ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nanoglass/introduction.tex
(Generate patch)

Comparing trunk/nanoglass/introduction.tex (file contents):
Revision 3208 by chuckv, Thu Aug 23 17:47:31 2007 UTC vs.
Revision 3259 by gezelter, Fri Oct 12 21:21:04 2007 UTC

# Line 1 | Line 1
1 < \input{header.tex}
2 < \section{Introduction}
3 <
4 < \input{footer.tex}
1 >
2 >
3 > %!TEX root = /Users/charles/Desktop/nanoglass/nanoglass.tex
4 >
5 > \section{Introduction}
6 >
7 > Excitation of the plasmon resonance in metallic nanoparticles has
8 > attracted enormous interest in the past several years. This is partly
9 > due to the location of the plasmon band in the near IR for particles
10 > in a wide range of sizes and geometries. Living tissue is nearly
11 > transparent in the near IR, and for this reason, there is an
12 > unrealized potential for metallic nanoparticles to be used in both
13 > diagnostic and therapeutic settings.\cite{West:2003fk,Hu:2006lr} One
14 > of the side effects of absorption of laser radiation at these
15 > frequencies is the rapid (sub-picosecond) heating of the electronic
16 > degrees of freedom in the metal. This hot electron gas quickly
17 > transfers heat to the phonon modes of the particle, resulting in a
18 > rapid heating of the lattice of the metal particles.  Since metallic
19 > nanoparticles have a large surface area to volume ratio, many of the
20 > metal atoms are at surface locations and experience relatively weak
21 > bonding. This is observable in a lowering of the melting temperatures
22 > of these particles when compared with bulk metallic
23 > samples.\cite{Buffat:1976yq,Dick:2002qy} One of the side effects of
24 > the excitation of small metallic nanoparticles at the plasmon
25 > resonance is the facile creation of liquid metal
26 > droplets.\cite{Mafune01,HartlandG.V._jp0276092,Link:2000lr,Plech:2003yq,plech:195423,Plech:2007rt}
27 >
28 > Much of the experimental work on this subject has been carried out in
29 > the Hartland, El-Sayed and Plech
30 > groups.\cite{HartlandG.V._jp0276092,Hodak:2000rb,Hartland:2003lr,Petrova:2007qy,Link:2000lr,plech:195423,Plech:2007rt}
31 > These experiments mostly use the technique of time-resolved optical
32 > pump-probe spectroscopy, where a pump laser pulse serves to excite
33 > conduction band electrons in the nanoparticle and a following probe
34 > laser pulse allows observation of the time evolution of the
35 > electron-phonon coupling. Hu and Hartland have observed a direct
36 > relation between the size of the nanoparticle and the observed cooling
37 > rate using such pump-probe techniques.\cite{Hu:2004lr} Plech {\it et
38 > al.} have use pulsed x-ray scattering as a probe to directly access
39 > changes to atomic structure following pump
40 > excitation.\cite{plech:195423} They further determined that heat
41 > transfer in nanoparticles to the surrounding solvent is goverened by
42 > interfacial dynamics and not the thermal transport properties of the
43 > solvent.  This is in agreement with Cahill,\cite{Wilson:2002uq}
44 > but opposite to the conclusions in Reference \citen{Hu:2004lr}.
45 >
46 > Since these experiments are carried out in condensed phase
47 > surroundings, the large surface area to volume ratio makes the heat
48 > transfer to the surrounding solvent a relatively rapid process. In our
49 > recent simulation study of the laser excitation of gold
50 > nanoparticles,\cite{VardemanC.F._jp051575r} we observed that the
51 > cooling rate for these particles (10$^{11}$-10$^{12}$ K/s) is in
52 > excess of the cooling rate required for glass formation in bulk
53 > metallic alloys.\cite{Greer:1995qy} Given this fact, it may be
54 > possible to use laser excitation to melt, alloy and quench metallic
55 > nanoparticles in order to form glassy nanobeads.
56 >
57 > To study whether or not glass nanobead formation is feasible, we have
58 > chosen the bimetallic alloy of Silver (60\%) and Copper (40\%) as a
59 > model system because it is an experimentally known glass former, and
60 > has been used previously as a theoretical model for glassy
61 > dynamics.\cite{Vardeman-II:2001jn} The Hume-Rothery rules suggest that
62 > alloys composed of Copper and Silver should be miscible in the solid
63 > state, because their lattice constants are within 15\% of each
64 > another.\cite{Kittel:1996fk} Experimentally, however Ag-Cu alloys are
65 > a well-known exception to this rule and are only miscible in the
66 > liquid state given equilibrium conditions.\cite{Massalski:1986rt}
67 > Below the eutectic temperature of 779 $^\circ$C and composition
68 > (60.1\% Ag, 39.9\% Cu), the solid alloys of Ag and Cu will phase
69 > separate into Ag and Cu rich $\alpha$ and $\beta$ phases,
70 > respectively.\cite{Banhart:1992sv,Ma:2005fk} This behavior is due to a
71 > positive heat of mixing in both the solid and liquid phases. For the
72 > one-to-one composition fcc solid solution, $\Delta H_{\rm mix}$ is on
73 > the order of +6~kJ/mole.\cite{Ma:2005fk} Non-equilibrium solid
74 > solutions may be formed by undercooling, and under these conditions, a
75 > compositionally-disordered $\gamma$ fcc phase can be
76 > formed.\cite{najafabadi:3144}
77 >
78 > Metastable alloys composed of Ag-Cu were first reported by Duwez in
79 > 1960 and were created by using a ``splat quenching'' technique in
80 > which a liquid droplet is propelled by a shock wave against a cooled
81 > metallic target.\cite{duwez:1136} Because of the small positive
82 > $\Delta H_{\rm mix}$, supersaturated crystalline solutions are
83 > typically obtained rather than an amorphous phase. Higher $\Delta
84 > H_{\rm mix}$ systems, such as Ag-Ni, are immiscible even in liquid
85 > states, but they tend to form metastable alloys much more readily than
86 > Ag-Cu. If present, the amorphous Ag-Cu phase is usually seen as the
87 > minority phase in most experiments. Because of this unique
88 > crystalline-amorphous behavior, the Ag-Cu system has been widely
89 > studied. Methods for creating such bulk phase structures include splat
90 > quenching, vapor deposition, ion beam mixing and mechanical
91 > alloying. Both structural \cite{sheng:184203} and
92 > dynamic\cite{Vardeman-II:2001jn} computational studies have also been
93 > performed on this system.
94 >
95 > Although bulk Ag-Cu alloys have been studied widely, this alloy has
96 > been mostly overlooked in nanoscale materials. The literature on
97 > alloyed metallic nanoparticles has dealt with the Ag-Au system, which
98 > has the useful property of being miscible on both solid and liquid
99 > phases. Nanoparticles of another miscible system, Au-Cu, have been
100 > successfully constructed using techniques such as laser
101 > ablation,\cite{Malyavantham:2004cu} and the synthetic reduction of
102 > metal ions in solution.\cite{Kim:2003lv} Laser induced alloying has
103 > been used as a technique for creating Au-Ag alloy particles from
104 > core-shell particles.\cite{Hartland:2003lr} To date, attempts at
105 > creating Ag-Cu nanoparticles have used ion implantation to embed
106 > nanoparticles in a glass matrix.\cite{De:1996ta,Magruder:1994rg} These
107 > attempts have been largely unsuccessful in producing mixed alloy
108 > nanoparticles, and instead produce phase segregated or core-shell
109 > structures.
110 >
111 > One of the more successful attempts at creating intermixed Ag-Cu
112 > nanoparticles used alternate pulsed laser ablation and deposition in
113 > an amorphous Al$_2$O$_3$ matrix.\cite{gonzalo:5163} Surface plasmon
114 > resonance (SPR) of bimetallic core-shell structures typically show two
115 > distinct resonance peaks where mixed particles show a single shifted
116 > and broadened resonance.\cite{Hodak:2000rb} The SPR for pure silver
117 > occurs at 400 nm and for copper at 570 nm.\cite{HengleinA._jp992950g}
118 > On Al$_2$O$_3$ films, these resonances move to 424 nm and 572 nm for
119 > the pure metals. For bimetallic nanoparticles with 40\% Ag an
120 > absorption peak is seen between 400-550 nm. With increasing Ag
121 > content, the SPR shifts towards the blue, with the peaks nearly
122 > coincident at a composition of 57\% Ag. Gonzalo {\it et al.} cited the
123 > existence of a single broad resonance peak as evidence of an alloyed
124 > particle rather than a phase segregated system.  However, spectroscopy
125 > may not be able to tell the difference between alloyed particles and
126 > mixtures of segregated particles.  High-resolution electron microscopy
127 > has so far been unable to determine whether the mixed nanoparticles
128 > were an amorphous phase or a supersaturated crystalline phase.
129 >
130 > Characterization of glassy behavior by molecular dynamics simulations
131 > is typically done using dynamic measurements such as the mean squared
132 > displacement, $\langle r^2(t) \rangle$. Liquids exhibit a mean squared
133 > displacement that is linear in time (at long times). Glassy materials
134 > deviate significantly from this linear behavior at intermediate times,
135 > entering a sub-linear regime with a return to linear behavior in the
136 > infinite time limit.\cite{Kob:1999fk} However, diffusion in
137 > nanoparticles differs significantly from the bulk in that atoms are
138 > confined to a roughly spherical volume and cannot explore any region
139 > larger than the particle radius ($R$). In these confined geometries,
140 > $\langle r^2(t) \rangle$ approaches a limiting value of
141 > $3R^2/40$.\cite{ShibataT._ja026764r} This limits the utility of
142 > dynamical measures of glass formation when studying nanoparticles.
143 >
144 > However, glassy materials exhibit strong icosahedral ordering among
145 > nearest-neghbors (in contrast with crystalline and liquid-like
146 > configurations). Local icosahedral structures are the
147 > three-dimensional equivalent of covering a two-dimensional plane with
148 > 5-sided tiles; they cannot be used to tile space in a periodic
149 > fashion, and are therefore an indicator of non-periodic packing in
150 > amorphous solids. Steinhart {\it et al.} defined an orientational bond
151 > order parameter that is sensitive to icosahedral
152 > ordering.\cite{Steinhardt:1983mo} This bond order parameter can
153 > therefore be used to characterize glass formation in liquid and solid
154 > solutions.\cite{wolde:9932}
155 >
156 > Theoretical molecular dynamics studies have been performed on the
157 > formation of amorphous single component nanoclusters of either
158 > gold,\cite{Chen:2004ec,Cleveland:1997jb,Cleveland:1997gu} or
159 > nickel,\cite{Gafner:2004bg,Qi:2001nn} by rapid cooling($\thicksim
160 > 10^{12}-10^{13}$ K/s) from a liquid state. All of these studies found
161 > icosahedral ordering in the resulting structures produced by this
162 > rapid cooling which can be evidence of the formation of an amorphous
163 > structure.\cite{Strandburg:1992qy} The nearest neighbor information
164 > was obtained from pair correlation functions, common neighbor analysis
165 > and bond order parameters.\cite{Steinhardt:1983mo} It should be noted
166 > that these studies used single component systems with cooling rates
167 > that are only obtainable in computer simulations and particle sizes
168 > less than 20\AA. Single component systems are known to form amorphous
169 > states in small clusters,\cite{Breaux:rz} but do not generally form
170 > amorphous structures in bulk materials.
171 >
172 > Since the nanoscale Ag-Cu alloy has been largely unexplored, many
173 > interesting questions remain about the formation and properties of
174 > such a system. Does the large surface area to volume ratio aid Ag-Cu
175 > nanoparticles in rapid cooling and formation of an amorphous state?
176 > Nanoparticles have been shown to have a size dependent melting
177 > transition ($T_m$),\cite{Buffat:1976yq,Dick:2002qy} so we might expect
178 > a similar trend to follow for the glass transition temperature
179 > ($T_g$). By analogy, bulk metallic glasses exhibit a correlation
180 > between $T_m$ and $T_g$ although such dependence is difficult to
181 > establish because of the dependence of $T_g$ on cooling rate and the
182 > process by which the glass is formed.\cite{Wang:2003fk} It has also
183 > been demonstrated that there is a finite size effect depressing $T_g$
184 > in polymer glasses in confined geometries.\cite{Alcoutlabi:2005kx}
185 >
186 > In the sections below, we describe our modeling of the laser
187 > excitation and subsequent cooling of the particles {\it in silico} to
188 > mimic real experimental conditions. The simulation parameters have
189 > been tuned to the degree possible to match experimental conditions,
190 > and we discusss both the icosahedral ordering in the system, as well
191 > as the clustering of icosahedral centers that we observed.

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines