interfacial segregation at cu-rich precipitates in a high...

9
Interfacial segregation at Cu-rich precipitates in a high-strength low-carbon steel studied on a sub-nanometer scale Dieter Isheim * , Michael S. Gagliano, Morris E. Fine, David N. Seidman Department of Materials Science and Engineering, Northwestern University, 2220 Campus Drive, Evanston, IL 60208-3108, USA Received 13 January 2005; received in revised form 25 September 2005; accepted 4 October 2005 Available online 19 December 2005 Abstract The local composition of small, coherent Cu-rich precipitates with a metastable body-centered cubic structure in a ferritic a-Fe matrix of a high-strength low-carbon steel was studied by conventional atom-probe tomography. The average diameter, ÆDæ, of the precipitates is 2.5 ± 0.3 nm at a number density of (1.1 ± 0.3) · 10 24 m 3 after direct aging at 490 °C for 100 min to a near-peak hardness condition, yielding a value of 84 Rockwell G. Besides Cu, the precipitates contain 33 ± 1 at.% Fe and are enriched in Al (0.5 ± 0.1 at.%). Nickel and Mn are significantly segregated at the a-Fe matrix/precipitate heterophase interfaces. The Gibbsian interfacial excesses relative to Fe and Cu are 1.5 ± 0.4 atoms nm 2 for Ni and 1.0 ± 0.3 atoms nm 2 for Mn. The reduction of the interfacial free energy, calculated utilizing the Gibbs adsorption isotherm, is 16 mJ m 2 for Ni and 11 mJ m 2 for Mn. Ó 2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved. Keywords: Interface segregation; Gibbsian interfacial excess; Ferritic steels; Precipitation hardening; Nanostructure 1. Introduction Copper precipitation plays an important role in the hardening of high-strength low-carbon steels [1–4]. The formation of Cu-rich precipitates during thermal aging has been studied by many techniques: atom-probe field- ion microscopy (APFIM) [5–12], small-angle neutron scat- tering (SANS) [8,13,14], high-resolution and conventional electron microscopy (HREM and CTEM) [15–17], theoret- ical modeling [13], and computer simulation [18]. APFIM, HREM, CTEM, and SANS studies indicate that small Cu precipitates have, at diameters less than about 5 nm, a metastable body-centered cubic (bcc) structure and are coherent with the ferritic a-Fe matrix [5,14–17]. In the 5–10 nm diameter range, the precipitates transform to a 9R structure [15–17], and eventually, at even larger diameters, gradually into the stacking-fault-free face-centered cubic (fcc) equilibrium structure of pure Cu. Interfacial segregation of Ni and Mn at the a-Fe matrix/ precipitate interfaces has been observed by APFIM for Cu precipitates 5–10 nm in diameter in Fe–Cu–Ni and Fe– Cu–Ni–Mn alloys [7–9,19]. At this size, however, the pre- cipitates are larger than the critical diameter for transfor- mation and loss of coherency of about 5 nm, and thus likely have transformed to a 9 R or faulted fcc structure and are no longer bcc and coherent with the a-Fe matrix. After neutron irradiation, Ni and Mn have also been found, employing APFIM and atom-probe tomography (APT), to be enriched in the interior and in the vicinity of smaller Cu-rich clusters and precipitates in nuclear pres- sure vessel steels and model alloys [9,19–24]. Thermody- namics, kinetics, loss of coherency, and structural transformation behavior under neutron irradiation, how- ever, are considerably different from the case of thermal decomposition. There is little quantitative information about the segregation of solute atoms at the interfaces of coherent bcc Cu-rich precipitates smaller than 5 nm in diameter in thermally aged alloys. We address this question employing APT to investigate Cu-rich precipitates in a newly developed high-strength low-carbon steel. A formalism is 1359-6454/$30.00 Ó 2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved. doi:10.1016/j.actamat.2005.10.023 * Corresponding author. Tel.: +1 847 491 3575; fax: +1 847 467 2269. E-mail address: [email protected] (D. Isheim). www.actamat-journals.com Acta Materialia 54 (2006) 841–849

Upload: others

Post on 09-May-2020

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Interfacial segregation at Cu-rich precipitates in a high ...arc.nucapt.northwestern.edu/Refbase/Files/IsheimCuPPTActaMat2005.pdfInterfacial segregation at Cu-rich precipitates in

Interfacial segregation at Cu-rich precipitates in a high-strengthlow-carbon steel studied on a sub-nanometer scale

Dieter Isheim *, Michael S. Gagliano, Morris E. Fine, David N. Seidman

Department of Materials Science and Engineering, Northwestern University, 2220 Campus Drive, Evanston, IL 60208-3108, USA

Received 13 January 2005; received in revised form 25 September 2005; accepted 4 October 2005Available online 19 December 2005

Abstract

The local composition of small, coherent Cu-rich precipitates with a metastable body-centered cubic structure in a ferritic a-Fe matrixof a high-strength low-carbon steel was studied by conventional atom-probe tomography. The average diameter, ÆDæ, of the precipitatesis 2.5 ± 0.3 nm at a number density of (1.1 ± 0.3) · 1024 m�3 after direct aging at 490 �C for 100 min to a near-peak hardness condition,yielding a value of 84 Rockwell G. Besides Cu, the precipitates contain 33 ± 1 at.% Fe and are enriched in Al (0.5 ± 0.1 at.%). Nickel andMn are significantly segregated at the a-Fe matrix/precipitate heterophase interfaces. The Gibbsian interfacial excesses relative to Fe andCu are 1.5 ± 0.4 atoms nm�2 for Ni and 1.0 ± 0.3 atoms nm�2 for Mn. The reduction of the interfacial free energy, calculated utilizingthe Gibbs adsorption isotherm, is 16 mJ m�2 for Ni and 11 mJ m�2 for Mn.� 2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Interface segregation; Gibbsian interfacial excess; Ferritic steels; Precipitation hardening; Nanostructure

1. Introduction

Copper precipitation plays an important role in thehardening of high-strength low-carbon steels [1–4]. Theformation of Cu-rich precipitates during thermal aginghas been studied by many techniques: atom-probe field-ion microscopy (APFIM) [5–12], small-angle neutron scat-tering (SANS) [8,13,14], high-resolution and conventionalelectron microscopy (HREM and CTEM) [15–17], theoret-ical modeling [13], and computer simulation [18]. APFIM,HREM, CTEM, and SANS studies indicate that small Cuprecipitates have, at diameters less than about 5 nm, ametastable body-centered cubic (bcc) structure and arecoherent with the ferritic a-Fe matrix [5,14–17]. In the5–10 nm diameter range, the precipitates transform to a 9Rstructure [15–17], and eventually, at even larger diameters,gradually into the stacking-fault-free face-centered cubic(fcc) equilibrium structure of pure Cu.

Interfacial segregation of Ni and Mn at the a-Fe matrix/precipitate interfaces has been observed by APFIM for Cuprecipitates 5–10 nm in diameter in Fe–Cu–Ni and Fe–Cu–Ni–Mn alloys [7–9,19]. At this size, however, the pre-cipitates are larger than the critical diameter for transfor-mation and loss of coherency of about 5 nm, and thuslikely have transformed to a 9 R or faulted fcc structureand are no longer bcc and coherent with the a-Fe matrix.After neutron irradiation, Ni and Mn have also beenfound, employing APFIM and atom-probe tomography(APT), to be enriched in the interior and in the vicinityof smaller Cu-rich clusters and precipitates in nuclear pres-sure vessel steels and model alloys [9,19–24]. Thermody-namics, kinetics, loss of coherency, and structuraltransformation behavior under neutron irradiation, how-ever, are considerably different from the case of thermaldecomposition. There is little quantitative informationabout the segregation of solute atoms at the interfaces ofcoherent bcc Cu-rich precipitates smaller than 5 nm indiameter in thermally aged alloys. We address this questionemploying APT to investigate Cu-rich precipitates in a newlydeveloped high-strength low-carbon steel. A formalism is

1359-6454/$30.00 � 2005 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

doi:10.1016/j.actamat.2005.10.023

* Corresponding author. Tel.: +1 847 491 3575; fax: +1 847 467 2269.E-mail address: [email protected] (D. Isheim).

www.actamat-journals.com

Acta Materialia 54 (2006) 841–849

Page 2: Interfacial segregation at Cu-rich precipitates in a high ...arc.nucapt.northwestern.edu/Refbase/Files/IsheimCuPPTActaMat2005.pdfInterfacial segregation at Cu-rich precipitates in

presented to describe quantitatively segregation by extract-ing the relative Gibbsian excess, which is the relevant ther-modynamic quantity to measure segregation at an interfacein a system with more than two components, from localcompositions measured by APT.

Cu-rich precipitates are studied in a newly developedMo- and Cr-free low-carbon steel with an excellent tensilestrength, up to 700 MPa (100 ksi). A Charpy V-notch testimpact fracture toughness of 64 J (47 ft lb) at �40 �C wasobtained for specimens aged to a near-peak hardness con-dition with a yield stress of 712 MPa [3,25–28]. For a ver-sion of this steel with 0.1 wt.% Ti added, the Charpyfracture energy at �79 �C was greater than 360 J (264 ft lb),since the specimen bent without breaking in the Charpyapparatus [26]. Commercial Cu precipitation-hardenedsteels need to contain Ni to prevent segregation of Cu tograin boundaries and thus prevent hot-shortness, that is,crack formation during hot rolling, and Mn additions togetter sulfur impurities. From a technological point of viewCu alloyed steels containing Ni and Mn are of significantinterest. Essentially the steel was developed by eliminatingthe Mo and Cr from the U.S. Navy�s high-strength low-car-bon HSLA 100 steel and optimizing the precipitation-hard-ening process [25].

The APT analyses presented here reveal a significantinterfacial segregation of Ni andMn at coherent bcc Cu-richprecipitates with an average diameter, ÆDæ, of 2.5 ± 0.3 nm.The Gibbsian interfacial excesses relative to Fe and Cu,determined by the methodology introduced in Section 2.2,are 1.5 ± 0.4 atoms nm�2 for Ni and 1.0 ± 0.3 atoms nm�2

for Mn.

2. Methods

2.1. Experimental procedures

The steel was fabricated at the Ispat-Inland Inc.Research Center, East Chicago, Indiana. A 45.5 kg(100 lb) laboratory heat was vacuum induction melted,cast into molds, hot rolled to 13 mm thick plates, andair cooled. Subsequent hot rolling to 2.5 mm was followedby a cold rolling step to the final thickness of 1.6 mm. Theaverage composition of this steel, as provided by theIspat-Inland Inc. Research Center, is given in Table 1.Following an austenitization treatment at 1100 �C for0.5 h, the specimens were quenched directly from this tem-perature to different aging temperatures and aged for100 min [3]. Direct aging was conducted for a fundamen-tal study of the nucleation and growth of precipitates inthis steel [3,4].

The hardness was measured as an indicator for strengthand to track the course of the precipitation. A Rockwellhardness tester was utilized, applying a 10 kg minor load,150 kg major load, and a 1.6 mm ball corresponding tothe Rockwell G scale. At least five hardness measurementswere recorded for each sample tested, with the averagebeing logged as the actual hardness value, and the standarddeviation as the uncertainty of that value.

FIM tips were prepared by electropolishing, using ini-tially a solution of 10 vol.% perchloric acid in acetic acidat 20 V dc at room temperature, and then a solution of2 vol.% perchloric acid in butoxyethanol at 15–12 V dc atroom temperature for the final tip preparation.

Conventional APT analyses were performed at a residualpressure of 3 · 10�8 Pa, utilizing a pulse–voltage-to-dc–voltage ratio of 20% at a pulse repetition frequency of1500 Hz. Visualization and data evaluation of the three-dimensional atom-by-atom datasets were performed withcustom software, ADAM 1.5, developed at NorthwesternUniversity [29]. Composition profiles with respect to dis-tance from the a-Fe matrix/precipitate heterophase inter-face were obtained with ADAM 1.5 using the proximityhistogram method or proxigram for short [30,31]. A10 at.% Cu isoconcentration surface served as a referencesurface for the proxigram analysis. This concentrationthreshold provided a reliable and topologically stable repre-sentation of the precipitates, that is, the changes of the mor-phology, size, and number of the precipitates were smallwhen varying the threshold around 10 at.% Cu. Isoconcen-tration surfaces were generated with ADAM 1.5 employingthe adaptive marching cube algorithm. The parameters usedto obtain a noise-free isoconcentration surface in ADAM1.5 were a voxel size of 0.8–1.0 nm and a delocalization dis-tance of 1.0 nm [32]. No additional noise suppression wasapplied, that is, the confidence sigma parameter of ADAM1.5 [29–32] was kept at a value of zero.

The diameter, D, of a precipitate containing n atoms inthe reconstruction was equated to the diameter of the vol-ume equivalent sphere:

D ¼ 6

pnXf

� �1=3

; ð1Þ

where the atomic volume, X, is equal to 1.178 · 10�2 nm3

for bcc Fe, and the overall detection and reconstructionefficiency, f, is estimated to be 0.6. The number of atoms,n, belonging to a precipitate was determined employingthe envelope method [33], based on clusters with more than20 Cu atoms separated not farther than a maximum sepa-ration distance of 0.5 nm and employing an envelope gridspacing of 0.15 nm.

Table 1Average composition of the high-strength low-carbon steel under investigation as determined by chemical analysis provided by Ispat-Inland Inc

Fe C Si Nb Mn Cu Ni Al

Concentration (wt.%) Balance 0.059 0.49 0.079 0.49 1.37 0.82 0.034Concentration (at.%) Balance 0.27 0.97 0.047 0.50 1.20 0.78 0.07

842 D. Isheim et al. / Acta Materialia 54 (2006) 841–849

Page 3: Interfacial segregation at Cu-rich precipitates in a high ...arc.nucapt.northwestern.edu/Refbase/Files/IsheimCuPPTActaMat2005.pdfInterfacial segregation at Cu-rich precipitates in

2.2. Determination of the relative Gibbsian excess of solute

at an interface

At constant temperature and pressure, the Gibbs adsorp-tion isotherm for a planar heterophase interface in a systemwith two phases, a and b, and n P 3 components is [34]

dr ¼ �Xn

i¼3

Crelativei dli; ð2Þ

where r is the interfacial free energy, li the chemical poten-tial of component i, and the relative Gibbsian interfacialexcess of component i is given by

Crelativei ¼ Ci � C1

cai cb2 � cbi c

a2

ca1cb2 � cb1c

a2

� C2

ca1cbi � cb1c

ai

ca1cb2 � cb1c

a2

; ð3Þ

where cki is the concentration of component i in phasek = a, b. The quantity Crelative

i is independent of the positionof the dividing interface [35] and can be calculated employ-ing Eq. (3) from the measured Gibbsian excesses, Ci, ob-tained for a specific choice of the position of an interface.The values Ci are determined utilizing proxigram concen-tration profiles for a specific interface [30,31] from

Ci ¼ qDxXp

m¼1

cim � cik� �

; ð4Þ

where q is the known atomic density, Dx is the distance be-tween the p concentration data points in the proxigram, andk = a on the matrix side and k = b on the precipitate side ofthe heterophase interface. The determination of interfacialexcesses Ci for one-dimensional concentration profilesacross grain boundaries in a two-component system by asimilar method has been discussed elsewhere [36,37].

One advantage of writing the Gibbs adsorption theoremin the form of Eq. (2) is that the (n � 3) values of dli(i = 3, . . . ,n) can be varied independently, and therefore acoefficient for the reduction of the interfacial free energyat a respective chemical potential, li, is given by

oroli

����T ;p;l3;...;li�1;liþ1;...;ln

¼ �Crelativei . ð5Þ

This coefficient can be obtained directly from the interfacialexcesses, Crelative

i , determined from Eq. (3). Using Henry�slaw for dilute solutions, Eq. (5) can be rewritten in the form

oroci

����T ;p;l3;...;li�1;liþ1;...;ln

¼ �kTCrelative

i

ci; ð6Þ

which is used to determine the coefficient of interfacial en-ergy reduction at a concentration ci due to segregation ofcomponent i at the heterophase interface, with a Gibbsianinterfacial excess Crelative

i .

3. Results

Fig. 1 displays an isochronal hardness curve for 100 minof direct aging at each different temperature. The maxi-

mum value for the hardness is 85.2 ± 0.1 Rockwell G,which is obtained at an aging temperature of 500 �C. Eachhardness value is the average of five independent measure-ments, and the error is the standard deviation of the fivevalues. A near-peak hardness condition with a high num-ber-density of nanometer-sized precipitates was studiedby conventional APT. The Cu-rich precipitates formedduring aging at 490 �C for 100 min are distinctively visiblein a three-dimensional atom-by-atom reconstruction,Fig. 2, displaying the positions of individual Cu atoms inthe analysis volume.

The spatial distributions of the solute components Cu,Ni, Si, Al, and Mn are shown in the three-dimensionalreconstructions in Fig. 3, representing a subset of the dataof Fig. 2, containing three Cu precipitates. It is seen qual-itatively that Ni, Al, and Mn are enriched at the locationsof the Cu-rich precipitates. When interpreting Fig. 3,however, account must be taken of the fact that thereconstruction represents atomic density rather than con-centration. In fact, during field evaporation, the precipi-tates develop a surface slightly recessed with respect tothe average curvature of a tip�s surface, that is, a surfacewith a larger radius of curvature, since Cu has a lowerevaporation field (30 V nm�1) than does Fe (36 V nm�1)[38]. Consequently, atoms field-evaporating from theCu-rich precipitates are projected toward the detector ata lower magnification than atoms field-evaporated fromthe a-Fe matrix. This causes an apparent higher atomicdensity for the precipitates than the matrix in the three-dimensional reconstruction, since the parameters usedfor the reconstruction are calibrated with respect to thea-Fe matrix. Fortunately, this effect affects only the twolateral dimensions of the reconstruction. The third dimen-sion, the depth, vertical in Fig. 2, is not affected and theprecipitates therefore appear slightly elongated alongthis direction. This distortion of a reconstructed Cu

Fig. 1. Isochronal hardness curve for 100 min aging time at temperaturesin the range 380–690 �C, after quenching from the austenitization treatmentat 1100 �C for 0.5 h [3]. The filled circle marks the condition used for theAPT investigations.

D. Isheim et al. / Acta Materialia 54 (2006) 841–849 843

Page 4: Interfacial segregation at Cu-rich precipitates in a high ...arc.nucapt.northwestern.edu/Refbase/Files/IsheimCuPPTActaMat2005.pdfInterfacial segregation at Cu-rich precipitates in

precipitate in an a-Fe matrix, resulting from this localmagnification effect, has been investigated in greater detailemploying computer simulation [39].

To determine the average precipitate diameter, ÆDæ,independently of the distortions due to the local magnifica-tion effect, the diameter of a volume equivalent undistortedsphere is calculated from the number of atoms in each pre-cipitate employing Eq. (1). Only precipitates fully con-tained in the reconstruction volume, that is, which arenot cut by the surface of the reconstructed volume, aretaken into consideration. The average diameter isÆDæ = 2.5 ± 0.3 nm. At this diameter, the precipitates areexpected to have a metastable bcc structure and to be fullycoherent with the a-Fe matrix.

The reconstructed volume displayed in Fig. 2 contains30 Cu-rich precipitates, of which 15 are cut by the surfacesof the reconstruction volume. Counting the cut precipitatesas one-half, a number density, Nv, of (1.1 ± 0.3) · 1024 m�3

is obtained.The proxigram method [30,31] is used to determine both

matrix and precipitate composition and interfacial segrega-tion. Since the proxigram measures compositions withrespect to distance from an isoconcentration surface, withno restrictions on the morphology of a precipitate, theelongated shape of the precipitates due to the local magni-fication effect does not affect their measured composition.Likewise, for the proxigram, it does not matter if a precip-itate is cut by the surface(s) of the reconstructed volume.Isoconcentration surfaces drawn at the 10 at.% Cu levelare used to delineate the Cu-rich precipitates and to definethe origin of the proxigram, serving as a fiducial mark formeasuring distances into a precipitate (positive values) anda-Fe matrix (negative values). Fig. 4 displays the proxi-gram concentration profiles for an individual precipitate.The center region of this precipitate is represented by thelast data point on the right-hand side of the diagram (posi-tive distances) and is enriched in Cu to 69 ± 4 at.%,Fig. 4(b), and contains 29 ± 4 at.% Fe, Fig. 4(a). Nickeland Mn are enriched in the interfacial region near the ori-gin of the diagram, Fig. 4(c) and (d), with peak concentra-tions of 2.8 ± 0.4 at.% Ni and 2.6 ± 0.4 at.% Mn,respectively, which is a significant enrichment by a factorof 3 and 5, respectively, with respect to the a-Fe matrixconcentrations of 0.95 and 0.55 at.%. The precipitate con-tains 0.5 ± 0.2 at.% Al, averaged from the last two datapoints in Fig. 4(e), which is an enrichment by a factor of10 with respect to the a-Fe matrix concentration of only0.05 at.% Al.

To improve the statistics and to extract quantitativeaverage values for concentrations and interfacial excesses,a proxigram of all precipitates in the reconstructed volume,Fig. 2, was calculated and is displayed in Fig. 5. Since allprecipitates are enveloped by a 10 at.% Cu isoconcentra-tion surface, the proxigram corresponds to a weightedsuperposition of concentration profiles across the interfacesof all precipitates, with all profiles aligned with the 10 at.%Cu level at the origin of the integral proxigram. A completeset of phase concentrations for all elements for Cu-rich pre-cipitates and the ferritic a-Fe matrix derived from Fig. 5 isprovided in Table 2. The center region of the precipitates,

Fig. 2. Positions of Cu atoms in an APT reconstruction, 14 · 14 · 101 nm3

in size, obtained from a steel aged at 490 �C for 100 min. Body-centeredcubic Cu-rich precipitates with an average diameter of 2.5 ± 0.3 nm haveformed, at a number density of (1.1 ± 0.3) · 1024 m�3.

844 D. Isheim et al. / Acta Materialia 54 (2006) 841–849

Page 5: Interfacial segregation at Cu-rich precipitates in a high ...arc.nucapt.northwestern.edu/Refbase/Files/IsheimCuPPTActaMat2005.pdfInterfacial segregation at Cu-rich precipitates in

that is, the plateau region consisting of the last two datapoints at the right-hand side of Fig. 5, has an average Cuconcentration of 64.0 ± 1.0 at.% and an average Fe con-centration of 33 ± 0.9 at.%. These concentrations are veryclose to the ones given above for a single precipitate. AnFe concentration this high in precipitates withÆDæ = 2.5 ± 0.3 nm or less is consistent with previousatom-probe studies [6,7] of Cu-rich precipitates of similar

size in thermally aged alloys. SANS experiments, however,indicate that precipitates with ÆDæ larger than 1 nm areessentially devoid of any Fe [13,24], even though some Fe(�10 at.%) can obviously not be accounted for by theSANS technique, as discussed in [24]. The Ni and Mn

Fig. 3. APT reconstructions of three Cu-rich precipitates, showing the spatial distributions of Cu, Ni, Si, Al, and Mn atoms. The reconstructed volume is2 nm in thickness and has a lateral cross-section of 15 · 14 nm2.

Fig. 4. Proxigram concentration profiles for (a) Fe, (b) Cu, (c) Ni, (d) Mn,(e) Al, and (f) Si for an individual Cu-rich precipitate. The origin is definedby a 10 at.% Cu isoconcentration surface. Fig. 5. Proxigram concentration profiles for (a) Fe, (b) Cu, (c) Ni, (d) Mn,

(e) Al, and (f) Si for all Cu-rich precipitates in Fig. 2, with respect to a10 at.% Cu isoconcentration surface. The cross-hatched areas in (c), (d),(e), and (f) correspond to the Gibbsian interfacial excesses of Ni, Mn, Al,and Si, respectively, at the interfaces of the precipitates.

D. Isheim et al. / Acta Materialia 54 (2006) 841–849 845

Page 6: Interfacial segregation at Cu-rich precipitates in a high ...arc.nucapt.northwestern.edu/Refbase/Files/IsheimCuPPTActaMat2005.pdfInterfacial segregation at Cu-rich precipitates in

concentrations in the centers of the precipitates,1.2 ± 0.2 at.% Ni and 0.76 ± 0.16 at.% Mn, are notstrongly enhanced with respect to the a-Fe matrix concen-trations of 0.95 ± 0.01 at.% Ni and 0.55 ± 0.01 at.% Mn,respectively. Aluminum, however, is, as noted above fora single precipitate, strongly enriched, by a factor of 10,in the interior of the precipitates with 0.50 ± 0.10 at.% Alin the precipitate center versus 0.047 ± 0.003 at.% Al inthe a-Fe matrix. The core of the Cu-rich precipitates isdepleted in Si, with a core concentration of0.60 ± 0.14 at.% Si versus a matrix concentration of0.99 ± 0.01 at.%.

An a-Fe matrix/precipitate interface location is indi-cated in Fig. 5 by the vertical line. The interface positionis arbitrarily chosen at the location with the mean valueof matrix and precipitate Cu concentration, and thus it isnot identical with the origin of the proxigram. As discussedpreviously, relative interfacial excesses, Crelative

i , are indepen-dent of the specific position of the dividing interface. Also,because of the possibility of an ion trajectory overlap effect,due to the reduced local magnification, the interface mayappear broader in the proxigram concentration profile thanin reality. Any artificial intermixing across the heterophaseinterface that results in a linear superposition of matrix andprecipitate atoms does not, however, affect the individualinterfacial excesses, Ci, since the matrix and precipitateconcentrations are used as reference concentrations, whichare subtracted in Eq. (4). Nickel and Mn show significantsegregation peaks at the interface with concentrations ofup to 2.5 at.% Ni and 1.8 at.% Mn, respectively, Fig. 5(c)and (d), corresponding to enhancement factors of 2.6 forNi and 3.3 for Mn.

Interfacial excesses, Ci, are determined from the areaunder the concentration curves as plotted in Fig. 5,employing Eq. (4), the precipitate and matrix concentra-tions given in Table 2, and the atomic density of Fe(85 nm�3). Significant positive relative Gibbsian interfacialexcesses with respect to Fe and Cu, Crelative

i , exist for Ni(1.5 ± 0.4 atoms nm�2) and Mn (1.0 ± 0.3 atoms nm�2).When normalized by the atomic density of a (110) latticeplane of the ferritic a-Fe matrix, 17 atoms nm�2, these val-ues correspond to cumulative projected excesses of0.09 ± 0.02 monolayers (ML) for Ni and 0.06 ± 0.02 MLfor Mn. The normalization of the Gibbsian excess to unitsof ML, however, does not imply that the solute atoms areactually located in one atomic layer. Aluminum has only aslight tendency to segregate at the interface, with an excessof 0.24 ± 0.23 atoms nm�2 (0.01 ± 0.01 ML), and Si, if

there is an interaction with the interface at all, has a slighttendency toward depletion at the interface with a negativeexcess of 0.10 ± 0.11 atoms nm�2 (0.006 ± 0.006 ML). Seg-regation and depletion at an interface are thermodynami-cally both possible, according to the Gibbs adsorptionisotherm.

4. Discussion

While Al segregates only weakly at the a-Fe matrix/precipitate heterophase interface, if at all, it is stronglyenriched in the inner portions of the precipitates,0.5 ± 0.1 at.% Al, compared with a value of only0.047 ± 0.003 at.% Al in the a-Fe matrix. These valuesof matrix and precipitate concentrations, cm and cp,result in a partitioning ratio, K = cp/cm, of K = 10 ± 2.A value of K this large indicates a strong attractive inter-action for Al atoms in bcc Cu. The binary Al–Cu systempossesses a stable bcc Cu(Al) solid solution phase in theconcentration range 18–29.4 at.% Al [40] with a latticeparameter of 0.29564 nm [41]. This indicates that Aladditions stabilize thermodynamically the bcc Cu phase,even though fcc Cu can dissolve up to 19.7 at.% Al[40]. Knowledge of this stabilizing effect of Al on bccCu precipitates may be of technological importance foroptimizing the beneficial effects of bcc Cu precipitatesin steels.

Niobium and C are not present in the entire three-dimen-sional reconstruction displayed in Fig. 2. This is consistentwith the formation of large niobium carbide precipitates,at a number density below 1022 m�3, which is too small avalue to be detected by the random-area analysis techniqueemployed by conventional APT. Niobium carbide precipi-tates are detectable using a selected area analysis approach[42] and by LEAPTM tomography.

Next, we discuss the consequences of limited solid-statediffusion for the kinetics of precipitation and interfacialsegregation. The three-dimensional root-mean-squared(RMS) diffusion distance is

ffiffiffiffiffiffiffiffiffiffiffihx23Di

p¼ ffiffiffiffiffiffiffiffiffiffi

6DXtp

, where DX

is the impurity diffusivity of the solute species at 490 �Cand t is time; for impurity diffusion of atomic species Xin ferromagnetic a-Fe the values are listed in Table 3[43–48]. The three-dimensional RMS diffusion distance cal-culated for Cu, 6.2 nm, is almost identical to half the aver-age surface-to-surface inter-precipitate spacing

hSi ¼ 3

4pNV

� ��1=3

� 1

2hDi

Table 2Average compositions (at.%) of Cu-rich precipitates and ferritic a-Fe matrix as measured by APT

Fe C Si Nb Mn Cu Ni Al

Interior of precipitates 33.0 ± 0.9 ND 0.60 ± 0.14 ND 0.76 ± 0.16 64.0 ± 1.0 1.2 ± 0.2 0.50 ± 0.10Ferrite matrix 96.96 ± 0.02 ND 0.99 ± 0.01 ND 0.55 ± 0.01 0.50 ± 0.01 0.95 ± 0.01 0.047 ± 0.003

The concentration values for the precipitates are based on 3017 atoms, and for the a-Fe matrix on 636,202 atoms; ND, not detected. The error given foreach concentration is rc ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffifficð1� cÞ=Np

, where N is the total number of atoms detected.

846 D. Isheim et al. / Acta Materialia 54 (2006) 841–849

Page 7: Interfacial segregation at Cu-rich precipitates in a high ...arc.nucapt.northwestern.edu/Refbase/Files/IsheimCuPPTActaMat2005.pdfInterfacial segregation at Cu-rich precipitates in

of 6.1 nm. This implies that the diffusion zones of neighbor-ing precipitates overlap and the formation of Cu-rich pre-cipitates has proceeded past the growth stage. Hence, thedecomposition reaction is most likely in the growth andcoarsening regime. The three-dimensional RMS diffusiondistance for Ni, 8.6 nm, is slightly larger, and the onesfor Al (39 nm), Mn (23 nm), and Si (35 nm) are much lar-ger than the diffusion distance of Cu, implying that Ni, Al,Mn, and Si can redistribute on a length scale that is largerthan that for Cu.

For a discussion of a possible kinetically induced, diffu-sion-limited interfacial segregation due to solute pile-up infront of the moving matrix/precipitate interface, we use theone-dimensional RMS diffusion distance,

ffiffiffiffiffiffiffiffiffiffiffihx21Di

p¼ffiffiffiffiffiffiffiffiffiffi

2DXtp

, which provides a conservative estimate for a com-parison with the extent of the interfacial segregation zonemeasured in the proxigram concentration profile becausethe atoms are diffusing in three dimensions, since the prox-igram folds the three-dimensional precipitate geometry intoa one-dimensional profile [30]. The one-dimensional RMSdiffusion distances of the solute elements (see Table 3) is3.6 nm for Cu, 5.0 nm for Ni, 22 nm for Al, 13 nm forMn, and 20 nm for Si. All these diffusion distances are sig-nificantly larger than the spatial extent of the zone of inter-facial segregation of 1–2 nm, and, consequently, solutediffusion in the matrix is sufficient to establish at the leastlocal metastable equilibrium, and thus the solute enrich-ment at the precipitate interfaces is not dominated by a dif-fusion-limited pile-up effect in front of the moving interfaceof a growing precipitate. A thermodynamic driving forcemust therefore exist for the observed interfacial segregation.

For a discussion of a potential elastic interactionbetween solute atoms and the a-Fe matrix/precipitate inter-face, we first estimate the linear lattice parameter misfit, e,between the a-Fe matrix and the bcc Cu-rich precipitates.The lattice parameter for bcc a-Fe at room temperatureis a0 = 0.28665 nm [49], and, with the thermal expansiondata given by [50], a0 = 0.28598 nm at 0 K anda0 = 0.28953 nm at 490 �C. For bcc Cu or Cu-rich bccCu–Fe solid solutions there is, however, to our knowledgeno directly experimentally measured lattice parameter

available. Lattice parameter measurements of mechanicallyalloyed Cu–Fe alloys [51] suggest a positive deviation fromVegard�s rule, with a similar atomic volume for both fccand bcc solid solutions. The lattice parameter values givenin [51] result in a lattice parameter mismatch of 0.6%between bcc a-Fe and bcc Cu–33 at.% Fe. A lattice param-eter a0 equal to 0.2884 nm has recently been calculated forbcc Cu–33 at.% Fe using cluster-expansion generalized-gradient approximation density-functional theory calcula-tions [52], resulting in e = 0.85%, at 0 K. An earlier molec-ular dynamics simulation [53], employing empiricalpotentials, calculated a value of a0 equal to 0.29607 nmfor pure bcc Cu at 0 K, equivalent to the comparativelylarge value of e = 3.2%. Better potentials, however, resultin smaller values, a0 = 0.2885 nm or a0 = 0.2880 nm, corre-sponding to e = 0.9 or 0.7% [53–55], respectively, at 0 K.Assuming that the thermal expansion coefficient of bccCu, which is not known, is not very different from bcc a-Fe, then e is less than about 1% at all temperatures consid-ered. Values of e smaller than 1% are also in agreementwith estimates of the bcc Cu lattice parameter from fccCu (a0 = 0.36146 at room temperature) by simple geomet-ric considerations: Firstly, e = 0.3% results by equating theatomic volumes of bcc and fcc Cu. Secondly, e = �0.1%results by calculating the next-neighbor distance and takinginto account, following Goldschmidt [56], a 3% contractionfor a reduction of the number of next-neighbor atoms from12 to 8. Thus, since the linear lattice parameter mismatch,e, between bcc a-Fe and bcc Cu–33 at.% Fe is about 1% orless and the linear mismatch of Ni or Mn solute atoms in a-Fe is only 1.52 or 1.60%, respectively [57], elastic interac-tions cannot account for the interfacial segregation, as dis-cussed in Ref. [58], and very likely electronic (chemical)effects come into play [61].

Eq. (6) is used for calculating a coefficient of the reduc-tion in interfacial free energy

oroci

����T ;p;l3;...;li�1;liþ1;...;ln

at a solute concentration ci of component i, due to theGibbsian interfacial excess of this solute component. Thevalue for Ni is �1579 mJ m�2 (at.fr.)�1 and for Mn it is�1620 mJ m�2 (at.fr.)�1. These values are more than oneorder of magnitude larger than the value of �114 mJ m�2

(at.fr.)�1 reported for the segregation of Si at grain bound-aries at 550 �C in an Fe–3 at.% Si alloy investigated by one-dimensional atom-probe microscopy [36], but more thanone order of magnitude smaller than the value of�100,000 mJ m�2 (at.fr.)�1 obtained for segregation of Sbto grain boundaries in Fe–0.03 at.% Sb at 350–450 �C byAuger spectroscopy [59] and �119,000 mJ m�2 (at.fr.)�1

measured for the segregation of P to the grain boundariesof an Fe–0.034 at.% P alloy employing a technique basedon equilibrating the contractile forces of thin foils [60].

Approximating the unknown dependence of the interfa-cial free energy on solute segregation by a linear relationship,

Table 3One- and three-dimensional RMS diffusion distances,

ffiffiffiffiffiffiffiffiffiffiffihx21Di

p¼ ffiffiffiffiffiffiffiffiffiffiffi

2DXtp

andffiffiffiffiffiffiffiffiffiffiffihx23Di

p ¼ ffiffiffiffiffiffiffiffiffiffiffi6DXt

p, where DX is the diffusivity of solute species X at

490 �C and t = 100 min, for self-diffusion of Fe, and impurity diffusion ofthe solute elements Cu, Ni, Al, Mn, and Si, in ferromagnetic a-Fe

DX [m2s�1] Ref.ffiffiffiffiffiffiffiffiffiffiffihx21Di

q¼ ffiffiffiffiffiffiffiffiffiffiffi

2DXtp

;t = 100 min [nm]

ffiffiffiffiffiffiffiffiffiffiffihx23Di

q¼ ffiffiffiffiffiffiffiffiffiffiffi

6DXtp

;t = 100 min [nm]

Fe 3.0 · 10�23 [43] 0.6 1.0Cu 1.1 · 10�21 [44] 3.6 6.2Ni 2.0 · 10�21 [45] 5.0 8.6Al 4.1 · 10�20 [46] 22 39Mn 1.5 · 10�20 [47] 13 23Si 3.5 · 10�20 [48] 20 35

DX is calculated from the activation energy and pre-exponential factorprovided by the respective reference listed.

D. Isheim et al. / Acta Materialia 54 (2006) 841–849 847

Page 8: Interfacial segregation at Cu-rich precipitates in a high ...arc.nucapt.northwestern.edu/Refbase/Files/IsheimCuPPTActaMat2005.pdfInterfacial segregation at Cu-rich precipitates in

a value of the total reduction of the interfacial free energydue to the given interfacial excesses is estimated by multi-plying the coefficient of reduction by the solute concentra-tion. With the values given, the interfacial energy of the a-Fe matrix/precipitate heterophase interface is reduced by16 mJ m�2 due to Ni segregation, and by 11 mJ m�2 dueto the segregation of Mn. The sum of these values(27 mJ m�2) represents 10% of the interfacial energy of270 mJ m�2 of bcc Cu precipitates in a binary Fe–1.38 at.% Cu alloy determined from precipitation kinetics[13]. These values for the reduction of interfacial freeenergy are comparable with the 10 mJ m�2 reduction mea-sured by APT for Mg segregation at coherent interfaces ofAl3Sc precipitates in an Al–2.2 Mg–0.12 Sc at.% alloy [61].

5. Summary and conclusions

1. APT was utilized to characterize nanometer-sized Cu-rich precipitates with a metastable bcc structure thatformed in the ferritic a-Fe matrix of a recently developedprecipitation-hardened high-strength low-carbon steel.

2. After direct isothermal aging at 490 �C for 100 min to anear-peak hardness condition of 84 Rockwell G, the pre-cipitates have an average diameter of 2.5 ± 0.3 nm at anumber density of (1.1 ± 0.3) · 1024 m�3.

3. The core regions of the Cu-rich precipitates contain64 ± 1 at.% Cu and 33 ± 1 at.% Fe. Al is enriched inthe precipitates to 0.5 ± 0.1 at.%, which implies a parti-tioning ratio of 10 ± 2 with respect to the0.047 ± 0.003 at.% Al in the a-Fe matrix. Nickel andMn do not partition significantly; and the precipitatesare depleted in Si.

4. Nickel and Mn are segregated at the a-Fe matrix/precip-itate heterophase interface with Gibbsian interfacialexcesses relative to Fe and Cu of 1.5 ± 0.4 nm�2 and1.0 ± 0.3 nm�2, respectively. Aluminum has only aslight tendency toward segregation, and Si is depletedat this heterophase interface.

5. A coefficient of reduction of interfacial energy is calcu-lated directly from the relative interfacial excesses. Thecoefficients are �1579 mJ m�2 (at.fr.)�1 for Ni and�1620 mJ m�2 (at.fr.)�1 for Mn. Total reductions ofthe interfacial free energy of 16 mJ m�2 for Ni and 11mJ m�2 for Mn are estimated. The sum of these energiesis 27 mJ m�2, which amounts to 10% of the interfacialenergy of 270 mJ m�2 [13].

Acknowledgements

The authors thank Dr. Shrikant P. Bhat, Ispat-InlandInc., for the provision of the steel used in this study andthe bulk chemical composition. The atom-probe tomo-graphic experiments were performed in the NorthwesternUniversity Center for Atom-Probe Tomography (NUC-APT). The instruments in this Center were purchased withfunds from the Office of Naval Research and the National

Science Foundation�s instrumentation programs. Financialsupport by the Office of Naval Research, Grant Nos.N000014-03-1-0252 and N00014-99-1-0601, Drs. GeorgeYoder and Julie Christodoulou, program directors, isgratefully acknowledged.

References

[1] Smith CS, Palmer EW. Trans AIME 1933;105:133.[2] Hornbogen E. Trans ASM 1964;57:120.[3] Gagliano MS. Ph.D. Thesis, Northwestern University, 2002.[4] Gagliano MS, Fine ME. Metall Trans A 2004;35A:2323.[5] Goodman SR, Brenner SS, Low JR. Metall Trans 1973;4:2363.[6] Goodman SR, Brenner SS, Low JR. Metall Trans 1973;4:2371.[7] Worall GM, Buswell JT, English CA, Hetherington MG, Smith

GDW. J Nucl Mater 1987;148:107.[8] Osamura K, Okuda H, Asano K, Furusaka M, Kishida K, Kurosawa

F, et al. ISIJ Int 1994;34:346.[9] Pareige PJ, Russell KF, Miller MK. Appl Surf Sci 1996;94/95:362.[10] Miller MK, Russell KF, Pareige P, Starink MJ, Thompson RC.

Mater Sci Eng 1998;A250:49.[11] Murayama N, Sugiyama M, Hara T, Tamehiro H. Mater Trans JIM

1999;40:268.[12] Murayama M, Katayama Y, Hono K. Metall Trans A 1999;30A:

345.[13] Kampmann R, Wagner R. In: Janot C, Petry W, Richter D, Springer

T, editors. Atomic transport and defects in metals by neutronscattering. Berlin: Springer-Verlag; 1986. p. 73.

[14] Osamura K, Okuda H, Takashima M, Asano K, Furusaka M. MaterTrans JIM 1993;34:305.

[15] Monzen R, Jenkins ML, Sutton AP. Philos Mag A 2000;A80:711.[16] Othen PJ, Jenkins ML, Smith GWD, Phythian WJ. Philos Mag Lett

1991;64:383.[17] Othen PJ, Jenkins ML, Smith GWD. Philos Mag A 1994;A70:1.[18] Soisson F, Barbu A, Martin G. Acta Mater 1996;44:3789.[19] Pareige P, Stoller RE, Russell KF, Miller MK. J Nucl Mater

1997;249:165.[20] Auger P, Pareige P, Welzel S, Van Duysen JC. J Nucl Mater

2000;280:331.[21] Pareige P, Miller MK. Appl Surf Sci 1996;94/95:370.[22] Miller MK, Russell KF, Stoller RE, Pareige P. Report NUREG/CR-

6629, US Nuclear Regulatory Commission, 2000.[23] Miller MK. Atom probe tomography. New York, NY: Kluwer

Academic/Plenum; 2000. p. 165.[24] Miller MK, Wirth BD, Odette GR. Mater Sci Eng 2003;A353:133.[25] Fine ME, Ramanathan R, Vaynman S, Bhat SP. In: Proc. Int. Symp.

Low-Carbon Steel of the 90s, Philadelphia (PA); 1993. p. 511.[26] Vaynman S, Fine ME. In: Afshani RI, Bodnar RL, editors. Proc. Int.

Symp. on Steel for Fabricated Structures. Chicago (IL): ASMAnnual Meeting; 1999. p. 59.

[27] Vaynman S, Fine ME, Ghosh G, Bhat SP. In: Chong KP, editor.Materials for the New Millennium: Proceedings of the FourthEngineering Conference, Nov. 10–14, 1996, Washington, DC. NewYork, NY: American Society of Civil Engineers; 1996. p. 1551.

[28] Vaynman S, Fine ME, Bhat SP. Materials Science and Technology2004 Conference Proceedings, Vol. I. New Orleans, LA: AIST andTMS; 2004. p. 417.

[29] Hellman OC, Vandenbroucke J, Blatz du Rivage J, Seidman DN.Mater Sci Eng A 2002;327:29.

[30] Hellman OC, Vandenbroucke JA, Rusing J, Isheim D, Seidman DN.Microsc Microanal 2000;6:437.

[31] Hellman OC, Seidman DN. Mater Sci Eng A 2002;327:24.[32] Hellman OC, Blatz du Rivage J, Seidman DN. Ultramicrosc

2003;95:199.[33] Miller MK. Atom probe tomography. New York, NY: Kluwer

Academic/Plenum; 2000. p. 158.

848 D. Isheim et al. / Acta Materialia 54 (2006) 841–849

Page 9: Interfacial segregation at Cu-rich precipitates in a high ...arc.nucapt.northwestern.edu/Refbase/Files/IsheimCuPPTActaMat2005.pdfInterfacial segregation at Cu-rich precipitates in

[34] Defay R, Prigogine I, Bellemans A, Everett DH. Surface tension andadsorption. New York, NY: Wiley; 1966. p. 89.

[35] Dregia SA, Wynblatt P. Acta Metall Mater 1991;39:771.[36] Krakauer BW, Seidman DN. Acta Mater 1998;46:6145.[37] Seidman DN. Annu Rev Mater Res 2002;32:235.[38] Muller EW, Tsong TT. Field ion microscopy, principles and

applications. New York: Elsevier; 1969. p. 72.[39] Vurpillot F, Bostel A, Blavette D. Appl Phys Lett 2000;76:3127.[40] Massalski TB. Binary alloy phase diagrams. second ed. Materials

Park, OH: ASM International; 1990. p. 141.[41] Pearson WB. A handbook of lattice spacings and structures of metals

and alloys, Vol. 2. Oxford: Pergamon Press; 1976. p. 111.[42] Isheim D. Unpublished research.[43] Mehrer H, Lubbehusen M. Defect Diffusion Forum 1989;66–69:591.[44] Salje G, Feller-Kniepmeier M. J Appl Phys 1977;48:1833.[45] Hirano KI, Cohen M, Averbach DL. Acta Metall 1961;9:440.[46] Akimova IA, Mironov VM, Pokoyev AV. Phys Met Metallogr

(English transl) 1983;56:175.[47] Nohara K, Hirano KI. Trans Iron Steel Inst Jpn 1971(Suppl 11):

1267.

[48] Bergner D, Khaddour Y, Lorx S. Defect Diffusion Forum 1989;66–69:1407.

[49] Massalski TB. Binary alloy phase diagrams. Second ed. MaterialsPark, OH: ASM International; 1990. p. T10.

[50] Touloukian YS, Kirby RK, Taylor RE, Desai PD. Thermal expan-sion. New York, NY: IFI/Plenum; 1975. p. 157.

[51] Ma E, Atzmon M, Pinkerton FE. J Appl Phys 1993;74:955.[52] Liu JZ, van de Walle A, Ghosh G, Asta M. Phys Rev B

2005;72:144109.[53] Harry T, Bacon DJ. Acta Mater 2002;50:195.[54] Osetsky YN, Serra A. Philos Mag A 1996;73:249.[55] Ludwig M, Farkas D, Pedraza D, Schmauder S. Modelling Simul

Mater Sci Eng 1998;6:19.[56] Goldschmidt VM. Z Phys Chem 1928;133:397.[57] King HW. J Mater Sci 1966;1:79.[58] Isheim D, Seidman DN. Surf Interf Anal 2004;36:569.[59] Fine ME, Marcus HL. Metall Trans 1971;2:1473.[60] Hondros ED. Proc R Soc A 1965;286:479.[61] Marquis EA, Seidman DN, Asta M, Woodward CM, Ozolins V. Phys

Rev Lett 2003;91:36101.

D. Isheim et al. / Acta Materialia 54 (2006) 841–849 849