Open access

Calorimetric Determination of Heat Capacity, Entropy and Enthalpy of Mixed Oxides in the System CaO–SrO–Bi2O3–Nb2O5–Ta2O5

Written By

Jindřich Leitner, David Sedmidubský, Květoslav Růžička and Pavel Svoboda

Submitted: 11 November 2011 Published: 23 January 2013

DOI: 10.5772/54064

Chapter metrics overview

3,536 Chapter Downloads

View Full Metrics

1. Introduction

Mixed oxides in the system CaO–SrO–Bi2O3–Nb2O5–Ta2O5 possess many extraordinary electric, magnetic and optical properties for which they are used in fabrication of various electronic components. For example Sr2(Nb,Ta)2O7 and (Sr,Ca)Bi2(Nb,Ta)2O9 are used for ferroelectric memory devices, CaNb2O6, Sr5(Nb1–xTax)4O15 and Bi(Nb,Ta)O4 for microwave dielectric resonators and Ca2Nb2O7 as non-linear optical materials and hosts for rare-earth ions in solid-state lasers. Ternary strontium bismuth oxides SrBi2O4, Sr2Bi2O5, and Sr6Bi2O9 are of considerable interest due to a visible light driven fotocatalytic activity.

To assess the thermodynamic stability and reactivity of these oxides under various conditions during their preparation, processing and operation, a complete set of consistent thermodynamic data, including heat capacity, entropy and enthalpy of formation, is necessary. Some of these data are available in literature. Akishige et al. [1] have been measured the heat capacities of Sr2Nb2O7 and Sr2Ta2O7 single crystals in the temperature range 2-600 K. The results have been only plotted and the values of S°m(298) have not been calculated. A commensurate transformation of Sr2Nb2O7 at TINC = 495 K has been observed accompanied by changes in enthalpy and entropy of ΔH = 291 J mol-1 and ΔS = 0.587 J K-1 mol-1. The heat capacity of Sr2Nb2O7 has been also measured by Shabbir at al. [2] in the temperature range 375-575 K. They have observed a phase transition at TINC = 487 ± 2 K connected with ΔH = 147 ± 14 J mol-1 and ΔS = 0.71 ± 0.10 J K-1 mol-1. The heat capacities of polycrystalline and monocrystalline SrBi2Ta2O9 and Sr0,85Bi2,1Ta2O9 have been measured by Onodera at al. [3–5] at 80-800 K. Morimoto at al. [6] have reported the results of the heat capacity measurements of SrBi2(NbxTa1-x)2O9 (x = 0, 1/3, 2/3 a 1). The temperature dependences of heat capacities show lambda-transitions with maxima at the Currie temperature TC = 570 ± 1 K, 585 ± 2 K, 625 ± 3 K a 690 ± 2 K for x = 0, 1/3, 2/3 and 1, respectively. Using EMF (electromotive force) measurements, Raghavan has obtained the values of the Gibbs energy of formation from binary oxides, ΔoxG, for some niobates [7,8] and tantalates [9,10] of calcium. His results are summarized in Table 1. The same technique has been employed by Dneprova et al. [11] for ΔoxG measurement for CaNb2O6 and Ca2Nb2O7. Their results presented in Table 1 are not significantly different from the results of Raghavan. Using the CALPHAD approach [13], Yang et al. [14] have assessed thermodynamic data for various mixed oxides in the SrO–Nb2O5 system. The same approach has been used by Hallstedt et al. for the assessment of thermodynamic data of mixed oxides in the systems CaO–Bi2O3 [14] and SrO–Bi2O3 [15]. Besides equilibrium data, values of the enthalpy of formation [16] of mixed oxide have been considered. Later on, these systems have been studied by EMF method by Jacob and Jayadevan [17,18] and temperature dependences of ΔoxG for various mixed oxides have been derived. These data have been included into the thermodynamic re-assessment of the CaO-SrO-Bi2O3 system [19].

OxideΔoxG
(kJ mol–1)
T
(K)
ΔoxH
(kJ mol–1)
ΔoxS
(J K-1 mol–1)
Ref.
CaNb2O6–75.82 – 0.03345T1245-1300–75.8233.45[7]
Ca2Nb2O7–178.441256[8]
Ca3Nb2O8–209.941256[8]
CaTa4O11–36.982 – 0.029T1250-1300–36.9829.0[9]
CaTa2O6–65.141250[10]
Ca2Ta2O7–102.821250[10]
Ca4Ta2O9–165.051250[10]
CaNb2O6–175.73 + 0.02259T1100-1276–175.73–22.59[11]
Ca2Nb2O7–212.54 – 0.02218T1100-1350–212.5422.18[11]
Sr2Nb10O27–1125.69 + 0.35069T298-5000–1125.69–350.69[12]
SrNb2O6325.04 + 0.05865T298-5000–325.04–58.65[12]
Sr2Nb2O7–367.43 + 0.03993T298-5000–367.43–39.93[12]
Sr5Nb4O14–746.72 + 0.05101T298-5000–746.72–51.01[12]
Ca5Bi14O26–125.90 – 0.055T298-1300–125.955.0[19]
CaBi2O4–27.60 – 0.003T298-1300–27.63.0[19]
Ca4Bi6O13–97.60 – 0.008T298-1300–97.68.0[19]
Ca2Bi2O5–42.20 – 0.003T298-1300–42.23.0[19]
SrBi2O4–63.86 – 0.0018T298-1300–63.861.8[19]
Sr2Bi2O5–118.75 + 0.024T298-1300–118.75–24.0[19]
Sr3Bi2O6–109.60 + 0.0024T298-1300–109.60–2.4[19]

Table 1.

Published values of ΔoxG, ΔoxH a ΔoxS for some mixed oxides in the system CaO-SrO-Bi2O3-Nb2O5-Ta2O5

This review brings a summary of our results [20–30] focused on calorimetric determination of heat capacity, entropy end enthalpy of mixed oxides in the system CaO–SrO–Bi2O3–Nb2O5–Ta2O5. Temperature dependences of molar heat capacity in a broad temperature range were evaluated from the experimental heat capacity and relative enthalpy data. Molar entropies at T = 298.15 K were calculated from low temperature heat capacity measurements. Furthermore, the results of calorimetric measurements of the enthalpies of drop-solution in a sodium oxide-molybdenum oxide melt for several stoichiometric mixed oxides in the above mentioned system are reported from which the values of enthalpy of formation from constituent binary oxides were derived. Finally, some empirical estimation and correlation methods (the Neumann-Kopp’s rule, entropy-volume correlation and electronegativity-differences method) for evaluation of thermodynamic data of mixed oxides are tested and assessed.

Advertisement

2. Experimental

Nineteen mixed oxides in the system CaO–SrO–Bi2O3–Nb2O5–Ta2O5 with stoichiometry CaBi2O4, Ca4Bi6O13, Ca2Bi2O5, SrBi2O4, Sr2Bi2O5, CaNb2O6, Ca2Nb2O7, SrNb2O6, Sr2Nb2O7, Sr2Nb10O27, Sr5Nb4O15, BiNbO4, BiNb5O14, BiTaO4, Bi4Ta2O11, Bi7Ta3O18, Bi3TaO7, SrBi2Nb2O9, and SrBi2Ta2O9 were prepared, characterized and examined. The samples were prepared by conventional solid state reactions from high purity precursors (CaCO3, SrCO3 Bi2O3, Nb2O5 and Ta2O5). A three step procedure was used consisting of an initial calcination run of mixed powder precursors and subsequent double firing of prereacted mixtures pressed into pellets. The phase composition of the prepared samples was checked by X-ray powder diffraction (XRD). XRD data were collected at room temperature with an X’Pert PRO (PANalytical, the Netherlands) θ-θ powder diffractometer with parafocusing Bragg-Brentano geometry using CuKα radiation (λ = 1.5418 nm). Data were scanned over the angular range 5–60° (2θ) with an increment of 0.02° (2θ) and a counting time of 0.3 s step–1. Data evaluation was performed by means of the HighScore Plus software.

The PPMS equipment 14 T-type (Quantum Design, USA) was used for the heat capacity measurements in the low temperature region [31-35]. The measurements were performed by the relaxation method [36] with fully automatic procedure under high vacuum (pressure ~10–2 Pa) to avoid heat loss through the exchange gas. The samples were compressed powder pellets. The densities of the samples were about 65 % of the theoretical ones.

The samples were mounted to the calorimeter platform with cryogenic grease Apiezon N (supplied by Quantum Design). The procedure was as follows: First, a blank sample holder with the Apiezon only was measured in the temperature range approx. 2–280 K to obtain background data, then the sample plate was attached to the calorimeter platform and the measurement was repeated in the same temperature range with the same temperature steps. The sample heat capacity was then obtained as a difference between the two data sets. This procedure was applied, because the heat capacity of Apiezon is not negligible in comparison with the sample heat capacity (~8 % at room temperature) and exhibits a peak-shaped transition below room temperature [37]. The manufacturer claims the precision of this measurement better then 2 % [38]; the control measurement of the copper sample (99.999 % purity) confirmed this precision in the temperature range 50–250 K. However, the precision of the measurement strongly depends on the thermal coupling between the sample and the calorimeter platform. Due to unavoidable porosity of the sample plate this coupling is rapidly getting worse as the temperature raises above 270 K and Apiezon diffuses into the porous sample. Consequently, the uncertainty of the obtained data tends to be larger.

A Micro DSC III calorimeter (Setaram, France) was used for the heat capacity determination in the temperature range of 253–352 K. First, the samples were preheated in a continuous mode from room temperature up to 352 K (heating rate 0.5 K min–1). Then the heat capacity was measured in the incremental temperature scanning mode consisting of a number of 5–10 K steps (heating rate 0.2 K min–1) followed by isothermal delays of 9000 s. Two subsequent step-by-step heating were recorded for each sample. Synthetic sapphire, NIST Standard reference material No. 720, was used as the reference. The uncertainty of heat capacity measurements is estimated to be better than ±1 %.

Enthalpy increment determinations were carried out by drop method using high-temperature calorimeter, Multi HTC 96 (Setaram, France). All measurements were performed in air by alternating dropping of the reference material (small pieces of synthetic sapphire, NIST Standard reference material No. 720) and of the sample (pressed pellets 5 mm in diameter) being initially held at room temperature, through a lock into the working cell of the preheated calorimeter. Endothermic effects are detected and the relevant peak area is proportional to the heat content of the dropped specimen. The delays between two subsequent drops were 25–30 min. To check the accuracy of measurement, the enthalpy increments of platinum in the temperature range 770–1370 K were measured first and compared with published reference values [39]. The standard deviation of 22 runs was 0.47 kJ mol–1, the average relative error was 2.0 %. Estimated overall accuracy of the drop measurements is ±3 %.

The heats of drop-solution were determined using a Multi HTC 96 high-temperature calorimeter (Setaram, France). A sodium oxide-molybdenum oxide melt of the stoichiometry 3Na2O + 4MoO3 was used as the solvent. The ratio of solute/solvent varied from 1/250 up to 1/500. The measurements were performed at temperatures of 973 and 1073 K in argon or air atmosphere. The method consists in alternating dropping of the reference material (small spherules of pure platinum) and of the sample (small pieces of pressed tablets 10–40 mg), being initially held near room temperature (T0), through a lock into the working cell (a platinum crucible with the solvent) of the preheated calorimeter at temperature T. Two or three samples were examined during one experimental run. The delays between two subsequent drops were 30–60 min. The total heat effect (ΔdsH) includes the heat of solution (ΔsolH), the heat content of the sample (ΔTH), and, for the carbonates, the heat of decomposition (ΔdecompH) to form solid CaO or SrO and gaseous CO2. Using appropriate thermochemical cycles, the values of the enthalpy of formation of mixed oxides from the binary oxides and from the elements at 298 K were evaluated. The temperature dependence of the heat capacity of platinum [39] was used for the calculation of the sensitivity of the calorimeters.

2.1. Characterization of prepared samples

The XRD analysis revealed that the prepared samples were without any observable diffraction lines from unreacted precursors or other phases. The lattice parameters of the oxides were evaluated by Rietveld refinement [40] and are summarized in Table 2 together with the values of theoretical density calculated from the lattice parameters.

2.2. Evaluation of temperature dependence of heat capacity at low temperatures

The fit of the low-temperature heat capacity data (LT fit) consists of two steps. Assuming the validity of the phenomenological formula Cpm = βT 3 + γelT, at T → 0 where β is proportional to the inverse cube root of the Debye temperature ΘD and γelT is the Sommerfeld term, we plotted the Cpm/T vs. T 2 dependence for T < 8 K to estimate the ΘD and γel values. Since all compounds under study are semiconductors with a sufficiently large band gap, the non-zero γel values are supposed to be either due to some metallic impurities or to a series of Schottky-like transitions resulting from structure defects. Nevertheless, they are negligible in most cases (typically < 0.5 mJ K–2 mol–1) and can be ignored in further analysis. As an example, the results of heat capacity measurements on CaNb2O6 and LT fit for T < 10 K is shown in Fig. 1.

Oxidea (nm)b (nm)c (nm)α (°)β (°)γ (°)d (g cm–3)Ref.
CaBi2O41.661431.157811.3991590134.03906.631[20]
Ca4Bi6O130.593081.735120.721929090906.540[20]
Ca2Bi2O51.010741.012491.04618116.88107.1692.986.468[20]
SrBi2O41.926350.434370.614449095.50907.392[29]
Sr2Bi2O51.429350.617150.764789090906.628[29]
CaNb2O61.496980.574720.522029090904.760[26]
Ca2Nb2O70.768531.335870.54959909098.294.496[26]
SrNb2O60.772090.559301.098219090.37905.174[24]
Sr2Nb2O70.395442.677350.570049090905.206[26]
Sr2Nb10O273.7153.6970.39439090905.653a)
Sr5Nb4O150.565760.565761.1453690901205.490[27]
BiNbO40.568931.17280.499159090907.297[21]
BiNb5O141.767621.720720.396109090904.948b)
BiTaO40.563941.17760.496269090909.149[21]
Bi4Ta2O110.661590.765280.98781101.3990.1089.999.306[28]
Bi7Ta3O183.401620.760540.6635490109.16909.395[28]
Bi3TaO70.547110.547110.547119090909.327[28]
SrBi2Nb2O90.551600.550872.510209090907.275[22]
SrBi2Ta2O90.552240.552662.501249090908.801[22]

Table 2.

Structural characterization of prepared samples

In the second step of the LT fit, both sets of the Cpm data (relaxation time + DSC) were considered. Analysis of the phonon heat capacity was performed as an additive combination of Debye and Einstein models. Both models include corrections for anharmonicity, which is responsible for a small, but not negligible, additive term at higher temperatures and which accounts for the difference between isobaric and isochoric heat capacity. According to literature [41], the term 1/(1 – αT ) is considered as a correction factor.

The acoustic part of the phonon heat capacity is described using the Debye model

CphD=9R1αDT(TΘD)30xDx4exp(x)[exp(x)1]2dxE1

where R is the gas constant, ΘD is the Debye characteristic temperature, αD is the coefficient of anharmonicity of acoustic branches and xD = ΘD/T. Here the three acoustic branches are taken as one triply degenerate branch. Similarly, the individual optical branches are described by the Einstein model

Figure 1.

Temperature dependence of Cpm/T function for CaNb2O7 at low temperatures

CphEi=R1αEiTxEi2exp(xEi)[exp(xEi)1]2E2

where αEi and xEi = ΘEi/T have analogous meanings as in the previous case. Several optical branches are again grouped into one degenerate multiple branch with the same Einstein characteristic temperature and anharmonicity coefficient. The phonon heat capacity then reads

Cph=CphD+i=13n3CphEiE3

All the estimated values were further treated by a simplex routine and a full non-linear fit was performed on all adjustable parameters.

The values of relative enthalpies at 298.15 K, Hm(298.15) – Hm(0), were evaluated from the low-temperature Cpm data (LT fit) by numerical integration of the Cpm(T) dependences from zero to 298.15 K. Standard deviations (2σ) were calculated using the error propagation law. The values of standard molar entropies at 298.15 K, Sm(298.15), were derived from the low-temperature Cpm data (LT fit) by numerical integration of the Cpm(T)/T dependences from zero to 298.15 K. A numerical integration was used with the boundary conditions Sm = 0 and Cpm = 0 at T = 0 K. Standard deviations (2σ) were calculated using the error propagation law. All calculated values are summarized in Table 3.

OxideCpm(298)
(J K–1 mol–1)
Hm(298)–Hm(0)
(J mol–1)
Sm(298)
(J K–1 mol–1)
ΔoxS
(J K-1 mol–1)
Ref.
CaBi2O4151.326470 ± 158188.5 ± 3.31.9[20]
Ca4Bi6O13504.185079 ± 507574.1 ± 8.8–23.8[20]
Ca2Bi2O5197.433735 ± 201231.3 ± 2.96.6[20]
SrBi2O4155.629601 ± 169206.1 ± 1.14.0[29]
Sr2Bi2O5201.938199 ± 219261.2 ± 1.45.5[29]
CaNb2O6171.828159 ± 170167.3 ± 0.9–8.1[26]
Ca2Nb2O7218.135631 ± 215212.4 ± 1.2–1.1[26]
SrNb2O6170.228722 ± 174173.9 ± 0.9–17.0[24]
Sr2Nb2O7216.637977 ± 266238.5 ± 1.3–5.9[26]
Sr2Nb10O27746.8124150 ± 740759.7 ± 4.1–33.9[27]
Sr5Nb4O15477.283340 ± 490524.5 ± 2.8–18.4[27]
BiNbO4121.322120 ± 134147.9 ± 0.85.0[21]
BiNb5O14386.862639 ± 362397.2 ± 2.1–25.8[23]
BiTaO4119.322021 ± 132149.1 ± 0.83.3[21]
Bi4Ta2O11363.266566 ± 384449.6 ± 2.39.5[28]
Bi7Ta3O18602.7109760 ± 634743.0 ± 3.88.6[28]
Bi3TaO7235.244265 ± 254304.3 ± 1.610.0[28]
SrBi2Nb2O9286.449230 ± 292327.2 ± 1.7–12.2[22]
SrBi2Ta2O9286.649060 ± 289339.2 ± 1.8–5.9[22]

Table 3.

Heat capacity, relative enthalpy, entropy and entropy of formation from binary oxides at temperature 298.15 K of various mixed oxides in the system CaO–SrO–Bi2O3–Nb2O5–Ta2O5

A comparison is given in Table 4 of the values of entropy of formation from binary oxides ΔoxS at 298 K calculated from our results and those from literature. The values of ΔoxS are calculated using the relation

ΔoxS=Sm(MO)ibiSm(BO,i)E4

where Sm(MO) and Sm(BO,i) stand for the molar entropies of a mixed oxide and a binary oxide i, respectively, and bi is a constitution coefficient representing the number of formula units of a binary oxide i per formula unit of the mixed oxide. The following values were used for calculation: Sm(CaO,298.15 K) = 38.1 J K–1 mol–1 [42], Sm(SrO,298.15 K) = 53.58 J K–1 mol–1 [43], Sm(Bi2O3,298.15 K) = 148.5 J K–1 mol–1 [44], Sm(Nb2O5, 298.15) = 137.30 J K–1 mol–1 [45] Sm(Ta2O5, 298.15) = 143.09 J K–1 mol–1 [45]. Furthermore, Sm(Sr2Nb2O7, 298.15) = 238.5 J K–1 mol–1 from this work can be directly compared with the value 232.37 J K–1 mol–1 obtained by numeric integration of the Cpm(T)/T dependences from zero to 298.15 K given in Ref. [1]. It should be noted that the values of entropy assessed by thermodynamic optimization of phase equilibrium data are generally considered as less reliable as the values derived from low temperature heat capacity measurements. It is due to possible strong correlation between the enthalpy and entropy contributions to the Gibbs energy. So the obvious discrepancies between our values and data from assessments [12,19] could be explain in this way.

OxideΔoxS a)
(J K-1 mol–1)
ΔoxS
(J K-1 mol–1)
Ref.
CaNb2O6–8.133.45[7]
–22.59[11]
Ca2Nb2O7–1.122.18[11]
Sr2Nb10O27–34.0–350.69[12]
SrNb2O6–17.0–58.65[12]
Sr2Nb2O7–6.0–39.93[12]
Sr5Nb4O14–18.0–51.01[12]
CaBi2O41.93.0[19]
Ca4Bi6O13–23.88.0[19]
Ca2Bi2O56.63.0[19]
SrBi2O44.01.8[19]
Sr2Bi2O55.5–24.0[19]

Table 4.

The values of entropy of formation from binary oxides at 298.15 K: a comparison of our results and data from literature

It should be noted that the thorough analysis of the Debye and Einstein contributions to the heat capacities reveals that the different vibrational modes contribute to the total values of ΔoxS to a different extent and partial compensation is possible in some cases.

2.3. Evaluation of heat capacity at temperatures above 298 K

For the assessment of temperature dependences of Cpm above room temperature, the heat capacity data from DSC and the enthalpy increment data from drop calorimetry were treated simultaneously by the linear least-squares method (HT fit). The temperature dependence of Cpm was considered in the form

Cpm=A+BT+C/T2E5

thus the related temperature dependence of ΔHm(T) = Hm(T) – Hm(T0) is given by equation

ΔHm(T)=Hm(T)Hm(T0)=T0TCpmdT=A(TT0)+B(T2T02)/2C(1/T1/T0)E6

The sum of squares which is minimized has the following form

F=i=1N(Cp)wi2[Cpm,iABTiC/Ti2]2++j=1N(ΔH)wj2[ΔHm,jA(TjT0,j)B(Tj2T0,j2)/2+C(1/Tj1/T0,j)]2minE7

where the first sum runs over the Cpm experimental points while the second sum runs over the ΔHm experimental points. Different weights wi (wj) were assigned to individual points calculated as wi = 1/δi (wj = 1/δj) where δij) is the absolute deviation of the measurement estimated from overall accuracies of measurements (1 % for DSC and 3 % for drop calorimetry). Both types of experimental data thus gain comparable significance during the regression analysis. To smoothly connect the LT fit and HT fit data the values of Cpm(298.15) from LT fit were used as constraints and so Eq. (7) is modified

Fconstr=Fλ[Cpm(298.15)A298.15BC/298.152]minE8

The numerical values of parameters A, B and C are now obtained by solving a set of equations deduced as derivatives of Fconstr with respect of these parameters and a multiplier λ which are equal to zero at the minimum of Fconstr. Assessed values of parameters A, B and C of Eq. (4) for mixed oxides are presented in Table 5.

As an example, the results of heat capacity measurements and relative enthalpy measurements on Bi7Ta3O18 [28] are shown in Fig. 2. Empirical estimation according to the Neumann-Kopp’s rule (NKR) is also plotted for comparison.

The empirical Neumann-Kopp’s rule (NKR) is frequently used for estimation of unknown values of the heat capacity of mixed oxides [46–48]. According to NKR, heat capacity of a mixed oxide is calculated as a sum of heat capacities of the constituent binary ones

Cpm(MO)=ibiCpm(BO,i)E9

It was concluded [47,48] that NKR predicts the heat capacities of mixed oxides remarkably well around room temperature but the deviations (mostly positive) from NKR become substantial at higher temperatures. Mean relative error of the estimated values of Cpm(298.15 K) is 1.4 %. Calculated temperature dependences of ΔoxCpm = Cpm(MO) – biCpm(BO,i) for various mixed oxides in the systems CaO–Nb2O5, SrO–Nb2O5 and Bi2O3–Ta2O5 are shown in Fig. 3.

OxideCpm = A + B·T + C/T 2 (J K–1 mol–1)Temperature range (K)Ref.
A103 B10–6 C
CaBi2O4157.16138.750–1.546298-1000[20]
Ca4Bi6O13550.808114.890–7.201298-1200[20]
Ca2Bi2O5226.09633.374–3.432298-1100[20]
SrBi2O4161.9745.936–1.7832298–1100[29]
Sr2Bi2O5197.4887.463–1.9282298–1200[29]
CaNb2O6200.4034.32–3.45298-1500[26]
Ca2Nb2O7257.2036.21–4.435298-1400[26]
SrNb2O6200.4729.37–3.473298-1500[24]
Sr2Nb2O7248.0043.50–3.948298-1400[26]
Sr2Nb10O27835.351227.648–13.904298-1400[27]
Sr5Nb4O15504.796147.981–6.376298-1400[27]
BiNbO4 a)128.62833.400–1.991150-1200[21]
BiNb5O14455.84060.160–7.734298-1400[23]
BiTaO4 b)133.59425.390–2.734150-1200[21]
Bi4Ta2O11445.85.451–7.489298–1400[28]
Bi7Ta3O18699.052.762–9.956298–1400[28]
Bi3TaO7251.667.05–3.237298–1400[28]
SrBi2Nb2O9324.47063.710–5.076298-1400[22]
SrBi2Ta2O9320.22064.510–4.700298-1400[22]

Table 5.

Parameters of temperature dependence of molar heat capacities of various mixed oxides in the system CaO–SrO–Bi2O3–Nb2O5–Ta2O5

Figure 2.

Temperature dependence of heat capacity (a) and relative enthalpy (b) of Bi7Ta3O18 (3NR means the Dulong-Petit limit).

2.4. Evaluation of enthalpy of formation

The heats of drop-solution for the calcium and strontium carbonates and for the bismuth and niobium oxides were measured first. These data are necessary for the evaluation of the ΔoxH values for the mixed oxides, and furthermore, these data could be compared with the literature data [49–52]. For the AECO3 carbonates, the measured heat effect consists of three contributions:

ΔdsH(AECO3,T)=ΔTH(AECO3,T0T)+ΔdecompH(AECO3,T)+ΔsolH(AEO,T)E10

The measurements were performed at 973 K. The values of ΔdsH(AECO3, 973 K) are given in Table 6 along with the values of ΔdsH(AEO, 973 K), which were derived based on the following thermochemical cycle (T0 ≈ 298 K):

AECO3(s,T0)AEO(melt,T) + CO2(g,T),ΔdsH(AECO3)E11
AECO3(s,T0)AEO(s,T0) + CO2(g,T0),ΔdecompH(AECO3)E12
CO2(g,T0)CO2(g,T),ΔTH(CO2)E13
AEO(s,T0)AEO(melt,T),ΔdsH(AEO)E14
ΔdsH(AEO)=ΔdsH(AECO3 ΔdecompH(AECO3ΔTH(CO2)E15

The values ΔdecompH(CaCO3, 298 K) = 178.8 kJ mol–1, ΔdecompH(SrCO3, 298 K) = 233.9 kJ mol–1 and ΔTH(CO2, 298 → 973 K) = 32.0 kJ mol–1 [53] were used for the calculations.

Next, the ΔdsH values of the binary oxides Bi2O3 and Nb2O5 were measured. Because the dissolution of Nb2O5 and of the mixed oxides at 973 K proceeds rather slowly, the higher temperature of 1073 K was used. The measured values ΔdsH are also given in Table 6.

The experimental values of ΔdsH for SrCO3 and CaCO3 are in quite good agreement with the literature data [49–51]. On the other hand, our results and the published [52] values of ΔdsH(Nb2O5) are quite different. It should be noted that a more endothermic value ΔdecompH(SrCO3, 298 K) = 249.4 kJ mol–1 is presented in the literature [45], which results in more exothermic value for ΔdsH(SrO) by 15.5 kJ mol–1.

ΔdsH for the mixed oxides was measured at 1073 K. The following thermochemical cycle was used for the calculation of ΔoxH for calcium and strontium niobates (T0 ≈ 298 K):

AExNb2O5+x(s,T0)xAEO(melt,T) + Nb2O5(melt,T),ΔdsH(AEO)E16
AEO(s,T0)AEO(melt,T),ΔdsH(AEO)E17
Nb2O5(s,T0)Nb2O5(melt,T),ΔdsH(Nb2O5)E18
xAEO(s,T0) + Nb2O5(s,T0)AExNb2O5+x(s,T0),ΔoxH(AExNb2O5+x)E19

Figure 3.

Temperature dependences of ΔoxCpm for various mixed oxides in the systems CaO–Nb2O5, and SrO–Nb2O5 (a) and Bi2O3–Ta2O5 (b)

ΔoxH(AExNb2O5+x)=xΔdsH(AEO) + ΔdsH(Nb2O5ΔdsH(AExNb2O5+x)E20

An analogous scheme was applied to calculate ΔoxH(BiNbO4). All of the experimental and calculated values are summarized in Table 7. The ΔoxH(298 K) values derived from high-temperature EMN measurements [7,8,11] for the CaO-Nb2O5 oxides and the assessed values from the phase diagram for the SrO-Nb2O5 oxides [12] are also presented in Table 7.

SubstanceT (K)ΔdsH (kJ mol–1) a)ΔdsH (kJ mol–1)
CaCO3973128.4 ± 10.1 (10)119.70 ± 1.02 b)
CaO973–82.39–90.70 ± 1.69 b)
CaO1073–77.04 c)
SrCO3973131.4 ± 9.1 (7)130.16 ± 1.66 d)
134.48 ± 1.89 e)
SrO973–134.47–135.82 ± 2.48 d)
–131.42 ± 1.89 e)
SrO1073–129.25 f)
Bi2O397326.0 ± 2.9 (12)
Bi2O3107339.6 g)
Nb2O51073141.8 ± 6.0 (11)91.97 ± 0.78 h)

Table 6.

Enthalpy of drop-solution in 3Na2O + 4MoO3 melts [30]

SubstanceT (K)ΔdsH (kJ mol–1) a)ΔoxH(298 K)
(kJ mol–1) b)
ΔoxH (298 K)
(kJ mol–1)
CaNb2O61073196.8 ± 20.7 (8)–132.0 ± 23.8–159.8 c)
–130.1 d)
Ca2Nb2O71073195.7 ± 27.8 (8)–208.0 ± 31.9–147.3 c)
–177.5 e)
SrNb2O61073180.50 ± 15.7 (4)–167.9 ± 19.1–325.0 f)
Sr2Nb2O71073167.54 ± 34.7 (4)–289.2 ± 37.5–367.4 f)
BiNbO41073132.61 ± 8.9 (7)–41.9 ± 11.1

Table 7.

Enthalpies of drop-solution in 3Na2O + 4MoO3 melt (ΔdsH) and enthalpy of formation from constituent binary oxides (ΔoxH) [30]

Our values for the calcium niobates are in good agreement with Raghavan’s data [7,8], while the data from Dneprova et al. [11] are quite different. Moreover, a relation, ΔoxH(CaNb2O6) > ΔoxH(Ca2Nb2O7), that holds for the values from the work of Dneprova et al. is rather unexpected. The ΔoxH values for strontium niobates obtained based on the binary SrO-Nb2O5 phase diagram evaluation [12] are substantially more exothermic than our calorimetric data. These large differences in the ΔoxH values are not surprising in view of simultaneous differences in the ΔoxS values from the assessment [12] and those derived from low temperature dependences of the molar heat capacity of SrNb2O6 and Sr2Nb2O7 [24,26].

Advertisement

3. Empirical correlation SV

A linear correlation between the standard molar entropy at 298.15 K and the formula unit volume Vf.u has been proposed by Jenkins and Glaser [54–56]. This approach was used in this work for mixed oxides in the CaO–SrO–Bi2O3–Nb2O5–Ta2O5 system. The linear relation is obvious (see Fig. 4) and the straight line almost naturally passes through the origin:

Sm(JK1mol1)=1680.5Vf.u.(nm3f.u.1)E21

The average relative error in entropy is 8.2 %, the binary oxides CaO and Nb2O5 show the deviations around 20 %. It should be noted that, in this set of values, the simple analogy of NKR (Eq.(9)) provides a better prediction with an average relative error in entropy of 4.2 %.

Eq. (21) can be used for estimation of missing data. So, the estimated value Sm(Sr2Ta2O7) = 256.06 J K–1 mol–1 can be compared with the value 245.41 J K–1 mol–1 obtained by numeric integration of the Cpm(T)/T dependences from zero to 298.15 K given in Ref. [1] (relative deviation of –4.3 %). Simple calculation Sm(Sr2Ta2O5) = 2Sm(SrO) + Sm(Ta2O5) = 250.25 J K–1 mol–1 gives more reliable value (relative deviation 2.0 %).

Advertisement

4. Empirical estimation of enthalpy of formation

There are other mixed oxides in the system CaO–SrO–Bi2O3–Nb2O5–Ta2O5 for which the values of enthalpy of formation ΔfH or enthalpy of formation from binary oxides ΔoxH have not yet been determined. As a rough estimate, the values of ΔoxH calculated according to an empirical method proposed by the authors [56] can be used. In the case of Ca, Sr and Bi niobates the following relation holds for ΔoxH:

ΔoxHnNb+nMe=296.5αyxNbxMeδ(XNbXMe)2E22

where XNb and XMe (Me = Ca, Sr or Bi) are Pauling’s electronegativities of the relevant elements, xNb and xMe are the molar fractions of the oxide-forming elements (xNb = nNb/(nNb + nMe) etc.), y is the number of oxygen atoms per one atom of oxide-forming elements and α and δ are the model parameters. Using Pauling’s electronegativities, XNb = 1.60, XCa = 1.00, XSr = 0.95, and XBi = 2.02, and the calorimetric values of ΔoxH obtained in this work, the values of α = 2.576 and δ = 1.50 were derived from the least-squares fit. The estimated ΔoxH values for calcium and strontium niobates are shown in Fig. 5. The values of ΔoxH that were calculated according to an empirical method proposed by Zhuang et al. [57] are displayed for comparison.

Figure 4.

Correlation between the standard molar entropy at 298.15 K and the formula unit volume Vf.u. for various mixed oxides in the system CaO–SrO– Bi2O3–Nb2O5–Ta2O5 (data from table Table 3)

Advertisement

5. Conclusion

The above presented data derived from calorimetric measurements became the basis for thermodynamic database FS-FEROX [58] compatible with the FactSage software [59,60]. Missing data for other stoichiometric mixed oxides were estimated by the empirical methods described before: the Neumann-Kopp’s rule for heat capacities, the entropy-volume correlation for molar entropies and electronegativity-differences method for enthalpies of formation. At the same time, thermodynamic description of a multicomponent oxide melt was obtained analyzing relevant binary phase diagrams published in literature. The database and the FactSage software were subsequently used for various equilibrium calculations including binary T-x phase diagrams and ternary phase diagrams in subsolidus region. Thermodynamic modeling of SrBi2Ta2O9 and SrBi2Nb2O9 thin layers deposition from the gaseous phase were also performed to optimize the deposition conditions.

Figure 5.

Values of enthalpy of formation of the mixed oxides from constituent binary oxides in the CaO–Nb2O5 (a) and SrO–Nb2O5 (b) systems (lines serve only as a guide for the eyes)

Advertisement

Acknowledgment

This work was supported by the Ministry of Education of the Czech Republic (research projects N° MSM6046137302 and N° MSM6046137307). Part of this work was also supported from the Grant Agency of the Czech Republic, grant No P108/10/1006. Low temperature experiments were performed in MLTL (http://mltl.eu/), which is supported within the program of Czech Research Infrastructures (project no. LM2011025).

References

  1. 1. AkishigeYShigematsuHTojoTKawajiHAtakeT2005Specific heat of Sr2Nb2O7 and Sr2Ta2O7J. Therm. Anal. Calorim. 81537540
  2. 2. ShabbirGKojimaS2003Acoustic and thermal properties of strontium pyroniobate single crystalsJ. Phys. D: Appl. Phys. 3610361039
  3. 3. OnoderaAYoshioKMyintC. CKojimaSYamashitaHTakamaT1999Thermal and structural studies of phase transitions in layered perovskite SrBi2Ta2O9. Jpn. J. Appl. Phys. 3856835685
  4. 4. OnoderaAYoshioKMyintC. CTanakaMHironakaKKojimaS2000Thermal behavior in feroelectric SrBi2Ta2O9 thin films. Ferroelectrics 241159166
  5. 5. YoshioKOnoderaAYamashitaH2003Ferroelectric phase transition and new intermediate phase in bi-layered perovskite SrBi2Ta2O9. Ferroelectrics 2846574
  6. 6. MorimotoKSawaiSHisanoKYamamotoT1999Simultaneous measurements of specific heat capacity and dielectric constant of ferroelectric SrBi2(NbxTa1-x)2O9 ceramics. Ferroelectrics 227133140
  7. 7. RaghavanS1991Thermodynamic stability of monocalcium niobate using calcium fluoride solid electrolyte galvanic cell. Trans. Indian Inst. Met. 44285286
  8. 8. RaghavanS1992Thermodynamics of formation of high calcium niobates from emf measurements. J. Alloys Compd. 179: L25L27.
  9. 9. RaghavanS1991Electrochemical determination of the stability of calcium ditantalate. Indian J. Technol. 29313314
  10. 10. RaghavanS1992Thermodynamics of the formation of high calcium tantalates from emf measurements. J. Alloys Compd. 189: L39L40.
  11. 11. DneprovaV. GRezukhinaT. NGerasimovY. I1968Thermodynamic properties of some calcium niobates.DoklAkad. Nauk SSSR 178135137
  12. 12. YangYYuHJinZ1999Thermodynamic calculation of the SrO-Nb2O5 system.J. Mater. Sci. Technol. 15203207
  13. 13. SaundersNMiodownikA. P1998Calphad (Calculation of phase diagrams): A comprehensive guide.Pergamon Materials Series, 1Pergamon.
  14. 14. HallstedtBRisoldDGaucklerL. J1997Thermodynamic assessment of the bismuth-calcium-oxygen system. J. Am. Ceram. Soc. 8026292636
  15. 15. HallstedtBRisoldDGaucklerL. J1997Thermodynamic assessment of the bismuth-strontium-oxygen system. J. Am. Ceram. Soc. 8010851094
  16. 16. IdemotoYShizukaKYasudaYFuekiK1993Standard enthalpies of formation of member oxides in the Bi-Sr-Ca-Cu-O systemPhysica C 2113644
  17. 17. JacobK. TJayadevanK. P1997Combined use of oxide and fluoride solid electrolytes for the measurement of Gibbs energy of formation of ternary oxides: System Bi-Ca-OMater. Trans. JIM 38427436
  18. 18. JacobK. TJayadevanK. P198System Bi-Sr-O: Synergistic measurements of thermodynamic properties using oxide and fluoride solid electrolytes. J. Mater. Res. 1319051908
  19. 19. HallstedtBGaucklerL2003Revision of the thermodynamic descriptions of the Cu-O, Ag-O, Ag-Cu-O, Bi-Sr-O, Bi-Ca-O, Bi-Cu-O, Sr-Cu-O, Ca-Cu-O and Sr-Ca-Cu-O systemsCALPHAD27177191
  20. 20. AbrmanPSedmidubskýDStrejcAVonkaPLeitnerJ2002Heat capacity of mixed oxides in the Bi2O3-CaO system. Thermochim. Acta 38117
  21. 21. HamplMStrejcASedmidubskýDRužickaKHejtmánekJLeitnerJ2006Heat capacity, enthalpy and entropy of bismuth niobate and bismuth tantalate. J. Solid State Chem. 1797780
  22. 22. LeitnerJHamplMRužickaKSedmidubskýDSvobodaPVejpravováJ2006Heat capacity, enthalpy and entropy of strontium bismuth niobate and strontium bismuth tantalate. Thermochim. Acta 450105109
  23. 23. HamplMLeitnerJRužickaKStrakaMSvobodaP2007Heat capacity and heat content of BiNb5O14. J. Thermal Anal. Calorimetry 87553556
  24. 24. LeitnerJHamplMRužickaKStrakaMSedmidubskýDSvobodaP2008Thermodynamic properties of strontium metaniobate SrNb2O6. J. Thermal. Anal. Calorimetry 91985990
  25. 25. LeitnerJHamplMRužickaKStrakaMSedmidubskýDSvobodaP2008Heat capacity, enthalpy and entropy of strontium niobate Sr2Nb2O7 and calcium niobate Ca2Nb2O7. Thermochim. Acta 4753338
  26. 26. LeitnerJRužickaKSedmidubskýDSvobodaP2009Heat capacity, enthalpy and entropy of calcium niobates. J. Thermal. Anal. Calorimetry 95397402
  27. 27. LeitnerJŠipulaIRužickaKSedmidubskýDSvobodaP2009Heat capacity, enthalpy and entropy of strontium niobates Sr2Nb10O27 and Sr5Nb4O15. J. Alloys Compd. 4813539
  28. 28. LeitnerJJakešVSoferZSedmidubskýDRužickaKSvobodaP2011Heat capacity, enthalpy and entropy of ternary bismuth tantalum oxides. J. Solid State Chem. 184241245
  29. 29. LeitnerJSedmidubskýDRužickaKSvobodaP2012Heat capacity, enthalpy and entropy of SrBi2O4 and Sr2Bi2O5. Thermochim. Acta. 5316065
  30. 30. LeitnerJNevrivaMSedmidubskýDVonkaP2011Enthalpy of formation of selected mixed oxides in a CaO-SrO-Bi2O3-Nb2O5 system. J. Alloys Compd. 50949404943
  31. 31. LashleyJ. CHundleyM. FMiglioriASarraoJ. LPagliusoP. GDarlingT. WJaimeMCooleyJ. CHultsW. LMoralesLThomaD. JSmithJ. LBoerio-goatesJWoodwardB. FStewartG. RFisherR. APhillipsN. E2003Critical examination of heat capacity measurements made on a Quantum Design physical property measurement system. Cryogenics 43369378
  32. 32. DachsEBertoldiC2005Precision and accuracy of the heat-pulse calorimetric technique: low-temperature heat capacities of milligram-sized synthetic mineral samplesEur. J. Mineral. 17251259
  33. 33. MarriottR. AStancescuMKennedyC. AWhiteM. A2006Technique for determination of accurate heat capacities of volatile, powdered, or air-sensitive samples using relaxation calorimetryRev. Sci. Instrum. 77: 096108 EOF096108 EOFpp).
  34. 34. KennedyC. AStancescuMMarriottR. AWhiteM. A2007Recommendations for accurate heat capacity measurements using a Quantum Design physical property measurement systemCryogenics47107112
  35. 35. ShiQSnowC. LBoerio-goatesJWoodfieldB. F2010Accurate heat capacity measurements on powdered samples using a Quantum Design physical property measurement systemJ. Chem. Thermodyn. 4211071115
  36. 36. HwangJ. SLinK. JTienC1997Measurement of heat capacity by fitting the whole temperature response of a heat-pulse calorimeterRev. Sci. Instrum. 6894101
  37. 37. SchnelleWEngelhardtJGmelinE1999Specific heat capacity of Apiezon N high vacuum grease and of Duran borosilicate glassCryogenics39271275
  38. 38. Quantum DesignPhysical Property Measurement System- Application Note, http://www.qdusa.com/sitedocs/productBrochures/heatcapacity-he3.pdfaccessed 21-12-2011
  39. 39. ArblasterJ. W1994The thermodynamic properties of platinum on ITS-90, Platinum Metals Rev. 38119125
  40. 40. Rodriguez-carvajalJ1993Recent advances in magnetic structure determination by neutron powder diffraction.Physica B 1925569
  41. 41. MartinC. A1991Simple treatment of anharmonic effects on the specific heatJ. Phys.: Condens. Matter 359675974
  42. 42. TaylorJ. RDinsdaleA. T1990Thermodynamic and phase diagram data for the CaO-SiO2 systemCALPHAD147188
  43. 43. RisoldDHallstedtBGaucklerL. J1996The strontium-oxygen systemCALPHAD20353361
  44. 44. RisoldDHallstedtBGaucklerL. JLukasH. LFriesS. G1995The bismuth-oxygen system. J. Phase Equilib. 16223234
  45. 45. KnackeOKubaschewskiOHesselmannK1991Thermochemical Properties of Inorganic Substancesnd Ed. Berlin: Springer.
  46. 46. QiuLWhiteM. A2001The constituent additivity method to estimate heat capacities of complex inorganic solidsJ. Chem. Educ. 7810761079
  47. 47. LeitnerJChuchvalecPSedmidubskýDStrejcAAbrmanP2003Estimation of heat capacities of solid mixed oxidesThermochim. Acta 3952746
  48. 48. LeitnerJVonkaPSedmidubskýDSvobodaP2010Application of Neumann-Kopp rule for the estimation of heat capacity of mixed oxides. Thermochim. Acta 497713
  49. 49. Helean KB, Navrotsky A, Vance ER, Carter ML, Ebbinghaus B, Krikorian O, Lian J, Wang LM, Catalano JG (2002) Enthalpies of formation of Ce-pyrochlore, Ca0.93Ce1.00Ti2.035O7.00, U-pyrochlore, Ca1.46U4+0.23U6+0.46Ti1.85O7.00 and Gd-pyrochlore, Gd2Ti2O7: three materials relevant to the proposed waste form for excess weapons plutonium. J. Nuclear Mater. 303: 226-239.
  50. 50. ChengJNavrotskyA2004Energetics of magnesium, strontium, and barium doped lanthanum gallate perovskitesJ. Solid State Chem. 177126133
  51. 51. XuHNavrotskyASuYBalmerM. L2005Perovskite Solid Solutions along the NaNbO3-SrTiO3 Join: Phase Transitions, Formation Enthalpies, and Implications for General Perovskite Energetics. Chem. Mater. 1718801886
  52. 52. PozdnyakovaINavrotskyAShilkinaLReznitchenkoL2002Thermodynamic and structural properties of sodium lithium niobate solid solutionsJ. Am. Ceram. Soc. 85379384
  53. 53. RobbieR. AHemingwayB. S1995Thermodynamic properties of minerals and related substances at 298.15 K and 1 bar pressure and at higher temperatures, U.S. Geological Survey Bulletin, 2131Washington.
  54. 54. Jenkins HDBGlasser L (2003Standard absolute entropy, S°298, values from volume or density. 1. Inorganic materials. Inorg.Chem. 4287028708
  55. 55. Jenkins HDBGlasser L (2006Volume-based thermodynamics: Estimations for 2:2 salts,Inorg. Chem. 4517541756
  56. 56. VonkaPLeitnerJ2009A method for the estimation of the enthalpy of formation of mixed oxides in Al2O3-Ln2O3 systems. J. Solid State Chem. 182744748
  57. 57. ZhuangWLiangJQiaoZShenJShiYRaoG1998Estimation of the standard enthalpy of formation of double oxideJ. Alloys Compd. 267610
  58. 58. LeitnerJSedmidubskýDVonkaP2009Thermodynamic database for the oxide system CaO-SrO-Bi2O3Nb2O5-Ta2O5. CALPHAD XXXVIII, 17.-22.5.2009, Praha, ČR.
  59. 59. BaleC. WChartrandPDegterovS. AErikssonGHackKBen Mahfoud R, Melançon J, Pelton AD, Petersen S (2002FactSage thermochemical software and databases. Calphad 26189228
  60. 60. BaleC. WBélisleEChartrandPDecterovS. AErikssonGHackKJungI. HKangY. BMelançonJPeltonA. DRobelinCPetersenS2009FactSage thermochemical software and databases- recent developments. Calphad 33295311

Written By

Jindřich Leitner, David Sedmidubský, Květoslav Růžička and Pavel Svoboda

Submitted: 11 November 2011 Published: 23 January 2013