In situ controlled modification of the helium density ... - Laurent Pizzagalli

Mar 25, 2014 - possibility to manipulate molecular-sized objects using the focused electron beam of a STEM, as it could be the case with the electron tweezers ...
3MB taille 4 téléchargements 329 vues
In situ controlled modification of the helium density in single helium-filled nanobubbles M.-L. David, K. Alix, F. Pailloux, V. Mauchamp, M. Couillard, G. A. Botton, and L. Pizzagalli Citation: Journal of Applied Physics 115, 123508 (2014); doi: 10.1063/1.4869213 View online: http://dx.doi.org/10.1063/1.4869213 View Table of Contents: http://scitation.aip.org/content/aip/journal/jap/115/12?ver=pdfcov Published by the AIP Publishing

[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: 194.167.47.1 On: Tue, 25 Mar 2014 16:25:28

JOURNAL OF APPLIED PHYSICS 115, 123508 (2014)

In situ controlled modification of the helium density in single helium-filled nanobubbles M.-L. David,1,2,a) K. Alix,1 F. Pailloux,1,2 V. Mauchamp,1 M. Couillard,2,3,b) G. A. Botton,2,3 and L. Pizzagalli1

1 Institut Pprime, UPR 3346 CNRS-Universit e de Poitiers, SP2MI, 86962 Futuroscope-Chasseneuil cedex, France 2 Canadian Centre for Electron Microscopy, Mc Master University, 1280 Main Street West, Hamilton, Ontario L8S 4M1, Canada 3 Department of Materials Science and Engineering, Mc Master University, 1280 Main Street West, Hamilton, Ontario L8S 4M1, Canada

(Received 6 January 2014; accepted 10 March 2014; published online 25 March 2014) We demonstrate that the helium density and corresponding pressure can be modified in single nano-scale bubbles embedded in semiconductors by using the electron beam of a scanning transmission electron microscope as a multifunctional probe: the measurement probe for imaging and chemical analysis and the irradiation source to modify concomitantly the pressure in a controllable way by fine tuning of the electron beam parameters. The control of the detrapping rate is achieved by varying the experimental conditions. The underlying physical mechanisms are discussed; our experimental observations suggest that the helium detrapping from bubbles could be interpreted in terms of direct ballistic collisions, leading to the ejection of the helium atoms from C 2014 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4869213] the bubble. V

I. INTRODUCTION

Transmission electron microscopy (TEM) is a powerful tool to simultaneously induce electron irradiation modifications of materials and to observe in situ the resulting damage evolution. Scanning Transmission Electron Microscopes (STEM) with sub-angstrom probe and fitted with spectrometers expand further the application of electron microscopy as it allows to combine high spatial resolution imaging and elemental analysis, while still providing the electron beam to modify concomitantly the matter at the atomic level. STEM/TEM has been used for instance to study irradiationinduced extended defects1 or void formation2 in materials, curling and closure of graphitic networks,3 melting of particles,4 radiolysis to knock-on damage transition in zeolites,5 or nanoparticle nucleation and growth.6 One of the exciting applications is to finely control the damage formation in order to sculpt matter at the atomic level and to fabricate new classes of nanostructures embedded in materials with potential chemical and biological applications. The drilling of solid state nanopores in materials is an outstanding example of this.7–9 Furthermore, the possibility to manipulate molecular-sized objects using the focused electron beam of a STEM, as it could be the case with the electron tweezers, is also raising great interest.10,11 Rare-gas nanoscale bubbles can be formed in a wide variety of materials following the agglomeration of vacancies and rare gas introduced in high concentration. They are an important class of defects in solid state materials that are both of technological and fundamental relevance as they can

a)

Electronic mail: [email protected] Present address: National Research Council Canada, 1200 Montreal Road, Ottawa, Ontario K1A 0R6, Canada.

b)

0021-8979/2014/115(12)/123508/9/$30.00

modify the electrical, optical, and mechanical properties of materials. For instance, these bubbles give rise to remarkable features in the optical properties of metals.12 In nanofluidics, they are an alternative route to create nanoholes.13 They are also of significant interest in the microelectronic and photovoltaic fields as they are used as gettering centers for metallic impurities in silicon,14 for annihilation of dislocations in GaN,15 or for the fabrication of ultrathin buried oxide layers in silicon.16 On the other hand, the formation of helium bubbles in metals, alloys, and in UO2 is a major issue in the field of materials used for nuclear energy production. Helium bubbles are indeed responsible for the embrittlement of the structural materials17 used in nuclear reactors. This has triggered a very recent interest on interfaces in composite materials that can act as sinks for radiation-induced defects.18,19 Moreover, rare gas bubbles cause the degradation of the mechanical and thermal properties of nuclear fuel (UO2).20 While the previously cited works make use of electron microscopy as a characterization tool of gas bubbles, we focus here on both the characterization and the use of electron microscopy as a tool to modify such nanostructures. This paper thus focuses on helium nanoscale bubbles embedded in elemental semiconductors silicon and germanium. In these materials, helium bubbles can be created by high fluence helium implantation. One interesting aspect of this technique is that the mean characteristics of a bubble distribution (bubble density, size, and shape) can be tailored at the microscopic scale by varying the implantation (current, energy, temperature, fluence) and annealing conditions.21,22 However, using this method, it is not possible to control the physical properties of a single bubble. This would be, for instance, of primary importance for the development of new bottom-up strategies for the production of spatially ordered nano-objects.23 Indeed, the idea to use the stress induced by

115, 123508-1

C 2014 AIP Publishing LLC V

[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: 194.167.47.1 On: Tue, 25 Mar 2014 16:25:28

123508-2

David et al.

a single helium-filled defect to fabricate self-assembly of nano-objects has emerged recently.24,25 In this stress-field engineering concept, the stress is due to the pressure in the defect, which cannot be controlled by the implantation and annealing conditions. We show here that the helium density and corresponding pressure can be modified in single nano-scale bubbles by using the electron beam of a STEM as a multifunctional probe: a measurement probe for imaging and chemical analysis26–28 and an irradiation source to modify concomitantly the pressure in a controllable way by fine tuning of the electron beam parameters. II. EXPERIMENTAL DETAILS A. Experiments

We prepare systems consisting of dispersed helium nanoscale bubbles embedded in elementary semiconductors. The bubbles were created by helium implantation in p-type (001) silicon and germanium samples by high fluence implantation followed by a thermal annealing stage. The implantation and annealing conditions were chosen to create a layer of bubbles of different diameters. For both materials, the current was set to 10 lA and the implantations were carried out at room temperature (RT). For silicon, the fluence was of 7  1016 cm2 and the energy of 50 keV. For germanium, the fluence was of 5  1016 cm2 and the energy of 30 keV. The samples were then submitted to an annealing at 700  C for silicon and 500  C for germanium during 30 min, under vacuum in a tubular furnace. In this paper, the word “cavity” refers to helium-filled as well as to empty cavity, whereas the word “bubble” is used exclusively for helium-filled cavity. Cross-sectional TEM samples were prepared by mechanical polishing down to 10 lm and ion milling in a GATANPIPS apparatus at low energy (2.5 keV Ar) and low incidence (68 ). To minimize irradiation damage, a final step was performed at 64 for 2 min. In situ electron irradiation and spatially resolved Electron Energy Loss Spectroscopy (STEM-EELS) experiments were performed using a FEI Titan cubed 80–300 microscope fitted with a monochromator and image and probe Cs correctors. In order to monitor the helium detrapping from bubbles, the high voltage was set to either 80 or 200 keV and the current in the probe ranged between 40 and 70 pA. Spectrum images with a high spatial sampling (typically 60  60 pixel2) and with 2048 channels were acquired using the following conditions: acquisition time of 100 ms per spectrum, energy dispersion of 0.02 eV/channel. All these experiments were conducted at RT. B. Methods

The presence of cavities (filled or not) is evidenced using either the high-angle annular dark-field (HAADF) contrast or by filtering on energy ranges corresponding to the cavity plasmon or silicon/germanium bulk plasmon. This allows to identify the morphology of the bubbles and to determine their diameter by averaging the diameters measured in the three images, leading to an accuracy between 1 and 2 nm. Moreover, for non-spherical bubbles, the diameter

J. Appl. Phys. 115, 123508 (2014)

is given as an average of the width and the length of the cavity, which may differ from 1 to 1.5 nm for the largest bubbles (about 20 nm in diameter). The helium density map is realized by filtering on the helium K-edge.27 Helium density, nHe, is estimated using the blue shift of the helium K-edge, DE, as compared to its position for the free atom28–30 nHe ¼

1 DE: C

(1)

This shift is attributed to the short range Pauli repulsion between electrons of neighboring atoms. The position of the helium K-edge is determined at the maximum of the extracted signal and the position for the free atom is taken as its smallest value determined in our experiments (once cavities have been emptied by the electron beam—21.4 eV). The helium K-edge was extracted using the following method. The raw data were first realigned on the elastic peak and subjected to multivariate statistic analysis (MSA)31 to remove the statistical noise. As the shift of the helium K-edge introduces an odd component in the signal, great care was taken to reconstruct the signal with a sufficiently large number of components. This leads to a reconstructed signal which is still a bit noisy. The multiple scattering was deconvoluted using a Fourier-log method32 and the silicon (germanium) plasmon was fitted using a sum of Lorentzian and Gaussian functions. A typical extracted spectrum is shown in Sec. III A together with the raw signal. It is worth noting that in previous studies by STEM-EELS,26,27,33 another approach based on the measurement of the He 1s ! 2p transition integration intensity was used. The method used here is more direct providing the value of the constant C is known with a sufficient accuracy. In the literature, values between 0.015 eV (Ref. 28) and 0.044 eV (Ref. 27) have been reported. These differences may be ascribed to several factors such as estimation of the thickness of bubble crossed by the electron beam (largest source of uncertainties), influence of the embedding material, surface effects not taken into account. In this paper, we use the C value we have recently determined for similar bubbles and conditions under study in silicon.28 Moreover, the influence of the material in terms of dielectric screening dominated by bulk plasmon is supposed to be equivalent in silicon and germanium since both plasmons have similar energies and full width at half maximum. Further, bubbles under study have similar morphology in both materials. The C value is thus estimated to be same for both systems. Finally, following the method initiated in Ref. 26 and used in Refs. 27 and 33, we use Eq. (2) as the equation of states to estimate the pressure.34 PLD ðnHe Þ ¼ anHe ½ð1  nHe Þ  ð1 þ nHe  2 2 n2He Þ þ nHe bð1  nHe Þ2 þ cð3  2nHe Þ 3 n2He  50ð1  nHe Þ 2 n2He ; PHD ðnHe Þ ¼

3b ð602=nHe =aÞ

2=3

(2)

 ð1  ð602=nHe =aÞ1=3 Þ

 exp½1:5ðc  1Þð1  ð602=nHe =aÞ1=3 Þ:

(3)

[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: 194.167.47.1 On: Tue, 25 Mar 2014 16:25:28

123508-3

David et al.

J. Appl. Phys. 115, 123508 (2014)

˚ 3), T is the In Eq. (2), nHe is expressed in (A 0:25  exp½0:145T 0:25 , temperature (T ¼ 300 K),  ¼ 56=T 1=3 a ¼ 0.0138062 T, b ¼ 170/T  1750/T, and c ¼ 0.1225  T0.555. For the specific cases where a very high helium density (>140 He/nm3) was determined, the pressure was derived using a Vinet fit35 on the 300 K data of Ref. 36, which is more accurate in this range, Eq. (3). In this equation, nHe is expressed in (nm3), a ¼ 11.78614, b ¼ 0.46756, and c ¼ 7.02903. III. RESULTS AND DISCUSSION A. The electron beam as a measurement probe

FIG. 1. STEM-HAADF images of bubbles created by (a) 7  1016 cm2, 50 keV, room temperature, helium implantation in (001) n-silicon and (b) 5  1016 cm2, 30 keV, room temperature, helium implantation in (001) p-germanium, followed, respectively, by a thermal annealing at 700  C and 500  C during 30 min. The direction where the surface of the sample is located is indicated by an arrow.

Fig. 1 shows typical HAADF images in cross-sectional view of our systems. In both samples, the bubbles are arranged in a layer, located between 100 and 500 nm from the surface. In silicon, the bubble diameter ranges from 5 to 20 nm. The largest bubbles (14–20 nm in diameter) are arranged along a line at the end of the bubble region, whereas smaller bubbles are dispersed in the region closer to the surface. In germanium, the bubbles exhibit diameter between 7 and 25 nm and are dispersed homogeneously in the bubble layer independently of their diameter. STEM-EELS measurements were carried out in the thinnest regions of the samples in order to study single nanobubbles and avoid the overlap of bubbles within the thickness of the sample. In most of the bubbles under study, the blue shift of the helium K-edge is between 1.2 and 1.65 eV, the corresponding estimated initial helium density lies between 80 and 110 He/nm3 and the corresponding pressure is in the range of 2–6 GPa. However, in some cases, very high density (170 He/nm3) and associated pressure (21 GPa) were also obtained. Fig. 2 is a typical example of helium density measurements realized on bubbles embedded in silicon.

FIG. 2. Helium cavities in silicon. (a) Energy-filtered image in the cavityplasmon energy range (12.3–13.3 eV) extracted from the spectrum image revealing a string of seven cavities. (b) Helium chemical map on the same area as (a) showing that only the two cavities labelled B1 and B3 contain helium. (c) EELS spectra extracted in the middle of B1 in black (the data are realigned and treated with Multivariate Statistical Analysis31) and the corresponding extraction of the helium-K-edge in light black. The small peak at 26 eV is an artefact of the CCD camera due to the intensity of the zero-loss peak. (d) Energy-filtered image in the cavity-plasmon energy range (13.2–14.2 eV) extracted from the spectrum image recorded on two tiny bubbles.

[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: 194.167.47.1 On: Tue, 25 Mar 2014 16:25:28

123508-4

David et al.

Figs. 2(a)–2(d) show the analysis for a string of seven large cavities. Here, cavities are clearly highlighted by filtering on the cavity plasmon energy range (12.3–13.3 eV), see Fig. 2(a), whereas the presence of helium is revealed by filtering around 23 eV, Fig. 2(b). From this analysis, it is clear that only two bubbles labelled B1 and B3 contain helium. Empty cavities (B2, B4–7) have likely been sectioned during the sample preparation as they can intersect the surface of the sample and helium has escaped. The helium density is estimated at 106 6 6 He/nm3 (associated pressure 5 GPa) in both cavities (14 and 18 nm in diameter). Furthermore, we succeeded in making quantitative measurements in bubbles as small as few nanometers in diameter. For instance, two smaller bubbles, 6 nm in diameter, labelled B8 and B9, are shown in Fig. 2(d). Only B9 contains helium; the estimated density is 80 6 7 He/nm3 corresponding to a pressure of 2.3 GPa. To perform these measurements, we have chosen specific experimental conditions so that the electron beam does not modify or damage our systems (low current-current density and low magnification). The results are thus reproductible and, under those conditions, the analysis is non destructive.

J. Appl. Phys. 115, 123508 (2014)

vertical profile in the largest bubble, 27 nm in diameter. As shown from the spectra, there is a clear red-shift from t1 (helium K-edge at 22.3 eV) to t4 (21.5 eV). It is worth noting that this variation is time-related rather than spatially related, as measurements obtained in the inverse sequence (from the bottom to the top of the bubble) would show the same trend with time rather than position. The relative helium density measured in the bubbles is mapped in Fig. 3(c). As shown, in the largest bubble, the estimated helium density decreases about 90% during the experiment (from 60 6 5 He/nm3 (1.1 GPa) at the beginning of the experiment to 7 6 6 He/nm3 (0.03 GPa) to the end.) We manage to induce helium density and pressure variation in sub-10 nm in diameter bubble as well, as it is demonstrated in the enlargement of Fig. 3(c) for the smallest bubble (9 nm in diameter). This small bubble experiences a 60% helium density variation during the experiment (from 173 6 8 He/nm3 (21 GPa) to 70 6 6 He/nm3 (2.4 GPa)). Finally, it is worth noting that the bubbles were systematically imaged using STEM-HAADF before and after each STEM-EELS experiment and we did not notice any visible modification of their morphology. This shows that while the electron beam induces helium detrapping from the bubble, it does not induce any destruction of the cavity.

B. Electron-beam-induced pressure modification in single helium nanobubbles

We now show that the electron beam can be used to modify the helium density and corresponding pressure in single nanobubbles. For that purpose, we have adjusted our experimental conditions in order to induce helium detrapping from the bubbles (explained in further details in Sec. III C). Fig. 3(a) shows the helium chemical map acquired on a group of three helium bubbles embedded in germanium. Fig. 3(b) shows the helium K-edge extracted from different successive positions labelled t1 to t4, extracted along a

C. Control of the detrapping rate

The helium detrapping rate can be controlled to a certain extent by varying the experimental conditions. The STEM probe parameters relevant here are the energy of the incident electrons, the current in the probe, the number of pixels recorded in the bubble, and the pixel dwell time. The current in the probe is the number of electrons impacting the sample per second. The pixel dwell-time is the time the STEM probe is held for each pixel in the image during the scan.

FIG. 3. Helium bubbles in germanium. (a) Helium chemical map filtered around 22 eV (60.5 eV). Three bubbles are clearly visible (b) Helium-K edge extracted at different vertical positions in the largest bubble, labelled t1, t2, t3, and t4. Note that both the 1s ! 2p and the 1s ! 3p transitions are visible. (c) Relative helium density map of the three bubbles, the helium density is normalized to the maximum helium density measured in the map. In the inset, helium density map of the smallest bubble normalized to the maximum helium density estimated in the bubble.

[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: 194.167.47.1 On: Tue, 25 Mar 2014 16:25:28

123508-5

David et al.

Moreover, we use the notion of residence time to characterize the time the STEM probe is held over the bubble. We define this time as the number of pixel recorded in the bubble times the pixel-dwell time. For a given current in the probe, the electron fluence is directly proportional to the residence time. Fig. 4 shows the relative helium density map determined on two bubbles of similar diameter in silicon. All the experimental conditions were kept constant except the residence time. The bubble in Fig. 4(a) was analysed using a residence time of 15 s, whereas the residence time was 5 times higher in the case of the bubble in Fig. 4(b). As shown, the residence time has a clear influence on the helium detrapping and it is possible to reduce the residence time so that no helium detrapping is visible. Moreover, to reveal a possible ageing effect, three successive scans were performed on the very same bubble. All scans were recorded after typically 10 to 15 min. The helium density variation as a function of the residence time is plotted Fig. 4(c) for the three scans. As shown, the two first scans were very effective in emptying the bubble: the estimated helium density decreases of about 90% (from 40 He/nm3 to 5 He/nm3—associated pressure from 0.4 to 0.02 GPa). During the third scan, the helium K-edge is barely detectable but the bubble still contains a few helium atoms per cubic nanometer. As seen, within the uncertainties of the experiments, the helium density

FIG. 4. Relative helium density map (normalized to the maximum helium density estimated in the bubble) determined in two helium bubbles in silicon of similar diameter. The data cube was acquired with (a) 500 pixels and (b) 2500 pixels with electron energy of 200 keV. (c) Relative helium density map (normalized to the maximum helium density estimated in the bubble) as a function of the residence time for three successive spectrum images performed on the bubble shown in (b).

J. Appl. Phys. 115, 123508 (2014)

measured at the end of a scan is the same as the helium density measured at the beginning of the next one, suggesting that no ageing effect occurs. The detrapping process can thus be simply stopped by shifting the electron beam away from the bubble. Fig. 5 shows the relative helium density plots as a function of the residence time for different experimental conditions. The use of a monochromator allows to modify in real time the current in the probe by a slight detuning without changing any other illumination conditions. This scheme was used to record the data plotted in Fig. 5(a), allowing to study the influence of the probe current while analysing the very same bubble. The current in the probe was increased by approximately a factor of 2 in the middle of the spectrum image. As shown, the helium detrapping rate is clearly enhanced by the increase of the probe current. Moreover, spectrum images were recorded in bubbles of same diameter and containing the same initial helium density using an electron probe of 80 keV and 200 keV. As shown in Fig. 5(b), the

FIG. 5. Relative helium density map (normalized to the maximum helium density estimated in the bubble) as a function of the residence time under different experimental conditions. Each point is an average of 8 points extracted horizontally in the middle of the bubble. The corresponding standard deviation gives the accuracy of the experiment. (a) Influence of the probe current for a 18 nm in diameter bubble in silicon. The probe current is increased by a factor of 1.5 between 70 and 160 s as compared to its value between 0 and 70 s. In the inset, the HAADF image recorded together with the spectrum image. (b) Influence of the electron beam energy for two bubbles in silicon, of similar diameter and containing the same initial helium density. The electron probe was set to 80 keV (open circle) and to 200 keV (full square).

[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: 194.167.47.1 On: Tue, 25 Mar 2014 16:25:28

123508-6

David et al.

helium detrapping rate clearly increases (an order of magnitude) with the decrease of the electron beam energy. Finally, the embedded material has a clear influence on the desorption rate. Helium desorption from a bubble embedded in silicon is compared to that from a bubble in germanium in Fig. 6. This is a typical example of what has been observed in all bubbles under study. Both bubbles are of similar diameter and they contain the same initial density, they were analysed using the same sampling, the same probe current, and 200 keV electrons. As shown, the helium detrapping rate is an order of magnitude higher in germanium than in silicon. At 80 keV, the desorption rate in germanium is so high that we did not find any experimental conditions under which helium density is measurable. It is worth noting that, keeping the experimental conditions constant, the helium detrapping rate may vary of a few percents from a bubble to another. This rate may indeed not only depend on the experimental conditions but also on the bubble characteristics: initial helium density, diameter of the bubble, position of the bubble with respect to other bubbles and in the thin foil. This last factor is important as it is linked to the influence of the free surfaces. This is a non negligible factor affecting the detrapping kinetics. Indeed, the migration energy of helium in silicon is small, 0.71 eV.37 Thus, once detrapped, helium atoms can migrate in the sample. They can then be trapped again by the bubble or another close one; they can stay in the sample as tetrahedral interstitials but can also migrate towards the surface. We thus measure the net result of helium detrapping by the electron beam and retrapping by diffusion.

D. Physical mechanisms of detrapping

Material modification under electron irradiation is intimately associated to the energy transferred by the electrons to the target atoms while traveling through the material. Two physical mechanisms are involved in the energy loss by a particle through the matter. Elastic scattering leads to the

FIG. 6. Relative helium density map (normalized to the maximum helium density estimated in the bubble) as a function of the residence time in two bubbles of similar diameter and initial helium density embedded in silicon (open circle) and germanium (full square).

J. Appl. Phys. 115, 123508 (2014)

direct transfer of momentum from electrons to the target nuclei, while inelastic scattering is due to the Coulomb interaction of incident electrons with the ones of the target atoms. These two mechanisms of energy loss are, respectively, described by the Mott and Massey cross-section38 and the Bethe formalism.39 In the energy range available in a TEM, both these cross-sections increase with a decrease of the energy of the incident particles, i.e., the electron-matter interaction is enhanced for lower electron energy. In the TEM, electron irradiation can lead to well known dramatic effects in the sample under examination: knock-on damage, electric charging, beam heating or radiolysis, see, for instance, Refs. 40 and 41. The release of helium has been previously observed under 10 keV helium implantation in silicon42 and 1 MeV electron irradiation in SiC.43 Comparing the nature of the incident particles and their energy, the detrapping mechanisms are probably not the same in these two experiments. Moreover, rare gas bubbles destruction due to interaction with fission fragments is well known to occur in UO2, this mechanism is termed re-solution (for a review, see Ref. 44). Two mechanisms are considered to explain this destruction of the bubbles: a ballistic mechanism45 resulting in the atom by atom ejection from the bubble and a mechanism driven by electronic excitations leading to the global destruction of the bubble.46 This last mechanism can be assimilated to thermal spikes.47 Recent molecular dynamic simulations show that both mechanisms play a role in the case of helium bubbles.48 In the present work, helium detrapping from nanobubbles is clearly induced by the electron beam. However, we do not observe any destruction of the bubbles: the HAADF images remain the same, before and after electron irradiation. Moreover, the detrapping rate increases with decreasing incident energy or with increasing current; further, it depends on the embedding material. In the energy range involved in the TEM, both elastic and inelastic interactions may be responsible for the helium detrapping from nanobubbles. Several scenarios are thus conceivable: helium detrapping through direct ballistic collisions (or secondary collisions), through a percolation path involving irradiation-induced point defects, or through a temperature increase due to beam heating. The sample being not homogeneous (helium embedded in silicon), the three dimensional heat conduction through the helium-embedding material interface certainly has to be taken into account in order to determine the temperature increment, DT, of helium in the bubble during electron irradiation. In a first attempt, DT can be estimated using the approach developed in Ref. 40 in the case of a thin homogeneous specimen (i.e., in our case, a layer of helium in the TEM). Using this approach, the heat generation due to the electron beam is simply balanced by the heat loss by conduction (the radiation term being negligible). The temperature increment is in this case given by    dQ 2t 1 0:58 þ 2 ln ; DT ¼ dt d 4pjh

(4)

where dQ dt is the heat deposited in the specimen by unit time, h is the bubble thickness, d the diameter of the probe, and j

[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: 194.167.47.1 On: Tue, 25 Mar 2014 16:25:28

123508-7

David et al.

J. Appl. Phys. 115, 123508 (2014)

is the thermal conductivity of helium (which depends on the helium density). The heat deposited in the specimen per second by the beam is given by the product of the stopping power  dE dx (the mean energy loss by an incident particle in the matter by unit length) times the number of incident particles   dQ dE I ¼  h ; (5) dt dx e where I is the current in the beam and e the electron charge. The stopping power can be expressed in terms of the stopping cross-section of inelastic scattering rst as follows:39   dE ¼ rst  ER  nHe ;  (6) dx where ER the Rydberg constant (¼13.606 eV). The stopping cross-section for elastic scattering being several times lower, it is not taken into account in our calculations. As shown, in this model, the temperature increment in the sample is directly proportional to the current in the probe which is low in STEM mode. Putting numbers into Eq. (4) leads to a temperature increment of less than 1 K in the case of a layer of helium of thickness 20 nm, using a probe diameter ˚ , and a thermal conductivity of helium gas in the of 0.7 A normal condition of temperature and pressure (0.15 W/m/K, Ref. 49). Detrapping through a percolation path created in the embedding material can also be envisaged. This mechanism requires the formation of a channel and thus the formation of point defects in the embedding material through the production of Frenkel pairs (a vacancy and a self interstitial) by the electron irradiation. Assuming the formation of point defects through elastic collisions, the maximum energy transferred from an electron (rest mass m, energy E in volt) to a nucleus of mass M is given by the relativistic expression Tm ¼

2EðE þ 2mc2 Þ : Mc2

where a0 is the Bohr radius (5.29  1011 m), Z is the atomic number of the target atom, a ¼ Z/137 and b is the relativistic qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi correction factor, b ¼ v=c ¼ 1  ð1 þ E=ðmc2 ÞÞ2 . Fig. 7(a) shows the plot of rd as a function of the energy of the incident electrons in the case of silicon (in black dash dot) and of germanium (in green). Ed being a highly anisotropic quantity, we have taken for these calculations the smallest values of Ed which correspond in the case of silicon and germanium, to the displacement of an atom in the h111i lattice direction: Ed(Si) ¼ 12.5 eV (Ref. 52) and Ed(Ge) ¼ 11.5 eV.53 As seen, for both embedded materials, rd decreases with increasing electron energy. This means that the helium detrapping rate should decrease with the increased electron energy. Moreover, for silicon, the electron energy threshold for creation of a Frenkel pair is 140 keV. Thus, there should be no detrapping for an electron energy of 80 keV. In the case of germanium, the threshold for the formation of Frenkel pairs (300 keV) is not even reached with 200 keV electrons. Moreover, we do not observe any apparent modification of the bubble surroundings after the STEM-EELS experiment. Based on this analysis, it turns out that a detrapping through a percolation mechanism is very unlikely. Finally, we investigate direct helium detrapping through binary collisions, i.e., an electron transfers a momentum to a helium atom, which is directly ejected from the bubble

(7)

The displacement cross-section rd(E) is evaluated as the integrated product of dr(T), the differential cross-section for energy transfer T from an electron of energy E through elastic collision, and of Pd(T) the probability to produce a displacement due to energy transferred T. dr(T) is given by the Mott and Massey expression.38 Moreover, in the case of electrons through solid, Pd(T) is often taken equal to 1, which means that cascade effects are neglected. The integration is done over all possible values of T, i.e., between the threshold displacement energy Ed and the maximum energy that can be transferred to a target atom Tm. rd(E) is then given by the McKinley and Feshbach expression50,51 !" rffiffiffiffiffiffi 2 1  b Tm Tm 2 2 2 rd ¼ 4pa0 ER Z þ 2pab 4 2 4 Ed Ed m cb   2  Tm   ð1 þ 2pabÞ ; (8)  b þ pab ln Ed

FIG. 7. (a) Displacement cross-section, rd, as a function of the energy of the incident electrons in the case of silicon (dashed dotted black) with Ed ¼ 12.5 eV (Ref. 52) and in the case of germanium (green) with Ed ¼ 11.5 eV (Ref. 53). The arrows indicate the energies at which the experiments were conducted. (b) Helium detrapping cross-section, rdt in the case of helium bubble in silicon (in black with a detrapping energy, Edt of 1.8 eV), in the case of helium bubble in germanium (in red with a Edt ¼ 1 eV), and in the case of helium bubble in a material with a high hypothetical detrapping energy (in blue with Edt ¼ 5 eV).

[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: 194.167.47.1 On: Tue, 25 Mar 2014 16:25:28

123508-8

David et al.

J. Appl. Phys. 115, 123508 (2014)

without any thermalisation with the other helium atoms in the bubble. In first approximation, assuming no retrapping process, the helium density in a bubble as a function of time can be expressed as t (9) nHe ðtÞ ¼ nHe0 exp  ; s

remains, however, difficult since we measure the net result of helium detrapping and retrapping and this last mechanism depends on many unknown factors (in particular, the position of the bubble with respect to the surface and to other bubbles). IV. CONCLUSION

with the constant rate s¼

1 ; ðurdt Þ

(10)

where nHe0 is the initial helium density, u is the dose rate given by u ¼ eI pd1 2 . The detrapping cross-section, rdt, is proportional to the probability for a helium atom to be emitted from a bubble. It can be evaluated in a similar manner as the displacement cross-section but in this case, the displacement energy is replaced by the detrapping energy, i.e., the minimum energy that has to be transferred to a helium atom in order to eject it out of the bubble. As a lower bound, we assume the detrapping energy is equal to the desorption energy although it should slightly depend on temperature and pressure in the bubble.54 This energy has been determined both experimentally55,56 and theoretically:54 Edt ¼ 1.8 eV in silicon and is smaller in germanium Edt ¼ 1 eV.57 Fig. 7(b) shows the detrapping cross-section of helium as a function of the detrapping energy, so as a function of the embedding material. The detrapping cross-section in the case of a high hypothetical detrapping energy (5 eV) is also shown for comparison. As seen, whatever the detrapping energy, the displacement cross-section is decreasing when electron energy increases. This is the case even if we consider a high detrapping energy of 5 eV, for instance (blue curve on Fig. 7(b)). At first glance, this behavior is surprising: it is opposite to what is observed when calculating displacement cross-section (see, for instance, the displacement cross-section for silicon). This is, however, the behavior observed when the displacement energy is low (typically around 1 eV),9 or when the target atoms are light. One of these two cases is not often encountered. Moreover, at a given electron energy, the detrapping cross-section for helium in germanium is approximately twice that for helium in silicon. This behavior would explain why for a given electron energy, helium detrapping is much more efficient in germanium than in silicon and for a given embedded material, why it is much more efficient with 80 keV electrons than with 200 keV electrons. Moreover, we clearly see that the current density is the relevant electron beam parameter in this model. In microscopes with sub-angstrom probe, the current density and thus the dose-rate is very high due to the small probe size (0.6–1.1  1025 e–/cm2/s in the experiments performed here), so high enough to counterbalance a low detrapping cross-section. This is in line with the observations of Fig. 5(a), where an increase of the current and thus of the dose-rate leads to an increase of the detrapping rate. Thus, this qualitative analysis suggests that the irradiation-induced helium detrapping observed in this study could be interpreted in terms of direct elastic collisions between helium atoms and electrons. A quantitative analysis

In summary, we have demonstrated that spatially resolved EELS can be used to modify the helium density and corresponding pressure in single helium nanobubble embedded in an elemental semiconductor. We have shown first that using a low spatial sampling (a small residence time), it is possible to study the bubbles without damaging them. Moreover, we have studied in details the experimental conditions under which helium detrapping occurs. We have clearly evidenced that the detrapping rate is strongly dependent on the incident electron energy, current, and on the embedding material. Our observations suggest that the underlying mechanism is a direct helium detrapping through ballistic collisions, leading to the ejection of the helium atoms from the bubble. ACKNOWLEDGMENTS

We gratefully acknowledge Marc Marteau for performing the implantations and Erwan Oliviero and Christiane Jaouen for fruitful discussions. The Region Poitou-Charentes is acknowledged for financial support. G.A.B. is grateful to NSERC for supporting part of this work. The STEM and EELS work were carried out at the Canadian Centre for Electron Microscopy, a facility supported by NSERC and McMaster University. 1

C. F. Lima and A. Howie, Philos. Mag. 34, 1057 (1976). W. Xu, Y. Zhang, G. Cheng, W. Jian, P. C. Millet, C. C. Koch, S. N. Mathaudhu, and Y. Zhu, “In situ atomic scale observation of irradiation induced void formation,” Nat. Commun. 4, 2288 (2013). 3 D. Ugarte, “Curling and closure of graphitic networks under electron irradiation,” Nature 359, 707 (1992). 4 T. Yokota, M. Murayama, and J. Howe, “In situ tem investigation of melting in submicron al-si alloy particles under electron irradiation,” Phys. Rev. Lett. 91, 265504 (2003). 5 O. Ugurlu, J. Haus, A. A. Gunawan, M. G. Thomas, S. Maheshwari, M. Tsapatsis, and K. A. Mkhoyan, “Radiolysis to knock-on damage transition in zeolites under electron beam irradiation,” Phys. Rev. B 83, 113408 (2011). 6 T. J. Woehl, I. A. James, E. Evans, W. D. Ristenpart, and N. D. Browning, “Determination of the mechanisms controlling nanoparticle nucleation and growth,” ACS Nano 6, 8599 (2012). 7 A. Storm, J. Chen, X. Ling, H. W. Zandbergen, and C. Dekker, Nature Mater. 2, 537 (2003). 8 M.-Y. Wu, R. M. M. Smeets, M. Zandbergen, U. Ziese, D. Krapf, P. E. Batson, N. H. Dekker, C. Dekker, and H. W. Zandbergen, “Control of shape and material composition of solid-state nanopores,” Nano Lett. 9, 479 (2009). 9 H.-M. Kim, M.-H. Lee, and K.-B. Kim, “Theoretical and experimental study of nanopore drilling by a focused electron beam in transmission electron microscopy,” Nanotechnology 22, 275303 (2011). 10 V. Oleshko and J. Howe, “Are electron tweezes possible,” Ultramicroscopy 111, 1599 (2011). 11 P. E. Batson, A. Reyes-Coronado, R. G. Barrera, A. Rivacoba, P. M. Echenique, and J. Aizpurua, “Plasmonic nanobilliards: Controlling nanoparticle movement using forces induced by swift electrons,” Nano Lett. 11, 3388 (2011). 2

[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: 194.167.47.1 On: Tue, 25 Mar 2014 16:25:28

123508-9 12

David et al.

R. Dhaka and S. Barman, “Plasmon excitations by photoelectron emission from rare gas nanobubbles in aluminum,” Phys. Rev. Lett. 104, 036803 (2010). 13 A. D. Maestro, M. Boninsegni, and I. Affleck, Phys. Rev. Lett. 106, 105303 (2011). 14 G. Petersen, S. Myers, and D. Follstaedt, Nucl. Instrum. Methods, Phys. Res. B 127/128, 301 (1997). 15 D. Alquier, C. Bongiorno, F. Roccaforte, and V. Raineri, Appl. Phys. Lett. 86, 211911 (2005). 16 X. Ou, R. K€ogler, A. M€ ucklich, W. Skorupa, W. M€ oller, X. Wang, and L. Vines, “The use of nanocavities for the fabrication of ultrathin buried oxide layers,” Appl. Phys. Lett. 94, 011903 (2009). 17 H. Trinkaus and B. N. Singh, “Helium accumulation in metals during irradiation—Where do we stand?,” J. Nucl. Mater. 323, 229 (2003). 18 N. Li, E. Fu, H. Wang, J. Carter, L. Shao, S. Maloy, A. Misra, and X. Zhang, “He ion irradiation damage in fe/w nanolayer films,” J. Nucl. Mater. 389, 233 (2009). 19 A. Kashinath, A. Misra, and M. J. Demkowicz, “Stable storage of helium in nanoscale platelets at semicoherent interfaces,” Phys. Rev. Lett. 110, 086101 (2013). 20 S. Guilbert, T. Sauvage, H. Erramli, M. F. Barthe, P. Desgardin, G. Blondiaux, C. Corbel, and J. Piron, J. Nucl. Mater. 321, 121 (2003). 21 M. L. David, M. F. Beaufort, and J. F. Barbot, “Effect of implant temperature on defects created using high fluence of helium in silicon,” J. Appl. Phys. 93, 1438 (2003). 22 E. Oliviero, M. Beaufort, and J. Barbot, “Influence of dose rate on bubble formation by high energy he implantation in silicon,” J. Appl. Phys. 90, 1718 (2001). 23 W. Lu and C. M. Lieber, “Nanoelectronics from the bottom up,” Nature Mater. 6, 841 (2007). 24 S. Reboh, M. F. Beaufort, J. F. Barbot, J. Grihle, and P. F. P. Fichnter, Appl. Phys. Lett. 93, 022106 (2008). 25 S. Reboh, J. F. Barbot, M. Vallet, M. F. Beaufort, F. Rieutord, F. Mazen, N. Cherkashin, P. F. P. Fichtner, and J. Grilhe, “Nanoscale organization by elastic interactions between h and he platelets in si,” J. Appl. Phys. 114, 073517 (2013). 26 C. Walsh, J. Yuan, and L. Brown, Philos. Mag. A 80, 1507 (2000). 27 D. Taverna, M. Kociak, O. Stephan, A. Fabre, E. Finot, B. Decamps, and C. Colliex, Phys. Rev. Lett. 100, 035301 (2008). 28 M.-L. David, F. Pailloux, V. Mauchamp, and L. Pizzagalli, “In situ probing of helium desorption from individual nanobubbles under electron irradiation,” Appl. Phys. Lett. 98, 171903 (2011). 29 J. Rife, S. Donnelly, A. Lucas, J. Gilles, and J. Ritsko, Phys. Rev. Lett. 46, 1220 (1981). 30 A. Lucas, J. Vigneron, S. Donnelly, and J. Rife, Phys. Rev. B 28, 2485 (1983). 31 N. Bonnet, J. Microsc. 190, 2 (1998). 32 R. F. Egerton, Electron Energy-Loss Spectroscopy in the Electron Microscope (Plenum Press, New York, 1996). 33 S. Frechard, M. Walls, M. Kociak, J. Chevalier, J. Henry, and D. Gorse, “Study by eels of helium bubbles in a martensitic steel,” J. Nucl. Mater. 393, 102 (2009). 34 H. Trinkhaus, Radiat. Eff. 78, 189 (1983). 35 P. Vinet, J. R. Smith, J. Ferrante, and J. H. Rose, “Temperature effects on the universal equation of state of solids,” Phys. Rev. B 35, 1945–1953 (1987). 36 P. Loubeyre, R. LeToullec, J. P. Pinceaux, H. K. Mao, J. Hu, and R. J. Hemley, “Equation of state and phase diagram of solid 4He from

J. Appl. Phys. 115, 123508 (2014) single-crystal x-ray diffraction over a large P-T domain,” Phys. Rev. Lett. 71, 2272–2275 (1993). 37 A. Charaf Eddin, G. Lucas, M. F. Beaufort, and L. Pizzagalli, “Dft calculation of the stability and mobility of noble gas atoms in silicon,” Comput. Mater. Sci. 44, 1030 (2009). 38 N. Mott and H. Massey, The Theory of Atomic Collision (Clarendon Press, Oxford, 1965). 39 M. Inokuti, “Inelastic collision of fast charged particles with atoms and molecules—The bethe theory revisited,” Rev. Mod. Phys. 43, 297 (1971). 40 R. Egerton, P. Li, and M. Malac, “Radiation damage in the TEM and SEM,” Micron 35, 399 (2004). 41 R. Egerton, “Control of radiation damage in the TEM,” Ultramicroscopy 127, 100 (2013). 42 K. Wittmaack, “In situ observation of gas reemission and blister rupture during helium implantation in silicon,” Appl. Phys. Lett. 92, 051907 (2008). 43 A. Morono, J. Manzano, M. Malo, and E. R. Hodgson, “Ionizing radiation induced helium release for hot pressed sic,” J. Nucl. Mater. 417, 1038 (2011). 44 D. Olander and D. Wongsawaeng, “Re-solution of fission gas—A review: Part I. Intragranular bubbles,” J. Nucl. Mater. 354, 94 (2006). 45 R. S. Nelson, “The stability of gas bubbles in an irradiation environment,” J. Nucl. Mater. 31, 153 (1969). 46 J. A. Turnbull and R. M. Cornell, “The re-solution of gas atoms from bubbles during the irradiation of uo2,” J. Nucl. Mater. 36, 161 (1970). 47 K. Govers, C. L. Bishop, D. C. Parfitt, S. E. Lemehov, M. Verwerft, and R. W. Grimes, “Molecular dynamics study of xe bubble re-solution in uo2,” J. Nucl. Mater. 420, 282 (2012). 48 D. Parfitt and R. Grimes, “Predicted mechanisms for radiation enhanced helium resolution in uranium dioxide,” J. Nucl. Mater. 381, 216 (2008). 49 E. Bich, J. Millat, and E. Vogel, “The viscosity and thermal properties of pure monoatomic gases from their normal boiling point up to 5000 k in the limit of zero density and at 0.101325 mpa,” J. Phys. Chem. Ref. Data 19, 1289 (1990). 50 W. A. McKinley and H. Feshbbach, “The coulomb scattering of relativistic electrons by nuclei,” Phys. Rev. 74, 1759 (1948). 51 J. Reyes-Gasga and R. Garcia-Garcia, “Analysis of the electron-beam radiation damage of tem samples in the acceleration energy range from 0.1 to 2 mev using the standard theory for fast electrons,” Radiat. Phys. Chem. 64, 359 (2002). 52 E. Holmstr€ om, A. Kuronen, and K. Nordlund, “Threshold defect production in silicon determined by density functional theory molecular dynamics simulations,” Phys. Rev. B 78, 045202 (2008). 53 E. Holmstr€ om, K. Nordlund, and A. Kuronen, “Threshold defect production in germanium determined by density functional theory molecular dynamics simulations,” Phys. Scr. 81, 035601 (2010). 54 A. Charaf Eddin and L. Pizzagalli, “First-principles calculations of helium and neon desorption from cavities in silicon,” J. Phys.: Condens. Matter 24, 175006 (2012). 55 S. Godey, E. Ntsoenzok, T. Sauvage, A. van Veen, F. Labohm, M. Beaufort, and J. Barbot, “Helium desorption from cavities induced by high energy 3he and 4he implantation in silicon,” Mater. Sci. Eng. B 73, 54 (2000). 56 E. Oliviero, M. David, M. Beaufort, J. Barbot, and A. van Veen, “On the effect of implantation temperature in helium implanted silicon,” Appl. Phys. Lett. 81, 4201 (2002). 57 L. Pizzagalli, private communication (2013).

[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: 194.167.47.1 On: Tue, 25 Mar 2014 16:25:28