inorganic compounds\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoSTRUCTURAL
CHEMISTRY
ISSN: 2053-2296

PbZn1/3Nb2/3O3 at 4.2 and 295 K

CROSSMARK_Color_square_no_text.svg

aSchool of Engineering, University of Newcastle, New South Wales 2308, Australia, and bISIS Facility, Rutherford Appleton Laboratory, Chilton, Didcot, Oxfordshire OX11 0QX, England
*Correspondence e-mail: erich.kisi@newcastle.edu.au

(Received 10 February 2006; accepted 3 April 2006; online 24 May 2006)

The structure of the relaxor ferroelectric lead zinc niobium trioxide, Pb(Zn1/3Nb2/3)O3, known as PZN, was determined at 4.2 and 295 K from very high resolution neutron powder diffraction data. The material is known for its extraordinary piezoelectric properties which are closely linked to the structure. Pseudo-cubic lattice parameters have led to considerable controversy over the symmetry of the structure, which was found to be rhombohedral in the space group R3m at both temperatures. Atomic coordinates have been determined for the first time. They show that, whereas the deviation of the rhombohedral angle from 90° approaches zero at 295 K, the atomic coordinates do not approach typical cubic positions and hence the polarization remains high.

Comment

Recently, the crystal structures of the perovskite relaxor ferroelectric Pb(Zn1/3Nb2/3)O3 (lead zinc niobate or PZN) and its alloys with PbTiO3 (PZN-xPT) have been widely debated in relation to the exceptional piezoelectric properties of these materials (e.g. Park & Shrout, 1997[Park, S.-E. & Shrout, T. R. (1997). J. Appl. Phys. 82, 1804-1811.]). A curious aspect of the crystal structure–property relationship in these materials is that the maximum piezoelectric response is along [001], whereas the spontaneous polarization is along [111] of the nominally rhombohedral crystals. Single-domain single crystals are not available and so previous work has relied on synchrotron X-ray (Noheda et al., 2001[Noheda, B., Cox, D. E., Shirane, G., Park, S.-E., Cross, L. E. & Zhong, Z. (2001). Phys. Rev. Lett. 86, 3891-3894.]; Noheda, Cox & Shirane, 2002[Noheda, B., Cox, D. E. & Shirane, G. (2002). Proceedings of the 10th International Meeting on Ferroelectrics (IMF-10), Madrid, September 2001; published in Ferroelectrics (2002), 267, 147-155.]; Cox et al., 2001[Cox, D. E., Noheda, B., Shirane, G., Uesu, Y., Fujishiro, K. & Yamada, Y. (2001). Appl. Phys. Lett. 79, 400-402.]) and neutron diffraction (Ohwada et al., 2001[Ohwada, K., Hirota, K., Rehrig, P. W., Gehring, P. M., Noheda, B., Fujii, Y., Park, S.-E. & Shirane, G. (2001). J. Phys. Soc. Jpn, 70, 2778-2783.]) reciprocal-space scans around a limited number of reflections from polydomain single crystals, and on X-ray powder diffraction (Ohwada et al., 2001[Ohwada, K., Hirota, K., Rehrig, P. W., Gehring, P. M., Noheda, B., Fujii, Y., Park, S.-E. & Shirane, G. (2001). J. Phys. Soc. Jpn, 70, 2778-2783.]; Noheda, Zhong et al., 2002[Noheda, B., Zhong, Z., Cox, D. E., Shirane, G., Park, S.-E. & Rehring, P. (2002). Phys. Rev. B, 65, 224101.]; La-Orauttapong et al., 2002[La-Orauttapong, D., Noheda, B., Ye, Z.-G., Gehring, P. M., Toulouse, J., Cox, D. E. & Shirane, G. (2002). Phys. Rev. B, 65, 144101.]). All of the previous experimental studies have dealt only with the lattice symmetry, due to a focus of the work on the polarization rotation hypothesis for the large piezoelectric response (Fu & Cohen, 2000[Fu, H. & Cohen, R. E. (2000). Nature (London), 403, 281-283.]). The materials are pseudo-cubic, making the determination of the true symmetry extremely difficult, even with three-axis diffractometers (Noheda, Zhong et al., 2002[Noheda, B., Zhong, Z., Cox, D. E., Shirane, G., Park, S.-E. & Rehring, P. (2002). Phys. Rev. B, 65, 224101.]). Park & Shrout (1997[Park, S.-E. & Shrout, T. R. (1997). J. Appl. Phys. 82, 1804-1811.]) first suggested that an electric field-induced phase transition occurred during piezoelectric cycling of PZN-8%PT and that this facilitates the large piezoelectric response. This was rapidly supported by X-ray diffraction studies (Paik et al., 1999[Paik, D.-S., Park, S.-E., Wada, S., Liu, S.-F. & Shrout, T. R. (1999). J. Appl. Phys. 85, 1080-1083.]; Durbin et al., 1999[Durbin, M. K., Jacobs, E. W., Hicks, J. C. & Park, S.-E. (1999). Appl. Phys. Lett. 74, 2848-2850.]). The observed symmetry was pseudo-tetra­gonal. However, Durbin et al. (1999[Durbin, M. K., Jacobs, E. W., Hicks, J. C. & Park, S.-E. (1999). Appl. Phys. Lett. 74, 2848-2850.]) noted that the true symmetry was more likely to be monoclinic. Noheda et al. (2001[Noheda, B., Cox, D. E., Shirane, G., Park, S.-E., Cross, L. E. & Zhong, Z. (2001). Phys. Rev. Lett. 86, 3891-3894.]) also concluded that a field-induced transition was responsible and proposed that the monoclinic space group is Pm. The same or a very similar mechanism has been proposed for all PZN-xPT crystals in the range 0 < x < 9%.

The counterpoint has come from Kisi et al. (2003[Kisi, E. H., Piltz, R. O., Forrester, J. S. & Howard, C. J. (2003). J. Phys. Condens. Matter, 15, 3631-3640.]) who argued, on group theoretical and physical property grounds, that the large response is more likely to be founded on the very soft elastic constants of the materials than on the lattice symmetry. The observed monoclinic symmetry under an electric field directed along [001] is exactly as expected for the conventional piezoelectric distortion of a rhombohedral crystal. However, this does not constitute a phase transition, nor does it explain the large piezoelectric response. An understanding of the connection between the crystal structure and the piezoelectric properties can only be derived from measured ion coordinates. These can then be used to compute the electric polarization by methods such as that given by Darlington et al. (1994[Darlington, C. N. W., David, W. I. F. & Knight, K. S. (1994). Phase Transitions, 48, 217-236.]).

This paper reports the full crystal structure of the end member, PZN, for the first time. The use of very high resolution neutron diffraction ensures both that the O-ion positions are reliably determined in the presence of far heavier elements such as Pb, and that the problem of pseudo-symmetry is minimized. Data at both 4.2 and 295 K were used to enhance the discrimination between different structural models. Visual inspection of the 111, 200 and 222 reflections in the high-resolution neutron diffraction patterns supported the conventional rhombohedral symmetry. The 111 and 222 reflections are split approximately in the ratio 3:1 and 200 remains unsplit. This aspect has been discussed by Kisi et al. (2005[Kisi, E. H., Forrester, J. S. & Knight, K. S. (2005). J. Phys. Condens. Matter, 17, L381-L384.]) in relation to reports of a cubic `X-phase' in this system. Rietveld refinements were initially conducted using the 4.2 K data, due to the larger rhombohedral distortion at this temperature. Refinements in the space group R3m, illustrated in Fig. 1[link](a), gave reasonable agreement. The major misfit is due to slightly anisotropic reflection widths, in which 200 is broader than 111 by ≃ 0.2 times the full width at half-maximum. Geometrically, it is possible for a monoclinic distortion both to split the 111 reflection (β > 90°) and to broaden 200 (abc) if a, b and c are sufficiently similar. Given that monoclinic phases have been reported in this material, refinements were attempted in all of the monoclinic space groups recently proposed [Cm, Cc (Ic) and Pm], with unsatisfactory results. No improvement in the agreement was observed even when a considerable number of extra parameters were added to the refinement. The best fit, shown in Fig. 1[link](b), used the monoclinic structure in Cm (Frantti et al., 2003[Frantti, J., Eriksson, S., Hull, S., Lantto, V., Rundlof, H. & Kakihana, M. (2003). J. Phys. Condens. Matter, 15, 6031-6041.]). Octa­hedral rotation in combination with ferroelectric cation displacements was ruled out by the absence of R-point (e.g. R3c) or M-point (e.g. Cc) superlattice reflections. This was confirmed by Rietveld refinement, which converged to unrotated octa­hedral positions regardless of the initial conditions. A two-phase mixture of rhombohedral (R3m) and tetra­gonal (P4mm) structures was also tested for completeness. The final refinements were undertaken in R3m using the anisotropic broadening model of Stephens (1999[Stephens, P. W. (1999). J. Appl. Cryst. 32, 281-289.]), as implemented in GSAS (Larson & Von Dreele, 1986[Larson, A. C. & Von Dreele, R. B. (1986). GSAS. Report LAUR 86-748. Los Alamos National Laboratory, New Mexico, USA.]), to model the effect of inter­domain strains due to the microscopic ferroelectric domain structure within the crystallites. Only one additional parameter, S400, was required. The final fit is illustrated in Fig. 1[link](c). A projection of the structure at 4.2 K perpendicular to the polar axis, [111], is illustrated in Fig. 2[link](a) and an equivalent view of the undistorted cubic structure is shown in Fig. 2[link](b). The Pb atoms are aligned, illustrating the large relative displacement of the Zn/Nb and O ions.

Structure refinements using the 295 K data were undertaken using the 4.2 K structure as a starting point. The refined structure at 295 K is the same as that at 4.2 K, with slightly reduced rhombohedral distortion and cation displacements. The departure of the rhombohedral angle from 90° diminishes with increasing temperature as the transition to cubic at approximately 400 K is approached. However, the atom coordinates remain relatively unchanged. Similar behaviour has been observed as phase transitions are approached in some other ferroelectrics, such as BaTiO3 (Darlington et al., 1994[Darlington, C. N. W., David, W. I. F. & Knight, K. S. (1994). Phase Transitions, 48, 217-236.]). This is somewhat contrary to the widely held notion that the ferroelectric distortion controls the crystal structure at all levels, i.e. that the lattice distortion should always scale with the ion positions and electric polarization. PZN has an extremely small distortion from cubic and yet has a large polarization, in contrast with, say, PbTiO3, which has a large spontaneous strain but considerably smaller polarization. These observations indicate that, in ferroelectric materials, the relationship between the lattice parameters and atomic coordinates (and hence electric polarization) is more complex than in the non-ferroelectric perovskites.

[Figure 1]
Figure 1
Results of Rietveld refinements using the 4.2 K data in the range 1–2.4 Å for models in (a) space group R3m without anisotropic broadening correction, (b) space group Cm and (c) space group R3m with the anisotropic broadening parameter S400 refined. Data are given as +, the calculated profile as a solid line and the difference profile as a solid line below. Vertical markers above the difference profile indicate the calculated Bragg reflection positions. Data in the range 2–2.4 Å are plotted according to the right-hand scale. Insets show details of the fit to the important 400 and 222 reflections.
[Figure 2]
Figure 2
Views of (a) the 4.2 K structure and (b) the undistorted cubic structure, both projected perpendicular to the polar axis, [111].

Experimental

The flux growth method of Mulvihill et al. (1996[Mulvihill, M. L., Park, S.-E., Risch, G., Li, Z., Uchino, K. & Shrout, T. R. (1996). Jpn J. Appl. Phys. 35, 3984-3990.]) was used to grow single crystals of composition Pb(Zn1/3Nb2/3)O3. Crystals in the size range 0.5–15 mm were extracted from the flux in a hot HNO3 solution. The chemical composition of the crystals and their homogeneity were verified using scanning electron microscopy-based energy dispersive spectroscopy (EDS) analysis. Using an agate pestle and mortar, crystals were lightly crushed to a size that passed through a coarse sieve (143 µm), so as to avoid strain and particle-size broadening. Good powder averaging was obtained using approximately 2 ml of powder sample.

Neutron powder diffraction patterns were recorded in transmission on the HRPD diffractometer (resolution Δd/d ≃ 4 × 10−4) at the ISIS facility, Rutherford Appleton Laboratory, England. Samples were held in a thin-walled aluminium can within a liquid helium cryostat. Diffraction patterns were recorded from 30 to 130 ms at room temperature and at 4.2 K for 20 and 90 min, respectively.

PZN at 4.2 K

Crystal data
  • PbZn1/3Nb2/3O3

  • Mr = 337.40

  • Trigonal, R 3m

  • a = 4.06048 (6) Å

  • α = 89.8693 (4)°

  • V = 66.95 (1) Å3

  • Z = 1

  • Dx = 8.367 Mg m−3

  • Time-of-flight neutron radiation

  • λ = 0.65–2.45 Å

  • T = 4.2 K

  • Specimen shape: powder

  • Specimen prepared at 101.3 kPa

  • Specimen prepared by cooling from 1473 to 1173 K at 1 K min−1 and at 50 K min−1 thereafter

  • Particle morphology: <143 μm particles, pale yellow

Data collection
  • HRPD beamline, ISIS, Didcot

  • Specimen mounting: 11 mm aluminium slab can

  • Scan method: time-of-flight

Refinement
  • Rp = 0.062

  • Rwp = 0.074

  • Rexp = 0.040

  • S = 2.48

  • Wavelength of incident radiation: 0.65–2.45 Å

  • Excluded region(s): none

  • Profile function: pseudo-Voight convoluted with double exponential incorportating the anisotropic microstrain broadening model of Stephens (1999[Stephens, P. W. (1999). J. Appl. Cryst. 32, 281-289.])

  • 34 parameters

  • (Δ/σ)max = 0.01

  • Preferred orientation correction: none

PZN at 295 K

Crystal data
  • PbZn1/3Nb2/3O3

  • Mr = 338.93

  • Trigonal, R 3m

  • a = 4.06240 (8) Å

  • α = 89.8693 (4)°

  • V = 67.04 (1) Å3

  • Z = 1

  • Dx = 8.396 Mg m−3

  • Time-of-flight neutron radiation

  • λ = 0.65–2.45 Å

  • T = 295 K

  • Specimen shape: powder

  • Specimen prepared at 101.3 kPa

  • Specimen prepared at 101.3 kPa

  • Specimen prepared by cooling from 1473 to 1173 K at 1 K min−1 and at 50 K min−1 thereafter

  • Particle morphology: <143 μm particles, pale yellow

Data collection
  • HRPD beamline, ISIS, Didcot

  • Specimen mounting: 11 mm aluminium slab can

  • Specimen mounted in transmission mode

  • Scan method: time-of-flight

Refinement
  • Rp = 0.075

  • Rwp = 0.088

  • Rexp = 0.080

  • S = 1.49

  • Wavelength of incident radiation: 0.65–2.45 Å

  • Excluded region(s): none

  • Profile function: pseudo-Voight convoluted with double exponential incorportating the anisotropic microstrain broadening model of Stephens (1999[Stephens, P. W. (1999). J. Appl. Cryst. 32, 281-289.])

  • 37 parameters

  • (Δ/σ)max = 0.03

  • Preferred orientation correction: none

Rietveld analyses used the program GSAS (Larson & Von Dreele, 1986[Larson, A. C. & Von Dreele, R. B. (1986). GSAS. Report LAUR 86-748. Los Alamos National Laboratory, New Mexico, USA.]). Data from the small crystal sample recorded by both the high-resolution backscattering bank and the high-intensity 90° detector bank were used simultaneously. Typically, lattice parameters, atom coordinates (Zn/Nb, O), isotropic displacement parameters (Zn/Nb), anisotropic displacement parameters (Pb, O), scale, eight polynomial background parameters and two peak profile parameters were used. The 4.2 K data were used to refine the Zn/Nb ratio of the B cation site and this ratio was held constant during refinements based on the 295 K data. In addition, the anisotropic peak-broadening model of Stephens (1999[Stephens, P. W. (1999). J. Appl. Cryst. 32, 281-289.]) was applied.

For both compounds, data collection: ISIS software (local programs); program(s) used to refine structure: GSAS (Larson & Von Dreele, 1986[Larson, A. C. & Von Dreele, R. B. (1986). GSAS. Report LAUR 86-748. Los Alamos National Laboratory, New Mexico, USA.]); molecular graphics: CaRine Crystallography (Bondias & Monceau, 2005[Bondias, C. & Monceau, D. (2005). CaRine Crystallography. Distributed by Divergent SA, Compiegne, France.]) and GRAPHER (Golden Software, 2005[Golden Software (2005). GRAPHER. Golden Software Inc., Golden, Colorado, USA.]); software used to prepare material for publication: GSAS.

Supporting information


Comment top

Recently, the crystal structures of the perovskite relaxor ferroelectric Pb(Zn1/3Nb2/3)O3 (lead zinc niobate or PZN) and its alloys with PbTiO3 (PZN-xPT) have been widely debated in relation to the exceptional piezoelectric properties of these materials (e.g. Park & Shrout, 1997). A curious aspect of the crystal structure–property relationship in these materials is that the maximum piezoelectric response is along [001], whereas the spontaneous polarization is along [111] of the nominally rhombohedral crystals. Single-domain single crystals are not available and so previous work has relied on synchrotron X-ray (Noheda et al., 2001; Noheda, Cox & Shirane, 2002; Cox et al., 2001) and neutron diffraction (Ohwada et al., 2001) reciprocal-space scans around a limited number of reflections from poly domain single crystals, and on X-ray powder diffraction (Ohwada et al., 2001; Noheda, Zhong et al., 2002; La-Orauttapong et al., 2002). All of the previous experimental studies have dealt only with the lattice symmetry, due to a focus of the work on the polarization rotation hypothesis for the large piezoelectric response (Fu & Cohen, 2000). The materials are pseudo-cubic, making the determination of the true symmetry extremely difficult, even with three-axis diffractometers (Noheda, Zhong et al., 2002). Park & Shrout (1997) first suggested that an electric field-induced phase transition occurred during piezoelectric cycling of PZN-8%PT and that this facilitates the large piezoelectric response. This was rapidly supported by X-ray diffraction studies (Paik et al., 1999; Durbin et al., 1999). The observed symmetry was pseudo-tetragonal. However, Durbin et al. (1999) noted that the true symmetry was more likely to be monoclinic. Noheda et al. (2001) also concluded that a field-induced transition was responsible and proposed that the monoclinic space group is Pm. The same or a very similar mechanism has been proposed for all PZN-xPT crystals in the range 0 < x < 9%.

The counterpoint has come from Kisi et al. (2003) who argued, on group theoretical and physical property grounds, that the large response is more likely to be founded on the very soft elastic constants of the materials than on the lattice symmetry. The observed monoclinic symmetry under an electric field directed along [001] is exactly as expected for the conventional piezoelectric distortion of a rhombohedral crystal. However, this does not constitute a phase transition, nor does it explain the large piezoelectric response. An understanding of the connection between the crystal structure and the piezoelectric properties can only be derived from measured ion coordinates. These can then be used to compute the electric polarization by methods such as that given by Darlington et al. (1994).

This paper reports the full crystal structure of the end member, PZN, for the first time. The use of very high resolution neutron diffraction ensures both that the O ion positions are reliably determined in the presence of far heavier elements such as Pb, and that the problem of pseudo-symmetry is minimized. Data at both 4.2 K and 295 K were used to enhance the discrimination between different structural models. Visual inspection of the 111, 200 and 222 reflections in the high-resolution neutron diffraction patterns supported the conventional rhombohedral symmetry. The 111 and 222 reflections are split approximately in the ratio 3:1 and 200 remains un-split. This aspect has been discussed by Kisi et al. (2005) in relation to reports of a cubic `X-phase' in this system. Rietveld refinements were initially conducted using the 4.2 K data, due to the larger rhombohedral distortion at this temperature. Refinements in space group R3m, illustrated in Fig. 1(a), gave reasonable agreement. The major misfit is due to slightly anisotropic reflection widths, in which 200 is broader than 111 by 0.2 times the full width at half-maximum. Geometrically, it is possible for a monoclinic distortion both to split the 111 reflection (β > 90°) and to broaden 200 (abc) if a, b and c are sufficiently similar. Given that monoclinic phases have been reported in this material, refinements were attempted in all of the monoclinic space groups recently proposed [Cm, Cc (Ic) and Pm], with unsatisfactory results. No improvement in the agreement was observed even when a considerable number of extra parameters were added to the refinement. The best fit, shown in Fig. 1(b), used the monoclinic structure in Cm (Frantti et al., 2003). Octahedral rotation in combination with ferroelectric cation displacements was ruled out by the absence of R-point (e.g. R3c) or M-point (e.g. Cc) superlattice reflections. This was confirmed by Rietveld refinement, which converged to un-rotated octahedral positions regardless of the initial conditions. A two-phase mixture of rhombohedral (R3m) and tetragonal (P4 mm) structures was also tested for completeness. The final refinements were undertaken in R3m using the anisotropic broadening model of Stephens (1999), as implemented in GSAS (Larson & Von Dreele, 1986), to model the effect of interdomain strains due to the microscopic ferroelectric domain structure within the crystallites. Only one additional parameter, S400, was required. The final fit is illustrated in Fig. 1(c). A projection of the structure at 4.2 K perpendicular to the polar axis, [111], is illustrated in Fig. 2(a) and an equivalent view of the undistorted cubic structure is shown in Fig. 2(b). The Pb atoms are aligned, illustrating the large relative displacement of the Zn/Nb and O ions.

Structure refinements using the 295 K data were undertaken using the 4.2 K structure as a starting point. The refined structure at 295 K is the same as that at 4.2 K, with slightly reduced rhombohedral distortion and cation displacements. The departure of the rhombohedral angle from 90° diminishes with increasing temperature as the transition to cubic at approximately 400 K is approached. However, the atom coordinates remain relatively unchanged. Similar behaviour has been observed as phase transitions are approached in some other ferroelectrics, such as BaTiO3 (Darlington et al., 1994). This is somewhat contrary to the widely held notion that the ferroelectric distortion controls the crystal structure at all levels, i.e. that the lattice distortion should always scale with the ion positions and electric polarization. PZN has an extremely small distortion from cubic and yet has a large polarization, in contrast with, say, PbTiO3, which has a large spontaneous strain but considerably smaller polarization. These observations indicate that, in ferroelectric materials, the relationship between the lattice parameters and atom coordinates (and hence electric polarization) is more complex than in the non-ferroelectric perovskites.

Experimental top

The flux growth method of Mulvihill et al. (1996) was used to grow single crystals of composition Pb(Zn1/3Nb2/3)O3. Crystals in the size range 0.5–15 mm were extracted from the flux in a hot HNO3 solution. The chemical composition of the crystals and their homogeneity were verified using scanning electron microscopy-based EDS [Please give in full] analysis. Using an agate pestle and mortar, crystals were lightly crushed to a size that passed through a coarse sieve (143 µm), so as to avoid strain and particle-size broadening. Good powder averaging was obtained using 2 ml of powder sample.

Neutron powder diffraction patterns were recorded in transmission on the HRPD diffractometer (resolution Δd/d 4 × 10−4) at the ISIS facility, Rutherford Appleton Laboratory, UK. Samples were held in a thin-walled aluminium can within a liquid helium cryostat. Diffraction patterns were recorded from 30 to 130 ms at room temperature and at 4.2 K for 20 min and 90 min, respectively.

Refinement top

Rietveld analyses used the program GSAS (Larson & Von Dreele, 1986). Data from the small crystal sample recorded by both the high-resolution backscattering bank and the high-intensity 90° detector bank were used simultaneously. Typically, lattice parameters, atom coordinates (Zn/Nb, O), isotropic displacement parameters (Zn/Nb), anisotropic displacement parameters (Pb, O), scale, eight polynomial background parameters and two peak profile parameters were used. The 4.2 K data were used to refine the Zn/Nb ratio of the B cation site and this ratio was held constant during refinements based on the 295 K data. In addition, the anisotropic peak-broadening model of Stephens (1999) was applied.

Computing details top

For both compounds, program(s) used to refine structure: GSAS.

Figures top
[Figure 1] Fig. 1. Results of Rietveld refinements using the 4.2 K data in the range 1–2.4 Å for models in (a) space group R3m without anisotropic broadening correction, (b) space group Cm and (c) space group R3m with the anisotropic broadening parameter S400 refined. Data are given as +, the calculated profile as a solid line and the difference profile as a solid line below. Vertical markers above the difference profile indicate the calculated Bragg reflection positions. Data in the range 2–2.4 Å are plotted according to the right-hand scale. Insets show details of the fit to the important 400 and 222 reflections.
[Figure 2] Fig. 2. Views of (a) the 4.2 K structure and (b) the undistorted cubic structure, both projected perpendicular to the polar axis [111].
(PZN4K) lead zinc niobium trioxide top
Crystal data top
Nb0.61O3PbZn0.39Dx = 8.367 Mg m3
Mr = 337.40Time-of-flight neutron radiation
Trigonal, R3mT = 4 K
a = 4.06048 (6) ÅParticle morphology: <143 µm particles
c = 4.06048 Åpale yellow
V = 66.95 (1) Å3?, ? × ? × ? mm
Z = 1
Data collection top
HRPD, ISIS, Didcot
diffractometer
Scan method: time of flight
Radiation source: ISIS, Spallation2θfixed = 168.33
Specimen mounting: 11mm Al slab canDistance from specimen to detector: 1000 mm
Data collection mode: transmission
Refinement top
Least-squares matrix: fullExcluded region(s): no excluded regions
Rp = 0.06234 parameters
Rwp = 0.0740 restraints
Rexp = ?Weighting scheme based on measured s.u.'s
R(F2) = 0.06901(Δ/σ)max = 0.01
χ2 = 6.150Background function: GSAS Background function number 1 with 12 terms. Shifted Chebyshev function of 1st kind 1: 0.439789 2: 2.990220E-03 3: 4.821580E-02 4: 1.914300E-02 5: 3.191740E-02 6: 3.362380E-02 7: 1.183440E-02 8: 7.640780E-03 9: 1.057330E-0210: 2.050820E-0211: -4.850290E-0312: 1.058700E-02
? data points
Crystal data top
Nb0.61O3PbZn0.39V = 66.95 (1) Å3
Mr = 337.40Z = 1
Trigonal, R3mTime-of-flight neutron radiation
a = 4.06048 (6) ÅT = 4 K
c = 4.06048 Å?, ? × ? × ? mm
Data collection top
HRPD, ISIS, Didcot
diffractometer
Scan method: time of flight
Specimen mounting: 11mm Al slab can2θfixed = 168.33
Data collection mode: transmissionDistance from specimen to detector: 1000 mm
Refinement top
Rp = 0.062χ2 = 6.150
Rwp = 0.074? data points
Rexp = ?34 parameters
R(F2) = 0.069010 restraints
Special details top

Experimental. resolution Δd/d 4x10-4 Diffraction patterns recorded from 30 to 130 ms

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/UeqOcc. (<1)
Pb10.50.50.50.04227
Zn10.0258 (3)0.0258 (3)0.0258 (3)0.0142 (3)*0.389
Nb10.0258 (3)0.0258 (3)0.0258 (3)0.0142 (3)*0.611
O10.5438 (5)0.0547 (4)0.0547 (4)0.02313
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Pb10.0423 (5)0.0423 (5)0.0423 (5)0.0146 (5)0.0146 (5)0.0146 (5)
O10.0239 (8)0.0227 (5)0.0227 (5)0.0004 (6)0.0004 (6)0.0045 (8)
Geometric parameters (Å, º) top
Pb1—Nb13.343 (2)Pb1—O1i2.8907 (3)
Pb1—Nb1i3.4572 (6)Nb1—O1iii1.964 (2)
Pb1—Nb1ii3.5787 (7)Nb1—O12.110 (3)
Pb1—O12.565 (2)
O1—Pb1—O1iv66.28 (7)O1—Nb1—O1iv83.3 (1)
O1iii—Nb1—O1170.6 (2)Pb1—O1—Nb190.73 (7)
O1iii—Nb1—O1v96.5 (1)Pb1—O1—Nb1vi98.7 (1)
O1iii—Nb1—O1iv89.70 (2)Nb1—O1—Nb1vi170.6 (2)
Symmetry codes: (i) x, y, z+1; (ii) x, y+1, z+1; (iii) x1, y, z; (iv) z, x, y; (v) z, x1, y; (vi) x+1, y, z.
(295K) top
Crystal data top
Nb0.61O3PbZn0.39Z = 1
Mr = 338.93Dx = 8.396 Mg m3
Trigonal, R3mTime-of-flight neutron radiation, λ = 0.65-2.45 Å
a = 4.06240 (8) ÅT = 295 K
c = 4.0624 Åpale yellow
V = 67.04 (1) Å3?, ? × ? × ? mm
Data collection top
HRPD beamline, ISIS, Didcot
diffractometer
Refinement top
Least-squares matrix: full? data points
Rp = 0.07537 parameters
Rwp = 0.0880 restraints
Rexp = ?(Δ/σ)max = 0.03
R(F2) = 0.08346Background function: GSAS Background function number 1 with 12 terms. Shifted Chebyshev function of 1st kind 1: 0.236557 2: 9.505910E-03 3: 2.018040E-02 4: 6.956430E-03 5: 8.509230E-03 6: 1.341600E-02 7: 3.919290E-03 8: 2.020580E-03 9: 5.622050E-0410: 8.049800E-0311: -3.196390E-0312: 3.955620E-03; GSAS Background function number 4 with 4 terms. Power series in Q**2n/n! 1: 0.441369 2: -1.651550E-03 3: 4.525360E-05 4: 2.316550E-05
χ2 = 2.220
Crystal data top
Nb0.61O3PbZn0.39V = 67.04 (1) Å3
Mr = 338.93Z = 1
Trigonal, R3mTime-of-flight neutron radiation, λ = 0.65-2.45 Å
a = 4.06240 (8) ÅT = 295 K
c = 4.0624 Å?, ? × ? × ? mm
Data collection top
HRPD beamline, ISIS, Didcot
diffractometer
Refinement top
Rp = 0.075χ2 = 2.220
Rwp = 0.088? data points
Rexp = ?37 parameters
R(F2) = 0.083460 restraints
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/UeqOcc. (<1)
Pb10.50.50.50.0494
Zn10.0216 (5)0.0216 (5)0.0216 (5)0.0167 (5)*0.39
Nb10.0216 (5)0.0216 (5)0.0216 (5)0.0167 (5)*0.61
O10.5377 (7)0.0469 (6)0.0469 (6)0.0269
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Pb10.0495 (8)0.0495 (8)0.0495 (8)0.012 (1)0.012 (1)0.012 (1)
O10.023 (1)0.0288 (7)0.0288 (7)0.0013 (9)0.0013 (9)0.005 (1)
Geometric parameters (Å, º) top
Pb1—Nb13.362 (3)Pb1—O1i2.8868 (4)
Pb1—Nb1i3.466 (1)Nb1—O1iii1.977 (4)
Pb1—Nb1ii3.574 (1)Nb1—O12.094 (4)
Pb1—O12.614 (3)
O1—Pb1—O1iv65.3 (1)O1—Nb1—O1iv84.6 (2)
O1iii—Nb1—O1172.5 (3)Pb1—O1—Nb190.5 (1)
O1iii—Nb1—O1v95.3 (2)Pb1—O1—Nb1vi97.1 (2)
O1iii—Nb1—O1iv89.80 (2)Nb1—O1—Nb1vi172.5 (3)
Symmetry codes: (i) x, y, z+1; (ii) x, y+1, z+1; (iii) x1, y, z; (iv) z, x, y; (v) z, x1, y; (vi) x+1, y, z.

Experimental details

(PZN4K)(295K)
Crystal data
Chemical formulaNb0.61O3PbZn0.39Nb0.61O3PbZn0.39
Mr337.40338.93
Crystal system, space groupTrigonal, R3mTrigonal, R3m
Temperature (K)4295
a, c (Å)4.06048 (6), 4.060484.06240 (8), 4.0624
α (°)89.8693 (4)89.9226 (5)
V3)66.95 (1)67.04 (1)
Z11
Radiation typeTime-of-flight neutronTime-of-flight neutron, λ = 0.65-2.45 Å
µ (mm1)?
Specimen shape, size (mm)?, ? × ? × ??, ? × ? × ?
Data collection
DiffractometerHRPD, ISIS, Didcot
diffractometer
HRPD beamline, ISIS, Didcot
diffractometer
Specimen mounting11mm Al slab can?
Data collection modeTransmission?
Scan methodTime of flight?
2θ values (°)2θfixed = 168.332θmin = ? 2θmax = ? 2θstep = ?
Distance from specimen to detector (mm)1000
Refinement
R factors and goodness of fitRp = 0.062, Rwp = 0.074, Rexp = ?, R(F2) = 0.06901, χ2 = 6.150Rp = 0.075, Rwp = 0.088, Rexp = ?, R(F2) = 0.08346, χ2 = 2.220
No. of data points??
No. of parameters3437

Computer programs: GSAS.

 

Acknowledgements

The authors extend thanks to Dr Christopher Howard for reviewing the work prior to submission. This work was supported by the Australian Research Council and the Access to Major Research Facilities Programme.

References

First citationBondias, C. & Monceau, D. (2005). CaRine Crystallography. Distributed by Divergent SA, Compiegne, France.  Google Scholar
First citationCox, D. E., Noheda, B., Shirane, G., Uesu, Y., Fujishiro, K. & Yamada, Y. (2001). Appl. Phys. Lett. 79, 400–402.  Web of Science CrossRef CAS Google Scholar
First citationDarlington, C. N. W., David, W. I. F. & Knight, K. S. (1994). Phase Transitions, 48, 217–236.  CrossRef CAS Web of Science Google Scholar
First citationDurbin, M. K., Jacobs, E. W., Hicks, J. C. & Park, S.-E. (1999). Appl. Phys. Lett. 74, 2848–2850.  Web of Science CrossRef CAS Google Scholar
First citationFrantti, J., Eriksson, S., Hull, S., Lantto, V., Rundlof, H. & Kakihana, M. (2003). J. Phys. Condens. Matter, 15, 6031–6041.  Web of Science CrossRef CAS Google Scholar
First citationFu, H. & Cohen, R. E. (2000). Nature (London), 403, 281–283.  PubMed CAS Google Scholar
First citationGolden Software (2005). GRAPHER. Golden Software Inc., Golden, Colorado, USA.  Google Scholar
First citationKisi, E. H., Forrester, J. S. & Knight, K. S. (2005). J. Phys. Condens. Matter, 17, L381–L384.  Web of Science CrossRef CAS Google Scholar
First citationKisi, E. H., Piltz, R. O., Forrester, J. S. & Howard, C. J. (2003). J. Phys. Condens. Matter, 15, 3631–3640.  Web of Science CrossRef CAS Google Scholar
First citationLa-Orauttapong, D., Noheda, B., Ye, Z.-G., Gehring, P. M., Toulouse, J., Cox, D. E. & Shirane, G. (2002). Phys. Rev. B, 65, 144101.  Web of Science CrossRef Google Scholar
First citationLarson, A. C. & Von Dreele, R. B. (1986). GSAS. Report LAUR 86-748. Los Alamos National Laboratory, New Mexico, USA.  Google Scholar
First citationMulvihill, M. L., Park, S.-E., Risch, G., Li, Z., Uchino, K. & Shrout, T. R. (1996). Jpn J. Appl. Phys. 35, 3984–3990.  CrossRef CAS Web of Science Google Scholar
First citationNoheda, B., Cox, D. E. & Shirane, G. (2002). Proceedings of the 10th International Meeting on Ferroelectrics (IMF-10), Madrid, September 2001; published in Ferroelectrics (2002), 267, 147–155.  Google Scholar
First citationNoheda, B., Cox, D. E., Shirane, G., Park, S.-E., Cross, L. E. & Zhong, Z. (2001). Phys. Rev. Lett. 86, 3891–3894.  Web of Science CrossRef PubMed CAS Google Scholar
First citationNoheda, B., Zhong, Z., Cox, D. E., Shirane, G., Park, S.-E. & Rehring, P. (2002). Phys. Rev. B, 65, 224101.  Web of Science CrossRef Google Scholar
First citationOhwada, K., Hirota, K., Rehrig, P. W., Gehring, P. M., Noheda, B., Fujii, Y., Park, S.-E. & Shirane, G. (2001). J. Phys. Soc. Jpn, 70, 2778–2783.  Web of Science CrossRef CAS Google Scholar
First citationPaik, D.-S., Park, S.-E., Wada, S., Liu, S.-F. & Shrout, T. R. (1999). J. Appl. Phys. 85, 1080–1083.  Web of Science CrossRef CAS Google Scholar
First citationPark, S.-E. & Shrout, T. R. (1997). J. Appl. Phys. 82, 1804–1811.  CrossRef CAS Web of Science Google Scholar
First citationStephens, P. W. (1999). J. Appl. Cryst. 32, 281–289.  Web of Science CrossRef CAS IUCr Journals Google Scholar

© International Union of Crystallography. Prior permission is not required to reproduce short quotations, tables and figures from this article, provided the original authors and source are cited. For more information, click here.

Journal logoSTRUCTURAL
CHEMISTRY
ISSN: 2053-2296
Follow Acta Cryst. C
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds