighting the carbide-matrix interface indicated by the arrow after oil quenching and
tempering 4340 steel at 450 °C for 2 h and (b and c) proximity histograms showing the C, Cr, Mn, Mo, Si, Al, Ni and P levels associated with the cementite
and the matrix. Error bars are provided every fifth data point in the proximity histograms. APT results.

ductility and/or toughness. Along with sometimes dramatic the quench (to room temperature or perhaps an intermedi-
mechanical property changes, quenching and tempering ate temperature below Ms for subsequent thermal process-
also results in subtle microstructural changes. Three dis- ing) will set the stage for subsequent tempering reactions,
tinct tempering stages are generally accepted for carbon such as carbide formation that occurs during the first stage
steels, including: stage I tempering from 100 to 250 °C of tempering [1,28,29].
involving the formation of transition carbide and reduction APT showed a carbon-enriched plate with a maximum
of the martensite carbon content; stage II tempering from carbon level exceeding the 25 at.% anticipated for equilib-
200 to 300 °C involving the decomposition of retained aus- rium cementite after tempering at 325 °C, traditionally a
tenite into ferrite and cementite; and stage III tempering stage III tempering temperature. Cr, Mn and Mo did not
from 250 to 350 °C involving the replacement of transition partition within the carbide or tempered martensite. Simi-
carbide and low-carbon martensite by cementite and ferrite larly, Al, Ni and P did not partition, but Si was rejected
[1,6,13,26,27]. Importantly, these tempering stages overlap from the carbide into the matrix. After tempering an Fe–
and the progression of microstructural changes that occur C–Mn–Si steel at 350 or 400 °C for 1.8  103 s (0.5 h),
are subtle and significantly influenced by alloying element Babu et al. reported no Mn or Si partitioning, suggesting
additions. early stage paraequilibrium growth of cementite, whereas
TEM revealed lath martensite, instances of twinning, tempering at 350 °C for 6.48  105 s (180 h) or 500 °C for
and indications of transition carbide that may have formed 0.3  103 s (0.83 h) resulted in Si rejection from and Mn
close to the martensite start (Ms) temperature during oil enrichment into the cementite [10]. Zhu et al. reported that
quenching in this 4340 steel. APT revealed ultra-fine-scale quenching and tempering of 4340 steel at 300 or 350 °C for
carbon clustering after quenching with maximum carbon 1 h resulted in features containing 12–14 at.% carbon, with-
levels ranging from 12 to 14 at.%, which are inconsistent out Cr, Mn or Ni partitioning or substantial Si redistribu-
with those anticipated for transition carbides (e.g. e, e0 or tion [11]. The carbon levels identified by Zhu et al. are
g, with anticipated carbon levels near 30 at.% or above). much lower than those obtained for the plate-like carbide
The experimentally obtained carbon levels appear to be in Fig. 4 and agree better with those obtained after quench-
influenced by quench rate, as shown qualitatively in ing (Fig. 3). This discrepancy may be associated with
Fig. 3. Importantly, the carbon clustering produced during quench rate and/or sample size variations, or perhaps carbon
24 A.J. Clarke et al. / Acta Materialia 77 (2014) 17–27

Fig. 7. (a) C atom map and 10 at.% C isoconcentration surfaces (red) defining the carbide–matrix interface after oil quenching and tempering 4340 steel at
575 °C for 2 h and atom maps for (b) Cr, (c) Mn, (d) Mo, (e) Si, (f) Al and (g) Ni, including the 10 at.% C isoconcentration surfaces, and (h) P segregation.
(i and j) Proximity histograms from the upper isoconcentration surface showing complex element partitioning in the cementite and ferrite matrix. Error
bars are provided every fifth data point in the proximity histograms. APT results. (For interpretation of the references to color in this figure legend, the
reader is referred to the web version of this article.)
A.J. Clarke et al. / Acta Materialia 77 (2014) 17–27 25

Fig. 8. (a and b) Ferrite (matrix) and cementite after oil quenching and tempering 4340 steel at 575 °C for 2 h. (c) Selected-area diffraction for the ½0 1 1a
zone axis and (d) a simulated electron diffraction pattern for the matrix (filled circles) and cementite (open circles). Some rotation of the cementite
diffraction data with respect to the matrix is observed in (c)—note splitting of reflections at 1 and 7 o’clock in (c), suggesting coherency loss during
spheroidization and rotation away from the Bagaryatsky orientation relationship. Bright-field TEM images and selected-area electron diffraction.

detection during APT [30–33]. Systematic studies on the After tempering at 450 or 575 °C, APT revealed plate-
location of substitutional elements with respect to transition shaped carbides containing 25 at.% C, the amount
carbide are limited, but Barnard et al. reported that after tem- expected in cementite. Cr, Mn and Mo clearly concentrated
pering an Fe–C–Si–Mn–Cr steel for 24 h at 250 °C, epsilon in the cementite close to the ferrite–carbide interface, with
carbide was not enriched with Si [34]. concentrations decreasing into the cementite (e.g. Fig. 7).
Although the elevated carbon level in Fig. 4 suggests that The Si, Al, Ni and P trends with respect to the cementite
transition carbide is possible, APT alone cannot be used to and ferrite were less clear after tempering at 450 °C, but
conclusively assign whether transition carbide or cementite Si appears to be rejected from the cementite. After
is present. Cementite was confirmed by TEM selected-area tempering at 575 °C, Si, Al, Ni and P were all noticeably
diffraction and microdiffraction for the 325 °C condition rejected from the cementite, persisting near the ferrite–
(Fig. 5), but more complex electron diffraction was also carbide interface. After tempering 4340 steel at 400 °C for
observed, perhaps indicating cementite-like carbide and/or 1 h, Zhu et al. reported carbide with carbon levels near
twin-related structure. Based upon the available TEM and 14 at.%, without significant Cr, Mn, Ni or Si partitioning
APT shown here, the carbide in Fig. 4 is more likely [11]. Tempering at 400 °C for 10 h resulted in a carbon level
early-stage cementite with initial Si redistribution, which exceeding 20 at.% and apparent Si and Ni rejection from
was reported by Babu et al. [10]. Although Babu et al. sug- and Mn and Cr enrichment into the carbide [11]. The Cr,
gest the presence of cementite, their selected-area diffraction Mn, Si and Ni trends observed in the current work agree
reveals a f2 0 0ga carbide habit plane, whereas the selected- with the results of Zhu et al. [11]. The current results
area diffraction in Fig. 5 reveals a f1 1 2ga cementite habit provide direct evidence for staged cementite growth, with
plane. Low-temperature tempering and more complemen- early paraequilibrium growth based only on carbon
tary APT/TEM microstructural studies may help to unam- diffusion, followed by initial silicon redistribution and then
biguously clarify the details associated with the carbides growth influenced by the diffusion of carbide-forming
present after tempering at 325 °C. elements to and within the cementite.
26 A.J. Clarke et al. / Acta Materialia 77 (2014) 17–27

While APT revealed complex substitutional element tempering. Attainment of desirable mechanical properties
partitioning and P segregation near the ferrite–carbide in steels such as 4340 is related to the redistribution of sol-
interface, the TEM images from the steel tempered at ute and subtle microstructural changes that occur during
575 °C for 2 h showed coarsened cementite with indications tempering. APT and TEM were used to systematically
of spheroidization and loss of the Bagaryatsky orientation examine the chemical and structural changes that occur
relationship (Fig. 8) compared to more typical elongated at the atomic and nanoscale in 4340 steel after quenching
carbides that result from lower-temperature tempering and tempering at 325, 450 or 575 °C for 2 h, resulting in
(Fig. 4), although it is important to note that complex car- the following conclusions:
bides were still present in the 575 °C, 2 h temper condition.
P segregation at the carbide–matrix interface has been sug- (1) Lath martensite, instances of twinning and indica-
gested as a contributing mechanism for slowing cementite tions of e0 (g) transition carbide were identified in
growth and retarding the spheroidization of elongated car- 4340 steel after oil quenching using TEM. Transition
bide structures [35,36]. The subtle cementite spheroidiza- carbides may have formed close to the martensite
tion at 575 °C suggests that this temperature is indeed at start (Ms) temperature during quenching.
the upper limit of the temperature range (generally 475– (2) APT revealed ultra-fine-scale carbon clustering in
575 °C) where solute mobility effectively retains elongated 4340 steel, with maximum carbon levels of 12–14 at.%.
carbide structures. Additionally, the enrichment of Si, Al Quench rate appears to influence the extent of carbon
and Ni at the ferrite–cementite interface suggests that the clustering, setting the initial microstructural condi-
relative mobility of these elements may also contribute to tion for subsequent tempering reactions.
the retention of elongated carbides in alloyed steels such (3) From APT, tempering 4340 steel at 325 °C resulted in
as 4340. Alloy effects on carbide morphology have signifi- a carbon-enriched plate (>25 at.%) without Cr, Mn,
cant ramifications for the mechanical performance of Mo, Al, Ni or P partitioning and the initiation of Si
quenched and tempered steels. For instance, James et al. rejection from the carbide.
showed that elongated carbides retained during medium- (4) TEM electron diffraction was used to identify
temperature tempering could effectively limit the decrease cementite after tempering at 325 or 450 °C. The
in strength during tempering without significantly decreas- Bagaryatsky orientation relationship for cementite
ing ductility [37]. However, Materkowski and Krauss and the tempered martensite was determined, and
reported a plateau in toughness during Stage III tempering precipitates with internal structure after tempering
due to easy crack propagation along elongated brittle car- at 325 or 450 °C were observed. Complex electron
bides [38]. Comparing high- and low-P 4340 steels, Mater- diffraction sometimes indicated cementite-like precip-
kowski and Krauss also showed a decrease in toughness for itates and/or twin-related structure.
the high-P steel over the entire tempering range examined (5) APT after tempering at 450 or 575 °C showed
[38]. cementite with carbon levels near 25 at.% and com-
Si and Al are commonly used to modify carbide forma- plex element partitioning. Cr, Mn and Mo unambig-
tion temperatures and kinetics in low-carbon steels, thereby uously partitioned to the cementite. Si appears to be
promoting carbon availability for enriching and stabilizing rejected from cementite at 450 °C, without significant
austenite (e.g. during austempering to produce transforma- Al, Ni or P partitioning trends. Tempering at 575 °C
tion-induced plasticity (TRIP) steels with carbon-enriched results in noticeable rejection of Si, Al, Ni and P from
retained austenite). Direct experimental evidence that Al the cementite.
and Si behave similarly with respect to partitioning from (6) The APT and TEM results presented here provide
cementite during martensite tempering is shown in Fig. 7, direct evidence for staged cementite growth, where
making Al an effective alloying addition to help prevent early-stage growth likely occurs under paraequilibri-
cementite formation during processing. Although Al is um conditions, followed by initial silicon redistribu-
often used to partially replace Si in alloying strategies for tion and subsequent alloying element redistribution
promoting austenite retention, the results presented here during late-stage growth.
suggest that P plays a similar role in retarding cementite (7) APT clearly revealed P segregation at ferrite–cementite
development. Thus, alloying strategies including P, espe- interfaces and the formation of Cottrell atmospheres
cially in conjunction with Si and/or Al, may present an within the cementite.
additional approach for controlling cementite development (8) From TEM, coarsened cementite existed after tem-
and austenite retention, keeping in mind the possibility of pering at 575 °C, with indications of spheroidization
embrittlement associated with P segregation to austenite and loss of the Bagaryatsky orientation relationship.
grain boundaries [38–42].

5. Conclusions Acknowledgements

Alloying of medium-carbon steels permits a range of The authors gratefully acknowledge support from Los
strength and toughness combinations to be achieved after Alamos National Security, LLC, operator of the Los
A.J. Clarke et al. / Acta Materialia 77 (2014) 17–27 27

Alamos National Laboratory under contract number DE- [17] Chang L, Barnard SJ, Smith GDW. The segregation of carbon atoms
AC52-06NA25396 with the US Department of Energy and to dislocations in low-carbon martensites: studies by atom field ion
microcopy and atom probe microanalysis. In: Krauss G, Repas PE,
the Innovative Manufacturing Initiative of the Office of editors. Gilbert R. Speich symposium proceedings: fundamentals of
Advanced Manufacturing (AMO) at the US Department aging and tempering in bainitic and martensitic steel products. War-
of Energy (Award DE-EE0005765). Atom probe tomogra- rendale, PA: Iron and Steel Society; 1992. p. 19.
phy research (M.K.M. and K.A.P.) was supported through [18] Wilde J, Cerezo A, Smith GDW. Scr Mater 2000;43:39.
a user project supported by Oak Ridge National Labora- [19] Cochardt AW, Schoek G, Wiedersich H. Acta Metall 1955;3:533.
[20] Bhadeshia HKDH. Worked examples in the geometry of crystals. 2nd
tory’s Center for Nanophase Materials Sciences (CNMS), ed. London: Institute of Materials; 2001.
which is sponsored by the Scientific User Facilities [21] Zhou DS, Shiflet GJ. Metall Mater Trans A 1992;23A:1259.
Division, Office of Basic Energy Sciences, US Department [22] Bagaryatsky YA. Dokl Akad Nauk SSSR 1950;73:1161.
of Energy. J.E. Scott, T.J. Tucker, T.V. Beard, and R.W. [23] Isaichev IV. Zh Tekh Fiz 1947;17:835.
Hudson are thanked for specimen heat-treatment and [24] Nagakura S, Hirotsu Y, Kusunoki M, Suzuki T, Nakamura Y.
Metall Trans A 1983;14A:1025.
machining assistance. Helpful discussions with Professor [25] Nagakura S, Suzuki T, Kusunoki M. Trans Jpn Inst Metals 1981;22:
S.W. Thompson regarding TEM electron diffraction data 699.
and discussions with Professors G.D.W. Smith, J.G. [26] Speich GR. Tempered ferrous martensitic structures. In: Metallogra-
Speer, E. De Moor, B.C. De Cooman, D.K. Matlock, phy, structures and phase diagrams. Metals handbook, Vol. 8. Mate-
and S.S. Babu and Dr. F.G. Caballero are also gratefully rials Park, OH: American Society for Metals; 1973. p. 202.
[27] Speich GR, Leslie WC. Metall Trans 1972;3:1043.
acknowledged. [28] Genin JMR, Flinn PA. Trans Metall Soc AIME 1968;242:1419.
[29] Kurdjumov GV, Khachaturyan AG. Metall Trans 1972;3:1069.
References [30] Marceau RKW, Choi P, Raabe D. Ultramicroscopy 2013;132:239.
[31] Miyamoto G, Shinbo K, Furuhara T. Scr Mater 2012;67:999.
[1] Krauss G. STEELS: processing, structure, and performance. Materi- [32] Takahashi J, Kawakami K, Kobayashi Y. Ultramicroscopy 2011;111:
als Park, OH: ASM International; 2005. 1233.
[2] Grange RA, Hribal CR, Porter LF. Metall Trans A 1977;8A:1775. [33] Yao L, Gault B, Cairney JM, Ringer SP. Philos Mag Lett 2010;90:
[3] Zhu C, Cerezo A, Smith GDW. Ultramicroscopy 2009;109:545. 121.
[4] Grossman MA, Bain EC. Principles of heat treatment. 5th ed. Metals [34] Barnard SJ, Smith GDW, Garratt-Reed AJ, Vander Sande J. Atom
Park, OH: American Society of Metals; 1964. probe studies: (1) the role of silicon in tempering of steel and (2) low
[5] Saeglitz M, Krauss G. Metall Mater Trans A 1997;28A:377. temperature chromium diffusivity in bainite. In: Aaronson HI,
[6] Williamson DL, Schupmann RG, Materkowski JP, Krauss G. Metall Laughlin DE, Sekerka RF, Wayman CM, editors. Proceedings of
Trans A 1979;10A:379. an international conference on solid-solid phase transforma-
[7] Lee H-C, Krauss G. Intralath carbide transition in martensitic tions. Warrendale, PA: Metallurgical Society of AIME; 1982. p. 881.
medium carbon steels tempered between 200 and 300 °C. In: Krauss [35] Rellick JR, McMahon CJ. Metall Trans 1974;5:2439.
G, Repas PE, editors. Gilbert R. Speich symposium proceedings: [36] Brown EL, Krauss G. Metall Trans A 1986;17A:31.
fundamentals of aging and tempering in bainitic and martensitic steel [37] James BA, Matlock DK, Krauss G. Interactive effects of phosphorus
products. Warrendale, PA: Iron and Steel Society; 1992. p. 39. and tin on carbide evolution and fatigue properties of 5160 steel. 38th
[8] Williamson DL, Nakazawa K, Krauss G. Metall Trans A 1979; mechanical processing and steel working conference proceedings, Vol.
10A:1351. XXXIV. Warrendale, PA: Iron and Steel Society; 1997. p. 579.
[9] Hirotsu Y, Nagakura S. Trans Jpn Inst Metals 1974;15:1074. [38] Materkowski JP, Krauss G. Metall Trans A 1979;10A:1643.
[10] Babu SS, Hono K, Sakurai T. Metall Mater Trans A 1994;25A: [39] Chen HC, Era H, Shimizu M. Metall Trans A 1989;20A:437.
499. [40] Traint S, Pichler A, Hauzenberger K, Stiaszny P, Werner E. Influence
[11] Zhu C, Xiong XY, Cerezo A, Hardwicke R, Krauss G, Smith GDW. of silicon, aluminum, phosphorus and copper on the phase transfor-
Ultramicroscopy 2007;107:808. mations of low alloyed TRIP-steels. In: De Cooman BC, editor.
[12] Miller MK. Atom probe tomography. New York: Kluwer Academic/ International conference on trip-aided high strength ferrous alloys—
Plenum Press; 2000. proceedings. Aachen, Mainz: GRIPS; 2002. p. 121.
[13] Speich GR. Trans Metall Soc AIME 1969;245:2553. [41] Gallagher MF, Speer JG, Matlock DK. Microstructure development
[14] Taylor KA, Chang L, Olson GB, Smith GDW, Cohen M, Vander in TRIP-sheet steels containing Si, Al, and P. 44th mechanical
Sande JB. Metall Trans A 1989;20A:2717. processing and steel working conference proceeding, Vol. XL. War-
[15] Taylor KA, Olson GB, Cohen M, Vander Sande JB. Metall Trans A rendale, PA: Iron and Steel Society; 2002. p. 153.
1989;20A:2749. [42] Zhao L, van Dijk NH, Peekstok ER, Tegus O, Brück E, Sietsma J.
[16] Leslie WC, Sober RJ. ASM Trans Q 1967;60:459. Mater Sci Forum 2007;539–543:4321.

You might also like