PHYS07200604023 Prabodh Kumar Kuiri - Homi Bhabha National ...
PHYS07200604023 Prabodh Kumar Kuiri - Homi Bhabha National ...
PHYS07200604023 Prabodh Kumar Kuiri - Homi Bhabha National ...
Transform your PDFs into Flipbooks and boost your revenue!
Leverage SEO-optimized Flipbooks, powerful backlinks, and multimedia content to professionally showcase your products and significantly increase your reach.
ENERGETIC Au IRRADIATION EFFECTS ON<br />
NANOCRYSTALLINE ZnS FILMS DEPOSITED<br />
ON Si AND Au NANOPARTICLES EMBEDDED<br />
IN SILICA GLASS<br />
By<br />
PROBODH KUMAR KUIRI<br />
INSTITUTE OF PHYSICS<br />
BHUBANESWAR, INDIA<br />
A THESIS SUBMITTED TO THE<br />
BOARD OF STUDIES IN PHYSICAL SCIENCES<br />
IN PARTIAL FULFILLMENT OF REQUIREMENTS<br />
FOR THE DEGREE OF<br />
DOCTOR OF PHILOSOPHY<br />
OF<br />
HOMI BHABHA NATIONAL INSTITUTE<br />
April 2010
Statement by Author<br />
This dissertation has been submitted in partial fulfillment of requirements for an<br />
advanced degree at <strong>Homi</strong> <strong>Bhabha</strong> <strong>National</strong> Institute (HBNI) and is deposited in the<br />
Library to be made available to borrowers under rules of the HBNI.<br />
Brief quotations from this dissertation are allowable without special permission,<br />
provided that accurate acknowledgement of source is made. Requests for permission<br />
for extended quotation from or reproduction of this manuscript in whole or in part<br />
may be granted by the Competent Authority of HBNI when in his or her judgment the<br />
proposed use of the material is in the interests of scholarship. In all other instances,<br />
however, permission must be obtained from the author.<br />
(Probodh <strong>Kumar</strong> <strong>Kuiri</strong>)<br />
ii
Declaration<br />
I, hereby declare that the investigation presented in the thesis has been carried out<br />
by me. The work is original and has not been submitted earlier as a whole or in part<br />
for a degree / diploma at this or any other Institution / University.<br />
(Probodh <strong>Kumar</strong> <strong>Kuiri</strong>)<br />
iii
.<br />
To My PARENTS<br />
iv
Acknowledgements<br />
I take this opportunity to express my sincer gratitude to my advisor, Prof. D. P. Mahapatra,<br />
for his constant encouragement, support, and invaluable guidance throughout<br />
my research period at the Institute of Physics. I learnt many things from his love for<br />
Physics, intuative way of looking into the problems, and constant drive for perfection.<br />
He has helped me grow both academically and personally. I am also thankful to him<br />
for giving me complete freedom in my academic and social life during the period of<br />
this work. Without his help this dissertation would have not been possible. I heartly<br />
thank him and his family for making my stay valuable in the Institute.<br />
I am grateful to Dr. D. Kanjilal and Prof. N. C. Mishra for their encouragement,<br />
suggestions, and help. It was always a nice experience while discussing physics with<br />
them. My sincer thanks to Dr. B. S. Acharya, Prof. S. M. Bhattacharyya, Prof.<br />
G. Tripathy, Dr. A. Tripathi, Prof. S. K. Pal, and Dr. G. Kuri for their effective<br />
discussions and help received at various stages of this work.<br />
I have spent many enjoyable hours working with my fellow group members. I am<br />
happy to thank them all, both past and present, for their support and friendship in<br />
and out of the lab., including Dr. S. Mohapatra, Dr. B. Joseph, Mr. H. P. Lenka, Dr.<br />
J. Ghatak, and Mrs. G. Sahu.<br />
I take this opportunity to thank all the experimental condensed matter group<br />
faculty members of the Institute of Physics (IOP), Bhubaneswar for their generous<br />
support in extending different facilities. Special thanks are to Prof. P. V. Satyam,<br />
Prof. S. N. Sahu, and Prof. S. Varma in this regard. I am grateful to Mr. A.<br />
K. Behera, a person who possesses a vast amount of practical knowledge. I learned<br />
something new every time we worked in lab. Sincere thanks to Mrs. R. Dash, Mr. M.<br />
Baliarsingh, and Mr. D. B. Sahi for providing me a homely atmosphere in our lab. I<br />
thank Dr. S. N. Saragni for doing several optical measurements for me. I thank all the<br />
accelerator and other technical staff members of IOP and Inter University Accelerator<br />
Centre, New Delhi for their help during carrying out different experiments. I would<br />
v
like to specially mention the names of M. Majhi, P. K. Biswal, A. Sahu, K. C. Patra,<br />
S. K. Choudhary, A. K. Dash, and P. C. Marandi. I acknowledge the help received<br />
from all scholars of the experimental condensed matter group whenever I was in need.<br />
It is my pleasure to thank my seniours, scholars, friends, and others who have<br />
made my stay in the beautiful campus of IOP a memorable and pleasing one. Special<br />
thanks to Sumalay-da, Jay, Prabir, Anupam, and Kuntala-di for their encouragements,<br />
help, company, and standing by me in good and bad times. I also thank Sandip-da,<br />
Anand, Dipak (Mishra), Raghu, Srikumar-da, Tanay-da, Manas, Sadhana, Ajay, and<br />
Chitrasen.<br />
I would like to thank my friends and others, outside IOP, who have helped me in<br />
one way or other. Special thanks to my friend-cum-elder brother, Manoranjan-da for<br />
his valuable advice, useful comments, and help whenever I needed.<br />
My sincer regards and respect are due to all my teachers, especially Mr. B. B.<br />
Mishra, who had taught me at various stages of my educational career. I am grateful<br />
to late M. M. <strong>Kuiri</strong>, Dr. S. Chattopadhyay, Prof. S. K. Pradhan, and Prof. A. K.<br />
Bhattacharyya for their encouragement to pursue research work.<br />
I am indebted to the former director, Dr. Y. P. Viogy, the present director, Prof.<br />
A. Khare of IOP, the principal Dr. D. N. Mahata of A. M. College, Jhalda, and all<br />
the members (academic and non-academic) of both the institutions for their timely<br />
help, cooperation, and support whenever needed. Special thanks are to all the people<br />
who have maintained the excellent library and computer facilities at IOP.<br />
Finally, my sincer gratitude to my grand mother, parents, in-laws, wife, and brothers<br />
for their love and unwavering support. They have always been around when I<br />
needed them and they continue to support me in my decisions. I would have never<br />
made it this far without them.<br />
Date : April 6, 2010<br />
Bhubaneswar<br />
(Probodh <strong>Kumar</strong> <strong>Kuiri</strong>)<br />
vi
List of Publications / Preprints<br />
In journals:<br />
1 Low energy O induced redistribution of nanosized gold inclusions in SiO 2 grown<br />
on Si(100),<br />
B. Joseph, S. Mohapatra, B. Satpati, H. P. Lenka, P. K. <strong>Kuiri</strong> and D. P.<br />
Mahapatra,<br />
Nucl. Instrum. Methods Phys. Res. B 227, 559 (2005).<br />
2 Size saturation in low energy ion beam synthesized Au nanoclusters and their<br />
size redistribution with O irradiation,<br />
B. Joseph, S. Mohapatra, H. P. Lenka, P. K. <strong>Kuiri</strong> and D. P. Mahapatra,<br />
Thin Solid Films 492, 35 (2005).<br />
*3 Observation of ZnS nanoparticles sputtered from ZnS films under 2 MeV Au<br />
irradiation,<br />
P. K. <strong>Kuiri</strong>, B. Joseph, J. Ghatak, H. P. Lenka, G. Sahu, B. S. Acharya and<br />
D. P. Mahapatra,<br />
Nucl. Instrum. Methods Phys. Res. B 248, 25 (2006).<br />
*4 Effect of Au irradiation energy on ejection of ZnS nanoparticles from ZnS film,<br />
P. K. <strong>Kuiri</strong>, J. Ghatak, B. Joseph, H. P. Lenka, G. Sahu, N. C. Mishra, A.<br />
Tripati, D. Kanjilal and D. P. Mahapatra,<br />
J. Appl. Phys. 101, 014313 (2007).<br />
5 Effect of 100 MeV Au irradiation on embedded Au nanoclusters in silica glass,<br />
B. Joseph, J. Ghatak, H. P. Lenka, P. K. <strong>Kuiri</strong>, G. Sahu, N. C. Mishra and D.<br />
P. Mahapatra, Nucl. Instrum. Methods Phys. Res. B 256, 659 (2007).<br />
6 Low energy C n cluster ion induced damage effects in Si(100) substrates,<br />
H. P. Lenka, B. Joseph, P. K. <strong>Kuiri</strong>, G. Sahu and D. P. Mahapatra,<br />
Nucl. Instrum. Methods Phys. Res. B 256 , 665 (2007).<br />
vii
7 Synthesis of alloy metal nanoclusters in silica glass by sequential ion implantation,<br />
B. Joseph, H. P. Lenka, P. K. <strong>Kuiri</strong>, D. P. Mahapatra and R. Kesavamoorthy,<br />
Int. J. Nanoscience. 6, 432 (2007).<br />
8 Anomalous diffusion of Au in mega-electron-volt Au implanted SiO 2 /Si(100),<br />
S. Mohapatra, J. Ghatak, B. Joseph, H. P. Lenka, P. K. <strong>Kuiri</strong> and D. P.<br />
Mahapatra,<br />
J. Appl. Phys. 101, 063542 (2007).<br />
9 Formation and growth of SnO 2 nanoparticles in silica glass by Sn implantation<br />
and annealing,<br />
P. K. <strong>Kuiri</strong>, H. P. Lenka, J. Ghatak, G. Sahu, B. Joseph and D. P. Mahapatra,<br />
J. Appl. Phys. 102, 024315 (2007).<br />
10 MeV Au irradiation induced nanoparticle formation and recrystallization in a<br />
low energy Au implanted Si layer,<br />
G. Sahu, B. Joseph, H. P. Lenka, P. K. <strong>Kuiri</strong>, A. Pradhan and D. P. Mahapatra,<br />
Nanotechnology 18, 495702 (2007).<br />
*11 Observation of a universal aggregation mechanism and a possible phase transition<br />
in Au sputtered by swift heavy ions,<br />
P. K. <strong>Kuiri</strong>, B. Joseph, H. P. Lenka, G. Sahu, J. Ghatak, D. Kanjilal and D.<br />
P. Mahapatra,<br />
Phys. Rev. Lett. 100, 245501 (2008).<br />
12 Study of low energy Si − 5 and Cs − implantation induced amorphization effects in<br />
Si(100),<br />
H. P. Lenka, B. Joseph, P. K. <strong>Kuiri</strong>, G. Sahu, P. Mishra, D. Ghose and D. P.<br />
Mahapatra,<br />
J. Phys. D.: Appl. Phys. 41, 215305 (2008).<br />
13 The mechanism of ion induced amorphization in Si,<br />
H. P. Lenka, U. M. Bhatta, P. K. <strong>Kuiri</strong>, G. Sahu, B. Joseph, B. Satpati and<br />
D. P. Mahapatra,<br />
arXiv:0811.0806v4 [cond-mat.mtrl-sci] (2008).<br />
14 Effect of oxidation temperature on the photoluminescence of Sn implanted and<br />
annealed SiO 2 /Si,<br />
viii
P. K. <strong>Kuiri</strong> and M Ghosh,<br />
Adv. Sci. Lett. 2, 35 (2009).<br />
*15 Tuning size and shape of Au nanoparticles embedded in silica glass by swift heavy<br />
ion irradiation,<br />
P. K. <strong>Kuiri</strong>, B. Joseph, J. Ghatak, H. P. Lenka and G. Sahu<br />
Adv. Sci. Lett. (accepted) (2010).<br />
*16 Modification of nanocrystalline ZnS thin films by Au ion irradiation: Role of<br />
energy loss processes,<br />
P. K. <strong>Kuiri</strong>, H. P. Lenka, G. Sahu, D. Kanjilal and D. P. Mahapatra,<br />
(to be published).<br />
In proceedings:<br />
1 A low energy negative ion implanter for surface modification studies at IOP,<br />
Bhubaneswar,<br />
B. Joseph, A. K. Behera, S. Mohapatra, P. <strong>Kuiri</strong>, and D. P. Mahapatra,<br />
Solid State Phys. (India) 46, 271 (2003).<br />
2 Study of Au-Ag alloy nanoparticles in silica glass synthesized by sequential low<br />
energy ion implantation,<br />
B. Joseph, S. Mohapatra, H. P. Lenka, P. K. <strong>Kuiri</strong>, and D. P. Mahapatra,<br />
Proceedings of the DAE Symposium on Solid State Phys., Amritsar, India, December<br />
2004.<br />
——————————————————————-<br />
* indicates papers on which this thesis is based.<br />
ix
Synopsis<br />
A major part of the modern technology depends on materials with precise controlled<br />
properties. Ion implantation is a favoured method to achieve controlled modification<br />
of surface and near-surface regions of various materials [1]. It is a key doping technique<br />
in semiconductor processing, and it is used to induce phase formation, amorphization,<br />
grain growth, structural transformations, tailoring optical band gap or mixing in a<br />
wide range of materials. Also energetic ion irradiation of solids leads to the ejection<br />
of atoms, molecules, and sometimes clusters from the surface of the target. The<br />
observation of stable, intact clusters of atoms and molecules, due to ion impact [2],<br />
is rather surprising because the energies involved in the ion-target collisions, that<br />
generate sputtered atoms, are typically much larger than the binding energies (∼ 1–2<br />
eV) holding the clusters together. Hence it is difficult to understand how stable clusters<br />
can form during ion sputtering. Eversince the first observation, the fundamental<br />
question of the cluster emission mechanism has attracted considerable attention both<br />
experimentally [3] and theoretically [4,5,6].<br />
As an ion penetrates a solid, it loses energy due to two distinct interactions [1].<br />
At low energies, it loses energy by elastic, nuclear collisions with the target atoms<br />
described by the well known Rutherford cross sections for the Coulomb interaction<br />
between the charged atomic nuclei. At higher energies, it loses energy inelastically<br />
by exciting and ionizing electrons of target atoms. Depending upon the incident ion<br />
energy and ion-solid combinations, there can be production of elastic [7] or inelastic [8]<br />
thermal spikes leading to local temperature rise well above the melting temperature of<br />
the materials. In fact, inelastic thermal spikes can sometimes result in the formation of<br />
a cylindrically-shaped molten region around the ion track. These are typically several<br />
nm in diameter [8]. This process can therefore bring about structural changes in the<br />
target material. Here, we have utilized energetic ion beams for controlled modifications<br />
of nanocrystalline ZnS films and Au nanoparticles (NPs). We have also looked at<br />
ejection of NPs from these materials. Attempts have been made to understand the<br />
x
underlying physics through a study of the size distributions of ejected particles.<br />
We have studied the effect of Au irradiation on structural, optical, and surface<br />
morphological properties of nanocrystalline ZnS films deposited on Si and silica glass<br />
substrates [9], using a simple thermal evaporation of ZnS powder. The as-deposited<br />
films were of wurtzite structures with an optical band gap of 3.56 eV. X-ray diffraction<br />
measurements indicated the presence of compressive stress with a grain size of ∼ 17.5<br />
nm. Irradiation at MeV energies have been found to result in a considerable grain<br />
growth and a decrease in dislocation density. In addition, the irradiated films were<br />
found to be under tensile stress. The tensile stress produced in the film is found to<br />
be much higher at 100 MeV than at 2 MeV. There is also a change in optical band<br />
gap primarily due to irradiation induced changes of the grain size and the formation<br />
of trap levels. On the other hand no observable changes in grain size or optical band<br />
gap could be seen following a lower energy (35 keV) Au irradiation. The samples<br />
have also been studied using atomic force microscopy. The 35 keV and 2 MeV Au<br />
irradiations are found to produce rough surfaces the characteristics of which have<br />
been studied through the analysis of the power spectral density. At large spatial<br />
frequency, q, (at small length scales) the Fourier exponent, γ (determined from the<br />
fall off going as q −γ ) has been found to have values close to 2.80, 3.11, 3.72, and 2.85<br />
for the as-deposited and the samples irradiated with Au at 35 keV, 2 MeV, and 100<br />
MeV, respectively. The lower values close to 3 could be understood in a model where<br />
evaporation and condensation dominate, the higher value corresponding to a surface<br />
diffusion dominated process.<br />
The effect of Au irradiation on isolated and embedded nanostructures system [10]<br />
has also been studied. For this, spherical Au NPs (of average size 7 nm) were synthesized<br />
in the near surface region of silica glass samples by implanting 32 keV Au − ions<br />
to a fluence of 4×10 16 cm −2 , followed by annealing at 850 ◦ C in air for 1 h. The silica<br />
glass samples so prepared, containing Au NPs are found to show a surface plasmon<br />
resonance (SPR) band centered around 543 nm in the optical absorption spectrum.<br />
The silica glass samples with embedded NPs were subjected to Au irradiations (at<br />
room temperature) at 10 and 100 MeV. The higher energy irradiation was found to<br />
result in an elongation of the embedded Au NPs along the beam direction. At higher<br />
irradiation fluence there was Au loss from the silica matrix. Up to a fluence of 5×10 13<br />
ions cm −2 , the smaller NPs (diameter < 9 nm) are found to grow in size while larger<br />
ones deformed anisotropically along the ion beam direction. The anisotropy in the<br />
larger NPs have been seen to increase with increase in ion fluence. Optical absorption<br />
xi
spectra corresponding to the elongated Au NPs show two SPR bands corresponding to<br />
transverse and longitudinal modes. Further, longitudinal SPR band has been seen to<br />
shift towards higher wavelength with increase in fluence. This shift of the longitudinal<br />
SPR band is a direct consequence of the further elongation of the NPs [11]. However,<br />
in the sample irradiated at a fluence of 1×10 14 ions cm −2 , the Au NPs are seen to be<br />
almost spherical in shape, with a large interparticle separation. The average size is<br />
found to be ∼ 9.2 nm, which is little larger than that of the unirradiated samples. In<br />
fact what is found is a two-dimensional array of small Au NPs, very near the silica<br />
surface which result in an SPR band with a single peak at around 561 nm. This SPR<br />
peak position appeared to be red-shifted as compared to that of the as-grown sample.<br />
Using a simple simulation [12] that uses, NP size, density, and dielectric constant of<br />
the host together with optical density of the NPs, it has been shown that the shift of<br />
the SPR band corresponding to the spherical NPs is, mainly, the result of the change<br />
in matrix dielectric constant, due to the dissolution of Au in the form of atom and/or<br />
small clusters, instead of the growth of the NPs after ion irradiation.<br />
The results on change of shape (from nearly spherical to elongated ones) from<br />
ion irradiation, can be understood within the framework of the inelastic thermal spike<br />
induced ion track formation [8]. Smaller NPs within the ion track completely evaporate<br />
and collect around existing small clusters leading Oswald ripening of small NPs. In<br />
any case they can grow to a size comparable to the track diameter, which is about<br />
8.5 nm in silica glass in the present experimental conditions [8]. On the other hand<br />
larger ones can undergo a melting and due to pressure imbalance can get squeezed<br />
to elongated shapes along the ion tracks which are formed along the beam direction.<br />
The material loss at higher fluence occurs from a squeezing out and vaporization of<br />
NPs formed near the surface.<br />
Following the above study, it was also necessary to see whether the material which<br />
is ejected during ion irradiation really contained any nanoclusters. This is because<br />
the size distribution in that case can be connected to the underlying mechanism of<br />
emission. Therefore, in further pursuance of our investigations we have collected the<br />
ejected particles on catcher foils placed suitably in the reflection geometry with an<br />
aim to look at their size distribution. This has been done in both cases concerning<br />
irradiation of ZnS films and embedded Au NPs in silica glass as mentioned earlier.<br />
In each case the size distribution has been determined from transmission electron<br />
microscopy.<br />
For ZnS [13,14], with 35 keV Au irradiation no NP larger than 1 nm could be<br />
xii
observed on the catcher foil and therefore no size distribution study could be carried<br />
out. However, we have observed wurtzite ZnS NPs on the catcher foils due to MeV Au<br />
irradiations. The ejected NPs have been found to have sizes (diameter) lying between 2<br />
to 7 nm. The most probable size is seen to be around 3−3.5 nm. For particle sizes ≥ 3<br />
nm, the distributions show a power law decay in the form of Y (n) ∼ n −δ ; n and δ being<br />
the number of atoms in a NP and the power law exponent, respectively. In case of Au<br />
irradiation at 2 MeV, the values of exponent, δ, are found to be around 2.60 ± 0.09,<br />
whereas the same for 100 MeV irradiation is found to be around 3.45 ± 0.09. After<br />
correcting the results for cluster breakup effects, the values of power law exponent<br />
reduced to about 2.0 and 2.6 for the above two irradiation energies. A δ value close<br />
to 2 indicates that the NPs are produced when shock waves, generated by subsurface<br />
displacement cascades, ablate the surface [4]. On the other hand, Coulomb explosion<br />
followed by thermal spike [15] induced vaporization of ZnS seems to be the dominant<br />
mechanism regarding material removal at 100 MeV. In such a case the evaporated<br />
material can cool down going into the fragmentation region ejecting clusters with a<br />
power law exponent close to 7/3 [5]. Interestingly enough, no NP could be found<br />
in the catcher foil even during the 100 MeV Au irradiation when the sample was<br />
maintained at liquid nitrogen temperature. This suppression on cluster ejection at low<br />
temperature can be explained as due to an enhanced thermal conductivity resulting<br />
in fast heat dissipation that suppresses thermal spike effects.<br />
However, the scenario become completely different in case of sputtering from embedded<br />
nanostructures [16]. For 10 MeV Au irradiation of Au NPs, embedded in<br />
silica glass, no NP could be observed on the catcher foils. However, crystalline Au<br />
NPs of size 1−20 nm could be seen on the catcher foils for 100 MeV Au irradiation<br />
case. Size distribution, of the ejected NPs, has been seen to follow a similar power<br />
law decay as seen in the sputtered ZnS NPs. Here, we have seen the existence of two<br />
power law exponents: 3/2 for smaller NPs, below 12.5 nm in size, and 7/2 for larger<br />
NPs. The first case can be rationalized as occurring from a steady state aggregation<br />
process [17], independent of cluster size. The later case may come from a dynamical<br />
transition to another steady state where aggregation and evaporation rates are size<br />
dependent. To understand the aggregation of clusters during sputtering at very high<br />
energy (100 MeV), inelastic thermal spike [8] concept has been invoked. Production<br />
of thermal spike in silica, due to 100 MeV Au ions, can lead to a local temperature<br />
rise well above the vaporization temperature of Au. In such a case evaporated Au<br />
clusters can exchange particle between nucleation sites, within the framework of a<br />
xiii
mass-aggregation model, that takes diffusion, aggregation on contact, and dissociation.<br />
This can result in a size independent steady state aggregation process with a δ<br />
value of 3/2 [17]. This happens to be a very general case corresponding to a broad<br />
class of phenomena, such as fish schooling, distribution of wealth, cloud formation,<br />
and polymer gels. On the other hand, a phase transition can occur when the aggregation<br />
and evaporation rates become size dependent leading to a δ value of 7/2 at the<br />
critical point [18]. Since the system indicates a steady state scenario there is no need<br />
to correct the exponent against any breakup effects as applicable for the ZnS NPs<br />
ejected during Au irradiations.<br />
REFERENCES<br />
[1] M. Nastasi, J.W. Mayer and J.K. Hirvonen, Ion-Solid Interactions: Fundamentals and<br />
Applications, Cambridge University Press, (1996).<br />
[2] R.E. Honig, J. Appl. Phys. 29, 549 (1958).<br />
[3] R. Birtcher, S.E. Donnely and K. Nordlund (E. Knystautas, Ed.), Engineering Thin Films<br />
and Nanostructures with Ion Beams, Taylor & Francis, Florida, (2005), Ch. 1.<br />
[4] I.S. Bitensky and E.S. Parilis, Nucl. Instrum. Methods Phys. Res. B 21, 26 (1987).<br />
[5] H.M. Urbassek, Nucl. Instrum. Methods Phys. Res. B 31, 541 (1988).<br />
[6] K.O.E. Henriksson, K. Nordlund and J. Keinonen, Phys. Rev. B 71, 014117 (2005).<br />
[7] J.A. Brinkman, Am. J. Phys. 24, 246 (1956).<br />
[8] M. Toulemonde, J.M. Costantini, Ch. Dufour, A. Meftah, E. Paumier and F. Studer,<br />
Nucl. Instrum. Methods Phys. Res. B 116, 37 (1996).<br />
[9] P.K. <strong>Kuiri</strong>, H.P. Lenka, G. Sahu, D. Kanjilal and D.P. Mahapatra, [to be published].<br />
[10] P.K. <strong>Kuiri</strong>, B. Joseph, J. Ghatak, H.P. Lenka and G. Sahu, Adv. Sci. Lett. (accepted)<br />
(2010).<br />
[11] W. Kreibig and M. Vollmer, Optical Properties of Metal Clusters (Springer Series in<br />
Materials Science Vol. 25) Springer, Berlin (1995).<br />
[12] M.A. Gracia, J. Llopis and S.E. Paje, Chem. Phys. Lett. 315, 313 (1999).<br />
[13] P.K. <strong>Kuiri</strong>, B. Joseph, J. Ghatak, H.P. Lenka, G. Sahu, B.S. Acharya and D.P. Mahapatra,<br />
Nucl. Instrum. Methods Phys. Res. B 248, 25 (2006).<br />
[14] P.K. <strong>Kuiri</strong>, J. Ghatak, B. Joseph, H.P. Lenka, G. Sahu, D.P. Mahapatra, A. Tripathi,<br />
D. Kanjilal and N.C. Mishra, J. Appl. Phys. 101, 014313 (2007).<br />
[15] E.M. Bringa and R.E. Johnson, Phys. Rev. Lett. 88, 165501 (2002).<br />
[16] P.K. <strong>Kuiri</strong>, B. Joseph, H.P. Lenka, G. Sahu, J. Ghatak, D. Kanjilal and D.P. Mahapatra,<br />
Phys. Rev. Lett. 100, 245501 (2008).<br />
[17] H. Takayasu, Phys. Rev. Lett. 63, 2563 (1989).<br />
[18] R.D. Vigil, R.M. Ziff and B. Lu, Phys. Rev. B 38, 942 (1988).<br />
xiv
Contents<br />
Statement by Author<br />
Declaration<br />
Acknowledgements<br />
List of Publications / Preprints<br />
Synopsis<br />
ii<br />
iii<br />
v<br />
vii<br />
x<br />
1 Introduction 1<br />
1.1 Interests and applications of ion beams in materials science . . . . . . . 1<br />
1.2 Ion-solid interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2<br />
1.2.1 Basic concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2<br />
1.2.2 Ion penetration and stopping . . . . . . . . . . . . . . . . . . . 5<br />
1.2.3 Nuclear thermal spike . . . . . . . . . . . . . . . . . . . . . . . 8<br />
1.2.4 Coulomb explosion and electronic thermal spike . . . . . . . . . 9<br />
1.3 Sputtering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />
1.3.1 Sputtering mechanisms . . . . . . . . . . . . . . . . . . . . . . . 12<br />
1.3.2 Size distribution of clusters ejected during ion irradiation . . . . 13<br />
2 Experimental techniques 19<br />
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />
2.2 Sample preparation techniques . . . . . . . . . . . . . . . . . . . . . . . 19<br />
2.2.1 Vacuum evaporation thin film deposition setup . . . . . . . . . . 19<br />
2.2.2 Ion implantation and irradiation . . . . . . . . . . . . . . . . . . 20<br />
2.3 Characterization techniques . . . . . . . . . . . . . . . . . . . . . . . . 23<br />
2.3.1 Rutherford backscattering spectrometry . . . . . . . . . . . . . 23<br />
2.3.2 X-ray diffraction . . . . . . . . . . . . . . . . . . . . . . . . . . 27<br />
xv
2.3.3 Transmission electron microscopy . . . . . . . . . . . . . . . . . 29<br />
2.3.4 Atomic force microscopy . . . . . . . . . . . . . . . . . . . . . . 32<br />
2.3.5 Optical absorption . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />
3 Modification of nanocrystalline ZnS films by Au irradiation 38<br />
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38<br />
3.2 Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />
3.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />
3.3.1 Energy loss processes . . . . . . . . . . . . . . . . . . . . . . . 40<br />
3.3.2 Thickness and composition of the ZnS films . . . . . . . . . . . 41<br />
3.3.3 Structural properties . . . . . . . . . . . . . . . . . . . . . . . . 41<br />
3.3.4 Optical properties . . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />
3.3.5 Surface morphology . . . . . . . . . . . . . . . . . . . . . . . . . 47<br />
4 Ejection of ZnS nanoparticles from ZnS films by Au irradiation 52<br />
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52<br />
4.2 Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53<br />
4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />
4.3.1 Ejection of ZnS NPs due to 2 MeV Au irradiation . . . . . . . . 54<br />
4.3.2 Ejection of ZnS NPs due to 100 MeV Au irradiation . . . . . . . 58<br />
4.3.3 A power law behaviour in nanoparticle size distribution . . . . . 60<br />
4.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62<br />
4.4.1 NP ejection mechanisms . . . . . . . . . . . . . . . . . . . . . . 62<br />
4.4.2 Temperature dependence of ZnS NP ejection under 100 MeV<br />
Au irradiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />
5 Anisotropic deformation of Au nanoparticles in silica glass by swift<br />
heavy ion irradiation 67<br />
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />
5.2 Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68<br />
5.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />
5.3.1 Synthesis of Au NPs in silica glass . . . . . . . . . . . . . . . . . 69<br />
5.3.2 Shape deformation of the NPs: TEM studies . . . . . . . . . . . 71<br />
5.3.3 Outward movement and loss of Au: RBS studies . . . . . . . . . 73<br />
5.3.4 Shape deformation of the NPs: OA studies . . . . . . . . . . . . 74<br />
5.3.5 Dependence of longitudinal SPR peak position on aspect ratio . 76<br />
xvi
5.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77<br />
6 Observation of a universal aggregation mechanism in Au sputtered<br />
by swift heavy ions 82<br />
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />
6.2 Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84<br />
6.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86<br />
6.3.1 Ejection of Au NPs: Dependence on ion fluence . . . . . . . . . 86<br />
6.3.2 Crystal structure of the ejected NPs . . . . . . . . . . . . . . . . 87<br />
6.3.3 Loss of Au content in the target: RBS studies . . . . . . . . . . 87<br />
6.3.4 A power law behaviour in NP size distribution . . . . . . . . . . 88<br />
6.4 Discussion: NP ejection mechanisms . . . . . . . . . . . . . . . . . . . 89<br />
7 Summary and conclusions 92<br />
xvii
List of Figures<br />
1.1 Schematic diagram illustrating the basic ion-solid interaction processes. . . . 3<br />
1.2 Schematic diagram illustrating the basic materials modification processes:<br />
(a) implantation; (b) damage due to collision cascades; (c) sputtering; and<br />
(d) atomic mixing across an interface (shown as a dashed line). . . . . . . . 4<br />
1.3 Variation of nuclear and electronic energy losses as a function of incident ion<br />
energy in case of Au in ZnS. All the data were extracted using SRIM code [31]. 5<br />
1.4 Schematic presentation of the ion depth distribution. . . . . . . . . . . . . 7<br />
1.5 Schematic diagram showing multiple vacancy production under a primary<br />
knock-on (an atom which is struck and displaced from its lattice site by a<br />
high energy particle) leading to a displacement spike production. . . . . . . 8<br />
1.6 Time evolution of an ion track due to Coulomb explosion and thermal spike<br />
production by the high energy heavy ion. This figure is adapted from Ref. [46] 10<br />
1.7 Schematic presentation of the shock wave model with two processes: 1 – primary<br />
compressed region, 2 – region of shock wave propagation. The particles<br />
in the shaded region are ejected. . . . . . . . . . . . . . . . . . . . . . . 14<br />
1.8 Schematic presentation of different sputtering scenario for cluster formation<br />
in a p−V-diagram. A target volume is heated up by ion bombardment and<br />
thus attains a high pressure (points 1, 2 and 3). This figure is adapted from<br />
Ref. [23]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />
2.1 Schematic diagram showing the thermal evaporation (resistive heating) experimental<br />
setup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20<br />
2.2 Schematic of the low energy negative ion implanter facility. BPM – beam<br />
profile monitor, FC – Faraday cup, TMP – Turbo-molecular pump. . . . . . 21<br />
2.3 Schematic diagram of the 3.0 MV 9SDH-2 tandem Pelletron accelerator facility<br />
at the Institute of Physics, Bhubaneswar. . . . . . . . . . . . . . . . 22<br />
2.4 Schematic presentation of an elastic collision between an energetic projectile<br />
and a target atom. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />
xviii
2.5 Schematic diagram showing Rutherford backscattering spectrometry experimental<br />
setup. PREAMP – pre-amplifier, AMP – amplifier, ADC – analog<br />
to digital converter, MCA – multi-channel analyzer. . . . . . . . . . . . . . 26<br />
2.6 Schematic diagram illustrating the x-ray diffractometer setup. . . . . . . . . 28<br />
2.7 Block diagram of a transmission electron microscope. . . . . . . . . . . . . 30<br />
2.8 Schematic diagram illustrating an atomic force microscope setup with a scanner<br />
attached with the sample holder. . . . . . . . . . . . . . . . . . . . . . 32<br />
2.9 van der Waals force as a function of tip to sample surface distance. . . . . . 33<br />
2.10 Schematic diagram of the spectrophotometer used for optical absorption<br />
measurements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />
3.1 Variation of energy losses (S n , S e , and S T ) against ZnS depth for 100 MeV<br />
Au ion at normal incidence. Inset: Same for 35 keV and 2 MeV Au ions. All<br />
the energy loss data are extracted using SRIM code [31]. . . . . . . . . . . 40<br />
3.2 RBS spectrum obtained with the as-grown ZnS film on Si(100) substrate.<br />
The Zn, S and Si edges are marked in the figure. Inset shows the RBS<br />
experimental geometry. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41<br />
3.3 The XRD spectra of the as-grown and the Au irradiated ZnS films. The<br />
spectra corresponding to irradiated samples are shifted vertically downward<br />
for clarity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42<br />
3.4 Variation of the strain in as-grown and Au ion irradiated films as a function<br />
of energy deposition in the ZnS surface. . . . . . . . . . . . . . . . . . . . 45<br />
3.5 (a) Plots of (αhν) 2 versus photon energy, hν, for the as-grown and the Au<br />
irradiated ZnS films. Intersections of the straight lines with the abscissa, as<br />
shown by the arrows indicate the band gaps of the films. Inset shows the<br />
experimental OA spectra of the samples. (b) Variation of the band gap as a<br />
function of total energy deposition in the ZnS matrix for different irradiation<br />
energies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />
3.6 2×2 µm 2 three-dimensional AFM images of (a) as-grown, (b) 35 keV, (c) 2<br />
MeV, and (d) 100 MeV Au irradiated ZnS surfaces. Vertical scale is 10 nm<br />
per division. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47<br />
3.7 Variation of the PSD function with the spatial frequency of ZnS surfaces<br />
for as-grown and Au ion irradiated ZnS films. In the high q region, the<br />
experimental data are fitted with a function as given in Eqn. 3.6. The slopes<br />
yield the parameter γ for each case. . . . . . . . . . . . . . . . . . . . . . 49<br />
4.1 Schematic diagram illustrating the sputtering experiment. . . . . . . . . . . 54<br />
xix
4.2 HRTEM micrographs of the NPs corresponding to 2 MeV Au irradiation to<br />
the fluences of (a) 1×10 11 and (b) 5×10 13 ions cm −2 . . . . . . . . . . . . . 54<br />
4.3 SAED patterns of the NPs corresponding to 2 MeV Au irradiation to a<br />
fluence of 5×10 13 ions cm −2 . The diffraction rings corresponding to various<br />
planes are indicated in the figure. . . . . . . . . . . . . . . . . . . . . . . 55<br />
4.4 Bright-field TEM micrographs of ZnS NPs collected on catcher foils. (a),<br />
(b), (c), and (d) correspond to the fluences of 1×10 12 , 5×10 13 , 1×10 14 , and<br />
1×10 15 ions cm −2 with corresponding size distributions shown in (e), (f),<br />
(g), and (h), respectively. The histograms in (e) – (h) have been fitted using<br />
a log-normal distribution function (Eqn. 4.1), which are also plotted in the<br />
figure as solid curves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56<br />
4.5 Variation of (a) the width of the particle size distribution and (b) the catcher<br />
foil surface coverage by the ejected NPs as a function of fluence. . . . . . . 57<br />
4.6 HRTEM micrograph of a NP corresponding to 100 MeV Au irradiation to a<br />
fluence of 1×10 13 ions cm −2 . . . . . . . . . . . . . . . . . . . . . . . . . . 58<br />
4.7 (a) Bright-field planar TEM micrograph of ZnS NPs collected on catcher foil<br />
during 100 MeV Au irradiation at room temperature. (b) shows corresponding<br />
size distribution of the ZnS NPs. . . . . . . . . . . . . . . . . . . . . . 59<br />
4.8 Bright-field TEM micrograph taken on the catcher foil for the 35 keV Au<br />
irradiation to a fluence of 1×10 14 ions cm −2 . . . . . . . . . . . . . . . . . 59<br />
4.9 Measured number of collected particles (having diameter ≥3 nm) as a function<br />
of hemisphere volume. Experimental data are fitted, with a function<br />
given in Eqn. 4.2, and shown in the figure as a straight line for each set of<br />
data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />
4.10 Snapshots from a simulation of sputtering of Au at different times for 20 keV<br />
Xe incident on Au. The labels A and B show clusters that are fragmenting,<br />
and C illustrates late sputtering (a cluster separating from the surface after<br />
the displacement cascade has ended). This figure is adapted from Ref. [21]. . 61<br />
4.11 SRIM simulated collision cascades as a function of depth for (a) 35 keV and<br />
(b) 2 MeV Au in ZnS at an incident angle of 45 ◦ with the target normal. . . 64<br />
5.1 Bright-filed XTEM micrographs of the Au NPs in silica glass after 32 keV<br />
Au − ions impantation to a fluence of 4×10 16 cm −2 (a) before and (b) after<br />
annealing in air at 850 ◦ C for 1 h. (c) and (d) show the corresponding size<br />
distributions, respectively. Inset in (a) shows an HRTEM image of a Au NP<br />
indicating the particles are crystalline in nature. . . . . . . . . . . . . . . . 70<br />
xx
5.2 OA spectrum of Au implanted and annealed silica glass samples. . . . . . . 71<br />
5.3 (a)−(d) Bright-filed XTEM micrographs of the as-grown and 100 MeV Au<br />
irradiated samples with fluences of 2×10 13 , 5×10 13 , and 1×10 14 ions cm −2 ,<br />
respectively; (e)−(h) show the corresponding size distributions, respectively.<br />
Note that (b) and (c) are having same scale as in (a). Dashed lines are<br />
indicating the surfaces. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />
5.4 Bright-field planar TEM micrograph of the 100 MeV Au irradiated sample<br />
for the fluence of 1×10 14 ions cm −2 . . . . . . . . . . . . . . . . . . . . . . 73<br />
5.5 (a) RBS spectra taken from the as-grown samples before and after SHI irradiations.<br />
The Au and Si RBS profiles are indicated in the figure. (b) Au<br />
sputtered from the samples after SHI irradiation with different ion fluences. 74<br />
5.6 OA spectra taken on the (a) as-grown and SHI irradiated samples corresponding<br />
to the fluences of (b) 2×10 13 , (c) 5×10 13 , and (d) 1×10 14 ions<br />
cm −2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />
5.7 OA spectrum (filled circles) of the as-grown sample after 2×10 13 ions cm −2<br />
irradiation along with the fitted (solid line) curve using two Gaussian functions<br />
with proper background substraction (dashed lines). . . . . . . . . . . 76<br />
5.8 Dependence of the absorption maximum of the longitudinal plasmon resonance<br />
against the average aspect ratio as determined by XTEM. The solid<br />
line is the linear fit to the experimental data, except the 1×10 14 ions cm −2<br />
irradiated one. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77<br />
5.9 Simulated OA spectra of the Au NPs embedded in silica matrix for different<br />
NP sizes (d) and medium dielectric constants (ǫ m ). Inset shows the variation<br />
of SPR peak position with dielectric constant (filled circles). The solid line<br />
is the linear fit to the simulated data. . . . . . . . . . . . . . . . . . . . . 80<br />
6.1 Bright-field XTEM micrographs of the Au NPs in silica glass after 32 keV<br />
Au − ions implantation to a fluence of 4×10 16 cm −2 followed by annealing<br />
in air at 850 ◦ for 1 h. (b) show the corresponding size distribution with a<br />
Gaussian fitting. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84<br />
6.2 Schematic diagram illustrating the sputtering experiment. . . . . . . . . . . 85<br />
6.3 Bright-field planer TEM micrographs of the ejected Au particles on the<br />
catcher foils collected during ion irradiation with irradiation fluences of (a)<br />
2×10 13 , (b) 5×10 13 , and (c) 1×10 14 ions cm −2 , respectively. (d) Coverage<br />
of the catcher foils surface by the ejected NPs. Solid line is the linear fit to<br />
the experimentally obtained data. . . . . . . . . . . . . . . . . . . . . . . 86<br />
xxi
6.4 (a) HRTEM micrograph of a NP and (b) SAED pattern of the NPs collected<br />
on the catcher foil corresponding to a Au irradiation fluence of 1×10 14 ions<br />
cm −2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />
6.5 The Au part of the RBS spectra as measured on the targets before and after<br />
the irradiations. The irradiation fluences are in the unit of ions cm −2 . Inset<br />
shows the experimental arrangement. . . . . . . . . . . . . . . . . . . . . 88<br />
6.6 Size distribution in terms of the number of clusters of particular size (cluster<br />
yield) plotted against the number of atoms in the cluster. Dotted, dotdashed,<br />
and continuous lines correspond to power law decays with δ values<br />
of 3/2, 7/3, and 7/2, respectively. . . . . . . . . . . . . . . . . . . . . . . 88<br />
xxii
Chapter 1<br />
Introduction<br />
1.1 Interests and applications of ion beams in materials<br />
science<br />
A major part of the modern technology depends on materials with precise controlled<br />
properties. Ion implantation is a favoured method to achieve controlled modification<br />
of surface and near-surface regions of various materials. In fact, ion beams have proven<br />
to be versatile tools for processing materials and synthesizing unique structures for<br />
electronic [1], optoelectronic [2], and magnetic [3] device applications. Introduction of<br />
ion species in the near surface of solids can be used to alter and tailor many material<br />
properties. For instance, He implantations, with energy in the range of 10 to 180 keV,<br />
into Ni has been found to result in an increase in hardness as much as approximately<br />
seven times than that of unirradiated Ni [4]. Implantation of 250 keV Ge + ions into<br />
SiC has been seen to result in a decrease in resistivity with increase in Ge fluence [5].<br />
Co implantation at 30 keV into thin Fe 70 Co 8 Si 12 B 10 films deposited on SiO 2 has been<br />
found to show a stress-induced magnetic anisotropy in the ferromagnetic thin film [6].<br />
It is possible to tune the band-gap of semiconductor materials by using ion implantation.<br />
For example, 18 keV P + implantation into InAs/InP heterostructures has been<br />
shown to result in a shift in photoluminescence peak towards higher wavelength. This<br />
shift in peak position increases with increase in ion fluence [7]. C implantation into<br />
Si(111) at 1 MeV energy has been found to result in enhanced gettering of Au [8].<br />
Recently, using ion implantation, a new route for synthesis of nanostructure materials<br />
inside a host has been reported [9]. For example, 32 keV Au − ions implantation to<br />
a fluence of 8×10 16 cm −2 into silica glass has been found to result in Au nanocluster<br />
formation with an average size of about 6 nm [9].<br />
Defect generation during ion-solid interactions adds a further dimension to this<br />
1
Introduction 2<br />
processing capability enabling kinetic mechanisms to synthesize metastable phases<br />
normally unobtainable under equilibrium thermodynamic conditions. For example,<br />
mixing of thermally immiscible layers of Fe/Ag using ion irradiation has been reported<br />
[10]. In this case, 85 keV Ar + ions were irradiated in multilayers of Fe/Ag in the<br />
fluence range of 5×10 14 – 3×10 16 ions cm −2 . It was seen that the mixing increased<br />
with increase in fluence. In certain cases, involving both metals and insulating systems,<br />
irradiation with high energy heavy ions has been found to result in the formation of ion<br />
tracks, a few nm in diameter and few µm in length [11]. These can be used as templates<br />
for growth of nanoparticles (NPs) inside them. In addition, anisotropic deformation<br />
of Ag NPs, embedden in silica, has also been observed by 8 MeV Si irradiation [12].<br />
The anisotropy in shape results in change in optical properties. Also energetic ion<br />
irradiation of solids leads to the ejection of materials from the surface of the target.<br />
This process of removal of materials, either atoms or chunks of atoms (clusters), is<br />
known as sputtering [13]. Sputtering is used for cleaning solid surfaces [13], coating<br />
thin films (where the sputtered material is deposited on another body) [14], etc. It is<br />
also used in the analysis of materials using secondary ion mass spectroscopy [15].<br />
Very recently, sputtering of metal as well as alloy metal NPs have been observed<br />
[16]. This observation of stable, intact clusters of atoms and molecules, due to ion<br />
impact, is rather surprising because the energies involved in the ion-target collisions,<br />
that generate sputtered atoms, are typically much larger than the binding energies (∼<br />
1 – 2 eV) holding the clusters together [17]. Hence it is difficult to understand how<br />
stable clusters can form during ion sputtering. Eversince the first observation, the<br />
fundamental question of the cluster emission mechanism has attracted considerable<br />
attention both experimentally [13, 16, 18, 19] and theoretically [20, 21, 22, 23].<br />
Due to all the above interests, during the last several decades, considerable attention<br />
has been paid to ion-solid interactions, leading to a detailed understanding<br />
of the important processes, such as ion penetration, energy loss, radiation damage,<br />
sputtering of clusters, grain growth in thin films, and irradiation-induced mixing.<br />
1.2 Ion-solid interactions<br />
1.2.1 Basic concepts<br />
When a beam of energetic charged particles is directed onto the surface of a solid,<br />
the fast projectiles interact with the stationary near-surface atoms of the solid. These
Introduction 3<br />
interactions can lead to ion penetration into the subsurface region of the solid, rearrangement<br />
of the near-surface target atoms, and the emission of particles and other<br />
forms of radiation from the surface [24]. When an ion penetrates a solid, it undergoes<br />
a succession of binary collisions with target atoms and surrounding electrons, losing<br />
energy at each encounter. The incident ion may conceivably be reflected back out of<br />
the surface during a single “hard” collision or from many correlated collisions. However,<br />
for ions of incident energy above a few keV, it is most probable that the ion will<br />
finally come to rest within the solid. The transfer of energy from projectile ion to<br />
the solid can be conveniently divided into two independent processes, namely nuclear<br />
collisions and electronic collisions [25, 26]. The former process constitutes collisions<br />
between the incident ion and lattice atoms where conservation of energy and momentum<br />
apply (so-called elastic collisions). The latter electronic process depicts the<br />
interaction of fast ions with lattice electrons (so-called inelastic collisions). Electronic<br />
collisions result in ionization and excitation processes, which can lead to emission of<br />
radiation in the form of characteristic x-rays, optical photons, and Auger or secondary<br />
electrons. In Fig. 1.1, the types of emission process which can result from both nuclear<br />
and electronic collisions are illustrated schematically.<br />
Ion beam<br />
Nuclear processes<br />
Electronic processes<br />
Backsc−<br />
attered<br />
atoms<br />
Nuclear<br />
reaction<br />
products<br />
Sputtered<br />
target<br />
atoms<br />
Photons<br />
Phonons<br />
Electrons<br />
Figure 1.1: Schematic<br />
diagram illustrating the<br />
basic ion-solid interaction<br />
processes.<br />
Solid<br />
Electronic<br />
and Nuclear<br />
energy−loss<br />
processes
Introduction 4<br />
Many<br />
ions<br />
(a)<br />
Single ion<br />
(b)<br />
Modified composition<br />
(c)<br />
Sputtering<br />
Atomic<br />
displacement<br />
Surface<br />
erosion<br />
Collision<br />
cascade<br />
Figure 1.2: Schematic diagram illustrating<br />
the basic materials modification<br />
processes: (a) implantation;<br />
(b) damage due to collision cascades;<br />
(c) sputtering; and (d) atomic mixing<br />
across an interface (shown as a dashed<br />
line).<br />
(d)<br />
Thin film<br />
Substrate<br />
Figure 1.2 (a) shows the ion implantation process, which can lead to the build up<br />
of a concentration profile of foreign atoms within a solid, thus altering the near-surface<br />
composition [27]. The distribution of foreign atoms depends on the ion energy and the<br />
stopping (nuclear and electronic) processes which determine how individual ions slow<br />
down and where they ultimately come to rest. Figure 1.2 (b) illustrates that “hard”<br />
nuclear collisions can result in displacement of lattice atoms from their regular sites. A<br />
single heavy ion can lead to displacement of several tens, and even hundreds, of lattice<br />
atoms within a volume surrounding the ion trajectory, termed a collision or displacement<br />
cascade. As a consequence, ion bombardment can create considerable structural<br />
damage to the material [28]. Also atoms at or near the surface may receive sufficient<br />
energy to surmount the surface potential barrier and leave the solid. This sputtering<br />
process [29], results in a flux of ejected atomic and/or molecular species, which can
Introduction 5<br />
be neutral or in various charge states [13]. Schematic illustration of this process is<br />
given in Fig. 1.2 (c) and discussed in details in Section 1.3. Finally, in Fig. 1.2 (d) we<br />
illustrate the process of atomic mixing in which solid atoms can be transported across<br />
an interface within dimensions of the collision cascade at temperatures below those<br />
at which normal diffusion processes would operate. One consequence of this effect is<br />
shown for a thin film of one material on a substrate of another: ion bombardment can<br />
produce appreciable intermixing with the substrate material.<br />
1.2.2 Ion penetration and stopping<br />
Energy loss processes<br />
As an ion penetrates a solid, it loses energy due to two distinct interactions [25,<br />
30]. At lower energies, it loses energy by elastic, nuclear collisions with the target<br />
atoms (nuclear stopping) described by the well-known Rutherford cross sections for<br />
the Coulomb interaction between the screened charges of the atomic nuclei. At higher<br />
energies, it loses energy inelastically by exciting and ionizing electrons of target atoms<br />
(electronic stopping). In this process, incident ion transfers its energy to the lattice<br />
electrons inelastically. It is expected that the lattice atoms are not displaced in the case<br />
of electronic excitation. At low ion velocities, where the nuclear energy-loss dominates,<br />
the energy loss process is accurately described by the electrostatic interaction between<br />
the projectile and the target atoms using a Thomas-Fermi potential [25, 26]. At<br />
higher ion velocities, the electronic energy-loss processes are better described by the<br />
Bethe-Bloch formalism [25, 26]. Both stopping processes strongly depend on the<br />
kinetic energy of the ion. Figure 1.3 shows a graph of the nuclear and electronic<br />
Energy Loss [keV/nm]<br />
30<br />
25<br />
20<br />
15<br />
10<br />
5<br />
Nuclear energy loss<br />
Electronic energy loss<br />
Figure 1.3: Variation of nuclear and<br />
electronic energy losses as a function<br />
of incident ion energy in case of Au in<br />
ZnS. All the data were extracted using<br />
SRIM code [31].<br />
0<br />
10 0 10 1 10 2 10 3 10 4 10 5 10 6 10 7 10 8<br />
Energy [keV]
Introduction 6<br />
energy losses as a function of ion energy for Au irradiation of a ZnS target. As<br />
can be clearly seen in Fig. 1.3, the stopping for Au in ZnS matrix is dominated by<br />
the nuclear energy loss for ion energies smaller than 1 MeV. In this region, due to<br />
collisions with the incident ions, target atoms are displaced from their original positions<br />
[32], leading to subsequent nuclear collision cascades. This can lead to structural<br />
changes, such as the generation of point defects and the amorphization of crystalline<br />
materials (and vice-versa). This happens when the Rutherford scattering cross section<br />
is highest, i.e., for heavy incident ions at energies of typically 10−100 keV. On the<br />
other hand, for higher ion energies (> 4 MeV in Fig. 1.3) energy loss is dominated by<br />
electronic stopping, resulting in a rise in local lattice temperatures, as described above.<br />
Depending upon the type of material and ion energy, this can lead to the formation of<br />
a cylindrically-shaped molten region around the ion track, with typical diameter in the<br />
range of few nm [33]. A fast cooling of the molten region can, therefore, bring about<br />
structural changes in the target material, such as amorphization, defect generation,<br />
and structural transformations. One prominent effect of electronic stopping processes<br />
is anisotropic deformation [12], a macroscopic materials shape change that is the result<br />
of microscopic processes occurring in the thermal spike of individual ion impacts.<br />
Ion Range and Range distributions<br />
An ion traversing a solid gradually loses its energy in every monolayer or so following<br />
the above energy loss processes, finally coming to rest at a depth below the surface.<br />
This is knows as “ion implantation” as mentioned in Section 1.2.1. Ion implantation<br />
is widely used in materials engineering for the introduction of atoms into surface<br />
layer of a solid substrate by bombardment of the solid with ions of desired species<br />
in the keV to MeV energy range. Use of the implantation technique affords the<br />
possibility of introducing a wide range of atomic species, thus making it possible to<br />
obtain impurity concentrations and distributions of particular interest; in many cases,<br />
these distributions would not be otherwise attainable [34].<br />
One of the most important parameters of implantation process is the “range” of<br />
the implanted ions which correspond to the total distance traveled by an ion before<br />
coming into rest as given by [25]<br />
R =<br />
∫ E0<br />
0<br />
[( ) dE<br />
+<br />
dx<br />
nucl.<br />
( ) ] −1<br />
dE<br />
dE, (1.1)<br />
dx<br />
elec.<br />
where E 0 is the incident ion energy at the target surface, x is the distance measured
Introduction 7<br />
along the ion path, ( )<br />
dE<br />
and ( )<br />
dE<br />
are the nuclear and electronic energy losses,<br />
dx nucl. dx elec.<br />
respectively, of the ion at energy E. Since energy is lost by the projectile to the target<br />
atoms and electrons in a series of discrete collisions, the energy loss per collision and<br />
hence the total path length will have a statistical spread of values. This leads to a<br />
near-Gaussian distribution of stopping distances. The mean length of projection of<br />
the ion path on its incident direction is known as the projected range, R p . This is<br />
schematically shown in Fig. 1.4. In ion implantation studies one is actually interested<br />
Z<br />
log scale<br />
ion<br />
distribution<br />
∆Rp<br />
Ion beam<br />
Rp<br />
∆Rt<br />
X<br />
Y<br />
R<br />
Figure 1.4: Schematic presentation of the ion depth distribution.<br />
in the entire depth distribution of the implanted atoms and, therefore, a knowledge of<br />
the range straggling (fluctuation) is also necessary. For ions of moderate energies, in<br />
the absence of channeling, the range distribution is approximately Gaussian in shape<br />
and can be described by a single extra parameter namely the fluctuation (standard<br />
deviation) ∆R p , in the projected range along the direction of R p . ∆R p is called the<br />
longitudinal range straggling and the fluctuation along the direction perpendicular<br />
to this is known as transverse range straggling, ∆R t . A schematic diagram showing<br />
range, R, projected range, R p , and range stragglings (∆R p and ∆R t ) of the implanted<br />
atoms is shown in Fig. 1.4. These parameters of ion implantation are important for<br />
device modeling and design, as well as for the process optimization and control in<br />
device fabrication. A large amount of experimental and theoretical work has been
Introduction 8<br />
devoted to the task of understanding the energy-loss processes that govern the range<br />
distribution, and it is now possible to predict fairly accurately most of the factors<br />
involved in the ion implantation process [25, 35].<br />
1.2.3 Nuclear thermal spike<br />
In the elastic collision regime, with few keV energy, an ion upon collision with a lattice<br />
atom can result in displacement of the lattice atom. The minimum energy imparted<br />
to the lattice atoms as required for this is known as the displacement threshold which<br />
is about 10 – 50 eV [34]. At much higher energy transfers the displaced atoms can lead<br />
to collision cascades leading to a larger number of displaced atoms. Multiple collisions<br />
between recoiling atoms can be described by linear transport theory as independent<br />
binary collision events as long as the density of recoiling atoms is low. However,<br />
when the projectile mass and/or the atomic mass of the substrate are large, these<br />
assumptions fail as nonlinear processes occur [36]. When the mean free path between<br />
collisions is of the order of the atomic spacing, a highly disturbed volume known as<br />
an atomic displacement spike or energy spike [Fig. 1.5] is created in less than 10 −13 s.<br />
Figure 1.5: Schematic diagram showing<br />
multiple vacancy production under<br />
a primary knock-on (an atom which<br />
is struck and displaced from its lattice<br />
site by a high energy particle) leading<br />
to a displacement spike production.<br />
This is followed by a thermal spike phase (which lasts for about 10 −11 s) of a small<br />
volume in which the energy density may be several eV/atom. Here the temperature<br />
goes well above the melting temperature and in some cases going even beyond the<br />
vaporization temperature of the material. This is known as elastic or nuclear thermal<br />
spike. In addition to this there is also a corresponding pressure spike. Molecular<br />
dynamics (MD) simulations, using 10 keV Au on Au(001) surface do indicate such a<br />
scenario associated with cascade induced melting and viscous flow [37].
Introduction 9<br />
1.2.4 Coulomb explosion and electronic thermal spike<br />
However, the scenario become completely different when a high energy heavy ion (swift<br />
heavy ion) passes through a material. The energy loss in such a situation (involving inelastic<br />
interactions with target electrons) is given by the electronic stopping power S e .<br />
After the initial energy-transfer by high energy ions, electrons escape the central region<br />
of the ion path. Depending on the ionization density and on the charge-neutralization<br />
time, the mutual repulsion of positively charged target ions may convert a significant<br />
amount of the stored electronic potential energy into atomic motion. This is known<br />
as the Coulomb explosion [38, 39, 40, 41]. Coulomb explosion is significant only if the<br />
charge-neutralization time exceeds 10 −14 s for light target atoms and 10 −13 s for heavy<br />
atoms.<br />
In such a case the initially excited and ionized electrons transfer their energy to<br />
the lattice atoms (in about 10 −12 s) by electron-phonon coupling. This would induce<br />
a local rise in lattice temperature [42]. In this scenario, the electron and atom system<br />
are not in thermal equilibrium with each other. In such a case, the space and time<br />
evolutions of the electronic and lattice temperatures, T e and T, are governed by a set<br />
of coupled nonlinear differential equations [43] (in cylindrical geometry) as given by<br />
and<br />
ρC e (T e ) δT e<br />
δt = δ [<br />
K e (T e ) δT ]<br />
e<br />
δr δr<br />
ρC(T) δT<br />
δt = δ [<br />
K(T) δT ]<br />
δr δr<br />
+ K e(T e ) δT e<br />
r δr − g(T e − T) + A(r), (1.2)<br />
+ K(T) δT<br />
r δr − g(T e − T), (1.3)<br />
where C e , C and K e , K are the specific heat and thermal conductivity for the electronic<br />
system and lattice, respectively, ρ is the materials density, g is the electron-phonon<br />
coupling, A(r) is the energy brought on the electronic system in a time considerably<br />
less than the electronic thermalization time, and r is the radius in cylindrical geometry<br />
with the heavy ion path as axis. This description of rise of local lattice temperature<br />
and its dissipation to surround medium is well known as inelastic or electronic thermal<br />
spike model [43].<br />
As a result of Coulomb explosion and inelastic thermal spike production, there can<br />
be formation of amorphized latent tracks (of few nm in diameter and several µm in<br />
length) [33, 44] coming from sudden cooling of the molten zone. Using MD simulations,<br />
Bringa and Johnson have shown that the Coulomb explosion and the thermal spikes<br />
correspond to the early and late stages of the ionization track produced in a solid by a<br />
fast incident ion [45]. Figure 1.6 displays a schematic view of the time dependence of
Introduction 10<br />
Figure 1.6: Time evolution of an<br />
ion track due to Coulomb explosion<br />
and thermal spike production by the<br />
high energy heavy ion. This figure is<br />
adapted from Ref. [46]<br />
the ion-track evolution. The upper part shows the rapidly passing projectile (dashed<br />
arrow). Once the projectile has reached its equilibrium charge state, there will be only<br />
minor fluctuations of its internal state and it will move with constant velocity along<br />
a straight-line trajectory until deep inside the solid. Thus, the projectile ion acts<br />
as a well defined and virtually instantaneous source of strongly localized electronic<br />
excitation. According to Bethe’s equipartition rule, about 50 % of the total electronic<br />
energy is deposited inside the so-called intra-track radius of about 1 nm around the<br />
projectile path at projectile energies of a few MeV per nucleon [46]. Excitation times<br />
are 10 −19 to 10 −17 s for inner-shell processes and reach 10 −16 s for collective electronic<br />
excitations (plasmon production). The lattice atoms, displaced following thermal spike<br />
production, freeze out and may lead to permanent rearrangements. In the bulk, this<br />
may lead to structural or chemical modifications. At the surface craters or blisters on<br />
an atomic scale can be produced. These are shown schematically in the last part of<br />
Fig. 1.6.
Introduction 11<br />
1.3 Sputtering<br />
In 1852 W. R. Grove, studying electric discharges between metal electrodes in various<br />
gas atmospheres [47], observed that material was removed from the negative electrode<br />
(the cathode) and got deposited on the glass walls containing the electrodes and the<br />
gas mixture. The above phenomenon corresponds to the simplest erosion process of<br />
a surface called sputtering [13]. In the sputtering process, atoms, molecules, and<br />
sometimes clusters are ejected from the outer surface layers of the target. Sputtering<br />
is quantitatively described by the sputtering yield, Y , which is defined as the average<br />
number of sputtered target atoms per incident ion, i.e.,<br />
Number of sputtered atoms<br />
Y = (1.4)<br />
Number of incident ions<br />
The sputtering yield, Y , depends on the ion incident angle, the energy of the ion, the<br />
masses of the ion and target atoms, the surface topography, and the surface binding<br />
energy of atoms in the target. For a crystalline target the orientation of the crystal<br />
axes with respect to the target surface is relevant.<br />
Depending upon the ion-solid interactions, there can be various types of sputtering<br />
processes. Among these, physical sputtering is well studied in the materials science.<br />
This type of sputtering is driven by momentum exchange between the energetic ion<br />
and atoms in the material, due to collisions [13]. This type of sputtering takes place<br />
only when the surface atoms of the substrate receive from the incident ion an energy<br />
which is greater than the surface binding energy. In certain cases a chemical reaction<br />
between the ion and the target atoms may result in lowering of the binding energy<br />
resulting in enhancement in sputtering yield [48].<br />
Apart from these, there are another two types of sputtering, viz. electronic sputtering<br />
and potential sputtering. The term electronic sputtering means either sputtering<br />
induced by energetic electrons, or sputtering due to very high-energy or highly charged<br />
heavy ions which lose energy to the solid mostly by electronic stopping processes. In<br />
the last case electronic excitations cause sputtering [49]. On the other hand, in the<br />
case of multiply charged projectile ions, a particular form of electronic sputtering can<br />
take place which has been termed potential sputtering [50]. In this case a low energy<br />
ion in a high charge state can interact with a surface forming a hollow atom with<br />
empty inner orbitals. Impact of a hollow atom on the surface, in some cases, can<br />
lead to sputtering of target atoms through the release of the potential energy stored.<br />
In the present case we will confine ourselves to sputtering from nuclear or electronic<br />
collisions as observed at low and high ion energies, respectively.
Introduction 12<br />
1.3.1 Sputtering mechanisms<br />
Sputtering from elastic collision<br />
For some time, the sputtering process was thought of as an evaporation process which<br />
implied that sputtering was due to very high local temperatures created at or very<br />
near the target surface by the bombarding ions. Later, however, measurements on<br />
the dependence of sputtering yield on the angle of incidence of the bombarding ions<br />
established that sputtering is really a momentum transfer process [29, 51]. There have<br />
been a number of theories and models to explain the main features of the sputtering<br />
process, observed at lower ion energy, but the most widely accepted one is that developed<br />
by Sigmund [29]. The underlying concepts and the results of the Sigmund’s<br />
theory is discussed briefly below.<br />
As mentioned earlier, in case of low energy (keV), nuclear energy loss dominates. In<br />
such a case, incident ion transfers its energy to the target atoms leading to generation<br />
of collision cascades. Assuming that the cascades are produced as a result of binary<br />
collisions, for particles incident normal to a substrate surface, the sputtering yield<br />
at a given energy, E 0 , is described by a linear function of nuclear energy loss. This<br />
sputtering yield, Y (E 0 ), can be written as<br />
Y (E 0 ) = ΛF D (E 0 ), (1.5)<br />
where F D is the energy deposition per unit length due to nuclear processes at the<br />
surface, and Λ is a term which depends on the properties of target and the state of<br />
the surface. Λ is given by<br />
Λ = 3 1<br />
4π 2 NC 0 U 0<br />
where N is the target density, U 0 is the surface binding energy, and C 0 is a constant<br />
defined as<br />
C 0 = 1 ( ) (<br />
2 πλ 0a 2 M1 2Z 1 Z 2 e 2 )<br />
M 2 a<br />
Here M 1 and M 2 are the ion and target atom masses respectively, Z 1 and Z 2 being<br />
the ion and target atomic numbers, and a is the screening radius.<br />
Electronic sputtering<br />
Sigmund sputtering theory was quite successful in explaining the sputtering yields<br />
for single elemental targets for various ion-target combinations. However, this theory<br />
completely fails to explain enhanced sputtering as observed in certain cases as well
Introduction 13<br />
as the sputtering of large clusters of atoms and molecules where nonlinear effects<br />
of energy deposition are required to be included to explain the observed data [16,<br />
18, 19, 20, 21]. As one enters the realm of nonlinear energy transport processes,<br />
the mechanisms leading to sputtering become less obvious and, consequently, less<br />
amenable to theoretical treatment discussed above. In spite of that, several nonlinear<br />
models have been proposed [52]. Among them, the thermal spike theory is most<br />
often used in this regime [53, 54]. This theory assumes that the moving atoms in<br />
the target achieve, quickly, a state of local thermal equilibrium. Therefore, if the<br />
density is assumed to remain constant, the local temperature suffices to determine the<br />
thermodynamical state of the system. The deposited energy then relaxes according to<br />
the heat conduction equation (Eqns. 1.2 and 1.3) and the sputtering can be readily<br />
obtained as the flux of atoms evaporated from the hot surface. One testable outcome<br />
of this theory is that, under quite general conditions, for large dE/dX, the sputtering<br />
yield for a cylindrical excitation geometry is a quadratic function of the deposited<br />
energy, i.e. Y ∼ (dE/dX) 2 [55]. Such a quadratic behaviour of Y was in fact observed<br />
in many experiments, and it appeared to be so firmly established that a larger-thanlinear<br />
sputtering yield often is sufficient to conclude that a thermal spike occurred.<br />
1.3.2 Size distribution of clusters ejected during ion irradiation<br />
Recently, several researchers have observed the ejection of large clusters in the sputtered<br />
species during ion bombardment [16, 17, 18, 19, 20, 21]. Most of these studies,<br />
carried out at keV energies, show a monotonically decreasing yield distribution, which<br />
closely follows an inverse power law decay with increasing cluster size.<br />
In the last several decades, a lot of efforts has been made to understand cluster<br />
ejection mechanisms, resulting from energetic ion bombardment. In almost all cases<br />
the size distributions have been found to follow power law distributions with different<br />
decay exponents [16, 17, 18, 19]. With low energy ions, for small clusters, the decay<br />
exponents have been found to lie between 4 to 8 [19]. However, at higher energy it<br />
seems to be close to 2 which has been argued as coming due to shock wave induced<br />
emissions [16]. MD simulations [20, 21] do indicate cluster emissions however, with<br />
some breakup during the flight. On the theoretical front there are two models viz.<br />
the one based on generation of shock waves [22] and a thermodynamic liquid-gas type<br />
phase transition model [23] for the clustering of sputtered material. Both these models
Introduction 14<br />
predict decay exponents close to 2 and there are no results, available in the literature,<br />
to distinguish between the two. There are also non-equilibrium aggregation processes<br />
that can play a role in determining the cluster size distribution. In these cases, there<br />
is a competition between aggregation and breakup leading to entirely different size<br />
distributions [56]. Such processes are expected to be present in sputtered material<br />
coming from thermal spikes induced by very high energy ions. There have been no<br />
data on this as well. Some basic features of the above mentioned models are given<br />
below.<br />
Shock wave model<br />
One of the most successful model for cluster ejection is shock wave model developed<br />
by Bitensky and Parilis [22]. For low energy heavy ion irradiation on materials, it is<br />
plausible that the mean free displacement collision paths of cascading atoms become<br />
of the order of the atomic spacing or less than it. In this situation, the nonlinear interactions<br />
become significant and the available kinetic energy may be instantaneously<br />
released locally. If this is the case, recoiling atoms in the region may be energetic<br />
enough to form a fire ball locally. If the average velocity of these atoms exceeds the<br />
sound velocity in the target medium, atoms of the cascade may play the role of a<br />
“hammer”compressing the surrounding medium towards the outer direction following<br />
a spherical geometry (Fig. 1.7). In case of high energy heavy ion irradiation, particu-<br />
Projectiles<br />
Solid<br />
Casade process<br />
1<br />
Figure 1.7: Schematic presentation of<br />
the shock wave model with two processes:<br />
1 – primary compressed region,<br />
2 – region of shock wave propagation.<br />
The particles in the shaded region are<br />
ejected.<br />
2<br />
larly of insulators, the accumulated energy is initially the potential energy of Coulomb<br />
repulsion of the ionized atoms. In a non-metallic solid the time for its neutralization is<br />
rather long and after 10 −12 s, this potential energy is transferred through a Coulomb
Introduction 15<br />
explosion into kinetic energy which is spent in track creation, formation of high density<br />
collision cascades and shock wave creation. In this case, the shock wave propagates<br />
perpendicular to the cylindrical ion track. The necessary condition for shock wave<br />
generation is the achievement of a critical mean kinetic energy of atoms ǫ c , which is<br />
the free parameter of the theory [22].<br />
Once a shock wave is formed, it propagates towards the surface. Upon reaching<br />
the surface it gets reflected. When the pressure at this surface exceeds a critical value;<br />
a part of the surface is unloaded (shaded region in the Fig. 1.7). The result of this<br />
unloading gives the sputtering yield in the shock wave model. Ejection of clusters<br />
following this model leads to an inverse power-law size distribution [22] i.e.,<br />
Y (n) ∼ n −δ , (1.6)<br />
where, n is the number of atoms in a cluster and δ is the exponent with a value of<br />
5/3 or 7/3. The first exponent corresponds to the sputtering of clusters due to the<br />
formation of collision cascades by single ion, whereas the second exponent corresponds<br />
to the sputtering of clusters due to the formation of collision cascades by the impacts<br />
of multiple ions.<br />
Shock waves production has also been seen in MD simulations [37] involving the<br />
impact of 10 keV Au on Au(001) surface. However, in that case there is a viscous flow<br />
of molten Au atoms coming from the pressure spike generated. Cluster emission data<br />
[16] from Au thin films, irradiated using four different ions (Ar, Ne, Kr, and Au) with<br />
energies in the range of 400 to 500 keV, have been seen to follow the size distribution<br />
as given in Eqn. 1.6 with δ values as given by the shock wave model of Bitensky and<br />
Parilis [22].<br />
Equilibrium liquid-gas type phase transition model<br />
Another important cluster ejection model is the thermodynamic equilibrium liquidgas<br />
type phase transition model proposed by Urbassek [23]. Main assumption of<br />
this model is that after the high energy atoms in a collision cascade, produced by<br />
energetic ion impact, are emitted from the near surface regions of the targets, the<br />
energized cascade volume will begin to equilibrate thermally. We depict this in the<br />
p−V diagram of Fig. 1.8 as a vertical transition, assuming that collisional sputtering<br />
did not substantially alter the relevant volume. The high pressure that builds up<br />
leads to an expansion of the heated volume into the vacuum, and its further evolution<br />
is essentially controlled by the temperature: If the temperature T is greater than a
Introduction 16<br />
Figure 1.8: Schematic presentation of<br />
different sputtering scenario for cluster<br />
formation in a p−V-diagram. A target<br />
volume is heated up by ion bombardment<br />
and thus attains a high pressure<br />
(points 1, 2 and 3). This figure is<br />
adapted from Ref. [23].<br />
critical temperature T c , the heated material gasifies and may flow out of the target.<br />
Moreover, in the course of its development, the system will cool down, and may enter<br />
the coexistence region. Now it is decisive, whether it enters at the large volume side<br />
(V > V c ) or at the low volume side (V < V c ). This is obviously to a large part<br />
determined by T c of the system. In the first case, the gasified system may form<br />
clusters by nucleation from the gas phase (agglomeration). In the latter case, the<br />
liquid system may break up due to the expansion and form a vapour (fragmentation),<br />
as a result of liquid-gas phase transition, leading to clusters ejection. This equilibrium<br />
model predicts transitions from a exponential decay to power law decay as the phase<br />
transition occurs closer to the critical point. The dependence of the cluster yield Y (n)<br />
on the cluster size, n, is given by<br />
Y (n) = Y 0 n −δ exp[(−∆Gn − 4πn 2/3 r 2 σ)/kT], (1.7)<br />
where ∆G is the difference of the Gibbs free energies of the liquid and gas phase, k is<br />
Boltzmann’s constant, T is the temperature of the energized region, Y 0 is the sputter<br />
yield (a constant for a given projectile-target system), r is the cluster radius, σ is the<br />
surface tension, and δ is the exponent. At equilibrium ∆G is zero and at the critical<br />
point the surface tension vanishes. It follows that predicted cluster size distributions<br />
are very similar for the shock wave model. This model also predicts a power-law size<br />
distribution of the sputtered particles with a δ value in between 2 and 2.5. For a van<br />
der Waals gas the exponent, δ, has a value of 7/3.
Introduction 17<br />
Nonequilibrium aggergation mechanisms<br />
Another one of the important cluster formation model is nonequilibrium aggregation<br />
model [56, 57, 58, 59] which was not studied earlier in the sputtering process. In<br />
this model, there is a competition between aggregation and breakup or evaporation, a<br />
delicate balance between the two leading to a variety of steady state mass distributions<br />
[56]. The nonequilibrium aggregation model was considered to be in a discretized space<br />
and time [57]. That is, particles with integer mass are placed on each site of the twodimensional<br />
lattice, and they aggregate by random-walk process with discrete time<br />
steps. At the beginning of each time step, there are particles on every site of the<br />
lattice. All of them independently jump to randomly chosen sites according to a given<br />
probability. If two or more particles come together at a site after the jump, they are<br />
combine to form a new particle with a conserved mass. Then a particle with unit<br />
mass is added to every site. Thus the evolution of one time step is completed and the<br />
process of aggregation without considering any fragmentation has been repeated. This<br />
model then shows that the asymptotic distribution of mass always follows a power law<br />
behaviour. Later the analysis was extended to include both positive and negative<br />
values for the dynamical variable which was taken to be charge rather than mass [58].<br />
In this scenario, if m(j, n) be the charge of the particle on site j at the nth time step<br />
then the aggregation process can be represented by the following stochastic equation<br />
for m(j, n):<br />
m(j, n + 1) = ∑ W jk (n)m(k, n) + I(j, n), (1.8)<br />
k<br />
where I(j, n) denotes the charge injected at the jth site at time n, and W jk (n) is a<br />
stochastic variable which is equal to 1 when the particle on the kth site jumps to the<br />
jth site and which is equal to 0 otherwise. Since one particle cannot go to two different<br />
sites in a single time step, W jk (n) must be normalized as ∑ j W jk (n) = 1. In case of<br />
W jk (n) = 1/N with probability 1/N, where N is the total number of sites, which is<br />
taken to be infinity in the calculation. This is what corresponds to the mean field<br />
limit. The distribution of m can be obtained by introducing the r-body characteristic<br />
function<br />
Z 1 (ρ, n + 1) = Φ(ρ) ∑ a r Z 1 (ρ, n) r , (1.9)<br />
r=0<br />
where a r ≡ ( N r )(1/N)r (1 − 1/N) N−r . In the limit N → ∞, the steady-state solution<br />
of Eqn. 1.9 is obtained as<br />
Z 1 (ρ) = 1 − √ 2 < I > 1/2 |ρ| 1/2 i −1/2 + . . . for < I >≠ 0 (1.10)
Introduction 18<br />
This equation gives the following solution for m ≫< I > > 0 (or for m ≪< I > <<br />
0):<br />
p(m) ∝ |m| −3/2 , (1.11)<br />
where p(m) is the probability density.<br />
A similar result has also been obtained by Majumdar et al. [59] for aggregation<br />
in a mass conserving mean field type site-site interaction model. It has also been<br />
shown that small changes in the breakup parameters do not affect the decay exponent<br />
[60, 61]. This happens to be a very general case corresponding to a broad class of<br />
phenomena. As shown by Bonabeau et al. [62, 60], fish schools, with breakup and<br />
injection, show a similar aggregation, the size distribution showing an inverse power<br />
law behaviour with an exponent of 3/2. This is seen even in economics related to<br />
distribution of wealth [61]. Hence the nonequilibrium aggregation process is known to<br />
be Universal in nature [63].
Chapter 2<br />
Experimental techniques<br />
2.1 Introduction<br />
In the era of the advancement of electronic devices, fabrication and characterization<br />
techniques play an important role. As the device size is shrinking to nanometer scale,<br />
size, shape, and local environment of the building blocks (say transistors) will eventually<br />
strongly influence the device performance. The above fact has given a big push to<br />
research related to the synthesis of nanostructured materials, either on or embedded<br />
in different substrates. A variety of experimental tools are being used for both the<br />
fabrication and the characterization of nanostructured materials.<br />
In this chapter, we provide a brief description of some of the synthesis and characterization<br />
techniques that are used for preparation and characterization of the samples<br />
used in the present studies. We begin with a short description of the vacuum evaporation<br />
thin film deposition setup which was used for preparation of ZnS films. This<br />
is followed by a description of the ion implantation/irradiation setups which were essentially<br />
particle accelerators used for acceleration of heavy ions in the keV and MeV<br />
energy ranges. Following this, we will discuss briefly, the principles of ion scattering, x-<br />
ray diffraction, transmission electron microscopy, atomic force microscopy, and optical<br />
absorption spectroscopy. These are the techniques used for sample characterization.<br />
2.2 Sample preparation techniques<br />
2.2.1 Vacuum evaporation thin film deposition setup<br />
A schematic diagram of the vacuum coating unit (HINDHIVAC, Bangalore), available<br />
at the Institute of Physics (IOP), Bhubaneswar, is shown in Fig. 2.1. This unit<br />
19
Experimental techniques 20<br />
Thickness monitor<br />
Water<br />
cooled<br />
chamber<br />
wall<br />
Vent and gas<br />
flow valves<br />
Sample holder<br />
Shutter<br />
Mo boat<br />
To vacuum<br />
pumps<br />
Figure 2.1: Schematic diagram<br />
showing the thermal evaporation<br />
(resistive heating) experimental<br />
setup.<br />
I<br />
Power<br />
supply<br />
is equipped with three resistive heating facilities and an electron beam evaporation<br />
facility (not shown in the figure). All these are housed in a relatively large vacuum<br />
chamber with a base pressure in the low 10 −6 mbar range. There is a sample holder and<br />
a quartz crystal thickness monitor (that could be placed close to the sample) mounted<br />
inside the chamber. For typical applications, the material is loaded in a Molybdenum<br />
(Mo) boat, which is resistively heated by passing a high current through it. During<br />
heating, the material, to be deposited, melts and evaporates from the boat. At this<br />
low pressure ∼ 10 −6 mbar , the mean free path of the evaporated atoms is larger<br />
than the Mo boat to substrate distance. Therefore the evaporated particles travel in<br />
straight lines from the evaporation source towards the substrate kept above. In the<br />
present arrangement the sample holder is always at room temperature. A shutter,<br />
available just below the sample holder, allows to start or stop the material deposition<br />
on the substrates.<br />
2.2.2 Ion implantation and irradiation<br />
The Au implantations carried out on various samples, as discussed in this thesis,<br />
have been performed using the low energy negative ion implanter facility at IOP,<br />
Bhubaneswar [64]. The higher energy ion irradiation experiments, carried out on ZnS<br />
thin films and Au implanted silica glass samples have been carried out using the implantation<br />
beamline of the 3 MV tandem Pelletron accelerator (9SDH-2, NEC, USA)
Experimental techniques 21<br />
at IOP [65] and the materials science beamline of the 15 MV tandem Pelletron accelerator<br />
at the Inter University Accelerator Centre, New Delhi [66]. A brief description<br />
of the ion implantation and the irradiation facilities are given below.<br />
Low energy negative ion implanter facility<br />
valve<br />
cathode<br />
cooling<br />
000000000000<br />
111111111111<br />
000000000000<br />
111111111111<br />
000000000000<br />
111111111111<br />
000000000000<br />
111111111111<br />
000000000000<br />
111111111111<br />
000000000000<br />
111111111111<br />
00 11<br />
00 11<br />
00 11<br />
magnet<br />
gate Cs vapor vent cryo gate<br />
valve pump valves<br />
ion ion acce− pump<br />
source −leration + tee<br />
ionizer<br />
power<br />
steerer quadrupole<br />
singlet<br />
drift<br />
tube<br />
BPM<br />
slit<br />
FC<br />
einzel<br />
lens<br />
000000000000<br />
000000000000<br />
111111111111<br />
111111111111 cyro<br />
000000000000<br />
111111111111 pump<br />
000000000000<br />
111111111111<br />
pump + tee<br />
scanner<br />
000000000000<br />
111111111111<br />
000000000000<br />
111111111111<br />
000000000000<br />
111111111111 TMP rotary<br />
000000000000<br />
111111111111 pump<br />
000000000000<br />
111111111111<br />
gate<br />
valves<br />
target<br />
holder<br />
chamber<br />
Figure 2.2: Schematic of the low energy negative ion implanter facility. BPM – beam profile<br />
monitor, FC – Faraday cup, TMP – Turbo-molecular pump.<br />
A schematic of the low energy negative ion implanter facility used is shown in Fig. 2.2.<br />
This machine is equipped with a SNICS-II ion source made by NEC, USA [67, 68].<br />
SNICS stands for source of negative ions by cesium sputtering and the source is capable<br />
of producing ions from almost all the elements in the periodic table, except the inert<br />
gas elements. In SNICS, Cs vapours from a heated reservoir are thermally ionized<br />
by a tantalum ionizer and are accelerated towards a cold cathode, in order to sputter<br />
out the cathode material. A part of the Cs vapour also gets deposited on the cathode<br />
surface in the form of a film. The sputtered materials while coming out of the cathode<br />
get an extra electron while passing through the Cs film. A positive potential is applied<br />
to the extractor and focus electrode assembly to extract the negative ions in the form<br />
of a beam. The entire source is kept floating at a high negative potential. The ions<br />
thus extracted get accelerated to the ground potential. If C, E, and B are the cathode,<br />
extractor and bias potentials respectively, then the total negative potential at which<br />
negative ions are generated is given by V 0 = C + F + B. SNICS is known to provide<br />
singly charged ions and therefore the energy of the ions is given by V 0 eV (V 0 is usually<br />
between 10–60 keV). To shift the beam suitably, an electrostatic X-Y steerer assembly
Experimental techniques 22<br />
is provided. This is followed by an electrostatic quadruple singlet which focuses the<br />
beam in the vertical plane and puts it into a 45 ◦ bending magnet. This magnet is<br />
being used for the ion mass selection. Following the magnet, there is a drift section at<br />
the end of which there is a beam profile monitor (BPM) and a Faraday cup followed<br />
by an Einzel lens. The BPM provides a continuous oscilloscope display of the shape<br />
and position of the beam, the Faraday cup is used to monitor the beam current.<br />
After a drift section there is an Einzel lens which is used to focus the beam onto the<br />
target inside a scattering chamber. This beam is raster scanned over the sample by<br />
using another X-Y steerer placed before the scattering chamber. The entire system<br />
is maintained at high vacuum (10 −6 mbar) using two cryo pumps, a turbo molecular<br />
pump followed by a rotary pump, as shown in Fig. 2.2. This implanter facility can<br />
provide singly charged negative ions with a maximum energy of 60 keV.<br />
High energy ion implantation/irradiation facilities<br />
Figure 2.3: Schematic diagram of the 3.0 MV 9SDH-2 tandem Pelletron accelerator facility<br />
at the Institute of Physics, Bhubaneswar.<br />
A schematic diagram of the 3 MV tandem accelerator along with various beamlines,
Experimental techniques 23<br />
available at IOP, Bhubaneswar, is shown in Fig. 2.3. The machine is equipped with<br />
two ion sources – one RF ion source (alphatros) used for providing inert gas ions and<br />
the other one, a Cs sputtering ion source (a multi cathode SNICS). The negative ions<br />
produced by the ion source are accelerated to a maximum energy of 75 keV before<br />
entering the machine through a 90 ◦ injector magnet. This mass selected low energy<br />
beam is focused using an Einzel lens and directed into the accelerating column of<br />
the Pelletron machine. The Pelletron is a typical double ended Van-de-Grafff with<br />
the belts replaced by chains of metallic pellets separated by insulating connectors.<br />
The pellets carry the induced charge more efficiently into the terminal since charges<br />
reside on the inner surface of each pellet.In total there are two chains carrying charges<br />
continuously to the terminal. In the high voltage terminal of the Pelletron, the negative<br />
ions are converted to the positive ions by making them pass through a gas-stripper.<br />
The positive ions formed (of different charge states) get further accelerated towards<br />
the other end of the machine kept at ground potential. After the acceleration, the ions<br />
pass through an analyzing magnet for selecting the required ions with right energy.<br />
Finally the switching magnet directs the beam into the desired beamline. For ion<br />
implantation or irradiation at few MeV, the beamline at -15 ◦ , as shown in Fig. 2.3,<br />
is used. This beamline is equipped with a raster scanner operating at 517 Hz in the<br />
horizontal direction and 64 Hz in the vertical direction.<br />
A similar tandem accelerator facility with a maximum terminal voltage of 15 MV,<br />
available at the Inter University Accelerator Centre, New Delhi, has been used to<br />
irradiate samples at much higher energy using Swift Heavy Ions (SHI) (e.g. 100 MeV<br />
Au 8+ ). The working principle of the this machine is almost similar to the 3 MV tandem<br />
Pelletron accelerator, described above. The implantation/irradiation chambers are<br />
maintained at high vacuum using turbo-molecular pumps and are equipped with the<br />
necessary electrical connections for accurate beam current measurements. The target<br />
holders have two degrees of freedom viz. z-position and tilt angle. In each case several<br />
samples can be loaded at a time.<br />
2.3 Characterization techniques<br />
2.3.1 Rutherford backscattering spectrometry<br />
One of the most important analytical methods in material science using ion beams is<br />
the Rutherford backscattering spectrometry (RBS) [69]. RBS is based on the hard
Experimental techniques 24<br />
sphere collisions between atomic nuclei and derives its name from Lord Ernest Rutherford,<br />
who in 1911 was the first to present the concept of atoms having nuclei. The<br />
technique involves shooting a monoenergetic beam of light ions (proton, He + or He ++ )<br />
into a target and detect the number and energy of particles in a backward direction.<br />
With this information, it is possible to determine atomic mass of the scatterer and<br />
elemental concentrations as a function of depth starting from the surface.<br />
When a sample is bombarded with a beam of high energy light particles (few keV<br />
< E < few MeV), the interaction between the projectile and the target atoms, can be<br />
modeled accurately as an elastic collision using classical physics. There are four basic<br />
factors that are involved in the process [69]. These are given below.<br />
• The energy of the projectile after the collision can be related to its energy before<br />
the collision by means of a kinematic factor. This factor brings in the much<br />
needed target-mass dependence.<br />
• The interaction between the projectile and the target atom can be described as<br />
an elastic Coulomb collision between two isolated particles and expressed in term<br />
of a scattering cross section. Knowing the cross section, the scattering yield can<br />
be connected to the target mass concentration.<br />
• The energy loss of the energetic particle, as it traverses the scattering medium,<br />
is expressed in terms of the stopping cross section. Through this one determines<br />
the energy of the ion just before it gets scattered in a given depth. It also is used<br />
to determine the final energy at which the ion comes out losing energy along its<br />
path.<br />
• The energy loss is a statistical process. Mono-energetic projectiles assume an<br />
energy distribution after penetrating a certain depth in the target. This uncertainty<br />
in energy loss is known as energy straggling.<br />
The above four concepts form the basis of RBS analysis. The kinematic factor provides<br />
RBS an elemental detection capability, the scattering cross section gives it the ability<br />
of a quantitative measurement, the stopping cross section results in the capability for<br />
the depth analysis and energy straggling sets limits on mass and depth resolutions.<br />
When a projectile atom with mass M 1 and energy E 0 undergoes an elastic collision<br />
with a stationary atom of mass M 2 (M 2 > M 1 ) in a target, energy is transferred<br />
from the projectile to the target. The interaction between two atoms can be properly<br />
described by a simple elastic collision of two isolated particles. The interaction is
Experimental techniques 25<br />
basically Coulomb interaction between two nuclei of charges Z 1 and Z 2 , provided<br />
the following two conditions are satisfied: (i) the projectile energy E 0 must be much<br />
larger than the binding energy of the atoms in the targets and (ii) nuclear reactions<br />
and resonances must be absent. The energy transfers in the elastic collision between<br />
two isolated particles can be obtained by applying the principles of conservation of<br />
energy and momentum. A schematic diagram of the scattering process is shown in<br />
Fig. 2.4. The energy (E 1 ) of the projectile after the collision with the target atom is<br />
scattering angle<br />
E 1, M 1<br />
collimated<br />
E 0, M 1 beam<br />
M 2<br />
/<br />
E , M 2<br />
Figure 2.4: Schematic presentation<br />
of an elastic collision<br />
between an energetic<br />
projectile and a target atom.<br />
related to its energy before collision, E 0 and is given by [69]<br />
where k =<br />
angle.<br />
[<br />
E 1 = k E 0 , (2.1)<br />
] 2<br />
M 1 cosθ+(M2 2−M2 1 sin2 θ) 1/2<br />
M 1 +M 2<br />
is the kinematic factor and θ is the scattering<br />
If Q is the total number of incident particles that hit a target and dQ is the<br />
number of backscattered particles recorded by the detector, placed subtending a solid<br />
angle dΩ at an angle θ with the direction of incidence, then the differential scattering<br />
cross section dσ<br />
dσ<br />
dΩ = 1 Nt<br />
[ dΩ<br />
dQ/dΩ<br />
Q<br />
is related to the number of detected particles dQ by the relation<br />
]<br />
, where N is the volume density of atoms in the target and t is its<br />
thickness. So Nt is the number of target atoms per unit area. Thus, the number of<br />
particles detected by a detector with 100 % efficiency, subtending a solid angle dΩ is<br />
given by<br />
Y = QNt dσ dΩ (2.2)<br />
dΩ<br />
The differential scattering cross section for an elastic collision between two atoms is<br />
given by Rutherford’s formula<br />
[<br />
dσ<br />
dΩ = Z1 Z 2 e 2 ]<br />
[ 2<br />
4 [1 − (<br />
M 1<br />
M 2<br />
sinθ) 2 ] 1/2 + cosθ ] 2<br />
4E 0 sin 4 [ ]<br />
θ<br />
1/2<br />
(2.3)<br />
1 − (<br />
M 1<br />
M 2<br />
sinθ) 2
Experimental techniques 26<br />
Simulations of the RBS spectra are needed for the interpretation of the experimental<br />
results. This is because the composition in the depth direction can not be<br />
derived, directly, from the experimental data. Several algorithms [70, 71, 72] based on<br />
the above physical principles are available for this purpose .<br />
RBS setup<br />
All the ion scattering experiments as presented in this thesis have been carried out in<br />
a multi-purpose scattering chamber, equipped with a sample manipulator under high<br />
vacuum conditions (∼ 10 −6 Torr). This chamber sits in the 45 o beam line (schematically<br />
shown in Fig. 2.3) of the 3 MV tandem Pelletron accelerator (9SDH-2, NEC,<br />
USA) facility at IOP [73]. For RBS experiments pertaining to the studies presented<br />
in this thesis, He ions of 1.35 to 3.0 MeV energy, with beam currents in the range of<br />
1 to 15 nA have been used.<br />
current<br />
integrator<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
sample<br />
slit 2 slit 1<br />
mass selected mono−energetic<br />
ion beams from accelerator<br />
scattered ions<br />
detector<br />
ADC AMP PREAMP<br />
MCA<br />
detector<br />
bias<br />
Figure 2.5: Schematic diagram showing Rutherford backscattering spectrometry experimental<br />
setup. PREAMP – pre-amplifier, AMP – amplifier, ADC – analog to digital converter,<br />
MCA – multi-channel analyzer.<br />
A schematic diagram of the RBS experimental setup is shown in Fig. 2.5. A well<br />
collimated mono-energetic beam of ions (usually He + or He ++ ), obtained from the<br />
accelerator is made to fall on a target. A silicon surface barrier detector placed in<br />
front of the target detects the scattered ions coming at a backscattering angle. A<br />
good setup allows for target tilt (changing the angle of incidence) and an angular<br />
movement of the detector leading to a change in the scattering angle. The distance
Experimental techniques 27<br />
from the sample can also be adjusted resulting in a change in solid angle (usually kept<br />
Experimental techniques 28<br />
sample surface at a grazing angle. Especially for the analysis of thin films, XRD<br />
techniques have been developed for which the primary beam enters the sample with<br />
a very small angle of incidence. In its simplest variant, this configuration is denoted<br />
by GIXRD that stands for grazing incidence XRD. The small entrance angle causes<br />
the path traveled by the x-rays to significantly increase and the structural information<br />
contained in the diffractogram to stem primarily from the thin surface layer. In order<br />
to evaluate the mean grain size (D) of the film, Scherrer’s formula [74],<br />
D =<br />
0.94 λ<br />
β cos(θ B )<br />
(2.6)<br />
is often used, where λ, β, and θ B are incident radiation wavelength, full width at<br />
half maximum of the diffraction peak (in radians), and the Bragg diffraction angle,<br />
respectively.<br />
XRD setup<br />
Detector<br />
Divergent<br />
slits<br />
Axis<br />
X−rays<br />
Receiving<br />
slits<br />
Theta<br />
2 Theta<br />
Figure 2.6: Schematic diagram<br />
illustrating the x-ray<br />
diffractometer setup.<br />
X−ray<br />
source<br />
Theta<br />
Sample<br />
A schematic diagram of the XRD setup used in the present studies is shown in Fig. 2.6.<br />
A well collimated monochromatic beam of x-rays (Cu K α radiation, i.e., λ=1.54<br />
Å from a Panalytical made X ′ PERT powder x-ray diffractometer) are directed towards<br />
the sample. The interaction of the incident rays with the sample produces<br />
constructive interference (and a diffracted ray) when Bragg’s law, i.e., Eqn. 2.4 is<br />
satisfied. These diffracted x-rays are then detected, processed, and counted. In case<br />
of pollycrystalline or powder samples, by scanning through a range of 2θ angles, all<br />
possible diffraction directions of the lattice could be included for analysis. Conversion
Experimental techniques 29<br />
of the diffraction peaks to d-spacings allows identification of the materials because<br />
of unique d-spacings. Typically, this is achieved by comparison of d-spacings with<br />
standard reference patterns.<br />
2.3.3 Transmission electron microscopy<br />
Like XRD, transmission electron microscopy (TEM) provides informations, such as<br />
crystal structure and microstructure of materials [75]. Apart from these, TEM observation<br />
can provide direct imaging of the structure of the particles, from which size,<br />
shape, and their arrangements can be obtained. This technique can be used for the<br />
study of defects in the atomic-scale. TEM uses a high energy electron beam transmitted<br />
through a very thin specimen to image and analyze its microstructure with atomic<br />
scale resolution. The resolution of an optical microscope is limited by the wavelength<br />
of light used and hence limited to the order of a micrometer. However, electrons accelerated<br />
to hundreds of keV have wavelengths much smaller than that of light. At an<br />
energy of 200 keV, electrons have a wavelength of 0.025 Å. However, the resolution of<br />
a TEM is limited, by the aberrations inherent in electromagnetic lenses, to about 1 to<br />
2 Å and hence can reveal details down to the atomic scale.<br />
Energetic electrons incident on the specimen undergo various elastic and inelastic<br />
scatterings due to interaction with the atoms of the specimen [75]. Most of the electrons<br />
are elastically scattered by the nuclei of atoms in the specimen. Some electrons<br />
are inelastically scattered by the electrons in the specimen as well. Energetic electrons<br />
passing through the specimen near the nuclei are somewhat accelerated towards the<br />
nuclei causing small local reductions in wavelength, resulting in a small phase change.<br />
In this way, information about the specimen structure get transferred to the phase of<br />
the electrons. For getting high resolution images only the elastically scattered electrons<br />
are of importance. The inelastically scattered electrons contribute mostly to<br />
the background. This can be removed by inserting an energy filter in the microscope<br />
between the specimen and the recording device. The inelastically scattered electrons<br />
can sometimes be useful. Analysis of them can be helpful in obtaining chemical information<br />
from very small regions in the spectrum. From the energy loss spectrum of the<br />
inelastically scattered electrons or the filtered part of the spectrum, this information<br />
can be extracted.<br />
Electron diffraction from a crystalline lattice can be described as a kinematic scattering<br />
process that meets the wave reinforcement and interference conditions given
Experimental techniques 30<br />
in the Bragg equation. The patterns are formed by diffraction of an electron beam<br />
transmitted through the thin specimen. The electron diffraction yields spot patterns<br />
from single crystals, ring patterns from fine-grains and randomly oriented crystallites<br />
and superimposed ring and spot patterns from larger-grain polycrystalline films.<br />
TEM setup<br />
A transmission electron microscope consists of an illumination system, specimen stage<br />
and imaging system analogous to a conventional light microscope. A schematic diagram<br />
of a TEM is shown in Fig. 2.7. At the top, there is an electron gun, producing<br />
filament<br />
000 111 0000<br />
000 111 0000<br />
000 111 0000<br />
000 111 0000 1111<br />
1111<br />
1111<br />
1111<br />
000 111<br />
1100 0000 1111<br />
000 111<br />
0000 1111<br />
000 111<br />
0000 1111<br />
000 111 0000 1111<br />
000 111 0000 1111<br />
wehnelt cap<br />
(negative potenial)<br />
space charge<br />
gun corss over<br />
electron beam<br />
anode plate<br />
(positive potenial)<br />
000000000000000000000<br />
111111111111111111111first condenser lens<br />
000000000000000000000<br />
111111111111111111111<br />
00000000000<br />
11111111111<br />
00000000000<br />
11111111111<br />
00000000000<br />
11111111111<br />
00000000000<br />
11111111111<br />
1111100000<br />
1111100000<br />
second condenser lens<br />
Condenser aperture<br />
sample<br />
0000 1111<br />
00000000000<br />
11111111111objective lens<br />
00000000000<br />
11111111111objective aperture<br />
selectred area aperture<br />
Figure 2.7: Block diagram<br />
of a transmission electron microscope.<br />
00000000000<br />
11111111111first intermediate lens<br />
00000000000<br />
11111111111<br />
00000000000<br />
11111111111<br />
00000000000<br />
11111111111<br />
00000000000<br />
11111111111<br />
second intermediate lens<br />
00000000000<br />
11111111111projector lens<br />
00000000000<br />
11111111111<br />
000000000000<br />
111111111111 main screen (phospher)<br />
a stream of monoenergetic electrons. In the system we have used, the electrons were<br />
generated using a LaB 6 electron source by thermionic emission. The stream of electrons<br />
from the electron gun is focused to a small, thin, coherent beam by the use of<br />
the first and the second condenser lenses. The first lens largely determines the “spot
Experimental techniques 31<br />
size”; the general size range of the final beam spot that strikes the sample. The second<br />
lens is used for varying the size of the spot on the sample, changing it from a wide<br />
dispersed spot to a pinpoint beam. The beam is restricted by the condenser aperture<br />
(user selectable), knocking out high angle electrons (those far from the optic axis).<br />
The beam strikes the specimen and part of it is transmitted. This transmitted<br />
portion is focused by the objective lens into an image. Optional objective and selected<br />
area metal apertures can restrict the beam. The objective aperture enhances<br />
the contrast by blocking out high-angle diffracted electrons, the selected area aperture<br />
enabling the user to examine the periodic diffraction of electrons by ordered<br />
arrangements of atoms in the sample. The image is passed down the column through<br />
the intermediate and projector lenses, being enlarged all the way. Finally the image<br />
strikes the phosphor screen allowing the user to see the image. The darker areas of<br />
the image represent those areas of the sample that fewer electrons were transmitted<br />
through (they are thicker or denser). The lighter areas of the image represent those<br />
areas of the sample through which more electrons were transmitted through (they are<br />
thinner or less dense). A detailed description of the transmission electron microscopy<br />
technique is given in Ref. [75].<br />
Sample preparation for TEM<br />
Due to the strong interaction between electrons and matter, the specimens have to<br />
be rather thin (
Experimental techniques 32<br />
energy ion beam at grazing incidence. Ar ions of energy 1.5 keV are used for this<br />
purpose.<br />
2.3.4 Atomic force microscopy<br />
Atomic force microscopy (AFM) is an extensively used techniques in the field of fundamental<br />
research, technology and bio-medicine [76]. It does not require the sample<br />
to be electrically conductive unlike scanning electron microscopy. AFM is applicable<br />
to any type of materials – insulators, semiconductors as well as conductors. Because<br />
it uses a probe which is very small and sharp, AFM can be used to acquire highresolution<br />
real-space topographic images of various sample surfaces. Hence it is a very<br />
useful technique to monitor or characterize the morphology and topological changes of<br />
various ion irradiated surfaces [76]. A schematic diagram of AFM is shown in Fig. 2.8.<br />
The AFM probes the surface of a sample with a sharp tip, a couple of µm long and<br />
Display and<br />
feedback<br />
electronics<br />
Sample surface<br />
Position<br />
sensitive<br />
photodiode<br />
Laser<br />
Cantilever<br />
& Tip<br />
Figure 2.8: Schematic diagram illustrating<br />
an atomic force microscope<br />
setup with a scanner attached with the<br />
sample holder.<br />
PZT Scanner<br />
often less than 10 nm in diameter. The tip is located at the free end of a cantilever,<br />
that is 100 to 200 µm long. Forces between the tip and the sample surface cause the<br />
cantilever to bend or deflect. A detector measures the cantilever deflection as the tip<br />
is scanned over the sample or the sample is scanned under the tip. The measured<br />
cantilever deflection allow a computer to generate a map of surface topography.<br />
An AFM can be used to study insulators and semiconductors as well as conductors<br />
as mentioned above. Several forces typically contribute to the deflection of an AFM<br />
tip and hence the cantilever. The force most commonly associated with AFM is the<br />
interatomic van der Waals force. The dependence of the van der Waals force on the
Experimental techniques 33<br />
Tip is in hard contact<br />
with the surface −<br />
repulsive regime<br />
Tip is far from the surface<br />
− no deflection<br />
Force<br />
0<br />
Figure 2.9: van der Waals<br />
force as a function of tip to<br />
sample surface distance.<br />
Tip is pulled towards the<br />
surface − attractive regime<br />
Probe distance from sample (z distance)<br />
distance between the tip and the sample surface is schematically shown in Fig. 2.9.<br />
Two distinct regimes are labeled in the figure namely attractive and repulsive regimes<br />
(contact and non-contact regimes). In the contact regime, the cantilever is held less<br />
than a few Å from the sample surface and the interatomic force between the cantilever<br />
and sample is repulsive. In the non-contact regime, the cantilever is held above the<br />
surface at about tens to hundreds of Å from the sample surface. The interatomic force<br />
between the cantilever and the sample, in this mode, is attractive (largely a result of<br />
the long-range van der Waals interaction).<br />
In addition to the contact and the non-contact mode there is a third mode of<br />
operation corresponding to the tapping mode which is an intermittent-contact mode.<br />
In this case the vibrating cantilever is brought closer to the sample surface so that<br />
at the bottom of its travel just barely hits or “taps” the sample. This mode is a<br />
compromise between the contact and non-contact modes where the cantilever is made<br />
to oscillate so that the tip remains very close to the sample for a short time and then<br />
going far away for a short time. The compromise between the two allows one to scan<br />
soft adsorbate on a substrate with better resolution than in the non-contact mode but<br />
with a small interaction (and consequently less modification of the samples) between<br />
the tip and adsorbate as in the non-contact mode.<br />
2.3.5 Optical absorption<br />
Optical absorption by metal particles<br />
In the beginning of the 20 th century, Gustav Mie solved Maxwell’s equations for an<br />
electromagnetic light wave interacting with a metal [77]. When applying the correct
Experimental techniques 34<br />
boundary conditions for a metallic sphere, the calculations gave a series of multipole<br />
oscillations (dipole, quadrupole, etc.) for the extinction and scattering cross sections<br />
of the particle as a function of particle radius. Using the macroscopic, frequencydependent<br />
complex dielectric constant of the bulk material of the metal sphere, he<br />
was able to explain the beautiful colors of stained glass first systematically studied by<br />
Michel Faraday in 1857 [78].<br />
For very small metal particles in an optical medium, the fields can be considered<br />
homogeneous (constant) over the particle, where only the low-order modes contribute<br />
to the extinction. Then, the particles can be imagined to be placed in a uniform static<br />
electric field (quasistatic regime). When there is negligible scattering, extinction cross<br />
section is same as the absorption cross section and the extinction cross section, under<br />
the dipolar approximation, is given by [79]<br />
σ ext = 12πR3 ωǫ 3/2<br />
m<br />
c<br />
ǫ 2 (ω)<br />
[ǫ 1 (ω) + 2ǫ m ] 2 + ǫ 2 (ω)<br />
(2.7)<br />
In the above expression R is the radius of the metal particle, ω is the angular frequency<br />
of the incident light, c, is the speed of light in vacuum, ǫ 1 (ω) and ǫ 2 (ω) are the<br />
real and imaginary parts of the metal dielectric function, ǫ, and ǫ m is the dielectric<br />
function of the matrix. A resonance in the spectrum of σ ext will occur whenever<br />
the condition ǫ 1 (ω) = -2ǫ m is satisfied. This resonance is well known as the surface<br />
plasmon resonance (SPR) which occurs due to a collective oscillation of the conduction<br />
electrons of the metallic system in response to an exciting electromagnetic radiation.<br />
As can be seen from Eqn. 2.7, the extinction cross section depends on the dielectric<br />
function, ǫ, of the metal particles. However, no analytical theory gives a correct<br />
description of the variation of the dielectric function with frequency, ω. Although the<br />
Drude model and the Drude-Sommerfield model [80] are effective in explaining many<br />
of the metal properties, a correct description of the dielectric constant is still lacking.<br />
The main weakness of the “free” or “nearly free electron model” [80] is its inability to<br />
account for the intraband transitions. These intraband transitions are very important<br />
in case of Au- and Ag-like noble metals. In addition, for small particles (nm size),<br />
some extra effects, such as quantum confinement [81] can play a significant role.<br />
Therefore, in case of very small particles (sizes ≤ 50 nm), to get the correct size<br />
dependence of the absorption cross section, one needs to use the dielectric function for<br />
the metal, corrected for the variation of electron mean free path with the radius.(For<br />
example, electron mean free path of bulk Ag is around 40 nm [82].) A phenomenological<br />
formula for the complex dielectric function, incorporating such a correction is
Experimental techniques 35<br />
given by [83]<br />
ǫ(ω, R) = ǫ bulk (ω) +<br />
ωp<br />
2 ωp<br />
2 −<br />
ω 2 + iωγ 0 ω 2 + iω(γ 0 + Av F /R)<br />
(2.8)<br />
In the above equation, R is the radius of the nanocluster, ǫ bulk (ω) = ǫ 1 (ω) + iǫ 2 (ω) is<br />
the bulk dielectric function of the metal particle, ω p is the plasma frequency, A is a<br />
constant usually taken as 1.0, and v F is the Fermi velocity. The factor γ 0<br />
is defined<br />
as v F /l ∞ , where l ∞ is the mean free path of the electrons in the bulk material. Using<br />
Eqn. 2.8 in Eqn. 2.7, one can estimate the optical properties of the particles of a<br />
given size. For calculations involving Cu, Ag, and Au, the bulk dielectric constants,<br />
as measured experimentally by Jonhson and Christy [82], are used.<br />
Optical absorption by semiconductors<br />
When a semiconductor is illuminated, photons can make the valence electrons of an<br />
atom go to a higher electronic energy levels in the conduction band. In this process,<br />
the incident photon is absorbed by an electron. The fundamental absorption process<br />
is strongly affected by whether the gap between the valence band maximum and the<br />
conduction band minimum (i.e. optical band gap) is a direct or an indirect [80] one.<br />
In a direct gap semiconductor, an electron executes a vertical transition at the band<br />
gap. Energy is conserved according to<br />
¯hω = E f − E i , (2.9)<br />
where, E f and E i are the final and initial state energies, respectively, of the material<br />
and ¯hω is photon energy. For an indirect gap material (like Si), there is a requirement<br />
of an additional momentum to reach the conduction band minimum at a non-zero wave<br />
vector. This momentum comes from interaction with a phonon. Then the statement<br />
of energy conservation then becomes<br />
¯hω = E f − E i ± ¯hΩ, (2.10)<br />
where, ¯hΩ is the energy of the phonon, and the plus and minus signs correspond<br />
to phonon emission or absorption, respectively. Phonon emission dominates at low<br />
temperatures. The need for an additional interaction with the phonon makes indirect<br />
absorption far less probable than direct absorption which means that is the indirect<br />
absorption is a weaker process.<br />
The absorption coefficient for a direct or indirect band-gap semiconductor can be<br />
calculated from<br />
α = C ∑ n i n f P if , (2.11)
Experimental techniques 36<br />
where, C is a constant, P if is the probability of the transition from the initial to the<br />
final state, n i is the electron density in the initial state and n f is the density of the<br />
empty states. The summation is over all states separated by the photon energy. For<br />
a direct gap semiconductor, this relation gives an absorption coefficient [84] as<br />
α = A(¯hω − E gap ) p , (2.12)<br />
where, A is a constant independent of ω (frequency of the photon) and the exponent<br />
p is 1/2 or 3/2 depending on quantum selection rules for the particular material. For<br />
example, the value of p for ZnS is 1/2 [84]. Thus using Eqn. 2.12, one can find out<br />
the band gap of a semiconductor by measuring the optical absorption coefficient, α.<br />
This has been used in our analysis.<br />
Optical absorption spectroscopy<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
m1<br />
m7<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
m8<br />
000 111<br />
000 111<br />
000 111<br />
m9<br />
sample<br />
compartment<br />
w3 w2<br />
w3<br />
sample<br />
reference<br />
w2<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
m3<br />
m4<br />
m5<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
C<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
000 111<br />
m6<br />
s3<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
m2<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
s2<br />
G1<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
00 11<br />
G2<br />
s1<br />
w1<br />
0000 1111<br />
0000 1111<br />
0000 1111<br />
0000 1111<br />
source<br />
0000 1111<br />
0000 1111<br />
0000 1111<br />
0000 1111<br />
0000 1111<br />
0000 1111<br />
0000 1111<br />
detector<br />
w1−w3 window plate<br />
G1−G2 grating<br />
C − chopper mirror<br />
s1−s3 slits<br />
m1−m9 mirrors<br />
Figure 2.10: Schematic diagram of the spectrophotometer used for optical absorption measurements.<br />
Optical spectroscopy is related to the absorption and emission of light by matter<br />
[85]. A spectrophotometer can be used for studying the optical absorption properties<br />
of embedded metal NPs and semiconductor thin films prepared in and the surface of<br />
transparent media. In the present case, as will be shown in the subsequent chapters, we
Experimental techniques 37<br />
are concerned with Au particles embedded in silica glass and ZnS thin films deposited<br />
on silica glass. For studies related to these samples a Shimadzu UV-3101PC UV-VIS-<br />
NIR dual beam spectrophotometer was used. A schematic diagram of the optics of<br />
the spectrometer is shown in Fig. 2.10. This spectrometer uses a deuterium or halogen<br />
lamp as the photon source. Different frequencies of the light are selected using a grating<br />
assembly. In one of the arms of the spectrophotometer, the sample is kept while the<br />
other is loaded with a reference sample. In studies involving metal nanoparticles in<br />
silica glass or thin film deposited on silica glass, we have used a pristine sample (silica<br />
glass piece from the sam batch) without any film deposited or ion implanted into it<br />
as the reference. Depending on the energy either a photomultiplier tube or a PbS<br />
detector is used for the detection of the outcoming photons. The spectrophotometer<br />
is connected to a personal computer equipped with the control and data acquisition<br />
software. Before starting the measurements one needs to make sure that the two<br />
identical reference samples show zero absorption (which is carried out through the<br />
baseline correction and auto zero options available with the control software).
Chapter 3<br />
Modification of nanocrystalline ZnS<br />
films by Au irradiation<br />
3.1 Introduction<br />
ZnS has been extensively investigated as an important wide-band-gap semiconductor<br />
(E g ∼ 3.6 eV at room temperature [86]). It is one of the oldest and probably one of the<br />
most important materials in the electronic industry with a wide range of applications<br />
in light-emitting diodes, photovoltaic cells, dielectric filters, and efficient phosphors in<br />
flat-pannel display. It is an excellent material for band-gap engineering. At ambient<br />
conditions, ZnS shows two structural polymorphs: cubic sphalerite and hexagonal<br />
wurtzite [87]. Synthesis of nanoscale ZnS, due to its marked varied properties, which<br />
come from quantum-confinement effect, has received a lot of attention in the recent<br />
years [86, 88, 89, 90]. To prepare ZnS thin films, a variety of techniques, such as<br />
molecular beam epitaxy [91], pulsed laser deposition [92], metalorganic chemical vapor<br />
deposition [93], chemical bath deposition [94], and thermal evaporation [95, 96] have<br />
been utilized. Among these, the technique of thermal evaporation is a very simple<br />
one suitable for good crystalline thin film preparation. This technique also provides a<br />
possibility to grow thin films with grain sizes in the nanoscale regime.<br />
As far as the ZnS is concerned, there are not many studies in literature that have<br />
investigated ion irradiation induced modifications. In a recent work, a change in<br />
morphology, of ZnS colloidal particles, from spherical to prolate shape under 4 MeV<br />
Xe irradiation has been reported [97]. Chowdhury et al. [98] have carried out such a<br />
study, on ZnS nanoparticles embedded in a polymer matrix, using Ni ions . They have<br />
observed an increase in size of the crystallites and a reduction in optical band gap with<br />
38
Modification of nanocrystalline ZnS films by Au irradiation 39<br />
increase in ion fluence. In an another study, Mohanta et al. have observed a tunable<br />
surface state emission from Mn-doped ZnS nanoparticles due to O irradiation [99].<br />
For both the studies, very high irradiation energies viz. 160 and 80 MeV, respectively,<br />
have been used. Since such studies are very limited it was necessary to carry out a<br />
systematic study of ion irradiation effects on ZnS over a wide range of energies. Results<br />
of this study on modifications in structural as well as optical properties and surface<br />
morphology are presented in this chapter [100]. Details regarding the experiment and<br />
results are given in the following sections.<br />
3.2 Experiments<br />
ZnS films (thickness ∼ 600 nm) were deposited, at room temperature, on cleaned<br />
Si(100) (with native oxide) and silica glass substrates by thermal evaporation of 99.99<br />
at. % pure ZnS powder (Materials Research Corporation, New York) using a vacuum<br />
coating unit at a working base pressure of ∼ 7×10 −6 Torr. The evaporant (ZnS) was<br />
placed in a molybdenum (Mo) boat at a distance of about 15 cm from the substrates.<br />
The thickness and the rate of deposition of the films were measured in-situ using a<br />
quartz crystal thickness monitor. The rate of deposition of the film was maintained<br />
at about 1.5 nm s −1 . This way prepared films were irradiated, at room temperature,<br />
with 35 keV Au − , 2 MeV Au 2+ , and 100 MeV Au 8+ ions at normal incidence with<br />
the sample surface. In each case the ion fluence was always kept fixed to a value of<br />
1×10 14 ions cm −2 with a maximum beam current density of ∼ 0.04 µA cm −2 to avoid<br />
the beam induced heating effects. The ions beam were raster scanned over an area of<br />
1×1 cm 2 on the sample surfaces for uniform irradiation.<br />
Rutherford backscattering spectrometry (RBS), using 3 MeV He 2+ ions was used<br />
to determine the composition and thickness of the as-grown and the irradiated films.<br />
RBS was carried out with a surface barrier detector at a backscattering angle of 160 ◦ .<br />
Crystal structure of the films were analyzed, before and after ion irradiations, using<br />
a Panalytical make X ′ PERT powder x-ray diffractometer with CuK α radiation (λ =<br />
0.154 nm) over a 2θ scan range of 10 ◦ −65 ◦ . Optical absorption (OA) measurements<br />
were carried out on the as-grown and the irradiated ZnS films, deposited on silica<br />
glass substrates, in the wavelength range of 250−800 nm using a Shimadzu UV 3101<br />
PC UV-VIS-NIR dual spectrophotometer. The measurements were carried out in<br />
the transmission mode with a blank silica glass sample (from the same batch) in the<br />
reference line. Irradiation induced changes in the surface morphology of the ZnS/Si
Modification of nanocrystalline ZnS films by Au irradiation 40<br />
samples were studied using a Nanoscope IIIA atomic force microscope (AFM) with a<br />
silicon nitride cantilever operated in tapping mode. All the measurements were carried<br />
out at room temperature.<br />
3.3 Results and Discussion<br />
3.3.1 Energy loss processes<br />
During irradiation, the incident ions impart their energy to the target material through<br />
both electronic and nuclear energy loss processes. The energy lost to the film results in<br />
modifications in the properties and it is important to estimate the relative importance<br />
of the energy loss processes. Using this in mind we have carried out a calculation<br />
of the energy loss by Au in ZnS using SRIM [31] at 35 keV, 2, and 100 MeV. The<br />
results, as a function of depth, are shown in Fig. 3.1. In the figure, nuclear energy<br />
1600<br />
600<br />
o<br />
Energy Loss [eV A -1<br />
]<br />
1400<br />
1200<br />
1000<br />
800<br />
600<br />
400<br />
200<br />
S<br />
S n + S e<br />
n<br />
Se<br />
E. L. [eV A ]<br />
-1<br />
o<br />
400<br />
200<br />
35 keV<br />
0<br />
0 1000 2000 3000 4000<br />
o<br />
Depth [A]<br />
0<br />
0 20000 40000 60000 80000 100000<br />
o<br />
Depth [A]<br />
2 MeV<br />
Figure 3.1: Variation of<br />
energy losses (S n , S e , and<br />
S T ) against ZnS depth for<br />
100 MeV Au ion at normal<br />
incidence. Inset: Same for<br />
35 keV and 2 MeV Au ions.<br />
All the energy loss data are<br />
extracted using SRIM code<br />
[31].<br />
loss (S n ), electronic energy loss (S e ), and sum of the both (S T = S n + S e ) are shown<br />
as a function of ZnS depth for the above three energies. One can see the value of S T ,<br />
in case of 35 keV Au irradiation, rapidly decreases with increase in depth. Here the<br />
major contribution is from S n . However, S T decreases slowly up to a depth of about<br />
8000 nm for 100 MeV Au irradiation. Beyond ∼ 8000 nm, S T is nearly constant. In<br />
this case, the major contribution comes from S e . On the other hand, for 2 MeV Au<br />
irradiation, S T is almost constant up to a depth of 300 nm. Beyond this S T decreases<br />
rapidly and finally Au atoms stop at a depth of about 400 nm. Here, both S n and
Modification of nanocrystalline ZnS films by Au irradiation 41<br />
S e are contributing to S T significantly. Therefore any change that takes place in the<br />
material properties is due to both electronic and nuclear energy deposition processes.<br />
3.3.2 Thickness and composition of the ZnS films<br />
Figure 3.2 shows the RBS spectrum of the as-grown ZnS film deposited on Si(100).<br />
Composition and thickness of the film were determined by simulating the RBS spectrum<br />
using GISA simulation code [72]. A good fit to the observed RBS spectrum<br />
could be obtained considering a single film with Zn to S concentration ratio as 1:1<br />
with a total thickness of around 615 nm. The above fit to the data is also shown in<br />
Fig. 3.2. This indicates that the as-grown film was stoichiometric in nature.<br />
Backscattered Yield (arb. unit)<br />
2000<br />
1500<br />
1000<br />
500<br />
. . . . Experimental<br />
Gisa fitted<br />
Si<br />
0<br />
200 300 400 500 600 700 800<br />
S<br />
Si ZnS<br />
Channel Number<br />
3 MeV He 2+<br />
Zn<br />
o<br />
20<br />
Detector<br />
Figure 3.2: RBS spectrum obtained<br />
with the as-grown ZnS film on<br />
Si(100) substrate. The Zn, S and Si<br />
edges are marked in the figure. Inset<br />
shows the RBS experimental geometry.<br />
3.3.3 Structural properties<br />
Figure 3.3 shows the XRD spectra of the as-grown and the Au irradiated ZnS films<br />
for various Au irradiation energies viz. 35 keV, 2, and 100 MeV, respectively. The<br />
spectrum corresponding to the as-grown film exhibits a peak at 2θ = 28.58 ◦ with a<br />
very narrow full width at half maximum (FWHM). This peak is identified as either of<br />
a cubic crystal structure with a preferred (111) orientation or a wurtzite structure with<br />
a preferred (002) orientation. This is because the reflections from these two planes<br />
cannot be resolved, using our x-ray diffractometer, due to their similar inter-planner<br />
spacing [101]. The presence of only one peak, indicating a preferential orientation,<br />
could be due to the growth of crystallites perpendicular to the substrate surface resulting<br />
from controlled nucleation process associated with the slow deposition rate.<br />
The observe narrow FWHM also indicates that the ZnS films were having good crystalline<br />
quality. At this point it is also important to estimate the mean grain size, D,
Modification of nanocrystalline ZnS films by Au irradiation 42<br />
as-grown<br />
Intensity [arb. units]<br />
35 keV<br />
2 MeV<br />
Figure 3.3: The XRD spectra of the<br />
as-grown and the Au irradiated ZnS<br />
films. The spectra corresponding to irradiated<br />
samples are shifted vertically<br />
downward for clarity.<br />
100 MeV<br />
10 20 30 40 50 60<br />
2θ o<br />
from the FWHM of the observed peak. This can be done using Scherrer’s formula<br />
[74], as given by<br />
D =<br />
0.94 λ<br />
β cos(θ β ) , (3.1)<br />
where λ, β, and θ β are CuK α radiation wavelength, the FWHM of the diffraction<br />
peak (in radians), and the Bragg diffraction angle, respectively. The estimated grain<br />
size of the as-grown film is found to be 17.4 nm (Table 3.1) which indicates that<br />
the crystallites were very small in size. It must be mentioned that the FWHM of<br />
the observed peak has been corrected due to instrumental broadening of ∼ 0.06 ◦ for<br />
determination of the mean grain size.<br />
Table 3.1: Parameters obtained form the XRD spectra for the as-grown and the Au irradiated<br />
ZnS films.<br />
Irradiation Peak Grain size Grain size<br />
energy position before correction after correction<br />
(2θ ◦ ) (nm) (nm)<br />
as-grown 28.58 17.2 17.4<br />
35 keV 28.58 17.2 17.4<br />
2 MeV 28.44 27.9 28.3<br />
100 MeV 28.31 46.7 49.5<br />
From Fig. 3.3, it is also observed that there is almost no change in intensity and<br />
FWHM of the XRD peak after an irradiation at 35 keV. On the other hand, compared<br />
to the as-grown case, there is a small increase in intensity and decrease in FWHM for
Modification of nanocrystalline ZnS films by Au irradiation 43<br />
the 2 MeV irradiation case. In case of the 100 MeV Au irradiated sample the peak<br />
intensity is seen to increases further with a further decreases in FWHM as compared to<br />
the other cases. The increasing peak intensity and narrowing of the peak is indicative<br />
of irradiation induced grain growth. The estimated grain sizes of the irradiated films,<br />
as obtained using Scherrer’s formula, are shown in Table 3.1. A direct correlation<br />
between an increase in grain size with irradiation energy is clearly seen.<br />
One can see from Fig. 3.1 that for 35 keV Au ion in ZnS, the S n and the S e values<br />
at the surface are 2.91 and 0.24 keV nm −1 , respectively. For 2 MeV Au ion, they<br />
are 3.66 and 1.91 keV nm −1 , respectively. The fact that we could not see any grain<br />
growth at 35 keV while there are grain growth for the Au irradiation at 2 MeV makes<br />
one feel that S e plays a dominant role regarding the observed grain growth. At both<br />
35 keV and 2 MeV although around 3 keV nm −1 , due to S n , is deposited in the ZnS<br />
film, in case of the lower ion energy, the ion range is limited to only 20 nm. Therefore,<br />
all energy is deposited in 20 nm where grain growth may be expected. However, this<br />
depth is too small compared to the total film thickness which is seen by the x-rays.<br />
Therefore the observed x-ray signal is mostly that coming from the film not affected<br />
by the irradiation. That is why the signal from the 32 keV Au irradiated film is<br />
almost identical to that obtained from the as-grown sample. However, in case there<br />
are thermal spikes produced solely from nuclear energy loss [37], there would be some<br />
melting associated with flow of materials towards the surface. This would result in an<br />
enhancement in surface roughness as compared to the as-grown sample. As would be<br />
shown later this is indeed true. Compared to this, at 2 MeV, both S e and S n play a<br />
role although S n dominates. Because of this, one would expect a larger material flow<br />
resulting in a higher surface roughness. At MeV energies, electrons in a cylindrical<br />
region around the ion path get excited and ionized. In case of ZnS, which has a low<br />
hole mobility [102], there is a possibility of Coulomb explosion coming from a higher<br />
density of positive charges in the ionized region. If the hole mobility is higher then<br />
the positive charges can get neutralized quickly not effecting a Coulomb explosion<br />
and a subsequent thermal spike [45]. However, there is an increase in grain growth<br />
coming from localized melting are recrystallization. Here energy deposition from both<br />
electronic excitations and nuclear collisions contribute.<br />
At 100 MeV with a much higher value of S e (∼ 15.8 keV nm −1 ), one would expect<br />
Coulomb explosion to take place followed by an inelastic thermal spikes [45] resulting in<br />
localized melting and recrystallization. Obviously, at 100 MeV one would have melting<br />
over bigger zones as compared to those produced at 2 MeV. In case of inelastic thermal
Modification of nanocrystalline ZnS films by Au irradiation 44<br />
spikes, produced by swift heavy ions (energy 1 MeV/amu), it has been shown that<br />
the ion track diameter over which melting occurs increases with ion energy [33]. This<br />
can result in larger grains over the entire film thickness. But with much reduced S n<br />
value ∼ 0.4 keV nm −1 , one would not expect any material flow towards the surface<br />
coming from nuclear collision effects. The surface roughness is thus expected to be<br />
comparable to that of the as-grown sample.<br />
From a closer look at Fig. 3.3, one can also see that there is a shift of the XRD<br />
peak towards lower values of 2θ with increase in irradiation energy from 35 keV to<br />
2 and 100 MeV. Further the XRD peak for the as-grown (or the 35 keV Au irradiated)<br />
sample is seen to be at a higher value of 2θ as compared to bulk ZnS [101]. In<br />
general, a shift in the peak position from that of its corresponding bulk material indicates<br />
the development of stress in the film [103] during deposition and post-deposition<br />
treatments. The observed strain can be due to several origins: (a) from the strained<br />
regions in the films (e.g. grain boundaries, dislocations, voids, impurities, etc.), (b)<br />
from the film-substrate interface (due to difference in thermal expansion coefficients<br />
between the film and the underlying substrate), (c) from the film-vacuum interface<br />
(surface stress, adsorption, etc.) or (d) from dynamic processes (e.g. recrystallization,<br />
interdiffusion, etc.) [104]. In the present case the strain is mostly due to point (a)<br />
as mentioned above. The strain, ǫ, developed in the films can be estimated using the<br />
relation [105]<br />
ǫ = d − d 0<br />
d 0<br />
× 100%, (3.2)<br />
where d is the lattice-spacing of the ZnS film and d 0 is the unstrained bulk latticespacing.<br />
A shift of the diffraction peaks to lower angles indicates the presence of<br />
tensile strain whereas a shift towards higher angles indicates a compressive strain<br />
[74, 106]. Based on this, the as-grown and the 35 keV Au irradiated films are found<br />
to have a compressive strain. This is because the observed XRD peak positions of<br />
the ZnS films are slightly higher than that of the bulk ZnS [101]. This compressive<br />
strain is modified into tensile strain after MeV Au irradiations. Further, the tensile<br />
strain, observed in the MeV irradiated films is seen to increase with increase in MeV<br />
energy. The modification of internal compressive stress to tensile stress is mainly due<br />
to the irradiation induced grain growth and annihilation of preexisting defects, such<br />
as dislocations in the as-grown films. That an increase in grain size, does result in a<br />
changeover from compressive strain to tensile strain, has been seen in ZnS thin films<br />
grown using the thermal evaporation technique [96]. The variation of the strain in
Modification of nanocrystalline ZnS films by Au irradiation 45<br />
going from the as-grown to the irradiated ZnS films with energy deposited in the film<br />
surface is shown in Fig. 3.4.<br />
0.8<br />
0.6<br />
<br />
(d)<br />
Strain [%]<br />
0.4<br />
(c)<br />
0.2<br />
(a) as-grown<br />
0.0<br />
(b) 35 keV<br />
(a)<br />
-0.2<br />
(c) 2 MeV<br />
(b)<br />
(d) 100 MeV<br />
-0.4<br />
0 4 8 12 16<br />
Energy Loss at Surface [keV/nm]<br />
Figure 3.4: Variation of the strain in<br />
as-grown and Au ion irradiated films<br />
as a function of energy deposition in<br />
the ZnS surface.<br />
3.3.4 Optical properties<br />
Based on the above XRD results it was not possible to conclude about the crystal<br />
structure of the ZnS film. It is also well known that cubic and wurtzite ZnS have<br />
different band gaps. For cubic ZnS the band gap is 3.66 eV [86] while for wurtzite<br />
structure it is 3.54 eV [94]. That prompted us to carry out an OA measurement on<br />
this system. The absorption spectra of the as-grown and the irradiated ZnS films<br />
are shown in inset of Fig. 3.5(a). The irradiation caused a tailing of the absorbance<br />
towards higher wavelength region, which also increases with increase in energy. At 35<br />
keV, we have not seen any observable change in the OA spectrum as compared to the<br />
as-grown film and, therefore, the corresponding results are not shown in the figure.<br />
To estimate the band gap of the films, we have used the standard expression for direct<br />
transition between two parabolic bands as given by [84]<br />
α 2 = C(hν − E g ), (3.3)<br />
where α, C, hν, and E g are absorption coefficient, a constant, incident photon energy,<br />
and the optical energy gap, respectively. The value of optical band gap, E g , is determined<br />
from the intercept on the energy axis, in the plot of (αhν) 2 versus hν. The data<br />
for various films are shown in Fig. 3.5(a). For the estimation of the value of α, we<br />
have made use of the relation α = 2.303 A/t [107], where A and t correspond to the<br />
measured absorbance and the film thickness, respectively. The estimated band gap<br />
for the as-grown film is found to be about 3.56 eV, which compares favorably with
Modification of nanocrystalline ZnS films by Au irradiation 46<br />
( α h ν ) 2 -1 2<br />
(10 eV cm )<br />
4<br />
8<br />
6<br />
4<br />
Absorbance [a. u.]<br />
0.4<br />
0.3<br />
0.2<br />
0.1<br />
0<br />
300 400 500<br />
Wavelength [nm]<br />
3.35<br />
2 as-grown<br />
3.30<br />
2 MeV<br />
100 MeV<br />
100 MeV (a) 0.0 3.0x10 6 6.0x10 6 9.0x10 6<br />
0<br />
2 2.5 3 3.5 4 4.5<br />
Energy Deposition [eV]<br />
Photon Energy, h ν [eV]<br />
Band Gap [eV]<br />
3.60<br />
3.55<br />
3.50<br />
3.45<br />
3.40<br />
as-grown<br />
35 keV<br />
2 MeV<br />
Figure 3.5: (a) Plots of (αhν) 2 versus photon energy, hν, for the as-grown and the Au<br />
irradiated ZnS films. Intersections of the straight lines with the abscissa, as shown by the<br />
arrows indicate the band gaps of the films. Inset shows the experimental OA spectra of the<br />
samples. (b) Variation of the band gap as a function of total energy deposition in the ZnS<br />
matrix for different irradiation energies.<br />
<br />
(b)<br />
the literature value of 3.54 eV for wurtzite ZnS at room temperature [94]. Therefore,<br />
together with the XRD data, we conclude that the ZnS films were having wurtzite<br />
crystal structure with the diffraction peak, as seen in Fig. 3.3, at (2θ =) 28.58 ◦ coming<br />
from the (002) reflection.<br />
Figure 3.5(b) shows the variation of the band gap, E g , of the irradiated films as<br />
a function of total energy deposition in the ZnS matrix as obtained from the SRIM<br />
simulation [31], for the present experimental conditions. One can see, there is no<br />
change in E g in case of a 35 keV Au irradiated film as compared to the as-grown film.<br />
On the other hand the band gap, E g , decreases to a value as small as about 3.0 eV<br />
for a 100 MeV Au irradiated film. The figure also shows a consistent decrease of E g<br />
with increase in total energy deposition or in other words band gap is red-shifted with<br />
increase in irradiation energy.<br />
It must be mentioned here that in chalcogenide materials the lone pair orbital forms<br />
the valence band, while the conduction band is formed by the antibonding orbital. It<br />
is known that ion irradiation produces lattice damage in the form of point defects such<br />
as vacancies, antisite defects, and interstitials. In such a situation, the reduction in<br />
the band gap with increase in irradiation energy, may arise due to the effect of band<br />
tailing, owing to the defects produced during irradiation. In the high-energy regime,<br />
i.e., at 100 MeV the production of defects is dominated by electronic excitations only
Modification of nanocrystalline ZnS films by Au irradiation 47<br />
and the influence of nuclear energy loss is insignificant. Therefore, the incident ions<br />
excite the electrons from the lone pair and bonding states to higher energy states.<br />
Vacancies created in these states are immediately filled by the outer electrons with<br />
Auger processes that in turn induce more holes in the lone pair and bonding orbital<br />
leading to a vacancy cascade process. In this process, bond breaking or ionization of<br />
atoms is easier to occur which leads to a change in the local structure order causing a<br />
decrease in the optical band gap [108].<br />
3.3.5 Surface morphology<br />
(a)<br />
(b)<br />
(c)<br />
(d)<br />
Figure 3.6: 2×2 µm 2 three-dimensional AFM images of (a) as-grown, (b) 35 keV, (c) 2<br />
MeV, and (d) 100 MeV Au irradiated ZnS surfaces. Vertical scale is 10 nm per division.
Modification of nanocrystalline ZnS films by Au irradiation 48<br />
Figure 3.6 shows the three-dimensional AFM images of the as-grown and the Au irradiated<br />
surfaces of ZnS films deposited on Si(100) substrates. In the image corresponding<br />
to the as-grown film [Fig. 3.6(a)], one can see an almost uniform distribution of small<br />
structures on the surface. Three other images corresponding to the surface topography<br />
of the Au (35 keV, 2, and 100 MeV) irradiated films are shown in Figs. 3.6 (b), (c),<br />
and (d), respectively. One can see there is clear variation in the film surface roughness.<br />
The 35 keV Au irradiated surface shows a clear growth in surface structures with<br />
respect to the unirradiated sample. The surface features in the form of nano-hillocks<br />
are seen to be the largest in the 2 MeV irradiation case. Au irradiation at 100 MeV<br />
is seen to result in smaller structures as compared to the 2 MeV irradiation case.<br />
Usually the features on as-grown and irradiated surfaces are described through a<br />
height-height correlation function which contains three important roughness parameters:<br />
(i) the vertical correlation length σ rms , (ii) the lateral correlation length η 0 , and<br />
(iii) the roughness exponent α [109]. The lateral correlation length, η 0 , describes the<br />
lateral characteristics of the surface, the roughness exponent, α, describing the static<br />
scaling properties. The most commonly reported parameter of surface roughness i.e.,<br />
σ rms or the root-mean-square (rms) roughness, characterizes the surface, only along<br />
the vertical direction. This is defined as standard deviation of the surface height profile,<br />
h(x, y), at each point (x, y) of a reference surface plane from the mean height<br />
(< h >) as given by,<br />
[ 1<br />
σ rms =<br />
N<br />
] 1/2 N∑<br />
(h i − < h >) 2 , (3.4)<br />
i=1<br />
where N is the number of pixels, h i = h(x, y) being the height at the i th pixel. The σ rms<br />
values have been found to be 0.75, 0.96, 1.49, and 0.91 nm, for the surfaces obtained<br />
with as-grown and Au irradiations as 35 keV, 2, and 100 MeV, respectively. At<br />
lower energies there will be elastic thermal spike induced by nuclear collision cascades,<br />
leading to shock wave [22, 110, 111] and material flow [112, 113] towards the surface.<br />
If F d is the damage energy deposited per unit length and ǫ m is the average energy per<br />
atom at the melting point, then as shown by Averback and Ghaly [114], the parameter<br />
(F D /ǫ) 2 determines whether there would be melting or not. In the present case with<br />
2 MeV Au, F D (can be taken as S n ) has a value close to 3.66 keV nm −1 . Taking the<br />
melting point of ZnS to be 2023 K, one can estimate ǫ m to be 0.5 eV. Thus (F D /ǫ m ) 2<br />
turns out to be 6000 2 which is quite high indicating surface melting. At 100 MeV,<br />
the electronic energy loss dominates leading to inelastic thermal spikes produced in a<br />
cylindrical region around the ion track [33, 115]. This occurs when the energy taken
Modification of nanocrystalline ZnS films by Au irradiation 49<br />
by the electrons goes to the lattice through electron-phonon coupling. In case the<br />
system has lower thermal conductivity as in ZnS there can be localized melting with<br />
evaporation and/or plastic flow [115, 116]. However, this is seen to result in much<br />
reduced surface damage.<br />
However, σ rms is not sufficient to provide the complete information of the surface<br />
modifications. This is because of σ rms gives information along the vertical direction<br />
only, and hence, morphology parallel to surface plane is left out. In view of this, a<br />
power spectral density (PSD) analysis is often used to look at surface features and<br />
their possible origin [117]. The PSD analysis is accomplished by radially averaging<br />
the square magnitude of the coefficients of the two-dimensional Fourier transform of<br />
the digitized surface profile and is defined by [118]<br />
C(q) = 1 L 2 ∫ ∫ d 2 r<br />
2π e−iq.r 〈h(r)〉 2 , (3.5)<br />
where L 2 , q, and h(r) are the scanned area of length L, the spatial frequency, and the<br />
height at the position r = (x, y), respectively.<br />
0.1<br />
PSD [A<br />
4<br />
]<br />
o<br />
0.01<br />
0.001<br />
■<br />
●<br />
▲<br />
▲<br />
as-grown<br />
35 keV<br />
2 MeV<br />
100 MeV<br />
1 10 100<br />
Spatial Frequency [ m ]<br />
µ -1<br />
Figure 3.7: Variation of the PSD<br />
function with the spatial frequency of<br />
ZnS surfaces for as-grown and Au ion<br />
irradiated ZnS films. In the high q region,<br />
the experimental data are fitted<br />
with a function as given in Eqn. 3.6.<br />
The slopes yield the parameter γ for<br />
each case.<br />
The PSD curves of as-grown and irradiated ZnS surfaces are shown in Fig. 3.7.<br />
The variations in magnitude of the PSD function indicate the perturbations on the ion<br />
irradiated surface which increase up to the irradiation energy of 2 MeV and decrease<br />
for 100 MeV. The surface corrugation defined as the slope of a line connecting two<br />
points on the surface, becomes small for points separated by a length longer than<br />
the correlation length, η 0 (= 1/q 0 ). For lengths larger than correlation length the<br />
surface can be considered to be flat. Thus, PSD is expected to be independent of q<br />
for q < q 0 . For lengths smaller than η 0 (i.e. q > q 0 ), corrugations become significant
Modification of nanocrystalline ZnS films by Au irradiation 50<br />
and no special value of q is characteristic of the surface morphology, and hence, the<br />
PSD curve displays a power-law dependence [109]:<br />
C(q) = A q −γ , (3.6)<br />
where A is a constant and γ is the power-law exponent.<br />
The power spectra as shown in Fig. 3.7 can be divided into two distinct regions:<br />
the low frequency part resembles the uncorrelated white noise, while the straight line<br />
high frequency part represents the correlated surface features. The estimated values of<br />
γ, obtained by fitting the experimental data using Eqn. 3.6, are found to be 2.80±0.03,<br />
3.11±0.04, 3.72±0.06, and 2.85±0.03 for as-grown, 35 keV, 2, and 100 MeV Au ion<br />
irradiated ZnS films, respectively. The value of γ is seen to be the highest for the 2<br />
MeV irradiation case decreasing to a lower value for the 100 MeV irradiation.<br />
The evolution of surface morphology during ion irradiation is the result of a balance<br />
between multiple roughening and smoothing processes. There is a stochastic<br />
roughening process due to random arrival of the ions on the surface. The smoothing<br />
mechanisms on the surface include volume diffusion, viscous flow, sputter redeposition,<br />
and sputter removal affected by shadowing processes [109]. However, to know the<br />
dominant mechanism behind the observed morphological changes in the ZnS surfaces,<br />
due to ion irradiation, one can use the values of γ.<br />
Using a simple linear dimensional analysis, Herring [119] has shown that γ has<br />
values of 1, 2, 3, and 4 representing four modes of surface transport viz. viscous flow,<br />
evaporation-condensation, volume diffusion, and surface diffusion, respectively. In the<br />
case of single predominant process, the process can be easily identified. The index, γ,<br />
is further related to the roughness scaling exponent, α, through the relation γ = 2(α<br />
+ 1) [109].<br />
For the as-grown ZnS film, the value of γ is about 2.8 which is close to 3. This<br />
means the surface morphology of the as-grown film is governed by mainly volume<br />
diffusion. This is expected as the films were deposited using thermal evaporation<br />
of ZnS powder. During continuous film growth ZnS molecules evaporated from ZnS<br />
powder seem to be diffusing into the earlier deposited film showing features of volume<br />
diffusion. The 35 keV Au irradiated sample is also expected to show volume diffusion<br />
coming from nuclear collision cascades formed near the surface. For this γ has a value<br />
of 3.11 again close to 3. On the other hand the 2 MeV Au irradiated sample shows a<br />
γ value of 3.7 which is close to 4, indicating surface diffusion. In fact this is expected<br />
since there is a spreading of material on the surface coming from underneath by the
Modification of nanocrystalline ZnS films by Au irradiation 51<br />
elastic thermal spikes. A trace of volume diffusion is also expected here which probably<br />
is the reason why γ is slightly less than 4. On the other hand at 100 MeV, the ZnS<br />
film surface temperature, in a localized region around the ion trajectory, may go well<br />
above the vaporization temperature because of the inelastic thermal spike production<br />
[33]. This would lead to surface evaporation which seems to produce smaller surface<br />
features as compared to the 2 MeV irradiation case. This may be the reason why<br />
the corresponding surface roughness indicate a γ value of 2.8 (close to 3). However,<br />
as can be seen for all the four cases, involving one un-irrdaiated and three irradiated<br />
surfaces the γ values have been found to range between 2.8 to 3.7. This implies the<br />
surface scaling exponent, α has a value lying between 0.4 and 0.86. This indicate all<br />
the surfaces to be self-affine.
Chapter 4<br />
Ejection of ZnS nanoparticles from<br />
ZnS films by Au irradiation<br />
4.1 Introduction<br />
Energetic ion irradiation of solids leads to ejection of atoms, molecules, and/or clusters.<br />
This phenomenon, known as sputtering, takes place due to the energy transfer<br />
to the atoms in the target by the incident ions through electronic interactions as well<br />
as nuclear collision cascades produced in the target. Simple sputtering of atoms coming<br />
from nuclear collision cascades is well described by Sigmund’s linear theory [29].<br />
Recently, enhanced sputtering by ion irradiation has been reported by several groups<br />
[16, 18, 111, 120, 121, 122, 123, 124, 125, 126, 127]. Sigmund’s theory fails to explain<br />
these results and nonlinear effects were considered to explain the cluster ejection processes.<br />
Most of these studies show a monotonically decreasing yield distribution, which<br />
closely follows an inverse power law with increasing cluster size. At present there are<br />
two models that predict power law cluster yield distributions. One of them is based<br />
on shock wave model [22] while the other is based on a thermodynamic equilibrium<br />
description of the cluster formation process [23]. The first one predicts a power law<br />
with an exponent, δ, equal to 5/3 or 7/3, whereas the second one predicts a δ value<br />
close to 7/3. Molecular dynamics (MD) simulations carried out in metals with ion energies<br />
in the keV range do show emission of clusters in the sputtered species following<br />
an inverse power law in size distribution [20, 21]. Simulations also show a break up<br />
in the larger clusters during their flight. The exponent of the final (or metastable)<br />
clusters size distribution has been found to be 1 to 1.7 times the same for the nascent<br />
(or new born) clusters, i.e., the clusters existing just after ejection [21]. Correcting<br />
52
Ejection of ZnS nanoparticles from ZnS films by Au irradiation 53<br />
the results for such an effect the nascent power law exponent comes down closer to 2.<br />
Because of the fact δ is either close to 2 or lower, the shock wave induced emission<br />
model is preferred over the thermodynamic model which predicts a higher value.<br />
Energetic ion irradiation of a non-elemental compound material is expected to<br />
result, with a certain probability, in the formation of localized spikes that would result<br />
in throwing out materials with the same atomic coordination as in the target. To check<br />
for this we have carried out a study of sputtering from ZnS films using 35 keV, 2, and<br />
100 MeV Au ions [111, 120]. In this chapter we report on the observation of ZnS<br />
nanoclusters, collected on Cu grids coated with amorphous carbon films (hereafter<br />
referred as catcher foil), placed in the reflection geometry.<br />
4.2 Experiments<br />
ZnS thin films (thickness of ∼ 600 nm) were deposited on Si(100) substrates (with native<br />
oxide) by thermal evaporation of 99.99 at.% pure ZnS powder (Materials Research<br />
Corporation, New York) using a vacuum coating unit at a working base pressure of<br />
∼7×10 −6 Torr. The thickness and the rate of deposition of the films were measured<br />
in situ using a quartz crystal thickness monitor. This way prepared films were having<br />
wurtzite (002) crystal structure with an optical band gap of 3.56 eV. The thickness<br />
and composition of the as-grown film were measured using Rutherford backscattering<br />
spectrometry, which showed the film were having ∼ 615 nm thick with Zn and<br />
S concentration ratio as one. Details of these studies were discussed in Chapter 3.<br />
After the deposition, the as-grown ZnS films on Si(100) were irradiated with Au at<br />
three different energies viz. 35 keV (Au − ), 2 MeV (Au 2+ ), and 100 MeV (Au 8+ ) at<br />
an incidence angle of ≈ 45 ◦ . In all the cases, the irradiation fluence was varied from<br />
1×10 11 to 1×10 15 ions cm −2 at a beam current density of ∼ 0.04 µA cm −2 . Au irradiations<br />
at 35 keV and 2 MeV were carried out at room temperature, and 100 MeV Au<br />
irradiations were carried out at two different temperatures viz. room temperature and<br />
liquid nitrogen temperature. The particles sputtered during irradiation were collected<br />
on catcher foils, commonly used for transmission electron microscopy (TEM) studies.<br />
The separations between the catcher foils and the sample surfaces were ∼ 0.5 cm for<br />
35 keV as well as for 2 MeV. However, for 100 MeV Au irradiation the separation<br />
between the catcher foil and the sample surface was ∼ 1 cm. In each of the cases the<br />
catcher foil was placed making an angle of ∼ 30 ◦ (with sample surface) in a reflection<br />
geometry as shown in Fig. 4.1. TEM measurements were carried out to study the size
Ejection of ZnS nanoparticles from ZnS films by Au irradiation 54<br />
Au beam<br />
45 0<br />
ZnS film<br />
Catcher<br />
foil<br />
Sputtered<br />
particles<br />
Figure 4.1: Schematic diagram illustrating<br />
the sputtering experiment.<br />
Si substrate<br />
distributions of the sputtered nanoparticles (NPs) collected on the catcher foils using<br />
a JEOL 2010 UHR TEM (operating at 200 kV). Also high resolution TEM (HRTEM)<br />
and selected area electron diffraction (SAED) techniques were employed to determine<br />
the crystal structure of the sputtered NPs.<br />
4.3 Results<br />
4.3.1 Ejection of ZnS NPs due to 2 MeV Au irradiation<br />
Crystal structure of the ejected NPs<br />
a<br />
b<br />
Figure 4.2: HRTEM micrographs of the NPs corresponding to 2 MeV Au irradiation to the<br />
fluences of (a) 1×10 11 and (b) 5×10 13 ions cm −2 .<br />
Figure 4.2(a) shows an HRTEM micrograph of a large nanocluster on catcher foil,<br />
obtained with a sample irradiated with 2 MeV Au 2+ ions to a fluence of 1×10 11 cm −2 .<br />
For the above low fluence we have observed only very few NPs on the catcher foil.
Ejection of ZnS nanoparticles from ZnS films by Au irradiation 55<br />
The HRTEM image clearly shows the crystalline rows of particles with an inter-planar<br />
spacing of 0.307±0.006 nm which closely agrees with the bulk ZnS lattice spacing for<br />
both the cubic (111) and the hexagonal (002) planes. A typical HRTEM micrograph<br />
taken from a NP corresponding to a higher fluence of 5×10 13 ions cm −2 is shown in<br />
Fig. 4.2(b). The image shows, again, the crystalline rows of the particles with an<br />
inter-planar spacing of 0.310±0.006 nm.<br />
However, to further check the structure, an SAED measurement was carried out on<br />
the catcher foil. The obtained SAED pattern taken on the catcher foil corresponding<br />
to a fluence of 5×10 13 ions cm −2 is shown in Fig. 4.3. The image shows spots and<br />
Figure 4.3: SAED patterns of the NPs<br />
corresponding to 2 MeV Au irradiation<br />
to a fluence of 5×10 13 ions cm −2 . The<br />
diffraction rings corresponding to various<br />
planes are indicated in the figure.<br />
rings which have been identified as corresponding to the (002), (101), (110) and (300)<br />
planes of hexagonal ZnS. These are indicated in the figure. The diffraction rings from<br />
(002) and (101) planes are seen to overlap in the form of a broad ring because of very<br />
close inter-planar spacings. In HRTEM we did not see (300) plane due to very small<br />
d-spacing (0.11 nm) which is smaller than the lattice resolution (0.14 nm) of our TEM.<br />
But in diffraction measurement the spots corresponding to (300) plane are very clearly<br />
seen. Thus the NPs on the catcher foil were unambiguously identified as ZnS having<br />
a wurtzite structure.<br />
Size distribution of the ejected NPs<br />
Plane-view bright-field TEM micrographs of particles collected on the catcher foils,<br />
during Au irradiations of ZnS/Si samples, are shown in Fig. 4.4, where (a), (b), (c),<br />
and (d) correspond to the fluences of 1×10 12 , 5×10 13 , 1×10 14 , and 1×10 15 ions cm −2 ,<br />
respectively. The corresponding size distributions are shown in same figure in the right<br />
panels (e), (f), (g), and (h), respectively. In all cases the particle size is seen to vary
Ejection of ZnS nanoparticles from ZnS films by Au irradiation 56<br />
Frequency (%)<br />
40<br />
30<br />
20<br />
10<br />
(e)<br />
0<br />
2 3 4 5 6 7<br />
Particle Size / nm<br />
Frequency (%)<br />
40<br />
30<br />
20<br />
10<br />
(f)<br />
c<br />
Frequency (%)<br />
0<br />
2 3 4 5 6 7<br />
Particle Size / nm<br />
40<br />
30<br />
20<br />
10<br />
(g)<br />
d<br />
Frequency (%)<br />
0<br />
40<br />
30<br />
20<br />
10<br />
2 3 4 5 6 7<br />
Particle Size / nm<br />
(h)<br />
0<br />
2 3 4 5 6 7<br />
Particle Size / nm<br />
Figure 4.4: Bright-field TEM micrographs of ZnS NPs collected on catcher foils. (a),<br />
(b), (c), and (d) correspond to the fluences of 1×10 12 , 5×10 13 , 1×10 14 , and 1×10 15 ions<br />
cm −2 with corresponding size distributions shown in (e), (f), (g), and (h), respectively. The<br />
histograms in (e) – (h) have been fitted using a log-normal distribution function (Eqn. 4.1),<br />
which are also plotted in the figure as solid curves.
Ejection of ZnS nanoparticles from ZnS films by Au irradiation 57<br />
from 2 to 7 nm. For a quantitative comparison, the particle size distributions have<br />
been fitted with a log-normal distribution function of the form<br />
f(x) =<br />
( ) 2<br />
1<br />
√ e − ln(x/xc)<br />
√<br />
2w<br />
, (4.1)<br />
2πxw<br />
where x c and w are the most probable size and the width of the size distribution,<br />
respectively.<br />
From the left panels of Fig. 4.4, the particle number density in a given frame<br />
is seen to increase with increase in fluence. For fluences, at and below 5×10 13 ions<br />
cm −2 , there is practically no overlap between observed NPs. The corresponding size<br />
distributions are seen to remain almost the same with a mean size of 3.26 nm and a<br />
width of ∼ 0.35 nm. With increase in fluence, beyond 5×10 13 ions cm −2 , a slight shift<br />
in the mean towards a higher value of ∼ 3.63 nm with a corresponding increase in the<br />
width has been observed. The variation of the width of the particle size distributions<br />
with fluence is shown in Fig. 4.5(a).<br />
The width remains practically the same up to<br />
Width of the particle<br />
size distribution (nm)<br />
0.55<br />
0.5<br />
0.45<br />
0.4<br />
0.35<br />
0.3<br />
(a)<br />
1 10 100 1000<br />
12<br />
Ion Fluence (10 ions cm<br />
-2<br />
)<br />
Surface coverage (%)<br />
80<br />
60<br />
40<br />
20<br />
0<br />
(b)<br />
0 250 500 750 1000<br />
12<br />
Fluence (10 ions cm -2 )<br />
Figure 4.5: Variation of (a) the width of the particle size distribution and (b) the catcher<br />
foil surface coverage by the ejected NPs as a function of fluence.<br />
a fluence of 5×10 13 ions cm −2 , beyond which there is a noticeable rise. The rise in<br />
width and mean size appear to be primarily due to overlapping of sputtered particles<br />
(atoms, molecules and clusters) on the catcher foil. Birtcher et al. [16], in their in-situ<br />
TEM studies of Au NPs, have seen mass flow and growth in size of sputtered NPs<br />
under continuous irradiation of Au thin films with 400 keV Xe ions. At this point it is<br />
also important to note that for fluences in the range of 1×10 12 to 1×10 14 ions cm −2 ,<br />
the percentage of the surface (on the catcher foil) covered by the NPs has been found<br />
to grow linearly from a value of about 5 % to about 32 % [Fig. 4.5(b)]. A similar<br />
behavior has also seen by Birtcher et al. [16]. However, there is a deviation at the
Ejection of ZnS nanoparticles from ZnS films by Au irradiation 58<br />
highest fluence considered where the NPs have been found to cover about 67 % of the<br />
catcher foil area. We argue, the above deviation is due to some changes in surface<br />
topography of the irradiated ZnS film coming from continuous irradiation. In fact, we<br />
have seen a change in surface morphology leading to an increase in surface roughness<br />
of the as-deposited ZnS film after 2 MeV Au irradiation to a fluence of 1× 10 14 ions<br />
cm −2 [Fig. 3.6 (c)]. Such changes are expected to affect the sputtering yield [22].<br />
For example, let us imagine a situation where a high density cascade is formed in a<br />
certain depth under a cone causing the emergence of a shock wave. As it propagates<br />
towards the vortex, the cumulation effect increases the wave energy density. When<br />
it reaches a certain magnitude, the cone top breaks off and flies away leading to a<br />
different sputtering yield as compared to that in case of a plane surface [22].<br />
4.3.2 Ejection of ZnS NPs due to 100 MeV Au irradiation<br />
Crystal structure of the ejected NPs<br />
To check the crystal structures of the ZnS NPs ejected during 100 MeV Au irradiation<br />
at room temperature, we have also carried out some HRTEM measurements.<br />
Figure 4.6 shows an HRTEM micrograph of a NP, deposited on the catcher foil. The<br />
HRTEM image clearly shows lattice planes with an inter-planar spacing of 0.311±0.006<br />
nm. This together with earlier data clearly indicates that the ejected NPs are having<br />
wurtzite ZnS structure.<br />
0.311 nm<br />
Figure 4.6: HRTEM micrograph of a NP<br />
corresponding to 100 MeV Au irradiation<br />
to a fluence of 1×10 13 ions cm −2 .<br />
Size distribution of the ejected NPs<br />
Figure 4.7(a) shows a bright-field planar TEM micrograph of the sputtered particles<br />
on the catcher foil due to 100 MeV Au ion irradiation to a fluence of 1×10 13 ions<br />
cm −2 at room temperature. The observed particles were 2−5 nm in size with a most
Ejection of ZnS nanoparticles from ZnS films by Au irradiation 59<br />
a<br />
Frequency (%)<br />
40<br />
30<br />
20<br />
10<br />
(b)<br />
0<br />
1 2 3 4 5 6<br />
Particle size / nm<br />
Figure 4.7: (a) Bright-field planar TEM micrograph of ZnS NPs collected on catcher foil<br />
during 100 MeV Au irradiation at room temperature. (b) shows corresponding size distribution<br />
of the ZnS NPs.<br />
probable size of about 3 nm [Fig. 4.7(b)]. It was observed that the average number<br />
of NPs per frame (i.e., number density) was about 61. However, the average number<br />
of NPs per frame was about 57 in case of 2 MeV Au irradiation to a fluence of<br />
5×10 13 ions cm −2 . Therefore, a comparison of the TEM images of NPs taken from<br />
the catcher foils corresponding to 2 MeV Au ion irradiation with a fluence of 5×10 13<br />
ions cm −2 [Fig. 4.4(b)] and 100 MeV Au ion irradiation with a fluence of 1×10 13 ions<br />
cm −2 [Fig. 4.7(a)] showed that the particle (number) densities were nearly same even<br />
though the distance between the catcher foil and the sample surface for later case was<br />
two times that in case of 2 MeV.<br />
Figure 4.8: Bright-field TEM micrograph<br />
taken on the catcher foil for<br />
the 35 keV Au irradiation to a fluence<br />
of 1×10 14 ions cm −2 .<br />
It is important to mention here is that we have not observed any NP larger than<br />
1 nm to be present on the catcher foil after 100 MeV Au irradiation of ZnS films kept<br />
at liquid nitrogen temperature. In addition, for the 35 keV Au irradiation case, we<br />
did not observe any NP to be present on the catcher foil. A typical bright-field planar
Ejection of ZnS nanoparticles from ZnS films by Au irradiation 60<br />
TEM images taken on the catcher foil for the above case is shown in Fig. 4.8.<br />
4.3.3 A power law behaviour in nanoparticle size distribution<br />
In order to infer about possible mechanisms behind the NPs ejection, it was decided to<br />
look at the size distributions of the ejected particles. For this, many TEM micrographs<br />
of the same frame size (as shown in Fig. 4.4 and Fig. 4.7) were analyzed for each fluence.<br />
The number of visible particles of different sizes, for each frame, was determined using<br />
the ImageJ software [128]. At 2 MeV, Au irradiation was carried out at four different<br />
fluences, F1, F2, F3, and F4 corresponding to 1×10 12 , 5×10 13 , 1×10 14 , and 1×10 15<br />
ions cm −2 , respectively. Corresponding to these four fluences we have analyzed a total<br />
of 52, 342, 455, and 392 particles of different sizes for obtaining the corresponding size<br />
distributions. However, there is only one fluence, F5, which is 1×10 13 ions cm −2 for<br />
the 100 MeV irradiation in which case there were a total of 223 particles observed in<br />
different frames. The results for various irradiations at different fluences, as mentioned<br />
above, were grouped into 0.5 nm bins to get the yield, Y , as a function of NP volume.<br />
Figure 4.9 shows the results, where the size distributions are plotted taking the<br />
100<br />
Frequency [%]<br />
10<br />
1<br />
0.1<br />
▼ 100 MeV<br />
2 MeV<br />
12 2<br />
◆ 1x10 /cm<br />
13 2<br />
■ 5x10 /cm<br />
14<br />
▲<br />
2<br />
1x10 /cm<br />
15<br />
● 1x10 /cm 2<br />
10 100<br />
Hemisphere Volume [nm 3 ]<br />
Figure 4.9: Measured number of<br />
collected particles (having diameter<br />
≥3 nm) as a function of hemisphere<br />
volume. Experimental data<br />
are fitted, with a function given in<br />
Eqn. 4.2, and shown in the figure as<br />
a straight line for each set of data.<br />
frequency (in terms of % of the measured number of particles) as a function of NP<br />
volume. For the volume calculation all the particles were assumed to be hemispherical<br />
in nature. To avoid any overlapping among the data points the size distributions for<br />
fluences F1, F3, F4 and F5 were shifted in the frequency axis. Following earlier works<br />
on cluster size distributions [16, 18, 21, 22, 23, 121], we have tried to fit our data on
Ejection of ZnS nanoparticles from ZnS films by Au irradiation 61<br />
cluster yield, Y , as obtained for larger sizes, to power law distributions of the form<br />
Y (n) = Y 0 n −δ , (4.2)<br />
where n is the number of atoms in a cluster; Y 0 and δ being fitting parameters.<br />
The power law fits to our experimental data (for the particles of diameter ≥3 nm)<br />
yield 2.56, 2.70, 2.59, 2.14, and 3.45 as the values for the exponent δ for fluences<br />
F1, F2, F3, F4 at 2 MeV and fluence F5 corresponding to irradiation at 100 MeV,<br />
respectively. The errors on the estimated δ values are found to be ≤ ± 0.09. As has<br />
been mentioned in the introduction (Section 4.1), MD simulations of clusters emission<br />
do show fragmentation of clusters during their flight path [21]. Figure 4.10 shows<br />
Figure 4.10: Snapshots from a simulation of sputtering of Au at different times for 20 keV<br />
Xe incident on Au. The labels A and B show clusters that are fragmenting, and C illustrates<br />
late sputtering (a cluster separating from the surface after the displacement cascade has<br />
ended). This figure is adapted from Ref. [21].<br />
snapshots of a MD simulation of sputtering of Au at different times (between 16 and
Ejection of ZnS nanoparticles from ZnS films by Au irradiation 62<br />
19 ps) for 20 keV Xe incident on Au. The fragmentation of nascent clusters has<br />
been shown to have an important effect on the size distribution. It is to be noted<br />
that the experimentally observed power law exponents are valid only for the final<br />
clusters, formed after the break up of nascent clusters in the flight. Therefore it is<br />
inappropriate to compare these experimental values of the exponent δ, as obtained for<br />
the final clusters, to those predicted by theory. However, there is no way to detect<br />
clusters or carryout measurements at the rapid time scale of cluster formation within<br />
few ps of the ion impact. In such a scenario the only way out is to use a correction<br />
factor from MD simulations.<br />
In the MD simulation [21] as mentioned earlier, it has been shown that the ratio,<br />
r, of the exponent δ final for the final clusters size distribution to δ nascent that for the<br />
nascent clusters is equal to 1.2±0.2 for 15 keV Xe ion irradiation on Ag. The same<br />
has been found to be 1.3±0.4 for 20 keV Xe ion irradiation on Au. In yet another<br />
simulation study, carried out using the embedded atom method (EAM) for Ar ion<br />
irradiation on fcc metals with energies between 100 eV and 5 keV, the same ratio r<br />
has been estimated to be 1.3 and 1.4 [20]. Following the assumption made in ref.<br />
[21], that the ratios r have a real physical basis and are transferable to other ion and<br />
substrate types, we have estimated the nascent exponents for the present experiment<br />
taking r=1.3. The estimated values of the nascent exponents are then found to be 1.97,<br />
2.08, 1.99, 1.65, and 2.66 for the cases F1, F2, F3, F4, and F5 respectively. One can<br />
see that both the final and the nascent exponents for the ion fluence F4, corresponding<br />
to 1×10 15 Au cm −2 , are slightly smaller than those for the other fluences at the same<br />
incident ion energy of 2 MeV. This is mainly due to the overlapping of the emitted<br />
NPs on the catcher foil [120]. It is to be noted that the exponent for 2 MeV Au<br />
irradiation, with the exception of the case with the highest fluence, are all close to 2<br />
while that for 100 MeV Au irradiation is distinctly higher.<br />
4.4 Discussion<br />
4.4.1 NP ejection mechanisms<br />
Sputtering from compound targets during ion bombardment can take place in the<br />
form of individual atoms, molecules, and sometimes small clusters consisting of a<br />
few atoms. It has been shown that ion bombardment induces compositional change<br />
in the sputtered surface mainly due to preferential sputtering of one element over
Ejection of ZnS nanoparticles from ZnS films by Au irradiation 63<br />
another and radiation induced segregation [129, 130]. In such a case, one can expect<br />
dissimilar sputtering of constituent atoms, mainly due to their mass and binding<br />
energy differences. In addition, there may be a possibility of having a thermodynamic<br />
reaction between the sputtered atoms on the catcher foils leading to the formation of<br />
non-stoichiometric clusters. In contrast to this, our HRTEM micrographs and SAED<br />
patterns show that the larger NPs correspond to wurtzite ZnS which is stoichiometric.<br />
As has been shown in Fig. 4.2 (a) crystalline NPs of size ∼ 5 nm could be seen<br />
on the catcher foil even at a low irradiation fluence of 1×10 11 ions cm −2 . We must<br />
also mention that the ion current density of 0.04 µA cm −2 amounts to the incidence of<br />
about 10 −3 Au 2+ ions nm −2 sec −1 . This means the observed effect is only due to single<br />
ion impact. At a fluence of 1×10 11 ions cm −2 we would expect much less sputtering of<br />
individual atoms, molecules or small clusters which would be also spatially separated.<br />
Therefore one would not expect any reaction taking place on the catcher foil leading<br />
to the formation of larger ZnS NPs with sizes ∼ 5 nm. In fact, at that fluence we<br />
did not find many of those to get a meaningful size distribution. Further, the ejection<br />
of small clusters can be understood based on a statistical model in which cluster<br />
emission is treated as a more or less uncorrelated ejection of constituent atoms. Since<br />
the statistical probability for independent ejection of many atoms into essentially the<br />
same phase space interval is exceedingly small, it is apparent that clusters containing<br />
more than a few atoms cannot be formed in such a way [18].<br />
Although experimental results regarding cluster ejection, in some cases, could be<br />
quantitatively reproduced by MD simulations, a rigorous theoretical explanation of the<br />
observed power law drop in the size distributions is yet to be found. Further, there<br />
is a strong variation in the exponent, δ, for small and large clusters which cannot be<br />
accounted for in any (presently) available theoretical models.<br />
It must be added here that for 35 keV Au ions in ZnS the nuclear and the electronic<br />
energy losses, S n and S e , respectively, as determined from SRIM simulation [31] are<br />
2.91 and 0.24 keV nm −1 , respectively. For 2 MeV Au ions they are 3.66 and 1.91 keV<br />
nm −1 , respectively. The fact that we could not get any NP on the catcher foil at 35<br />
keV while there were NPs for Au irradiations at 2 MeV means S e plays a dominant role<br />
regarding clusters ejection in ZnS. This is particularly so because S n values are not too<br />
far different in the two cases. The SRIM calculations also show that the displacement<br />
cascades in the ZnS for 35 keV Au ions, incident at 45 ◦ , are very close (≤10 nm) to<br />
the target surface [Fig. 4.11(a)]. This would also imply that within the Bitensky and<br />
Parilis model “spherical” shock waves, formed purely from nuclear collision cascades,
Ejection of ZnS nanoparticles from ZnS films by Au irradiation 64<br />
(a)<br />
35 keV Au in ZnS<br />
(b)<br />
2 MeV Au in ZnS<br />
incident ion trajectory<br />
displaced Zn atoms<br />
displaced S atoms<br />
0<br />
10 20<br />
Target Depth [nm]<br />
0<br />
200 400<br />
Target Depth [nm]<br />
600<br />
Figure 4.11: SRIM simulated collision cascades as a function of depth for (a) 35 keV and<br />
(b) 2 MeV Au in ZnS at an incident angle of 45 ◦ with the target normal.<br />
would originate at depths close to 10 nm or less. Not finding any NP at 35 keV would<br />
also mean these shock waves, if produced, die down almost totally while traversing a<br />
material. On the other hand, with 2 MeV Au ions dense displacement cascades coming<br />
from nuclear collisions will be produced in depths greater than 50 nm [Fig. 4.11(b)].<br />
This does not mean that a 2 MeV Au ion would not undergo a collision generating a<br />
collision cascade in the near surface region. All that it means is that denser collision<br />
cascades, leading to sizable displacements which can result in a shock waves, are<br />
mostly produced at larger depths (> 50 nm). The corresponding shock waves would<br />
certainly not be able to reach the surface. However, 2 MeV Au ions lose about 2 keV<br />
nm −1 through electronic excitations and ionization. With low hole mobility this can<br />
result in Coulomb explosion and “cylindrical” shock waves production in a transverse<br />
direction (around the ion track) in the near surface region. This would happen in the<br />
Bitensky and Parilis model particularly when the incident ion falls at an acute angle<br />
with the sample surface.<br />
At 2 MeV, the Coulomb explosion may result in the production of a thermal spike<br />
[45] resulting in local melting and mass flow onto the surface. If the temperature goes<br />
above that required for vaporization then material in vapor phase can be ejected. Upon<br />
cooling down there may be cluster formation and in such a case the thermodynamic<br />
model of Urbassek would be valid leading to a δ value close to 7/3. Since all values<br />
of δ nascent as obtained at 2 MeV, for various fluences, are just about 2 it is unlikely<br />
that the thermodynamic model is valid. However, at 100 MeV incident energy, S e<br />
is about 16 keV nm −1 . The thermal spike produced because of Coulomb explosion
Ejection of ZnS nanoparticles from ZnS films by Au irradiation 65<br />
at this irradiation energy would result in vaporization of a large chunk of material<br />
around the ion track. The evaporated material can thermalize and undergo cooling<br />
resulting in cluster formation. The δ final value observed in this particular case has<br />
been found to be about 3.45. If we take an r value of 1.5, which can be the case<br />
within errors quoted [21], the δ nascent value can be shown to be almost equal to 7/3.<br />
Therefore the NPs emitted from 100 MeV Au irradiation do seem to go in line with<br />
the thermodynamic equilibrium liquid-gas-type phase transition model rather than<br />
the Bitensky and Parilis model. As mentioned earlier, with a dose 1/5 times that at<br />
2 MeV and the catcher foil set at twice the distance, the emitted particle flux from<br />
the surface at 100 MeV was almost the same as that at 2 MeV. This means at the<br />
same fluence, irradiation at 100 MeV results in an order of magnitude higher flux of<br />
particle emission, as compared to that at 2 MeV. So much of material liberated from<br />
the irradiated surface is expected to form a gas that may expand, cool and condense,<br />
finally fragmenting into clusters which get deposited on the catcher foil.<br />
4.4.2 Temperature dependence of ZnS NP ejection under 100<br />
MeV Au irradiation<br />
The difference in the NPs sputtering with 100 MeV Au ions at two different irradiation<br />
temperatures can be understood using a two component inelastic thermal spike model<br />
(see Section 1.2.4 of Chapter 1). This model predicts a local temperature rise, within<br />
a very short time, during the transfer of energy from the excited electrons to the<br />
lattice atoms via electron-phonon coupling, g. The coupling constant g is inversely<br />
proportional to the thermal conductivity, K(T), and is given by [125]<br />
g = π4 (k B z n l v s ) 2<br />
, (4.3)<br />
18K(T)<br />
where k B is the Boltzman constant, v s the sound velocity, n l the atomic density, and z<br />
the number of electrons participating in the thermal spike. It is observed that thermal<br />
conductivity strongly depends upon temperature [131]. In the case of wurtzite ZnS,<br />
at 80 and 300 K, the values of the thermal conductivities are about 0.95 and 0.27 W<br />
cm −1 K −1 , respectively [131]. Therefore, the thermal conductivity at 80 K, which is<br />
mainly due to electrons, is around 3.5 times that of the same at 300 K i.e.at room<br />
temperature. It therefore means that g, the electron-phonon coupling parameter, will<br />
have lower value at 80 K than the same at room temperature. This means at 80 K<br />
only a small amount of energy will be given to a localized region in the lattice, most
Ejection of ZnS nanoparticles from ZnS films by Au irradiation 66<br />
of the energy being taken away by the surrounding medium. On the other hand,<br />
at room temperature more energy is given to the lattice which stays confined for a<br />
relatively longer time because of lower thermal conductivity resulting in the production<br />
of a thermal spike. This would result in vaporization of material which can result in<br />
clusters ejection following the thermodynamic equilibrium model of liquid-gas type<br />
phase transition.
Chapter 5<br />
Anisotropic deformation of Au<br />
nanoparticles in silica glass by swift<br />
heavy ion irradiation<br />
5.1 Introduction<br />
Metallic nanoparticles (NPs) embedded in silica are interesting materials because of<br />
their nonlinear optical properties with potential technological applications in optoelectronics.<br />
The optical properties of these NPs are mainly determined by their surface<br />
plasmon resonance (SPR) [132]. When metallic NPs are excited by electromagnetic<br />
radiation, their conduction electrons exhibit collective oscillations known as localized<br />
SPRs. The wavelength λ max corresponding to the optical absorption maximum of the<br />
SPR is highly dependent on the size, shape, dielectric properties of the metallic NP,<br />
and their surrounding. Because of the SPR, Ag and Au NPs embedded in optical<br />
media have been found to result in an enhancement of the electric field near their<br />
vicinity, that can be used for pattern transfer and nanolithography. In certain cases,<br />
illuminations with unfocused light allows optically addressing particles either individually<br />
or specific controlled configuration, depending upon wavelength, polarization or<br />
the incident angle of exciting radiation [133].<br />
For anisotropically shaped metal NPs embedded in optical media there is a splitting<br />
of the SPR band with increase in aspect ratio (AR) [12, 134, 135]. For a large<br />
enough AR, one of the plasmon bands, in certain cases, have been shown to shift well<br />
into the infrared, thus enabling the system to be important in applications involving<br />
telecommunication [136]. There is also an enhancement in nonlinear optical response<br />
67
Anisotropic deformation of Au NPs by SHI irradiation 68<br />
when compared to spherically shaped NPs [137].<br />
However, it is difficult to synthesize NPs, embedded in a matrix, with the desired<br />
AR using conventional chemical and physical routes. This is because of the requirement<br />
of minimum surface energy for stability, which is true for spherical particle.<br />
In such a scenario, MeV ion irradiation provides a unique possibility for shaping of<br />
NPs with the desired AR. In these applications, it is usually the electronic energy<br />
loss which results in inelastic thermal spikes that can create channels in the dielectric<br />
medium containing the NPs. In case the spikes are formed in the NPs they<br />
may melt and get squeezed into the ion tracks formed resulting in elongated structures.<br />
In view of this, in recent times, a lot of attention has been paid to study<br />
swift heavy ion (SHI) (E/amu ≥ 1 MeV) irradiation induced modification of NPs<br />
[12, 115, 134, 138, 139, 140, 141, 142, 143]. For example, embedded spherical Co NPs<br />
were found to be elongated under SHI irradiation resulting in anisotropic magnetic<br />
properties [115]. 30 MeV Se irradiation of metallic dielectric colloids with an Au core<br />
and silica shell structure has been found to result in different effects on the core and<br />
the shell. The irradiation turned the spherical silica shells into oblate ellipsoids while<br />
the spherical metal cores transformed into prolate ellipsoids [134]. Similarly, 30 MeV<br />
Cu irradiation has been found to induce an anisotropic plastic deformation in colloids<br />
consisting of a silica core (diameter 300–500 nm) and Au shell (thickness 20–60 nm)<br />
structure. The shape of the spherical colloids changed into oblate ellipsoids, with<br />
the final degree of anisotropy depending upon the ion fluence [142]. Also, embedded<br />
spherical Au NPs were found to be deformed anisotropically under 120 MeV self-ion<br />
irradiation [138]. The elongated Au NPs showed a splitting up of the SPR bands.<br />
In most of these studies, anisotropic deformation of NPs have been reported, without<br />
any growth of the particles. In a very recent work, Mishra et al. have shown<br />
growth of Au NPs due to 90 MeV Ni irradiation [144]. Their study showed that the<br />
size of Au NP increased with increase in fluence. However, the authors did not observe<br />
any shape deformation of the NPs. In contrast to these, in the present study,<br />
we demonstrate that size and shape of Au NPs in silica matrix can be controlled,<br />
simultaneously, using SHI irradiations.<br />
5.2 Experiments<br />
Au NPs were synthesized, inside high purity silica glass, by implantation of 32 keV<br />
Au − ions to a fluence of 4×10 16 cm −2 . The low energy implantations were performed
Anisotropic deformation of Au NPs by SHI irradiation 69<br />
at normal beam incidence, at room temperature, using the low energy negative ion<br />
implantation facility available at Institute of Physics (IOP), Bhubaneswar [64]. A set<br />
of as-implanted samples were annealed in air at 850 ◦ C for 1 h for obtaining embedded<br />
Au NPs of a certain reasonable size. Hereafter, the annealed samples will be referred<br />
to as the “as-grown” samples. Three of the as-grown samples were irradiated, at room<br />
temperature, using 100 MeV Au 8+ ions at normal incidence to fluences F2, F5, and F10<br />
with values of 2×10 13 , 5×10 13 , and 1×10 14 ions cm −2 , respectively. This was done<br />
using the 15 MV Pelletron accelerator at Inter University Accelerator Centre, New<br />
Delhi. For comparison, an as-grown samples was irradiated with 10 MeV Au 4+ ions to<br />
a fluence of F10 at IOP, Bhubaneswar. All the samples were subjected to transmission<br />
electron microscopy (TEM) measurements to image the Au NPs inside silica matrix<br />
using a JEOL 2010 UHR TEM (operating at 200 kV) at IOP, Bhubaneswar. The TEM<br />
samples were prepared by the standard preparation techniques as described in Section<br />
2.3.3 in Chapter 2. The Au content in silica, before and after SHI irradiation, was<br />
estimated by Rutherford backscattering spectrometry (RBS) with 1.35 MeV 4 He + ions<br />
using the 3 MV Pelletron accelerator at IOP, Bhubaneswar. The detector was placed<br />
at 135 ◦ with respect to the incident beam direction. Such a glancing angle geometry<br />
was used to improve the surface sensitivity. Optical absorption (OA) measurements<br />
were carried out on the as-grown and the SHI irradiated samples using a Shimadzu<br />
PC3101 UV-VIS-NIR dual beam spectrophotometer, available at IOP. OA spectra<br />
were recorded in the transmission mode with a blank silica glass in the reference line.<br />
5.3 Results<br />
5.3.1 Synthesis of Au NPs in silica glass<br />
A typical bright-field cross-sectional TEM (XTEM) micrograph and the corresponding<br />
NP size distribution for the sample implanted with 4×10 16 ions cm −2 , are shown in<br />
Fig. 5.1(a) and (c), respectively. NPs with larger size (∼ 6 nm) are found to be closer<br />
to the surface, with smaller clusters lying in the deeper region. High resolution TEM<br />
(HRTEM) image of one of the particles, shown in the inset of Fig. 5.1(a) shows that<br />
the NPs are crystalline in nature. The lattice spacing in the NP has been found to be<br />
about 2.2 Å, which is very close to (111) inter-planar spacing of the fcc Au crystal.<br />
Annealing the as-implanted samples in air at 850 ◦ C for 1 h has been found to increase<br />
the NP size, which is shown in Fig. 5.1(b). The corresponding NP size distribution is
Anisotropic deformation of Au NPs by SHI irradiation 70<br />
5 nm<br />
Surface<br />
(a)<br />
20<br />
(c)<br />
Frequency [%]<br />
15<br />
10<br />
5<br />
0.22 nm<br />
0<br />
1 2 3 4 5 6 7 8<br />
Diameter [nm]<br />
(b)<br />
15<br />
(d)<br />
Surface<br />
Frequency [%]<br />
10<br />
5<br />
0<br />
2 4 6 8 10 12 14 16<br />
Diameter [nm]<br />
Figure 5.1: Bright-filed XTEM micrographs of the Au NPs in silica glass after 32 keV Au −<br />
ions impantation to a fluence of 4×10 16 cm −2 (a) before and (b) after annealing in air at<br />
850 ◦ C for 1 h. (c) and (d) show the corresponding size distributions, respectively. Inset in<br />
(a) shows an HRTEM image of a Au NP indicating the particles are crystalline in nature.<br />
shown in Fig. 5.1(d). The XTEM micrograph also shows that there is redistribution<br />
of the NP size and size distribution resulting in a Guassian-like distribution. Fitting<br />
of a Gaussian function to the size distribution leads to a mean size of 6.6±0.2 nm with<br />
a standard deviation of 2.4±0.1 nm. The growth of the NP size is attributed as the<br />
Ostwald ripening process which occurs at the expense of the smaller particles.<br />
As has been mentioned earlier, metal NPs can be probed by the OA spectroscopy.<br />
This is because the OA technique is a very effective tool to measure the metal NP size<br />
and shape very accurately [134, 135, 143] through the SPR. Figure 5.2 shows the OA<br />
spectra for the Au implanted glass samples before and after annealing. The spectrum<br />
corresponding to the as-implanted sample reveals the presence of Au NPs in silica glass<br />
[143] with an SPR peak around 543 nm, which is due to the free electron oscillations in<br />
small metal particles when excited by electromagnetic radiation [145]. Annealing the
Anisotropic deformation of Au NPs by SHI irradiation 71<br />
absorbance (arb. units)<br />
as-implanted<br />
annealed<br />
Figure 5.2: OA spectrum of<br />
Au implanted and annealed silica<br />
glass samples.<br />
320 400 480 560 640 720<br />
wavelength (nm)<br />
as-implanted sample leads to an increase in the SPR peak intensity with a noticeable<br />
reduction in the full width at half maximum (FWHM) of the SPR peak as shown in<br />
Fig. 5.2. Such an enhancement in the SPR peak with a corresponding reduction in its<br />
FWHM is a clear indication of the growth in size of NPs inside the matrix [145].<br />
5.3.2 Shape deformation of the NPs: TEM studies<br />
Figure 5.3(b) shows the bright-field XTEM micrograph of the as-grown sample irradiated<br />
with 100 MeV Au 8+ ions at a fluence F2. Clearly, after irradiation, there is an<br />
elongation of Au NPs along the ion beam direction (with a prolate shape). The image<br />
also shows only bigger NPs are elongated, whereas the smaller ones remain spherical<br />
in shape. For a comparison, XTEM micrograph of Au NPs and their size distribution<br />
corresponding to the as-grown sample is also shown in Fig. 5.3(a) and (e), respectively.<br />
Increasing the fluence to F5 (two and half time that of F2) is seen to result in a further<br />
elongation of the NPs (with increased anisotropy). This is shown in Fig. 5.3(c). The<br />
estimated average ARs of the elongated NPs are about 1.21 and 1.55 for irradiations<br />
with fluences F2 and F5, respectively. The TEM images also show a growth of NPs<br />
with increase in fluence. However, in case of the sample irradiated with the highest<br />
fluence F10, the TEM image shows only spherical NPs with a very large interparticle<br />
distance [Fig. 5.3(d)]. A Gaussian fitting of the corresponding size distribution
Anisotropic deformation of Au NPs by SHI irradiation 72<br />
15<br />
(e)<br />
Frequency [%]<br />
10<br />
5<br />
20 nm<br />
(a)<br />
0<br />
2 4 6 8 10 12 14 16<br />
Diameter [nm]<br />
30<br />
40<br />
30<br />
(f)<br />
Frequency [%]<br />
20<br />
10<br />
Frequency [%]<br />
20<br />
10<br />
(b)<br />
(c)<br />
(d)<br />
Frequency [%]<br />
0<br />
4 8 12 16 20 24<br />
Long Axis Length [nm]<br />
30<br />
25<br />
20<br />
15<br />
10<br />
5<br />
0<br />
8 12 16 20 24<br />
Long Axis Length [nm]<br />
40<br />
Frequency [%]<br />
0<br />
4 8 12 16 20<br />
Short Axis Length [nm]<br />
40<br />
30<br />
20<br />
10<br />
(g)<br />
0<br />
4 8 12 16 20<br />
Short Axis Length [nm]<br />
(h)<br />
10 nm<br />
Frequency [%]<br />
30<br />
20<br />
10<br />
0<br />
4 6 8 10 12 14 16<br />
Diameter [nm]<br />
Figure 5.3: (a)−(d) Bright-filed XTEM micrographs of the as-grown and 100 MeV Au<br />
irradiated samples with fluences of 2×10 13 , 5×10 13 , and 1×10 14 ions cm −2 , respectively;<br />
(e)−(h) show the corresponding size distributions, respectively. Note that (b) and (c) are<br />
having same scale as in (a). Dashed lines are indicating the surfaces.
Anisotropic deformation of Au NPs by SHI irradiation 73<br />
[Fig. 5.3(h)] gives a mean value of 9.2±0.1 nm with a standard deviation of 0.98±0.09<br />
nm. Note that this average Au NP size is slightly larger than that of NPs in the the<br />
as-grown (unirradiated) sample which is about 6.6 nm. Also, for the highest fluence<br />
(F10) irradiated sample, a very narrow distribution is found. This corresponds to a<br />
standard deviation of 0.98 nm, which is much lower than that for NPs in the as-grown<br />
sample which is about 2.4 nm. Interestingly, at the irradiation fluence F10, what has<br />
been seen is a single layer, (i.e., a two-dimensional array) of well separated Au NPs<br />
in the matrix, just below the silica glass surface. This is seen very clearly in a planar<br />
TEM micrograph taken of this sample as shown in Fig. 5.4. The image clearly shows<br />
well separated NPs. The image also clearly shows a contrast difference among the<br />
NPs. The larger NPs are noticeably darker than the smaller NPs. This indicates that<br />
the Au NPs are present almost in the same plane.<br />
20 nm<br />
Figure 5.4: Bright-field planar TEM micrograph<br />
of the 100 MeV Au irradiated<br />
sample for the fluence of 1×10 14 ions<br />
cm −2 .<br />
It is important to mention here that the XTEM images show a decrease in Au<br />
content after SHI irradiation and the Au loss increases with increase in ion fluence. It<br />
is also important to mention here that no observable modification has been seen in the<br />
Au NPs after an irradiation of the as-grown sample by 10 MeV Au ions to fluence of<br />
1×10 14 ions cm −2 (F10). This loss in Au which occurs through an outward movement<br />
and subsequent evaporation is discussed in the following section.<br />
5.3.3 Outward movement and loss of Au: RBS studies<br />
Variation of Au concentration in the depth direction inside silica glass at different<br />
SHI irradiation fluences were determined through an analysis of RBS data. Figure<br />
5.5(a) shows RBS spectra of the samples before and after 100 MeV Au irradiations.<br />
To compare the data for different samples, all the spectra have been normalized using<br />
the silicon profile which is marked in the figure. The figure clearly shows a decrease
Anisotropic deformation of Au NPs by SHI irradiation 74<br />
Normalized Yield [counts]<br />
250 (a)<br />
200<br />
150<br />
100<br />
50<br />
Si<br />
■ ■ ■<br />
▲ ▲ ▲<br />
● ● ●<br />
▼ ▼ ▼<br />
as-grown<br />
13<br />
2x10<br />
13<br />
5x10<br />
14<br />
1x10<br />
Au<br />
0<br />
400 500 600 700 800<br />
Channel Number<br />
Au Sputtered [%]<br />
100<br />
80<br />
(b)<br />
60<br />
40<br />
20<br />
0<br />
0 2 4 6 8 10<br />
13<br />
Fluence [10 ions cm<br />
-2<br />
]<br />
Figure 5.5: (a) RBS spectra taken from the as-grown samples before and after SHI irradiations.<br />
The Au and Si RBS profiles are indicated in the figure. (b) Au sputtered from the<br />
samples after SHI irradiation with different ion fluences.<br />
of backscattered yield corresponding to the Au signal with increase in fluence. The<br />
spectra also show a shift in the Au signal, coming from deep inside the matrix, towards<br />
higher channel number with increase in fluence. The glancing angle geometry for the<br />
RBS measurements helped to observe the small shift in channel number. Decrease<br />
and shift, towards higher channel number, of backscattered yield corresponding to<br />
Au indicate reduction and outward movement of Au towards the silica glass surface.<br />
Figure 5.5(b) shows the variation of Au content inside the silica matrix as a function<br />
of fluence. As can be seen from the figure, there is a gradual increase in the amount of<br />
sputtered Au with increase in fluence. For a SHI irradiation fluence as given by F10,<br />
it is found that almost 90 % of the Au earlier present in the as-grown sample is lost.<br />
5.3.4 Shape deformation of the NPs: OA studies<br />
Figure 5.6 (a), (b), (c), and (d) show the OA spectra corresponding to the unirradiated<br />
(as-grown) and the SHI irradiated samples for the fluences F2, F5, and F10, respectively.<br />
One can clearly see an SPR peak which is very prominent in the unirradiated<br />
sample. From Fig. 5.6, one can see a reduction of SPR peak intensity with increase<br />
in fluence. It is also important to note, from XTEM and RBS data [Figs. 5.3 and 5.5,<br />
respectively], that there is a reduction of Au content in the matrix after irradiation<br />
which increases with fluence. This reduction of the SPR peak intensity is mainly due<br />
to the reduction in the Au content in the matrix. Clearly, OA spectrum corresponding<br />
to a fluence F5 shows two distinct SPR peaks. In fact, the spectrum for fluence F2 (as<br />
will be shown later), is a combination of two peaks which are found to be separated
Anisotropic deformation of Au NPs by SHI irradiation 75<br />
(a)<br />
(b)<br />
Absorbance [arb. unit]<br />
(c)<br />
(d)<br />
Figure 5.6: OA spectra<br />
taken on the (a)<br />
as-grown and SHI irradiated<br />
samples corresponding<br />
to the fluences<br />
of (b) 2×10 13 , (c)<br />
5×10 13 , and (d) 1×10 14<br />
ions cm −2 .<br />
400 500 600 700 800 400 500 600 700 800<br />
Wavelength [nm]<br />
at fluence F5. On the other hand the OA spectrum corresponding to a fluence F10<br />
shows a single peak (corresponding to nearly spherical NPs) of very low intensity. For<br />
quantitative analysis, we have fitted all the OA spectra using Gaussian function(s)<br />
riding over a linear background. One such fitting as obtained for the sample irradiated<br />
with Au fluence F2, is shown in Fig. 5.7. The fitting parameters, viz. number of<br />
SPR peaks and peak positions together with their FWHMs for different samples are<br />
listed in Table 5.2.<br />
Table 5.2: Fitting parameters obtained from the optical absorption spectra for as-grown<br />
and irradiated Au NPs in silica glass.<br />
Fluence Peak position FWHM<br />
(ions cm −2 ) (nm) (nm)<br />
as-grown 543 81.2<br />
2×10 13 532 569 73.5 129.9<br />
5×10 13 542 608 65.1 40.9<br />
1×10 14 561 78.3<br />
The deconvoluted spectrum for sample with fluence F2, as shown in Fig. 5.7 show<br />
two peaks. The one at higher wavelength is found to be redshifted with increase in SHI
Anisotropic deformation of Au NPs by SHI irradiation 76<br />
Absorbance [arb. unit]<br />
● ● ● Experimental<br />
Fitted<br />
Figure 5.7: OA spectrum (filled<br />
circles) of the as-grown sample after<br />
2×10 13 ions cm −2 irradiation<br />
along with the fitted (solid line)<br />
curve using two Gaussian functions<br />
with proper background substraction<br />
(dashed lines).<br />
0<br />
400 450 500 550 600 650 700 750<br />
Wavelength [nm]<br />
fluence (showing two clear peaks at F5). There has also been an earlier SHI irradiation<br />
study on Au-core silica-shell particles using 30 MeV Cu ions with a fluence of 2×10 14<br />
cm −2 [134]. The measured optical extinction spectrum showed a broad peak in the<br />
long-wavelength region (600−650 nm). The two peak structure in OA spectrum has<br />
been shown to be a characteristic of elongated Au NPs. Similar two peak structures<br />
in OA spectrum have also been seen in chemically prepared elongated Au NPs [135].<br />
It is also well known that Au nanorods have two plasmon resonance absorptions, one<br />
due to the transverse oscillation of electrons regardless of the aspect ratio, the other<br />
absorption coming from the longitudinal oscillation of the electrons, whose wavelength<br />
maximum depends strongly on the aspect ratio. The present results are in agreement<br />
with this.<br />
5.3.5 Dependence of longitudinal SPR peak position on aspect<br />
ratio<br />
Figure 5.8 shows the variation of the absorption maximum of longitudinal SPR as a<br />
function of average AR. The linear fit of the experimental data (excluding the data<br />
obtained from the sample irradiated with a fluence of F10) is obtained for the following<br />
equation:<br />
λ max = 120.48AR + 422.33, (5.1)<br />
where λ max is the absorption maximum of the longitudinal SPR. In an earlier work,<br />
Link et al. have shown a linear dependence of the absorption maximum of longitudinal
Anisotropic deformation of Au NPs by SHI irradiation 77<br />
λ max [nm]<br />
610<br />
600<br />
590<br />
580<br />
570<br />
560<br />
550<br />
540<br />
13<br />
5x10<br />
1x10 14<br />
2x10<br />
13<br />
as-grown<br />
1 1.1 1.2 1.3 1.4<br />
Aspect Ratio<br />
1.5 1.6<br />
Figure 5.8: Dependence of the absorption<br />
maximum of the longitudinal plasmon<br />
resonance against the average aspect<br />
ratio as determined by XTEM.<br />
The solid line is the linear fit to the<br />
experimental data, except the 1×10 14<br />
ions cm −2 irradiated one.<br />
SPR of Au nanorods as a function of AR, and is given by [135]<br />
λ max = (53.71AR − 42.29)ǫ m + 495.14, (5.2)<br />
where ǫ m is the medium dielectric constant. Taking into account the medium dielectric<br />
constant ǫ m = 2.26 of the surrounding medium (silica glass [146]), Eqn. 5.2 gives<br />
λ max = 121.39AR + 399.57. One can very clearly see that the slope of the linear fit to<br />
the experimental data agrees very well with that obtained from the theoretical estimation.<br />
The shift/intercept can be taken care of through a single point normalization.<br />
5.4 Discussion<br />
From the XTEM and OA data, it is clearly seen that the Au NPs are deformed in<br />
prolate spheroids with their major axis along the ion-beam direction for the lowest<br />
and intermediate fluences (F2 and F5), whereas there are spherical Au NPs in silica<br />
glass after Au irradiation to a fluence of F10. It seems the highest fluence irradiated<br />
sample have gone through all the intermediate stages of the Au NPs as in the F2 and<br />
F5 irradiated samples and finally left out with spherical NPs. Notice that this effect<br />
has never been observed previously. On the other hand, anisotropic deformation has<br />
been seen for Au and Co NPs embedded in SiO 2 after irradiation with Au [138] and<br />
In [115] ions, respectively. Recently, deformation mechanism has also been observed<br />
in Ag NPs implanted in SiO 2 films after irradiation with Si [12]. However, in contrast<br />
with the above we have seen an outward movement and loss of Au in the present<br />
experiment.<br />
In most of these studies, inelastic thermal spike concepts [125] were invoked to
Anisotropic deformation of Au NPs by SHI irradiation 78<br />
explain the observed results. In the thermal spike model, due to the passage of a SHI,<br />
the electronic subsystem is assumed to be excited first which thermalizes in a very short<br />
time and then the excitation energy is transferred to the lattice by electron-phonon<br />
coupling (Section 1.2.4 of Chapter 1). Due to a large energy deposition in a short<br />
time, the material undergoes deformation leading to chemical and physical changes in<br />
nm sized regions. However, the deformation mechanism in all the above cases is not<br />
well understood. For many years, the plastic flow model had been successfully used<br />
to understand the anisotropic deformation of amorphous materials (such as silica)<br />
subjected to ion irradiation [116, 141]. It is shown in the previous works [12, 140]<br />
that the plastic flow model cannot be applied directly to the metallic NPs. In fact,<br />
in the present study, it is expected that each individual Au ion impact in silica leads<br />
to the formation of thermal spike which results in plastic flow of silica [141]. This<br />
plastic flow induces an in-plane stress perpendicular to the ion beam direction [116].<br />
The combined effect of stress [116] and thermal spike [115] in Au-silica nanocomposite<br />
leads to the elongation of Au NPs. If the elongation is only due to the pressure effect,<br />
then the smaller NPs should also elongate, which has not been observed in present<br />
case. Recently, an important refinement to the plastic flow model was proposed by<br />
D’Orléans and co-workers [115]. These authors have invoked the mechanism of track<br />
formation combined with the plastic flow model, assuming that the NPs melts and<br />
flow into the ion track.<br />
The passage of SHI (100 MeV Au) through thin silica layer containing the gold<br />
NPs (the composition of Au, Si, and O has been obtained to be 13%, 29%, and 58%<br />
by fitting the RBS spectrum, corresponding to the as-grown sample, using GISA3.9<br />
code [72]) deposits the electronic energy of 13.5 keV nm −1 [31]. It is known from the<br />
work done by Toulemonde et al. [147] that an electronic energy loss of ∼ 13.5 keV<br />
nm −1 in silica creates latent track with a diameter ∼ 8.5 nm. We, therefore, expect<br />
that 100 MeV Au ions in the present case create the track of diameter about 8.5<br />
nm. The material within the ion tracks goes to the molten state due to thermal spike<br />
produced by the energetic ion for a short duration of time (∼ 10 −12 s). The movement<br />
of Au atoms and smaller particles towards bigger particles leads to Ostwald ripening<br />
of Au NPs in the latent track around the ion path. Normally, it is expected that<br />
the growth of particles should not exceed the track diameter, whereas in the present<br />
case the particle growth is observed to be occurring with fluence little larger than the<br />
track diameter. This can be explained by the fact that the temperature of silica in the<br />
annular region around the ion track is high enough to melt the Au particles transiently.
Anisotropic deformation of Au NPs by SHI irradiation 79<br />
The high temperature of silica leads to Ostwald ripening of Au NPs. As the fluence<br />
increases, the overlapping of latent tracks occurs, which leads to further agglomeration<br />
of Au NPs within overlapped ion tracks. To understand the shape deformation, as<br />
suggested earlier by D’Orléans et al. [115], the molten material can be squeezed out<br />
due to the difference in volume expansion and compressibility between liquid metal<br />
and matrix. This would result in deformation in shape. Therefore, it is more likely<br />
that the NPs grow when their size is smaller than the track size and elongate when<br />
their size is larger than track size. The observed movement of the Au NPs towards<br />
the silica glass surface by the SHI cannot be understood within the framework of<br />
a thermal spike model alone. This is because of the thermal spike model does not<br />
include density and pressure effects [125]. One needs to consider the pressure pulse<br />
coming from fast density changes in the ion track to explain the observed movement of<br />
Au towards silica glass surface. Similar movement of Au, towards the sample surface,<br />
during 90−350 MeV Ar-, Kr-, Xe- and Au-ions irradiation of Au marker layer in NiO<br />
has also been reported [148].<br />
The process of elongation depends on the product of S e Φ, where S e and Φ are<br />
electronic energy loss and fluence, respectively. The ARs achieved in the previous<br />
studies were ∼ 3.5 and ∼ 8.4 for 120 MeV Au (S e =14.1 keV nm −1 and Φ=3×10 13<br />
ions cm −2 ) [138] and 30 MeV Se (S e =6.2 keV nm −1 and Φ=2×10 14 ions cm −2 ) [134]<br />
irradiations of silica-Au nanocomposite, respectively. It is clear from these data that<br />
when the product of S e Φ changes by approximately three times, the AR changes by a<br />
similar factor. In the present case, the values of S e Φ for the fluences of F2 and F5 are<br />
13.5 keV nm −1 ×2×10 13 ions cm −2 and 13.5 keV nm −1 ×5×10 13 ions cm −2 , respectively.<br />
These S e Φ have resulted in the average ARs of 1.21 and 1.55, respectively, which are<br />
quite smaller than that one would expect from the earlier studies [134, 138]. In fact, in<br />
the present case, increasing the product S e Φ by a factor of 2.5 did not lead to increase<br />
the AR by similar factor. We argue that these deviations are mainly due to the loss<br />
of Au from the silica matrix. Increase in Au loss with increase in fluence resulted in<br />
observed lower values of ARs. If there was no Au loss, in the present case, we might<br />
have also been observing a linear relation between S e Φ and AR.<br />
In case of highest fluence, F10, irradiated sample, Au NPs are spherical in shape<br />
instead of anisotropic. These observations can be understood as follows. Continuous<br />
irradiation of SHI leads to reduction of Au content inside the silica matrix (Fig. 5.5),<br />
which resulted in very less amount of Au available to form NPs. It is known that the<br />
elongation of the NPs is observed only when the NP size become larger than the track
Anisotropic deformation of Au NPs by SHI irradiation 80<br />
diameter in the matrix [115]. The average NP size (∼ 9 nm) observed in case of the<br />
highest fluence is comparable with the track size (∼ 8.5 nm), created by 100 MeV<br />
Au ions inside silica glass matrix, and hence NP deformation has not been observed.<br />
Rather, the small amount (∼ 10 % of the implanted material) of Au form the spherical<br />
NPs, from its melt phase, with average size of 9.2 nm.<br />
At a SHI irradiation fluence of F10, a redshift of the SPR peak position has been<br />
observed by about 18 nm as compared to the as-grown sample. Similar redshift of<br />
the SPR band corresponding to Au NPs embedded in silica has also been reported,<br />
recently, by Mishra et al. [144]. Such a redshift in its position with decrease in<br />
FWHM is known to be due to an increase in the size of the NP [143, 135]. In fact, size<br />
distributions of the XTEM data [Figs. 5.3(e) and 5.3(h)] clearly show slight growth<br />
(2.6 nm) of the average size of as-grown Au NPs after SHI irradiation to a fluence of<br />
F10. But such a small growth in size of the Au NPs alone can not account for an<br />
SPR redshift of ∼ 18 nm. This can be easily seen from a simple simulation of the<br />
OA spectra for the Au NPs embedded in silica glass. This simulation were carried<br />
out using the Eqn. 2.7 that uses, NP size, dielectric constant of the host, and the<br />
dielectric constant of the NPs. Figure 5.9 shows the simulated OA spectra of the Au<br />
Absorbance [arb. units]<br />
■ ■<br />
▲ ▲<br />
● ●<br />
SPR peak [nm]<br />
630<br />
600<br />
570<br />
540<br />
2 3 4 5 6<br />
Dielectric constant<br />
d = 9.2 nm, εm= 3.25<br />
d = 9.2 nm, εm= 2.26<br />
d = 6.6 nm, ε m= 2.26<br />
500 550 600 650 700<br />
Wavelength [nm]<br />
Figure 5.9: Simulated OA spectra of<br />
the Au NPs embedded in silica matrix<br />
for different NP sizes (d) and medium<br />
dielectric constants (ǫ m ). Inset shows<br />
the variation of SPR peak position<br />
with dielectric constant (filled circles).<br />
The solid line is the linear fit to the<br />
simulated data.<br />
NPs embedded in silica glass for different NP sizes and medium dielectric constants,<br />
keeping all the other parameters fixed. Details of the calculation is reported elsewhere<br />
[149]. For ǫ m = 2.26, about 2 nm redshift of SPR peak position is seen due to a growth<br />
of NP from 6.6 to 9.2 nm. However, for the NPs of size 9.2 nm, a change in ǫ m from<br />
2.26 to 3.25 leads to a large redshift in SPR peak position. In case of ǫ m = 3.25, the<br />
SPR peak appears at about 561 nm as observed in the present experiment for the
Anisotropic deformation of Au NPs by SHI irradiation 81<br />
fluence of F10. The inset of Fig. 5.9 shows the variation of the SPR peak position<br />
with the dielectric constant for the NPs of size 9.2 nm. A linear increase of SPR<br />
peak position with increase in dielectric constant is seen. Therefore, we attribute the<br />
observed redshift of SPR peak position as a result of the change in dielectric constant<br />
of the matrix after Au irradiation to a fluence of F10. We argue that the dissolution<br />
of Au (in the form of atoms and/or small clusters containing only few atoms) in silica<br />
glass resulted in a change of medium dielectric constant after such high fluence SHI<br />
irradiation. This argument is supported by the XTEM [Figs. 5.3(c) and (d)] and<br />
RBS [Fig. 5.5(b)] data. RBS data showed nearly 10 % of Au lost from the matrix<br />
by increasing the fluence from F5 to F10. However, it seems from XTEM data that<br />
Au loss is more as compared to that obtained by RBS studies for the same increase<br />
in fluence. This is because of RBS is an absolute measurement that includes all the<br />
Au regardless of cluster size. These observations suggested that some of Au is in the<br />
form of atoms and/or small clusters in the matrix which could not be imaged by our<br />
TEM. It is also important to mention here that the highest fluence SHI irradiation<br />
can result in compaction of the silica leading to an increase in density and hence an<br />
increase in dielectric constant of the matrix [150]. This can result in a redshift of the<br />
SPR bands of Au NPs embedded in silica glass.
Chapter 6<br />
Observation of a universal<br />
aggregation mechanism in Au<br />
sputtered by swift heavy ions<br />
6.1 Introduction<br />
The observation of emission of clusters of few atoms and molecules from energetic ion<br />
impact was reported way back in 1958 [17]. This observation was quite surprising<br />
since the cluster binding energies are of the order of 1−2 eV, too small compared<br />
to the energies of the colliding ions. Since then there has been a lot of studies on<br />
the subject with an aim to understand the basic mechanisms behind cluster emission<br />
[16, 18, 19, 111, 120, 123].In most cases, involving low energy ion impact [16, 18, 19],<br />
the cluster yield, Y (n), as a function of number of constituent atoms, n, has been<br />
found to follow an inverse power law, Y (n) ∼ n −δ , δ being a decay exponent. For<br />
small clusters with very few atoms, detected using time-of-flight, δ has been found to<br />
lie between 4 and 8 [19]. However, using transmission electron microscopy (TEM), it<br />
has recently been possible to study the size distribution of large clusters, collected on<br />
catcher foils, placed suitably during ion irradiations [16, 111, 120].<br />
As has been mentioned earlier, on the theoretical side there is a shock-wave model<br />
[22] where overlapping collision cascades, from low energy heavy ion impact, can result<br />
in the production of shock waves in a medium. These can propagate to the surface. In<br />
some cases they can result in emission of a chunk of material with essentially the same<br />
atomic coordination as the target. This mechanism, yields a value of δ close to 2. It<br />
has been argued to be a valid one where irradiation of Au thin films by four different<br />
82
Observation of a universal aggregation mechanism in sputtered Au 83<br />
ions, viz. Ne, Ar, Kr, and Au at energies between 400 and 500 keV has been found<br />
to yield a δ value close to 2, independent of the sputtering yield [16]. There is also<br />
a thermodynamic model [23] where the cascade of atomic displacements, produced<br />
in the near surface region of a target, can thermalize and expand into vacuum. The<br />
temperature is supposed to go beyond a critical temperature, T C . Upon expansion<br />
and cooling the material can undergo a liquid-gas phase transition. A large liquid<br />
drop thus formed can fragment producing a mass distribution with a δ value close to<br />
7/3.<br />
As compared to continuous media (films or bulk), irradiation of nm sized metal<br />
islands or metal nanoparticles (NPs), embedded in a matrix, with a possibility of<br />
melting and evaporation, form a different class of systems. In this chapter, we show<br />
that the size distribution of clusters emitted from such a system, under swift heavy<br />
ion (SHI) irradiation, falls under a universal class of aggregation [151]. These systems<br />
possess non-equilibrium steady state solutions of mass distributions in the form of<br />
inverse power laws. In all such cases there is a competition between aggregation and<br />
breakup or evaporation, a delicate balance between the two leading to a variety of<br />
steady state mass distributions [56] with exponent, δ = 3/2 [57] or 7/2 [63]. Using a<br />
simple two-dimensional lattice model with jump between nearest sites and aggregation,<br />
Takayasu et al. [57] have shown that the asymptotic distribution of mass or size always<br />
follows a power law only with the injection of a unit mass (monomer) at each site.<br />
Without the injection of monomers the solution, in the infinite time limit, corresponds<br />
to an aggregate with all the particles sticking together. Later the analysis was extended<br />
to include both positive and negative values for the dynamical variable which was taken<br />
to be charge rather than mass [58]. The model included both injection and evaporation<br />
(through pair creation injection of unit positive and negative charges). The kinetics of<br />
aggregation were studied using a mean field theory. The system was found to reach a<br />
steady state with a charge distribution following a power law with δ = 3/2 [58]. This<br />
happens to be a very general case corresponding to a broad class of phenomena. As<br />
shown by Bonabeau et al. [60, 62], fish schools, with breakup and injection, show a<br />
similar aggregation, the size distribution showing a decay with δ = 3/2. This is seen<br />
even in economics related to distribution of wealth [61]. Such a result has also been<br />
obtained by Majumdar et al. [59] for aggregation in a mass conserving mean field type<br />
site-site interaction model. It has also been shown that small changes in the breakup<br />
parameters do not affect the decay exponent [60, 61]. There is however a cutoff size<br />
which depends upon the competition between aggregation and breakup, and the time
Observation of a universal aggregation mechanism in sputtered Au 84<br />
scales associated with them.<br />
Here we show, Au atoms, sputtered from vaporized Au NPs, follow a steady state<br />
aggregation as mentioned above. The observed mass distribution is found to be in the<br />
form of a truncated power law with a decay exponent of 3/2. Beyond a critical size<br />
there is a drop off which appears to be again in the form of a power law with a much<br />
larger exponent. Under the present irradiation conditions, the temperature of the NPs<br />
is known to go well above the vaporization temperature. But the mass distribution<br />
shows power law decays quite different from what has been suggested for a liquid-gas<br />
type phase transition [23]. Since the system indicates a steady state scenario there<br />
is no need to correct the mass distribution against any breakup effects as applicable<br />
for cluster emission at lower irradiation energies [21]. The results also indicate the<br />
thermal spike production from electronic energy loss to be an essential requirement<br />
for the present observations.<br />
6.2 Experiments<br />
For the present study, nearly spherical Au NPs were synthesized in silica glass, very<br />
near to the matrix surface, by implanting 32 keV Au − ions to a fluence of 4×10 16<br />
cm −2 followed by annealing in air at 850 ◦ for 1 h. The Au implantation was carried<br />
out using a low energy negative ion implantation facility at the Institute of Physics<br />
(IOP), Bhubaneswar. A bright-field cross-sectional TEM micrograph of the silica<br />
glass sample containing Au NPs is shown in Fig. 6.1(a). One can see a buried layer of<br />
(a)<br />
Surface<br />
Frequency [%]<br />
15<br />
10<br />
5<br />
(b)<br />
0<br />
2 4 6 8 10 12 14 16<br />
Diameter [nm]<br />
Figure 6.1: Bright-field XTEM micrographs of the Au NPs in silica glass after 32 keV Au −<br />
ions implantation to a fluence of 4×10 16 cm −2 followed by annealing in air at 850 ◦ for 1 h.<br />
(b) show the corresponding size distribution with a Gaussian fitting.
Observation of a universal aggregation mechanism in sputtered Au 85<br />
nearly spherical NPs with particle size decreasing from about 15 to 2 nm, progressively<br />
with depth. Figure 6.1(b) shows the corresponding size distribution. Hereafter, these<br />
samples will be referred to as the “target”. Three of the targets were irradiated, at<br />
room temperature, using 100 MeV Au 8+ ions at normal incidence to fluences F2, F5,<br />
and F10 with values of 2×10 13 , 5×10 13 , and 1×10 14 ions cm −2 , respectively. This was<br />
done using the 15 MV Pelletron accelerator at Inter University Accelerator Centre,<br />
New Delhi. For comparison, a target was irradiated with 10 MeV Au 4+ ions to a<br />
fluence as given by F10 using the 3 MV Pelletron accelerator at IOP, Bhubaneswar.<br />
During the irradiation of the targets, sputtered particles were collected using catcher<br />
foils (in the form of carbon coated Cu grids usually used for TEM studies) placed at a<br />
distance of ∼ 1 cm in front of the target at an angle of ∼ 15 ◦ with the sample surface.<br />
The schematic diagram of the sputtering experiment is shown in Fig. 6.2. In each case,<br />
Sputtered<br />
particles<br />
Au beam<br />
Catcher<br />
foil<br />
Au NPs in<br />
silica glass<br />
Silica glass<br />
Figure 6.2: Schematic diagram illustrating<br />
the sputtering experiment.<br />
both involving 10 and 100 MeV Au irradiations, the ion beams were raster scanned over<br />
an area of 1×1 cm 2 for uniform irradiation. All the Au implantation and irradiations<br />
were carried out at room temperature. The the catcher foils with collected particles<br />
were imaged using a JEOL 2010 UHR TEM operating at 200 kV. High resolution<br />
TEM (HRTEM) and selected area electron diffraction (SAED) measurements were<br />
also carried out to infer about the crystal structure of the NPs. The Au content in the<br />
targets were checked before and after SHI irradiation using Rutherford backscattering<br />
spectrometry (RBS) employing with 1.35 MeV 4 He + ions using the 3 MV Pelletron<br />
accelerator at IOP, Bhubaneswar. A surface barrier detector was placed at an angle of<br />
135 ◦ with respect to the incident beam direction to collect the backscattered particles.
Observation of a universal aggregation mechanism in sputtered Au 86<br />
(a)<br />
(b)<br />
(c)<br />
Surface Coverage (%)<br />
8<br />
7<br />
6<br />
5<br />
4<br />
3<br />
2<br />
(d)<br />
2 4 6 8 10<br />
13<br />
Fluence [10 ions cm<br />
-2]<br />
Figure 6.3: Bright-field planer TEM micrographs of the ejected Au particles on the catcher<br />
foils collected during ion irradiation with irradiation fluences of (a) 2×10 13 , (b) 5×10 13 , and<br />
(c) 1×10 14 ions cm −2 , respectively. (d) Coverage of the catcher foils surface by the ejected<br />
NPs. Solid line is the linear fit to the experimentally obtained data.<br />
6.3 Results<br />
6.3.1 Ejection of Au NPs: Dependence on ion fluence<br />
Bright-field planer TEM micrographs of the particles collected on the catcher foils<br />
following SHI irradiations to the fluences of 2×10 13 , 5×10 13 , and 1×10 14 ions cm −2<br />
are shown in Figs. 6.3(a), (b), and (c), respectively. The larger particles are noticeably<br />
darker in the micrograph indicating them to be three-dimensional entities. The<br />
particles are seen to have sizes ranging from about 1 to 20 nm. Figures 6.3(a)−(c)<br />
also show an increase in NP number density with increase in fluence. Using different<br />
frames and knowing the size of the NPs it is possible to determine the fraction of the<br />
frame area covered by the observed NPs. Figure 6.3(d) shows this surface coverage in<br />
terms of percentage of area covered in a given frame, by the ejected NPs. The data<br />
shown for each fluence also correspond to an average over several frames. This figure
Observation of a universal aggregation mechanism in sputtered Au 87<br />
shows a linear increase in the coverage with increase in irradiation fluence. In fact,<br />
the surface coverage of Au is seen to increase from about 3% to about 8% in going<br />
from the lowest to the highest irradiation fluence. It must be mentioned here, we have<br />
not observed any NP on the catcher foil following the 10 MeV Au irradiation where<br />
inelastic thermal spike induced sputtering of Au is found to be much reduced.<br />
6.3.2 Crystal structure of the ejected NPs<br />
a<br />
0.219 nm<br />
b<br />
(111)<br />
(002)<br />
Figure 6.4: (a) HRTEM<br />
micrograph of a NP and (b)<br />
SAED pattern of the NPs<br />
collected on the catcher foil<br />
corresponding to a Au irradiation<br />
fluence of 1×10 14<br />
ions cm −2 .<br />
Figure 6.4(a) shows an HRTEM micrograph of an ejected NP on the catcher foil corresponding<br />
to a fluence of 2×10 13 ions cm −2 . The HRTEM image clearly shows the<br />
crystalline nature of the particles with an inter-planar spacing of 0.219±0.004 nm<br />
which closely agrees with Au lattice spacing for cubic (111) plane. Similar HRTEM<br />
measurements on the NPs collected on the catcher foils for other fluences also indicated<br />
crystalline nature of the NPs. However, to further check the structure, SAED<br />
measurements were carried out on the same catcher foil. A typical SAED micrograph<br />
taken on the catcher foil is shown in Fig. 6.4(b). The image shows ring patterns with<br />
dots indicating the crystalline nature of the NPs present. The dots are identified as<br />
corresponding to the (111) and (002) planes of face-centered-cubic Au. These are<br />
indicated in the figure.<br />
6.3.3 Loss of Au content in the target: RBS studies<br />
Figure 6.5 shows the RBS spectra taken from the targets before and after Au irradiations.<br />
To compare the data for different samples, all the spectra have been normalized<br />
using the silicon profile. The figure clearly shows almost no loss of Au in the target<br />
following 10 MeV Au irradiation. However, a decrease of backscattered yield is seen,<br />
in the RBS spectra corresponding to the targets after 100 MeV Au irradiations, with
Observation of a universal aggregation mechanism in sputtered Au 88<br />
Normalized Yield [counts]<br />
250<br />
200<br />
150<br />
100<br />
50<br />
● ●<br />
unirradiated<br />
10 MeV<br />
◆ ◆ 1x10 14<br />
100 MeV<br />
▲ ▲ 13<br />
2x10<br />
■ ■ 13<br />
5x10<br />
▼ ▼ 1x1014<br />
SiO<br />
2<br />
+<br />
He ions<br />
0<br />
725 750 775 800 825<br />
Channel Number<br />
increase in fluence. Decrease of backscattered yield indicates loss of Au in the target.<br />
The Au loss is seen to increase with increase in ion fluence.<br />
6.3.4 A power law behaviour in NP size distribution<br />
Cluster Yield<br />
10<br />
10<br />
10<br />
10<br />
3<br />
2<br />
1<br />
0<br />
Fluence (ions cm -2)<br />
13<br />
■ 2x10<br />
●<br />
13<br />
5x10<br />
▲ 1x10<br />
14<br />
104 5<br />
10<br />
Number of Atoms<br />
45 o<br />
Detector Figure 6.5: The Au part of the<br />
Embedded Au RBS spectra as measured on the<br />
targets before and after the irradiations.<br />
The irradiation fluences<br />
are in the unit of ions cm −2 . Inset<br />
shows the experimental arrangement.<br />
Figure 6.6: Size distribution in<br />
terms of the number of clusters<br />
of particular size (cluster yield)<br />
plotted against the number of<br />
atoms in the cluster. Dotted,<br />
dot-dashed, and continuous<br />
lines correspond to power<br />
law decays with δ values of 3/2,<br />
7/3, and 7/2, respectively.<br />
The size distributions of the NPs collected on the catcher foils, for various SHI fluences,<br />
are shown in Fig. 6.6 where the number of observed Au NPs of different sizes, Y (n),<br />
has been plotted against n, the number of atoms in the n−atom cluster or NP. To<br />
generate these data, many TEM micrographs of the same frame size [as shown in Figs.<br />
6.3(a)−6.3(c)] were analyzed taking each NP to be spherical in shape. The number<br />
of atoms in an n−atom NP is estimated multiplying the volume of the NP by the
Observation of a universal aggregation mechanism in sputtered Au 89<br />
number density of atoms in bulk Au. A total of 1313, 2009, and 2047 particles were<br />
considered for SHI fluences of F2, F5, and F10 respectively. Superimposed on the data<br />
is a straight line representing a power law distribution with a δ value of 3/2 (dotted<br />
line). This is seen to agree with the data very well up to a cutoff size of about 12.5 nm<br />
(corresponding to about 62000 atoms) beyond which evaporation or breakup effects are<br />
dominant. This region corresponding to larger clusters and is affected by fluctuations<br />
because of progressively small number of particles observed. For this region we have<br />
also shown a power law decay with a δ value of 7/2 (continuous line). The data<br />
points seem to follow this behaviour. What is more important is that the data for<br />
three different fluences, taken on three different samples, show the same behavior.<br />
As shown in Fig. 5.3, SHI irradiation results in a change in the size distribution of<br />
the embedded Au NPs. But this does not seem to affect the size distribution of the<br />
sputtered Au NPs. In the following section we present a discussion on various aspects<br />
of the observed phenomenon and the conditions under which such aggregation takes<br />
place.<br />
6.4 Discussion: NP ejection mechanisms<br />
The first and formost requirement is the ejection of Au atoms from embedded Au NPs<br />
under SHI irradiation. This can happen due to the formation of localized inelastic<br />
thermal spikes [33] produced in the NPs resulting in their vaporization. As has been<br />
mentioned earlier, in the inelastic thermal spikes model, passage of a SHI results in<br />
excitation and ionization of electrons in a cylindrical region around the ion path. This<br />
energy at first gets distributed into the electronic system through electron-electron<br />
interactions in a time scale of ∼ 10 −13 s. Electron-lattice interactions cause a part of<br />
this energy to go to lattice atoms resulting in a temperature rise. The temperature<br />
of the thermal spike thus generated depends upon (i) the volume in which energy<br />
imparted by the SHI diffuses due to mobility of the hot electron gas and (ii) the<br />
strength of electron-phonon coupling that determines the efficiency of the transfer of<br />
energy from electronic system to the lattice. Scattering of the excited electrons from<br />
the NP surface and matrix-NP interface will reduce the mobility of the electrons and<br />
will increase the electron-phonon coupling. Further, in small nanostructures density<br />
of structural defects is expected to be very much higher [153] as compared to the bulk<br />
material, which will further contribute to the reduction of the electron mobility in<br />
NPs. Therefore, as compared to the bulk material, the temperature of the thermal
Observation of a universal aggregation mechanism in sputtered Au 90<br />
spike is expected to be much higher in the case of NPs, thus enhancing the effects<br />
of electronic energy loss, S e , leading to a large temperature rise which sometimes can<br />
go beyond the evaporation temperature of the materials. There is also no dissipation<br />
of heat into the insulating surrounding matrix. Simulations indicate this to be true<br />
leading to vaporization of smaller particles [115], some of which get attached to other<br />
NPs leading to Ostwald ripening. Cross-sectional TEM images taken on the targets,<br />
after SHI irradiations at various fluences, do show this (Fig. 5.3). Such a phenomenon<br />
does not seem to happen with 10 MeV Au ions where electronic stopping (2.5 keV<br />
nm −1 ) is way below that at 100 MeV (13.5 keV nm −1 ) [31]. The vaporized material<br />
must also come out of the matrix for the observed aggregation to take place. In silica<br />
glass SHI irradiation can result in a melting of a cylindrical zone around the beam<br />
path which, because of the pressure imbalance [115], can result in a squeezing out of<br />
the evaporated Au atoms (Fig. 5.3).<br />
In such a case, it is not obvious that the expanding system could be in a thermalized<br />
state, with a temperature above a critical value, T C , before going through<br />
a liquid-gas phase transition as proposed by Urbassek [23]. In case it goes through<br />
such a transition, through the liquid gas co-existence region, there would be droplet<br />
formation. The size distribution of the sputtered clusters is then expected to show a<br />
power law decay with a δ value as given by Fisher’s critical exponent, τ, which lies<br />
between 2 and 2.5. For a Van der Waal’s gas the exponent τ has a value of 7/3.<br />
Compared to this, in the present case a δ value of 3/2 has been obtained for smaller<br />
clusters which changes over to 7/2 for larger ones. In fact there is no indication of an<br />
exponent close to 7/3 which is also shown in Fig. 6.6 (dot-dashed line) for comparison.<br />
This rules out, any possible liquid-gas type phase transition taking place in the<br />
vaporized Au system resulting from SHI irradiation. On the other hand, exchange of<br />
particles between nucleation sites, within the framework of a mass-aggregation model,<br />
that takes diffusion, aggregation on contact, and dissociation, can result in a steady<br />
state aggregation process with a δ value of 3/2 [59]. What is more interesting is that<br />
the exponent 3/2 appears in a variety of systems, constituting a universal class where<br />
mass plays the role of a control parameter. Mass conservation with steady injection<br />
of monomers is all that is required with other details of aggregation or breakup not<br />
being important. These results are also in disagreement with simulation results on<br />
cluster emission from Au NPs under self-ion irradiation at lower energies [154] where<br />
electronic stopping effects are neglected.<br />
It is also important to understand the reason behind the cutoff observed in the
Observation of a universal aggregation mechanism in sputtered Au 91<br />
cluster size distribution at about 12.5 nm and the change over taking place at that<br />
point. Such a cutoff can come from splitting of larger clusters [62]. But as long as<br />
aggregation and evaporation rates are independent of mass or size, the δ value, up<br />
to the cutoff, remains 3/2. On the other hand a change over can occur when the<br />
aggregation and evaporation rates become mass dependent. As shown by Vigil et<br />
al. [63], for a critical value of the ratio of the two rates, there can be a transition<br />
between a steady state mass distribution and geletion (infinite mass aggregate). A<br />
power law decay, with a δ value of 7/2, has been shown to occur at the transition<br />
point. It appears, in the present case, at cluster sizes greater than 12.5 nm, somehow<br />
both the aggregation and evaporation rates become mass dependent leading to a phase<br />
transition from one steady state behaviour with a δ value of 3/2 to another with a δ<br />
value of 7/2. In the absence of such a transition the mass distribution, beyond the<br />
cutoff, would have followed the straight line for the power lay decay with a δ value of<br />
3/2. This is expected in a model which does not include splitting [62]. Instead it has<br />
been found to follow a new decay line. At the moment it is not clear as to how such<br />
a transition takes place.<br />
However, based on the present results, it is difficult to rule out the occurrence<br />
of a liquid-gas type phase transition in sputtered material for any general ion-target<br />
combination where the ion energy also plays a crucial role. In fact some experimental<br />
data do exist in support of the above model [155]. What has been shown here is the<br />
existence of a new mechanism of aggregation in sputtered particles, not shown earlier.
Chapter 7<br />
Summary and conclusions<br />
This thesis deals with two set of studies using energetic Au beams. The first set of<br />
studies involve the modification of nanocrystaline ZnS films and ejection of nanoclusters<br />
from them during ion irradiation. The second set involves the modification of<br />
silica embedded Au nanoparticles (NPs) and ejection of nanoclusters from these NPs.<br />
In both cases ejection of NPs have been observed and attempts have been made to understand<br />
the underlying mechanism of emission using the size distribution of emitted<br />
particles.<br />
The first study involves changes in structural, optical, and surface morphological<br />
properties of nanocrystalline ZnS films induced by 35 keV, 2 MeV, and 100 MeV<br />
Au ions. In this, irradiation induced modifications in properties of ZnS thin films,<br />
deposited through a simple thermal evaporation technique, have been studied using X-<br />
ray diffraction (XRD), optical absorption (OA), and atomic force microscopy (AFM).<br />
XRD and OA measurements on the as-grown film indicated the film to have a wurtzite<br />
crystal structure with a grain size and an optical band gap of about 17.5 nm and 3.56<br />
eV, respectively. MeV Au irradiation has been found to result in an increase in grain<br />
size and a decrease in band gap. There is also a change over from compressive stress in<br />
the strating film to tensile stress after MeV Au irradiation. The effects of irradiation<br />
are seen to be stronger at 100 MeV than that at 2 MeV. However, no change in grain<br />
size and band gap have been observed in following irradiation at the keV energy. The<br />
roughness produced on the sample surface have been studied through an analysis of<br />
the power spectral density data as determined from AFM topographies. In the large<br />
spatial frequency, q, (at small length scales) the Fourier exponent, γ, has been found to<br />
have values 2.80, 3.11, 3.72, and 2.85 for the as-deposited and the samples irradiated<br />
with Au at 35 keV, 2 MeV, and 100 MeV, respectively. The values close to 3 could be<br />
92
Summary and conclusions 93<br />
understood in a model where evaporation and condensation dominate the higer value<br />
corresponding to a surface diffusion dominated process.<br />
The effect of Au irradiation on isolated and embedded nanostructures system has<br />
also been studied. For this, spherical Au NPs (of average size 7 nm) were synthesized<br />
in the near surface region of silica glass samples by implanting 32 keV Au − ions to<br />
a fluence of 4×10 16 cm −2 , followed by annealing at 850 ◦ C in air for 1 h. The silica<br />
glass samples so prepared, containing Au NPs are found to show a surface plasmon<br />
resonance (SPR) band centered around 543 nm in the optical absorption spectrum.<br />
The silica glass samples with embedded NPs were subjected to Au irradiations (at<br />
room temperature) at 10 and 100 MeV. The higher energy irradiation was found to<br />
result in an elongation of the embedded Au NPs along the beam direction. At higher<br />
irradiation fluence there was Au loss from the silica matrix. Up to a fluence of 5×10 13<br />
ions cm −2 , the smaller NPs (diameter < 9 nm) are found to grow in size while larger<br />
ones deformed anisotropically along the ion beam direction. The anisotropy in the<br />
larger NPs have been seen to increase with increase in ion fluence. Optical absorption<br />
spectra corresponding to the elongated Au NPs show two SPR bands corresponding to<br />
transverse and longitudinal modes. Further, longitudinal SPR band has been seen to<br />
shift towards higher wavelength with increase in fluence. This shift of the longitudinal<br />
SPR band is a direct consequence of the further elongation of the NPs. However, in<br />
the sample irradiated at a fluence of 1×10 14 ions cm −2 , the Au NPs are seen to be<br />
almost spherical in shape, with a large interparticle separation. The average size is<br />
found to be ∼ 9.2 nm, which is little larger than that of the unirradiated samples. In<br />
fact what is found is a two-dimensional array of small Au NPs, very near the silica<br />
surface which result in an SPR band with a single peak at around 561 nm. This SPR<br />
peak position appeared to be red-shifted as compared to that of the as-grown sample.<br />
Using a simple simulation that uses, NP size, density, and dielectric constant of the<br />
host together with optical density of the NPs, it has been shown that the shift of the<br />
SPR band corresponding to the spherical NPs is, mainly, the result of the change in<br />
matrix dielectric constant, due to the dissolution of Au in the form of atom and/or<br />
small clusters, instead of the growth of the NPs after ion irradiation.<br />
The results on change of shape (from nearly spherical to elongated ones) from ion<br />
irradiation, can be understood within the framework of the inelastic thermal spike<br />
induced ion track formation. Smaller NPs within the ion track completely evaporate<br />
and collect around existing small clusters leading Oswald ripening of small NPs. In<br />
any case they can grow to a size comparable to the track diameter, which is about
Summary and conclusions 94<br />
8.5 nm in silica glass in the present experimental conditions. On the other hand<br />
larger ones can undergo a melting and due to pressure imbalance can get squeezed<br />
to elongated shapes along the ion tracks which are formed along the beam direction.<br />
The material loss at higher fluence occurs from a squeezing out and vaporization of<br />
NPs formed near the surface.<br />
Following the above study, it was also necessary to see whether the material which<br />
is ejected during ion irradiation really contained any nanoclusters. This is because<br />
the size distribution in that case can be connected to the underlying mechanism of<br />
emission. Therefore, in further pursuance of our investigations we have collected the<br />
ejected particles on catcher foils placed suitably in the reflection geometry with an<br />
aim to look at their size distribution. This has been done in both cases concerning<br />
irradiation of ZnS films and embedded Au NPs in silica glass as mentioned earlier.<br />
In each case the size distribution has been determined from transmission electron<br />
microscopy.<br />
For ZnS, with 35 keV Au irradiation no NP larger than 1 nm could be observed on<br />
the catcher foil and therefore no size distribution study could be carried out. However,<br />
we have observed wurtzite ZnS NPs on the catcher foils due to MeV Au irradiations.<br />
The ejected NPs have been found to have sizes (diameter) lying between 2 to 7 nm.<br />
The most probable size is seen to be around 3−3.5 nm. For particle sizes ≥ 3 nm, the<br />
distributions show a power law decay in the form of Y (n) ∼ n −δ ; n and δ being the<br />
number of atoms in a NP and the power law exponent, respectively. In case of Au<br />
irradiation at 2 MeV, the values of exponent, δ, are found to be around 2.60 ± 0.09,<br />
whereas the same for 100 MeV irradiation is found to be around 3.45 ± 0.09. After<br />
correcting the results for cluster breakup effects, the values of power law exponent<br />
reduced to about 2.0 and 2.6 for the above two irradiation energies. A δ value close<br />
to 2 indicates that the NPs are produced when shock waves, generated by subsurface<br />
displacement cascades, ablate the surface. On the other hand, Coulomb explosion<br />
followed by thermal spike induced vaporization of ZnS seems to be the dominant<br />
mechanism regarding material removal at 100 MeV. In such a case the evaporated<br />
material can cool down going into the fragmentation region ejecting clusters with<br />
a power law exponent close to 7/3. Interestingly enough, no NP could be found<br />
in the catcher foil even during the 100 MeV Au irradiation when the sample was<br />
maintained at liquid nitrogen temperature. This suppression on cluster ejection at low<br />
temperature can be explained as due to an enhanced thermal conductivity resulting<br />
in fast heat dissipation that suppresses thermal spike effects.
Summary and conclusions 95<br />
However, the scenario become completely different in case of sputtering from embedded<br />
nanostructures. For 10 MeV Au irradiation of Au NPs, embedded in silica<br />
glass, no NP could be observed on the catcher foils. However, crystalline Au NPs of<br />
size 1−20 nm could be seen on the catcher foils for 100 MeV Au irradiation case. Size<br />
distribution, of the ejected NPs, has been seen to follow a similar power law decay as<br />
seen in the sputtered ZnS NPs. Here, we have seen the existence of two power law<br />
exponents: 3/2 for smaller NPs, below 12.5 nm in size, and 7/2 for larger NPs. The<br />
first case can be rationalized as occurring from a steady state aggregation process,<br />
independent of cluster size. The later case may come from a dynamical transition<br />
to another steady state where aggregation and evaporation rates are size dependent.<br />
To understand the aggregation of clusters during sputtering at very high energy (100<br />
MeV), inelastic thermal spike concept has been invoked. Production of thermal spike<br />
in silica, due to 100 MeV Au ions, can lead to a local temperature rise well above the<br />
vaporization temperature of Au. In such a case evaporated Au clusters can exchange<br />
particle between nucleation sites, within the framework of a mass-aggregation model,<br />
that takes diffusion, aggregation on contact, and dissociation. This can result in a<br />
size independent steady state aggregation process with a δ value of 3/2. This happens<br />
to be a very general case corresponding to a broad class of phenomena, such as fish<br />
schooling, distribution of wealth, cloud formation, and polymer gels. On the other<br />
hand, a phase transition can occur when the aggregation and evaporation rates become<br />
size dependent leading to a δ value of 7/2 at the critical point. Since the system<br />
indicates a steady state scenario there is no need to correct the exponent against any<br />
breakup effects as applicable for the ZnS NPs ejected during Au irradiations.
Bibliography<br />
[1] T.W. Kim, D.C. Choo, J.H. Shim and S.O. Kang, Appl. Phys. Lett. , 80 2168 (2002).<br />
[2] Y.K. Kim, A.J. Danner, J.J. Raftery (Jr.) and K.D. Choquette, IEEE J. Quantum<br />
Elect. 11, 1292 (2005).<br />
[3] B.N. Dev, S. Bera, B. Satpati, D.K. Goswami, K. Bhattacharjee, P.V. Satyam, K.<br />
Yamashita, O.M. Liedke, K. Potzger, J. Fassbender, F. Eichhorn, R. Gröetzschel, Microelectron.<br />
Eng. 83, 1721 (2006).<br />
[4] J.A. Knapp, D.M. Follstaedt and S M. Myers, J. Appl. Phys. 103, 013518 (2008).<br />
[5] M.S. Rafique and N. Tahir, Vacuum 81, 1062 (2007).<br />
[6] N. Martin, J. McCord, A. Gerber, T. Strache, T. Gemming, I. Mönch, N. Farag, R.<br />
Schöfer, J. Fassbender, E. Quandt and L. Schultz, Appl. Phys. Lett. 94, 062506 (2009).<br />
[7] B. Salem, V. Aimez, D. Morris, A. Turala, P. Regreny and M. Gendry, Appl. Phys.<br />
Lett. 87, 241115 (2005).<br />
[8] S. Mohapatra, B. Joseph, D.P. Mahapatr and P. Chakraborty, Nucl. Instrum. Methods<br />
Phys. Res. B 217, 578 (2004).<br />
[9] B. Joseph, S. Mohapatra, B. Satpati, H.P. Lenka, P.K. <strong>Kuiri</strong>, D.P. Mahapatra, Nucl.<br />
Instrum. Methods Phys. Res. B 227, 559 (2005).<br />
[10] S. Amirthapandian, B.K. Panigrahi, S. Rajagopalan, A. Gupta, K.G.M. Nair, A.K.<br />
Tyagi and A. Narayanasamy, Phys. Rev. B 69, 165411 (2004).<br />
[11] K. Awazu, S. Ishii, K. Shima, S. Roorda and J. L. Brebner, Phys. Rev. B 62, 3689<br />
(2000).<br />
[12] A. Oliver, J.A. Reyes-Esqueda, J.C. Cheanf-Wong, C.E. Román-Velázquez, A. Crespo-<br />
Sosa, I. Redríguez-Fernández, J.A. Seman and C. Noguez, Phys. Rev. B 74, 245425<br />
(2006).<br />
96
Bibliography 97<br />
[13] R. Behrisch (Ed.), Sputtering by Particle Bombardment I, II & III, (Springer, Berlin,<br />
Heidelberg, 1983).<br />
[14] J.E. Greene, A. Rockett, and J.-E. Sundgren, Beam-Solid Interactions and Transient<br />
Processes Symposium (Mater. Res. Soc., Pittsburgh, Pennsylvania, U.S.A., 1987), pp.<br />
5973.<br />
[15] A. Benninghoven, F. G. Rüdenauer and H.W. Werner, Secondary Ion Mass Spectrometry:<br />
Basic Concepts, Instrumental Aspects, Applications, and Trends (Wiley, New<br />
York, 1987).<br />
[16] R. Birtcher, S.E. Donnely and K. Nordlund (E. Knystautas, Ed.), Engineering Thin<br />
Films and Nanostructures with Ion Beams (Taylor & Francis, Florida, 2005) Ch. 1.<br />
[17] R.E. Honig, J. Appl. Phys. 29, 549 (1958).<br />
[18] C. Staudt and A. Wucher, Phys. Rev. B 66, 075419 (2002).<br />
[19] C. Staudt, R. Heinrich and A. Wucher, Nucl. Instrum. Methods Phys. Res. B 164/165,<br />
677 (2000).<br />
[20] A. Wucher, Nucl. Instrum. Methods Phys. Res. B 83, 79 (1993).<br />
[21] K.O.E. Henriksson, K. Nordlund and J. Keinonen, Phys. Rev. B 71, 014117 (2005).<br />
[22] I.S. Bitensky and E.S. Parilis, Nucl. Instrum. Methods Phys. Res. B 21, 26 (1987).<br />
[23] H.M. Urbassek, Nucl. Instrum. Methods Phys. Res. B 31, 541 (1988).<br />
[24] J.S. Williams, Rep. Prog. Phys. 49, 491 (1986).<br />
[25] J.F. Ziegler, J.P. Biersack and U. Littmark, The Stopping and Range of Ions in Matter<br />
(Pergamon Press, New York, 1985).<br />
[26] J. Linhard, M. Scharff, Phys. Rev. 124, 128 (1961); J. Linhard, M. Scharff and H.E.<br />
Schiott, Mat. Fys. Medd. Dan. Vid. Selsk 33, 14 (1963).<br />
[27] J.W. Mayer, L. Eriksson and J.A. Davis, Ion implantation in semiconductors (Academic<br />
Press, New York, 1970).<br />
[28] G. Dearnaley, J.H. Freeman, R.S. Nelson and J. Stehan, Ion Implantation, (North-<br />
Holland, Amsterdam, 1973).<br />
[29] P. Sigmund, Phys. Rev. 148, 383 (1969); Phys. Rev. 187, 187 (1969).
Bibliography 98<br />
[30] R. Spohr, Ion Tracks and Microtechnology, principles and applications (Friedr. Vieweg<br />
& Sohn Verlagsgesellschaft mbH, Braunschweig, 1990).<br />
[31] SRIM 2006: a version of TRIM program. This package is available at www.srim.org.<br />
The physical principles of the program and the basic fortran code are given in Ref. [25].<br />
[32] The target atom is displaced in case the transferred energy is larger than the displacement<br />
energy of the atom: Threshold displacement energies in typical solids are of the<br />
order of 10 – 50 eV (see Ref. [34]).<br />
[33] M. Toulemonde, C. Trautmann, E. Balanzat, K. Hjort, A. Weidinger, Nucl. Instrum.<br />
Methods Phys. Res. B 216, 1 (2004).<br />
[34] M. Nastasi and J.W. Mayer, Ion Implantation and Synthesis of Materials, (Springer,<br />
2006).<br />
[35] K. Nordlund, Comp. Mat. Sci. 3, 448 (1995).<br />
[36] J.A. Brinkman J. Appl. Phys. 25, 961 (1954); Am. J. Phys. 24, 246 (1956).<br />
[37] M. Ghaly and R. S. Averback, Phys. Rev. Lett. 72, 364 (1994).<br />
[38] R.L. Fleischer, P.B. Price and R.M. Walker, Nuclear Tracks in Solids (University of<br />
California Press, Berkeley, CA, 1975).<br />
[39] R.E. Johnson and W.L. Brown, Nucl. Instrum. Methods Phys. Res. B 198, 103 (1982).<br />
[40] A. Akkermann, J. Levinson, D. Ilberg and Y. Lifshitz ( R. Baragiola Ed.), in Ionization<br />
of Solids by Heavy Particles (NATO Advanced Study Institutes Series 306, Plenum<br />
Press, New York, 1992) p. 431.<br />
[41] D. Lesueur and A. Dunlop, Radiant. Eff. Deffects Solids 126, 163 (1993).<br />
[42] M. Toulemonde, C. Dufour, and E. Paumier, Phys. Rev. B 46 14362 (1992).<br />
[43] J.G. Fujimoto, J.M. Liu, E.P. Ippen and N. Bloembergen, Phys. Rev. Lett. 53, 1837<br />
(1984).<br />
[44] H.D. Mieskes, W. Assmann, F. Grüner, H. Kucal, Z.G. Wang, M. Toulemonde, Phys.<br />
Rev. B 67, 155414 (2003); M. Toulemonde, W. Assmann, C. Trautmann, F. Grüner,<br />
H.D, Mieskes, H. Kucal, Z.G. Wang, Nucl. Instrum. Methods Phys. Res. B 212, 346<br />
(2003).<br />
[45] E.M. Bringa and R.E. Johnson, Phys. Rev. Lett. 88, 165501 (2002).
Bibliography 99<br />
[46] G. Schiwietz, K. Czerski, M. Roth, F. Staufenbiel and P.L. Grande, Nucl. Instrum.<br />
Methods Phys. Res. B 225, 4 (2004).<br />
[47] W.R. Grove, Phil. Trans. Roy. Soc. London 142, 87 (1852).<br />
[48] W. Jacob, C. Hopf and M. Schlüter Phys. Scr. T124, 32 (2006).<br />
[49] T. Schenkel1, M.A. Briere, H. Schmidt-Böcking, K. Bethge and D.H. Schneider, Phys.<br />
Rev. Lett. 72, 2481 (1997).<br />
[50] T. Neidhart, F. Pichler, F. Aumayr, H.P. Winter, M. Schmid and P. Varga, Phys. Rev.<br />
Lett. 74, 5280 (1995).<br />
[51] M. W. Thompson, Phys. Mag. 4, 139 (1959).<br />
[52] C.T. Reimann, Mat. Fys. Medd. K. Dan. Vidensk. Selsk. 43, 351 (1993) and references<br />
therein.<br />
[53] R.E. Johnson, M. Pospieszalka and W.L. Brown, Phys. Rev. B 44, 7263 (1991).<br />
[54] M.A. Jakas and E. M. Bringa, Phys. Rev. B 62, 824 (2000).<br />
[55] M. M. Jakas, Radiat. Eff. Defects Solids 152, 157 (2000).<br />
[56] W. White, J. Colloid Interface Sci. 87, 204 (1982).<br />
[57] H. Takayasu, I. Nishikawa and H. Tasaki, Phys. Rev. A 37, 3110 (1988).<br />
[58] H. Takayasu, Phys. Rev. Lett. 63, 2563 (1989).<br />
[59] S.N. Majumdar, S. Krishnamurthy and M. Barma, Phys. Rev. Lett. 81, 3691 (1998).<br />
[60] E. Bonabeau and L. Dagorn, Phys. Rev. E 51, R5220 (1995).<br />
[61] H. Yamamoto, T. Ohtsuki, A. Fujihara and S. Tanimoto, J. Phys. Conference Series<br />
31, 59 (2006).<br />
[62] E. Bonabeau, L. Dagorn and P. Fréon, Proc. Natl. Acad. Sci. USA 96, 4472 (1999).<br />
[63] R.D. Vigil, R.M. Ziff and B. Lu, Phys. Rev. B 38, 942 (1988).<br />
[64] B. Joseph, A.K. Behera, S. Mohapatra, P.K. <strong>Kuiri</strong>, D.P. Mahapatra, Solid State Physics<br />
(India) 46, 271 (2003).<br />
[65] K. Sekhar, P.V. Satyam, G. Kuri, D.P. Mahapatra and B.N. Dev, Nucl. Instrum. Methods<br />
Phys. Res. B 73, 63 (1993).
Bibliography 100<br />
[66] D. Kanjilal, S. Chopra, M.M. Narayanan, Indira S. Iyer, Vandana Jha, R. Joshi and<br />
S.K. Datta, Nucl. Instrum. Methods Phys. Res. A 328, 97 (1994).<br />
[67] H. Zhang, Ion sources (Science Press, Beijing, Springer-Verlag, Berlin, 1999).<br />
[68] Details regarding SNICS source can be found in: R. Middleton, Nucl. Instr. Methods<br />
Phys. Res. 214, 139, (1983) or SNICS ion source manual, NEC, USA.<br />
[69] W.K. Chu, J.W. Mayer, M.A. Nicolet, Backscattering spectrometry (Academic Press,<br />
New York, 1978).<br />
[70] M. Mayer, AIP Conference Proceedings, 475, 541 (1999).<br />
[71] RUMP: A simulation package for backscattering spectra for ion beam analysis with<br />
MeV ions, J.W. Mayer, L.R. Doolittle, M.O. Thompson and R.C. Cochran of Cornell<br />
University, developed the original FORTRAN code.<br />
[72] J. Saarilahti and E. Rauhala, Nucl. Instrum. Methods Phys. Res. B 64, 734 (1992).<br />
[73] K. Sekhar, G. Kuri, P.V. Satyam, B. Sundaravel D.P. Mahapatra and B.N. Dev Indian<br />
J. Phys. 684, 1 (1994).<br />
[74] B.D. Cullity, Elements of X-ray Diffraction (Addison-Wesley, Palp Alto, 1978).<br />
[75] D.B. Williams and C.B. Carter, Transmission Electron Microscopy, Vol. I-IV (Plenum<br />
Press, New York, 1996).<br />
[76] S.N. Magonov and M.H. Whangbo, Surface analysis with STM and AFM : Experimental<br />
and theoretical aspects of image analysis (VCH, New York, 1996).<br />
[77] G. Mie, Ann. Phys. 25, 377 (1908).<br />
[78] M. Faraday, Philos. Trans. R. Soc. 147, 145 (1857).<br />
[79] C.F. Bohren and D.R. Huffman, Absorption and Scattering of Light by Small Particles<br />
(Wiley Interscience, New-York, 1983).<br />
[80] N.W. Ashcroft and N.D. Mermin, Solid State Physics (Harcourt College Publishers,<br />
1976).<br />
[81] H. Fröhlich, Physica 4, 406 (1937)<br />
[82] P.B. Johnson and R.W. Christy, Phys. Rev. B 6, 4370 (1972).<br />
[83] H. Hovel, S. Fritz, A. Hilger, U. Kreibig and M. Vollmer, Phys. Rev. B 48, 18178 (1993).
Bibliography 101<br />
[84] F.A. Davis and N.F. Mott, Philos. Mag. 22, 903 (1970).<br />
[85] N.V. Tkachenko, Optical Spectroscopy, Methods and Instrumentation (Elsevier, Amsterdam,<br />
2006).<br />
[86] R.N. Bhargava, D. Gallagher, X. Hong and A. Nurmikko, Phys. Rev. Lett. 72, 416<br />
(1994).<br />
[87] B. Gilbert, B.H. Frazer, H. Zhang, F. Huang, J.F. Banfield, D. Haskel, J.C. Lang, G.<br />
Srajer and G. De Stasio, Phys. Rev. B 66, 245205 (2002).<br />
[88] V.V. Nikesh, A. Dharmadhikari, H. Ono, S. Nozaki, G.R. <strong>Kumar</strong> and S. Mahamuni,<br />
Appl. Phys. Lett. 84, 4602 (2004).<br />
[89] D. Haranath, N. Bhalla, H. Chander, Rashmi, M. Kar and R. Kishore, J. Appl. Phys.<br />
96, 6700 (2004).<br />
[90] I. Sarkar, M.K. Sanyal, S. Kar, S. Biswas, S. Banerjee, S. Chaudhuri, S. Takeyama, H.<br />
Mino and F. Komori, Phys. Rev. B 75, 224409 (2007).<br />
[91] K. Ichino, T. Onishi, Y. Kawakami and S. Fujita, J. Cryst. Growth 138, 28 (1994).<br />
[92] S. Yano, R. Schroeder, B. Ullrich and H. Sakai, Thin Solid Films 423, 273 (2003).<br />
[93] Z.Z. Zhang, D.Z. Shen, J.Y. Zhang, C.X. Shan, Y.M. Lu, Y.C. Liu, B.H. Li, D.X. Zhao,<br />
B. Yao and X.W. Fan, Thin Solid Films 513, 114 (2006).<br />
[94] J. Cheng, D. Fan, H. Wang, B. Liu, Y. Zhang and H. Yan, Semicond. Sci. Technol. 18,<br />
676 (2003).<br />
[95] V. Dimitrova and J. Tate, Thin Solid Films 365, 134 (2001).<br />
[96] P. Prathap, N. Revathi, Y.P.V. Subbaiah and K.T.R. Reddy, J. Phys.: Condens. Matter<br />
20, 035205 (2008).<br />
[97] E. Snoeks, A. van Blaaderen, T. van Dillen, C.M. van Kats, K. Velikov, M.L.<br />
Brongersma and A. Polman, Nucl. Instrum. Methods Phys. Res. B 178, 62 (2001).<br />
[98] S. Chowdhury, A.M.P. Hussain, G.A. Ahmed, F. Singh, D.K. Avasthi and A. Choudhury,<br />
Mater. Res. Bull. 43, 3495 (2008).<br />
[99] D. Mohanta, G.A. Ahmed, A. Choudhury, F. Singh and D.K. Avasthi, J. Nanopart.<br />
Res. 8, 645 (2006).
Bibliography 102<br />
[100] P.K. <strong>Kuiri</strong>, H.P. Lenka, G. Sahu, B.S. Acharya, D. Kanjilal, N.C. Mishra and D.P.<br />
Mahapatra, (to be published) (2010).<br />
[101] Joint Committee of Powder Diffraction Standards (JCPDS), International Center for<br />
Diffraction Data (1601 Park Lane, Swarthmore, PA 1994) p. 19081.<br />
[102] O. Madelung, U. Rössler and M. Schulz (Landolt-Börnstein, Ed. ), Group III Condensed<br />
Matter Numerical Data and Functional Relationships in Science and Technology<br />
Volume 41 (Springer-Verlag, 1999).<br />
[103] R. Sathyamoorthy, S. Chandramohan, P. Sudhagar, D. Kanjilal, D.Kabiraj, and K.<br />
Asokan, Sol. Energy Mater. Sol. Cells 90, 2297 (2006).<br />
[104] D. Bhattacharyya and M.J. Carter, Thin Solid Films 288, 176 (1996).<br />
[105] H.C. Ong, A.X.E. Zhu and G.T. Du, Appl. Phys. Lett. 80, 941 (2002).<br />
[106] L.I. Maissel and R. Glang, Handbook of the Thin Film Technology (Mc Graw Hill, New<br />
York, 1970).<br />
[107] A.A. Kaplanski and R.M. Macforland, Spectroscopy of Solids Containing Rare Earth<br />
Ions (Elsevier Publishers, Amsterdam 1987); A.A. Kaminskii, Crystalline Laser (CRC,<br />
New York 1996).<br />
[108] M.S. Kamboj, G. Kaur, R. Thangaraj and D.K. Avasthi, J. Phys. D: Appl. Phys. 35,<br />
477 (2002).<br />
[109] E.A. Eklund, E.J. Snyder and R.S. Williams, Surf. Sci. 285, 157 (1993).<br />
[110] L.E. Rehn, R.C. Birtcher, S.E. Donnelly, P.M. Baldo and L. Funk, Phys. Rev. Lett.<br />
87, 207601 (2001).<br />
[111] P.K. <strong>Kuiri</strong>, J. Ghatak, B. Joseph, H.P. Lenka, G. Sahu, D.P. Mahapatra, A. Tripathi,<br />
D. Kanjilal and N.C. Mishra, J. Appl. Phys. 101, 014313 (2007).<br />
[112] K. Nordlund, J. Keinonen, M. Ghaly and R.S. Averback, Nature (London) 398, 49<br />
(1999).<br />
[113] R.C. Birtcher and S.E. Donnelly, Phys. Rev. Lett. 77, 4374 (1996).<br />
[114] R.S. Averback and M Ghaly J. Appl. Phys. 76, 3908 (1994).<br />
[115] C. D’Orléans, J. P. Stoquert, C. Estournès, C. Cerruti, J.J. Grob, J.L. Guille, F. Hass,<br />
D. Muller and M. Richard-Pouet, Phys. Rev. B 67, 220101(R) (2003).
Bibliography 103<br />
[116] H. Trinkaus and A.I. Ryazanov, Phys. Rev. Lett. 74, 5072 (1995).<br />
[117] S.G. Mayr and R.S. Averback, Phys. Rev. B 68, 075419 (2003).<br />
[118] R. Petri, P. Brault, O. Vatel, D. Henry, E. Andre, P. Dumas and F. Salvan, J. Appl.<br />
Phys. 75, 7498 (1994).<br />
[119] C. Herring, J. Appl. Phys. 21, 301 (1950).<br />
[120] P.K. <strong>Kuiri</strong>, B. Joseph, J. Ghatak, H.P. Lenka, G. Sahu, B.S. Acharya and D.P. Mahapatra,<br />
Nucl. Instrum. Methods Phys. Res. B 248, 25 (2006).<br />
[121] B. Satpati, J. Ghatak, P.V. Satyam and B.N. Dev, J. Appl. Phys. 98, 064904 (2005).<br />
[122] R. Heinrich and A. Wucher, Nucl. Instrum. Methods Phys. Res. B 140, 27 (1998).<br />
[123] B. Satpati, J. Ghatak, B. Joseph, T. Som, D. Kabiraj, B.N. Dev and P.V. Satyam,<br />
Nucl. Instrum. Methods Phys. Res. B 244, 278 (2005).<br />
[124] H.H. Andersen, A. Brunelle, S. Della-Negra, J. Depauw, D. Jacquet, Y. Le Beyec, J.<br />
Chaumont and H. Bernas, Phys. Rev. Lett. 80 5433 (1998).<br />
[125] H.D. Mieskes, W. Assmann, F. Gruner, H. Kucal, Z.G. Wang and M. Toulemonde,<br />
Phys. Rev. B 67, 155414 (2003).<br />
[126] M. Toulemonde, W. Assmann, C. Trautmann and F. Gruner, Phys. Rev. Lett. 88,<br />
057602 (2002).<br />
[127] J. Eriksson, P. Demirev, P. Hakansson, R.M. Papaleo and B.U.R. Sundqvist, Phys.<br />
Rev. B 54, 15025 (1996).<br />
[128] Available from .<br />
[129] Y. Yuba, S. Hazama and K. Gemo, Nucl. Instrum. Methods Phys. Res. B 206, 648<br />
(2003).<br />
[130] M. Tanemura, M. Ukita and F. Okuyama, Surf. Sci. 426, 141 (1999).<br />
[131] G.A. Slack, Phys. Rev. B 6, 3791 (1972).<br />
[132] W.L. Barnes, A. Dereux and T.W. Ebbesen, Nature 424, 824 (2003).<br />
[133] A.F. Koenderink, J.V. Hernández, F. Robicheaux, L.D. Noordam and A. Polman,<br />
Nano Lett. 7, 745 (2007).
Bibliography 104<br />
[134] S. Roorda, T. van Dillen, A. Polman, C. Graff, A. van Blaaderen and B. J. Kooi, Adv.<br />
Mater. 16, 235 (2004).<br />
[135] S. Link and M.A. El-Sayed, J. Phys. Chem. B 103, 8410 (1999); J. Phys. Chem. B<br />
109, 10531 (2005).<br />
[136] V.A. Podolskiy, A.K. Sarychev and V.M. Shalaev, J. Nonlinear Opt. Phys. Mater. 11,<br />
65 (2002).<br />
[137] K.P. Yuen, M.F. Law, K.W. Yu and P. Sheng, Phys. Rev. E 56, R1322 (1997).<br />
[138] Y.K. Mishra, F. Singh, D.K. Avasthi, J.C. Pivin, D. Malinovska and E. Pippel, Appl.<br />
Phys. Lett. 91, 063103 (2007).<br />
[139] P.K. <strong>Kuiri</strong>, B. Joseph, J. Ghatak, H.P. Lenka and G. Sahu, Adv. Sci. Lett. (accepted)<br />
(2010).<br />
[140] J.-M. Lamarre, Z. Yu, C. Harkati, S. Roorda and L. Martinu, Thin Solid Films 479,<br />
232 (2005).<br />
[141] T. van Dillen, A. Polman, P.R. Onck and E. van der Giessen, Phys. Rev. B 71, 24103<br />
(2005).<br />
[142] J.J. Penninkhof, C. Graf, T. van Dillen, A.M. Vredenberg, A. van Blaaderen and A.<br />
Polman, Adv. Mater. 17, 1484 (2005).<br />
[143] B. Joseph, J. Ghatak, H.P. Lenka, P.K. <strong>Kuiri</strong>, G. Sahu, N.C. Mishra and D.P. Mahapatra,<br />
Nucl. Instrum. Methods Phys. Res. B 256, 659 (2007).<br />
[144] Y.K. Mishra, D.K. Avasthi, P.K. Kulriya, F. Singh, D. Kabiraj, A. Tripathi, J.C.<br />
Pivin, I.S. Bayer and A. Biswas, Appl. Phys. Lett. 90, 073110 (2007).<br />
[145] U. Kreibig and M. Vollmer, Optical Properties of Metal Clusters, Springer Series in<br />
Materials Science Vol. 25 (Springer, Berlin, 1995)<br />
[146] E.D. Palik (Ed.), Handbook of Optical Constants of Solids (Academic Press Inc,<br />
Florida, 1985)<br />
[147] M. Toulemonde, J.M. Costantini, Ch. Dufour, A. Meftah, E. Paumier and F. Studer,<br />
Nucl. Instrum. Methods Phys. Res. B 116, 37 (1996).<br />
[148] W. Bolse and J. Wiesner, Nucl. Instrum. Methods Phys. Res. B 218, 479 (2004).
Bibliography 105<br />
[149] B. Joseph, S. Mohapatra, H.P. Lenka, P.K. <strong>Kuiri</strong> and D.P. Mahapatra, Thin Solid<br />
Films 493, 35 (2005).<br />
[150] P.D. Townsend, P.J. Chandler and L. Zhang, Optical Effects of Ion Implantation (Cambridge<br />
University Press, Cambridge, 1994).<br />
[151] P.K. <strong>Kuiri</strong>, B. Joseph, H.P. Lenka, G. Sahu, J. Ghatak, D. Kanjilal and D.P. Mahapatra,<br />
Phys. Rev. Lett. 100, 245501 (2008).<br />
[152] M. Mayer, SIMNRA User Guide Technical Report IPP 9/113 Max Plank Institut Fur<br />
Plasmaphysik, Garching, Germany.<br />
[153] P.K. <strong>Kuiri</strong>, H.P. Lenka, J. Ghatak, G. Sahu, B. Joseph and D.P. Mahapatra, J. Appl.<br />
Phys. 102, 024315 (2007).<br />
[154] R. Kissel and H.M. Urbassek, Nucl. Instrum. Methods Phys. Res. B 180, 293 (2001).<br />
[155] H.T. Jonkman and J. Michl, J. Am. Chem. Soci. 103, 733 (1981); R.G. Orth, H.T.<br />
Jonkman, D.H. Powell and J. Michl, J. Am. Chem. Soci. 103, 6026 (1981).