• Aucun résultat trouvé

Revealing ‘invisible’ defects in implanted material with 3D Bragg ptychography

N/A
N/A
Protected

Academic year: 2021

Partager "Revealing ‘invisible’ defects in implanted material with 3D Bragg ptychography"

Copied!
33
0
0

Texte intégral

(1)

HAL Id: hal-02925871

https://hal-amu.archives-ouvertes.fr/hal-02925871

Preprint submitted on 31 Aug 2020

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Revealing ‘invisible’ defects in implanted material with 3D Bragg ptychography

Peng Li, Nicholas Phillips, Steven Leake, Marc Allain, Felix Hofmann, Virginie Chamard

To cite this version:

Peng Li, Nicholas Phillips, Steven Leake, Marc Allain, Felix Hofmann, et al.. Revealing ‘invisible’

defects in implanted material with 3D Bragg ptychography. 2020. �hal-02925871�

(2)

Revealing ‘invisible’ defects in implanted material

1

with 3D Bragg ptychography

2

Peng Li1, Nicholas W. Phillips2, Steven Leake3, Marc Allain1, Felix Hofmann2, 3

Virginie Chamard1,* 4

5

1 Aix-Marseille Univ, CNRS, Centrale Marseille, Institut Fresnel, Marseille, France.

6

2 Department of Engineering Science, University of Oxford, Parks Road, Oxford, OX1 3PJ, 7

8 UK

3 ESRF – The European Synchrotron, 71 Avenue des Martyrs, 38000 Grenoble, France.

9

*Correspondence to: V. Chamard, virginie.chamard@fresnel.fr 10

11 12

Abstract 13

Small ion-irradiation-induced defects can dramatically alter material properties and speed up 14

degradation. Unfortunately, most of the defects irradiation creates are below the visibility 15

limit of state-of-the-art microscopy. As such, our understanding of their impact is largely 16

based on simulations with major unknowns. Here we present a novel x-ray crystalline 17

microscopy approach, able to image with high sensitivity, nano-scale 3D resolution and 18

extended field of view, the lattice strains and tilts in crystalline materials. Using this 19

enhanced Bragg ptychography tool, we study the damage helium-ion-irradiation produces in 20

tungsten, revealing a series of crystalline details in the 3D sample. Our results show that few- 21

atom-large ‘invisible’ defects are likely isotropic in orientation and homogeneously 22

distributed. No defect-denuded region is observed close to the grain boundary. These results 23

open up exciting perspectives for the modelling of irradiation damage and the detailed 24

analysis of crystalline properties in complex materials.

25 26 27

(3)

Introduction

28

29

Atomic defects play a fundamental role in controlling the mechanical and physical properties 30

of crystalline materials, resulting in critical hurdles for advanced applications, as epitomized 31

in e.g., energy generation (nuclear1 or photo-voltaic2), energy storage,3 aerospace,4 32

micromechanics5 and semiconductor miniaturization.6 Native or induced defects localise 33

distortion of the crystal lattice and thereby reduce the overall strain energy.7 Conversely, the 34

behaviour of defects is strongly dependent on the microstructural environment, which 35

provides fantastic potential for tuning material properties.8 36

Understanding and exploiting defects in crystalline materials requires probing the material 37

structure from atomic- to macro-scale. At near atomic resolution, transmission electron 38

microscopy (TEM) is an essential method, which allows the direct visualisation of lattice 39

defects in two dimensions (2D) and even three dimensions (3D).9,10 While it is also sensitive 40

to the associated lattice strains, the investigations of these are restricted to 2D and only 41

possible for a small subset of samples (e.g., straight dislocations of edge character11). This is 42

detrimental for the understanding of the defect-defect interplays, since defects interact via 43

their 3D strain fields and resulting stresses.12 In inherently thin TEM samples, defects may 44

also be lost to nearby surfaces that act as strong defect sinks, thus reducing the apparent 45

defect density.13 46

A further challenge concerns the visibility of small defects: In structures containing more 47

than a few thousand atoms, TEM is insensitive to defects smaller than ~1.5 nm.14 However 48

crystals that contain a population of rather small defects with broad size distribution 49

constitute a large class of materials, such as the ones resulting from intense irradiation 50

exposure (materials for future fusion and fission technologies,1 accelerator targets15 or 51

modified surfaces for biocompatibility16). Defects produced by irradiation range from single 52

atom defects to clusters of tens of nanometres in size, where the number density of defects is 53

typically linked to the defect size by a power law with a negative exponent, i.e., the vast 54

majority of defects are below the visibility limit.17 Although small, these “invisible” defects 55

can dramatically change the mechanical properties of the material18 or its thermal transport,19 56

and may lead to irradiation-induced dimensional change.20,21 Recent simulations have 57

successfully predicted the presence and evolution of large populations of “invisible” defects 58

during irradiation.22 However, these predictions remain unverified by experiments. As such 59

there is an urgent need for techniques able to determine the “invisible” defect number density 60

and spatial distribution.

61

(4)

A promising alternative is to infer the concentration of those small defects by measuring the 62

distortions (i.e., strains) they cause in the crystal lattice. This idea has been successfully 63

demonstrated using micro-beam Laue diffraction.20,21 However the spatial resolution of this 64

technique (~0.5 µm in 3D) is insufficient to resolve the nano-scale spatial heterogeneities and 65

defect clustering predicted by simulations.22 66

In this context, we foresee that x-ray lens-less microscopy could play a major role.23 Since its 67

first demonstration in 2001,24 x-ray Bragg coherent diffraction imaging has been proven as a 68

powerful method to investigate crystalline properties of materials25 in various sample 69

environments.26,27 X-ray Bragg ptychography (BP), a recently developed method, combines 70

sensitivity to the atomic displacements of lattice planes and outstanding imaging 71

performances of ptychography. Thereby, it provides the means to image, in 3D, extended 72

crystalline materials, with high sensitivity to weak crystalline displacements and strong 73

robustness to large strain fields, those specifications being mandatory to cope with the 74

distinctive crystalline features of strain induced defects and native strains of e.g., irradiated 75

materials. To date, BP with 3D spatial resolution of 10 - 50 nm, and strain resolution on the 76

order of 10-4 has been demonstrated,28,29 and applied to complex problems in material 77

science, e.g., anti-phase domain boundaries in metallic alloy,30 stacking faults in 78

semiconductor quantum wire,31 and crystalline domains in biomineral32. The advent of 4th 79

generation synchrotrons, where BP will become available at several beamlines, should 80

provide a broader access to this microscopy method.

81

X-ray BP makes use of a series of 3D Bragg diffraction intensity data sets, measured in the 82

vicinity of a Bragg reflection, obtained by scanning the sample across a localised illumination 83

beam. The sample motion is designed to ensure a significant amount of probe overlap, the 84

redundancy in the data set allowing retrieving the otherwise inaccessible phase of the 85

diffracted field from the set of diffracted intensities.33 As long as the probe is known, the 86

phase recovery is performed with iterative algorithms, which further give access to the 87

sample 3D image. An effective complex-valued electron density34, specifically designed to 88

account for the Bragg geometry, is used to describe the sample as a function of r, the 3D 89

spatial coordinate. Its amplitude ⎪ρ(r)⎪, provides information about the morphology (or 90

density) of the scattering crystal domain. The spatially varying phase φhkl(r) associated with a 91

specific hkl Bragg vector Qhkl is linked to the atomic displacement field in the crystal, u(r), 92

by φhkl(r) = Qhkl · u(r).

93

(5)

However, current BP approaches suffer from substantial limitations. They require the 94

acquisition of an extended data set, which relies on the stability of the experimental set-up 95

over long measurement times (typically 6-12 hours). This necessitates the samples be 96

resistant to radiation damage, a particular challenge for biologically relevant materials.

97

Moreover, BP requires the 3D probe to be known prior to the sample electron density 98

reconstruction. Although some probe pre-characterisations can be performed via 99

ptychography of a test pattern (or similar) in the transmission geometry, the final BP image 100

quality fully depends on the amount of uncertainties introduced by the lack of detailed 101

knowledge on the probe used during the BP experiment. In forward direction ptychographic 102

imaging, this limitation has long been recognised, and simultaneous retrieval of the probe and 103

object is now the universal measurement standard.35 The additional complexity of the Bragg 104

geometry has thus far prevented this refinement in BP.

105

Here, we used an advanced BP approach to investigate TEM-invisible defects in a tungsten- 106

rhenium alloy sample implanted with helium ions. To enable these measurements, we 107

develop a simultaneous probe refinement strategy that makes it possible to improve the 108

retrieved image sensitivity and map a much larger field of view than previously attainable.

109

Notably, this was achieved without increasing the amount of collected data (and therefore the 110

total acquisition time). This approach allows us to view the extended sample and directly 111

compare implanted and non-implanted regions, highlighting details of the crystalline 112

structure, which include lattice damage from helium irradiation, dislocations and sample 113

preparation damage. The results are discussed in the context of understanding irradiation 114

damage processes within tungsten and their potential impacts for its use in the design of 115

future fusion reactor components.

116 117 118

Results 119

A tungsten, 1% rhenium alloy was manufactured by arc melting, producing a polycrystalline 120

material with grain size of a few hundred microns as determined by scanning electron 121

microscopy (SEM). The addition of rhenium mimics neutron-irradiation-induced 122

transmutation alloying.36 The material was further implanted with helium ions that modified a 123

~3 µm thick surface layer (see Methods Sec. 1) and generate neutron-like collision cascade 124

damage.20 Note that, during fusion reactor operation, helium is generated by transmutation 125

and also diffuses from the plasma. Our use of helium ion implantation effectively mimics 126

(6)

these two effects. Finally, using focussed ion beam (FIB) milling, a cross-section sample 127

containing a grain boundary was extracted from the bulk and a region of 15×8 µm2 was 128

thinned to ~ 0.45 µm thickness. The resulting sample is shown in Fig. 1a.

129

BP microscopy was performed at a third generation synchrotron beamline, schematically 130

presented in Fig. 1b and described in detail elsewhere.37 An 8 keV coherent x-ray beam was 131

focused down to about 200 nm at the sample position with a set of Kirkpatrick-Baez (KB) 132

mirrors. The sample was placed on a three-axis translation stage, allowing for nano- 133

positioning and nano-scanning. The stage was mounted on the top of a goniometer cradle, 134

used to orientate the sample to the Bragg conditions and to map the 3D components of the 135

scattered intensity distribution, by angularly scanning the so-called rocking curve.38 The 136

sample was scanned across the beam propagation direction, in steps sufficiently small to 137

collect partially redundant information. For each probe-to-sample position and each angle 138

along the rocking curve, coherent diffraction patterns were recorded on a 2D pixelated 139

detector, placed at an exit angle of twice the Bragg angle, at a sufficiently large distance to 140

ensure far-field regime detection. For our measurements, the (220) specular reflection of the 141

right-hand grain of Fig. 1a was probed. A region of 2×2 µm2 (enclosed in the red rectangle in 142

Fig. 1a) near the grain boundary and the implantation layer was selected for the ptychography 143

scan, so that both implanted and non-implanted areas could be imaged. Full experimental 144

details are provided in Methods Sec. 2.

145

As an example representative of the collected data, a 2D coherent diffraction pattern from the 146

He-implanted region is shown in Fig. 1c. Its asymmetry and extent are strong signatures of 147

the presence of strain and rotation fields within the illuminated sample volume.38 A set of 148

four patterns are further presented, corresponding to the intensity distribution obtained when 149

integrating the 2D patterns along the rocking curve, for each of the extreme positions of the 150

raster scan (Fig. 1d). One observes clearly the motion of the main peak towards lower Bragg 151

angles, resulting from the strain in the implanted region (i.e., on the 20th and 400th positions).

152

This suggests that the implantation results in a crystalline lattice swelling, in line with 153

previous reports.20,21 Moreover, the 3D representation of the intensity distribution allows the 154

investigation of the fringe structure. They arise from the interference from the two surfaces of 155

the sample foil and present a period of about 14 µm-1. This provides an initial estimation of 156

the sample thickness of ~0.45 µm, in excellent agreement with the thickness expected from 157

SEM.

158

(7)

To go further in the sample structure analysis, the whole data set is inverted to retrieve the 3D 159

sample electron density. This requires a detailed modelling of the scattering process: Under 160

the kinematic scattering approximation, the Bragg diffraction intensity distribution Ij(q) at the 161

jth given probe position 𝐫! is given by:

162

𝐼! 𝐪 = !𝑃(𝐫−𝐫!)𝜌(𝐫)e!𝐐𝒉𝒌𝒍∙𝐮(𝐫)e!𝐪∙𝐫𝑑𝐫

!

!

,

(1) 163

where q represents the 3D coordinates reciprocal spaces, P is the probe function. For the 164

numerical implementation, the direct and reciprocal spaces are introduced. To preserve the 165

data information, the reciprocal frame follows the measurement scanning space (q1, q2, q3).

166

Using a recently developed formalism39, this sampling depicts a direct space frame (r1, r2, 167

r3). Fig. 2a shows a schematic representation of both space coordinate frames. Those are the 168

spaces into which the inversion is performed before the final object reconstruction is plotted 169

in the orthogonal (x, y, z) sample frame (Fig. 1b).

170

Accessing the phase of the sample scattering function from intensity measurements requires 171

one to overcome the loss of the scattered field phase. In forward ptychography, the 3D 172

problem can be decomposed in a series of 2D problems, each of them being solved in the 173

plane which contains the two scanning ptychographic directions.40 BP inversion is more 174

intricate because it aims at solving a problem that is intrinsically 3D. Indeed, Eq. 1 cannot be 175

simplified into a series of 2D ptychographic problems, by e.g., separating the probe and the 176

object in the integral. However, only two scanning ptychographic directions are available, 177

transverse to the probe propagation direction (note that along the beam direction, the weak 178

focussing power of x-ray lenses produces an extremely elongated depth of focus). Therefore, 179

while forward ptychography allows the simultaneous retrieval of both probe and sample 180

functions,35 3D BP requires strong a priori knowledge of the probe.

181

The structure and size of the probe also have a major impact on the 3D intensity acquisition 182

in BP. To better illustrate this issue, we present the probe used during the experiment (see the 183

cross-section and calculated profile along the beam propagation direction, in Fig. 2b and c).

184

This probe function was characterized prior to our measurement (see Methods Sec. 2). The 185

cross-section presents a rather structured distribution, which includes a central spot of about 186

400×200 nm2 (intensity FWHM) along the horizontal and vertical planes, respectively and 187

some beam tails and secondary maxima that extend to a much bigger area, of about 4×4 µm2. 188

These features have to be considered when defining the rocking curve angular steps. Indeed 189

(8)

the numerical analysis of the data set, which is based on the use of the fast Fourier transform, 190

connects reciprocal and direct space through Fourier conjugation relations. Precisely, the 191

sampling angle must be chosen such that the illuminated volume (green line) is fully 192

contained in a cuboid defined in the (r1, r2, r3) reconstruction space (orange line) as shown in 193

Fig. 2a. Depending on the probe structure and on the criterion used to define the probe size, 194

the angular sampling Δθ required along the rocking curve varies substantially, as illustrated 195

in Fig. 2d, for vertical-plane diffraction geometry. This plot presents the sampling condition 196

as a function of the considered (i.e., vertical) probe width W and the sample thickness T, for a 197

Bragg angle θB of 43.9° (see Methods Sec. 3 and Eq. 2 for further details). For our sample, 198

the required angular sampling varies from a few 0.01° when only the probe central lobe is 199

accounted for, to about 0.001° when the full probe extent is considered (interestingly, this 200

behaviour is rather marginally affected by the sample thickness). While 0.001° angular steps 201

are mechanically accessible at Bragg diffraction optimized beamline set-ups, those small 202

steps are still challenging and result in a detrimental linear increase of the total measurement 203

time. To mitigate issues caused by long data collection, a series of compromises are usually 204

made, such as choosing an angular step only accounting for the central lobe sampling 205

condition and/or fewer ptychographic positions and/or smaller angular range, which reduce 206

the field of view and/or degrade the image quality. Furthermore, the structure of the probe 207

secondary maxima that develop far from the central lobe, i.e., corresponding to high 208

frequencies, is more prone to optics instabilities, limiting the use of prior knowledge of the 209

probe for faithful reconstruction. All these underline the need for retrieving the full probe 210

distribution function and the sample scattering function simultaneously, a novel strategy we 211

have implemented and applied in this work, to produce high-fidelity maps of lattice strains 212

and rotations.

213

To succeed with the simultaneous retrieval of the 3D sample scattering function and 214

illumination function, additional information needs to be brought to the inversion problem.

215

Considering the generally small numerical aperture of x-ray optics, which results in a self- 216

similar probe along the beam direction, a straightforward constraint can be derived on the 217

probe invariance. This inversion strategy allows recovering not only the information in the 218

close vicinity of the central probe lobe, but also the information arising from much larger 219

distances where only the tails (the series of secondary maxima) of the probe illuminate the 220

sample. According to the sampling principles described above and illustrated in Fig. 2d, this 221

would imply scanning the rocking curve in extremely small angular steps (in the order of a 222

(9)

few 0.001°). We circumvented this issue by further up-sampling the rocking curve and 223

retrieving the information in the missing planes (see Methods Sec. 4). This operation 224

corresponds to virtually inserting angular sampling points in-between the measured ones, and 225

to retrieving their diffraction patterns based on the intensity information in the measured 226

ones. Using this strategy, the field of view was increased and the image quality was 227

improved, revealing unprecedented structural details, as shown hereafter. The whole 228

reconstruction strategy and inversion details are given in Methods Sec. 4.

229

The iso-surfaces of the retrieved 3D electron density and the phase φ220(r) associated with the 230

displacement field (φ220(r) = Q220 · u(r), with Q220 being the Bragg vector for the (220) 231

reflection), are shown in Fig. 3. For sake of clarity, only the relevant part of the retrieved 232

image is shown, using a mask applied on the all presented maps (see Supplementary 233

information S2 for the mask definition). The presence of a strong edge (seen from the top left 234

corner to the bottom middle) in the electron density map corresponds to the expected crystal 235

grain boundary. While the sample density is rather homogenous in most parts of the retrieved 236

field of view, some fluctuations are visible. The pipe-like ones (see black arrows in Fig. 3c 237

and 3d) are well known features, which can be understood when considering the phase map.

238

Here, phase vortices are co-localised with the pipes of missing intensity, a characteristic 239

feature of dislocations in noise-limited BCDI data.27,41–43 Similarly, additional density 240

oscillations appear in regions, where the phase wraps rapidly, e.g., at the top right corner.33 A 241

few density voids are observed in the vicinity of the grain boundary, corresponding to 242

cavities in the crystal. Finally, at the limit of the raster scan, where both the counting statistics 243

and overlap constraint are reduced the reconstruction quality is degraded. The probe function 244

retrieved simultaneously from the BP data set (cross-section profile shown in Fig. 3e) 245

exhibits the expected characteristic features (e.g., the central lobe with size of 450×190 nm2, 246

intensity FWHM), in excellent agreement with the profile obtained separately and previously 247

shown in Fig. 2b. The missing triangular region at the bottom left of the probe reconstruction 248

is due to the limited extent of the crystal set by the grain boundary. Note the extent of the 249

field of views, of about 0.6×6×6 µm3 (z×y×x) for the 3D sample and 5×4 µm2 (p1×p2) for the 250

probe cross section. If the intensity information were not limited by the grain boundary (non- 251

scattering crystal), we estimate the field of view along x could be as large as 7.8 µm. Finally, 252

the spatial resolution of this 3D reconstruction is estimated as 37×40×39 nm3 along z, y and x 253

respectively (see Supplementary information S3).

254

(10)

As a matter of comparison, the same data set was used to perform a reconstruction using the 255

formerly employed approach,29,44 i.e., no angular up-sampling and using a reconstructed 256

result of the probe based on transmission ptychography. The 3D reconstruction numerical 257

volume was directly designed according to the conjugation relation applied to the experiment 258

parameters, in particular the angular step. The final results are shown in Figs. 3f-h. The 259

reduction in the field of view is particularly evident along the x direction (maximum field of 260

view of 4.6 µm compared to 7.8 µm for the angular up-sampled approach), which is the 261

direction mostly impacted by the angular sampling along the rocking curve. Although the 262

main focal spot is fully contained in the probe function, the beam tails are clearly truncated.

263

This impacts the quality of the object image, where the grain boundary is still visible, but 264

artefacts, such as fluctuations in both electron density and phase, are more abundant. The 265

previously evident dislocations are barely visible. This direct comparison underlines the merit 266

of the new inversion strategy we have proposed.

267

The high-quality extended reconstruction was used to analyse the structural features present 268

in the implanted crystal. To this aim, we used 3D phase map to extract the lattice strain 269

projected along Q220 and the lattice rotations about the x and y axes, hereafter referred to as 270

εzz, ωx and ωy, respectively (see Methods Sec. 5). The top edge of the reconstructed volume 271

corresponds to the former sample surface during ion-implantation (note that, the strain 272

increase observed at the very top surface is an artefact due to a bit of aliasing resulting from 273

the probe extent). The 3D strain map (Fig. 4 (a) and (b)) clearly shows the He-implantation 274

induced strain of about 3×10-4, evident to a depth of approximately 2.8 µm. In the vicinity of 275

the two FIB-processed surfaces, large positive strains, corresponding to a lattice expansion, 276

can be seen. These strains extend to similar depths (about 150 nm for the top surface and 100 277

nm for the bottom surface) and have similar magnitude (~1.5×10-3 and ~1.1×10-3 for the top 278

and bottom surfaces, respectively). Several dislocations are also observed in both implanted 279

and non-implanted layers. Besides these, εzz presents homogeneous behaviour in the whole 280

implanted layer including in the vicinity of the grain boundary. Finally, we would also 281

encourage the reader to view the Supplementary Videos as these provide an interactive way 282

to visualise the multi-dimensional dataset. SV1 shows the object iso-surface, whilst SV2 and 283

SV3 each show a perpendicular plane propagating through the object with the εzz, ωx and ωy

284

components displayed.

285

To further illustrate the interest of our BP result, characterisation using more routinely 286

applied strain microscopy methods, namely high-resolution electron back scattering 287

(11)

diffraction (HR-EBSD) and x-ray micro-beam Laue diffraction measurements, were 288

performed for the same region of the sample. These results are shown in the Supplementary 289

information S4. Whilst these methods are considered a mainstay of analyse for strain and tilt 290

distributions in polycrystalline materials, their limitations are obvious in the present case.

291

HR-EBSD presents a lateral spatial resolution of ~100 nm45 and is only surface sensitive 292

(down to a depth of about 10-20 nm) due to the short mean free path of backscattered 293

electrons.46 On the other hand, micro-beam Laue diffraction is bulk sensitive, but the 294

obtained 2D maps represent the integration of the strain and tilts information along the whole 295

thickness. Its transverse spatial resolution, which integrates the size of the focused beam, the 296

scanning precision and the step size, builds up to about 0.5 µm in the present case. Some 297

indication of positive lattice strain along the upper edge of the sample (where the implanted 298

layer is known to be) was recorded by HR-EBSD and micro-beam Laue diffraction.

299

However, neither approaches were able to definitively capture the presence of the implanted 300

layer with a high level of clarity or any indication of the dislocations and the fine details of 301

the strains as was possible using BP.

302 303 304

Discussion 305

The success of our approach relies on the capability of our new method to retrieve the far- 306

field information in planes, whose intensity distributions are not experimentally measured. Of 307

course, this process has a limit and a dedicated analytical work is on the way to further derive 308

it mathematically. It was numerically determined for this experiment (Supplementary 309

information S5). Qualitatively, the limit can be understood by considering that the numerical 310

aperture of the focusing optic broadens the signal along the corresponding directions in the 311

reciprocal space, in particular along the rocking curve direction. Therefore, the intensity 312

information obtained in a given detector plane carries additional features arising from the 313

broadening of the signal along the direction perpendicular to the detector plane. Interestingly, 314

we observed that the reconstruction quality using 7 angular measurements (up-sampling ratio 315

of 18 and angular step of 0.03°) is comparable to the quality of the image produced by the 316

former reconstruction strategy, using 42 angles. From this comparison, considering the gains 317

with respect to the measurement time (×6) and to the achieved field of view (×1.4 318

considering the limits imposed by the crystalline grain size or ×1.7 considering the whole 319

accessible field of view), a total gain of about 8-10 is estimated. This significant 320

(12)

improvement brought by our approach should further benefit from the advent of 4th 321

generation synchrotron sources and their expected gain (of about 100) in coherence flux.

322

Regarding the crystalline properties imaged in the sample, we note that the 3D lattice strain 323

and rotation maps obtained using our new approach are not accessible in any other way and 324

reveal important new features: In the vicinity of the top and bottom sample surfaces, large 325

positive strains, corresponding to a lattice expansion, can be seen. The strains extend to 326

similar depths (~120 nm), have similar values (~1.1-1.5×10-3) and are present over the entire 327

sample surfaces. Their spatial distribution and location are consistent with the effects of 328

residual FIB-induced damage. FIB imaging and machining are known to be able to cause 329

dramatic material changes such as introduction of lattice defects,47 large lattice strains,48 330

amorphization,49 and formation of Ga intermetallic.50 From electron microscopy, FIB damage 331

is often indistinguishable from the intrinsic defects and damage features of interest in the 332

sample.51,52 In the present case, it is worth noting that FIB-induced damage remains after 333

employing preventative and mitigating measures, routinely employed in fabricating strain 334

microscopy samples for electron microscopy including sacrificial capping, glancing ion 335

incidence angle and the removal of surface material via low energy polishing (see Methods 336

Sec. 1 for fabrication details). We observe that even these residual effects lead to large lattice 337

distortions. However, the beauty of our approach is that sample regions affected by FIB 338

damage can be unambiguously identified and then excluded from further analysis. This 339

dramatically simplifies data interpretation and gives certainty that features observed in the 340

sample are not artefacts of the preparation. In the following only sample volumes more than 341

150 nm from sample surfaces, and thereby unaffected by FIB, are considered.

342

The εzz lattice strain map (Fig. 4a and 4b) shows a substantial strain with an average value of 343

3×10-4 in the helium-implanted layer. This is in good agreement with low-resolution micro- 344

beam Laue measurements of helium-implantation-induced lattice strain,21 estimating strains 345

of about 2.4×10-4. Whilst there is a clear change in lattice strain from the He-implanted layer 346

to the un-implanted material beneath, there is no discernible, abrupt change in lattice rotation 347

at the boundary between the implanted and un-implanted material (Fig. 4c-4f). This is a very 348

important evidence, since, in principle, a shear deformation of the lattice during ion 349

implantation could lead to rotations about the x and z axes. Such a shearing would indicate 350

the preferential formation of defects with specific orientation.53 In tungsten, helium 351

implantation is expected to predominantly produce Frenkel pairs that are prevented from 352

recombining as He occupies the vacancies.20,54,55 The lack of sharp change in lattice 353

(13)

orientation at the un-implanted/implanted interface shows that defects are randomly oriented, 354

leading to a purely volumetric strain.53,56 This important result should greatly simplify the 355

simulation of irradiation-induced strain in reactors,56 which will be one of the main 356

mechanisms driving in-service degradation of fusion reactor armour.

357

Across the layer, the homogeneous behaviour of εzz indicates a uniform distribution of 358

defects, at least within the resolution and sensitivity limit of the present measurements.

359

Previously, the formation of a zone denuded of large irradiation-induced dislocations loops 360

was reported close to grain boundaries (within 20 to 50 nm) in self-ion irradiated material.57 361

Our results suggest that for small defects and defect clusters, such as those created by helium 362

ion-irradiation,20 this is not the case, limiting the potential effectiveness of grain size 363

reduction as a means of reducing irradiation defect accumulation.

364

Finally, the quality of our crystalline image allows us to consider in more detail the 365

dislocations visible in the reconstructed sample. An iso-surface rendering of the electron 366

density, as well as the associated phase variation in the sample mid-plane are shown in Fig.

367

4h and Supplementary information S6 for dislocation #1 (implanted layer) and #2 (un- 368

implanted region), respectively. Dislocations in tungsten have predominantly b=1/2〈111〉 369

Burgers vector, although dislocations with b=〈100〉 have also been reported.43 For a given 370

Bragg reflection, only dislocations for which Qhkl·b is non-zero are visible.58 For the present 371

(220) reflection, this means that the observed dislocations may have the following Burgers 372

vectors: 1/2[111], 1/2[11-1], [100] or [010]. Dislocation #2 shows a central pipe of missing 373

intensity.26 The phase variation around the dislocation line shows a vortex with a total phase 374

increase of 4π. The phase variation agrees well with the simulated phase for a dislocation in 375

the thickness direction, as shown in Supplementary information S6. In contrast, dislocation 376

#1 shows a more complex and surprising structure: Here two separate pipes of reduced 377

intensity are clearly identified, suggesting that in fact this dislocation corresponds to two 378

parallel dislocations, each with an associated phase vortex of 2π. This is the signature 379

expected from two partial dislocations linked by a stacking fault on the (-110) plane. The 380

presence of dissociated dislocations in tungsten is surprising since bcc metals have 381

comparatively high stacking fault energy.59 However, recent ab-initio calculations show that 382

in fact, the addition of Re can substantially reduce the stacking fault energy in tungsten, a 383

scenario possibly at play here.59 384

385

(14)

386 387

Conclusions 388

Thanks to the development of an upgraded BP approach, we obtain quantitative 3D maps of 389

lattice strain and rotation in a He-implanted Tungsten crystal. These maps, of unmatched 390

quality and extent, reveal numerous, otherwise inaccessible, crystalline features. Beyond a 391

FIB-damaged layer, which can now be unambiguously discarded from the analysis, we 392

observed strains and lattice rotations caused by ‘invisible’ helium-implantation-induced 393

defects, identified to be of random orientation. Surprisingly, we do not find a defect-denuded 394

region in the vicinity of the crystal boundary. These results provide new insights, essential for 395

predicting the effects of ion-irradiation on crystals. Moreover, they pave the way for highly 396

detailed investigation of complex next-generation crystalline materials, e.g., refractory high- 397

entropy alloys for extreme environments.60,61 398

399 400

(15)

References 401

1. Baluc, N. et al. Status of R&D activities on materials for fusion power reactors. Nucl.

402

Fusion 47, S696 (2007).

403

2. Hartman, K., Bertoni, M., Serdy, J. & Buonassisi, T. Dislocation density reduction in 404

multicrystalline silicon solar cell material by high temperature annealing. Appl. Phys. Lett.

405

93, 122108 (2008).

406

3. Singer, A. et al. Nucleation of dislocations and their dynamics in layered oxide 407

cathode materials during battery charging. Nat. Energy 3, 641–647 (2018).

408

4. Clemens, H. & Mayer, S. Design, Processing, Microstructure, Properties, and 409

Applications of Advanced Intermetallic TiAl Alloys. Adv. Eng. Mater. 15, 191–215 (2013).

410

5. Ma, E., Shen, T. D. & Wu, X. L. Less is more. Nat. Mater. 5, 515–516 (2006).

411

6. Mahajan, S. Defects in semiconductors and their effects on devices. Acta Mater. 48, 412

137–149 (2000).

413

7. Gratias, D. Imperfections in crystalline solids. By Wei Cai and William D. Nix. MRS- 414

Cambridge Materials Fundamentals. Cambridge University Press, 2016. Paperback Pp. 532.

415

Price GBP 44.99. ISBN 1107123135. Acta Crystallogr. Sect. B Struct. Sci. Cryst. Eng. Mater.

416

73, 1018–1018 (2017).

417

8. Pollock, T. M. & Tin, S. Nickel-Based Superalloys for Advanced Turbine Engines:

418

Chemistry, Microstructure and Properties. J. Propuls. Power 22, 361–374 (2006).

419 9. Xu, R. et al. Three-dimensional coordinates of individual atoms in materials revealed 420

by electron tomography. Nat. Mater. 14, 1099–1103 (2015).

421

10. Tanaka, M., Higashida, K., Kaneko, K., Hata, S. & Mitsuhara, M. Crack tip dislocations 422

revealed by electron tomography in silicon single crystal. Scr. Mater. 59, 901–904 (2008).

423

11. Hÿtch, M. J., Putaux, J.-L. & Pénisson, J.-M. Measurement of the displacement field 424

of dislocations to 0.03 Å by electron microscopy. Nature 423, 270–273 (2003).

425

12. Anderson, P. M., Hirth, J. P. & Lothe, J. Theory of Dislocations. (Cambridge University 426

Press, 2017).

427

13. Yu, H., Yi, X. & Hofmann, F. 3D reconstruction of the spatial distribution of 428

dislocation loops using an automated stereo-imaging approach. Ultramicroscopy 195, 58–68 429

(2018).

430

14. Zhou, Z., Jenkins, M. L., Dudarev, S. L., Sutton, A. P. & Kirk, M. A. Simulations of 431

weak-beam diffraction contrast images of dislocation loops by the many-beam Howie–

432

Basinski equations. Philos. Mag. 86, 4851–4881 (2006).

433

15. Hurh, P. The Radiation Damage in Accelerator Target Environments (RaDIATE) 434

Collaboration R&D Program - Status and Future Activities. (2017) doi:10.18429/JACoW- 435

IPAC2017-WEOCB3.

436

16. Kim, B.-S., Kim, J. S., Park, Y. M., Choi, B.-Y. & Lee, J. Mg ion implantation on SLA- 437

treated titanium surface and its effects on the behavior of mesenchymal stem cell. Mater.

438

Sci. Eng. C Mater. Biol. Appl. 33, 1554–1560 (2013).

439

17. Yi, X. et al. Direct observation of size scaling and elastic interaction between nano- 440

scale defects in collision cascades. EPL Europhys. Lett. 110, 36001 (2015).

441

18. Das, S., Yu, H., Tarleton, E. & Hofmann, F. Hardening and Strain Localisation in 442

Helium-Ion-Implanted Tungsten. Sci. Rep. 9, 18354 (2019).

443

19. Hofmann, F. et al. Non-Contact Measurement of Thermal Diffusivity in Ion-Implanted 444

Nuclear Materials. Sci. Rep. 5, 16042 (2015).

445

20. Hofmann, F. et al. Lattice swelling and modulus change in a helium-implanted 446

tungsten alloy: X-ray micro-diffraction, surface acoustic wave measurements, and multiscale 447

(16)

modelling. Acta Mater. 89, 352–363 (2015).

448

21. Das, S., Liu, W., Xu, R. & Hofmann, F. Helium-implantation-induced lattice strains and 449

defects in tungsten probed by X-ray micro-diffraction. Mater. Des. 160, 1226–1237 (2018).

450

22. Derlet, P. M. & Dudarev, S. L. Microscopic structure of a heavily irradiated material.

451

Phys. Rev. Mater. 4, 023605 (2020).

452

23. Phillips, N. W. et al. Nanoscale Lattice Strains in Self-ion-implanted Tungsten. Acta 453

Mater. (2020) doi:10.1016/j.actamat.2020.05.033.

454

24. Robinson, I. K., Vartanyants, I. A., Williams, G. J., Pfeifer, M. A. & Pitney, J. A.

455

Reconstruction of the Shapes of Gold Nanocrystals Using Coherent X-Ray Diffraction. Phys.

456

Rev. Lett. 87, 195505 (2001).

457 25. Pfeifer, M. A., Williams, G. J., Vartanyants, I. A., Harder, R. & Robinson, I. K. Three- 458

dimensional mapping of a deformation field inside a nanocrystal. Nature 442, 63–66 (2006).

459

26. Ulvestad, A. et al. Topological defect dynamics in operando battery nanoparticles.

460

Science 348, 1344–1347 (2015).

461

27. Clark, J. N. et al. Three-dimensional imaging of dislocation propagation during crystal 462

growth and dissolution. Nat. Mater. 14, 780–784 (2015).

463

28. Pateras, A. I. et al. Nondestructive three-dimensional imaging of crystal strain and 464

rotations in an extended bonded semiconductor heterostructure. Phys. Rev. B 92, 205305 465

(2015).

466 29. Chamard, V. et al. Strain in a silicon-on-insulator nanostructure revealed by 3D x-ray 467

Bragg ptychography. Sci. Rep. 5, 9827 (2015).

468

30. Kim, C. et al. Three-Dimensional Imaging of Phase Ordering in an Fe-Al Alloy by Bragg 469

Ptychography. Phys. Rev. Lett. 121, 256101 (2018).

470

31. Hill, M. O. et al. Measuring Three-Dimensional Strain and Structural Defects in a 471

Single InGaAs Nanowire Using Coherent X-ray Multiangle Bragg Projection Ptychography.

472

Nano Lett. 18, 811–819 (2018).

473

32. Mastropietro, F. et al. Revealing crystalline domains in a mollusc shell “single- 474

crystalline” prism. Nat. Mater. 2nd Rev.

475

33. Godard, P., Allain, M., Chamard, V. & Rodenburg, J. Noise models for low counting 476

rate coherent diffraction imaging. Opt. Express 20, 25914 (2012).

477

34. Takagi, S. A Dynamical Theory of Diffraction for a Distorted Crystal. J. Phys. Soc. Jpn.

478

26, 1239 (1969).

479

35. Thibault, P. et al. High-Resolution Scanning X-ray Diffraction Microscopy. Science 480

321, 379–382 (2008).

481

36. Gilbert, M. R. & Sublet, J.-C. Neutron-induced transmutation effects in W and W- 482

alloys in a fusion environment. Nucl. Fusion 51, 043005 (2011).

483

37. Leake, S. J. et al. The Nanodiffraction beamline ID01/ESRF: a microscope for imaging 484

strain and structure. J. Synchrotron Radiat. 26, 571–584 (2019).

485

38. Williams, G. J., Pfeifer, M. A., Vartanyants, I. A. & Robinson, I. K. Three-Dimensional 486

Imaging of Microstructure in Au Nanocrystals. Phys. Rev. Lett. 90, 175501 (2003).

487

39. Li, P. et al. General approaches for shear-correcting coordinate transformations in 488

Bragg coherent diffraction imaging. Part II. J. Appl. Crystallogr. 53, (2020).

489

40. Dierolf, M. et al. Ptychographic X-ray computed tomography at the nanoscale.

490

Nature 467, 436–439 (2010).

491

41. Favre-Nicolin, V. et al. Analysis of strain and stacking faults in single nanowires using 492

Bragg coherent diffraction imaging. New J. Phys. 12, 035013 (2010).

493

42. Takahashi, Y. et al. Bragg x-ray ptychography of a silicon crystal: Visualization of the 494

(17)

dislocation strain field and the production of a vortex beam. Phys. Rev. B 87, 121201 (2013).

495

43. Hofmann, F. et al. Nanoscale imaging of the full strain tensor of specific dislocations 496

extracted from a bulk sample. Phys. Rev. Mater. 4, 013801 (2020).

497

44. Mastropietro, F. et al. Revealing crystalline domains in a mollusc shell single- 498

crystalline prism. Nat. Mater. 16, 946–952 (2017).

499

45. Chen, D., Kuo, J.-C. & Wu, W.-T. Effect of microscopic parameters on EBSD spatial 500

resolution. Ultramicroscopy 111, 1488–1494 (2011).

501

46. Drouin, D. et al. CASINO V2.42: a fast and easy-to-use modeling tool for scanning 502

electron microscopy and microanalysis users. Scanning 29, 92–101 (2007).

503

47. Kiener, D., Motz, C., Rester, M., Jenko, M. & Dehm, G. FIB damage of Cu and possible 504 consequences for miniaturized mechanical tests. Mater. Sci. Eng. A 459, 262–272 (2007).

505

48. Hofmann, F. et al. 3D lattice distortions and defect structures in ion-implanted nano- 506

crystals. Sci. Rep. 7, 45993 (2017).

507

49. Giannuzzi, L. A., Geurts, R. & Ringnalda, J. 2 keV Ga+ FIB Milling for Reducing 508

Amorphous Damage in Silicon. Microsc. Microanal. 11, 828–829 (2005).

509

50. Yu, J., Liu, J., Zhang, J. & Wu, J. TEM investigation of FIB induced damages in 510

preparation of metal material TEM specimens by FIB. Mater. Lett. 60, 206–209 (2006).

511

51. Sato, T. et al. Low damage lamella preparation of metallic materials by FIB 512

processing with low acceleration voltage and a low incident angle Ar ion milling finish. J.

513 Microsc. n/a,.

514

52. Baram, M. & Kaplan, W. D. Quantitative HRTEM analysis of FIB prepared specimens.

515

J. Microsc. 232, 395–405 (2008).

516

53. Ma, P.-W. & Dudarev, S. L. Universality of point defect structure in body-centered 517

cubic metals. Phys. Rev. Mater. 3, 013605 (2019).

518

54. Sandoval, L., Perez, D., Uberuaga, B. P. & Voter, A. F. Competing Kinetics and He 519

Bubble Morphology in W. Phys. Rev. Lett. 114, 105502 (2015).

520

55. Boisse, J., Backer, A. D., Domain, C. & Becquart, C. S. Modeling of the self trapping of 521

helium and the trap mutation in tungsten using DFT and empirical potentials based on DFT.

522

J. Mater. Res. 29, 2374–2386 (2014).

523

56. Dudarev, S. L., Mason, D. R., Tarleton, E., Ma, P.-W. & Sand, A. E. A multi-scale model 524

for stresses, strains and swelling of reactor components under irradiation. Nucl. Fusion 58, 525

126002 (2018).

526

57. Zhang, Z., Yabuuchi, K. & Kimura, A. Defect distribution in ion-irradiated pure 527

tungsten at different temperatures. J. Nucl. Mater. 480, 207–215 (2016).

528

58. Williams, D. B. & Carter, C. B. Transmission Electron Microscopy: A Textbook for 529

Materials Science. (Springer US, 2009). doi:10.1007/978-0-387-76501-3.

530

59. Qian, J., Wu, C. Y., Fan, J. L. & Gong, H. R. Effect of alloying elements on stacking fault 531

energy and ductility of tungsten. J. Alloys Compd. 737, 372–376 (2018).

532

60. Senkov, O. N., Wilks, G. B., Miracle, D. B., Chuang, C. P. & Liaw, P. K. Refractory high- 533

entropy alloys. Intermetallics 18, 1758–1765 (2010).

534

61. El-Atwani, O. et al. Outstanding radiation resistance of tungsten-based high-entropy 535

alloys. Sci. Adv. 5, eaav2002 (2019).

536

62. Maiden, A. M. & Rodenburg, J. M. An improved ptychographical phase retrieval 537

algorithm for diffractive imaging. Ultramicroscopy 109, 1256–1262 (2009).

538

63. Berenguer, F. et al. X-ray lensless microscopy from undersampled diffraction 539

intensities. Phys. Rev. B 88, 144101 (2013).

540

64. Goodman, J. W. Introduction To Fourier Optics. (W.H.Freeman & Co Ltd, 2005).

541

(18)

65. Maiden, A., Johnson, D. & Li, P. Further improvements to the ptychographical 542

iterative engine. Optica 4, 736–745 (2017).

543

66. Chapman, H. N. et al. High-resolution ab initio three-dimensional x-ray diffraction 544

microscopy. JOSA A 23, 1179–1200 (2006).

545

67. Maddali, S. et al. General approaches for shear-correcting coordinate 546

transformations in Bragg coherent diffraction imaging. Part I. J. Appl. Crystallogr. n/a,.

547

68. Marchesini, S. et al. X-ray image reconstruction from a diffraction pattern alone.

548

Phys. Rev. B 68, 140101 (2003).

549

69. Guizar-Sicairos, M. et al. Phase tomography from x-ray coherent diffractive imaging 550

projections. Opt. Express 19, 21345–21357 (2011).

551 552 553

(19)

Methods 554

1 - Sample preparation 555

The sample was produced from a bulk polycrystalline tungsten-1wt% rhenium alloy, 556

manufactured by arc melting from high purity elemental powders. The sample surface was 557

mechanically ground and further polished using diamond paste. A final chemo-mechanical 558

polishing step (with 0.1 µm-colloidal silica suspension) was used to produce a high-quality 559

surface finish. The grain size in the polycrystal was ~500 µm. The sample was implanted 560

with helium ions to produce a 2.8 µm thick implanted layer. To achieve a relatively uniform 561

damage level (0.02 ± 0.003 dpa between 0 µm and 2.8 µm depth) and injected helium 562

concentration (310 ± 30 appm between 0 µm and 2.8 µm depth), implantation was performed 563

using a number of different ion energies up to 2 MeV. The details of the implantation are 564

provided elsewhere.21 For BP measurements, a small specimen (~25×14×0.5 µm3) was 565

extracted from the bulk using a focused ion beam (FIB) lift-out process. This preparation 566

method is adapted from that developed for the preparation of isolated micro-crystals for 567

BCDI.43 In order to minimise the effects of lattice damage induced during the lift-out process 568

a combination of protective capping, glancing angle milling, and low energy polishing was 569

used. Specifically, the surface of the bulk sample was first covered with carbonaceous 570

platinum, to further protect the tungsten from normal incidence angle milling ions, using an 571

electron-beam, which does not introduce any damage in the implanted material. This 572

produces a thin layer of about 0.1 – 0.2 µm, which appears as a dark line in the SEM picture 573

(Fig. 2a). On top of this layer, 2 µm of carbonaceous platinum are further deposited using the 574

gallium-ion-beam, which achieves a balance of sputtering and deposition. This is faster than 575

the electron-beam but more intrusive, justifying the prior deposition of the electron-beam 576

assisted deposition layer. The gallium-deposited platinum appears with a different contrast at 577

the top of the sample. Trenches on either side of the lift-out sample were made using FIB, 578

this was then undercut and attached to a micro-manipulator. The sample was finally attached 579

to a copper TEM-lift-out holder as 90 degrees to its original orientation i.e., the walls of the 580

trench became the top and bottom surfaces of the sample for all measurements. As the sample 581

was still rather thick at this point (1 – 2 µm), FIB was again used to further thin down the 582

sample. This was performed at 30 keV, with a current in the order of 240 pA or slightly 583

lower. Finally, the obtained lamella was polished with a 2 keV gallium ion beam to remove 584

most of the damage from the previous FIB steps. It results in a ~0.35 µm thinned region 585

(20)

around the implantation layer region, in the vicinity of the grain boundary. The implanted 586

layer starts from the dark line downwards (Fig. 2a). The (220) lattice planes of the interested 587

grain (further away from the copper support end in Fig. 2a) are parallel to the two lift-out 588

surfaces.

589 590

2 - Experimental details – Bragg ptychography 591

The x-ray beam of the ID01 beamline (ESRF) was filtered with a Si-111 double crystal 592

monochromator to provide a monochromatic beam with energy of 8 keV (or wavelength λ of 593

1.54 Å) with a bandwidth of about 1×10-4. Kirkpatrick-Baez mirrors were used to focus the 594

beam onto the sample, horizontally (H) and vertically (V). A pair of slits placed upstream the 595

mirrors were set to a width of 60×200 µm2 (H ×V), selecting a spatially coherent beam for 596

the experiment. From a dedicated (forward ptychography) beam characterization, the beam 597

profile was extracted and the central spot size (intensity FWHM) was measured as 400×200 598

nm2 (H×V). The resulting numerical aperture (NA) was estimated to be 1.9×10-4 and 3.7×10-4 599

(H×V) using the extent of the calculated over-focused beam and the depth of focus to 600

1950×540 µm2 (H×V). More details are given in the following Methods sections.

601

For the BP measurements, the specular (220) Bragg reflection of the grain away from the 602

copper support was chosen. It corresponds to a Bragg angle θB of 43.9˚. The intensity 603

patterns were measured with an area detector (Maxipix, 516×516 pixels, pitch size of 604

55×55𝜇m2) located at 1.4 m away from the sample. A total of 42 angular positions, with an 605

angular step size of 0.005˚, were taken along the rocking curve. At each of these angles, a set 606

of 20×20 positions, with a step size of 100 nm, was used for the raster scan. It corresponds to 607

an area of 2×2 µm2. The exposure time per frame was set to 0.2 seconds.

608

Immediately after BP measurements, the experimental geometry was changed to the forward 609

geometry to perform a conventional transmission ptychographic scan with a Siemens star 610

pattern, in order to retrieve the illumination profile.35 An Archimedean spiral scan of 513 611

positions, with a step size of 100 nm, was performed and for each position the detector was 612

exposed for 1s to measure the diffraction pattern. A sub-region of 334×334 pixels, centered 613

around the forward beam, was used to retrieve the probe profile at the focal plane using 614

standard approaches based on the extended ptychographical iterative engine algorithm 615

(ePIE).62 The initial guess for the probe was a simulated beam profile based on the 616

(21)

experimental setup, i.e. calculated via the Fourier transform of a rectangle that is the 617

numerical aperture (or NA) of the beam. The initial guess for the object was simply a uniform 618

matrix of 1s. The gap between the different sensor modules was masked and left free. The 619

reconstruction was run for 100 iterations, where the difference between the measured 620

diffraction patterns and the calculated ones is very small and further iterations provide very 621

little further convergence.

622 623

3 – Angular sampling along the rocking curve for 3D Fourier transform based Bragg 624

coherent diffraction imaging 625

Due to the use of a 3D discrete Fourier transform to describe the propagation between real 626

and reciprocal spaces, the angular sampling pitch Δθ and the window size R3 of the 627

illumination function along r3 are linked via ∆θ = 𝜆 ∕ 2R3sin(θB)cos(θB), according to the 628

sampling relations,39,63 where, considering the geometry depicted in Fig. 1a, 629

R3 = 𝚃cos(2𝜃B)/sin(𝜃B) + W/tan(𝜃B). Here 𝚃 is the sample thickness and W is the beam width 630

along p1. This leads to the following relation between the beam width W and the sampling 631

pitch Δθ: 632

W = 𝜆 ∕ 2∆θ[cos(θB)]2 - 𝚃cos(2𝜃B) ∕ cos(𝜃B) (2) 633

For the used angular step size of 0.005˚ with a sample thickness of 500 nm, W is about 1.7 634

µm. It is big enough to encompass the focal spot, but not the secondary maxima in the probe 635

tails. Those strong beam tails, which extend beyond the window size, result in an aliasing 636

artefact according to the Nyquist sampling theorem.64 Therefore, with this angular step size, 637

the quality of the reconstruction is limited.

638 639

4 – Reconstruction strategy 640

The overall reconstruction strategy stems from previous works in 3D Bragg 641

ptychography28,33,63 with some pivotal adaptations that are required to obtain the high quality, 642

stable reconstruction shown in Fig. 3. The reconstruction is performed via the following cost- 643

function to minimized over all the probe positions 644

(3)

645

L(⇢, P, b) = XJ

j=1

Lj(⇢, P, b)

(22)

where 𝜌, 𝑃 and 𝑏, respectively, the 3D electronic density (i.e., the sample), the 3D probe 646

function and the incoherent intensity background. Individual cost-functions are defined 647

by28,33 648

(4)

649

where 𝐼! 𝐪 and ℎ!(𝐪) = |Ψ!(𝐪)|!+𝑏(𝐪) are the measured and predicted photon counts 650

for the j-th probe position, respectively and Ψ! 𝐪 is the Fourier transform of the exit field.

651

The detector mask ω(q) is designed to discard hot or dead camera pixels but also to provide 652

the expected angular up-sampling factor (described in details below). The second, quadratic 653

term is a thickness-support regularization applied to the retrieved sample preventing the 654

density of the sample to build-up in 𝑆, the set of points outside the sample support. The 655

regularization parameters 𝜇! ≥ 0 adjust how the final solution should comply with this 656

support constraint; in the Bragg geometry, we note that this regularization is pivotal in 657

dealing with the intrinsic lack of diversity of the beam along the propagation direction.63 The 658

reconstruction strategy aims at retrieving simultaneously 𝜌, 𝑃 and 𝑏 from the set of 659

measurements 𝐼!(𝐪)

!!!

! via the minimization of Eq. 3.

660

1 - Presentation of the reconstruction algorithm 661

The reconstruction algorithm was derived from the ePIE62 approach, with specific 662

adaptations that allow the probe function to be accurately retrieved in the Bragg geometry.

663

More specifically, a probe and object updates associated with the current (updated) position 664

𝑗 ∈ 1⋯𝐽 are given by 665

, (5a)

666

667 (5b)

where 𝜕!;! and 𝜕!;! are the gradients of with respect to 𝜌 and 𝑃, respectively 668

(6a)

669

. (6b)

670

The projection and back-projection operators and , respectively, are acting altogether 671

along the direction of the incoming beam 𝐤!; the pair of operators is basically enforcing, in 672

the probe update (Eq. 5b), the invariance of the probe along the direction ki (described in 673

Lj(⇢, P, bj) := w(q)⇥⇣

Ij1/2(q) h1/2j (q)⌘ 2

+ µj

X

r2S

|⇢(r)|2

⇢(r) ⇢(r) + 1(r)⇥@⇢;j(r)

P(r) BkiRki P(r) + P1(r)⇥@P;j(r) L(⇢, P, b) =

XJ

j=1

Lj(⇢, P, b)

<latexit sha1_base64="zrUWBEs8P7B5QdLxBt90Q3QBYaE=">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</latexit>

@⇢;j(r) = P(r rj)⇥ 0

j(r) j(r)⇤

µj[1 S(r)]⇢(r)

@P;j(r) = ⇢(r)⇥ 0

j(r) j(r)⇤

@P;j(r) = B@kPiR;j(r) =ki(r) [ BkiR0(r)ki(r) [ (r)]0(r) (r)]

Références

Documents relatifs

+BZBOUI $)&#34;//&#34;(*3* ] ĉÍTF EF EPDUPSBU ] 6OJWFSTJUÊ EF -JNPHFT ] %ÊDFNCSF.. GPS UIF TBNQMF JSSBEJBUFE BU , JT DIBSBDUFSJTUJD PG UIF ėSTU TUFQ PG UIF EBNBHF CVJMEVQ

La dernière partie présente une hybridation entre les métaheuristiques de base (AG-RT) et les métaheuristiques avec une méthode analytique quasi-Newtonienne (RT-QN, AG-QN). Mots

Keywords: hexagonal ice, viscoplastic deformation, crystalline anisotropy, dislocation density, geometrically necessary dislocations, curvature tensor, lattice distortions,

In this article we present the framework for two computa- tionally efficient Fourier transformations that simultaneously offer a natural frame both for the sample (orthogonal

This two-part article series provides a generalized description of the scattering geometry of Bragg coherent diffraction imaging (BCDI) experiments, the shear distortion

The degeneracy of the ordering results in antiphase domain boundaries that, in addition to the general lattice strain, provide phase shifts of the diffracted beam depending on

The spatial distribution of the lattice distortion can be visualized through a series of X-ray diffraction topo- graphs recorded along the rocking curves, which allows real- space

5.2 , which presents the obtained results in the vicinity of the intense InP peak, we can see that the proposed diffraction patterns are well reproduced with respect to the