1: \documentstyle[aps,floats,graphicx]{revtex}
2:
3: \begin{document}
4:
5: \newcommand{\bec}{\begin{center}}
6: \newcommand{\ec}{\end{center}}
7: \newcommand{\be}{\begin{equation}}
8: \newcommand{\ee}{\end{equation}}
9: \newcommand{\beqn}{\begin{eqnarray}}
10: \newcommand{\eeqn}{\end{eqnarray}}
11: \newcommand{\bet}{\begin{table}}
12: \newcommand{\ent}{\end{table}}
13: \newcommand{\bib}{\bibitem}
14:
15:
16: %\baselineskip 4.2mm
17:
18: \wideabs{
19:
20: \title{
21: What is the real driving force of ion beam mixing?
22: }
23:
24:
25: \author{P. S\"ule$^1$, M. Menyh\'ard$^1$, K. Nordlund$^2$}
26: \address{$^1$ Research Institute for Technical Physics and Material Science,\\
27: Konkoly Thege u. 29-33, Budapest, Hungary, sule@mfa.kfki.hu/www.mfa.kfki.hu/~sule\\
28: $^2$ Accelerator Lab., Helsinki, Finland
29: }
30:
31: \date{\today}
32:
33: \maketitle
34:
35: \begin{abstract}
36: Molecular dynamics simulations have been used to study the driving force of
37: ion irradiation induced interfacial mixing in metal bilayers in which the
38: relative mass of the constituents is considerable.
39: We find no apparent effect of chemical forces, such as heat of mixing
40: or cohesive energy up to 8 keV ion energy, although a considerable
41: number of liquid and high energy particles (hot atoms) persist up to even $20$ ps during the thermal spike.
42: %In the pure elements the thermal spike survives up to $4$ ps.
43: This result is in direct conflict with the widely accepted theory of
44: thermal spike mixing (chemical interdiffusion model).
45: The supersaturation of vacancies also occurs and which induces a thermally activated
46: intermixing of the lighter constituent of the bilayer.
47: The delay and the decoupling of the intermixing of the light constituent is explained
48: as a backscattering effect at the interface:
49: the interface acts as a diffusional barrier for high energy particles due to the large
50: difference in atomic masses.
51: The heavier atoms are predominantly ejected to the overlayer at the beginning of
52: the thermal spike while the light atoms are injected to the bulk at the beginning of the
53: cooling period (in Ti/Pt) or during the thermal spike with some time delay (Al/Pt).
54: \\
55:
56:
57: {\em PACS numbers:} 79.20.Rf, 61.80.Az 61.80.Jh 61.82.Bg\\
58: 61.80.Az Theory and models of radiation effects\\
59: 61.80.Jh Ion radiation effects \\
60: 61.82.-d Radiation effects on specific materials\\
61: 61.82.Bg Metals and alloys\\
62: 66.30.-h Diffusion in solids (for surface and interface diffusion, see 68.35.Fx)\\
63: 79.20.Rf Atomic, molecular, and ion beam impact and interactions with surfaces\\
64: 81.40.Wx Radiation treatment (particle and electromagnetic) (see also 61.80.-x Physical radiation effects, radiation damage)\\
65:
66: {\em Keywords: Computer simulations, Ion-solid interaction, ion-beam mixing, molecular dynamics, mass effect, interfacial mixing, atomic migration,
67: amorphisation}
68: \end{abstract}
69: }
70:
71:
72: \section{Introduction}
73:
74: The rapid growth in the use of ion implantation and ion beam processing of materials
75: in the semiconductor industry and elsewhere, such as ion-beam sputtering,
76: cleaning, smoothing, depth profiling, etc. \cite{Gnaser}, or
77: global response of solids under irradiation (e.g. phase instability in driven alloys) \cite{Bellon,Okamoto},
78: makes it important to understand the radiation effects in materials and the displacement mechanism
79: of atoms during particle irradiation.
80: Despite the tremendous knowledge available in this field it is still not well understood
81: how to manipulate thin films, nanocrystalline materials, etc. in a proper way under ion irradiation driven conditions \cite{Samwer,Enrique}.
82:
83: Ion-beam assisted processing of materials always leads to ion-beam mixing (IM), hence it is a topic
84: of a considerable technological interest. One of the greatest challenges is to keep
85: IM in a controllable fashion and to drive systems to a desired atomic configuration.
86: IM is oftenly used for amorphization, when ion bombardment drives the system far
87: from its equilibrium \cite{Okamoto,Samwer}.
88: Ion beam bombardment in mechanically stressed thin films and radiation-induced viscous flow are also hot topics of the area
89: \cite{Mayr}.
90:
91:
92:
93: Two types of phenomenological models have been developed to describe
94: ion beam mixing.
95: The ballistic model assumes only the kinematic properties of the materials and the elastic collisions of energetic
96: particles (recoils) hence the deposited energy depends on the relative masses
97: of the colliding atoms \cite{Sigmund}. The effectivity of ballistic mixing (BM)
98: will depend on then the concentration of recoils with relatively
99: short mean free path which we call hot atoms.
100: BM tends to randomize atomic configurations leading to disordering of crystalline ordered
101: phases even in precipitating systems \cite{Enrique,Roussel}.
102: The slowing down of energetic particles, however,
103: leads to energy deposition and to local melting.
104: When a local volume is heated up induced by the slowed down
105: incoming ion and by the recoils, a thermal spike is created \cite{AverbackRubia}.
106:
107: The ballistic random diffusion process \cite{Roussel} might also be active still during the
108: TS and above a threshold energy leads to the formation of stable Frenkel pairs.
109: This results in
110: a supersaturation of the point defects and the enhancement of thermal diffusion \cite{Roussel}.
111: The competition between thermally activated and ballistic diffusional mixing might be a key
112: ingredient of IM \cite{Bellon,Roussel,Martin}.
113: In Martin's ballistic diffusional theory
114: atomic interdiffusion is influenced by atomic collisions and thermodynamic forces \cite{Martin}.
115: The latter is associated with thermally activated jumps of point defects.
116:
117: In addition,
118: thermochemical forces could be set in and influence thermally activated
119: atomic transitions \cite{AverbackRubia}.
120: In ref. \cite{Gades} it was explicitly demonstrated by MD simulations that within an
121: artifical Cu/Cu bilayer system the heat of mixing has a considerable effect on broadening.
122: Therefore the coexistence of chemical interdiffusion and BM could easily be
123: occurred (chemically guided collisional mixing \cite{Kelly}).
124:
125: The influence of thermodynamic driving forces on IM is summarized by Cheng \cite{Cheng} within a series of bilayer samples.
126: Although they found a considerable scatter in the data the trend seems to be unmistakable: the amount of IM depends linearly on $\Delta H_m$
127: as it was proposed by Johnson and co-workers \cite{Johnson}, and the slope of
128: the mixing versus heat of mixing ($\Delta H_m$) curve is proportional to the cascade temperature ($\sim
129: 10^4$ K) \cite{AverbackRubia}.
130: This TS model is based on the Vineyard model developed for thermally activated
131: atomic transitions \cite{Vineyard}, and relates IM to the thermodynamic quantities,
132: the cohesive energy ($E_{cohes}$) and $\Delta H_m$.
133: The model sharply distinguish between ballistic and TS mixing.
134: The chemical interdiffusion model is supported in a number of ways experimentally \cite{AverbackRubia,mix_exp}, however, mostly at high ion energies (several hundreds keV or even more).
135: The TS concept assumes that following a short cascade period a stable liquid phase occurs
136: which persists up to several ps and no high energy particles (recoils) survive the cascade period.
137:
138: The mechanism of IM, is, however, less extensively studied at lower energies experimentaly
139: \cite{Gnaser} and by atomistic simulations \cite{AverbackRubia,TS,Colla,Gades,Nordlund,Sule1,Sule2}.
140: What is more or less
141: accepted now that the thermal spike plays an essential role in IM
142: above $1$ keV ion energy \cite{Gnaser}. Although the magnitude of TS IM is uncertain with the increasing
143: ion energy.
144: While the TS model assumes that all the atomic transport occur during the TS, other
145: authors suggest that IM takes place afterwards, in the relaxation period, because of the motion
146: of radiation-induced defects \cite{Kelly}.
147: In our recent communication we studied the effect of $\Delta H_m$ on IM in Ti/Pt by
148: MD simulations and found no apparent effect of chemical forces on intermixing
149: \cite{Sule1}.
150: This is in contrast with other low-energy MD studies where the effect of $\Delta H_m$ was apparent
151: \cite{Gades,Nordlund}.
152: We discuss in detail in the Discussion section the possible reason of the discrepancy.
153:
154:
155: Low energy ion mixing (several keV or less) is even more complicated because in this case
156: the ion-solid interaction takes place close to the free surface and the
157: structural rearrangements at the surface and the bulk are coupled \cite{Sule1,Sule2}.
158: The atomic relocation processes are affected by the proximity of the surface
159: leading to a cooperative atomic transport as well as to surface cavity growth \cite{Sule2}.
160:
161: In this paper we report on molecular dynamics simulations up to 7 keV ion energy in two
162: metal bilayers. As in ref. \cite{Sule1} we found no apparent dependence of IM on $\Delta H_m$, although we explicitly demonstrate that an extended liquid ensemble (TS) is present at the interface. Contrary
163: to this a strong influence of the atomic mass ratio on interfacial mixing is explored.
164: Hence the role of the TS model can be questionized below $7$ keV ion energy.
165: We have chosen two system for detailed studies of IM: the already at low energies well studied
166: Ti/Pt \cite{Sule1} and the Al/Pt systems. These bilayers are chosen because of the considerable
167: difference in the atomic masses of the constituents and as such are suitable for the
168: study of the effect of the mass ratio on IM \cite{Sule2}.
169: We examine the relative role of ballistic mixing vs. thermally activated vacancy mechanism
170: as it was given in the macroscopic continuum model of Martin \cite{Bellon,Martin}.
171: A different picture could be emerged if
172: both processes are understood as fast diffusion processes in an atomistic description
173: which act against each other \cite{Bellon,Enrique}.
174:
175:
176:
177: The structure of this paper is as follows. First we outline the simulation method
178: used and define certain expressions used frequently throughout the paper.
179: In the discussion section we focus on the problem of the coexistence of the ballistic mechanism
180: and the thermal spike.
181: We also demonstrate the decoupling of the ballistic and thermally activated diffusional
182: mixing processes.
183: We discuss and explain the presence of a backscattering mass effect of the light particles
184: at the interface.
185:
186:
187: %have similar, while in the other system have different
188: %melting points.
189: %The melting point difference may lead to different recrystallization rates of the
190: %two materials \cite{Nordlund} hence to different diffusion behavior across the
191: %interface (diffusional strain).
192: %It has been suggested that in interfacial systems with melting point difference
193: %the interdiffusional transport is an inverse Kirkendall type with a vacancy mechanism \cite{Nordlund,Sutton}.
194:
195:
196: \section{The setup of the simulation}
197:
198: Classical molecular dynamics simulations \cite{Allen} were used to simulate the ion-solid interaction
199: using the PARCAS code developed by Nordlund {\em et al.} \cite{Nordlund_ref}.
200: Here we only shortly summarize the most important aspects.
201: The variable timestep
202: and the Berendsen temperature control is used \cite{Allen}. The bottom layers
203: are held fixed and used as heat sink (heat bath) to maintain the thermal equilibrium of the entire
204: system.
205: The detailed description of other technical aspects of the MD simulations are given in \cite{Nordlund_ref} and details specific to the current system in recent
206: communications \cite{Sule1,Sule2}.
207:
208: The Ti/Pt sample consists of $328000$ atoms for the interface (IF) system
209: with $16$ Ti top layers and a bulk which is Pt.
210: The lattice constants for Pt is $a \approx 3.92$ \hbox{\AA} and for Ti $a \approx
211: 2.95$ and $c \approx 4.68$ \hbox{\AA}.
212: At the interface (111) of the fcc crystal is parallel to (0001) of the hcp
213: crystal
214: and close packed directions are parallel \cite{Sutton}.
215: The interfacial system as a heterophase bicrystal and a composite object of
216: two different crystals with different
217: symmetry are created as follows:
218: the hcp Ti is put by hand on the (111) Pt bulk and various structures are probed
219: and are put together randomly. Finally that one is selected which has the smallest
220: misfit strain prior to the relaxation run.
221: The remaining misfit is properly minimized below $\sim 6 \%$ during the relaxation
222: process so that the Ti and Pt layers keep their original crystal structure and we get an
223: atomically sharp interface.
224: The corresponding Ti-Ti and Pt-Pt interatomic (first neighbour) distances are $2.89$ and $2.73$
225: \hbox{\AA} at
226: the interface.
227: The Ti (hcp) and Pt (fcc) layers at the interface initially are separated by $2.8$ {\AA} and allowed freely to relax during the simulations. The variation of the equilibrium Ti-Pt distance within a reasonable $2.6-3.0$ \hbox{\AA} interval in the interatomic potential does not affect the final results significantly. We find the average value of $d \approx 2.65$ \hbox{\AA} Ti-Pt distance in the various irradiation steps and also in the nonirradiated system after a careful relaxation process. We believe that the system is properly relaxed and equilibrated before the irradiation steps.
228:
229:
230: Another sample (Al/Pt) consists of 280000 atoms ($200 \times 200 \times 100 \hbox{\AA}$ for the interface (IF) system
231: with 6 fcc-Al top layers and a bulk which is fcc-Pt both with (111) orientation.
232: The interface is (111) oriented.
233: We find the size of the system is large enough to limit the influence of
234: the boundaries on the internal material transport.
235: The embedded atomic potential of Cleri and Rosato \cite{CR} is used and modified
236: to reproduce the interatomic distance for Al-Pt found in the AlPt alloy.
237: The heat of mixing for Al-Pt is also fitted to the experimental values \cite{Miedema},
238: although we find no significant dependence on it in accordance with earlier findings
239: in the TiPt system \cite{Sule1}.
240: This system size is capable of accepting as large as 8 keV ion bombardment.
241: In this paper we show results at 6 keV ion energy.
242: The entire interfacial system is equilibrated prior to the irradiation
243: simulations and the temperature scaled softly down towards zero at the
244: outermost three atomic layers during the cascade events \cite{Nordlund_ref}.
245: We believe that the system is properly relaxed and equilibrated before the irradiation steps.
246: We found no vacancies at the interface or elsewhere
247: in the system after the relaxation procedure.
248: We visually checked and found no apparent screw dislocations, misorientations or any
249: kind of undesired distortions (in general stress-generator dislocations) at the interface or elsewhere in the system which could result in
250: stress induced rearrangements in the crystal during the simulations.
251:
252: Collsion cascades were initiated by giving an Ar atom a kinetic energy of 1-7 keV at
253: 7 deg impact angle (grazing angle of incidence) and at randomly chosen impact positions in the
254: central region of the free surface.
255: The SRIM96 electronic stopping power \cite{SRIM} was used to describe energy loss to electrons for all
256: atoms with a kinetic energy higher then 10 eV.
257:
258: We used a many body potential, the type of an embedded-atom-method given by Cleri and
259: Rosato \cite{CR}, to describe interatomic interactions.
260: This type of a potential gives a very good description of lattice vacancies, including migration
261: properties and a reasonable description of solid surfaces and melting \cite{CR}.
262: Since the present work is mostly associated with the elastic properties,
263: melting behaviors, surface, interface and migration energies, we believe the model used should be suitable for this study.
264: The Ti-Pt and Al-Pt interatomic potential of the Cleri-Rosato \cite{CR} type is fitted to the experimental heat of
265: mixing of the corresponding alloy system. Further details are also given elsewhere \cite{Sule1,Sule2}.
266: To obtain a statistics, several events are generated ($\sim 5-10$), and the typical events are
267: analyzed in the paper. Therefore no averaging is carried out over the events.
268: The variation of mixing as a function of the impact position and various events is shown
269: in refs. \cite{Sule1,Sule2}.
270:
271:
272: Following the simulations we analyse the generated structural data base (movie file) which contains
273: the histories of the atomic positions using a code written for this purpose.
274: The analysis includes the identification of vacancies, liquid and mixed atoms, mean free path
275: of recoils and the extraction of information on local temperature of high energy particles
276: using the results of liquid analysis.
277:
278: Those atoms are recognised mixed (intermixed) which moved at least 0.5 monolayers ($\sim 1.4$ \hbox{\AA})
279: across the interface.
280:
281: The recognition of {\em vacancies} is done using a Voronoy polyhedron analysis
282: together with a Wigner-Seitz occupation analysis \cite{Nordlund_ref,Nordlund_PRB97}.
283: Within this approach we interpret voids, cavities or craters as vacancy clusters \cite{Sule2}.
284: Another problem is that in liquids, such as appears during the thermal spike,
285: the definition of vacancies is rather plausible since the empty cells within a liquid
286: are highly mobile. Hence during the TS we monitor only formally the time evolution of
287: the number of vacancies. This is important to indentify the appearance of the
288: supersaturation of vacancies, which is in fact the monitoring of the change of
289: the concentration of the empty cells within a liquid zone.
290:
291: {\em Liquid analysis} is also carried out and we call an energetic atom liquid if its
292: and its first neighbours kinetic energy exceed $k T_m$ where $T_m$ is the melting temperature.
293: Important requirement for a liquid atom is that
294: it should be a member of a real liquid ensemble, hence an individual fast moving hot atom or a recoil
295: can not be considered as liquid.
296: Though the quench rate of the liquid phase is large, it can persist normally up to
297: several ps. In certain cases (e.g. Al/Pt at 6 keV) we observe an extraordinary long
298: lifespan for the local melt reaching $20$ ps or even more.
299: It turned out that one has to be careful when defining the liquid atoms.
300: Strictly speaking we denote those atoms liquid for which $T_m < T_{local} < T_{TS}$, where
301: $T_{local}$ and $T_{TS}$ are the local temperature
302: and an arbitrary upper limit for the TS temperature, respectively.
303: The time dependent local temperature of the $i$th high energy particle ($T_{i,local}(t)$) can be given as follows:
304: \be
305: \frac{1}{2} m_i v_i^2(t) = \frac{3}{2} k T_{i,local}(t),
306: \ee
307: where $v_i$ and $m_i$ are the atomic velocity and mass of the $i$th high energy particle and $k$ is the Boltzmann constant.
308: We chose $T_{TS} \approx T_m+1000$ K, which seems to be a reasonable choice, because an atom
309: with $T_{local} >> T_{TS}$ is moving through the liquid/solid interface and migrates rapidly
310: in the interstitial space (recoil).
311: We also carry out recoil and hot atom analysis. For a hot atom we use the definition
312: $T_m+1000 < T_{local}$.
313: We find a considerable amount of hot atoms both in Ti/Pt and in Al/Pt at various ion energies. In Al/Pt under 6 keV bombardment the high energy particles persist up to $20$
314: ps.
315: We use the notation recoil for those individual high energy particles, which can have extraordinarily
316: high local temperature and very short lifetime ($\tau_{recoil} < 3-500$ Fs) with long mean free path
317: ($> 10$ \hbox{\AA}).
318: Hot atoms, however, can be taken as highly mobile high energy particles with medium mean free path
319: (around several times the lattice constant).
320: Therefore the hot atoms might thermalize a given region with a size of which is comparable with the
321: mean free path of the hot species.
322: The kinetic energy of a hot atom is below the threshold of a stable Frenkel pair formation
323: energy therefore the ballistic jumps of hot atoms lead to subthreshold events which, however,
324: are shown to contribute noticeably to the evolution of the damaged structure during IM (e.g. close Frenkel-pair recombinations, etc.) \cite{Wollenberger}.
325: Another basic difference between a recoil and a hot atom is that the former is
326: moving "freely" in the solid while the latter is a weakly bounded particle,
327: hence it interacts weakly with its neighbourhood.
328: It should also be mentioned that a brief definiton of hot atoms has given in
329: ref. \cite{Nordlund_ref} where those energetic atom are considered hot which
330: have experienced less then 3-5 lattice vibrations.
331: In this paper we prefer to select the energetic particles on the basis of their
332: local temperature regime.
333:
334:
335: In order to avoid confusion we define the expressions diffusion and migration.
336: These phenomena are widely used in the literature for different processes \cite{Shewmon}, although
337: their meaning often is not clearly defined. Even more difficult to define the
338: precise meaning of radiation enhanced diffusion (RED).
339: In the present article in general we use the expression diffusion for any kind of
340: atomic migration processes which are taken through a solid.
341: Usually diffusion is used for slow atomic migration process on a ms or longer timescale.
342: Therefore we use the "fast" or "ultrafast" diffusion expressions if the timescale
343: is shorter \cite{Okamoto,Martin,Sutton}.
344: If we do not specify the duration of the process, we prefer to use atomic migration.
345: Also we use the notion RED for those atomic transport processes which are taken place
346: within tens or hundreds of ps. This is typically the timescale of MD simulations.
347: Sometimes the RED is used for any kind of atomic transport phenomena in the literature
348: which appears under irradiation.
349: We prefer not to use RED for diffusion on a longer timescale (ms or longer) which is in our opinion can not be
350: distinguished from diffusion under parent conditions \cite{Shewmon}.
351: The expression ballistic diffusion \cite{Roussel} can also be used instead of
352: (ultra)fast diffusion.
353:
354:
355: %------------------------------------------------------
356:
357: \begin{figure}[hbtp]
358: %\begin{figure}[!t]
359: \begin{center}
360: \includegraphics*[height=5cm,width=6cm,angle=0.0]{tipt_kev7.eps}
361: \includegraphics*[height=5cm,width=6cm,angle=0.0]{alpt_yz_6kev.eps}
362: \includegraphics*[height=5cm,width=6cm,angle=0.0]{alpt_6kev_xz_2.5ps.eps}
363: %\includegraphics*[height=5cm,width=6cm,angle=0.0]{alpt_yz_liq_6kev.eps}
364: \caption[]{
365: The cross sectional view of the Ti/Pt system after a 7 keV irradiation
366: at the end of the relaxation process.
367: The overlayer is Ti and the substrate is the Pt.
368: {\em Middle panel}: The cross sectional view of Al/Pt after 6 keV
369: irradiation and 55 ps after the ion impact.
370: The thickness of the cross-sectional slabs are $10$ \hbox{\AA} in both cases.
371: {\em Lower panel}: The cross sectional view of Al/Pt after 6 keV
372: at 2.5 ps.
373: }
374: \label{xz}
375: \end{center}
376: \end{figure}
377:
378: %------------------------------------------------------
379:
380:
381:
382:
383: \section{Results and discussion}
384:
385: First we focus on the lack of the dependence of IM on chemical forces and then
386: we try to address a ballistic scenario for mixing without the usage of the TS concept
387: \cite{Cheng}.
388: Already we have presented in our previous communication \cite{Sule1}
389: that we find no dependence on $\Delta H_m$ in Ti/Pt, although we found
390: extensive IM at 1 keV ion energy.
391: The Cleri-Rosato many body potential can be fitted to the experimental heat of mixing
392: or can be varied in order to account for any values of $\Delta H_m$.
393: Hereby we do not give the technical details of the variation of $\Delta H_m$ in the
394: interatomic potential and refer to another ref. \cite{Sule1}.
395: Furthermore, we explored the decoupling of the intermixing of the constituents:
396: first the bulk Pt atoms are ejected to the top layers and at the end of the TS
397: the top layer Ti atoms are injected to the Pt bulk.
398: We had no clear cut explanation for this unexpected behavior of this system.
399: In this paper we would like to demonstrate that basically we find the same effects
400: at higher energies.
401:
402: In the next subsections we summarize the results obtained and outline certain possible
403: scenarios which might help to understand IM.
404: These are the following: competing dynamics between ballistic diffusion and
405: vacancy saturation induced reordering \cite{Martin}.
406: We explain the asymmetric mixing by a backscattering effect of the light atoms
407: at the interface which delays their intermixing.
408: We try to compare our results with experiments obtained for glassy and amorphised metals
409: in order to point out the parallels.
410: In any of the outlined mechanisms we discuss the essential role of hot atoms in IM.
411:
412:
413: \subsection{Intermixing and liquid analysis}
414:
415: The cross-sectional view of the irradiated Ti/Pt sample at 7 keV is shown after 30 ps
416: of the ion impact in FIG ~(\ref{xz}) which clearly demonstrate that the interface is
417: damaged and large amount of intermixing occured and remained there built-in.
418: A complete animation is available on the world wide web \cite{www}.
419: %after resolidification.
420: Liquid analysis provides us the information that the number of liquid atoms
421: is increased heavily with the increasing ion energy (FIG ~(\ref{liquid})).
422: Therefore there should be present a liquid phase which behaves as a real TS.
423: Contrary to this we find no apparent dependence on $\Delta H_m$.
424: Even if $\Delta H_m \ge 0$, considerable mixing occurs instead of segregation.
425: This is in direct conflict with the results obtained in ref. \cite{Gades}.
426: We explain the discrepancy with the following.
427: Gades and Urbassek studied the effect of $\Delta H_m$ on IM in an artifical system
428: of Cu/Cu bilayer \cite{Gades}. They varied $\Delta H_m$ within this system in which the first six
429: layers consist of natural Cu and the substrate of a modified Cu which has a positive
430: or negative heat of mixing towards Cu.
431: They found a strong effect of $\Delta H_m$ on broadening.
432: We attribute the presence of a chemically guided mixing to the 1:1 mass ratio, to the lack of lattice missmatch and and to the same cohesive energy in the
433: bulk and in the overlayer.
434: Within their computer experiment the only variable was $\Delta H_m$, while in a real
435: system this is far not the case.
436: In a general bilayer system couple of other parameters set in which may suppress the effect
437: of $\Delta H_m$.
438: The relocation energy of an atom in a solid is affected
439: by a couple of other parameters \cite{Allnat}.
440: E.g. in the case of Ti/Pt and Al/Pt there is
441: a considerable mass ratio, difference in cohesive energy and in the melting points of
442: the constituents not to mention the lattice missmatch.
443: Another problem with chemically guided IM is that $\Delta H_m$ could only affect
444: atomic transport processes when the atoms are moving relatively slowly with ordinary
445: diffusion. Fast moving particles are only weakly influenced by chemical forces and the chemical
446: environment affects only those species which are bounded at a lattice site.
447: We will show later on that the concentration of ballistic particles (hot atoms)
448: is sufficiently high in Ti/Pt and in Al/Pt during the TS to supress the effect of $\Delta H_m$.
449: Although the chemical interaction between liquid atoms is stronger, however, their
450: mobility might still be sufficiently high to be affected only weakly by $\Delta H_m$.
451: This is a subject, however, what should be carefully examined in the future and which goes
452: beyond the scope of the present article.
453:
454:
455:
456: We find a strong asymmetric behavior in mixing (FIG ~(\ref{mixing})).
457: The decoupling of the mixing of the top layer constituent from the TS period
458: is also apparent (FIG ~(\ref{mixing})).
459: Asymmetric mixing behavior has also been observed in Ni/Ag \cite{Colla} and in Cu/Ni or in Cu/Co
460: \cite{Nordlund}.
461: In our case first the heavier element Pt mixes and the intermixing of the light constituent Al or Ti
462: is initiated at 5 ps.
463: The mixing of Ti (Ti/Pt) continues after the TS period hence
464: no liquid atom is required for the mixing of Ti.
465: Pt mixing peaks at around 4 ps.
466: We find no intermixing hot or liquid Ti atoms.
467:
468: %------------------------------------------------------
469:
470: \begin{figure}[hbtp]
471: \begin{center}
472: \includegraphics*[height=4.5cm,width=6.cm,angle=0.0]{tipt_7kev_mix_vac.eps}
473: %\includegraphics*[height=4cm,width=5.5cm,angle=0.0]{alpt_mix_8kev.eps}
474: \includegraphics*[height=4.5cm,width=6.cm,angle=0.0]{alpt_6kev_mix_vac_adatoms.eps}
475: %\includegraphics*[height=4cm,width=5.5cm,angle=0.0]{niag_9kev_mix.eps}
476: \caption[]{
477: The time evolution of mixing and vacancy formation in Ti/Pt (upper figure) at 7 keV ion energy. Mixing profile in Al
478: /Pt (lower figure) together with the time evolution of vacancies at 6 keV ion energy.
479: The number of adatoms are also shown in the inset of the lower figure as a function of the time
480: (ps).
481: %{\em The lower figure}: Mixing profile in Ni/Ag after 9 keV bombardment. Note the strong
482: %mixing and demixing effects.
483: }
484: \label{mixing}
485: \end{center}
486: \end{figure}
487:
488: %------------------------------------------------------
489:
490:
491: The number of vacancies $N_{vac}$ exhibits a sharp peak at 4 ps (upper FIG ~(\ref{mixing})) which
492: corresponds to a highly diffuse liquid state in the Ti overlayers and
493: approximately the number of vacancies is 3 \% of the number of the liquid atoms
494: which indeed leads to the sudden decrease of the atomic density (not shown).
495: %The ultrafast ballistic diffusion of the Pt atoms to the overlayers might be induced by the
496: %supersaturation of vacancies.
497: The intermixing of the Ti atoms might be initiated by the saturation of Pt vacancy concentration at around 4 ps.
498: The supersaturation of point defects results in enhanced atomic mobility \cite{Bellon}:
499: a ballistic jumps are increasingly probable in the direction of nearby vacancies.
500: Replacement collision sequences might be inititated in this way (via subtreshold collisions)
501: which enhances atomic mixing \cite{Bellon}.
502: %The interface becomes more damaged on the Ti side then in the other side (upper FIG ~(\ref{xz})) what is
503: %due to the overlayer positioning of the TS.
504:
505: The results of the liquid analysis are also shown for Ti/Pt and Al/Pt at 7 and 6 keV
506: in FIG ~(\ref{liquid}).
507: The TS persists up to $\sim 20$ ps in Al/Pt, which is an unusually long lifetime.
508: The longest reported lifetime for a TS ($\tau_{TS}$) is 14 ps for Au at a similar ion energy regime in a 5 keV cascade event \cite{Nordlund_PRB97}.
509: For pure Al and for Pt a much shorter $\tau_{TS} \approx 4$ ps is obtained \cite{Nordlund_PRB97}.
510: Also, in pure elements one can find no damage at the end of the simulation hence the restoring forces (recrystallization) are strong \cite{Sule2}.
511: As already mentioned above,
512: in the bilayer systems, however, the damage rate is high, the recovery process is uncomplete within the timescale of the MD simulations.
513: Moreover the accumulated damage rate is so high that it might also be remained built-in
514: on a longer timescale.
515:
516: In case of Al/Pt the average local temperature far exceeds the melting point of Al ($\sim T_m \approx 982$ K) value and is stabilized around
517: $\sim T \approx 1800$ K.
518: When the number of liquid atoms is less then a critical
519: value ($\sim 50$), the local temperature starts to oscillate due to the
520: abrupt condensation of the liquid atoms at the end of the TS.
521: The TS exhibits in general an
522: ultrarapid quenching rate ($\sim 10^{14}-10^{15}$ K/s, $\sim 100-1000$ K/ps) which is much larger then
523: the values are given in the literature $\sim 10^{12}$ K/s for supercooled metallic glasses \cite{Samwer}),
524: leading to a quenched metastable solid solution at the interface with compositions lying beyond the
525: equilibrium solubility limit.
526: Others also propose that fast quenching, occurring the estimated rate of about
527: $10^{14}$ K/s provides ideal conditions to nucleate metastable phases, including
528: the amorphous one \cite{Ossi} or to the clustering of vacancies \cite{Sule2}.
529: We find also a broad amorphous interphase as demonstrated in the middle FIG ~(\ref{xz}).
530:
531:
532: Glassy metals exihibit a polymorphic melting point falls
533: below the ambient temperature and below the glass transition temperature of the
534: liquid phase.
535: Under these circumstances the oversaturated solution undergoes a crystal-to-glass
536: transformation \cite{Samwer}.
537: We find also a polymorphic melting point (liquid temperature) which lies between the parent melting point of
538: the constiuents.
539: The fast "cool" mixing of Ti might be induced by a glassy state saturated by vacancies.
540: The mixing of Ti is fast enough not to be uderstood by a simple vacancy mechanism \cite{Shewmon}.
541: Therefore we argue that a glassy sate might accelerate the diffusion of the Ti species.
542: The initialization of mixing of Ti coincides with the sudden fall of $N_{vac}$.
543: Hence the vacancy supersaturation initializes a fast intermixing which requires no
544: liquid atoms in Ti/Pt.
545: In Al/Pt the situation is somewhat different. Here the vacancies tend to aggregate while
546: in Ti/Pt a diffuse vacancy saturation occurs with little or no vacancy clustering which
547: favors the formation of a glassy sate.
548:
549: The broad intermixed interface in Al/Pt
550: (seen in FIG ~(\ref{xz})) is consistent with the ion-induced amorphization experiment carried
551: out on Al/Pt \cite{Gyulai,Buchanan}, in which a uniform mixing is achieved and confirmed the existence of an amorphous phase
552: at the interface over a large composition range.
553: In general it is well documented now that binary metal systems undergo solid-sate
554: amorphization by interdiffusion or by other type of mixing \cite{Samwer}.
555:
556: %------------------------------------------------------
557: \begin{figure}[hbtp]
558: %\begin{figure}[!t]
559: \begin{center}
560: \includegraphics*[height=4.5cm,width=6.cm,angle=0.0]{tipt_liq_731kev.eps}
561: \includegraphics*[height=4.5cm,width=6.cm,angle=0.0]{alpt_6kev_liquid_temp.eps}
562: %\includegraphics*[height=4cm,width=5.5cm,angle=0.0]{alpt_6kev_liq_hot_4ps.eps}
563: %\includegraphics*[height=4cm,width=5.5cm,angle=0.0]{niag_9kev_liq.eps}
564: \caption[]{
565: {\em Upper panel}:
566: The number of liquid atoms in Ti/Pt
567: at various ion energies up
568: to 7 keV and the local average temperature (K) at 7 keV ion energy (inset).
569: %{\em 2nd figure from above:}
570: {\em Lower panel}:
571: The number of liquid, hot atoms and recoils in Al/Pt as a function of the time (Fs).
572: {\em inset:} the time evolution of the local temperature in the liquid phase (K),
573: the number of the liquid atoms, recoils and hot particles.
574: }
575: \label{liquid}
576: \end{center}
577: \end{figure}
578: %------------------------------------------------------
579: %Irradiation experiments of various intermetallic compounds show
580: %that a large elastic softening and delatation strain due to disordering precede
581: %the chemical order leading to elastic instability \cite{Okamoto,Samwer}.
582:
583:
584: Another interesting feature of FIG ~(\ref{liquid}) is the time evolution of
585: recoils/hot atoms. The recoils dissapear at the end of the
586: cascade period ($\sim 0.3$ ps) \cite{AverbackRubia}.
587: The sharp peak at $\sim 0.3$ ps corresponds to the recoils.
588: However, we find that a large amount of high energy
589: particles (hot atoms) survive the cascade cooling.
590: This is a rather surprising result and it shows that the ballistic period with hot atoms
591: coexists with the TS in these bilayers.
592: In the inset of lower FIG ~(\ref{liquid}) we show that the average temperature of
593: the hot atoms far exceeds $T_m$.
594: %In the lower FIG ~(\ref{xz}) we show that a spatially extended liquid phase exists during the TS
595: %at 4 ps.
596: In FIG ~(\ref{liqhot}) we show the time evolution of hot and liquid
597: atoms separately for various atom types.
598: Not surprisingly the various atoms exhibit different time evolution
599: due to the large melting point difference.
600: The number of Pt hot and liquid atoms
601: delays rapidly while the number of Al energetic particles show a saturation at 5 ps.
602: It has been found that the solidification time of the cascade together with hot and liquid atoms
603: depends on the melting point of the material \cite{Nordlund,Nordlund_ref}.
604: Hence the intermixed Pt particles quench out more rapidly from the liquid zone then the Al atoms and remained mixed
605: atoms at the liquid zone boundary.
606: Nevertheless it should be mentioned that in Ti/Pt the number of Pt liquid and hot
607: atoms is considerably smaller then those of Ti because the displacement cascade
608: is localized in the overlayer and the Pt side of interface is hit only
609: slightly by recoils. This is because a relatively thick Ti overlayer is constructed
610: (with 16 monolayers).
611: %We carried out MD simulations on a system with 8 Ti monolayers, and get much
612: %more Pt hot atoms.
613: %------------------------------------------------------
614: \begin{figure}[hbtp]
615: \begin{center}
616: %\includegraphics*[height=4.5cm,width=6.cm,angle=0.0]{tipt_8kev_8overl_liq_hot.eps}
617: \includegraphics*[height=4.5cm,width=6.cm,angle=0.0]{tipt_7kev_liq12_hot12.eps}
618: \includegraphics*[height=4.5cm,width=6.cm,angle=0.0]{alpt_6kev_liq12_hot12.eps}
619: \caption[]{
620: {\em Upper panel:}
621: The time evolution of the number of hot and liquid atoms (inset)
622: in Ti/Pt at 8 keV Ar$^+$ irradiation.
623: {\em Lower panel:}
624: The time evolution of the number of liquid (inset) and hot atoms in Al/Pt at 6 keV
625: ion irradiation.
626: }
627: \label{liqhot}
628: \end{center}
629: \end{figure}
630: %------------------------------------------------------
631: % The advancing front of the Pt solidification front pushes most of the vacancies to
632: %the liquid Al zone in a similar way as it was found in the Co/Cu bilayer \cite{Nordlund}.
633: %In, general it was demonstrated in fcc metals, that the vacancies are pushed towards the
634: %center of the liquid region when the molten region resolidifies \cite{Nordlund_PRB97}.
635: In Al/Pt the Pt atoms exhibit a second hot and liquid atom peak (the first one is due to the recoils)
636: at 2 ps which coincides with the end of the fast mixing of Pt atoms (FIG ~(\ref{mixing})).
637: Hence the hot atom peak is a mixing peak due to the ballistically mixing hot Pt particles.
638: In Ti/Pt we see only the broadening of the first recoil peak hence the increasing mass difference
639: might induce the development of a second recoil (hot) peak.
640: %In Ti/Pt we see the recombination of the vacancies as a much slower process which
641: %goes beyond the end of the TS (FIG ~(\ref{mixing})).
642:
643: \subsection{Thermally activated diffusion and vacancies}
644:
645: In Ti/Pt we found that
646: the fast quench of the TS leads to extensive vacancy formation \cite{Sule1}, and
647: this is what induces the injection of Ti atoms to the bulk.
648: Hence we also support a vacancy mechanism for mixing with a light atom injection to the bulk.
649: The large concentration of vacancies is an indicative of a glassy interface \cite{Okamoto}.
650: It should be remarked that Pt intermixes to the top layers during the TS possibly
651: with a ballistic interdiffusion mechanism because we find a considerable amount of
652: mixing hot Pt atoms.
653: The interdiffusion of Pt leads to amorphous inter-layer growth at the original interface
654: and in this sense resemble to an anomalous fast diffusion \cite{Okamoto,Sutton}.
655: Moreover due to the elevated temperature of the TS, a thermally activated diffusion process might become
656: dominant.
657: Although at this moment it is difficult to determine whether thermally activated or ballistic
658: diffusion is the dominant atomic migration process during the TS.
659: %The atomic migration of Ti atoms accompanied by a counter-flux of vacancy-like defects
660: %which is a well-known feature of growing amorphous layers \cite{Samwer}.
661:
662: In Al/Pt we see that the number of vacancies peaks at 2.5 ps. A more elaborate
663: analysis of the structures at that time we see that vacancies are clustering
664: and temporarily form a cavity (lower FIG ~(\ref{xz}).
665: Hence in Al/Pt the supersaturation of vacancies leads to clustering while in Ti/Pt
666: we have still a higly diffuse liquid state.
667: The most of the vacancies are located in the overlayer.
668: Comparing the heat of formation of a vacancy $\Delta H_V$
669: we find that $\Delta H_V \approx 1.31$ and $0.74$ eV for Pt and for Al
670: \cite{Swalin}.
671: This is a marked difference and explains the large discrepancy in the number of vacancies
672: in these metals.
673: The value for Al is in the energy regime of the hot atoms hence the formation
674: of vacant sites throughtout the TS can be expected.
675: In Pt, however, we see that the vacancies dissappear rapidly (not shown).
676:
677: A competition between BM and vacancy saturation induced atomic transport (vacancy mixing)
678: might also be active in irradiated samples \cite{Roussel}.
679: The BM tends to randomize (disordering) the system leading to thermalization and vacancy
680: saturation. The thermally activated vacancy recombination or net flux through the
681: interface induces reordering \cite{Roussel}.
682: In our samples we see the dominant feature of ballistic mixing in both cases for the
683: Pt atoms.
684: In Ti/Pt, however, the Ti atoms intermix with a vacancy mechanism afterwards the TS.
685: In Al/Pt both species migrate with a ballistic diffusion.
686:
687: %------------------------------------------------------
688: \begin{figure}[hbtp]
689: \begin{center}
690: \includegraphics*[height=4.5cm,width=6.cm,angle=0.0]{tipt_1kev_mix_vac.eps}
691: \includegraphics*[height=4.5cm,width=6.cm,angle=0.0]{tipt_1kev_liq_temp.eps}
692: %\includegraphics*[height=4.5cm,width=6.cm,angle=0.0]{hot_mix1_meanpath.eps}
693: \includegraphics*[height=4.5cm,width=6.cm,angle=0.0]{mix_liq_hot.eps}
694: \caption[]{
695: {\em Upper panel:}
696: The number of vacancies, adatoms and intermixed atoms
697: in Ti/Pt at 1 keV Ar$^+$ irradiation as a function of time.
698: The time evolution of the numbers of interstitials
699: is also shown in the inset.
700: {\em Middle panel:}
701: The time evolution of the number of liquid and hot atoms.
702: The average temperature (K) is shown of the liquid and hot atoms
703: as a funtcion of time (ps)
704: {\em Lower panel:}
705: The time evolution of
706: the number of mixed hot and liquid atoms.
707: }
708: \label{tipt_vacancy}
709: \end{center}
710: \end{figure}
711: %------------------------------------------------------
712:
713:
714:
715:
716: \subsection{Mixing at 1 keV ion energy}
717:
718: In order to show that the long-lived hot atoms exist even at lower ion energies
719: we recall our results presented in ref. \cite{Sule1} and give further details
720: at 1 keV energy.
721: We find the same effect for 1 keV bombardment of Ti/Pt, where the intermixing of Ti
722: coexists with the formation of vacancies \cite{Sule1} (upper FIG
723: ~\ref{tipt_vacancy}).
724: On a longer timescale in upper FIG ~(\ref{tipt_vacancy}) we see the oscillation of $N_{mix,Ti}$ (mixed Ti atoms)
725: which is due to a short distance movement of these atoms at the interface.
726: The vacancies in the interface tend to nucleate as it is demonstrated
727: in ref. \cite{Sule2} leading to a cavity growth under repeated irradiations.
728: Therefore we see a tendency to cavity formation in these systems which
729: is coupled to a mixing induced material transport \cite{Sule2}.
730: The number of interstitial atoms (inset upper FIG ~(\ref{tipt_vacancy})) exhibits
731: a similar time evolution to that of the vacancies which indicates us that the
732: mixing of Ti atoms is accomplished by vacancy exchange mechanism via interstitial
733: atoms. The vacancy mechanism is well established as the dominant mechanism of diffusion
734: in fcc metals and alloys and has been shown to be operative in many hcp metals \cite{Shewmon}.
735: However, it must be admitted that the mixing of Ti is quite fast. More then
736: 20 Ti atoms intermix during 2.5 ps at the end of the TS and IM survives the
737: quench of the TS.
738: %An interstitial atom jumps to a vacancy and then to an adjacent interstitial position.
739:
740:
741: Two types of intermixing mechanism can be seen which are more or less decoupled in time from each other:
742: First the fast ballistic diffuser specie
743: intermixes to the overlayer during the TS partly to the interstitial
744: positions or partly to vacancies left behind by overlayer atoms which migrated to the surface (adatoms).
745: %------------------------------------------------------
746: \begin{figure}[hbtp]
747: %\begin{figure}[!t]
748: \begin{center}
749: \includegraphics*[height=4.5cm,width=6.cm,angle=0.0]{alpt_6kev_recoil_xz.eps}
750: \caption[]{
751: A crossectional slab of
752: hot atoms with a slab thickness of $10$ \hbox{\AA} at 5 ps in Al/Pt after 6 keV irradiation. Note that the Al recoils
753: are reflected at interface while Pt recoils enter the interface hence intermix
754: to the overlayers.
755: Intermixed hot Al atoms are also shown by light colored crosses after 5 ps.
756: }
757: \label{recoil_xz}
758: \end{center}
759: \end{figure}
760: %------------------------------------------------------
761:
762: After cooling of the TS adatoms return from the surface to the overlayer \cite{Sule1} and push Ti atoms
763: to the bulk.
764: It should be remarked that in the case of Al/Pt the intermixing of Al starts still during
765: the TS, however, in both cases sets in at around 5 ps in the high ion energy cases (FIG
766: ~(\ref{mixing})).
767: The timescale of this fast atomic migration is very short (fractions of ps) in agreement with
768: the findings obtained phenomenologically \cite{Ossi}.
769:
770:
771:
772: In the lower FIG ~(\ref{tipt_vacancy}) the number of the intermixed
773: liquid and hot atoms are shown.
774: One can see that a considerable amount of energetic particles are mixed time to time
775: to the overlayers. Interestingly no or only few hot and liquid Ti atoms are mixed to the bulk, although
776: they are present during the TS in the overlayers. We explain this by the
777: backscattering of the energetic light atoms (Ti) at the interface due to the
778: large atomic mass difference (see later).
779: This phenomenon is already discussed briefly in ref. \cite{Gades}, although they
780: reached the conclusion that the effect of mass is small.
781: Probably that is the reason of the decoupling of the mixing of the constituents and the
782: delay of IM of the light atoms.
783: The mean free path of the hot atoms is around $\sim 10$ \hbox{\AA} and the
784: corresponding velocity of these mixing energetic particles is a robust $\sim 0.5$ \hbox{\AA}/Fs
785: ($\sim 50$ m/s).
786: %These particles are really ultra fast species, and as such should intermix
787: %via ballistic diffusion \cite{Martin}.
788:
789: \subsection{Backscattering at the interface and mass effect}
790:
791: The backscattering phenomenon is shown in detail in FIG ~(\ref{recoil_xz}) in Al/Pt
792: after 6 keV bombardment. It can clearly be seen that the Al atoms are reflected at the
793: interface until 5 ps, however, after 5 ps they start to move beyond the interface.
794: The reflection can be understood by simple kinematic reasons: the large mass difference
795: between Pt and Al results in the scattering of the light atoms.
796: Both species are reflected at the solid/liquid zone boundary.
797: It should also be mentioned that we find the same scattering phenomenon in Ti/Pt.
798: The treshold time for Al or Ti intermixing is still rather puzzling.
799: By the moment we attribute it to a vacancy supersaturation induced mixing which is strong
800: enough driving force for moving beyond the interface the light atoms at 5 ps.
801: The reflecting particles are confined within the thermalized hot zone leading to superheating of the
802: liquid region ($T_{aver} \approx 4000$ K).
803: Superheating effects under pressure in encapsulated metal clusters has also been reported
804: \cite{Banhart}.
805: The subject of superheating is, however, beyond the scope of the present article and hereby
806: we simple note that the local melt of the TS is also under pressure put by the surrounding media hence superheating properties
807: might also be present.
808: The supersaturation of vacancies is also a consequence of the confinement induced thermalization.
809: The treshold time 5 ps for the light atom mixing coincides with the saturation of
810: the number of hot Al atoms (lower FIG ~(\ref{liquid})) and with the sudden decrease in the number of vacancies (FIG ~(\ref{mixing})).
811: One might be assumed that at 5 ps the interface becomes sufficiently damaged
812: and saturated by vacancies to be permeable for light energetic atoms.
813: The temporal confinement of the cascade in the overlayer is also found in ref. \cite{Gades},
814: although they found much smaller mass effect since the smallest mass ratio was 0.5.
815: We expect the similar confinement effect in the reverse case when the bulk is formed from the
816: light atoms and the overlayer is from the heavy atoms.
817: In this particular case the light hot atoms are reflected from the interface from below, hence
818: the superheated region has no connection with the surface.
819:
820: In FIG ~(\ref{liqhot}) the number of Al hot atoms $N_{Al,hot}$ peaks at 5 ps which supports the
821: idea of interfacial barrier climbing at 5 ps: when $N_{Al,hot}$ peaks
822: Al hot atoms break through the interface and deposit their kinetic energy at the Pt side.
823: Therefore the interface is a ballistic barrier for interfacial mixing of light atoms even if they
824: are hot atoms. Therefore the barrier results in the phenomenon what we call {\em retarded ballistic mixing}
825: due to the semipermeable interface.
826: When the overlayers are sufficiently frustrated by vacancy supersaturation,
827: the light atoms break through the interface.
828: This process is much slower, however, then the fast ballistic diffusion of Pt.
829: The backscattering effect is also interesting in that point of view that
830: the semipermeable interface divides the TS volume into to separated regions.
831: On the permeable side (Pt side) the liquid atoms are Pt atoms while on the Ti or Al side
832: the local melt is a mixture of the constituents.
833: Therefore there can be a concentration gradient towards the Pt side which can drive
834: the intermixing of Ti or Al according to the Fick's law \cite{Shewmon}.
835: In that case the driving force of ballistic mixing would be a concentration and
836: a thermal gradient which changes sign during the thermal spike (Al/Pt) or due to
837: the quench of the local melt (Ti/Pt).
838:
839:
840:
841:
842:
843:
844: \section{Conclusion}
845:
846: We demonstrated that a bilayer system exhibits enhanced intermixing properties and
847: an increased thermal spike lifetime when compared with the values obtained in pure elements.
848: We attribute this behaviour to the
849: sufficiently large difference in the atomic masses of the constituents (mass ratio).
850:
851: It is shown that, in a bilayer system under irradiation, ion beam mixing
852: is driven by a ballistic mechanism due to the unusually long lifetime of hot particles
853: in systems such as Al/Pt or Ti/Pt with a large mass ratio.
854: The interfacial mixing for the constituents are decoupled in time: predominantly
855: Pt mixes to the overlayer by ballistic diffusion (cascade mixing) and
856: the light constituent (Al or Ti) starts to
857: migrate to the bulk with some time delay.
858: We find that the energetic ligth atoms are backscattered from the interface
859: due to the large mass difference.
860: Hence the intermixing of them is delayed significantly.
861:
862: We find no sign of sensitivity of ion beam mixing to thermochemical properties (heat of mixing).
863: Instead strong random ballistic diffusional features rule the mechanism with a long lifetime of hot particles
864: hence with an enhanced cascade mixing properties.
865: A different mechanism is active for the light atom mixing in Ti/Pt and in Al/Pt.
866: In the former bilayer Ti is injected to the Pt bulk afterwards the TS by a vacancy mechanism
867: while the Al atoms intermix during the TS via a time delayed (retarded) ballistic diffusion.
868: Therefore, there is a great difference in the details which should be precisely studied
869: case by case.
870: Finally we would like to remark that in accordance with the theory of Martin \cite{Martin}
871: we find that ballistic diffusion and thermally activated vacancy movements
872: seem to be suitable
873: for understanding ion beam mixing.
874:
875:
876:
877:
878:
879:
880:
881: \section{acknowledgment}
882: {\small
883: This work is supported by the OTKA grants F037710, T043704
884: and T30430
885: from the Hungarian Academy of Sciences.
886: Computer time at the Center of Scientific Computing (NIIF) in
887: Budapest is gratefully acknowledged.
888: }
889:
890:
891:
892:
893: %\section{Conclusion}
894:
895: %\vspace{-2cm}
896:
897: \begin{thebibliography}{99}
898:
899:
900: \bib{Gnaser}
901: H. Gnaser, {\em Low-Energy Ion Irradiation of Solid Surfaces}, Solid-State Physics, {\bf 146} (1999), Springer.
902:
903:
904: \bib{Bellon}
905: G. Martin, P. Bellon, {\em Driven Alloys}, Solid State Phys., {\bf 50}, 189 (1997).
906:
907: \bib{Okamoto}
908: P. R. Okamoto, N. Q. Lam, L. E. Rehn, {\em Physics of Crystal-to-Glass Transformations}, Solid State Phys., {\bf 52}, 1 (1999).
909:
910: %\bib{Betz}
911: %G. Betz, G. K. Wehner, in {\em Sputtering by Particle Bombardment II},
912: %ed. by R. Behrisch (Springer, Berlin 1983).
913:
914: \bib{Samwer}
915: K. Samwer, H. J. Fecht and W. L. Johnson, {\em Amorphization in Metallic Systems},
916: in Glassy Metals III, eds. H. Beck, H.-J. Güntherodt, Springer, (1994).
917:
918: \bib{Enrique}
919: R. A. Enrique, P. Bellon, Phys. Rev. {\bf B60}, 14649 (1999).
920:
921: \bib{Mayr}
922: S. G. Mayr, R. S. Averback, Phys. Rev. {\bf B68}, 214105 (2003).
923:
924: \bib{Sigmund}
925: P. Sigmund, A. Gras-Marti, Nucl Instr. and Meth. {\bf 182/183}, 25 (1981).
926:
927: \bib{Roussel}
928: J-M. Roussel, P. Bellon, Phys. Rev. {\bf B65}, 144107-1 (2001).
929:
930: \bib{AverbackRubia}
931: R. S. Averback, T. Diaz de la Rubia, {\em Displacement Damage in Irradiated Metals and Semiconductors}, Solid State Pysics, {\bf 51}, 281 (1998).
932:
933:
934: \bib{Martin}
935: G. Martin, Phys. Rev. {\bf B30}, 1424 (1984).
936:
937: \bib{Cheng}
938: Y.-T. Cheng, Material Science Reports, {\bf 5}, 45 (1990).
939:
940: \bib{Kelly}
941: R. Kelly, A. Miotello, Nucl Instr. and Meth. {\bf B122}, 374. (1997)
942: A. Miotello, R. Kelly, {\em ibid}, 458. (1997)
943:
944:
945: \bib{TS}
946: T. Diaz de la Rubia, R. S. Averback, R. Benedek,W. E. King,
947: Phys. Rev. Lett. {\bf 59}, 1930 (1987).
948:
949:
950: \bib{Johnson}
951: W. L. Johnson, Y. T. Cheng, M. Van Rossum, and M.-A. Nicolet, Nucl Instr. and Meth. {\bf B7/8}, 657 (1985).
952:
953:
954: \bib{Vineyard}
955: G. H. Vineyard, Radiat. Eff. {\bf 29}, 245 (1976).
956:
957: \bib{mix_exp}
958: A. Crespo-Sosa, M. Munoz, J.-C. Cheang-Wong, A. Oliver, J. M. Saniger, J. G. Banuelos, Mat. Sci. Eng. {\bf B100}, 297 (2003),
959: W. Bolse, Mat. Sci. Eng. {\bf A253}, 194 (1998),
960: L. C. Wei, R. S. Averback,J. Appl. Phys. {\bf 81}, 613 (1997),
961: T. Weber, K. Lieb, J. Appl. Phys., {\bf 73}, 3499 (1993),
962: S.-J. Kim, M-A. Nicolet, R. S. Averback, D. Peak, Phys. Rev. {\bf B37}, 38 (1985),
963: T. A. Workman, Y. T. Cheng, W. L. Johnson, and M.-A. Nicolet, Appl. Phys. Lett. {\bf 50}, 1486 (1987).
964:
965:
966: \bib{Colla}
967: T. J. Colla, H. M. Urbassek, Phys. Rev. {\bf B63}, 104206 (2001).
968:
969: \bib{Gades}
970: H. Gades, H. M. Urbassek, Phys. Rev. {\bf B51}, 14559 (1995),
971: Nucl Instr. and Meth. {\bf B115}, 485 (1996).
972:
973: \bib{Nordlund}
974: K. Nordlund, R. S. Averback, Phys. Rev. {\bf B 59}, 20 (1999).
975:
976: \bib{Sule1}
977: P. S\"ule, M. Menyh\'ard, K. Nordlund, Nucl Instr. and Meth. {\bf B211}, 524 (2003),
978: condmat/0302262.
979:
980: \bib{Sule2}
981: P. S\"ule, M. Menyh\'ard, K. Nordlund, Nucl Instr. and Meth. B, accepted for publication,
982: condmat/0310238.
983:
984: \bib{Allen}
985: M. P. Allen, D. J. Tildesley, {\em Comupter Simulation of Liquids}, (Oxford Science Publications, Oxford 1989)
986:
987: %\bib{Mayr}
988: %S. G. Mayr, Y. Ashkenazy, K. Albe, and R. S. Averback, Phys. Rev. Lett. {\bf 90}, 055505-1 (2003), S. G. Mayr, R. S. Averback, Phys. Rev. {\bf B68}, 075419-1 (2003),
989: %{\bf B68}, 214105 (2003).
990:
991: \bib{Nordlund_ref}
992: PARCAS was written by K. Nordlund,
993: see e.g., K. Nordlund, R. S. Averback, Phys. Rev. {\bf B56}, 2421 (1997),
994: K. Nordlund, M. Ghaly, R. S. Averback, M. Caturla, T. Diaz de la Rubia, and J. Tarus, Phys. Rev. {\bf B57}, 7556 (1998).
995:
996: \bib{Shewmon}
997: P. G. Shewnon, {\em Diffusion in Solids}, McGraw-Hill Book, London, (1963).
998:
999: \bib{Sutton}
1000: A. P. Sutton, R. W. Balluffi, {\em Interfaces in Crystalline Materials}, Oxford Science Publications, Clarendon Press, Oxford, (1996).
1001:
1002: \bib{SRIM}
1003: J. F. Ziegler, 1996, SRIM-96 computer code, J. F. Ziegler, J. P. Biersack, U.
1004: Littmark, {\em The Stopping and Range of Ions in Solids}, Pergamon Press, (1985).
1005:
1006: \bib{CR}
1007: F. Cleri and V. Rosato, Phys. Rev. {\bf B48}, 22 (1993),
1008: M. S. Daw, S. M. Foiles, and M. I. Baskes, Mater. Sci. Rep. {\bf 9}, 251 (1993).
1009:
1010: \bib{Miedema}
1011: A. R. Miedema, Philips Tech. Rev. {\bf 36}, 217. (1976)
1012:
1013: \bib{Nordlund_PRB97}
1014: K. Nordlund, R. S. Averback, Phys. Rev. {\bf B56}, 2421 (1997).
1015:
1016: \bib{Wollenberger}
1017: H. Wollenberger, in {\em Vacancies and Interstitials in Metals}, edited by A. Seeger, D.
1018: Schumacer, W. Schilling, and J. Diehl (North-Holland, Amsterdam, 1970), p. 215.
1019:
1020: \bib{Allnat}
1021: A. R. Allnatt, A. B. Lidiard, {\em Atomic Transport in Solids}, Cambridge
1022: University
1023: Press, Cambridge, England, (1993)
1024:
1025: \bib{www}
1026: P. S\"ule, http://www.mfa.kfki.hu/$\sim$sule, 2004.
1027:
1028: \bib{Swalin}
1029: R. A. Swalin, {\em Thermodynamics of Solids}, Wiley, London, (1963).
1030:
1031: \bib{Gyulai}
1032: L. S. Hung, M. Nastasi, J. Gyulai, and J. W. Mayer, Appl. Phys. Lett., {\bf 42},
1033: 672 (1983).
1034:
1035: \bib{Buchanan}
1036: J. D. R. Buchanan, T. P. A. Hase, B. K. Tanner, P. J. Chen,
1037: L. Gan, C. J. Powel, W. F. Egelhoff, Jr., Phys. Rev. {\bf B66}, 104427 (2002),
1038: P. Gas, J. Labar, G. Clugnet, A. Kovacs, C. Bergman, P. Barna, J. Appl. Phys.,
1039: {\bf 90}, 3899 (2001).
1040:
1041: \bib{Ossi}
1042: P. M. Ossi, Surf. Sci., {\bf 554}, 1. (2004).
1043:
1044: \bib{Banhart}
1045: F. Banhart, E. Hern\'andez, M. Terrones, Phys. Rev. Lett., {\bf 90}, 185502-1, (2003).
1046:
1047: %\bib{Sule3}
1048: %P. S\"ule, unpublished.
1049:
1050:
1051: \end{thebibliography}
1052:
1053:
1054:
1055:
1056: \end{document}
1057: