cu–ga–in thermodynamics: experimental study, modeling, and implications...

18
Cu–Ga–In thermodynamics: experimental study, modeling, and implications for photovoltaics Christopher P. Muzzillo 1 Carelyn E. Campbell 2 Timothy J. Anderson 1 Received: 18 September 2015 / Accepted: 6 December 2015 / Published online: 17 December 2015 Ó Springer Science+Business Media New York 2015 Abstract Both experimental measurements and assess- ment of phase equilibria are reported in the Cu–Ga–In material system, which is an important constituent in growth of the thin film photovoltaic absorber Cu(In,Ga)Se 2 (CIGS). Characterization of four different alloys using inductively coupled plasma atomic emission spectroscopy, X-ray diffraction, scanning electron microscopy, energy dispersive spectroscopy, differential thermal analysis, and differential scanning calorimetry has been conducted, and high-tem- perature equilibration studies have been performed on 2 of those. The new data are qualitatively consistent with the previous nonequilibrium thin film Cu–Ga–In observations. A thermodynamic assessment of the ternary system has also been performed using a CALPHAD approach after re- assessing the Cu–In constituent. The model fits the ternary data well in addition to the Cu–Ga, Cu–In, and Ga–In binary data. Practical applications of the model to metal precursors in a CIGS selenization process are discussed. Using the assessed parameters, the model predicts that at temperatures typically used in CIGS processes, Cu–Ga–In films should undergo equilibrium phase transformations. Slight changes in composition are found to determine whether or not these transitions occur, and at what temperature. Equilibrated precursor films are calculated to have high c-Cu 9 (Ga,In) 4 phase content, where reducing equilibration and c- Cu 9 (Ga,In) 4 formation have previously been found to improve photovoltaic performance. Introduction The Cu–Ga–In material system has received recent attention as a result of its use as an absorber in thin film photovoltaic (PV) applications and could potentially be useful in lead- free solder alloys as well [1, 2]. PV devices utilizing Cu(In x Ga 1-x )Se 2 (CIGS) absorber films have demonstrated power conversion efficiencies up to 21.7 % [3]. CIGS is often synthesized in a two-step selenization process begin- ning with deposition of a metal precursor film and followed by reaction with selenium and/or sulfur. In addition to the potential low cost, this ‘selenization’ process has the advantage of using established metal deposition techniques to reduce complexity in the first step and thus simplify scale- up. Device efficiencies obtained with selenization processes, however, have lagged relative to the champion co-evapo- rated films, due largely to nonoptimal cation distributions, and poor adhesion and void accumulation at the rear inter- face [4]. The phase constitution and spatial variations in the composition of the Cu–Ga–In precursor film influence the final cation distribution, void formation, and chalcogeniza- tion reaction pathways. The thermochemistry and phase equilibria of the Cu–Ga–In material system are, therefore, of interest for both the design of ternary sputtering targets, as well as understanding reactions during deposition of the precursor film and the heating selenization process. Knowledge of the material’s equilibrium behavior can assist in the design of economical, robust processes for fabricating quality CIGS PV absorbers. As summarized in the modeling section, the thermody- namic properties and phase equilibria of the constituent & Christopher P. Muzzillo muzzillo@ufl.edu Timothy J. Anderson tim@ufl.edu 1 Department of Chemical Engineering, University of Florida, Gainesville, FL 32611, USA 2 National Institute of Standards and Technology, Gaithersburg, MD 20874, USA 123 J Mater Sci (2016) 51:3362–3379 DOI 10.1007/s10853-015-9651-3

Upload: others

Post on 04-Feb-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

  • Cu–Ga–In thermodynamics: experimental study, modeling,and implications for photovoltaics

    Christopher P. Muzzillo1 • Carelyn E. Campbell2 • Timothy J. Anderson1

    Received: 18 September 2015 / Accepted: 6 December 2015 / Published online: 17 December 2015

    � Springer Science+Business Media New York 2015

    Abstract Both experimental measurements and assess-

    ment of phase equilibria are reported in the Cu–Ga–In

    material system, which is an important constituent in growth

    of the thin film photovoltaic absorber Cu(In,Ga)Se2 (CIGS).

    Characterization of four different alloys using inductively

    coupled plasma atomic emission spectroscopy, X-ray

    diffraction, scanning electron microscopy, energy dispersive

    spectroscopy, differential thermal analysis, and differential

    scanning calorimetry has been conducted, and high-tem-

    perature equilibration studies have been performed on 2 of

    those. The new data are qualitatively consistent with the

    previous nonequilibrium thin film Cu–Ga–In observations.

    A thermodynamic assessment of the ternary system has also

    been performed using a CALPHAD approach after re-

    assessing the Cu–In constituent. The model fits the ternary

    data well in addition to the Cu–Ga, Cu–In, and Ga–In binary

    data. Practical applications of the model to metal precursors

    in a CIGS selenization process are discussed. Using the

    assessed parameters, the model predicts that at temperatures

    typically used in CIGS processes, Cu–Ga–In films should

    undergo equilibrium phase transformations. Slight changes

    in composition are found to determine whether or not these

    transitions occur, and at what temperature. Equilibrated

    precursor films are calculated to have high c-Cu9(Ga,In)4phase content, where reducing equilibration and c-

    Cu9(Ga,In)4 formation have previously been found to

    improve photovoltaic performance.

    Introduction

    The Cu–Ga–In material system has received recent attention

    as a result of its use as an absorber in thin film photovoltaic

    (PV) applications and could potentially be useful in lead-

    free solder alloys as well [1, 2]. PV devices utilizing

    Cu(InxGa1-x)Se2 (CIGS) absorber films have demonstrated

    power conversion efficiencies up to 21.7 % [3]. CIGS is

    often synthesized in a two-step selenization process begin-

    ning with deposition of a metal precursor film and followed

    by reaction with selenium and/or sulfur. In addition to the

    potential low cost, this ‘selenization’ process has the

    advantage of using established metal deposition techniques

    to reduce complexity in the first step and thus simplify scale-

    up. Device efficiencies obtained with selenization processes,

    however, have lagged relative to the champion co-evapo-

    rated films, due largely to nonoptimal cation distributions,

    and poor adhesion and void accumulation at the rear inter-

    face [4]. The phase constitution and spatial variations in the

    composition of the Cu–Ga–In precursor film influence the

    final cation distribution, void formation, and chalcogeniza-

    tion reaction pathways. The thermochemistry and phase

    equilibria of the Cu–Ga–In material system are, therefore, of

    interest for both the design of ternary sputtering targets, as

    well as understanding reactions during deposition of the

    precursor film and the heating selenization process.

    Knowledge of the material’s equilibrium behavior can assist

    in the design of economical, robust processes for fabricating

    quality CIGS PV absorbers.

    As summarized in the modeling section, the thermody-

    namic properties and phase equilibria of the constituent

    & Christopher P. [email protected]

    Timothy J. Anderson

    [email protected]

    1 Department of Chemical Engineering, University of Florida,

    Gainesville, FL 32611, USA

    2 National Institute of Standards and Technology,

    Gaithersburg, MD 20874, USA

    123

    J Mater Sci (2016) 51:3362–3379

    DOI 10.1007/s10853-015-9651-3

    http://crossmark.crossref.org/dialog/?doi=10.1007/s10853-015-9651-3&domain=pdfhttp://crossmark.crossref.org/dialog/?doi=10.1007/s10853-015-9651-3&domain=pdf

  • binary systems have been experimentally investigated and

    assessed extensively [5–57]. Very little experimental study,

    however, has been carried out for the ternary system. Most

    researchers studying Cu–Ga–In deposit thin film alloys

    with overall compositions around Cu47Ga13In40 onto

    unheated substrates. They typically contain the c-Cu9(Ga,In)4 and g-Cu16(Ga,In)9 ternary solid solutions as wellas solid elemental indium, while the Cu11In9 compound is

    less frequently observed. The c and g phases exhibitextensive ternary solubility and, therefore, span much of

    the phase diagram outside of the Ga-rich corner, which is

    dominated by the liquid phase. No ternary compounds have

    been reported. Common to most deposition processes is the

    formation of In-rich nodules of mostly (In) atop a Cu- and

    Ga-rich two-dimensional film consisting largely of c-Cu9(Ga,In)4. The subsequent chalcogenization of this c-Cu9(Ga,In)4 compound at the rear interface is often the

    rate-limiting step in chalcopyrite film formation [58] and

    has been suggested to be related to void formation [4].

    The Cu–Ga–In ternary phase diagram was predicted

    recently by combining Gibbs energy functions estimated

    from assessments of the three binary systems [7]. The

    prediction was consistent with the results of an equilibra-

    tion study on a single Cu41Ga19In40 alloy [inductively

    coupled plasma atomic emission spectroscopy (ICP-AES),

    differential thermal analysis (DTA), and X-ray diffrac-

    tometry (XRD)] [59]. At room temperature, the as-cast

    alloy exhibited XRD peaks that were attributed to c-Cu9(Ga,In)4, Cu11In9, and (In). Thermal transitions at 429

    and 926 K were attributed to c-Cu9(Ga,In)4 ? L $ c-Cu9(Ga,In)4 ? (In) and L $ c-Cu9(Ga,In)4 ? L transfor-mations, respectively. Additionally, a study of phase rela-

    tions in Cu–Ga–In thin films on glass/Mo substrates was

    reported by Purwins et al. [60]. The data therein may not be

    at equilibrium because of (1) thin film stress, (2) possible

    phase transitions during the slow cooling of samples, and

    (3) possible substrate contamination (the type of glass was

    not reported; soda-lime glass is often used for CIGS PV

    substrates, from which Na usually accumulates in the film

    up to *1 at.%). Despite these potential deviations fromtrue equilibrium, the study nevertheless presents stabilized

    phase existence regions of particular importance to PV

    device processing.

    Experimental

    Sample preparation and alloy annealing

    A set of four alloy samples (A–D) was prepared for equi-

    libration studies. Two of the alloy compositions were close

    to that found in absorber precursors (molar ratios Cu/

    (Ga ? In) * 1 and Ga/(Ga ? In) * 0.3). Sample A was

    Cu-poor (Cu48Ga10In43) and sample B was Cu-rich (Cu55Ga18In27), where the Ga and Cu molar fractions were

    chosen to span the range of interest for PV applications.

    The other two samples were intended to better define the

    three-phase region near very Cu-rich alloys (sample C:

    Cu71Ga15In14 and sample D: Cu82Ga9In9). Alloys were

    prepared by vacuum induction melting of pure Cu, Ga, and

    In (99.999 % metals basis). The overall composition was

    determined by ICP-AES (Table 1). The compositional

    uniformity was also probed on pieces from three different

    locations on the ingots. Differential scanning calorimetric

    (DSC) or DTA measurements were performed on all

    samples, and samples C and D were equilibrated at three

    temperatures (591, 711, and 950 K). Pieces of the metal

    alloy ingot were washed in heated trichloroethylene, ace-

    tone, and methanol, followed by deionized (DI) water rinse

    and N2 blow drying, and then loaded into fused quartz

    ampoules with 2-mm-thick walls. The ampoule chamber

    was evacuated and purged three times with 4 % H2 in N2(99.999 % minimum purity with\1 ppm O2 and H2O, and\0.5 ppm total hydrocarbons). Hydrogen was used toprevent formation of surface oxides during the long-time

    equilibration anneal. The ampoules were sealed by soft-

    ening the outer walls around their plugs with an oxygen

    and acetylene torch, and slightly reduced pressure

    (0.20–0.33 atm) in the ampoules was used to minimize

    explosion hazards at high temperature.

    High-temperature samples (950 K) were annealed for a

    period of approximately 1 week, while the lower temper-

    ature samples (591 and 711 K) were annealed for 3 weeks.

    These durations were chosen to be slightly longer than

    those reported for Cu–Ga and Cu–In alloy equilibration

    [34, 44]. The sealed ampoules were placed at the center of

    a resistively heated tube furnace with temperature con-

    trolled to within approximately ±1 K. After equilibration,

    the ampoules were dropped from the furnace directly into

    icy brine to quench the high-temperature equilibrium

    states. A filed powder of sample C equilibrated at 591 K

    was additionally annealed at 600 ± 10 K for 10 min under

    4 % H2 in N2 and then cooled quickly to room temperature

    to relieve micro-strain for subsequent collection of high-

    resolution XRD patterns. Bulk samples were filed and then

    polished using lapping film with 0.3 lm Al2O3 particles.Polished surfaces were then swabbed with an etchant of

    67 vol.% ethyl alcohol and 33 vol.% aqueous solution

    [19 wt.% ferric chloride (FeCl3) and 6 wt.% hydrochloric

    acid (HCl) in DI water].

    Alloy characterization

    Compositions were measured by ICP-AES on a Perkin

    Elmer Optima 3200RL system, after dissolving the alloys

    in nitric acid and calibrating with plasma standard solutions

    J Mater Sci (2016) 51:3362–3379 3363

    123

  • obtained from Alfa Aesar. Phase and structure analysis was

    performed by XRD using a PANalytical PRO powder

    diffractometer and Cu Ka radiation generated at 45 kV and

    40 mA. DTA was performed on a Mettler–Toledo ther-

    mogravimetric analysis (TGA)/SDTA 851 system under

    pure N2 flow (no greater than 200 ppm O2 impurity). Quick

    scans were first acquired from room temperature to 1273 K

    for sample C and 1373 K for sample D at a ramp rate of

    10 K/min, followed by scans in temperature regions of

    interest at a ramp rate of 5 K/min. Thermograms of empty

    alumina crucibles were subtracted from each scan. Peak

    onset temperatures did not change within the accuracy of

    the technique (estimated ±1 K) in the ramp rate range

    (5–10 K/min). DSC scans were taken with a Netzsch

    Instrument STA 449C using alumina-lined platinum pans

    with a purified (gettering furnace to\10-10 ppm oxygen)Ar purge flow. Heating and cooling scans were performed

    twice at 10 K/min from room temperature to 1423 K. The

    phase distributions and composition were determined using

    a field-emission gun scanning electron microscope (FE-

    SEM) and energy dispersive X-ray spectroscopy (EDS),

    respectively, on a JEOL 6335F FE-SEM at 15.0 kV fila-

    ment tension. Both secondary electron (SE) and backscat-

    tered electron (BSE) micrographs were captured to better

    distinguish phases.

    Experimental results

    The compositions of the ingots are summarized in Table 1,

    as determined by ICP-AES. Overall ingot composition data

    from EDS are not presented, but satisfactory agreement

    between the two techniques was observed. No evidence of

    oxidation for any sample was found by EDS or XRD.

    XRD

    The crystalline phases of equilibrated alloys C and D were

    identified at room temperature by powder XRD, with

    example patterns shown in Fig. 1. The XRD pattern of

    sample C equilibrated at 711 K (Fig. 1) was very similar to

    sample C at 591 and 950 K, as well as sample D at 950 K.

    The pattern of sample D equilibrated at 711 K (Fig. 1) was

    very similar to sample D at 591 K. Assigning the peaks

    unambiguously to particular phases is challenging in this

    system, as the candidate intermetallic phases have multiple

    overlapping peaks that can shift with alloy composition.

    Also, the g-Cu16In9 binary phase field is known to exhibitsuperstructure ordering, although neither standard powder

    XRD patterns nor unambiguous structures have been

    published for all the phasoid regions [49]. XRD was first

    performed on the bulk samples, so phase fraction specifi-

    cation is complicated by the possible presence of texture as

    well as peak shifts due to probable misalignment and long-

    range stresses. Bulk sample XRD scans revealed more

    peaks that were narrower and more intense, and thus were

    used to help identify phases. XRD scans on filed powders

    had substantially more disorder, yielding broader, and less

    intense peaks, but more reliable peak positions and relative

    intensities, and so were used to measure lattice constants

    and phase fractions (summarized in Table 2). The com-

    position of c-Cu9(Ga,In)4 was on the Cu-rich side, so itsstructure was taken to be that of the c1-Cu9Ga4 phase withCu69Ga31 composition determined by Mizutani et al. [40].

    The published diffraction pattern of quenched, high-tem-

    perature c-Cu9In4 permitted linear interpolation of thecubic lattice constant to be related to group III

    Table 1 Composition of alloys,as determined by ICP-AES

    Sample Nominal composition Composition (at.%)

    Cu Ga In

    Kim et al. [59] Cu41Ga19In40 40.72 19.44 39.84

    A Cu48Ga10In43 47.6 9.5 42.9

    B Cu55Ga18In27 54.5 18.2 27.3

    C Cu71Ga15In14 71.2 ± 2.3 15.0 ± 1.0 13.9 ± 1.5

    D Cu82Ga9In9 81.6 ± 2.5 9.4 ± 1.0 9.0 ± 1.5

    Fig. 1 XRD patterns for samples C and D (Cu71Ga15In14 andCu82Ga9In9, respectively) bulk samples annealed at 711 K with

    according crystalline phase assignments. Reference powder diffrac-

    tion patterns were calculated from the known structures and site

    occupancies for a-(Cu), c-Cu9(Ga,In)4 [40], and d-Cu7In3 [36],respectively

    3364 J Mater Sci (2016) 51:3362–3379

    123

  • composition. Lattice parameter variation due to Cu/

    (Ga ? In), or Cu/III, compositional change is assumed to

    be negligible relative to that of In/III changes. The d-Cu7In3 structure [36], or powder diffraction file (PDF) #

    01-073-8028, and experimental diffraction pattern [32], or

    PDF# 03-065-2249, have been published. Lattice parame-

    ters of the a-fcc(A1)-(Cu) phase with up to 10.4 at.% Inand separately 20.0 at.% Ga have also been published [45,

    53]. Lattice constants exhibit a linear relationship with

    group III composition for both systems. Although solubil-

    ity of both Ga and In in a-(Cu) has not been studiedexperimentally, a linear combination of the 2 separate

    lattice parameter-composition functions has been used to

    arrive at an approximate anticipated lattice parameter for a

    particular ternary composition. In this way, a latticeparameters have been calculated from EDS compositional

    information and compared with those measured by XRD in

    Table 2.

    SEM/EDS

    Example SE and BSE micrographs of the samples etched in

    FeCl3:HCl:H2O after annealing at 950 K are shown in

    Fig. 2. The differential etching contrasted the phases to

    reveal the microstructure. The etching process also

    revealed surface scratches due to imperfect polishing. In all

    cases, both SE and BSE images revealed two phases, with

    the darker-appearing region the major phase in sample D

    and the minor phase in sample C. Multiple spot EDS

    measurements acquired for each phase in all samples

    indicated that the darker phase had a composition that

    corresponded to a-(Cu), while the lighter phase corre-sponded to either c-Cu9(Ga,In)4 or separately d-Cu7In3,depending on the sample (see Table 2). Overall sample

    compositions were measured using EDS with a beam raster

    area much larger than individual phase domains, and these

    were commensurate with ICP-AES results. Overall EDS

    composition was combined with individual phase compo-

    sition to obtain estimates for phase fractions in each sample

    (see Table 2). Although EDS and XRD lattice parameter

    measurements have relatively low precision, rough agree-

    ment between the two drastically different characterization

    techniques is presented in Table 2 to strengthen the phase

    constitution and composition conclusions.

    DTA/DSC

    The significant peaks from DTA scans for samples C and

    D are presented in Fig. 3. The scan in this figure for

    sample D only covered the temperature range 675 to

    1060 K. A single phase transition was detected for sample

    C by DTA, with its onset at 990 K. This peak also dis-

    plays a possible shoulder which ends at 1044 K. Two

    peaks were observed for sample D, with onset tempera-

    tures of 913 K and 1002 K. The 913 K transition is cor-

    roborated by XRD scans showing a change from a ? dfor sample D equilibrated at 711 K to a ? c when equi-librated at 950 K. Example DSC heating curves for

    samples A and B are presented in Figs. 4 and 5, respec-

    tively. When scans were repeated substantial variability

    (*20 K) was observed in the cooling curve peaks, andthis was taken to be the result of supercooling, as there

    was little change in the repeated heating curve peaks.

    Heating curve peak onsets were therefore taken to cor-

    respond to phase transitions listed in Table 3. Both sam-

    ple A and B display transitions at *428 K and highertemperatures. For comparison, a DTA scan of a sample

    with Cu content (Cu41Ga19In40) lower than samples A–D

    from Kim et al. is reproduced in Fig. 6 [59]. Due to the

    ambiguity of assigning onset temperatures to the high-

    temperature peaks of Cu41Ga19In40 [59], sample A, and

    sample B calculated enthalpy curves have been overlaid

    for comparison in Figs. 4, 5, 6. The phase transitions

    which are predicted to correspond to these heat effects are

    discussed in section ‘‘Cu–Ga–In Assessment,’’ and com-

    pared with calculations in Table 3.

    Table 2 Comparison of phases and their composition and extent asdetected from XRD and SEM/EDS measurements at three equili-

    bration temperatures

    SEM/EDS XRD

    Sample C 591 K 18 % a; 82 % c

    a… a = 3.690 Åc… In/III = 57 %

    8 % a; 92 % c

    c… In/III = 48 %

    Sample D 591 K 40 % a; 60 % d

    a… a = 3.677 Å61 % a; 39 % d

    a… a = 3.670 ÅSample C 711 K 21% a; 79 % c

    a… a = 3.657 Åc… In/III = 61 %

    21 % a; 79 % c

    a… a = 3.663 Åc… In/III = 52 %

    Sample D 711 K 64 % a; 36 % d

    a… a = 3.649 Å57 % a; 43 % d

    a… a = 3.666 ÅSample C 950 K 24 % a; 76 % c

    a… a = 3.722 Åc… In/III = 54 %

    6 % a; 94 % c

    c… In/III = 43 %

    Sample D 950 K 45 % a; 55 % c

    a… a = 3.678 Åc… In/III = 74 %

    18 % a; 82 % c

    a… a = 3.686 Åc… In/III = 58 %

    EDS phase amounts were estimated by combining overall composi-

    tion with those of individual phases. Lattice constants were calculated

    assuming linear composition dependence. XRD phase amounts were

    estimated from relative peak areas

    J Mater Sci (2016) 51:3362–3379 3365

    123

  • Thermodynamic assessment

    Thermodynamic literature

    The three constituent binary systems have previously been

    assessed. The recent Cu–Ga assessment by Li et al. [34]

    has been used here with almost no modification; the only

    additional data considered being the new structural infor-

    mation for the c1-Cu9Ga4 compound [40]. The Ga-Inassessment by Anderson and Ansara has been used here

    without modification [5]. The subsequent Ga-In assessment

    by Reddy and Hajra introduces 2 more parameters in a less-

    common composition- and temperature-dependent excess

    Fig. 2 Scanning electron micrographs at 92000 magnification of samples equilibrated at 950 K, with a-(Cu) and c-Cu9(Ga,In)4 phases labeled.a SE image and b BSE image of sample C. c SE image and d BSE image of sample D

    Fig. 3 DTA heating curves for samples C and D, with peak onsettemperatures indicated

    Fig. 4 DSC heating curve (left axis) and calculated system enthalpy(right axis) for sample A, where mol refers to the chemical formula

    3366 J Mater Sci (2016) 51:3362–3379

    123

  • Gibbs energy function to marginally reduce residual error

    for selected data [50], relative to the prior assessment, and

    so has not been used. Two phase models (Cu11In9 and

    liquid) from the recent Cu–In assessment by Cao et al. have

    been used here without modification [11]. The re-assess-

    ment of the Cu–In binary system reported in this work

    considered the extensive equilibrium experimental data [6–

    10, 12, 13, 15–19, 21, 22, 24–26, 28–33, 35, 36, 39, 41–44,

    46–49, 51–53, 55–57], as well as several previous ther-

    modynamic assessments [11, 20, 27, 37, 38, 54]. The newly

    reported Cu10In7 phase has not been presently included in

    the model because no previous experimental reports have

    documented it, apparently due to pronounced kinetic bar-

    riers to its formation (requiring 9 mo of equilibration to

    surmount) [46]. A ternary Cu–Ga–In CALPHAD model

    has previously been fitted to the data of Purwins et al. and

    Kim et al. [59, 60], although the optimized parameters

    were not included in the report. The data from this work,

    Purwins et al. [60] and Kim et al. [59] were considered in

    the Cu–Ga–In optimization.

    Thermodynamic models

    Pure elements

    The pure solid �GfccCu ,�GortGa, and

    �GtetIn are taken as the a-fcc(A1)-(Cu), orthorhombic-Ga, and tetragonal(A6)-(In)-

    condensed phase reference Gibbs energy functions relative

    to the respective standard elemental reference enthalpies

    published by Dinsdale [61]. The pure element liquid phases

    and lattice stabilities have also been taken from Dinsdale.

    Gas-phase species of Cu, Cu2, Ga, Ga2, In, and In2 are

    included in the database, and their models are taken from

    the Scientific Group Thermodata Europe (SGTE) 1994

    substance database. The standard elemental reference

    functions have the following form.

    �Giu ¼ Gui � HSERi¼ aþ b � T þ c � T � ln T þ d � T2 þ e � T3 þ f � T�1

    þ g � T4 þ h � T5 þ . . .ð1Þ

    Here G, �, u, i, H, SER, a–h, and T represent Gibbsenergy, use of a standard elemental reference, phase

    structure, component, enthalpy, standard elemental refer-

    ence (298.15 K at 1 bar), estimated coefficients, and

    absolute temperature, respectively. The coefficients are

    typically fitted to experimental data and can be directly

    related to heat capacity and entropy.

    Stoichiometric compounds

    The expressions and estimated parameters for the Gibbs

    energy of Cu778Ga222 and CuGa2 suggested by Li et al.

    were used without further modification [34]. The Cu11In9phase model and parameters of Cao et al. were used [11]. It

    has been established that the d-Cu7In3 and g-LT-Cu16In9phases exist over *1 at.% compositional ranges [23, 54].However, to base the models for these phases on their

    structures and to produce more reliable extrapolations [62],

    it was necessary to model them as line compounds.

    Enthalpy data for d-Cu7In3 [7] (Fig. 7), as well as invariantreaction compositions and temperatures (Table 4), were

    used to optimize the parameters for d-Cu7In3 and g-LT-Cu16In9 listed in Table 5. The models use the following

    formalism.

    �GCumIIIn ¼ GCumIIIn � m � HSERCu � n � HSERIII ¼ aþ b � Tð2Þ

    Here III has been used to represent either Ga or In, and

    the stoichiometric coefficients are m and n. The optimized

    Fig. 6 DTA cooling curve (left axis) [59] and calculated systementhalpy (right axis) for the Cu41Ga19In40 alloy, where mol refers to

    the chemical formula

    Fig. 5 DSC heating curve (left axis) and calculated system enthalpy(right axis) for sample B, where mol refers to the chemical formula

    J Mater Sci (2016) 51:3362–3379 3367

    123

  • coefficients, a and b, are the enthalpy and negative entropy

    of formation, respectively.

    Solution phases

    A simple ideal-substitutional-solution model (ideal gas)

    has been used for the gas phase [63]. An ionic two-sub-

    lattice liquid model has been used with the metal ion

    species (Cu?1, Ga?3, and In?3) mixing on the cation

    sublattice and only vacancies (Va) on a second sublattice

    [64]. This model has been chosen so that anion species may

    be readily added for a Cu–Ga–In–Se database. Since there

    are no anions in the present material system, the ionic two-

    sublattice model is mathematically equivalent to a substi-

    tutional solution (also known as Redlich–Kister, or R–K

    model [65]), the only difference being the ionized species

    in place of the neutral atoms. The liquid parameters were

    therefore taken from the previous assessments of Li et al.

    for Cu–Ga [34], Cao et al. for Cu–In [11], and Anderson

    and Ansara for Ga-In [5]. As the respective assessments

    have achieved excellent fits to extensive liquid component

    activity measurements, the binary liquid parameters were

    not modified, and no ternary parameters were added.

    Equilibrium calculations for the ternary system extrapolate

    the binary R–K series parameters using the symmetric

    Muggianu method [63]. The R–K formalism is as follows:

    Gu ¼X

    i

    xi �� Gui þ R � T �X

    i

    xi � ln xi

    þX

    i

    X

    j[ ixi � xj �

    X

    m

    mLui;j � xi � xj

    � �m ð3Þ

    mLui;j ¼ aþ b � T ð4Þ

    Here R, m, and L represent the gas constant, parameterorder, and mixing parameter, respectively. For m only up to0, the R–K model simplifies to an ideal-substitutional-so-

    lution model, and for m only up to 1, it simplifies to asubregular-solution model [63].

    The f, c2-Cu9Ga4 and c3-Cu9Ga4 subregular-solutionmodels of Li et al. were used without further modification,

    Table 3 Comparison of Cu–Ga–In experimental thermal

    analyses and calculated phase

    transitions

    Sample Experimental Calculated

    Peak range (K) Transformation Temperature (K)

    Cu41Ga19In40 [59]

    Cu4322Ga1844In3834

    890–949 L $ c 885.8–924.2L $ c ? L 885.8–891.2

    425–455 L $ c ? (In) 374.8–407.5A

    Cu476Ga95In429

    860–946 L $ g 904.0–933.4L $ g ? L 904.0–912.2L $ c ? g ? L 904.0

    428 L $ c ? g ? (In) 427.7B

    Cu55Ga18In27

    825.9–1018.8 L $ c 825.9–1001.4L $ c ? L 904.2c ? L $ g 825.9

    427.3 L $ c ? g ? (In) 427.7C

    Cu711Ga150In139

    990–1044 L $ c 990.4–1048.9L $ a ? c 990.4–1014.4

    D

    Cu844Ga65In91

    1002–1027 L $ a 945.6–1176.7L $ a ? c 945.6–946.2

    913–939 c $ a ? d 912.7–917.3

    Compositions in italics were used in the calculations

    Fig. 7 Enthalpy of formation of the Cu7In3 alloy as a function oftemperature, with reference state at 298 K, where mol refers to the

    chemical formula. The line is calculated and the circles are

    experimental data [7]

    3368 J Mater Sci (2016) 51:3362–3379

    123

  • as they only exist on the Cu–Ga binary tie-line at equi-

    librium [34]. Here the ternary c-Cu9(Ga,In)4 phase is takento be separate from c2-Cu9Ga4 and c3-Cu9Ga4. The high-temperature disordered c0-Cu9Ga4 phase has been identi-fied experimentally by the observation of DSC transitions.

    However, it has been presently omitted because its relation

    to the c-Cu9(Ga,In)4 and c-Cu9In4 compounds (as distinct

    from the c1-Cu9Ga4 compound’s relation to them) isunclear. The substitutional-regular-solution model taken

    from Anderson and Ansara has been employed for the

    tetragonal(A6)-(In), or (In), solid phase, which incorpo-

    rates Ga but not Cu.

    An R–K formalism has been used for both the a, orfcc(A1)-(Cu), phase and the b, or bcc(A2)-Cu4(Ga,In),

    Table 4 Comparison ofexperimentally measured and

    calculated invariant reactions in

    the binary Cu–In system

    Cooling reaction Respective phase’s composition (at.% In) T (K) Reference

    a ? L $ b 9.5 20.8 18.3 988 [57]10.05 20.9 18.05 983 [21]

    7.5 20.0 19.0 984.5 [8]

    7.5 19.5 18.8 976.4 This work

    L $ c 29.1 29.1 – 958.2 [57]29.56 29.56 955.5 [51]

    29.4 29.4 957.3 [8]

    30.3 30.3 957.5 This work

    L $ b ? c 25.8 23.8 27.2 952 [57]– 24.5 – 949 [21]

    26.0 22.0 28.5 952.2 [8]

    26.9 24.4 28.9 953.8 This work

    c ? L $ g-HT 31.2 35.2 32.7 944 [57]32.0 35.6 33.3 943.4 [8]

    - 35.4 33.0 940 [23]

    31.9 37.4 35.2 944.5 This work

    c $ d 29.5 29.5 – 904.2 [57]29.8 29.8 905.4 [8]

    30.0 30.0 917.7 This work

    c $ d ? g-HT 30.8 30.5 32.7 888 [57]32.0 30.6 33.3 891.0 [8]

    – – 33.4 887 [23]

    31.8 30.0 35.1 896.3 This work

    c $ b ? d 27.9 22.1 28.2 889 [57]27.7 21.8 28.9 890 [51]

    – 22.0 29.0 893.3 [8]

    26.8 21.3 30.0 885.4 This work

    b $ a ? d 20.0 11.63 28.7 847 [57]20.15 10.90 – 847 [21]

    19.0 – 29.0 849.7 [8]

    – 10.90 – 848 [43]

    – 10.85 – 848 [24]

    19.6 9.3 30.0 847.1 This work

    d ? g-HT $ g-LT 32.0 35.0 34.5 661.8 [8]30.0 36.3 36.0 652.8 This work

    g-HT ? L $ Cu11In9 36.4 97.0 45.0 579.0 [8]38.6 95.9 45.0 574.3 This work

    g-HT $ g-LT ? Cu11In9 – – – 549.8 [8]38.5 36.0 45.0 554.7 This work

    L $ Cu11In9 ? (In) – – 100.0 428.7 [8]99.3 45.0 100.0 426.3 This work

    J Mater Sci (2016) 51:3362–3379 3369

    123

  • Table 5 Optimized Gibbsenergy expression parameters,

    where mol refers to the

    chemical formula

    Phase T range (K) Parameters (J/mol)

    a-fcc(A1)-(Cu,In) 298–3000 �GaCu ¼� GfccCu � HSERCu�GaIn ¼� GfccIn � HSERIn0LaCu;In ¼ �22452:5439 þ 39:4052846 � T1LaCu;In ¼ �35850:2348 � 9:1642906 � T2LaCu;In ¼ 62576:0783 � 39:7753474 � T

    b-bcc(A2)-Cu4In 298–3000 �GbCu ¼� GbccCu � HSERCu�GbIn ¼� GbccIn � HSERIn0L

    bCu;In ¼ �3632:122 þ 0:4448 � T

    1LbCu;In ¼ �47128:413 � 0:6313 � T

    2LbCu;In ¼ 13753:699 þ 11:2268 � T

    d-Cu7In3 298–3000 �GdCu7In3 � 0:7 �� GfccCu � 0:3 �� GtetIn ¼ �7991:308 þ 1:1703 � T

    c-Cu9In4 298–3000 �GcCu:Cu:Cu:Cu:In � 12 �� GfccCu �� GtetIn ¼ 0

    �GcCu:Va:Cu:Cu:In � 11 �� GfccCu �� GtetIn ¼ 0�GcCu:Cu:Cu:In:In � 9 �� GfccCu � 4 �� GtetIn ¼ �62519:633 � 11:0002 � T�GcCu:Cu:In:Cu:In � 9 �� GfccCu � 4 �� GtetIn ¼ 0�GcCu:Va:Cu:In:In � 8 �� GfccCu � 4 �� GtetIn ¼ �80900 � 3 � T�GcCu:Va:In:Cu:In � 8 �� GfccCu � 4 �� GtetIn ¼ 0�GcCu:Cu:In:In:In � 6 �� GfccCu � 7 �� GtetIn ¼ 0�GcCu:Va:In:In:In � 5 �� GfccCu � 7 �� GtetIn ¼ 0

    g-HT-Cu16In9 298–3000 �Gg�HTCu:Cu:In � 2 �� GfccCu �� GtetIn ¼ �19569:83 � 0:9607 � T�Gg�HTCu:Va:In �� GfccCu �� GtetIn ¼ �7199 þ 0:14 � T0L

    g�HTCu:Cu;Va:In ¼ �17436 þ 26:453 � T

    g-LT-Cu16In9 298–3000 �Gg�LTCu16In9 � 0:64 �� GfccCu � 0:36 �� GtetIn ¼ �8173:8 þ 1:38 � T

    a-fcc(A1)-(Cu,Ga,In) 298–3000 0LaGa;In ¼ 7369b-bcc(A2)-Cu4(Ga,In) 298–3000 0LbGa;In ¼ 150000d-Cu7(Ga,In)3 �GdCu:Ga � 0:7 �� GfccCu � 0:3 �� GortGa ¼

    �GdCu:In � 0:7 �� GfccCu � 0:3 �� GtetIn ¼�7991:308 þ 1:1703 � T

    1LdCu:Ga;In ¼ 17:04 � Tc-Cu9(Ga,In)4 298–3000 �GcCu:Cu:Cu:Ga:In � 9 �� GfccCu � 3 �� GortGa �� GtetIn ¼

    �GcCu:Cu:Cu:Ga:Ga � 9 �� GfccCu � 4 �� GortGa ¼� 166571:3935 � 46:786919 � T�GcCu:Va:Cu:Ga:In � 8 �� GfccCu � 3 �� GortGa �� GtetIn ¼�GcCu:Cu:Cu:Ga:Ga � 9 �� GfccCu � 4 �� GortGa ¼� 166571:3935 � 46:786919 � T

    0LcCu:Va:Cu:Ga;In:In ¼ 10:94 � T

    1LcCu:Va:Cu:Ga;In:In ¼ 13100

    g-HT-Cu16(Ga,In)9 298–3000 �Gg�HTCu:Cu:Ga � 2 �� GfccCu �� GortGa ¼�Gg�HTCu:Cu:In � 2 �� GfccCu �� GtetIn ¼ �19569:83 � 0:9607 � T�Gg�HTCu:Va:Ga �� GfccCu �� GortGa ¼ �17199 � 14:142 � T ¼�Gg�HTCu:Va:In �� GfccCu �� GtetIn � 10000 � 14:282 � T0L

    g�HTCu:Cu;Va:Ga ¼ �17436 þ 26:453 � T ¼

    0Lg�HTCu:Cu;Va:In � 120 þ 0:5 � T

    0Lg�HTCu:Cu:Ga;In ¼ �99950 þ 100 � T

    1Lg�HTCu:Va:Ga;In ¼ �12400 þ 101:8 � T

    3370 J Mater Sci (2016) 51:3362–3379

    123

  • phase [65]. While the previous Cu–Ga–In assessment

    employed a sublattice model for b with Cu0.73(Cu,Ga,In,Va)0.14(Ga,In)0.13, the R–K model based on the

    structure was used in this work to allow more realistic

    extrapolations and to employ sublattices suitable for dif-

    fusion modeling. Binary a and b parameters were takenfrom Li et al. for Cu–Ga and used without modification.

    Modifications were made to both a binary Cu–In parame-ters taken from Cao et al. and to b binary Cu–In parameterstaken from Hertz et al. [20]. A single ternary mixing

    parameter was added to the b description to destabilize andprevent ternary solubility, which has not been observed

    experimentally (Table 5). Also, a single ternary mixing

    parameter was added to the a phase to destabilize andprevent its appearance in the Ga-In binary tie-line.

    The optimized parameters of the binary c1-Cu9Ga4phase were taken from Li et al. although the sublattice was

    slightly modified. The new Cu5(Cu,Va)(Cu,Ga)3(Cu,Ga)3-

    Ga sublattice is based on the site occupancies published by

    Mizutani et al. [40], and it is more suitable for extrapola-

    tion and diffusion modeling than the previous assessment’s

    Cu0.654(Cu,Ga,In,Va)0.115(Ga,In)0.231 model [59]. The

    analogous sublattice model was also optimized for the c-Cu9In4 phase in the Cu–In binary description to fit exper-

    imental enthalpy data (Fig. 7) and to slightly improve

    phase boundaries (see Fig. 8). The ternary c-Cu9(Ga,In)4(or simply c) description combines the c1-Cu9Ga4 and c-Cu9In4 binary parameters into a single phase along with

    two new ternary c end-members to fit experimental solu-bility data (Table 5). These end-members were included to

    extend the single-phase c region to encompass the 623 KCu659Ga218In123 datum of Purwins et al. [60]. Specifically,

    the Cu5(Va)(Cu)3(Ga)3(In) and Cu5(Cu)(Cu)3(Ga)3(In)

    end-member parameters were used directly from the binary

    Cu5(Va)(Cu)3(Ga)3Ga and Cu5(Cu)(Cu)3(Ga)3Ga end-

    members, respectively, after appropriately replacing �GortGa

    with �GtetIn . The two new ternary c mixing parameters wereadded to pin the boundary between the two-phase c ? Land 3-phase c ? g ? L regions observed in the 623 Kisotherm (here g refers to the ternary g-Cu16(Ga,In)9 forbrevity) [60]. Those data are not technically equilibrated,

    but they are qualitatively corroborated by the experimental

    phase transitions observed in the neighboring Cu41Ga19In40and sample B alloys. The g phase was not observed byXRD in the Cu41Ga19In40 alloy [59], which is in agreement

    with the boundary drawn by Purwins et al. [60]. On the

    other hand, sample B’s DSC curve displays a low-tem-

    perature peak (corresponding to L $ c ? g ? (In) inTable 3), a possible shoulder around 826 K (corresponding

    to c ? L $ g), which is then followed by a broad thermalarrest up to 1018.8 K (L $ c in Table 3). These threetransitions in sample B compared with the two transitions

    and lack of g observation in Cu41Ga19In40 indicate that thethin film 623 K isotherm boundary approaches that of

    equilibrium (see phase diagram in Fig. 9). The1L

    cCu:Va:Cu:Ga;In:In mixing parameter also extends the single-

    phase c region to Ga-poor compositions as temperature isincreased. This was experimentally observed, as less Ga

    content was found in the c phase in sample C at 711 K thanfor 591 K. Stabilization of the c phase at these more Ga-poor compositions additionally at lower temperatures was

    avoided because it was found to narrow the two-phase

    a ? d region and lower the c $ d transition temperature,pushing both further away from experimental observation.

    No Ga solubility was observed by EDS or by XRD peak

    shifts in the d-Cu7(Ga,In)3 phase in sample D equilibrated at591 and 711 K. However, a Cu0.7(Ga)0.3 end-member with

    parameters equal to those of Cu0.7(In)0.3, along with a single

    first-order mixing parameter were added to its model

    Fig. 8 Close-up of the calculated Cu–In phase diagram with overlaidexperimental data

    Fig. 9 Calculated Cu–Ga–In 623 K isotherm with overlaid experi-mental data [59, 60]

    J Mater Sci (2016) 51:3362–3379 3371

    123

  • (Table 5). This single new parameter allowed the a ? d2-phase region to more closely approach the composition of

    sample D (Fig. 9), and it enabled the accurate prediction of

    the c $ a ? d transition temperature at 913.0 K (Table 3).The g phase field is centered at the Cu16In9 (Cu0.64In0.36)

    composition in the binary Cu–In phase diagram and con-

    sists of at least 5 equilibrium phases [23]. These phases

    span the range of approximately 33.0–38.0 at.% In [23,

    51], depending on temperature. The exact nature of these

    phases’ stability and equilibrium boundaries is unsolved;

    however, it is clear that they are all based on a simple P63/

    mmc substructure [35]. Like the previous Cu–In thermo-

    dynamic assessments, only an ordered low-temperature (g-LT-Cu16In9) phase and a disordered high-temperature (g-HT-Cu16In9) phase have been considered [11, 37, 38].

    XRD patterns indicate ternary solubility in a P63/mmc

    structure [60], and ternary solubility in LT superstructures

    has been documented [14, 58, 66–69]. Samples A and B,

    however, display no thermal transition that could corre-

    spond to g-HT-Cu16(Ga,In)9 $ g-LT-Cu16(Ga,In)9.Therefore, only g-HT is extended to Cu–Ga–In. An g-HTmodel similar to that used by Kao et al. was employed,

    except the energetically unfavorable VaIn was not allowed

    [27]. The following sublattice model has been used for the

    g-Cu16(Ga,In)9 phase.�Gg ¼ y00Cu � y000Ga ��G

    gCu:Cu:Ga þ y000In ��G

    gCu:Cu:In

    � �

    þ y00Va � y000Ga ��GgCu:Va:Ga þ y000In ��G

    gCu:Va:In

    � �

    þR �T � y00Cu � lny00Cu þ y00Va � lny00Va þ y000Ga � lny000Ga þ y000In � lny000In� �

    þ y00Cu �y00Va � y000Ga � 0LgCu:Cu;Va:Ga þ y000In � 0L

    gCu:Cu;Va:In

    � �

    þ y00Cu �y000Ga �y000In � 0LgCu:Cu:Ga;In

    þ y00Va �y000Ga �y000In � 1LgCu:Va:Ga;In � y000Ga � y000In

    � �

    ð5Þ

    The first two terms are the surface of reference end-

    members, the third term is the configurational entropy, and

    the final terms are the excess Gibbs energies. The

    Cu0.545(Cu,Ga,In,Va)0.122(Ga,In)0.333 model used in the

    previous Cu–Ga–In assessment was exchanged in favor of

    the structure-based Cu(Cu,Va)(Ga,In), which again is

    expected to yield more accurate extrapolations.

    As the Cu–Ga binary system contains no analogous

    phase, its parameters were taken to be equal to those

    optimized for Cu–In, only with �GortGa replacing�GtetIn . These

    2 binary phase descriptions were combined for the ternary

    g description. The single Cu–In binary mixing parameterwas unchanged and entered as a Cu–Ga binary mixing

    parameter. Two new temperature-dependent ternary mix-

    ing parameters were added to complete the optimization

    (Table 5). The 0LgCu:Cu:Ga;In parameter dominates stability in

    the Cu-rich corner. Stabilizing too much at lower

    temperature was found to further narrow the two-phase

    a ? d (d refers to d-Cu7(Ga,In)3) region, contrary toexperimental evidence. Moderate stabilization was neces-

    sary to increase the c $ a ? d transition temperaturecloser to the experimental value for sample D. Destabi-

    lization of this parameter at high temperature was neces-

    sary to match the experimental g solidus and liquiduscurves for samples A and B. The asymmetric 1L

    gCu:Va:Ga;In

    parameter was needed to destabilize g at Ga-rich compo-sitions. This parameter also determines the boundary

    between the two-phase g ? L and three-phase c ? g ? Lregions, determined experimentally in the thin film 623 K

    isotherm. Its destabilization at high temperature was also

    necessary to match experimentally determined g solidusand liquidus temperatures in samples A and B. A total of

    11 terms were added or modified in the assessment of the

    ternary system, not including those used to prevent for-

    mation of a and b in the Ga-In binary system.

    Results and discussion

    Thermodynamic calculations and parameter optimizations

    were performed with global Gibbs energy minimization

    software (Thermo-Calc and Pandat). The calculated binary

    phase diagrams are presented in Figs. 10, 11, 12.

    Cu–In assessment

    As can be seen from Fig. 8, the calculated g-HT high-temperature phase boundary does not sufficiently extend to

    the Cu-rich compositions to fit experimental measure-

    ments. However, improved fit to enthalpy measurements

    Fig. 10 Cu–Ga phase diagram calculated with the model used in thepresent work, from Li et al. [34]

    3372 J Mater Sci (2016) 51:3362–3379

    123

  • (Table 6; Fig. 7) with a sublattice that is based on the real

    structure was obtained, which is better suited for ternary

    extrapolation and diffusivity modeling. The phase bound-

    aries of the a-(Cu,In), b-Cu4In, and c-Cu9In4 have all beenimproved slightly (Fig. 8). The calculated invariant reac-

    tions are in line with the previous measurements and

    assessments (see Table 4). Overall fit to enthalpy data has

    been improved for the re-assessed phases (see Table 6).

    Cu–Ga–In assessment

    As can be seen from the calculated isotherm in Fig. 9,

    satisfactory fit to the data has been obtained. The thin film

    single-phase c datum at Cu659Ga218In123 composition [60]is actually predicted to contain two phases at equilibrium:

    89 % c and 11 % g. It is postulated that either lack ofaccuracy in the composition measurement technique (the

    technique is not reported) or the minor phase amounts in

    the already thin film were not sufficient to produce a dis-

    cernible XRD signal. A composition of Cu671Ga212In117,

    which probably lies within the error of the composition

    measurement technique, does lie in the calculated single-

    phase c region. The average composition of sample Ddetermined by ICP places it in the calculated three-phase

    a ? c ? d region at 623 K. However, if the composition isshifted to Cu8198Ga895In907 (which falls well within the

    error margin of the experimental technique), the model

    predicts it to fall in the two-phase a ? d region, inagreement with experiment.

    The phases identified by XRD for each sample were the

    same as those predicted by the thermodynamic model. The

    only crystalline phases observed in samples C and D were

    a, d, and c. The XRD pattern for Cu41Ga19In40 displayed cand (In) peaks as well as peaks which were attributed to

    Cu11In9 [59], but due to peak overlap with c, the presenceof Cu11In9 in that alloy is considered uncertain in the

    present work. The model has also been optimized against

    thermal analysis scans at the five compositions shown in

    Fig. 9. Better agreement was found when calculations used

    compositions which were slightly different from the as-

    measured values (see Tables 1, 3). The phase transitions

    observed with thermal analyses are predicted adequately by

    the model (Figs. 3, 4, 5, 6; Table 3). Notably, the largest

    deviations of the calculated transitions from experiment

    coincide with peaks which have ambiguous onsets. In these

    cases, calculated system enthalpies have been overlaid on

    experimental thermal scans to assist in the comparison

    (Figs. 4, 5). The Cu41Ga19In40 alloy displayed a thermal

    event with onset at 425 K, which is not predicted well by

    the model, and which is very similar to the 427.7 K

    eutectic L $ c ? g ? (In) transition temperature. Thismay be evidence that at lower temperatures, the c com-pound’s (c-Cu9(Ga,In)4; c1-Cu9Ga4; c2-Cu9Ga4; c3-Cu9Ga4) stability extends the three phase c ? g ? (In) regionto a more Ga-rich composition than predicted by the

    model. Further experimental evidence would be needed to

    substantiate this possibility. Overall, good binary system

    fits have been retained, while re-assessment of the Cu–In

    binary system has enabled appropriate ternary extrapola-

    tions. The few added ternary parameters were necessary to

    match experiments, and good fit was obtained.

    Practical applications

    The thermodynamic model may be used as a guide for

    choosing promising Cu–Ga–In precursors for selenization

    Fig. 11 Cu–In phase diagram calculated with the optimized model

    Fig. 12 Ga-In phase diagram calculated with the model used in thepresent work, from Anderson and Ansara [5]

    J Mater Sci (2016) 51:3362–3379 3373

    123

  • processes. For metal precursors sputtered onto Mo sub-

    strates, the formation of a Cu- and Ga-rich matrix with In-

    rich nodules on top is usually observed [4, 58, 70–73].

    Examples of various deposition methods that affect this

    morphological segregation are sequential sputtering of a

    Cu0.8Ga0.2/In bilayer [73], sequential in situ sputtering of

    350 ultrathin-stacked layers of Cu0.8Ga0.2 and pure In [4,

    70], sputtering from single ternary targets of varying

    compositions [72, 74–76], or simultaneously co-sputtering

    onto the same substrate from dual targets [77]. Similar

    results are obtained by co-deposition from effusion sources

    onto unheated substrates in the cases of CuGa/CuIn

    bilayers, Cu/Ga/In layers repeated four times, and the same

    repeated eight times [73]. The overall film compositions for

    these examples fall between 0.8\Cu/III\ 1.0 and0.2\Ga/III\ 0.3 [4, 70, 72, 73, 75–77]. The lack ofdependence on deposition sequence suggests that the

    observed phase segregation is correlated with overall film

    composition more than it is correlated with local compo-

    sitional modulations. Calculations for Cu/III = 0.9 and Ga/

    III = 0.3 at room temperature predict an equilibrated Cu–

    Ga–In system to contain c, g, and (In) with indium molefractions of 0.08, 0.23, and 1.00, respectively. It is

    speculated that sputtered and co-deposited films may tend

    toward equilibrium segregation of the two major phases

    (Cu–Ga-rich c and In-rich (In)) due to natural heating ofthe substrate induced by sputtered particle bombardment or

    energetic evaporant arrival, depending on the process.

    Liquid indium is known to have a significantly lower sur-

    face free energy than both liquid gallium and copper for all

    temperatures [78]. Liquid alloys containing mostly indium

    are expected to have surface tensions approaching that of

    pure indium. Therefore, the surface segregation of In-rich

    nodules may be related to the liquid phase, while its rela-

    tion to any solid phases remains unknown.

    Calculations indicate that overall cation compositions

    that are reasonably close to those of record efficiency CIGS

    cannot avoid (In) and In-rich liquid phases in equilibrated

    metal precursors below and above 428 K, respectively.

    Figure 13 shows phase fractions of systems as a function of

    temperature with Cu/III = 0.9 at different Ga/III. As can

    be seen, the g phase is avoided at all temperatures for Ga/III = 0.4 (more precisely, for all Ga/III C 0.34). Similarly,

    the c phase can be avoided at all temperatures for all Ga/III B 0.14. The c phase typically segregates toward the rearcontact in the as-deposited metal precursors and also

    Table 6 Comparison of experimental, previous assessments’ calculated, and the present assessment’s calculated standard formation enthalpiesin the binary Cu–In system, with pure solid fcc(A1)-Cu and pure solid tetragonal(A6)-In reference states, and mol refers to the chemical formula

    Exp. phase T (K) Mole fraction of In Experimental DH (J/mol) Previous calc. DH (J/mol) Present calc. DH (J/mol)

    b 903 0.2086 -431 [25] -976 [27], -48 [20], -1330 [37] -1017

    b 903 0.2194 -2423 [25] -976 [27], -1487 [37] -1099

    b 941 0.2086 -464 [25] -949 [27], -1330 [37] -1017

    c 903 0.3116 -6100 [25] -5667 [37] -6454 (-5958)

    c 941 0.2995 -3310 [25] -5147 [37] -5519

    c 941 0.3 -4424 [26] -6734 [27] -5532

    c 941 0.3087 -3874 [25] -4218 [20], -5542 [37] -5749

    c 941 0.3116 -6088 [25] -5667 [37] -5818

    g-HT 723 0.338 -8000 [30] -6573 [27] -7444 (-7134)

    g-HT 723 0.352 -7548 [55] -7267 [37] -7242 (-7134)

    g-HT 723 0.358 -7780 [55] -7110 [27] -7156 (-7134)

    g-HT 723 0.365 -7700 [55] -7185 [27] -7193

    g-HT 723 0.375 -7700 [30] -7132 [27] -7245

    g-HT 723 0.377 -7470 [55] -7132 [27] -7229 (-7247)

    g-HT 723 0.39 -7531 [30], -7382 [55] -6839 [27], -6540 [20] -6936 (-7247)

    g-HT 773 0.352 -7550 [55] -6933 [27], -7055 [20] -7174 (-7097)

    g-HT 773 0.358 -7450 [55] -7074 [27], -7417 [37] -7114

    g-HT 773 0.365 -7420 [55] -7096 [27], -7504 [37] -7193

    g-HT 773 0.375 -7264 [55] -7463 [37] -7172 (-7238)

    g-HT 773 0.377 -7470 [55] -7048 [27] -7122 (-7238)

    g-HT 903 0.3431 -6636 [26] -6477 [27], -5837 [20], -6915 [37] -6762 (-7000)

    g-HT 903 0.3476 -6402 [26] -6591 [27], -5916 [20], -7113 [37] -6896 (-7000)

    For conditions in two phase regions, calculated equilibrium system values are listed first followed by metastable single-phase values in

    parenthesis

    3374 J Mater Sci (2016) 51:3362–3379

    123

  • during the selenization, reducing the reaction rate and

    ultimately relating to the void formation at the back contact

    that may increase the series resistance [4, 58, 71, 72].

    Notably, an overall metal precursor composition of Cu/

    III = 0.9 and Ga/III = 0.14 in practice would probably

    still not avoid these phenomena, because the preferential

    reaction of In with Se during initial selenization is likely to

    enrich the rear interface to Ga/III[ 0.14. It is thereforeimpractical to avoid c formation in equilibrated metalprecursors with compositions of interest for high-efficiency

    PV devices. However, it may be possible to avoid equili-

    bration—a previous report found that as-deposited Cu/Ga/

    In films exhibited reduced c content relative to CuGa/Infilms [79], while equilibration (by annealing) caused both

    films to form similar amounts of c. As less c formation wasassociated with better PV performance in that study, low

    energy precursor processes may be advantageous.

    The undesired formation of voids at the rear interface

    may also be related to the nearly ubiquitous agglomeration

    of indium at the surface. As the agglomeration forms

    nodules, it is also likely a source of lateral nonuniformity.

    Assuming either of these phenomena is a result of In-rich

    liquid, the fraction of overall indium in the system which is

    present in the liquid would give a measure of the likelihood

    of In surface agglomeration. The model predicts that no

    Ga/III precursor compositions or temperatures can avoid a

    significant equilibrium amount of liquid In, with the

    exception of Ga/III approaching unity (that is, composi-

    tions which are presently of diminished interest for high-

    efficiency PV).

    Calculated isopleths at three different constant Ga/III are

    shown in Fig. 14. At lower processing temperatures (such

    as those reached by an unheated substrate during

    sputtering), the equilibrium phases change from

    CuGa2 ? c ? (In) to c ? (In) to c ? g ? (In) withincreasing Cu/III or decreasing Ga/III. The

    c ? L $ CuGa2 ? (In) invariant reaction occurs at374.8 K, and the L $ c ? g ? (In) eutectic reactionoccurs at 427.7 K; both these transitions shift to greater Cu/

    III with increasing Ga/III. This is a 50 �C change in initialliquid formation temperature which is caused by small

    compositional changes.

    The calculated ternary isotherms at 423, 823, and 1023 K

    are in Figs. 15, 16, and 17, respectively. As can be seen by

    visual inspection of the isotherms, compositions with Cu/

    III & 0.9 and Ga/III & 0.3 (In & 37; Ga & 16 at.%)

    Fig. 15 Calculated Cu–Ga–In 423 K isotherm. Bold lines indicatebinary tie-line phase existence regions

    Fig. 14 Calculated Cu–Ga–In isopleths, or vertical sections atconstant Ga/III = 0.25 (light-gray solid lines), 0.30 (dark-gray

    dashed lines), and 0.35 (black-short dashed lines)

    Fig. 13 Calculated phase fractions of g (red) and c (blue) at constantCu/III = 0.9 and Ga/III = 0.2 (solid lines), Ga/III = 0.3 (dashed

    lines), and Ga/III = 0.4 (dotted line) as a function of temperature

    J Mater Sci (2016) 51:3362–3379 3375

    123

  • exist quite near to or overlapping phase boundaries from low

    temperature up to the liquidus temperature (Figs. 15, 16,

    17). Generally speaking, the In-poor c and In-rich (In)phases predominate, while the g phase with moderate Incontent is calculated to be present in smaller amounts. The

    CuGa2 phase also appears at low Cu/III and/or high Ga/III

    (see also Fig. 14), in agreement with its experimental

    observation in Cu437Ga388In175 films [80]. The L $ CuGa2,L $ (In), and L $ g equilibrium transitions all occurwithin typical processing temperature ranges (\873 K), andthe product liquid phase is always very In-rich (see tie-lines

    in Fig. 9). The likely presence of In-enriched liquid may

    relate to enhanced In mobility in the early stages of a sel-

    enization temperature ramp. Enhanced In mobility may

    permit rapid diffusion to the surface, where it selenizes first,

    in excellent agreement with a recent in situ XRD study of

    Cu–Ga–In selenization [81].

    At later stages of chalcogenization, however, a possible

    intermetallic phase transition at around 773 K has been

    hypothesized based on observance of a step increase for

    incorporation of Ga into chalcogenide phase(s), and there-

    fore apparent Ga mobility [71]. A subsequent study found a

    step increase in the extent of reaction of Ga with S at around

    803–813 K in a selenization process [82]. There are only

    three possible explanations for the observed jump in reaction

    extent with only slightly increasing temperature: (1) the

    thermodynamic-limited case where temperature changes the

    Gibbs energy of reaction, (2) the diffusion-limited case

    where temperature probably causes a phase change to greatly

    increase Ga mobility, and (3) the reaction kinetic-limited

    case where temperature causes a change in activation

    energy. The first two cases are most likely because the

    reaction of gases with condensed phases is usually diffusion

    limited, and thermodynamics determines the favorability of

    competing reaction pathways. The Ga-S thermodynamic

    system has not been assessed to the authors’ knowledge, so

    Gibbs energy calculations for the Cu–Ga–In–S system

    cannot presently be performed for the thermodynamic-lim-

    ited case. The diffusion-limited case, though, would quali-

    tatively agree with the Cu–Ga–In model. Assuming the back

    of the films are effectively still chalcogen free and have Cu/

    III & 0.9 and Ga/III & 0.3 [82], the c ? L $ c ? g ? Ltransition can be expected to occur around 773–813 K,

    depending on precise Cu/III and Ga/III. This precise com-

    position dependence could also explain why the step in

    reaction extent with temperature was originally observed at

    773 K [71] and subsequently observed at 803–813 K [82].

    Importantly, the c ? L $ c ? g ? L transition tempera-ture shown in Fig. 14 is calculated to vary greatly with only

    slight changes in both Cu/III and Ga/III composition. For

    example, changing overall Cu composition by ± 0.5 at.%

    about Cu/III = 0.9 at a constant Ga/III = 0.3 changes the

    c ? L $ c ? g ? L transition temperature from 708 to801 K. Furthermore, changing overall Ga composition

    by ± 0.5 at.% about Ga/III = 0.3 at a constant Cu/III = 0.9

    changes the temperature from 822 to 695 K. As the sel-

    enization process is most likely condensed phase diffusion

    limited, and diffusion through liquids is usually orders of

    magnitude faster than through solids, the anticipated variance

    in L $ g transition temperature could have drastic reactionrate implications. If film composition of all cations is not

    controlled to better than the ambitious target of ±0.5 at.%,

    then significant reaction rate inhomogeneity can be expected.

    This is evidence that the equilibrium reaction pathways fol-

    lowed for metal precursors commonly used in selenization

    Fig. 17 Calculated Cu–Ga–In 1023 K isotherm. Bold lines indicatebinary tie-line phase existence regions

    Fig. 16 Calculated Cu–Ga–In 823 K isotherm. Bold lines indicatebinary tie-line phase existence regions

    3376 J Mater Sci (2016) 51:3362–3379

    123

  • processes are inherently non-robust, that is, highly sensitive to

    slight compositional changes which approach the detection

    limits of typical measurement techniques.

    The vapor pressures of all metal species were calculated

    for an equimolar mixture of Cu–Ga–In (Fig. 18). A system

    of Cu0.9Ga0.3In0.7 has very similar calculated vapor pres-

    sures. The vapor pressures vary greatly, with PIn[ 10PGa[ 100 PCu[ 100 PIn2 at typical processing tempera-tures. As temperature increases, the In partial pressure

    increases more rapidly than that of Se (not shown). How-

    ever, only in the case of extreme processing temperatures

    would the In vapor pressure be relevant (e.g., at 680 �C(950 K) In is just five orders of magnitude lower than

    weighted Se partial pressure). While some studies found

    material loss of In/Ga by evaporation to be an issue during

    selenization [83], a subsequent report attributed the appar-

    ent material loss to a characterization artifact [84]. The

    equilibrium vapor pressures indicate that while cation

    material loss by pure element evaporation should indeed be

    negligible, material loss by metal selenide molecule evap-

    oration could nevertheless be important. Specifically, the

    vapor pressure of In2Se at processing temperatures is

    greater than In by an order of magnitude [85], or substan-

    tially more [86–88]. Similarly, the vapor pressure of Ga2Se

    at processing temperatures is greater than Ga [85, 89]. It is

    therefore likely that In and Ga loss occurs by In2Se and

    Ga2Se formation and evaporation during selenization.

    Conclusions

    Equilibrium in the Cu–Ga–In material system has been

    studied. High-temperature equilibration experiments have

    been performed that map the ternary phase diagram at two

    different compositions and at three temperatures, as well as

    thermal analysis for five compositions. A Cu–Ga–In ther-

    modynamic model has been assessed that fits the equilib-

    rium data and nonequilibrium thin film observations

    reasonably well, which may indicate that typical precursor

    films approach equilibrium. With the model and thermo-

    dynamic calculation software, calculations may be per-

    formed to predict thermodynamic reaction pathways. This

    could be quite useful in the development of metal precursor

    films for selenization processing of CIGS PV absorbers.

    Specifically, at typical compositions and processing tem-

    peratures, metal precursor films are expected to undergo a

    phase transition that is highly sensitive to slight composi-

    tion changes. Small compositional fluctuations could

    therefore cause lateral nonuniformity in chalcogenized

    absorbers. Equilibrated films are calculated to have high ccontent. Reducing film equilibration could therefore reduce

    c formation, which has been shown experimentally, andwas associated with better PV performance [79]. Extension

    of the model to diffusivity modeling could further assist in

    understanding selenization reaction pathways.

    Acknowledgements The authors gratefully acknowledge fundingfrom the U.S. Department of Energy under FPACE contract DE-

    EE0005407. The authors would like to thank GE for graciously

    providing bulk alloys and performing DSC, and Thermo-Calc Soft-

    ware and CompuTherm LLC for providing software. This study was

    funded by Department of Energy FPACE grant DE-EE0005407.

    Compliance with ethical standards

    Conflict of interest The authors declare that they have no conflictof interest.

    References

    1. Chen G, Ma J, Geng Z (2005) Fabrication and properties of lead-free

    Sn–Ag–Cu–Ga solder alloy. Mater Sci Forum 475–479:1747–1750.

    doi:10.4028/http://www.scientific.net/MSF.475-479.1747

    2. Gierlotka W (2012) Thermodynamic description of the quater-

    nary Ag–Cu–In–Sn system. J Electron Mater 41:86–108. doi:10.

    1007/s11664-011-1757-z

    3. Jackson P, Hariskos D, Wuerz R et al (2015) Properties of

    Cu(In,Ga)Se2 solar cells with new record efficiencies up to

    21.7%. Phys Status Solidi (RRL) Rapid Res Lett 9:28–31. doi:10.

    1002/pssr.201409520

    4. Kim K, Hanket GM, Huynh T, Shafarman WN (2012) Three-step

    H2Se/Ar/H2S reaction of Cu–In–Ga precursors for controlled

    composition and adhesion of Cu(In, Ga)(Se, S)2 thin films. J Appl

    Phys 111:83710

    5. Anderson TJ, Ansara I (1991) The Ga-In (gallium-indium) sys-

    tem. J Phase Equilib 12:64–72

    6. Andrews KW, Hume-Rothery W (1941) On the a/b brass type of

    equilibrium. Proc Royal Soc Lond Ser A 178:464–473

    7. Bahari Z, Dichi E, Legendre B (1999) Heat content and heat

    capacity of Cu0.7In0.3 from 298 to 1273 K. Z Metallkd 90:55–59

    8. Bahari Z, Dichi E, Legendre B, Dugué J (2003) The equilibrium

    phase diagram of the copper–indium system: a new investigation.

    Thermochim Acta 401:131–138

    Fig. 18 Calculated vapor pressures of Cu1, Cu2, Ga1, Ga2, In1, andIn2 as a function of temperature in an equimolar system of Cu–Ga–In

    (Cu1Ga1In1) at atmospheric pressure

    J Mater Sci (2016) 51:3362–3379 3377

    123

    http://dx.doi.org/10.4028/http://www.scientific.net/MSF.475-479.1747http://dx.doi.org/10.1007/s11664-011-1757-zhttp://dx.doi.org/10.1007/s11664-011-1757-zhttp://dx.doi.org/10.1002/pssr.201409520http://dx.doi.org/10.1002/pssr.201409520

  • 9. Bhattacharya D, Masson B (1974) Thermodynamic properties of

    a Cu–In alloys by the method of atomic absorption. Metall Trans

    5:1357–1361

    10. Bolcavage A, Chen SW, Kao CR, Chang YA, Romig AD (1993)

    Phase equilibria of the Cu–In system I: Experimental investiga-

    tion. J Phase Equilib 14:14–21. doi:10.1007/bf02652157

    11. Cao S, Huang S, Chu M, Yue Q, Shen J (2007) Thermodynamic

    optimization of Cu–In system. Xiyou Jinshu 31:807

    12. Chatterjee SC, Gupta MP (1975) Lattice parameters of some

    binary and ternary copper alloys. J Appl Crystallogr 8:492–493

    13. Che GC, Ellner M (1992) Powder crystal data for the high-tempera-

    ture phases Cu4In, Cu9In4(h) and Cu2In(h). Powder Diffr 7:107–108

    14. Chen JS, Kolawa E, Nicolet MA, Ruiz RP (1991) Cu/In deposited

    at room temperature: morphology, phases and reactions. Solar

    Cells 30:451–458

    15. Corderoy DJH, Honeycombe RWK (1963–1964) Age-hardening

    in copper-base indium alloys. J Inst Met 92: 65–69

    16. De M (1969) Thermal expansion of some Cu- and Ag-base

    alloys. Indian J Phys 43:367–376

    17. Dichi E, Legendre B (2000) Enthalpy of formation of Cu0.70In0.30and Cu0.64In0.36 phases by solution calorimetry. Z Metallkd

    91:47–50

    18. Elding-Pontén M, Stenberg L, Lidin S (1997) The h-phase field of

    the Cu–In system. J Alloys Compd 261:162–171

    19. Fournelle RA, Clark JB (1972) The genesis of the cellular pre-

    cipitation reaction. Metall Trans 3:2757–2767

    20. Hertz J, El Aissaoui K, Bouirden L (2002) A thermodynamic

    optimization of the Cu–In system. J Phase Equilib 23:473–479

    21. Hume-Rothery W, Raynor GV, Reynolds PW, Packer HK (1940)

    The constitution and structure of alloys of intermediate compo-

    sition in the systems copper–indium, copper–aluminum, copper–

    germanium, silver–aluminum, and silver–germanium. J Inst Met

    66:209–239

    22. Jacob KT, Alcock CB (1973) Activity of indium in a-solid

    solutions of Cu ? In, Au ? In and Cu ? Au ? In alloys. Acta

    Metall 21:1011–1016

    23. Jain KC, Ellner M, Schubert K (1972) Über die Phasen in der

    Nähe der Zusammensetzung Cu64In36. Zeitschrift für Metal-

    lkunde 63:456–461

    24. Jones RO, Owen EA (1953–1954) X-ray determination of the

    alpha-phase boundary of the copper–indium alloy system. J Inst

    Met 82:445–448

    25. Kang T, Castanet R (1977) Étude thermodynamique du systéme

    binaire cuivre-indium I. Étude calorimétrique. J Less Common

    Met 51(1):125–135

    26. Kang T, Kehiaian HV, Castanet R (1976) Étude thermody-

    namique du systéme binaire cuivre-indium I. Étude calorimé-

    trique. J Calorim Anal Therm 7:1

    27. Kao CR, Bolcavage A, Chen S-L, Chen SW, Chang YA, Romig

    AD (1993) Phase equilibria of the Cu–In system II: thermody-

    namic assessment and calculation of phase diagram. J Phase

    Equilib 14:22–30. doi:10.1007/bf02652158

    28. Keppner W, Klas T, Körner W, Wesche R, Schatz G (1985)

    Compound formation at Cu–In thin-film interfaces detected by

    perturbed g-g angular correlations. Phys Rev Lett 54:2371

    29. Kirchner HOK, Mellor BG, Chadwick GA (1978) A calorimetric

    determination of the interfacial enthalpy of Cu–In and Cu–Al

    lamellar eutectoids. Acta Metall 26:1023–1031

    30. Kleppa OJ (1956) Heat of formation of solid and liquid binary

    alloys of copper with cadmium, indium, tin and antimony at 450�.J Phys Chem 60:852–858. doi:10.1021/j150541a005

    31. Knott S, Mikula A (2006) Calorimetric investigation of the binary

    Cu–In system. Int J Mater Res 97:1098–1101

    32. Koster AS, Wolff LR, Visser GJ (1980) Structure of copper–

    indium Cu7In3. Acta Crystallogr Sect B 36:3094–3096. doi:10.

    1107/s0567740880010886

    33. Kutsenok IB, Geiderikh VA, Gerasimov YI, Yalkanen KK (1983)

    Study of the thermodynamic properties of indium–copper alloys

    by an instantaneous EMF recording method. Russ J Phys Chem

    57:1639–1641

    34. Li J-B, Ji L, Liang JK et al (2008) A thermodynamic assessment

    of the copper–gallium system. Calphad 32:447–453. doi:10.1016/

    j.calphad.2008.03.006

    35. Lidin S, Larsson A-K (1995) A survey of superstructures in

    intermetallic NiAs-Ni2In-type phases. J Solid State Chem

    118:313–322

    36. Lidin S, Stenberg L, Elding-Pontén M (1997) The B8 type

    structure of Cu7In3. J Alloys Compd 255:221–226

    37. Liu HS, Liu XJ, Cui Y et al (2002) Thermodynamic assessment

    of the Cu–In binary system. J Phase Equilib 23:409–415

    38. Liu X, Liu H, Ohnuma I et al (2001) Experimental determination

    and thermodynamic calculation of the phase equilibria in the Cu–

    In–Sn system. J Electron Mater 30:1093–1103

    39. Masson DB, Bhattacharya D (1975) The effect of excess entropy

    on the activity coefficient of indium in a Cu–In alloys. Scripta

    Metall 9:207–210

    40. Mizutani U, Noritake T, Ohsuna T, Takeuchi T (2010) Hume-

    Rothery electron concentration rule across a whole solid solution

    range in a series of gamma-brasses in Cu–Zn, Cu–Cd, Cu–Al,

    Cu–Ga, Ni–Zn and Co–Zn alloy systems. Philos Mag

    90:1985–2008. doi:10.1080/14786430903246320

    41. Muschik T, Hehenkamp T (1987) Fest-flüssig-Gleichgewichte

    auf der Cu-Seite des Systems Cu–In. Metallkd 78:358–361

    42. Orbey N, Birkmire R, Russell T, Jones G (2000) Copper–indium

    alloy transformations. J Phase Equilib 21:509–513. doi:10.1007/

    s11669-000-0018-2

    43. Owen EA, Morris DP (1949–1950) The application of X-ray

    methods to the determination of phase boundaries in metallur-

    gical equilibrium diagrams. J Inst Met 76(2):145–168

    44. Owen EA, O’Donnell Roberts EA (1952–1953) The solubility of

    indium in copper. J Inst Met 81:479–480

    45. Owen EA, Roberts EW (1939) XXIX Factors affecting the limit

    of solubility of elements in copper and silver. Philos Mag Ser 7

    27:294–327. doi:10.1080/14786443908562230

    46. Piao S, Lidin S (2008) A new compound in the Cu–In system—

    the synthesis and structure of Cu10In7. Zeitschrift für anorganis-

    che und allgemeine Chemie 634:2589–2593. doi:10.1002/zaac.

    200800275

    47. Rajasekharan TP, Schubert K (1981) Kristallstruktur von Cu11In9. Z Metallkd 72:275–278

    48. Ramos de Debiaggi S, Cabeza GF, Toro CD, Monti AM, Som-

    madossi S, Guillermet AF (2011) Ab initio study of the structural,

    thermodynamic and electronic properties of the Cu10In7 inter-

    metallic phase. J Alloys Compd 509:3238–3245

    49. Ramos de Debiaggi S, Deluque Toro C, Cabeza GF, Fernández

    Guillermet A (2012) Ab initio comparative study of the Cu–In

    and Cu–Sn intermetallic phases in Cu–In–Sn alloys. J Alloys

    Compd 542:280–292

    50. Reddy SR, Hajra JP (1993) Thermodynamics and phase equilibria

    in the system Ga-In using multi-parameter functions. Calphad

    17:151–156

    51. Reynolds J, Wiseman WA, Hume-Rothery W (1951–1952) The

    equilibrium diagram of the system copper–indium in the region

    25–35 at.% indium. J Inst Met 80:637–640

    52. Stirling PH, Raynor GV (1955–1956) The copper-rich alloys of

    the system copper–aluminum–indium. J Inst Met 84:57–65

    53. Straumanis ME, Yu LS (1969) Lattice parameters, densities,

    expansion coefficients and perfection of structure of Cu and of

    Cu–In a phase. Acta Crystallogr A 25:676–682

    54. Subramanian PR, Laughlin DE (1989) The Cu–In (Copper–

    Indium) system. J Phase Equilib 10:554–568. doi:10.1007/

    bf02882415

    3378 J Mater Sci (2016) 51:3362–3379

    123

    http://dx.doi.org/10.1007/bf02652157http://dx.doi.org/10.1007/bf02652158http://dx.doi.org/10.1021/j150541a005http://dx.doi.org/10.1107/s0567740880010886http://dx.doi.org/10.1107/s0567740880010886http://dx.doi.org/10.1016/j.calphad.2008.03.006http://dx.doi.org/10.1016/j.calphad.2008.03.006http://dx.doi.org/10.1080/14786430903246320http://dx.doi.org/10.1007/s11669-000-0018-2http://dx.doi.org/10.1007/s11669-000-0018-2http://dx.doi.org/10.1080/14786443908562230http://dx.doi.org/10.1002/zaac.200800275http://dx.doi.org/10.1002/zaac.200800275http://dx.doi.org/10.1007/bf02882415http://dx.doi.org/10.1007/bf02882415

  • 55. Vinokurova GA, Geiderikh VA (1976) Thermodynamic proper-

    ties of indium–copper alloys. Russ J Phys Chem 50:1661

    56. Wallbrecht PC, Blachnik R, Mills KC (1981) The heat capacity

    and enthalpy of some hume-rothery phases formed by copper,

    silver and gold. Part III. Cu ? Ga, Ag ? Ga, Au ? Ga, Cu ? In,

    Ag ? In and Au ? In systems. Thermochim Acta 48:69–82

    57. Weibke F, Eggers H (1934) Das Zustandsdiagramm des Systems

    Kupfer-Indium. Zeitschrift für anorganische und allgemeine

    Chemie 220:273–292

    58. Hanket GM, Shafarman WN, McCandless BE, Birkmire RW

    (2007) Incongruent reaction of Cu–(InGa) intermetallic precur-

    sors in H2Se and H2S. J Appl Phys 102(07):4922

    59. Kim WK, Shen J, Chu M, Jung JH, Anderson TJ (2012) Pre-

    diction of the Cu–Ga–In ternary phase diagram. J Nanoelectron

    Optoelectron 7:425–429

    60. Purwins M, Enderle R, Schmid M et al (2007) Phase relations in

    the ternary Cu–Ga–In system. Thin Solid Films 515:5895–5898

    61. Dinsdale AT (1991) SGTE data for pure elements. Calphad

    15:317–425

    62. Schmid-Fetzer R, Andersson D, Chevalier PY et al (2007)

    Assessment techniques, database design and software facilities

    for thermodynamics and diffusion. Calphad 31:38–52. doi:10.

    1016/j.calphad.2006.02.007

    63. Lukas HL, Fries SG, Sundman B (2007) Computational ther-

    modynamics—the calphad method. Cambridge University Press,

    Cambridge

    64. Hillert M, Jansson B, Sundman B, Ågren J (1985) A two-sub-

    lattice model for molten solutions with different tendency for

    ionization. Metall Mater Trans A 16:261–266. doi:10.1007/

    bf02815307

    65. Redlich O, Kister AT (1948) Algebraic representation of ther-

    modynamic properties and the classification of solutions. Ind Eng

    Chem 40:345–348

    66. Brummer A, Honkimäki V, Berwian P, Probst V, Palm J, Hock R

    (2003) Formation of CuInSe2 by the annealing of stacked ele-

    mental layers–analysis by in situ high-energy powder diffraction.

    Thin Solid Films 437:297–307

    67. Gupta A, Isomura S (1998) Precursor modification for prepara-

    tion of CIS films by selenization technique. Solar Energy Mater

    Solar Cells 53:385–401

    68. Gupta A, Shirakata S, Isomura S (1994) Studies on CuIn pre-

    cursor for the preparation of CuInSe2 thin films by the seleniza-

    tion technique. Solar Energy Mater Solar Cells 32:137–149

    69. Prepelita P, Medianu R, Iacomi F, Sandu I (2011) Physico-

    chemical properties of CuInGa-ZnS heterostructure deposited.

    Rev Chim 62:905–907

    70. Alberts V (2007) A comparison of the material and device

    properties of homogeneous and compositional-graded Cu(In,

    Ga)(Se, S)2 chalcopyrite thin films. Semicond Sci Technol

    22:585

    71. Hanket GM, Kamada R, Kim WK, Shafarman WN (2008) Effect

    of reaction temperature on Cu(InGa)(SeS)2 formation by a

    sequential H2Se/H2S precursor reaction process. Photovoltaic

    Specialists Conference (PVSC), 33rd IEEE: 1. Doi:10.1109/pvsc.

    2008.4922470

    72. Huynh T, Buchanan W, Kim K, McCandless B, Shafarman W,

    Birkmire R (2011) Cu–In–Ga metal precursors sputter deposited

    from a single ternary target for Cu(lnGa)(SeS)2 film formation.

    Photovoltaic Specialists Conference (PVSC), 37th IEEE: 382.

    Doi:10.1109/pvsc.2011.6185974

    73. Krishnan R (2012) Rapid routes for synthesis of CIGS absorbers.

    Dissertation, University of Florida

    74. Chen GS, Yang JC, Chan YC, Yang LC, Huang W (2009) Another

    route to fabricate single-phase chalcogenides by post-selenization

    of Cu–In–Ga precursors sputter deposited from a single ternary

    target. Solar Energy Mater Solar Cells 93:1351–1355

    75. Haug V, Klugius I, Friedlmeier TM, Quintilla A, Ahlswede E

    (2011) Cu(In, Ga)Se2 thin-film solar cells based on a simple

    sputtered alloy precursor and a low-cost selenization step.

    J Photon Energy 1:018002

    76. Su C-Y, Ho W-H, Lin H-C, Nieh C-Y, Liang S-C (2011) The

    effects of the morphology on the CIGS thin films prepared by

    CuInGa single precursor. Solar Energy Mater Solar Cells

    95:261–263

    77. Song HK, Jeong JK, Kim HJ, Kim SK, Yoon KH (2003) Fabri-

    cation of CuIn1-xGaxSe2 thin film solar cells by sputtering and

    selenization process. Thin Solid Films 435:186–192

    78. Keene BJ (1993) Review of data for the surface tension of pure

    metals. Int Mater Rev 38:157–192

    79. Muzzillo CP, Mansfield LM, DeHart C, et al (2015) Differences

    between CuGa/In and Cu/Ga/In films for selenization. Photo-

    voltaic Specialist Conference (PVSC), 2015 IEEE 42nd: in press

    80. Muzzillo CP, Mansfield LM, Dehart C, et al (2014) The effect of

    Ga content on the selenization of co-evaporated CuGa/In films

    and their photovoltaic performance. Photovoltaic Specialist

    Conference (PVSC), 2014 IEEE 40th: 1649. Doi:10.1109/PVSC.

    2014.6925236

    81. Mainz R, Weber A, Rodriguez-Alvarez H, et al (2014) Time-

    resolved investigation of Cu(In,Ga)Se2 growth and Ga gradient

    formation during fast selenisation of metallic precursors. Progress

    in Photovoltaics: Research and Applications in press. Doi:10.

    1002/pip.2531

    82. Kim WK, Hanket GM, Shafarman WN (2011) Ga homogeniza-

    tion by simultaneous H2Se/H2S reaction of Cu–Ga–In precursor.

    Solar Energy Mater Solar Cells 95:235–238

    83. Kushiya K, Shimizu A, Yamada A, Konagai M (1995) Devel-

    opment of high-efficiency CuInxGa1-xSe2 thin-film solar cells by

    selenization with elemental Se vapor in vacuum. Jpn J Appl Phys

    34:54–60

    84. Klenk M, Schenker O, Alberts V, Bucher E (2001) Control of

    two-step growth processes of chalcopyrite thin films by X-ray

    fluorescence spectroscopy. Appl Surf Sci 173:62. doi:10.1016/

    S0169-4332(00)00876-X

    85. Berger LI, Strel’chenko SS, Bondar SA, Molodyk AD, Bala-

    nevskaya AE, Levedev VV (1969) ladkeybe lbccowbawbb ldo-qysx gokygpodolybrodsx coelbyeybq nbga BIII2CVI3. IzvestiyaAkademii nauk SSSR. Neorg Mater 5:872

    86. Brebrick RF, Su C-H (2002) Partial pressures for several In-Se

    compositions from optical absorbance of the vapor. J Phase

    Equilib 23:397–408. doi:10.1361/105497102770331343

    87. Colin R, Drowart J (1968) Mass spectrometric determination of

    dissociation energies of gaseous indium sulphides, selenides and

    tellurides. Trans Faraday Soc 64:2611–2621. doi:10.1039/

    tf9686402611

    88. Greenberg JH, Borjakova VA, Shevelkov VF (1973) Thermo-

    dynamic properties of In2Se. J Chem Thermodyn 5:233–237.

    doi:10.1016/S0021-9614(73)80083-7

    89. Piacente V, Bardi G, Di Paolo V, Ferro D (1976) The vapour

    pressure over Ga2S2 and Ga2Se2. J Chem Thermodyn 8:391–401.

    doi:10.1016/0021-9614(76)90080-X

    J Mater Sci (2016) 51:3362–3379 3379

    123

    http://dx.doi.org/10.1016/j.calphad.2006.02.007http://dx.doi.org/10.1016/j.calphad.2006.02.007http://dx.doi.org/10.1007/bf02815307http://dx.doi.org/10.1007/bf02815307http://dx.doi.org/10.1109/pvsc.2008.4922470http://dx.doi.org/10.1109/pvsc.2008.4922470http://dx.doi.org/10.1109/pvsc.2011.6185974http://dx.doi.org/10.1109/PVSC.2014.6925236http://dx.doi.org/10.1109/PVSC.2014.6925236http://dx.doi.org/10.1002/pip.2531http://dx.doi.org/10.1002/pip.2531http://dx.doi.org/10.1016/S0169-4332(00)00876-Xhttp://dx.doi.org/10.1016/S0169-4332(00)00876-Xhttp://dx.doi.org/10.1361/105497102770331343http://dx.doi.org/10.1039/tf9686402611http://dx.doi.org/10.1039/tf9686402611http://dx.doi.org/10.1016/S0021-9614(73)80083-7http://dx.doi.org/10.1016/0021-9614(76)90080-X

    Cu--Ga--In thermodynamics: experimental study, modeling, and implications for photovoltaicsAbstractIntroductionExperimentalSample preparation and alloy annealingAlloy characterization

    Experimental resultsXRDSEM/EDSDTA/DSC

    Thermodynamic assessmentThermodynamic literatureThermodynamic modelsPure elementsStoichiometric compoundsSolution phases

    Results and discussionCu--In assessmentCu--Ga--In assessmentPractical applications

    ConclusionsAcknowledgementsReferences