arxiv:1304.0068v1 [cond-mat.mtrl-sci] 30 mar 2013

12
High-temperature neutron diffraction and first-principles study of temperature-dependent crystal structures and atomic vibrations in Ti 3 AlC 2 , Ti 2 AlC, and Ti 5 Al 2 C 3 Nina J. Lane, 1 Sven C. Vogel, 2 El’ad N. Caspi, 1, * and Michel W. Barsoum 1, 1 Department of Materials Science & Engineering, Drexel University, Philadelphia, PA 19104, USA 2 Los Alamos Neutron Science Center, Los Alamos National Laboratory, Los Alamos, NM 87545, USA (Dated: April 2, 2013) Herein we report on the thermal expansions and temperature-dependent crystal structures of select ternary carbide MAX phases in the Ti–Al–C phase diagram in the 100–1000 C temperature range. A bulk sample containing 38(±1) wt.% Ti5Al2C3 (“523”), 32(±1) wt.% Ti2AlC (“211”), 18(±1) wt.% Ti3AlC2 (“312”), and 12(±1) wt.% (Ti0.5Al0.5)Al is studied by Rietveld analysis of high-temperature neutron diffraction data. We also report on the same for a single-phase sample of Ti3AlC2 for comparison. The thermal expansions of all the MAX phases studied are higher in the c direction than in the a direction. The bulk expansion coefficients – 9.3(±0.1) × 10 -6 K -1 for Ti5Al2C3, 9.2(±0.1) × 10 -6 K -1 for Ti2AlC, and 9.0(±0.1) × 10 -6 K -1 for Ti3AlC2 – are comparable within one standard deviation of each other. In Ti5Al2C3, the dimensions of the Ti–C octahedra for the 211-like and 312-like regions are comparable to the Ti–C octahedra in Ti2AlC and Ti3AlC2, respectively. The isotropic mean-squared atomic displacement parameters are highest for the Al atoms in all three phases, and the values predicted from first-principles phonon calculations agree well with those measured. I. INTRODUCTION Titanium carbide is one the hardest transition metal binary carbides known, which renders it resistant to wear and makes it a favorable material for applications such as drill bits and cutting tools. With a melting point of over 3000 C, it is thermally quite stable. However, transition metal binary carbides are brittle, difficult to machine, and highly susceptible to thermal shock. The ternary carbides known as M n+1 AX n (MAX) phases [1] (where n = 1,2, or 3 and M is a transition metal, A is a group A element mostly from groups 13 and 14, and X is either C or N) often overcome some of these shortcomings, while still being heat tolerant. As a class, the MAX phases have unusual – yet attractive and often unique – combinations of properties that bring together some of the best attributes of ceramics and metals [24]. Like metals, they are excellent electric and thermal conductors, with exceptional thermal shock resistance and damage tolerance [2, 3, 5]. In some cases, they are creep [69], oxidation [10, 11], and fatigue [12] resistant. Furthermore, they are elastically quite stiff, yet readily machinable [13]. In this work we are interested in the MAX phases in the Ti–Al–C system. Both Ti 2 AlC [Fig. 1(a)] and Ti 3 AlC 2 [Fig. 1(b)] have been relatively well studied. The crystal structure of Ti 2 AlC was first solved in the 1960s [14]; Ti 3 AlC 2 was discovered several decades later in 1994 [15]. Of the > 60 MAX phases known to date, Ti 2 AlC and Ti 3 AlC 2 are particularly attractive in terms * On sabbatical leave from Nuclear Research Centre - Negev, Beer- Sheva, Israel Michel W. Barsoum: [email protected] of high temperature applications. They are two of the most lightweight and oxidation resistant MAX phases [16, 17], and the accessibility and relative low cost of their raw materials render them the most promising for up-scaling and industrialization. In addition to Ti 2 AlC, henceforth referred to as 211, and Ti 3 AlC 2 , henceforth referred to as 312, there is also a Ti 5 Al 2 C 3 – or 523 – phase, which is in a category of higher- order MAX phases. The latter were first reported in 2004 in the Ti–Si–C system (i.e. Ti 5 Si 2 C 3 and Ti 7 Si 2 C 5 ) by Palmquist et al [18]. Wilhelmsson et al. [19] reported the existence of Ti 5 Al 2 C 3 in 2006. In both cases, these phases were only observed in local stacking sequences in transmission electron microscopy, TEM, micrographs. More recently, the stacking sequence of Ti 5 Al 2 C 3 was characterized using X-ray diffraction, XRD, in two studies [20, 21]. In our study [21], the characterized sample contained 43(±2) wt.% Ti 5 Al 2 C 3 . In another study [20], only a small amount of Ti 5 Al 2 C 3 was observed and neither weight nor volume fractions were reported. We note in passing that the first structure proposed by Wang et al. in Ref. [20] is totally wrong and unsubstantiated by the results shown in that paper (see Ref. [22]). It is also crucial to note that our paper was submitted a few weeks before that of Ref. [20]. Interestingly, Ref. [20] was submitted, reviewed, and accepted in exactly one week and published soon thereafter. It was only after our paper was published that the same group, working with a composition that was only 19.7 wt% Ti 5 Al 2 C 3 , made the case that the space group was R ¯ 3m [23]. Like most other MAX phases, Ti 2 AlC and Ti 3 AlC 2 both have layered hexagonal structures [Figs. 1(a) and 1(b), respectively] belonging to space group P 6 3 /mmc (No. 194). In both structures, Ti–C layers (Ti 2 C for Ti 2 AlC and Ti 3 C 2 for Ti 3 AlC 2 ) are interleaved between layers of Al. The Ti 5 Al 2 C 3 phase, on the other hand [Fig. arXiv:1304.0068v1 [cond-mat.mtrl-sci] 30 Mar 2013

Upload: others

Post on 07-May-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: arXiv:1304.0068v1 [cond-mat.mtrl-sci] 30 Mar 2013

High-temperature neutron diffraction and first-principles study oftemperature-dependent crystal structures and atomic vibrations in Ti3AlC2, Ti2AlC,

and Ti5Al2C3

Nina J. Lane,1 Sven C. Vogel,2 El’ad N. Caspi,1, ∗ and Michel W. Barsoum1, †

1Department of Materials Science & Engineering, Drexel University, Philadelphia, PA 19104, USA2Los Alamos Neutron Science Center, Los Alamos National Laboratory, Los Alamos, NM 87545, USA

(Dated: April 2, 2013)

Herein we report on the thermal expansions and temperature-dependent crystal structures of selectternary carbide MAX phases in the Ti–Al–C phase diagram in the 100–1000◦C temperature range. Abulk sample containing 38(±1) wt.% Ti5Al2C3 (“523”), 32(±1) wt.% Ti2AlC (“211”), 18(±1) wt.%Ti3AlC2 (“312”), and 12(±1) wt.% (Ti0.5Al0.5)Al is studied by Rietveld analysis of high-temperatureneutron diffraction data. We also report on the same for a single-phase sample of Ti3AlC2 forcomparison. The thermal expansions of all the MAX phases studied are higher in the c direction thanin the a direction. The bulk expansion coefficients – 9.3(±0.1) × 10−6 K−1 for Ti5Al2C3, 9.2(±0.1)× 10−6 K−1 for Ti2AlC, and 9.0(±0.1) × 10−6 K−1 for Ti3AlC2 – are comparable within onestandard deviation of each other. In Ti5Al2C3, the dimensions of the Ti–C octahedra for the 211-likeand 312-like regions are comparable to the Ti–C octahedra in Ti2AlC and Ti3AlC2, respectively.The isotropic mean-squared atomic displacement parameters are highest for the Al atoms in allthree phases, and the values predicted from first-principles phonon calculations agree well with thosemeasured.

I. INTRODUCTION

Titanium carbide is one the hardest transition metalbinary carbides known, which renders it resistant to wearand makes it a favorable material for applications suchas drill bits and cutting tools. With a melting pointof over 3000◦C, it is thermally quite stable. However,transition metal binary carbides are brittle, difficult tomachine, and highly susceptible to thermal shock. Theternary carbides known as Mn+1AXn (MAX) phases [1](where n = 1,2, or 3 and M is a transition metal, Ais a group A element mostly from groups 13 and 14,and X is either C or N) often overcome some of theseshortcomings, while still being heat tolerant. As a class,the MAX phases have unusual – yet attractive and oftenunique – combinations of properties that bring togethersome of the best attributes of ceramics and metals [2–4]. Like metals, they are excellent electric and thermalconductors, with exceptional thermal shock resistanceand damage tolerance [2, 3, 5]. In some cases, they arecreep [6–9], oxidation [10, 11], and fatigue [12] resistant.Furthermore, they are elastically quite stiff, yet readilymachinable [13].

In this work we are interested in the MAX phasesin the Ti–Al–C system. Both Ti2AlC [Fig. 1(a)] andTi3AlC2 [Fig. 1(b)] have been relatively well studied.The crystal structure of Ti2AlC was first solved in the1960s [14]; Ti3AlC2 was discovered several decades laterin 1994 [15]. Of the > 60 MAX phases known to date,Ti2AlC and Ti3AlC2 are particularly attractive in terms

∗ On sabbatical leave from Nuclear Research Centre - Negev, Beer-Sheva, Israel† Michel W. Barsoum: [email protected]

of high temperature applications. They are two of themost lightweight and oxidation resistant MAX phases[16, 17], and the accessibility and relative low cost oftheir raw materials render them the most promising forup-scaling and industrialization.

In addition to Ti2AlC, henceforth referred to as 211,and Ti3AlC2, henceforth referred to as 312, there is also aTi5Al2C3 – or 523 – phase, which is in a category of higher-order MAX phases. The latter were first reported in 2004in the Ti–Si–C system (i.e. Ti5Si2C3 and Ti7Si2C5) byPalmquist et al [18]. Wilhelmsson et al. [19] reportedthe existence of Ti5Al2C3 in 2006. In both cases, thesephases were only observed in local stacking sequences intransmission electron microscopy, TEM, micrographs.

More recently, the stacking sequence of Ti5Al2C3 wascharacterized using X-ray diffraction, XRD, in two studies[20, 21]. In our study [21], the characterized samplecontained 43(±2) wt.% Ti5Al2C3. In another study [20],only a small amount of Ti5Al2C3 was observed and neitherweight nor volume fractions were reported. We note inpassing that the first structure proposed by Wang etal. in Ref. [20] is totally wrong and unsubstantiated bythe results shown in that paper (see Ref. [22]). It isalso crucial to note that our paper was submitted a fewweeks before that of Ref. [20]. Interestingly, Ref. [20]was submitted, reviewed, and accepted in exactly oneweek and published soon thereafter. It was only after ourpaper was published that the same group, working witha composition that was only 19.7 wt% Ti5Al2C3, madethe case that the space group was R3m [23].

Like most other MAX phases, Ti2AlC and Ti3AlC2

both have layered hexagonal structures [Figs. 1(a) and1(b), respectively] belonging to space group P63/mmc(No. 194). In both structures, Ti–C layers (Ti2C forTi2AlC and Ti3C2 for Ti3AlC2) are interleaved betweenlayers of Al. The Ti5Al2C3 phase, on the other hand [Fig.

arX

iv:1

304.

0068

v1 [

cond

-mat

.mtr

l-sc

i] 3

0 M

ar 2

013

Page 2: arXiv:1304.0068v1 [cond-mat.mtrl-sci] 30 Mar 2013

2

c

a

c

c

a

a

Al

Al

Al

TiII TiI

TiIII

Ti

TiII

TiI

CI

CII C

C

Ti2AlC

Ti3AlC2

Ti5Al2C3 (a)

(b)

(c)

zTi

zTiI zC

zTiI zC

zTiIII zAl

FIG. 1. Structure of (a) Ti2AlC, (b) Ti3AlC2, and (c)Ti5Al2C3 showing lattice parameters a and c and the uniqueTi, Al, and C lattice sites as constrained by symmetry, alongwith the refined atomic position z parameters.

1(c)], consists of alternating Ti2C (“211-like”) and Ti3C2

(“312-like”) layers interleaved between Al layers. Due tothe shift in stacking sequence, three formula units needto be included in a unit cell. The lattice parameters arethus a=3.064(±0.002) A and c=48.23(±0.02) A [21].

While XRD and TEM have been used to characterizethe room temperature lattice parameters of Ti5Al2C3

and verify its stoichiometry and stacking sequence [21], acomplete structure refinement to determine bond lengthshas not yet been performed. Furthermore, its thermalexpansion coefficients, TECs, have not been measured.Previously, high-temperature neutron diffraction, HTND,was used to investigate the high-temperature stability ofTi2AlC [24, 25] and Ti3AlC2 [24, 26] up to 1550◦C, but inRefs. [24] and [25], only the temperature-dependent phasefractions were studied with no reports of high-temperaturecrystallographic data. In [26], the lattice parameters andTECs were reported for Ti3AlC2, but the bond lengthshave yet to be reported for both Ti2AlC and Ti3AlC2.

The original aim of this work was to determine thecrystal structure parameters of phase-pure Ti2AlC. How-ever, for reasons that are not fully understood, duringprocessing, Ti2AlC decomposed to yield Ti5Al2C3 andsome Ti3AlC2 [21]. And while obtaining the relevantcrystal structure parameters on such a multiphase sampleis not ideal we decided to proceed with this HTND studynevertheless for the following reasons:

(a) To date, neither we nor others have been able to syn-thesize phase-pure Ti5Al2C3, and the results obtainedhere are better than no results;

(b) Given the very similar crystal structures, elastic prop-erties, and TECs among the three phases (see below),the deviations from phase-pure behavior due to thepresence of other phases should be small;

(c) Careful Rietveld analysis can deconovolute the con-tributions from each phase.

These comments notwithstanding, to estimate the error inthe various parameters as a function of temperature, wecompare the results obtained on the multiphase samplewith a predominantly single-phase sample of Ti3AlC2 andshow that the differences, for the most part, are equal to(or less than) the experimental uncertainty.

Herein we use HTND on a sample consisting of 38(±1)wt.% Ti5Al2C3 (“523”), 32(±1) wt.% Ti2AlC (“211”),18(±1) wt.% Ti3AlC2 (“312”), and 12(±1) wt.% of anadditional intermetallic TiAl phase to determine thetemperature-dependent crystal structures of all three Ti–Al–C carbide phases. We report on the temperature evo-lution of the lattice parameters, isotropic thermal atomicdisplacement parameters, ADPs, and bond lengths duringboth heating and cooling for Ti2AlC and Ti3AlC2. Weverify our results by comparing the room temperaturelattice parameters and lattice expansions to previous stud-ies of predominantly single-phase Ti2AlC and Ti3AlC2

samples [15, 27–30]. We also present HTND results for apredominantly single-phase Ti3AlC2 sample for compari-son.

II. EXPERIMENTAL DETAILS

A. Sample synthesis

The sample used herein was prepared using pre-reactedTi2AlC powders that were commercially obtained (Kan-thal, Hallstahammar, Sweden). The powders were coldisostatic pressed, CIPed, at 200MPa and heated at a rateof 300◦C h−1 to 1500◦C, then sintered for 2 h under ahydrogen atmosphere.

The Ti3AlC2 sample was prepared by hot pressing,HPing, pre-reacted Ti2AlC (Kanthal, Hallstahammar,Sweden) and titanium carbide (Alfa Aesar,Ward Hill,MA) in a 1:1 ratio to make 3:1:2 stoichiometry of Ti:Al:C.Powders were ball milled for 24 h, placed in a graphite

Page 3: arXiv:1304.0068v1 [cond-mat.mtrl-sci] 30 Mar 2013

3

(a) Ti2AlC

(b) Ti3AlC2

(c) Ti5Al2C3

(d) Ti2AlC

(e) Ti3AlC2

(f) Ti5Al2C3

0 0 0 4 0 0 0 4 3 0 0 0

100 oC 1000 oC

3 0 0 0

0 0 0 4 3 0 0 0

0 0 0 12 3 0 0 0 0 0 0 12 3 0 0 0

0 0 0 4 3 0 0 0

FIG. 2. Pole figure for 000l and 3000 recalculated from the orientation distribution of HIPPO data for (a) Ti2AlC, (b) Ti3AlC2,and (c) Ti5Al2C3 at 100◦C, and (d) Ti2AlC, (e) Ti3AlC2, and (f) Ti5Al2C3 at 1000◦C. Sample cylinder axis is in the center ofpole figures.

die, and heated in a graphite-heated hot press under avacuum of 10−1 Torr at a rate of 500◦C h−1 to 1400◦C.It was held for 4 h under a pressure of ∼ 40 MPa beforecooling.

In both cases, bulk samples 9 mm in diameter and 3cm high were used for the HTND experiments.

B. High-temperature neutron diffraction

The HTND experiments were conducted on the High-Pressure Preferred Orientation (HIPPO) neutron diffrac-tometer [31–33] at the Lujan Neutron Scattering Center,Los Alamos National Laboratory. For both the multiphaseTi–Al–C sample and Ti3AlC2, bulk samples were placedin a vanadium, V, holder (9 mm diameter, 0.15 mm wallthickness), mounted in an ILL-type high-temperature vac-uum furnace with a V setup, and heated at a rate of 20◦Cmin−1. Data were collected every 100◦C during heatingfrom 100◦C to 1000◦C. For the multiphase sample only,data were collected every 200◦C upon cooling as well. Ateach measurement, the temperature was held constantduring data collection and neutrons were detected with42 panels of 3He detector tubes arranged on five ringswith nominal diffraction angles of 39◦, 60◦, 90◦, 120◦, and144◦. To allow texture analysis, the sample was measuredat rotation angles of 0◦, 22◦, and 45◦ around the verti-cal axis for each temperature, with a count time of 15

min per orientation for a total count time of 45 min pertemperature.

C. Structure refinement

A texture analysis using the Material Analysis UsingDiffraction, MAUD, code [32, 34] showed a mild fibertexture for all temperature points. As the texture wasweak, the data from the three rotations and the detectorrings were integrated, resulting in one histogram perdetector bank. Due to the large detector coverage ofHIPPO, this procedure is similar to spinning the sampleto randomize the preferred orientation. The neutrontime-of-flight data were analyzed as texture-free powderswith the Rietveld method using the General StructureAnalysis System, GSAS [35]. Only the higher resolutiondata from the 90◦, 120◦, and 144◦ banks were used in theanalysis. The MAUD refinements, which incorporatedpreferred orientation, gave parameters that were withinerror bars of those determined without including textureby Rietveld refinement with GSAS. Therefore, all resultsreported herein are from the GSAS refinements assumingrandom texture.

The gsaslanguage refinement script [36] was used toensure that identical refinement strategies were employedfor all temperatures. The instrument alignment (DIFCparameter in GSAS) was fixed for the backscattered (144◦)

Page 4: arXiv:1304.0068v1 [cond-mat.mtrl-sci] 30 Mar 2013

4

1.0 2.0 3.0 4.0

d-spacing (Å)

0

10

20

+ Observed Calculated (Ti0.5Al0.5)Al 523 312 211 Difference

523 (0 0 0 15)

523 (0 0 0 12)

211 (0 0 0 4)

312 (0 0 0 6)

30

0

10

20

523 (0 0 0 15)

211 (0 0 0 4)

312 (0 0 0 6) 523

(0 0 0 12)

(a) (b) N

orm

aliz

ed c

ount

s/μs

1.0 2.0 3.0 4.0

d-spacing (Å)

Nor

mal

ized

cou

nts/μs

FIG. 3. Rietveld analysis of neutron diffraction data measured on HIPPO at 100◦C for muliphase sample from (a) 90◦ detectorbank and (b) 144◦ detector bank. Raw data points are shown as red + symbols and the calculated profile is shown as a solidgreen line. Underneath, markers show calculated peak positions of each phase. From top to bottom: (Ti0.5Al0.5)Al (green), 523(blue), 312 (red), and 211 (black). Difference curve (Yobs − Ycalc) is shown in bottom of each panel as a solid purple line.

detector bank, which has the highest resolution, and re-fined for the 90◦ and 120◦ bank for the lowest temperaturerun at 100◦C. For subsequent runs, DIFC was fixed forall three banks. Refined parameters were 16 backgroundparameters of GSAS background function #1, latticeparameters of all phases, phase fractions, σ1 profile pa-rameter for peak width, atomic positions, and isotropicthermal motion parameters. The atomic positions inTi2AlC [z coordinate of Ti; see Fig. 1(a)] and in Ti3AlC2

[z coordinates of TiI and C; see Fig. 1(b)] were bothrefined in the P63/mmc space group.

For Ti5Al2C3, there are six unique atomic sites: TiI,TiII, and CI with 312-like stacking, TiIII and CII with211-like stacking, and one Al site between 211- and 312-stacked octahedra throughout the cell [see Fig. 1(c)]. Thebest space group to represent these positions was foundto be R3m; therefore, the z coordinates of TiI, TiIII, Al,and CI were refined according to the constraints inducedby the trigonal R3m space group.

In addition to the three MAX phase carbides, peakscorresponding to an intermetallic that crystallizes likeγ-TiAl, with tetragonal space group P4/mmm [37], werefound. Refinement of the site occupancy factor on the Tisite led to 50% Ti and 50% Al antisite defects, yieldinga stoichiometry of (Ti0.5Al0.5)Al. Given the nearly nullscattering intensity of the mixed site, and the low phasefraction of this phase, the isotropic thermal motion pa-rameters were constrained together to reduce the numberof variables.

D. First-principles calculations

First-principles phonon calculations were used to calcu-late the anisotropic mean-squared atomic displacements.

TABLE I. Profile agreement factors rfactors for Rietveld re-finements of neutron diffraction data collected during heatingand cooling for the multiphase Ti–Al–C sample.

T (◦C) wRp (%) χ2 Rexp

100 1.34 3.959 0.67200 1.31 4.042 0.65600 1.23 3.599 0.651000 1.17 3.331 0.64600† 1.29 3.998 0.65200† 1.4 4.697 0.65

† Data collected during cooling.

The DFT calculations were performed using the projector-augmented wave (PAW) method [38], as implementedin the Vienna Ab initio Simulation Package (VASP)code [39–41]. The generalized gradient approximation(GGA) exchange-correlation functional of Perdew-Burke-Ernzerhof (PBE) was used [42] with a cutoff of 500 eV.The total energy was converged to 10−8 eV, with a k-point grid of 12 × 12 × 4 for Ti2AlC and Ti3AlC2 and12 × 12 × 2 for T5Al2C3.

The real-space force constants were calculated usingdensity functional perturbation theory (DFPT) [43] im-plemented in the VASP code. The phonon frequencieswere calculated from the force constants with Phonopy[44], and the ADPs were calculated from the frequenciesand eigenvectors of the force constant matrix. Furtherdetails on the ADP calculations can be found in Ref. [45].

Page 5: arXiv:1304.0068v1 [cond-mat.mtrl-sci] 30 Mar 2013

5

10111213

0 200 400 600 800 1000 1200Temperature (°C)

36373839

17181920303132

Ti5Al2C3

Ti2AlC

Ti3AlC2

(Ti0.5Al0.5)Al

Com

posi

tion

(wt.

%)

FIG. 4. Compositions (wt.%) of each phase in the Ti–Al–Csample as a function of temperature upon heating and cooling.Dashed lines indicate limits for ±1% range.

III. RESULTS

As noted above, texture analysis (Fig. 2) showed amild (000l) fiber texture for all three MAX phases. Notsurprisingly, the texture did not change during heatingor cooling [compare Figs. 2(a), (b), (c) to Figs. 2(d),(e), (f), respectively]. The Rietveld fits for the neutrontime-of-flight data – integrated for the full detector ringsand the three measured orientations – are shown for thelowest temperature run at 100◦C for the 90◦ [Fig. 3(a)]and the 144◦ detector banks [Fig. 3(b)]. The calculated fit(solid green lines) and measured data (red plus signs) arecompared, with the difference curve plotted at the bottom(solid purple line). The markers above the differencecurve show the peak positions for the phases: from topto bottom, (Ti0.5Al0.5)Al (green), 523 (blue), 312 (red),and 211 (black). The higher-d-spacing peaks resultingfrom diffraction by the basal planes are labeled for the(0 0 0 6) peak of Ti3AlC2, the (0 0 0 4) peak of Ti2AlC,and the (0 0 0 15) and (0 0 0 12) peaks of Ti5Al2C3. Thelatter two peaks unambiguously identify Ti5Al2C3 as thedominant phase, as they cannot be accounted for by anyother known phase in the Ti–Al–C system.

The profile agreement factors for the Rietveld fitsare listed in Table I at select temperatures, giving theweighted profile R index, wRp, the goodness of fit, χ2,and the expected R factor, Rexp [46]. Good agreementis observed between the calculated and observed profiles,with one unidentified peak at 2.09 A whose origin remainsunclear. A broad peak at 3.80 A is present only in the90◦ bank [Fig. 3(a)], which is likely due to backgroundinterference.

The composition determined from Rietveld analysis is38(±1) wt. % Ti5Al2C3, 32(±1) wt.% Ti2AlC, 18(±1)

wt. % Ti3AlC2, and 12(±1) wt.% (Ti0.5Al0.5)Al. Thetemperature dependencies of the fractions of each phaseare plotted in Fig. 4, where the dashed lines indicate the±1 wt.% limits. The compositions generally stay within1 wt.% of the average value during heating and cooling,lending credibility to our data analysis and the resultinguncertainties for the compositions.

The temperature-dependent expansions of the lat-tice parameters and interatomic distances in Ti5Al2C3,Ti3AlC2, and Ti2AlC are shown in Figs. 6(a), (b), and(c), respectively. In Fig. 6(b), the parameters for the pre-dominantly single-phase Ti3AlC2 sample are also plottedas open symbols. The anisotropic TEC values are listedin Table II, along with those from previous studies onTi2AlC [28] and Ti3AlC2 [26, 29, 30].

To further compare the thermal expansions of the threephases, the temperature dependencies of (∆V/Vo)1/3 –where ∆V is the change in unit cell at temperature Tas compared to that at the reference temperature, 25◦C(extrapolated), V0 – are plotted in Fig 6(d). The resultsfor (Ti0.5Al0.5)Al are also shown. The slope of theselines yields the average thermal expansion αav. Fromthese results we find that for Ti2AlC, αav = 9.2(±0.1)× 10−6 K−1; for Ti3AlC2, αav = 9.0(±0.1) × 10−6 K−1;for Ti5Al2C3 αav = 9.3(±0.1) × 10−6 K−1. It is thusclear from Fig. 6(d) that the TECs of the three MAXphases are almost identical within the error bars.

The absolute values of the c and a lattice parameters[Figs. 5(a) and (b), Table III] are also comparable, but it isapparent that Ti3AlC2 has the highest a lattice parameterand Ti2AlC has the lowest, with that of Ti5Al2C3 fallingin between. The same is true of the c lattice parameters,after normalizing it by three to account for the threeformula units in Ti5Al2C3 [Fig. 5(b)].

The temperature dependences of the absolute valuesof the Ti–C and Al–Ti bonds are shown in Figs. 5(c)and (d), respectively. The bonds in 211, 312, and 523 areshown in black, red, and blue, respectively. Note that theabsolute range for the scale is the same for the graphsshown in Figs. 5(c) and 5(d). The extrapolated roomtemperature values for all bonds are also listed in TableIV, along with their expansion rates.

Due to the overlap of peaks in our data set we wereunable to determine the anisotropic displacements as wasdone in our previous HTND studies [45, 49–52]. Instead,we examine the isotropic ADPs, Uiso, which representthe mean-squared displacements of the atoms from theirequilibrium positions. Figures 7(a), (b), and (c) show thetemperature dependences of Uiso for the unique Ti (red),Al (blue), and C (black) atoms in Ti5Al2C3, Ti3AlC2, andTi2AlC, respectively. In Fig. 7(b) the values for predomi-nantly single-phase Ti3AlC2 are shown for comparison’ssake.

The values calculated with first-principles phonon cal-culations are shown as lines. In both experimental andcalculated results, the Al atom shows the highest am-plitude. Figure 8 compares the experimental and calcu-lated values of Uiso for the Al atoms in all three phases

Page 6: arXiv:1304.0068v1 [cond-mat.mtrl-sci] 30 Mar 2013

6

13

14

15

16

17

18

19

0 200 400 600 800 10001200

c (Å

)

Temperature (°C)

(a) (b)

(c) (d)

3.05

3.06

3.07

3.08

3.09

3.1

0 200 400 600 800 10001200

a (Å

)

Temperature (°C)

Ti3AlC2

Ti5Al2C3

Ti2AlC

2.08

2.1

2.12

2.14

2.16

2.18

2.2

0 200 400 600 800 10001200

r Ti-C

(Å)

Temperature ( °C)

TiII-C (312)

TiII-CI (523)

TiIII-CII (523)

Ti-C (211)

TiI-C (312) TiII-CI (523)

Ti3AlC2 (c)

Ti5Al2C3 (1/3 c)

Ti2AlC (c)

2.84

2.86

2.88

2.9

2.92

2.94

2.96

2.98

0 200 400 600 800 10001200

Al-Ti (211)Al-Ti

I (312)

Al-TiI (523)

Al-TiIII

(523)

Temperature ( °C)

r Al-T

i (Å)

FIG. 5. Temperature dependence of (a) a and (b) c lattice parameters, and (c) Ti–C and (d) Ti–Al interatomic distancesin Ti5Al2C3 (blue), Ti3AlC2 (red) and Ti2AlC (black). Refer to Fig. 1 for notations of Ti and C atoms. Errors are typicallysmaller than symbol size.

Ti5Al2C3, Ti3AlC2 and Ti2AlC in the multiphase sam-ple, along with Uiso for pure Ti3AlC2. Also shown arethe experimental Ueq values for the Al-containing phasesTi2AlN and Ti4AlN3, determined from in previous HTNDstudies [50, 52], along with calculated values for thosephases from Ref. [45]. The Uiso values for Al in the threeTi–Al–C phases are similar to each other, and slightlyhigher than Ueq of for Al in the Ti–Al–N phases, which isconsistent with the values determined from first-principlescalculations. Note that the calculated Uiso curves for Alin Ti5Al2C3, Ti2AlC, and Ti3AlC2 in Fig. 7 lie essentiallyon top of one another, and those for Al in Ti2AlN and

Ti4AlN3 are similar as well, but smaller than those of thecarbide phases.

IV. DISCUSSION

A. Lattice parameters, expansions, andanisotropies

Not surprising, the measured a- and scaled c-parametersof the 523 phase are in between those of the 211 and 312phases [Fig. 5(b)]. This is also consistent with the val-

Page 7: arXiv:1304.0068v1 [cond-mat.mtrl-sci] 30 Mar 2013

7

TABLE II. Thermal expansions from HTND for Ti–Al–C phases in the sample studied in this work, along with those fromother studies [26, 28–30, 47, 48] determined through HTXRD and dilatometry. Numbers in parentheses are estimated standarddeviations in the last significant digit of the refined parameters.

Phase αa αc αav Anisotropy Ref.(10−6 K−1) (10−6 K−1) (10−6 K−1) (αc/αa)

Ti5Al2C3 9.1(1)† 9.8(1)† 9.3(1)† 1.08(3)† This workTi2AlC 9.0(1)† 9.6(1)† 9.2(1)† 1.07(3)† This work

7.1(3)a 10.0(5)a 8.1(5)a,b; 8.2(2)c 1.41(4) Ref. [28]Ti3AlC2 8.6(1)† 9.7(1)† 9.0(1)† 1.13(3)† This work

7.6(1) 9.0(1) 8.1(1) 1.18(3) This work8.3(1)a 11.1(1)a 9.2(1)a,b; 7.9(5)c 1.33(1)a Ref. [29]- - 9.0(2)c - Ref. [30]8.5 10.2 9.2b 1.2 Ref. [26]

(Ti0.5Al0.5)Al 10.7(1)† 11.5(1)† 11.0(1) † 1.074(3)† This workγ-TiAl - - 10.0 - Ref. [47]

9.77 9.26 - - Ref. [48]

† Multiphase sample.a High-temperature XRD.b Assuming αav = (2/3αa + 1/3αc) = V

−1/30 dV 1/3/dT .

c Dilatometry.

ues from our first-principles calculations (Table III). Ourlattice parameters are in good agreement with literaturevalues for Ti2AlC [21, 27, 28], Ti5Al2C3 [21], and Ti3AlC2

[15, 21]. Their order of increase is likely due to intricaciesin charge transfer involved in bonding. The fact that thea lattice parameter scales with the number of Ti–C bondsis consistent with first principles calculations [21].

The overall expansions of the three MAX phases inthe multiphase sample are, within their error bars, nearlyequivalent (Table II). The a and c lattice expansions arequalitatively comparable, with the a lattice parametersand their thermal expansions all within 1% of eachother [Fig. 5(a)]. Consistent with previous studies, theexpansion in the c direction is greater than along thea. However, for reasons discussed below, in the presentstudy the degree of anisotropy is lower than in Refs.[28–30] (Table II). We now consider each of the phasesseparately.

Ti2AlC: The TECs along the a- and c-direction forthe Ti2AlC sample measured herein – 9.6(±0.1) × 10−6

K−1and 8.9(±0.1) × 10−6 K−1respectively [Fig. 6(c)] –fall in between those reported previously for Ti2AlC (Ref.[28]). The reason(s) for the discrepancy is unknown at thistime but could very well reflect differences in chemistry.Recent work in the literature suggests that Ti2AlC existsover a range of stoichiometries. For example, Bai et al.recently reported the existence of a Ti2AlCx phase wherex was as low as 0.69 [53, 54]. Herein, it is more likely thannot that the Ti2AlC is Al-deficient since it is believedthat the loss of Al is what triggers the transformation tothe 523 and possibly the 312 phase.

Ti3AlC2: The TEC values measured herein for the 312phase depended on sample. The predominantly single-

phase Ti3AlC2 sample has a lower expansion in bothdirections, resulting in a statistically significant lower αav

of 8.1(±0.1) × 10−6 K−1(Fig. 6). At 9.0(±0.1) × 10−6

K−1, αav for the 312 phase in the multiphase sample isabout 10% higher than in the single phase one.Ti5Al2C3: Since this is the first report on the

effect of temperature on the lattice parameters of the523 phase, there are no previous results to comparethem with. However, the fact that αav of this phaseis very comparable to the 211 and 312 phases is notsurprising and is consistent with the fact that the for-mer is comprised of the same building blocks as the latter.

Lastly, a few remarks on the expansions. The abil-ity to measure phase sensitive TECs is an advantage ofHTND, as compared to other methods such as dilatom-etry, that require pure phases to measure their volumeTECs. However, it is important to appreciate that theTEC values measured herein per force are less anisotropicthan those one would measure in loose powders. In thelatter case, the solid is free to expand, whereas when themeasurement is made on bulk solids, residual stresses canaccrue and reduce the values of the thermal expansionsin various directions. The effect is best appreciated whenthe TECs in the a and c directions are compared withthose measured on powder Ti2AlC samples [29].

B. Bond lengths

While the overall expansions and anisotropies in thethree MAX-like phases are comparable, the most inter-esting aspect of this work is the relationship betweenbond length evolution and the stacking of the octahedra

Page 8: arXiv:1304.0068v1 [cond-mat.mtrl-sci] 30 Mar 2013

8

0

0.25

0.5

0.75

1

1.25

1.5

1.75

0 200 400 600 800 1000 1200

acAl-Ti

ITi

I-C

TiII-C

Temperature (°C)

∆L/L

0 x 1

00 (%

)

multiphase (m) pure (p)

αa (m) = 8.6(1) x 10-6 K-1

αc (m) = 9.7(1) x 10-6 K-1

αa (p) = 7.6(1) x 10-6 K-1

αc (p) = 9.0(1) x 10-6 K-1

0

0.25

0.5

0.75

1

1.25

1.5

1.75

0 200 400 600 800 1000 1200

acAl-TiTi-C

Temperature (°C)

αa=8.9(1) x 10-6 K-1

αc=9.6(1) x 10-6 K-1

∆L/L

0 x 1

00 (%

)

(a) Ti5Al2C3 (b) Ti3AlC2

(c) Ti2AlC (d)

0

0.25

0.5

0.75

1

1.25

1.5

1.75

0 200 400 600 800 1000 1200

acAl-Ti

IAl-Ti

IIITi

I-C

ITi

II-C

ITi

III-C

II

∆L/L

0 x 1

00 (%

)

Temperature (°C)

αa=9.1(1) x 10-6 K-1

αc=9.8(1) x 10-6 K-1

0

0.2

0.4

0.6

0.8

1

1.2

0 200 400 600 800 1000 1200

Ti5Al

2C

3Ti

3AlC

2Ti

2AlC

Ti3AlC

2 (pure)

(Ti0.5

Al0.5

)Al

Temperature (°C)

∆(V

1/3 )/V

01/3 x

100

(%)

α(Ti5Al

2C

3) = 9.3(1) x 10-6 K-1

α(Ti2AlC) = 9.2(1) x 10-6 K-1

α(Ti3AlC

2) = 9.0(1) x 10-6 K-1

= 8.1 x 10-6 K-1 (pure)α[(Ti

0.5Al

0.5)Al] = 11.0(1) x 10-6 K-1

FIG. 6. Temperature dependences of the thermal strains of the lattice parameters and interatomic distances in (a) Ti5Al2C3, (b)Ti3AlC2, and (c) Ti2AlC for the multiphase sample. In (b), the results for predominantly single-phase Ti3AlC2 are also shown.(d) Volume expansion of Ti5Al2C3 (blue circles), Ti3AlC2 (red solid squares), Ti2AlC (black diamonds), and (Ti0.5Al0.5)Al(green triangles) in the multiphase sample, and pure Ti3AlC2 (open red squares). Room temperature (25◦C) values wereextrapolated to use for L0 and V0. Errors for volume expansions in (d) are typically smaller than symbol size.

among the three phases. In the literature, it is fairly wellestablished, both experimentally and theoretically, thatthe M–C bonds adjacent to the A layers (i.e. TiI-C inFig. 1) are shorter than those in the stoichiometric binaryMX, while the ones that are not, (viz. TiII–C, in Fig.1) are longer. Figure 5(c) and the Ti–C lengths in TableIV are fully consistent with this general conclusion. Notsurprisingly, the Ti–C bond lengths in the 211 slab inthe 523 phase are almost identical to those of the 211phase [Fig. 5(c)]. Similarly, the Ti–C bond lengths inthe 312 phase are very similar to those of the 312 slabsin the 523 phase [Fig. 5(c)]. This applies not only tothe absolute Ti–C bond lengths values but also to their

thermal expansions, which are quite comparable as well[Fig. 1(c)]. It should be noted that the longest bondsin the 312-stacked octahedra in Ti5Al2C3 are slightlyshorter than rTiII−C in Ti3AlC2, while the Ti–C bonds inthe 211-stacked octahedra in Ti5Al2C3 are slightly longerthan those in Ti2AlC [Fig. 5(c)]. This suggests that thestructure is slightly more uniform than the individual 211-and 312-stacked phases due to the interleaved nature ofthe stacking sequences. These comments notwithstanding,it is clear from Fig. 5(c) that the same structural unitsbehave similarly. These results are gratifying becausethey indirectly validate our Rietveld analysis.

The situation for the Ti-Al bonds is not as clear. Since

Page 9: arXiv:1304.0068v1 [cond-mat.mtrl-sci] 30 Mar 2013

9

0 200 400 600 800 1000 1200 1400

TiAlC

Temperature (K)

exp calc

(a) (b) (c)

Ti5Al2C3 Ti3AlC2 Ti2AlC

0

1

2

3

4

5

0 200 400 600 800 1000 1200

TiI

TiII

TiIII

AlC

IC

II

U x

100

(Å2 )

Temperature (K)

exp calc

0 200 400 600 800 1000 1200

TiI

TiII

AlC

Temperature (K)

exp pure calc

FIG. 7. Temperature dependence of isotropic thermal displacement parameter Uiso during heating and cooling for atoms in (a)Ti5Al2C3, (b) Ti3AlC2, and (c) Ti2AlC. Errors are typically smaller than symbol size.

the difference between the 211 and 312 phases is the num-ber of Ti–C octahedra between Al layers, it is expectedthat only the Ti–C bonds would be significantly affectedby the change in Ti–C stacking while the Al–Ti bondsshould be similar among the three phases. However, wefind that the Al–Ti bonds are clearly affected by stoi-chiometry [Fig. 5(d)]: rTi−Al is significantly longer in the211 phase [black diamonds in Fig. 5(d)] than in the 312phase [red crosses in Fig. 5(d)], while the opposite is trueof those bonds in the 523 phase; i.e., rAl−TiI > rAl−TiIII

in Ti5Al2C3 [compare blue circles and blue squares in5(d)]. The reasons for this state of affairs are not fullyunderstood, but are likely related to the following obser-vations:

(i) The Ti–C bonds are relatively stiff building blocksof the individual 312 and 211 units, as evidenced bythe fact that they stay relatively the same size as inthe original Ti3AlC2 and Ti2AlC phases when thestacking sequences are interleaved. The dimensionwithin the crystal that thus has the most flexibilityis the Al–Ti bond. Therefore, it is likely that the Al–Ti bond plays a role as an effective “compensatingspring” in the structure to minimize the crystalenergy. Furthermore, this role would be different –and most probably more dominant – in the morecomplex Ti5Al2C3 higher-order phase.

(ii) Among the possible factors that could be compen-sated for in the flexible Al–Ti bond discussed aboveare those related to constraints on the lattice param-eters – especially on the c-lattice parameter, whichessentially determines the Al–Ti bond length, giventhat the Ti–C octahedra are rigid blocks. In a sam-ple with multiple competing phases, it is likely thatthese effects are prominent and manifest themselves

in the Al–Ti bond.

(iii) The Al atoms in Ti2AlC and Ti3AlC2 lie in a mir-ror plane within the structures, while Al is notconstrained to mirror symmetry between the Ti–Catoms in the Ti5Al2C3 phase (see Fig. 1). Therefore,the changes in the Al–Ti distances in the 211-stackedand 312-stacked structures that occur when they areinterleaved to form 523 may be a consequence of thesymmetry break.

These comments notwithstanding, it is important tonote that the average Ti–Al bond length in 523 (2.863 A)is equal the average of the Al–Ti bond lengths in 312 and211 (also 2.863 A). The average expansions of those bondsare also similar (see below). While more work is neededto fully understand the Al–Ti bond length behavior, itcan be reasonably concluded that the dimensions of theTi–C units are consistent for a given stacking, regardlessof whether they are interleaved in a higher-order phaseor in a conventional MAX phase. Based on this fact andthe inconsistency of the Al–Ti bonds, it is further spec-ulated that the Al–Ti bonds serve to compensate otherenergy minimization factors for the crystal, especiallythose related to symmetry and lattice constraints.

C. Bond expansions

In Ti2AlC and Ti3AlC2, the Al–Ti bonds show thehighest expansion [see Table II and Figs. 6(b) and6(c)]. To our knowledge, there are no previous reports oftemperature-dependent bond lengths in any of the Ti–Al–C MAX phases to which to compare our results. However,in a previous HTND study of the nitride Ti2AlN, theAl–Ti bond also showed a higher expansion rate than the

Page 10: arXiv:1304.0068v1 [cond-mat.mtrl-sci] 30 Mar 2013

10

TABLE III. Temperature-dependent a and c lattice parameters from Rietveld refinement of neutron diffraction data collectedduring heating and cooling. Numbers in parentheses are estimated standard deviations in the last significant figure of the refinedparameters. Room temperature values are extrapolated to 25◦C from linear interpolation.

Ti5Al2C3 Ti2AlC Ti3AlC2

Temp. (◦C) a (A) c (A) a (A) c (A) a (A) c (A)RT (Ref.[15]) - - - - 3.0753 18.578RT (Ref. [27]) - - 3.065(4) 13.71(3) - -RT (Ref. [28]) - - 3.051 13.637 - -RT (Ref. [21]) 3.064(2) 48.23(2) 3.063 13.645 3.060 18.66RT (Ref. [21])a 3.068 48.186 3.067 13.75 3.083 18.661RTb 3.05678 48.189 3.05405 13.6422 3.06186 18.4994100 3.05926(6) 48.237(1) 3.05656(7) 13.6551(5) 3.06424(8) 18.5172(8)200 3.06184(6) 48.280(1) 3.05912(7) 13.6669(4) 3.06667(8) 18.5331(8)300 3.06419(6) 48.317(1) 3.06141(7) 13.6777(5) 3.06887(8) 18.5483(8)400 3.06642(6) 48.354(1) 3.06353(7) 13.6879(4) 3.07108(8) 18.5617(8)500 3.06911(6) 48.400(1) 3.06627(7) 13.7006(4) 3.07359(8) 18.5803(8)600 3.07208(6) 48.449(1) 3.06920(7) 13.7145(4) 3.07635(8) 18.5991(8)700 3.07501(6) 48.497(1) 3.07206(7) 13.7280(4) 3.07908(8) 18.6184(8)800 3.07814(6) 48.554(1) 3.07510(7) 13.7431(5) 3.08213(8) 18.6380(8)900 3.08141(6) 48.608(2) 3.07825(7) 13.7581(5) 3.08519(8) 18.6587(9)1000 3.08460(6) 48.667(2) 3.08126(7) 13.7743(5) 3.08818(8) 18.6811(9)800c 3.07876(6) 48.559(2) 3.07552(7) 13.7453(5) 3.08264(9) 18.6404(9)600c 3.07301(6) 48.459(2) 3.06978(7) 13.7180(5) 3.07730(9) 18.6025(9)400c 3.06770(6) 48.368(2) 3.06460(7) 13.6922(5) 3.07200(9) 18.5684(9)200c 3.06239(6) 48.279(2) 3.05928(7) 13.6672(5) 3.06712(9) 18.5337(9)

a DFT calculations.b Extrapolated value.c Data collected during cooling.

Ti–N bond [50]. Similarly, a higher expansion rate wasobserved for the A–M bonds than the M–C bonds inTi3SiC2 [49] and Cr2GeC [50]. This result is also consis-tent with a high-pressure XRD study of Ti3AlC2, wherethe Al–TiI bond was the most compressible, while theTiI–C and TiII–C bonds were more rigid [55].

In Ti5AlC2, the expansion rate of the Al–TiI bond –14.6 × 10−6 K−1– is the highest of all the bonds in thesample, but the Al–TiIII bond expansion is unexpect-edly low, at 5.5 × 10−6 K−1 (Table IV). Nonetheless,the average bond expansion in 523 (10.5 × 10−6 K−1)is still similar to the 211 and 312 average (11.3 × 10−6

K−1). Also note that the error bars for the Al–Ti bondexpansions in Ti5Al2C3 are the highest of those for thebonds in all phases (see Fig. 6). This uncertainty fur-ther suggests that the Al–Ti bond behavior is flexiblewithin the structure and indicates other crystal imperfec-tions and/or symmetry and lattice dimension effects, asdiscussed above.

D. Atomic displacement parameters

The results in Fig. 7 show that, like all other MAXphases studied to date, the A atom – Al in this case – isa rattler in that it vibrates with a significantly higher am-

plitude than the other atoms in the structures. The highADPs of Al, both calculated and experimental, relative tothe Ti and C atomic displacement values [Figs. 7(a)-(c)]are also consistent with the relatively weaker Al bondingevidenced by the higher Al–Ti expansion rates, at leastin Ti2AlC and Ti3AlC2 (Table IV) and the flexibility ofthe Al–Ti interaction, as discussed above.

Previous HTND studies of Ti3SiC2 [49, 51], Ti3GeC2

[49], Ti2AlN [50], Cr2GeC [50], and Ti4AlN3 [52] haveshown the same rattling phenomenon for the A-groupelement. A comparison of the vibrational behavior of Alwith two other HTND studies of the Al-containing nitridesTi2AlN and Ti4AlN3 (Fig. 8) further suggests that this“rattling” effect is independent of stacking sequence.

The results shown in Figs. 7 and 8 also clearly indicatethat from a theoretical point of view, the ADPs of thethree MAX phases should be very comparable. Giventhat the Al atoms in Ti–Al–N nitrides are also predictedto behave similarly to one another – but different fromthe carbides – in their vibrational amplitudes (see linesfor Ti2AlN and Ti4AlN3 in Fig. 8), the DFT calculationsindicate that the ADPs of Al should not be greatly in-fluenced by stoichiometry. Interestingly, the agreementbetween theoretical and experimental isotropic ADPs forthe Al atoms is quite good in all five compounds plottedin 8. The agreement for the other atoms is less good forreasons that are not well understood, but are typical of

Page 11: arXiv:1304.0068v1 [cond-mat.mtrl-sci] 30 Mar 2013

11

TABLE IV. Interatomic distances in Ti5AlC2, Ti2AlC, andTi3AlC2 in the multiphase sample from Rietveld refinement ofneutron diffraction data collected during heating and cooling,along with their expansions. All values are extrapolated to25◦C from linear interpolation. Numbers in parentheses areestimated standard deviations in the last significant digit ofthe refined parameters.

Phase Bond Bond length Bond expansion†

(A) (10−6 K−1)Ti5Al2C3 Al–TiI 2.878 14.6

Al–TiIII 2.848 5.5TiI–CI 2.079 8.8TiII–CI 2.180 8.1TiIII–CIII 2.106 10.0

Ti2AlC Al–Ti 2.873 10.8Ti–C 2.101 7.7

Ti3AlC2 Al–TiI 2.854 11.8TiI–C 2.087 6.8TiII–C 2.182 8.8

† Bond expansion: L−1 · dL/dT from least-squares fit of∆L/L0 vs. T .

the MAX phases [45].

V. SUMMARY AND CONCLUSIONS

Rietveld analysis of HTND data for a bulk samplecontaining Ti5Al2C3 [38(±1) wt.%], Ti2AlC [32(±1) wt.%], Ti3AlC2 [18(±1) wt. %], and (Ti0.5Al0.5)Al [12(±1)wt.%] has shown that Ti5Al2C3 exhibits similar thermalexpansion and thermal motion parameters as Ti2AlC andTi3AlC2. The thermal expansions for Ti5Al2C3 in the a-and c- directions, respectively, are αa = 9.1(±0.1) × 10−6

K−1and αc = 9.8(±0.1) × 10−6 K−1. In all three phases,the average expansion rates of all the Al–Ti bonds arehigher than the average Ti–C bond expansions. Ti5Al2C3

consists of alternating layers of 312- and 211-like stack-ing, where the 312 layers are similar to Ti3AlC2 and the211 layers are similar to Ti2AlC in dimensions and bondexpansions. The Al atoms in all three phases vibratewith higher amplitudes than the Ti and C atoms. Thiswork shows that Ti5Al2C3 exhibits similar properties toTi3AlC2 and Ti2AlC, two of the most promising MAXphases, which indicates that phase purity can be morerelaxed in processing when considering applications. Inaddition, this work shows that further studies on Ti5Al2C3

can lead to enhanced property optimization and engineer-ing for ternary carbides in the Ti–Al–C system.

ACKNOWLEDGMENTS

This work has benefited from the use of the LujanNeutron Scattering Center at LANSCE, which is funded

0

1

2

3

4

5

0 200 400 600 800 1000 1200 1400

U x

100

(Å2 )

Temperature (K)

Ti5Al2C3 Ti2AlC Ti3AlC2 Ti2AlN Ti4AlN3

exp calc mult. pure

FIG. 8. Temperature dependence of isotropic thermal motionUiso of Al in Ti5Al2C3 (blue circles), Ti2AlC (black diamonds),and Ti3AlC2 (red solid squares) in the multiphase (mult.)sample, and in pure in Ti3AlC2 (red open squares) determinedfrom HTND. Ueq of Al in predominantly single-phase Ti2AlN(green half-filled squares) and Ti4AlN3 (purple triangles) arealso shown, determined from HTND in Refs. [50] and [52],respectively. The lines refer to values calculated from first-principles phonon calculations, where the calculated values forTi2AlN and Ti4AlN3 are from Ref. [45]. Errors are typicallysmaller than symbol size.

by the U.S. Department of Energy’s Office of Basic EnergySciences. Los Alamos National Laboratory is operated byLos Alamos National Security LLC under DOE contractDE-AC52-06NA25396. This work was also supported bythe Army Research Office (W911NF-07-1- 0628). Theauthors would also like to thank D. J. Tallman and Dr.T. El-Raghy for providing the samples used in this work.

[1] M. W. Barsoum, Progress in Solid State Chemistry 28,201 (2000).

[2] M. W. Barsoum and T. El-Raghy, Journal of the AmericanCeramic Society 79, 1953 (1996).

[3] T. ElRaghy, A. Zavaliangos, M. W. Barsoum, and S. R.Kalidindi, Journal of the American Ceramic Society 80,513 (1997).

[4] M. W. Barsoum and T. El-Raghy, Journal of MaterialsSynthesis and Processing 5, 197 (1997).

Page 12: arXiv:1304.0068v1 [cond-mat.mtrl-sci] 30 Mar 2013

12

[5] A. Ganguly, T. Zhen, and M. W. Barsoum, Journal ofAlloys and Compounds 376, 287 (2004).

[6] D. J. Tallman, M. Naguib, B. Anasori, and M. W. Bar-soum, Scripta Materialia 66, 805 (2012).

[7] M. Radovic, M. W. Barsoum, T. El-Raghy, and S. Wieder-horn, Acta Materialia 49, 4103 (2001).

[8] M. Radovic, M. W. Barsoum, T. El-Raghy, and S. M.Wiederhorn, .

[9] T. Zhen, M. W. Barsoum, S. R. Kalidindi, M. Radovic,Z. M. Sun, and T. El-Raghy, .

[10] M. W. Barsoum, T. El-Raghy, and L. U. J. T. Ogbuji,Journal of the Electrochemical Society 144, 2508 (1997).

[11] M. Sundberg, G. Malmqvist, A. Magnusson, and T. El-Raghy, Ceramics International 30, 1899 (2004).

[12] Journal of the European Ceramic Society 32, 1813 (2012).[13] M. W. Barsoum and T. El-Raghy, American Scientist 89,

334 (2001).[14] W. Jeitschko, H. Nowotny, and F. Benesovsky, Monat-

shefte fur Chemie 94, 672 (1963).[15] M. A. Pietzka and J. C. Schuster, Journal of the American

Ceramic Society 79, 2321 (1996).[16] X. H. Wang and Y. C. Zhou, Corrosion Science 45, 891

(2003).[17] X. Wang and Y. Zhou, Oxidation of Metals 59, 303 (2003).[18] J. P. Palmquist, S. Li, P. O. A. Persson, J. Emmerlich,

O. Wilhelmsson, H. Hogberg, M. I. Katsnelson, B. Johans-son, R. Ahuja, O. Eriksson, L. Hultman, and U. Jansson,Physical Review B 70, 165401 (2004).

[19] O. Wilhelmsson, J. P. Palmquist, E. Lewin, J. Emmerlich,P. Eklund, P. O. A. . Persson, H. Hogberg, S. Li, R. Ahuja,O. Eriksson, L. Hultman, and U. Jansson, Journal ofCrystal Growth 291, 290 (2006).

[20] X. Wang, H. Zhang, L. Zheng, Y. Ma, X. Lu, Y. Sun,and Y. Zhou, Journal of the American Ceramic Society95, 1508 (2012).

[21] N. J. Lane, M. Naguib, J. Lu, L. Hultman, and M. W.Barsoum, Journal of the European Ceramic Society 32,3485 (2012).

[22] N. J. Lane, M. Naguib, J. Lu, P. Eklund, L. Hultman,and M. W. Barsoum, Journal of the American CeramicSociety 95, 3352 (2012).

[23] H. Zhang, X. Wang, Y. Ma, L. Sun, L. Zheng, andY. Zhou, Journal of Advanced Ceramics 1, 268 (2012).

[24] W. K. Pang, I. M. Low, B. H. O’Connor, A. J. Studer,V. K. Peterson, Z. M. Sun, and J. P. Palmquist, Journalof Physics: Conference Series 251, 012025 (2010).

[25] W. K. Pang, I. M. Low, B. H. O’Connor, V. K. Peterson,A. J. Studer, and J. P. Palmquist, Journal of Alloys andCompounds 509, 172 (2011).

[26] W. K. Pang, I. M. Low, and Z. M. Sun, Journal of theAmerican Ceramic Society 93, 2871 (2010).

[27] B. Manoun, F. X. Zhang, S. K. Saxena, T. El-Raghy, andM. W. Barsoum, Journal of Physics and Chemistry ofSolids 67, 2091 (2006).

[28] M. W. Barsoum, M. Ali, and T. El-Raghy, Metallurgicaland Materials Transactions A 31, 1857 (2000).

[29] T. H. Scabarozi, S. Amini, O. Leaffer, A. Ganguly,S. Gupta, W. Tambussi, S. Clipper, J. E. Spanier, M. W.

Barsoum, J. D. Hettinger, and S. E. Lofland, Journal ofApplied Physics 105, 013543 (2009).

[30] N. V. Tzenov and M. W. Barsoum, Journal of the Ameri-can Ceramic Society 83, 825 (2000).

[31] S. C. Vogel, C. Hartig, L. Lutterotti, R. B. Von Dreele,H. R. Wenk, and D. J. Williams, Powder Diffraction 19,65 (2004).

[32] H. R. Wenk, L. Lutterotti, and S. Vogel, Nuclear Instru-ments & Methods in Physics Research Section A 515, 575(2003).

[33] H. Reiche, S. Vogel, P. Mosbrucker, E. Larson, andM. Daymond, Review of Scientific Instruments 83, 053901(2012).

[34] H. R. Wenk, L. Lutterotti, and S. C. Vogel, PowderDiffraction 25, 283 (2010).

[35] A. Larson and R. V. Dreele, “GSAS General StructureAnalysis System,” Los Alamos, NM (2004).

[36] S. C. Vogel, Journal of Applied Crystallography 44, 873(2011).

[37] J. C. Schuster and M. Palm, Journal of Phase Equilibriaand Diffusion 27, 255 (2006).

[38] P. E. Blochl, Physical Review B 50, 17953 (1994).[39] G. Kresse and J. Hafner, Physical Review B 47, 558

(1993).[40] G. Kresse and J. Hafner, Physical Review B (1994).[41] G. Kresse and J. Furthmuller, Physical Review B 54,

11169 (1996).[42] J. P. Perdew, K. Burke, and M. Ernzerhof, Physical

Review Letters 77, 3865 (1996).[43] S. Baroni, S. de Gironcoli, and P. Dal Corso, A.and Gi-

annozzi, Reviews of Modern Physics 73, 515 (2001).[44] A. Togo, F. Oba, and I. Tanaka, Physical Review B.[45] N. J. Lane, S. C. Vogel, G. Hug, A. Togo, L. Chaput,

L. Hultman, and M. W. Barsoum, Physical Review B86, 214301 (2012).

[46] B. H. Toby, Powder Diffraction 21, 67 (2006).[47] T. A. Hahn, International Journal of Thermophysics 12,

711 (1991).[48] Y. He, R. B. Schwarz, T. Darling, M. Hundley, S. H.

Whang, and Z. M. Wang, Materials Science and Engi-neering: A 239240, 157 (1997).

[49] N. J. Lane, S. C. Vogel, and M. W. Barsoum, PhysicalReview B 82, 174109 (2010).

[50] N. J. Lane, S. C. Vogel, and M. W. Barsoum, Journal ofthe American Ceramic Society 94, 3473 (2011).

[51] M. W. Barsoum, T. El-Raghy, C. J. Rawn, W. D. Porter,H. Wang, E. A. Payzant, and C. R. Hubbard, Journal ofPhysics and Chemistry of Solids 60, 429 (1999).

[52] M. W. Barsoum, C. J. Rawn, T. El-Raghy, A. T. Procopio,W. D. Porter, H. Wang, and C. R. Hubbard, Journal ofApplied Physics 87, 8407 (2000).

[53] Y. Bai, X. He, Y. Li, C. Zhu, and S. Zhang, J. Mater.Res 24, 2528 (2009).

[54] Y. Bai, X. He, C. Zhu, and G. Chen, Journal of theAmerican Ceramic Society (2012).

[55] H. Zhang, X. Wu, K. G. Nickel, J. Chen, and V. Presser,Journal of Applied Physics 106, 013519 (2009).