Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

International Journal of Advanced Research in Engineering and Technology (IJARET)

Volume 13, Issue 7, July 2022, pp.1-11 Article ID: IJARET_13_07_001


Available online at https://iaeme.com/Home/issue/IJARET?Volume=13&Issue=7
ISSN Print: 0976-6480 and ISSN Online: 0976-6499
DOI: https://doi.org/10.17605/OSF.IO/DVZHT

© IAEME Publication Scopus Indexed

ION BEAMS' HYDRODYNAMIC APPROACH TO


THE GENERATION OF SURFACE PATTERNS
Shak Kareem1and Dr. Syed Shahnawaz Ali2
1Research Scholar, Department of Mathematics, Sri Satya Sai University of Technology and

Medical Science (SSSUTMS), Madhya Pradesh, India


2Professor, Department of Mathematics, Sri Satya Sai University of Technology and Medical

Science (SSSUTMS), Madhya Pradesh, India

ABSTRACT
Amorphous solids can flow given the right timescale. In lead pipes or glaciers, solid
flow can be seen in great detail, but it can also be manipulated by adding flaws. Ion
Beam Sputtering (IBS) is a method in which ions with energies between 0.1 and 10 keV
strike on a solid target, causing the formation of defects and their dynamics as well as
degrading its surface and creating ordered nanostructures.
Despite being technologically intriguing, a fundamental understanding of the
nanopattern creation processes that take place under IBS of amorphizable targets has
not been developed, with recent research on Si having mainly questioned knowledge
collected over the last two decades. In the past, a number of interfacial equations have
been presented to explain these phenomena. Typically, these equations involve non-
systematic addition of various contributions from surface diffusion, ion sputtering, mass
redistribution, etc. In order to create a broad framework into which various
mechanisms (such as viscous flow, stress, diffusion, or sputtering) can be added, under
general physical conservation rules, we here take advantage of the general idea of
solids flowing due to ion impacts. This approach enables a systematic evaluation of the
relevance and interaction of several physical factors influencing surface pattern
generation by IBS as opposed to the formulation of phenomenological interfacial
equations.
Key words: Cement, clinker, corrective material, Mayo Louti sand, standard sand.
Cite this Article: Shak Kareem and Syed Shahnawaz Ali, Ion Beams' Hydrodynamic
approach to the Generation of Surface Patterns, International Journal of Advanced
Research in Engineering and Technology (IJARET), 13(7), 2022, pp. 1-11.
https://iaeme.com/Home/issue/IJARET?Volume=13&Issue=7

1. INTRODUCTION
At least as early as the 1960s, observations of nano-scale patterns on the surfaces of solid tar-
gets that experience ion-beam sputtering (IBS) by ions with energies between 100 eV and 10

https://iaeme.com/Home/journal/IJARET 1 editor@iaeme.com
Ion Beams' Hydrodynamic approach to the Generation of Surface Patterns

keV have been made [1,2]. They either match objects that are amorphized by this kind of
irradiation, such semiconductors, or amorphous materials like glass [3]. Beyond a simple visual
parallel [4,5], the remarkable similarity to macroscopic structures, such as ripples on water or
sand dunes,1 highlighted in [2] highlights the fundamental interest of this class of formations.
Additionally, IBS has a lot of potential applications because to its great efficiency in producing
surface nanopatterns (ripples and dots) over broad areas on top of a range of targets, including
metals and insulators [5,6]. Specifically, in IBS, the complex ion-target interactions are what
cause the erosive process and the appearance of the patterns. Ions do this by inducing material
ejection (sputtering) from the surface, material redistribution, and material flow, as well as
permanent defects in a surface layer whose thickness is on the order of the average penetration
depth [8,9]. Since the late 1980s, Bradley and Harper's (BH) [10] work, which was based on
Sigmund's idea of linear collision cascades, has provided a conceptual framework for
understanding pattern creation in amorphous or amorphizable materials. Physically, a
morphological instability [11] is caused by a faster erosion rate (yield) at surface minima in
comparison to surface peaks, which results in the formation of a pattern for any angle of
incidence and ion energy value. The majority of experimental observations are now supported
by this theoretical paradigm. Minor discrepancies with this viewpoint have been explained by
secondary effects, which has allowed BH's hypothesis to be improved (see reviews in [5,6]).
However, it has been demonstrated that many tests on elemental targets are contaminated
[12], suggesting a connection between pattern development and a sizable concentration of
unwanted species. Due to this, in recent years, clean experiments have been specifically
designed to eliminate this characteristic [13–16], using Si targets as representative examples for
the broad class of substrates that became amorphous when exposed to radiation.
Interestingly, the results of these studies disprove the BH paradigm because different
morphological transitions between unstructured and patterned surfaces are seen as a function
of both incidence angle and ion energy E, supporting earlier observations at higher energies
[17] and defying one of the main predictions of BH theory. Ironically, whereas differential
sputtering and species segregation have likely lately been identified [18–21] as the primary
physical mechanisms causing morphological instability, the a priori simpler case of single-
component systems is still not fully understood.

Figure 1 Water ripples (top) of characteristic wavelength 1 cm vs. Ion Beam Sputtering (IBS) ripples (bottom)
on silicon. Water ripples were obtained by shaking a photography bucket filled with water (courtesy of R. Vida,
Universidad Pontificia Comillas). Silicon ripples were obtained by bombarding Silicon with Ar+. The width of
the observation window is 1 µm2(courtesy of L. Vázquez, Instituto de Ciencia deMateriales deMadrid-CSIC).

https://iaeme.com/Home/journal/IJARET 2 editor@iaeme.com
Shak Kareem and Syed Shahnawaz Ali

Additionally, none of the phenomenological models in use today that build on the BH
paradigm [22–24] can adequately represent the intricate morphological details revealed by
recent experiments on Si. These theories combine several physical mechanisms ad hoc and
disregard defect dynamics while describing the dynamics of the target surface using an effective
evolution equation. Even for more advanced physical models where the evolution of the surface
height is explicitly coupled to that of the density of defects whose transport is restricted to a
thin surface layer, similar to pattern formation on, say, a semiconductor, the same limitations
still apply when faced with experiments. For a current survey of aeolian sand dunes, see, for
example, [25-28] and [7].
We propose a general framework based on general conservation laws (such as those
governing mass or momentum) that allows to study systematically the role of the various
mechanisms involved in the process, in order to overcome the limitations of previous continuum
approaches to surface dynamics by IBS. We will be able to determine the physics underlying
the patterning of the eroded target surface using this method, which is reminiscent of classical
hydrodynamics. In conclusion, the primary contribution of this work is the presentation of a
general consistent physical framework for studying this system, which inherits the advantages
of conventional fluid mechanics issues for which it has proven to be quite effective in the past.

2. GOVERNING EQUATIONS
An impinging ion primarily interacts with the target's nuclear structure in the energy range
under consideration, producing cavities and interstitials. Ions and defects can form clusters,
diffuse at varying rates, or vanish by recombination. The overall impact of these processes is
tripartite and includes amorphizing [29], stressing the target [30], and shifting the mean atom
location within the material [31], as demonstrated by Molecular Dynamics (MD) simulations.
Additionally, as also observed in experiments [3], an amorphous layer quickly forms [32,30],
with its thickness and other mechanical properties like density, p, becoming stationary after a
fluence of order 1014 ions cm-2 (corresponding to a short period of exposure for a typical ion
flux of 1013 to 1015 ions cm-2 s1).A glacier, for example, experiences a time separation
between atomistic relaxation and dynamics because of the modest ion flux that drives it. a time
scale separation between atomistic relaxation rates (≈1ps-1 for collision cascades or
adatomhopping) and the 10 s times in which significant `morphological changes occur. This
fact legitimates a continuum description

2.1. Conservation of Mass


When the amorphous layer has already formed during the blasting process, that is the stage in
which we are most interested. Thus, following the initial transient as described, its density value
does not evolve further even though it differs from that of the pristine target. This fact can be
written mathematically as the following:
∇·V=0, (1)
where V is the velocity field of the fluidized layer, whose components are given by
V = ui + vj -wk(2)
In the language of fluid mechanics, the amorphous layer is said to be incompressible in the
steady state.

2.2. Conservation of Momentum


Conservation of momentum can be universally expressed as [33]

(3)

https://iaeme.com/Home/journal/IJARET 3 editor@iaeme.com
Ion Beams' Hydrodynamic approach to the Generation of Surface Patterns

and T stands for the stress tensor. In fact, the constitutive equation for this tensor includes
the physics of the IBS problem.
There are strong indications that the amorphous layer can be thought of as a highly viscous
fluid, therefore T can be represented as, even though a comprehensive characterisation of the
stress tensor in the blasted material is still lacking.
(4)
the components of the velocity field are u1, b, 3, u, v, and w, and Ts comprises all the terms
that are not related to hydrostatic pressure, pressure, or viscous flow (2).
Additionally, we can safely delete the left hand side of Eq. since the radiation induced
viscosity,, is roughly 15 orders of magnitude greater than the viscosity of water [9]. (3). As a
result, using this supposition for the amorphous layer and the conservation of mass, we can
formulate Eq. (3) as

Figure 2. Analogy between IBS and fluid flow down a inclined plane. By rotating the frame of
references anthelion impacts vertically, the IBS system can be seen as fluid flow down an inclined
plane. The body force b is equivalent to a non-
homogeneousgravityfieldthatdependsonthelocalslopevalueatthefreeinterface.
(5)
where b Ts is described as a body force operating primarily within the fluid layer. The
relevant physics, excluding viscous flow, are really contained in this force. It works because
we don't yet have a complete understanding of the stress that is caused underneath the surface.
The angular contribution, which depends on the local angle of incidence, and the force's
amplitude, fE, can be expected to be separated from each other. Mathematically
(6)
where is the surface's local slope and is the ion beam's incidence angle Here, fE has
dimensions of a gradient of stress and contains coarse-grained information on the impact of the
residual stress formed in the target as a result of ion-induced mass redistribution.
In three dimensions, and using standard spherical coordinates, we can write

https://iaeme.com/Home/journal/IJARET 4 editor@iaeme.com
Shak Kareem and Syed Shahnawaz Ali

the azimuthal angle is where (the angle in the plane of the tar- get with respect to the plane
of incidence of the ions). The disparity in the values of the amplitudes f’E and fE can lead to
possible anisotropies entering the dynamics. The damage caused by the collision cascade and,
eventually, the stress or strain that propels the amorphous layer are assumed to be directed in
the direction of the incoming ion (see Fig. 2). Due to the fact that there are only really two
crucial directions—parallel and perpendicular to the ion flux—fE is the same in the equations
for both x and z. The pre-factor does not change when we project the parallel contribution in
those directions (x and z).
This body force can be thought of as the word for effective gravity. As a result, in Fig. 2,
we use a cartoon to illustrate the comparison between an actual fluid flowing down an inclined
plane and an ion-induced solid flow. Once an angle rotation is completed in the reference frame
used in the lab, the two systems are mathematically equal. The associated gravitational fields
in both systems account for the majority of the differences. While gravity is a constant in the
standard fluid problem, it varies from point to point on the surface in the erosion system because
of the local surface orientation with regard to the ion beam. Consequently, one can acquire a
morphological instability for acceptable angles of incidence in the erosive case as opposed to
the fluid problem.

2.3. Boundary Conditions


Equations (1) through (9) need to be supplemented with appropriate boundary conditions that
account for the tension at the interfaces and the erosive terms. The amorphous-vacuum interface
(the free surface), h(a), and the amorphous-crystalline interface, h,h(a) are the two major system
boundaries for which we must specify the development in the most general situation h(c).
At the free surface, the balance of stress takes the form [33]

t and q are two mutually perpendicular vectors in the tangent plane to the surface, н is the
mean surface curvature, and is surface tension. Where n is the unit outward normal vector. The
projections along the appropriate directions of an external stress that is applied to the surface
are known as T (ext).
We'll assume that the surface tension is constant in this case. However, one might infer that
there is a surface tension gradient for multicomponent materials where one species segregates
at the surface, which may impact the dynamics of the surface (the so-called Marangoni effect
[33,34]).
Finally, a kinematic condition [33] at each interface lead coevolution equations for both
h(a)and h(c)as

https://iaeme.com/Home/journal/IJARET 5 editor@iaeme.com
Ion Beams' Hydrodynamic approach to the Generation of Surface Patterns

where w is the vertical component of the velocity field, as previously stated. The rates of
erosion and amorphization are denoted by the letters jam and jer, respectively. Eqs. (13) and
(14) mathematically describe that the rate at which material is removed from both phases is
equal to the difference between the vertical velocity field and the actual motion of the interface.
For instance, the erosive (BH-type) mechanisms enter through the jer directly, resulting in
an additive contribution to the complete dispersion relation (see Section 3). Since both
interfaces evolve in experiments at the same rates in the steady state, it is assumed for simplicity
that the rates of amorphization and erosion are equivalent. Additionally, we assume that the
crystalline-amorphous interface is flat for mathematical tractability. Under these presumptions,
a final boundary condition—often referred to as the no-slip boundary condition—is necessary,
by which we specify that, at h(c), the fluid travels at the same speed as the crystalline phase.

3. LINEAR ANALYSIS
A surface's linear stability is typically studied using the following three steps:
Finding the flat (zeroth order) interface solution is step one.
2. Making a periodic infinitesimal perturbation with wavelength = 2U/q to the solution
discovered in step one (q being the wavenumber).
3. Figuring out each wavenumber's amplification rate, q. (also known as linear dispersion
relation).
We shall outline these processes for the two-dimensional example in detail for the purpose
of simplicity (namely, the interface is a line and not a proper surface). We will mention the
dispersion relation for the complete three-dimensional example at the conclusion of our
derivation.

3.1. Flat Interface Solutions


We fix our reference frame on the free surface, where we designate by d the average thickness
of the amorphous layer and by h(x) the fluctuations of the free surface, and the crystalline-
amorphous interface is situated at z =- d. H(x) = 0 denotes the flat interface condition. By
include it in the velocity field equations, we discover additional characteristics of the flat
solution, including

To study the morphological stability of the flat interface, we look for solutions of the
governing equation sof the form

https://iaeme.com/Home/journal/IJARET 6 editor@iaeme.com
Shak Kareem and Syed Shahnawaz Ali

for an arbitrary wavenumber q = 2U/λ. Similarly, we assume that the pressure and the
velocity fields can be written as

and perturbatively solve the governing equations to linear order in the small parameter. Eqs
provide the precise form of the subsequent flat interface corrections, P1(z), u1(z), and w1(z).

3.2. Dispersion Relation


Finally, the dispersion relation can be extracted from the kine-matic condition. Specifically,
from Eqs. (14)and(18)weget
∂th=−u(x, 0)∂xh+w(x, 0)+jer→ωq =iqu0(0)+w1(0)+ωq,(22)
where ωq is directly derived from the erosive contribution jer; this is where, for example,
the Bradley and Harper classical theory can be applied [10].
Using Eqs. (15)–(21) and denoting real part with a single prime and imaginary part with a
double prime, we find:

3.2.1 Real Part: Stability of the flat Interface


A periodic disturbance to the flat solution with wavenumber q will grow or degrade
exponentially as stated by ωq whether qr is positive or negative, controlling the stability of the
flat interface (18). Important fluid dynamics findings are generalised in Eq. (23), which has also
been discussed in the context of the IBS problem for more precise function choices [7,35].For
example, it simplifies to Orchard's famous conclusion [36] on the flow of a viscous layer of any
thickness on top of an immobile substrate for the material fe-0. In particular, qd » 1 therefore
results in Mullins' [37] famed rate of flattening when relaxation occurs through bulk viscous
flow, ωq =− q/2µ which has already been used for various IBS experiments [38,39].
Eq. (23) is typically not polynomial, which suggests non-local influences in the dynamics
of the amorphous layer. When other non-local effects are present in the dynamics, this non-
locality might be significant (as redeposition, shadowing, or secondary ion scattering).
Additionally, the sign of ωq’ in our situation may vary with the incidence angle, θ, explaining
morphological changes from flat to patterned structures for various values of θ [13–16].

3.2.2. Imaginary part: In-plane propagation of the pattern


In those cases in which an unstable wavelength is selected(namely, ω’q >0 for some q), the
imaginary part of ωq provides information about the velocity of lateral in-plane propagation of
the pattern. Specifically, if the most unstable wavenumber is qc(namely, the one with the largest
positive value of ω’q), then the pattern propagates with velocity

https://iaeme.com/Home/journal/IJARET 7 editor@iaeme.com
Ion Beams' Hydrodynamic approach to the Generation of Surface Patterns

4. CONNECTION WITH OTHER CONTINUUM FRAMEWORKS


4.1. Two-field “hydrodynamic” theory
We want to point out that, even though we haven't discussed the dynamics of the amorphous-
crystalline interface, one can simplify the current full hydrodynamical formulation in the
shallow-water limit by referring to earlier theories of erosion, particularly the two-field model
approach pioneered in [43] and further developed in [25-28], see an overview in [7].
This method clearly describes the dynamics of the density R of moving species and the free
surface height, h, under the implicit presumption that the thickness of the surface layer where
material transport occurs is insignificant.
Proper rate equations, which were taken from the study of pattern development on Aeolian
sand dunes, can be used to describe the evolution of these two fields. Specifically,

(24)
(25)
where (1- ф ) represents the percentage of atoms that are really sputtered after being
knocked out of their equilibrium locations. Here, ex and ad represent the rates of atom addition
to and excavation from the (crystalline) immobile bulk, respectively.
R=h(a)−h(c), (26)
Then

Finally we just to define

To recover equation 24 exactly

5. CONCLUSION
In this study, we present a physical framework for IBS-generated surface pattern creation that
is based on traditional fluid mechanics. The theory's foundation in basic conservation principles
and the ability to include all the physics relating to the erosive process into suitable constitutive
equations for the stress (body force) and boundary conditions make it strong. This method
enables accurate accounting for a variety of phenomena, including viscous flow, erosion, and
ion-induced strain (damage).
The forces associated with the stress caused by the ion impacts do, in fact, affect the
morphological stability of the interface, as demonstrated by our examination of the governing
equations' linear stability. We have seen that some of the intricacies of the recently revealed
morphological diagram for Si can be explained by our description for a specific choice of the
angular dependency of the body force. In this regard, viscous flow appears to have a greater
impact on the surface dynamics than solely erosive contributions.
Finally, it can be observed that the suggested hydrodynamic framework generalises earlier
two-field formulations of IBS to the case of a flowing layer of any thickness. Such
phenomenological models have been effective in producing an equation that effectively
captures the nonlinear dynamics of the surface, sometimes reaching quantitative [4] and semi
quantitative [45] agreement with experiments. On the other hand, our current paradigm does
permit a methodical investigation of nonlinear effects using a lubrication theory methodology.

https://iaeme.com/Home/journal/IJARET 8 editor@iaeme.com
Shak Kareem and Syed Shahnawaz Ali

Future research will focus on the specifics of the constitutive equations that are appropriate
for various experimental situations. Additionally, the general non-linear theory will be
addressed using accepted practises [33] that ought to be able to take work into consideration.
nonlinear phenomena like pattern coarsening, roughness saturation, defect annihilation, and/or
defect formation.

REFERENCES
[1] R.L. Cunningham, et al., J. Appl. Phys. 31 (1960) 839.
[2] M. Navez, C. Sella, D. Chaperot, C. R. Acad. Sci. Paris 254 (1962) 240.
[3] H. Gnaser, in: Low Energy Ion Irradiation of Solid Surfaces, Springer, New York, 1998.
[4] J. Mu˜noz-García, R. Gago, L. Vázquez, J.A. Sánchez-Garcí a, R. Cuerno, et al., Phys. Rev.
Lett. 104 (2010) 026101.
[5] W.L. Chan, E. Chason, J. Appl. Phys. 101 (2007) 121301.
[6] J. Mu˜noz-García, L. Vázquez, R. Cuerno, J.A. Sánchez-García, M. Castro, R. Gago, in: Z.M.
Wang (Ed.), Towards Functional Nanomaterials, Springer, New York, 2009.
[7] R. Cuerno, M. Castro, J. Mu˜noz-García, R. Gago, L. Vázquez, Nucl. Instrum. Meth. Phys. Res.
B 269 (2011) 894.
[8] C.A. Volkert, J. Appl. Phys. 74 (1993) 7107.
[9] S.G. Mayr, Y. Ashkenazy, K. Albe, R.S. Averback, Phys. Rev. Lett. 90 (2003) 055505.
[10] R.M. Bradley, J.M.E. Harper, J. Vac. Sci. Technol. A 6 (1988) 2390.
[11] M. Cross, H. Greenside, in: Pattern Formation and Dynamics in Nonequilibrium Systems,
Cambridge University Press, Cambridge, 2009.
[12] G. Ozaydin, A.S. Ozcan, Y. Wang, K.F. Ludwig, H. Zhou, R.L. Headrick, D.P. Siddons, Appl.
Phys. Lett. 87 (2005) 163104.
[13] C.S. Madi, B. Davidovitch, H.B. George, S.A. Norris, M.P. Brenner, M.J. Aziz, Phys. Rev. Lett.
101 (2008) 246102.
[14] C.S. Madi, H.B. George, M.J. Aziz, J. Phys.: Condens. Matter 21 (2009) 224010.
[15] S. Macko, F. Frost, B. Ziberi, D.F. Förster, Th. Michely, Nanotechnology 21 (2010) 085301.
[16] S. Macko, F. Frost, M. Engler, D. Hirsch, T. Höche, J. Grenzer, T. Michely, New J. Phys. 13
(2011) 073017.
[17] G. Carter, V. Vishnyakov, Phys. Rev. B 54 (1996) 17647.
[18] V. Shenoy, W.L. Chan, E. Chason, Phys. Rev. Lett. 98 (2007) 256101.
[19] S. Le Roy, E. Sondergard, I.S. Nerbo, M. Kildemo, M. Plapp, Phys. Rev. B 81 (2010) 161401.
[20] R.M. Bradley, P.D. Shipman, Phys. Rev. Lett. 105 (2010) 145501.
[21] J. Zhou, M. Lu, Phys. Rev. B 82 (2010) 125404.
[22] M. Makeev, R. Cuerno, A.-L. Barabási, Nucl. Inst. Meth. Phys. Res. B 197 (2002) 185.
[23] S. Facsko, T. Bobek, A. Stahl, H. Kurz, T. Dekorsy, Phys. Rev. B 69 (2004) 153412.
[24] B. Davidovich, M. Aziz, M. Brenner, Phys. Rev. B 76 (2007) 205420.
[25] M. Castro, R. Cuerno, L. Vázquez, R. Gago, Phys. Rev. Lett. 94 (2005) 016102.
[26] J. Mu˜noz-García, M. Castro, R. Cuerno, Phys. Rev. Lett. 96 (2006) 086101.
[27] J. Mu˜noz-García, R. Cuerno, M. Castro, Phys. Rev. B 78 (2008) 205408.

https://iaeme.com/Home/journal/IJARET 9 editor@iaeme.com
Ion Beams' Hydrodynamic approach to the Generation of Surface Patterns

[28] J. Mu˜noz-García, R. Cuerno, M. Castro, J. Phys.: Condens. Matter 21 (2009) 224020.


[29] L. Pelaz, L.A. Marqués, J. Barbolla, J. Appl. Phys. 96 (2004) 5947.
[30] N. Kalyanasundaram, M. Wood, J.B. Freund, H.T. Johnson, Mech. Res. Commun. 35 (2008)
50.
[31] M. Moseler, P. Gumbsch, C. Casiraghi, A.C. Ferrari, J. Robertson, Science 309 (2005) 1545.
[32] M.C. Moore, N. Kalyanasundaram, J.B. Freund, H.T. Johnson, Nucl. Inst. Meth. Phys. Res. B
225 (2004) 241.
[33] A. Oron, S.H. Davis, S.G. Bankoff, Rev. Mod. Phys. 69 (1997) 931.
[34] R. Kree, Oral Communication at International Conference on Ion-Beam Induced
Nanopatterning of Materials (IINM-2011), 06–10 February, 2011.
[35] M. Castro, R. Cuerno, arXiv:1007.2144v1, 2011.
[36] S.E. Orchard, Appl. Sci. Res. 11A (1962) 451.
[37] W.W. Mullins, J. Appl. Phys. 30 (1959) 77.
[38] E. Chason, T.M. Mayer, B.K. Kellerman, D.T. Mcilroy, A.J. Howard, Phys. Rev. Lett. 72 (1994)
3040.
[39] F. Frost, R. Fechner, B. Ziberi, J. Völlner, D. Flamm, A. Schindler, J. Phys.: Condens. Matter
21 (2009) 224026.
[40] S.A. Norris, J. Samela, L. Bukonte, M. Backman, F. Djurabekova, K. Nordlund, C.S. Madi,
M.P. Brenner, M.J. Aziz, Nat. Commun. 2 (2011) 276.
[41] S.-P. Kim, B.-H. Kim, H. Kim, K.-R. Lee, Y.-C. Chung, J. Seo, J.-S. Kim, Nucl. Instrum. Meth.
Phys. Res. B 269 (2011) 2605.
[42] S. Vauth, S.G. Mayr, Phys. Rev. B 75 (2007) 224107.
[43] T. Aste, U. Valbusa, Physica A 332 (2004) 548, New J. Phys. 7 (2005) 122.
[44] B.J. Spencer, S.H. Davis, P.W. Voorhees, Phys. Rev. B 47 (1993) 9760.
[45] J.-H. Kim, N.-B. Ha, J.-S. Kim, M. Joe, K.-R. Lee, R. Cuerno, Nanotechnology 22 (2011)
285301.

https://iaeme.com/Home/journal/IJARET 10 editor@iaeme.com

You might also like