strength and texture of pt compressed to...

9
Strength and texture of Pt compressed to 63 GPa Susannah M. Dorfman, 1,a) Sean R. Shieh, 2 and Thomas S. Duffy 3 1 Earth and Planetary Science Laboratory, Ecole Polytechnique F ed erale de Lausanne, Station 3, CH-1015 Lausanne, Switzerland 2 Department of Earth Sciences, University of Western Ontario, London, Ontario N6A 5B7, Canada 3 Department of Geosciences, Princeton University, Princeton, New Jersey 08544, USA (Received 18 September 2014; accepted 28 January 2015; published online 10 February 2015) Angle- and energy-dispersive X-ray diffraction experiments in a radial geometry were performed in the diamond anvil cell on polycrystalline platinum samples at pressures up to 63 GPa. Observed yield strength and texture depend on grain size. For samples with 70–300-nm particle size, the yield strength is 5–6 GPa at 60 GPa. Coarse-grained (2-lm particles) Pt has a much lower yield strength of 1–1.5 GPa at 60 GPa. Face-centered cubic metals Pt and Au have lower strength to shear modulus ratio than body-centered cubic or hexagonal close-packed metals. While a 300-nm particle sample exhibits the h110i texture expected of face-centered-cubic metals under compression, smaller and larger particles show a weak mixed h110i and h100i texture under compression. Differences in texture development may also occur due to deviations from uniaxial stress under compression in the diamond anvil cell. V C 2015 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4907866] I. INTRODUCTION The deformation behavior of materials is fundamental to physics, engineering, and geoscience and is determined by both elastic and plastic properties. 1 While a variety of meth- ods can be employed to independently measure elasticity and plasticity at ambient conditions, X-ray diffraction is one of the only established techniques that can be used to explore these properties in situ at extreme pressures. 24 Existing diffraction-based methods use diffraction line shifts 24 and widths 5,6 to constrain material deformation but typically suf- fer from uncertainties due to confounding of multiple varia- bles and limited control and measurement of experimental conditions such as the applied stress. 5,7 The uncertainties in high-pressure measurements of strength and anisotropy can be evaluated by systematic study of simple materials such as platinum metal. Platinum (Pt) is a face-centered cubic (FCC) metal and does not undergo phase transitions up to multi-megabar pres- sures. 7,8 Due in part to its simple structure, Pt is widely used in high-pressure experiments as a pressure calibrant and laser absorber. 9 Nanoparticles of Pt 10 and other metals 11 have also been observed to have anomalously low compressibility, and this has possible implications for pressure calibration using metal powders. Recently, Pt has also been used as a devia- toric strain marker: the success of laser annealing or pressure media in minimizing deviatoric stresses has been judged by comparison to the strength of Pt as a function of pres- sure. 8,1215 The sensitivity of Pt to deviatoric stress due to its strength and elastic properties is also a source of systematic uncertainty in pressure calibration. 8,14 In previous high-pressure experimental studies, Kavner and Duffy concluded that Pt is one of the hardest FCC metals 16 and unusually strong for its shear modulus, 1 while Singh et al. found that Pt is weak. 7 The key difference between these studies may be grain size. In the former study, sample grain size was not characterized, but in the latter, much higher strength was observed for 20-nm Pt grains rela- tive to 300-nm grains. 7 The Hall-Petch effect, i.e., higher strength at nm-scale grain size, is consistent with this differ- ence. However, these two previous studies of Pt reached their different conclusions using different X-ray diffraction geometries (radial 16 vs. axial 7 ) and diagnostic techniques (diffraction line shifts 16 vs. line broadening 7 ). Diffraction-based measurements of deformation in a diamond anvil cell (DAC) can be performed in either an axial geometry (with incident X-rays along the loading axis) or in a radial geometry (with incident X-rays perpendicular to the loading axis). The radial geometry provides the most detailed measurement of strain at high pressures in the dia- mond anvil cell. Unlike the axial geometry used in the study by Singh et al., 7 this technique samples a wide range of crystallite orientations and is sensitive to elastic constants and texture as well as stress. Radial X-ray diffraction of plat- inum has been limited to pressures below 24 GPa so far. In this work, we re-examine the effect of grain size on the strength of platinum. We also extend measurements of the strength and deformation texture of Pt by X-ray diffraction in a radial geometry to 63 GPa. II. THEORETICAL BACKGROUND Compression of polycrystalline materials in the diamond anvil cell produces non-hydrostatic stresses due to the effects of both uniaxial loading (macro-stress) and of stress hetero- geneities at grain boundaries (micro-stress). 2,3 In the dia- mond anvil cell, the maximum stress, r 3 , is approximately aligned with the compression axis and the minimum stress, r 1 , is in the plane of the gasket. The differential stress, t, is a) Author to whom correspondence should be addressed. Electronic mail: susannah.dorfman@epfl.ch 0021-8979/2015/117(6)/065901/9/$30.00 V C 2015 AIP Publishing LLC 117, 065901-1 JOURNAL OF APPLIED PHYSICS 117, 065901 (2015) [This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP: 128.112.20.228 On: Fri, 03 Jul 2015 17:52:37

Upload: others

Post on 22-Aug-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Strength and texture of Pt compressed to 63GPaduffy.princeton.edu/sites/default/files/pdfs/2015/dorfman_jap_15.pdf · Strength and texture of Pt compressed to 63GPa Susannah M. Dorfman,1,a)

Strength and texture of Pt compressed to 63 GPa

Susannah M. Dorfman,1,a) Sean R. Shieh,2 and Thomas S. Duffy3

1Earth and Planetary Science Laboratory, Ecole Polytechnique F�ed�erale de Lausanne, Station 3, CH-1015Lausanne, Switzerland2Department of Earth Sciences, University of Western Ontario, London, Ontario N6A 5B7, Canada3Department of Geosciences, Princeton University, Princeton, New Jersey 08544, USA

(Received 18 September 2014; accepted 28 January 2015; published online 10 February 2015)

Angle- and energy-dispersive X-ray diffraction experiments in a radial geometry were performed

in the diamond anvil cell on polycrystalline platinum samples at pressures up to 63 GPa. Observed

yield strength and texture depend on grain size. For samples with 70–300-nm particle size, the yield

strength is 5–6 GPa at �60 GPa. Coarse-grained (�2-lm particles) Pt has a much lower yield

strength of 1–1.5 GPa at �60 GPa. Face-centered cubic metals Pt and Au have lower strength to

shear modulus ratio than body-centered cubic or hexagonal close-packed metals. While a 300-nm

particle sample exhibits the h110i texture expected of face-centered-cubic metals under

compression, smaller and larger particles show a weak mixed h110i and h100i texture under

compression. Differences in texture development may also occur due to deviations from uniaxial

stress under compression in the diamond anvil cell. VC 2015 AIP Publishing LLC.

[http://dx.doi.org/10.1063/1.4907866]

I. INTRODUCTION

The deformation behavior of materials is fundamental to

physics, engineering, and geoscience and is determined by

both elastic and plastic properties.1 While a variety of meth-

ods can be employed to independently measure elasticity

and plasticity at ambient conditions, X-ray diffraction is one

of the only established techniques that can be used to explore

these properties in situ at extreme pressures.2–4 Existing

diffraction-based methods use diffraction line shifts2–4 and

widths5,6 to constrain material deformation but typically suf-

fer from uncertainties due to confounding of multiple varia-

bles and limited control and measurement of experimental

conditions such as the applied stress.5,7 The uncertainties in

high-pressure measurements of strength and anisotropy can

be evaluated by systematic study of simple materials such as

platinum metal.

Platinum (Pt) is a face-centered cubic (FCC) metal and

does not undergo phase transitions up to multi-megabar pres-

sures.7,8 Due in part to its simple structure, Pt is widely used

in high-pressure experiments as a pressure calibrant and laser

absorber.9 Nanoparticles of Pt10 and other metals11 have also

been observed to have anomalously low compressibility, and

this has possible implications for pressure calibration using

metal powders. Recently, Pt has also been used as a devia-

toric strain marker: the success of laser annealing or pressure

media in minimizing deviatoric stresses has been judged by

comparison to the strength of Pt as a function of pres-

sure.8,12–15 The sensitivity of Pt to deviatoric stress due to its

strength and elastic properties is also a source of systematic

uncertainty in pressure calibration.8,14

In previous high-pressure experimental studies, Kavner

and Duffy concluded that Pt is one of the hardest FCC

metals16 and unusually strong for its shear modulus,1 while

Singh et al. found that Pt is weak.7 The key difference

between these studies may be grain size. In the former study,

sample grain size was not characterized, but in the latter,

much higher strength was observed for 20-nm Pt grains rela-

tive to 300-nm grains.7 The Hall-Petch effect, i.e., higher

strength at nm-scale grain size, is consistent with this differ-

ence. However, these two previous studies of Pt reached

their different conclusions using different X-ray diffraction

geometries (radial16 vs. axial7) and diagnostic techniques

(diffraction line shifts16 vs. line broadening7).

Diffraction-based measurements of deformation in a

diamond anvil cell (DAC) can be performed in either an

axial geometry (with incident X-rays along the loading axis)

or in a radial geometry (with incident X-rays perpendicular

to the loading axis). The radial geometry provides the most

detailed measurement of strain at high pressures in the dia-

mond anvil cell. Unlike the axial geometry used in the study

by Singh et al.,7 this technique samples a wide range of

crystallite orientations and is sensitive to elastic constants

and texture as well as stress. Radial X-ray diffraction of plat-

inum has been limited to pressures below 24 GPa so far. In

this work, we re-examine the effect of grain size on the

strength of platinum. We also extend measurements of the

strength and deformation texture of Pt by X-ray diffraction

in a radial geometry to 63 GPa.

II. THEORETICAL BACKGROUND

Compression of polycrystalline materials in the diamond

anvil cell produces non-hydrostatic stresses due to the effects

of both uniaxial loading (macro-stress) and of stress hetero-

geneities at grain boundaries (micro-stress).2,3 In the dia-

mond anvil cell, the maximum stress, r3, is approximately

aligned with the compression axis and the minimum stress,

r1, is in the plane of the gasket. The differential stress, t, is

a)Author to whom correspondence should be addressed. Electronic mail:

[email protected]

0021-8979/2015/117(6)/065901/9/$30.00 VC 2015 AIP Publishing LLC117, 065901-1

JOURNAL OF APPLIED PHYSICS 117, 065901 (2015)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

128.112.20.228 On: Fri, 03 Jul 2015 17:52:37

Page 2: Strength and texture of Pt compressed to 63GPaduffy.princeton.edu/sites/default/files/pdfs/2015/dorfman_jap_15.pdf · Strength and texture of Pt compressed to 63GPa Susannah M. Dorfman,1,a)

defined as the difference between these stresses, r3 � r1.

Differential stress reflects the non-hydrostatic stress compo-

nents that result in deviatoric strain. For a given lattice plane

in a crystallite, the deviatoric strain depends on its orienta-

tion relative to the compression axis and the elastic anisot-

ropy of the crystal. Microscopic deviatoric stresses also

produce micro-strains in the polycrystalline aggregate. Both

macro-strain and micro-strain are limited by the yield

strength of the material, which may vary with pressure, grain

size, and deformation rate.1 After yielding, the total strain is

a combination of elastic and plastic deformation.17–21

Lattice strain analysis of radial X-ray diffraction data

has been described in detail elsewhere;2–4 here we summa-

rize the major relations used to calculate strain and stress in

energy- and angle-dispersive diffraction modes. The varia-

tion of measured lattice spacings, dm(hkl), can be described

by a d-spacing resulting from the hydrostatic component of

stress, dp(hkl), plus a non-hydrostatic component:

dmðhklÞ ¼ dpðhklÞ½1þ ð1� 3 cos2wÞQðhklÞ�: (1)

The angle, w, is defined between the lattice plane normal

and the direction of maximum stress. w is related to the dif-

fraction angle, h, and the azimuthal orientation of the dia-

mond anvil cell relative to the detector, d, (defined as 0� at

the diamond anvil cell compression axis) as cos w¼ cos dcos h (Figure 1). Q(hkl) is a function of t and the shear modu-

lus G, where GR(hkl) is the modulus under iso-stress (Reuss

bound) conditions and GV under iso-strain (Voigt bound):

Q hklð Þ ¼ t

3

a

2GR hklð Þ þ1� a2GV

� �: (2)

The parameter a is a measure of the degree of stress/

strain continuity across crystallite boundaries. Under iso-

stress conditions, a¼ 1; measured values of a typically range

from 0.5 to 1, although it has been suggested that values

greater than 1 may occur depending on the elastic anisotropy

of the material.22,23 For materials of cubic symmetry, the fol-

lowing equations relate GR(hkl) and GV to the elastic compli-

ances, Sij, and the Miller indices of the lattice plane:

1

2GR¼ S11 � S12 � 3SC hklð Þ; (3)

S ¼ S11 � S12 �S44

2; (4)

1

2GV¼ 5

2

S11 � S12ð ÞS44

3 S11 � S12ð Þ þ S44½ � ; (5)

C hklð Þ ¼ h2k2 þ k2l2 þ l2h2

h2 þ k2 þ l2ð Þ2: (6)

In Eqs. (3) and (4), S is a measure of the elastic anisotropy of

a cubic crystal.

As a complement to lattice strain analysis, diffraction

line widths can be used to constrain micro-strain, g, and

grain size, D.5,7 For diffraction using monochromatic

X-rays,24,25

whkl cos hhklð Þ2 ¼ kD

� �2

þ g2hkl sin2hhkl: (7)

In this equation, whkl is the full width at half maximum

of the diffraction peak and k is the X-ray wavelength. The

slope of the linear relationship between (whklcos hhkl)2 and

FIG. 1. Schematic of radial-geometry

X-ray diffraction experiments in

energy-dispersive and angle-dispersive

modes.

065901-2 Dorfman, Shieh, and Duffy J. Appl. Phys. 117, 065901 (2015)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

128.112.20.228 On: Fri, 03 Jul 2015 17:52:37

Page 3: Strength and texture of Pt compressed to 63GPaduffy.princeton.edu/sites/default/files/pdfs/2015/dorfman_jap_15.pdf · Strength and texture of Pt compressed to 63GPa Susannah M. Dorfman,1,a)

sin2 hhkl terms is used to determine average micro-strain,

while the intercept provides the grain size. An additional cor-

rection is needed to account for instrumental broadening.

Lattice strain theory has been successfully applied to the

high-pressure deformation of metals, ceramics, and miner-

als.1 Both stress and elastic constants have been derived for

materials undergoing purely elastic deformation.16,23,26–29

However, on compression beyond the yield strength, elastic

constants were observed to deviate from values derived from

other techniques such as ultrasonic interferometry or

Brillouin spectroscopy.17,27 For this case, more complex

models have been developed to account for plastic deforma-

tion.30 When the elastic moduli are known independently,

the lattice strain technique has been demonstrated to provide

accurate estimates of non-hydrostatic stress.31

III. METHOD

Polycrystalline Pt samples were obtained with three dif-

ferent particle sizes. Particle shapes and sizes were charac-

terized by scanning electron microscopy. The bulk sample

(designated B) is made up of irregular platelets with median

Feret diameter �2 lm and minimum dimension �450 nm

(supplementary Figure 1).32 Nano-grained samples N1 and

N2 are sub-spherical particles with median diameters 300

and 70 nm and minimum diameters 70 and 40 nm, respec-

tively (supplementary Figures 2 and 3).32 We note that parti-

cle sizes measured by microscopy represent upper bounds on

grain size of these samples.

Each powdered sample was packed in a 100-lm sample

chamber drilled in an X-ray transparent beryllium (Be) gas-

ket. X-ray diffraction of the gasket material shows doublets

for each Be peak, suggesting the presence of both Be and an

exsolved Be-rich alloy phase due to impurities in the gasket.

We also observed hexagonal BeO, most likely resulting from

oxidation of Be grain surfaces. All samples were loaded in

symmetric or panoramic diamond anvil cells with 300-lm

diameter diamond culets. A 10–20-lm diameter piece of Mo

was loaded at the center of the diamond culet on one side as

position marker. Upon compression, sample chamber diame-

ters typically shrank to 50–80 lm.

Both energy-dispersive (EDXD) and angle-dispersive

(ADXD) X-ray diffraction experiments were performed at

beamline X17C of the National Synchrotron Light Source.

Both methods were used for experiments with samples B and

N1. Sample N2 was examined only with the ADXD method.

Diamond anvil cells were mounted such that incident X-rays

entered and exited the sample radially through the Be gasket

(Figure 1).

In energy-dispersive X-ray diffraction experiments in a

radial geometry (Figure 1),33 the symmetric DAC was

rotated around the beam axis to angles d¼ 0�, 24�, 35�, 45�,55�, 66�, and 90�. The incident X-ray beam size was �50

� 50 lm. Diffracted X-rays were collected using a solid-

state Ge detector positioned at diffraction angle

2h¼ 10�–12�. Detector energy calibration was performed

using fluorescent standards, and detector angle was cali-

brated with an Au standard.

In radial angle-dispersive X-ray diffraction experiments

(Figure 1),34 the panoramic DAC was mounted at a fixed

angle and the diffracted X-rays were collected on a charge-

coupled device (MarCCD). Due to space limitations in the

X17C experimental station, the detector was laterally

displaced relative to the beam center. As a result, diffraction

Inte

nsity

(ar

bitr

ary

units

)

2.42.22.01.81.61.41.21.0

d-spacing (Angstroms)

111

200

220

222

311

400

Be

Be

Be

Be

Be

Mo

Mo M

o

Be

0

24

35

45

55

66

48 GPa

δ=90º

FIG. 2. Energy-dispersive diffraction patterns for bulk Pt (sample B) at

48 GPa.

FIG. 3. Angle-dispersive diffraction pattern for nano-grained Pt sample N1

at 52 GPa. X and Y axes describe the locations of pixels on the CCD detec-

tor. Each pixel is 79.6 � 79.6 lm.

065901-3 Dorfman, Shieh, and Duffy J. Appl. Phys. 117, 065901 (2015)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

128.112.20.228 On: Fri, 03 Jul 2015 17:52:37

Page 4: Strength and texture of Pt compressed to 63GPaduffy.princeton.edu/sites/default/files/pdfs/2015/dorfman_jap_15.pdf · Strength and texture of Pt compressed to 63GPa Susannah M. Dorfman,1,a)

was observed at 2h values of more than 25�, but over only a

limited azimuthal range of 80�–190�. Detector distance and

orientation were calibrated with CeO2 and Au standards. The

wavelength of the monochromatic X-rays was 0.4066 A, and

beam size was �20 � 30 lm.

IV. RESULTS

Samples were compressed to maximum pressures of

48–63 GPa. After each pressure increase, samples were

allowed to relax for 30–120 min, or until the measured pres-

sure stabilized. Example diffraction patterns from an EDXD

measurement of sample B at 48 GPa are shown in Figure 2.

An example two-dimensional ADXD pattern of sample N1

at 52 GPa is shown in Figure 3. All diffraction lines corre-

spond to the Pt sample, Mo marker, and Be and BeO in the

gasket. Gasket diffraction peaks are highly textured and

invariant with pressure. Lattice d-spacings of Pt and the

weak Mo marker change with azimuth d from the loading

axis due to differences in stress. Texture development in Pt

can also be observed as intensity variations along the azi-

muthal angle, d, with different maxima for different lattice

orientations.

EDXD patterns were analyzed by fitting peaks to Voigt

lineshapes. ADXD patterns were integrated using FIT2D35 at

azimuthal intervals of 4�–5� over a range from d¼ 45�–145�,the maximum azimuthal range at which the (222) line is

observed at all pressures. Full-profile strain and texture anal-

ysis were performed with MAUD software36 (details are

given in supplementary material).32 Pressure was determined

from the lattice parameter of Pt as measured at d¼ 55�, the

angle at which the measured d-spacing corresponds to the

hydrostatic d-spacing under uniaxial conditions.3 As in

the previous studies of strength and deformation of Pt,7,16 we

calibrate pressure with the equation of state of Holmes et al.9

A. Line shift analysis

For each diffraction line, a linear relationship between

lattice spacing and the orientation function 1–3cos2w is

observed (Figure 4 and supplementary Figure 4).32

Curvature in this relationship may be due either to stress var-

iations in the diffraction volume or non-uniaxial stress. For

EDXD data, Q(hkl) was determined from the slope and inter-

cept of this linear relationship as in Eq. (1). For ADXD data,

Q(hkl) in Pt was refined from full-profiles in MAUD (Figure

5). Pt diffraction peak widths were fit by refining grain size

and micro-strain. We fit each Q(hkl) and the orientation of

the maximum strain. The angle between the maximum strain

and the compression axis, b, was 6��27� at maximum

FIG. 4. Lattice parameter and d-spac-

ings for Pt from EDXD for nano-

grained sample N1 at 61 GPa.

Spacings for (111) planes are shown in

red dots, (200)¼ blue squares,

(220)¼ yellow up-pointing triangles,

(311)¼ green down-pointing triangles,

and (222)¼ pink diamonds. Ranges of

d-spacings on plots are scaled to a con-

stant percentage of the value at

w¼ 55�.

065901-4 Dorfman, Shieh, and Duffy J. Appl. Phys. 117, 065901 (2015)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

128.112.20.228 On: Fri, 03 Jul 2015 17:52:37

Page 5: Strength and texture of Pt compressed to 63GPaduffy.princeton.edu/sites/default/files/pdfs/2015/dorfman_jap_15.pdf · Strength and texture of Pt compressed to 63GPa Susannah M. Dorfman,1,a)

pressures reached for both ADXD and EDXD measurements

(supplementary Figure 4).32 This deviation from uniaxial

stress has been observed in other studies37 and may be due to

deformation of the gasket hole at low pressures and bending

of diamond anvils at higher pressures. The relative impor-

tance of these two mechanisms changes from run to run. In

one sample, the angle b was observed to increase to 20�–30�

by �10 GPa (gasket deformation), while other samples

exhibited more continuous increase at higher pressures (dia-

mond bending). By correcting the orientation of the ADXD

data, the curvature of lattice spacing vs. 1–3cos2w can be

almost entirely removed (supplementary Figure 4),32 sug-

gesting that the stress orientation is the dominant cause of

observed non-linearity.

The observed lattice strain increases strongly with pres-

sure until the sample yields, above which pressure the lattice

strain increases at a weaker rate (Figure 6(a)). Difference

between EDXD and ADXD measurements represents run-to-

run differences in the magnitude and orientation of devia-

toric stress. Higher strain was observed for the nano-grained

samples N1 and N2 at all pressures relative to the bulk sam-

ple. For the bulk sample, we could not resolve the pressure

at which yielding occurred due to the low strain exhibited

and the size of the pressure steps. For samples N1 and N2,

yielding occurred at �9 GPa, similar to previous observa-

tions of sub-micron Pt.7,16

According to purely elastic models, the strain function

Q(hkl) is linearly related to the crystal orientation function

C(hkl) as in Eqs. (2)–(6). The slope of this linear relationship

is related to the elastic anisotropy, S, of Pt as well as the

magnitude of the differential stress. Due to the high elastic

anisotropy of Pt,8 the observed Q(hkl) depends strongly on

lattice plane (Figure 6(b)). The linear relationship between

Q(hkl) and C(hkl) can be used to determine the change in

elastic anisotropy with pressure. The measured elastic anisot-

ropy is lower than that predicted for Pt by ab initio density

functional theory.38 This discrepancy is due to effects of

plastic deformation on Q(hkl). Plastic deformation can be

neglected if elastic compliances Sij are fixed for analysis of

stress conditions.

While the elastic constants of Pt are well-constrained at

1 bar and high temperature, their evolution with pressure has

not been determined by experiment. Only the elastic modu-

lus C44 was measured by ultrasonic interferometry (and only

to pressures <0.3 GPa).39 The elastic constants of Pt were

computed by ab initio density functional theory up to

660 GPa.38 As the authors noted, the predicted ambient pres-

sure derivative of the C44 modulus, C440 ¼ dC44/P, is signifi-

cantly larger than the ultrasonic value.39 In this study (as in

the previous study by Singh et al.7), we use the ultrasonic

value for C440 and other zero-pressure elastic moduli and

FIG. 5. Full-profile refinement of Pt sample N1 at 52 GPa from MAUD pro-

gram (above¼fit, below¼ binned data). Pt diffraction lines are curved and

smoothly textured due to non-hydrostatic stress state and plastic deforma-

tion. The diffraction pattern from the gasket is composed of Be doublets

with large, randomly oriented grains and BeO with finer grains.

FIG. 6. a) For sample N1 (300 nm particles), the strain function Q(hkl) in

the (111) diffraction peak for ADXD (dots) and EDXD (squares) runs. The

change in slope at �9 GPa indicates yielding.1 (b) Variation of Q(hkl) for

different diffraction peaks at 61 GPa.

065901-5 Dorfman, Shieh, and Duffy J. Appl. Phys. 117, 065901 (2015)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

128.112.20.228 On: Fri, 03 Jul 2015 17:52:37

Page 6: Strength and texture of Pt compressed to 63GPaduffy.princeton.edu/sites/default/files/pdfs/2015/dorfman_jap_15.pdf · Strength and texture of Pt compressed to 63GPa Susannah M. Dorfman,1,a)

pressure derivatives from density functional theory.38,39 The

computed and ultrasonic values for zero-pressure elastic con-

stants and their pressure derivatives were extrapolated using

the finite strain equations.40 One additional parameter is

required to solve Eqs. (2)–(6): the stress-strain continuity pa-

rameter, a. Previous analysis of both peak width and line

shifts data for Pt suggested that the best value for a is 0.6.7

We calculate differential stress assuming a to be either 0.6 or

1 and observe a difference in t of �1.5%, much lower than

other uncertainties. Differential stress assuming a¼ 0.6 is

plotted in Figure 7.

Differential stress in the bulk Pt sample reaches a maxi-

mum of 1–1.5 GPa at 60 GPa (Figure 7). This is consistent

with axial diffraction measurements of linewidths that are

used to constrain the strength of Pt with 300-nm grain size7

but considerably lower than previous radial diffraction meas-

urements of line shifts.16 We concur with Singh et al.7 that

the difference between these studies is likely due to grain

size differences. The nano-scale samples, N1 and N2, exam-

ined in this study exhibited much higher yield strengths of

5–6 GPa at 60 GPa. These stresses are comparable to those

observed in Ref. 16 and slightly lower than those observed in

Ref. 7 for 20-nm grains, consistent with the particle size of

70–300 nm.

B. Peak width analysis

Diffraction peak widths typically increase with pressure

due to increasing micro-strain and stress gradient across the

sample. Grain size reduction may also increase peak widths

for nanocrystalline samples. Peak broadening in these

experiments was analyzed by peak fitting as well as analysis

using MAUD.

Observed peak widths were consistent with grain sizes

similar to particle sizes observed by SEM and remained

approximately constant with pressure (Figure 8 and supple-

mentary Figure 5).32 For �300-nm N1 sample particles,

grain size determined by peak widths was 200–400-nm.

Diffraction peak widths of the bulk-grained sample were

within error of the instrument resolution at loading. Fits to

the bulk sample peak widths at 20–60 GPa yield a grain size

of �1 lm, consistent with the particle size observed by

SEM. For the two finer-grained samples, the width of dif-

fraction peaks increased below the yield stress, decreased

immediately after yielding, and remained constant above

�10–20 GPa (Figure 8). Changes in peak widths are most

likely due to buildup of micro-strain before yielding and

release of micro-strain through plastic flow after yielding.

Other possible mechanisms for peak sharpening are grain

growth, which is unlikely at 300 K, or collapse of the gasket

hole, reducing the pressure gradient over the measured sam-

ple.41 No significant peak broadening is observed above the

12

10

8

6

4

2

0

Diff

eren

tial S

tres

s (G

Pa)

706050403020100

Pressure (GPa)

Nano-grains

Bulk grains

Method:Axial diffraction

20-nm grains (Ref. 7) 300-nm grains (Ref. 7)

Radial diffraction (Energy-dispersive) 300-nm particles (This study) Size unknown (Ref. 16) 2-μm particles (This study)

Radial diffraction (Angle-dispersive) 70-nm particles (This study) 300-nm particles (This study) 2-μm particles (This study)

FIG. 7. Maximum differential stress, or yield strength, measured for Pt sam-

ples of different grain size. Data from this work are shown with filled sym-

bols, with squares for EDXD and circles for ADXD (red¼ sample B,

green¼ sample N1, and cyan¼ sample N2). Data from previous work are

20-nm (open diamonds) and 300-nm (filled diamonds) grain sizes studied by

axial diffraction peak width analysis in Ref. 7 and unknown-size grains

(open squares) studied by EDXD in Ref. 16. Stress was calculated with

a¼ 0.6.

0.5

0.4

0.3

0.2

0.1

0.0

Diff

ract

ion

peak

wid

th a

t ψ=

90°

(deg

rees

)

50403020100

Apparent Pressure at ψ=90° (GPa)

111200

70 nm

300 nm

2 μm

FIG. 8. Diffraction peak widths measured for each particle size

(red¼ sample B, green¼ sample N1, and cyan¼ sample N2) in ADXD pat-

terns at w¼ 90�, corresponding to the axial geometry. Apparent pressure is

computed from d-spacings observed at w¼ 90�. Due to lattice strain, d-spac-

ings are larger and apparent pressure lower than those determined at

w¼ 54.7�, the orientation of approximately hydrostatic stress, by up to 8,

12, and 17 GPa for samples B, N1, and N2, respectively.

065901-6 Dorfman, Shieh, and Duffy J. Appl. Phys. 117, 065901 (2015)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

128.112.20.228 On: Fri, 03 Jul 2015 17:52:37

Page 7: Strength and texture of Pt compressed to 63GPaduffy.princeton.edu/sites/default/files/pdfs/2015/dorfman_jap_15.pdf · Strength and texture of Pt compressed to 63GPa Susannah M. Dorfman,1,a)

yield stress indicating no significant decrease in grain size

due to yielding.

C. Texture analysis

Nano-particle samples initially exhibited continuous,

untextured diffraction rings. For the bulk sample, diffraction

rings are randomly textured due to large grain size relative to

the X-ray beam. Upon compression to comparable pressures

of 49–52 GPa (Figures 9(a)–9(c)), stronger textures were

observed for the �300-nm particle size than for smaller or

larger particles. In addition, the nature of the texture devel-

opment differs with particle size. In sample N1 (300-nm

size), the grains are oriented to exhibit a maximum in (111)

at minimum stress (Figure 9(b)). However, a different

texture was observed in the other two samples, with a mini-

mum in the (111) planes at the minimum stress orientation

(Figures 9(a) and 9(c)).

These textures were evaluated from ADXD experiments

by refinement in MAUD software (Figures 9(d)–9(f)). For

sample N1, the texture index is observed to increase strongly

from 13 GPa, just above the yield pressure, to 34 GPa, above

which pressure it is saturated at �3 maximum of a random

distribution (m.r.d.). The deformation texture in this sample

is a h110i fiber texture typical of FCC metals under compres-

sion.42 The other two samples are weakly textured, with

texture index �1.1 at maximum pressures. The maximum

intensity is found in the h100i direction, more typical of

tension in FCC metals at ambient pressure.42

Different textures observed for different particle sizes

indicate changes in deformation mechanisms. The main

deformation mechanisms in cold-compressed powders are dis-

location creep and grain boundary processes. Which of these

mechanisms dominates depends on multiple factors including

grain size, pressure, temperature, and strain rate. Dislocation

creep is thought to dominate deformation in coarse (>100-nm

grain size) metals and produces strong lattice-preferred orien-

tation under compression. With decreasing grain size, disloca-

tions become increasingly unfavorable, and grain boundary

sliding and rotation dominate deformation in metals with

grain size smaller than �20-nm.43–46 Increasing pressure may

support formation of dislocations even in 3-nm particles.44

Unlike dislocation creep, grain rotation was reported to ran-

domize texture.15,44,47,48 In addition, this change in deforma-

tion mechanism at small grain sizes was suggested to cause a

decrease in strength.45,46

The bulk Pt sample in this study surprisingly does not

develop a strong compression texture, yet grain-boundary-

mediated processes are not expected to be dominant. The

differential stress supported by the sample is low, however,

and confirms that the bulk grains are relatively weak. In

addition, the irregular particles in this sample should inhibit

grain boundary sliding and rotation. The weak tension

texture in this sample may reflect low statistics due to large

grain size. Random texture from loading of the bulk sample

persists at the maximum pressures examined.

The elastic strain observed in the two Pt nano-powders

in this study is similar, but texture development differs. The

FIG. 9. a)–(c) Strain and texture in Pt 111 line at 49–52 GPa for samples with different particle sizes. The red horizontal line marks the orientation of the gasket

plane (90� from the DAC axis), which is the minimum stress direction for uniaxial compression. (d)–(f) Inverse pole figures showing texture development

along the compression axis at 49–52 GPa in each sample. Texture was fit in MAUD with the E-WIMV model and fixed cylindrical geometry (see supplemen-

tary material).32

065901-7 Dorfman, Shieh, and Duffy J. Appl. Phys. 117, 065901 (2015)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

128.112.20.228 On: Fri, 03 Jul 2015 17:52:37

Page 8: Strength and texture of Pt compressed to 63GPaduffy.princeton.edu/sites/default/files/pdfs/2015/dorfman_jap_15.pdf · Strength and texture of Pt compressed to 63GPa Susannah M. Dorfman,1,a)

70-nm and 300-nm particles are both stronger than previ-

ously measured 300-nm grain size Pt but weaker than 20-nm

grain size Pt.7 Since only the particle size is known, each

sample may be composed of clumps of smaller grains. The

weaker texture in the 70-nm particles may indicate that grain

sizes in this sample are small enough to promote grain

boundary sliding and rotation as a primary mechanism of

plastic deformation. Alternatively, the stronger texture in the

300-nm-particle sample may be due to the large deviation

from uniaxial stress in this experimental run (supplementary

Figure 4(a)). The h100i texture in the 70-nm and bulk Pt

samples would then be more representative for Pt under pure

uniaxial stress.

V. IMPLICATIONS

While earlier radial diffraction experiments16 reported

that platinum could be unusually strong for an FCC metal

with low shear modulus, consideration of grain size effects7

brings the behavior of this metal into agreement with the

general trends at high pressure displayed by other metals.

Our radial diffraction results for multiple grain sizes of Pt

confirm the low strength of bulk Pt observed by the axial dif-

fraction method by Singh et al.7 Like Au, another FCC

metal, Pt exhibits a differential stress/shear modulus ratio

lower than that of metals with body-centered cubic (BCC) or

hexagonal close-packed (HCP) structures (Figure 10).

The strong dependence of strength on grain size

increases the importance of careful characterization of grain

sizes of materials used as differential stress markers and cali-

brants. The sub-micron Pt samples examined in this work

were also used as differential stress markers in previous stud-

ies of the equation of state of various materials including

MgGeO3, CaF2, and Gd3Ga5O12.12,49,50 Lattice strain analy-

sis of Pt was performed using the axial geometry in these

cases. The effectiveness of Ne or He pressure media8 and

laser annealing12,49,50 for reducing deviatoric stresses at

Mbar pressures was evaluated by comparison to the yield

strength of the Pt marker. The high strength of sub-micron

Pt confirms that the observed differential stress in these pre-

vious studies was low relative to the yield strength, and

quasi-hydrostatic conditions were achieved. For all similar

measurements, the yield strength and grain size of the

marker are critical for this evaluation of deviatoric stress.

However, knowledge of the initial grain size of the

strain marker may not be sufficient to fully constrain the

strength at high pressures and temperatures. Grain size

changes with pressure and/or temperature may occur and be

difficult to detect. For example, reversible pressure-induced

grain size reduction has been reported previously but is not

yet understood;7,51 this may be due to the difficulty in sepa-

rating the effects of grain size, micro-strain, and plastic de-

formation on diffraction peak widths. In laser heating

studies, both reduced strength and grain growth are expected.

This uncertainty in grain size is thus a limitation to the use

of differential stress markers in diamond anvil cell

experiments.

This study provides an important test of different techni-

ques for characterizing deformation at high pressure. Our

results using multiple techniques demonstrate the consis-

tency of measurements of the strength of bulk and nano-

grained Pt by analysis of diffraction peak widths7 and both

energy-dispersive16 and angle-dispersive X-ray diffraction in

a radial geometry. The major sources of uncertainty in these

measurements are the poorly constrained high-pressure

elastic properties, stress-strain continuity factor, and non-

uniaxial stress in the diamond anvil cell. By measuring

diffraction across a wide and continuous range of orienta-

tions and stress conditions, radial ADXD provides the most

complete information for reducing these uncertainties.

VI. SUMMARY

Strength and deformation of Pt powder samples with

various grain sizes were examined at pressures up to 63 GPa

by radial diffraction experiments in both energy- and angle-

dispersive modes. For micron-scale grain size samples,

the yield strength of Pt reaches 1–1.5 GPa at �60 GPa.

For nanometer-scale grain sizes, the yield strength is much

larger: differential stress of 5–6 GPa is observed at

�60 GPa. These observations confirm that contrasting val-

ues of the strength of Pt in previous axial-geometry studies7

are the result of differences in grain size. Analysis of dif-

fraction peak widths suggests no significant grain size

reduction due to yielding or compression. Texture develop-

ment in the sample with �300-nm particles is strong and

typical of FCC metals under compression. For both larger

and smaller particle sizes, observed texture is weaker and

more typical of mixed compression and tension in FCC met-

als. Texture style may also depend on non-uniaxial stress.

Platinum, like FCC gold, exhibits a lower strength to shear

modulus ratio at high pressure than metals with BCC or

HCP structures including Mo, Re, and W. Smaller grain

sizes exhibit higher yield strength; characterization of the

grain size is thus critical to analysis of strain and stress in

high-pressure experiments.

3.0x10-2

2.5

2.0

1.5

1.0

0.5

0.0

Diff

eren

tial s

tres

s/S

hear

mod

ulus

706050403020100

Pressure (GPa)

Mo

Re

Au

W

Pt (this study)

FIG. 10. Strength to shear modulus ratio for metals at high pressures. Data

shown from this study are from the bulk sample. FCC Au26,52 and Pt are

weaker than BCC and HCP metals26,52,53 studied to date.

065901-8 Dorfman, Shieh, and Duffy J. Appl. Phys. 117, 065901 (2015)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

128.112.20.228 On: Fri, 03 Jul 2015 17:52:37

Page 9: Strength and texture of Pt compressed to 63GPaduffy.princeton.edu/sites/default/files/pdfs/2015/dorfman_jap_15.pdf · Strength and texture of Pt compressed to 63GPa Susannah M. Dorfman,1,a)

ACKNOWLEDGMENTS

Z. Chen, X. Hong, and S. Ghose supported beamline

operations at X17C. L. Miyagi provided helpful advice in

data analysis with MAUD. We thank A. K. Singh and two

anonymous reviewers for comments that improved this

manuscript. Electron microscopy was performed at the at the

Centre Interdisciplinaire de Microscopie Electronique at the

Ecole polytechnique f�ed�erale de Lausanne. This research

was partially supported by COMPRES, the Consortium for

Materials Properties Research in Earth Sciences under NSF

Cooperative Agreement No. EAR 10-43050. Use of the

National Synchrotron Light Source, Brookhaven National

Laboratory was supported by the U.S. Department of

Energy, Office of Science, Office of Basic Energy Sciences

under Contract No. DE-AC02-98CH10886.

1T. S. Duffy, AIP Conf. Proc. 955, 639–644 (2007).2A. K. Singh, J. Appl. Phys. 73, 4278 (1993).3A. K. Singh, C. Balasingh, H. Mao, R. J. Hemley, and J. Shu, J. Appl.

Phys. 83, 7567 (1998).4A. K. Singh, H. Mao, J. Shu, and R. J. Hemley, Phys. Rev. Lett. 80, 2157

(1998).5D. J. Weidner, Y. Wang, and M. T. Vaughan, Geophys. Res. Lett. 21, 753,

doi:10.1029/93GL03549 (1994).6A. K. Singh, E. Men�endez-Proupin, G. Guti�errez, Y. Akahama, and H.

Kawamura, J. Phys. Chem. Solids 67, 2192 (2006).7A. K. Singh, H.-P. Liermann, Y. Akahama, S. K. Saxena, and E.

Men�endez-Proupin, J. Appl. Phys. 103, 063524 (2008).8S. M. Dorfman, V. B. Prakapenka, Y. Meng, and T. S. Duffy, J. Geophys.

Res. 117, B08210, doi:10.1029/2012JB009292 (2012).9N. C. Holmes, J. A. Moriarty, G. R. Gathers, and W. J. Nellis, J. Appl.

Phys. 66, 2962 (1989).10A. S. Mikheykin, V. P. Dmitriev, S. V. Chagovets, A. B. Kuriganova, N.

V. Smirnova, and I. N. Leontyev, Appl. Phys. Lett. 101, 173111 (2012).11Q. F. Gu, G. Krauss, W. Steurer, F. Gramm, and A. Cervellino, Phys. Rev.

Lett. 100, 045502 (2008).12C. E. Runge, A. Kubo, B. Kiefer, Y. Meng, V. B. Prakapenka, G. Shen, R.

J. Cava, and T. S. Duffy, Phys. Chem. Miner. 33, 699 (2006).13Z. Mao, J. F. Lin, H. P. Scott, H. C. Watson, V. B. Prakapenka, Y. Xiao,

P. Chow, and C. McCammon, Earth Planet. Sci. Lett. 309, 179 (2011).14A. K. Singh, D. Andrault, and P. Bouvier, Phys. Earth Planet. Int.

208–209, 1 (2012).15B. Chen, K. Lutker, J. Lei, J. Yan, S. Yang, and H. Mao, Proc. Natl. Acad.

Sci. U.S.A. 111, 3350 (2014).16A. Kavner and T. S. Duffy, Phys. Rev. B 68, 144101 (2003).17D. J. Weidner, L. Li, M. Davis, and J. Chen, Geophys. Res. Lett. 31,

L06621, doi:10.1029/2003GL019090 (2004).18S. Merkel, J. Phys.: Condens. Matter 18, S949 (2006).19D. J. Weidner and L. Li, J. Phys.: Condens. Matter 18, S1061 (2006).

20D. J. Weidner, M. T. Vaughan, L. Wang, H. Long, L. Li, N. A. Dixon, and

W. B. Durham, Rev. Sci. Instrum. 81, 013903 (2010).21P. Raterron and S. Merkel, J. Synchrotron Radiat. 16, 748 (2009).22A. K. Singh, J. Appl. Phys. 106, 043514 (2009).23A. K. Singh and H.-P. Liermann, J. Appl. Phys. 109, 113539 (2011).24A. R. Stokes and A. J. C. Wilson, Proc. Phys. Soc. 56, 174 (1944).25J. I. Langford, J. Appl. Crystallogr. 4, 164 (1971).26T. S. Duffy, G. Shen, J. Shu, H.-K. Mao, R. J. Hemley, and A. K. Singh,

J. Appl. Phys. 86, 6729 (1999).27S. Speziale, S. R. Shieh, and T. S. Duffy, J. Geophys. Res. 111, B02203,

doi:10.1029/2005JB003823 (2006).28A. Kavner, Phys. Rev. B 77, 224102 (2008).29A. Kavner, Earth Planet. Sci. Lett. 214, 645 (2003).30S. Merkel, C. Tom�e, and H.-R. Wenk, Phys. Rev. B 79, 064110 (2009).31P. Raterron, S. Merkel, and C. W. Holyoke, Rev. Sci. Instrum. 84, 043906

(2013).32See supplementary material at http://dx.doi.org/10.1063/1.4907866 for

details of stress and texture modeling, scanning electron micrographs of

starting materials, and additional figures on lattice strain and peak

broadening.33H.-K. Mao and R. J. Hemley, High Pressure Res. 14, 257 (1996).34S. Merkel, H. R. Wenk, J. Shu, G. Shen, P. Gillet, H. Mao, and R. J.

Hemley, J. Geophys. Res. 107, 2271, doi:10.1029/2001JB000920 (2002).35A. P. Hammersley, S. O. Svensson, M. Hanfland, A. N. Fitch, and D.

H€ausermann, High Pressure Res. 14, 235 (1996).36L. Lutterotti, S. Matthies, and H. R. Wenk, Commission on Powder

Diffraction Newsletter 21, 14–15 (1999).37S. Merkel, N. Miyajima, D. Antonangeli, G. Fiquet, and T. Yagi, J. Appl.

Phys. 100, 023510 (2006).38E. Men�endez-Proupin and A. K. Singh, Phys. Rev. B 76, 054117 (2007).39S. N. Biswas, M. J. P. Muringer, and C. A. Ten Seldam, Phys. Status

Solidi A 141, 361 (1994).40F. Birch, J. Geophys. Res. 83, 1257, doi:10.1029/JB083iB03p01257 (1978).41K. Takemura, J. Appl. Phys. 89, 662 (2001).42U. F. Kocks, Texture and Anisotropy: Preferred Orientations in

Polycrystals and Their Effect on Materials Properties (Cambridge

University Press, 2000).43Z. Shan, E. A. Stach, J. M. K. Wiezorek, J. A. Knapp, D. M. Follstaedt,

and S. X. Mao, Science 305, 654 (2004).44B. Chen, K. Lutker, S. V. Raju, J. Yan, W. Kanitpanyacharoen, J. Lei, S.

Yang, H.-R. Wenk, H. Mao, and Q. Williams, Science 338, 1448 (2012).45J. Schiøtz, F. D. Di Tolla, and K. W. Jacobsen, Nature 391, 561 (1998).46J. Schiøtz and K. W. Jacobsen, Science 301, 1357 (2003).47J. Weissm€uller and J. Markmann, Adv. Eng. Mater. 7, 202 (2005).48K. E. Harris, V. V. Singh, and A. H. King, Acta Mater. 46, 2623 (1998).49S. M. Dorfman, F. Jiang, Z. Mao, A. Kubo, Y. Meng, V. B. Prakapenka,

and T. S. Duffy, Phys. Rev. B 81, 174121 (2010).50Z. Mao, S. M. Dorfman, S. R. Shieh, J. F. Lin, V. B. Prakapenka, Y.

Meng, and T. S. Duffy, Phys. Rev. B 83, 054114 (2011).51A. K. Singh, H. P. Liermann, S. K. Saxena, H. K. Mao, and S. U. Devi,

J. Phys.: Condens. Matter 18, S969 (2006).52T. S. Duffy, G. Shen, D. L. Heinz, J. Shu, Y. Ma, H.-K. Mao, R. J.

Hemley, and A. K. Singh, Phys. Rev. B 60, 15063 (1999).53D. He and T. S. Duffy, Phys. Rev. B 73, 134106 (2006).

065901-9 Dorfman, Shieh, and Duffy J. Appl. Phys. 117, 065901 (2015)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:

128.112.20.228 On: Fri, 03 Jul 2015 17:52:37