ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nanoglass/introduction.tex
Revision: 3230
Committed: Tue Sep 25 19:23:21 2007 UTC (16 years, 9 months ago) by gezelter
Content type: application/x-tex
File size: 11017 byte(s)
Log Message:
Edits

File Contents

# Content
1 %!TEX root = /Users/charles/Desktop/nanoglass/nanoglass.tex
2
3 \section{Introduction}
4
5 Excitation of the plasmon resonance in metallic nanoparticles has
6 attracted enormous interest in the past several years. This is partly
7 due to the location of the plasmon band in the near IR for particles
8 in a wide range of sizes and geometries. (Living tissue is nearly
9 transparent in the near IR, and for this reason, there is an
10 unrealized potential for metallic nanoparticles to be used in both
11 diagnostic and therapeutic settings.) One of the side effects of
12 absorption of laser radiation at these frequencies is the rapid
13 (sub-picosecond) heating of the electronic degrees of freedom in the
14 metal. This hot electron gas quickly transfers heat to the phonon
15 modes of the lattice, resulting in a rapid heating of the metal
16 particles.
17
18 Since metallic nanoparticles have a large surface area to volume
19 ratio, many of the metal atoms are at surface locations and experience
20 relatively weak bonding. This is observable in a lowering of the
21 melting temperatures and a substantial softening of the bulk modulus
22 of these particles when compared with bulk metallic samples. One of
23 the side effects of the excitation of small metallic nanoparticles at
24 the plasmon resonance is the facile creation of liquid metal
25 droplets.
26
27 Much of the experimental work on this subject has been carried out in
28 the Hartland and von~Plessen
29 groups.\cite{HartlandG.V._jp0276092,Hodak:2000rb,Hartland:2003lr,Petrova:2007qy,Link:2000lr}
30 These experiments mostly use the technique of time-resolved optical
31 pump-probe spectroscopy where a pump laser pulse serves to excite
32 conduction band electrons in the nanoparticle and a following probe
33 laser pulse allows observation of the time evolution of the
34 electron-phonon coupling. Hu and Hartland have observed a direct
35 relation between the size of the nanoparticle and the observed cooling
36 rate using such pump-probe techniques.\cite{HuM._jp020581+} Plech
37 {\it et al.} have use pulsed x-ray scattering as a probe to directly
38 access changes to atomic structure following pump
39 excitation.\cite{plech:195423} They further determined that heat
40 transfer in nanoparticles to the surrounding solvent is goverened by
41 interfacial dynamics and not the thermal transport properties of the
42 solvent.
43
44 Since these experiments are often carried out in condensed phase
45 surroundings, the large surface area to volume ratio makes the heat
46 transfer to the surrounding solvent also a relatively rapid process.
47 In our recent simulation study of the laser excitation of gold
48 nanoparticles,\cite{VardemanC.F._jp051575r} we observed that the
49 cooling rate for these particles (10$^{11}$-10$^{12}$ K/s) is in
50 excess of the cooling rate required for glass formation in bulk
51 metallic alloys. Given this fact, it may be possible to use laser
52 excitation to melt, alloy and quench metallic nanoparticles in order
53 to form glassy nanobeads.
54
55 To study whether or not glass nanobead formation is feasible, we have
56 chosen the bimetallic alloy of Silver (60\%) and Copper (40\%) as a
57 model system because it is an experimentally known glass former and
58 has been used previously as a theoretical model for glassy
59 dynamics.\cite{Vardeman-II:2001jn} The Hume-Rothery rules suggest that
60 alloys composed of Copper and Silver should be miscible in the solid
61 state, because their lattice constants are within 15\% of each
62 another.\cite{Kittel:1996fk} Experimentally, however Ag-Cu alloys are a
63 well-known exception to this rule and are only miscible in the liquid
64 state given equilibrium conditions. Below the eutectic temperature of
65 779 $^\circ$C and composition (60.1\% Ag, 39.9\% Cu), the
66 solid alloys of Ag and Cu will phase separate into Ag and Cu rich
67 $\alpha$ and $\beta$ phases, respectively. This behavior is due to a
68 positive heat of mixing in both the solid and liquid phases. For the
69 one-to-one composition fcc solid solution, $\Delta H_{\rm mix}$ is on the order
70 of +6~kJ/mole.\cite{Ma:2005fk} Non-equilibrium solid solutions may be
71 formed by undercooling, and under these conditions, a
72 compositionally-disordered $\gamma$ fcc phase can be formed.
73
74 Metastable alloys composed of Ag-Cu were first reported by Duwez in
75 1960 and were created by using a ``splat quenching'' technique in
76 which a liquid droplet is propelled by a shock wave against a cooled
77 metallic target.\cite{duwez:1136} Because of the small positive
78 $\Delta H_{\rm mix}$, supersaturated crystalline solutions are
79 typically obtained rather than an amorphous phase. Higher $\Delta
80 H_{\rm mix}$ systems, such as Ag-Ni, are immiscible even in liquid
81 states, but they tend to form metastable alloys much more readily than
82 Ag-Cu. If present, the amorphous Ag-Cu phase is usually seen as the
83 minority phase in most experiments. Because of this unique
84 crystalline-amorphous behavior, the Ag-Cu system has been widely
85 studied. Methods for creating such bulk phase structures include splat
86 quenching, vapor deposition, ion beam mixing and mechanical
87 alloying. Both structural
88 \cite{sheng:184203} and dynamic\cite{Vardeman-II:2001jn}
89 computational studies have also been performed on this system.
90
91 Although bulk Ag-Cu alloys have been studied widely, this alloy has
92 been mostly overlooked in nanoscale materials. The literature on
93 alloyed metallic nanoparticles has dealt with the Ag-Au system, which
94 has the useful property of being miscible on both solid and liquid
95 phases. Nanoparticles of another miscible system, Au-Cu, have been
96 successfully constructed using techniques such as laser
97 ablation,\cite{Malyavantham:2004cu} and the synthetic reduction of
98 metal ions in solution.\cite{Kim:2003lv} Laser induced alloying has
99 been used as a technique for creating Au-Ag alloy particles from
100 core-shell particles.\cite{Hartland:2003lr} To date, attempts at
101 creating Ag-Cu nanoparticles have used ion implantation to embed
102 nanoparticles in a glass matrix.\cite{De:1996ta,Magruder:1994rg} These
103 attempts have been largely unsuccessful in producing mixed alloy
104 nanoparticles, and instead produce phase segregated or core-shell
105 structures.
106
107 One of the more successful attempts at creating intermixed Ag-Cu
108 nanoparticles used alternate pulsed laser ablation and deposition in
109 an amorphous Al$_2$O$_3$ matrix.\cite{gonzalo:5163} Surface plasmon
110 resonance (SPR) of bimetallic core-shell structures typically show two
111 distinct resonance peaks where mixed particles show a single shifted
112 and broadened resonance.\cite{Hodak:2000rb} The SPR for pure silver
113 occurs at 400 nm and for copper at 570 nm.\cite{HengleinA._jp992950g}
114 On Al$_2$O$_3$ films, these resonances move to 424 nm and 572 nm for the pure metals. For
115 bimetallic nanoparticles with 40\% Ag an absorption peak is seen
116 between 400-550 nm. With increasing Ag content, the SPR shifts
117 towards the blue, with the peaks nearly coincident at a composition of
118 57\% Ag. Gonzalo {\it et al.} cited the existence of a single broad
119 resonance peak as evidence of a mixed alloy particle rather than a
120 phase segregated system. Unfortunately, they were unable to determine
121 whether the mixed nanoparticles were an amorphous phase or a
122 supersaturated crystalline phase. One consequence of embedding the
123 Ag-Cu nanoparticles in a glass matrix is that the SPR can be shifted
124 because of the nanoparticle-glass matrix
125 interaction.\cite{De:1996ta,Roy:2003dy}
126
127 Characterization of glassy behavior by molecular dynamics simulations
128 is typically done using dynamic measurements such as the mean squared
129 displacement, $\langle r^2(t) \rangle$. Liquids exhibit a mean squared
130 displacement that is linear in time (at long times). Glassy materials
131 deviate significantly from this linear behavior at intermediate times,
132 entering a sub-linear regime with a return to linear behavior in the
133 infinite time limit.\cite{Kob:1999fk} However, diffusion in nanoparticles
134 differs significantly from the bulk in that atoms are confined to a
135 roughly spherical volume and cannot explore any region larger than the
136 particle radius ($R$). In these confined geometries, $\langle r^2(t)
137 \rangle$ approaches a limiting value of $3R^2/40$.\cite{ShibataT._ja026764r} This limits the
138 utility of dynamical measures of glass formation when studying
139 nanoparticles.
140
141 However, glassy materials exhibit strong icosahedral ordering among
142 nearest-neghbors (in contrast with crystalline and liquid-like
143 configurations). Local icosahedral structures are the
144 three-dimensional equivalent of covering a two-dimensional plane with
145 5-sided tiles; they cannot be used to tile space in a periodic
146 fashion, and are therefore an indicator of non-periodic packing in
147 amorphous solids. Steinhart {\it et al.} defined an orientational
148 bond order parameter that is sensitive to icosahedral
149 ordering.\cite{Steinhardt:1983mo} This bond order parameter can
150 therefore be used to characterize glass formation in liquid and solid
151 solutions.\cite{wolde:9932}
152
153 Theoretical molecular dynamics studies have been performed on the
154 formation of amorphous single component nanoclusters of either
155 gold,\cite{Chen:2004ec,Cleveland:1997jb,Cleveland:1997gu} or
156 nickel,\cite{Gafner:2004bg,Qi:2001nn} by rapid cooling($\thicksim
157 10^{12}-10^{13}$ K/s) from a liquid state. All of these studies found
158 icosahedral ordering in the resulting structures produced by this
159 rapid cooling which can be evidence of the formation of an amorphous
160 structure.\cite{Strandburg:1992qy} The nearest neighbor information
161 was obtained from pair correlation functions, common neighbor analysis
162 and bond order parameters.\cite{Steinhardt:1983mo} It should be noted
163 that these studies used single component systems with cooling rates
164 that are only obtainable in computer simulations and particle sizes
165 less than 20\AA. Single component systems are known to form amorphous
166 states in small clusters,\cite{Breaux:rz} but do not generally form
167 amorphous structures in bulk materials. Icosahedral structures have
168 also been reported in nanoparticles, particularly multiply twinned
169 particles.\cite{Ascencio:2000qy}
170
171 Since the nanoscale Ag-Cu alloy has been largely unexplored, many
172 interesting questions remain about the formation and properties of
173 such a system. Does the large surface area to volume ratio aid Ag-Cu
174 nanoparticles in rapid cooling and formation of an amorphous state?
175 Would a predisposition to isosahedral ordering in nanoparticles also
176 allow for easier formation of an amorphous state and what is the
177 preferred ordering in a amorphous nanoparticle? Nanoparticles have
178 been shown to have size dependent melting
179 transition,\cite{Buffat:1976yq} so we would expect a similar trend to
180 follow for the glass transition temperature.
181
182 In the sections below, we describe our modeling of the laser
183 excitation and subsequent cooling of the particles {\it in silico} to
184 mimic real experimental conditions. The simulation parameters have
185 been tuned to the degree possible to match experimental conditions,
186 and we discusss both the icosahedral ordering in the system, as well
187 as the clustering of icosahedral centers that we observed.
188