ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nanoglass/introduction.tex
Revision: 3221
Committed: Thu Sep 6 20:44:02 2007 UTC (16 years, 9 months ago) by gezelter
Content type: application/x-tex
File size: 10152 byte(s)
Log Message:
more writing

File Contents

# User Rev Content
1 chuckv 3208 \section{Introduction}
2 gezelter 3217
3     Excitation of the plasmon resonance in metallic nanoparticles has
4     attracted enormous interest in the past several years. This is partly
5     due to the location of the plasmon band in the near IR for particles
6     in a wide range of sizes and geometries. (Living tissue is nearly
7     transparent in the near IR, and for this reason, there is an
8     unrealized potential for metallic nanoparticles to be used in both
9     diagnostic and therapeutic settings.) One of the side effects of
10     absorption of laser radiation at these frequencies is the rapid
11     (sub-picosecond) heating of the electronic degrees of freedom in the
12     metal. This hot electron gas quickly transfers heat to the phonon
13     modes of the lattice, resulting in a rapid heating of the metal
14     particles.
15    
16     Since metallic nanoparticles have a large surface area to volume
17     ratio, many of the metal atoms are at surface locations and experience
18     relatively weak bonding. This is observable in a lowering of the
19     melting temperatures and a substantial softening of the bulk modulus
20     of these particles when compared with bulk metallic samples. One of
21     the side effects of the excitation of small metallic nanoparticles at
22     the plasmon resonance is the facile creation of liquid metal
23     droplets.
24    
25     Much of the experimental work on this subject has been carried out in
26     the Hartland and von~Plessen
27 gezelter 3221 groups.\cite{HartlandG.V._jp0276092,Hodak:2000rb,Hartland:2003lr,HuM._jp020581+,Petrova:2007qy,plech:195423}
28 gezelter 3217 They have [BRIEF SURVEY OF THE EXPERIMENTAL WORK]
29    
30    
31     Since these experiments are often carried out in condensed phase
32     surroundings, the large surface area to volume ratio makes the heat
33     transfer to the surrounding solvent also a relatively rapid process.
34     In our recent simulation study of the laser excitation of gold
35 chuckv 3220 nanoparticles,\cite{VardemanC.F._jp051575r} we observed that the cooling rate for these
36 gezelter 3219 particles (10$^{11}$-10$^{12}$ K/s) is in excess of the cooling rate
37 gezelter 3217 required for glass formation in bulk metallic alloys. Given this
38     fact, it may be possible to use laser excitation to melt, alloy and
39     quench metallic nanoparticles in order to form metallic glass
40     nanobeads.
41    
42     To study whether or not glass nanobead formation is feasible, we have
43     chosen the bimetallic alloy of Silver (60\%) and Copper (40\%) as a
44     model system because it is an experimentally known glass former and
45     has been used previously as a theoretical model for glassy
46 gezelter 3219 dynamics.\cite{Vardeman-II:2001jn} The Hume-Rothery rules suggest that
47     alloys composed of Copper and Silver should be miscible in the solid
48     state, because their lattice constants are within 15\% of each
49 gezelter 3221 another.\cite{Kittel:1996fk} Experimentally, however Ag-Cu alloys are a
50 gezelter 3219 well-known exception to this rule and are only miscible in the liquid
51     state given equilibrium conditions. Below the eutectic temperature of
52     779 $^\circ$C and composition (60.1\% Ag, 39.9\% Cu), the
53     solid alloys of Ag and Cu will phase separate into Ag and Cu rich
54     $\alpha$ and $\beta$ phases, respectively. This behavior is due to a
55     positive heat of mixing in both the solid and liquid phases. For the
56     one-to-one composition fcc solid solution, $\Delta H$ is on the order
57 chuckv 3220 of +6~kJ/mole.\cite{Ma:2005fk} Non-equilibrium solid solutions may be
58 gezelter 3219 formed by undercooling, and under these conditions, a
59     compositionally-disordered $\gamma$ fcc phase can be formed.
60 gezelter 3217
61 gezelter 3219 Metastable alloys composed of Ag-Cu were first reported by Duwez in
62     1960 and were created by using a ``splat quenching'' technique in
63     which a liquid droplet is propelled by a shock wave against a cooled
64     metallic target.\cite{duwez:1136} Because of the small positive
65     $\Delta H$, supersaturated crystalline solutions are typically
66     obtained rather than an amorphous phase. Higher $\Delta H$ systems,
67     such as Ag-Ni, are immiscible even in liquid states, but they tend to
68     form metastable alloys much more readily than Ag-Cu. If present, the
69     amorphous Ag-Cu phase is usually seen as the minority phase in most
70     experiments. Because of this unique crystalline-amorphous behavior,
71     the Ag-Cu system has been widely studied. Methods for creating such
72     bulk phase structures include splat quenching, vapor deposition, ion
73     beam mixing and mechanical alloying. Both structural
74     \cite{sheng:184203} and dynamic\cite{Vardeman-II:2001jn}
75     computational studies have also been performed on this system.
76 chuckv 3218
77 gezelter 3219 Although bulk Ag-Cu alloys have been studied widely, this alloy has
78     been mostly overlooked in nanoscale materials. The literature on
79     alloyed metallic nanoparticles has dealt with the Ag-Au system, which
80     has the useful property of being miscible on both solid and liquid
81     phases. Nanoparticles of another miscible system, Au-Cu, have been
82     successfully constructed using techniques such as laser
83     ablation,\cite{Malyavantham:2004cu} and the synthetic reduction of
84     metal ions in solution.\cite{Kim:2003lv} Laser induced alloying has
85     been used as a technique for creating Au-Ag alloy particles from
86 gezelter 3221 core-shell particles.\cite{Hartland:2003lr} To date, attempts at
87 gezelter 3219 creating Ag-Cu nanoparticles have used ion implantation to embed
88     nanoparticles in a glass matrix.\cite{De:1996ta,Magruder:1994rg} These
89     attempts have been largely unsuccessful in producing mixed alloy
90     nanoparticles, and instead produce phase segregated or core-shell
91     structures.
92 chuckv 3218
93 gezelter 3219 One of the more successful attempts at creating intermixed Ag-Cu
94     nanoparticles used alternate pulsed laser ablation and deposition in
95     an amorphous Al$_2$O$_3$ matrix.\cite{gonzalo:5163} Surface plasmon
96     resonance (SPR) of bimetallic core-shell structures typically show two
97     distinct resonance peaks where mixed particles show a single shifted
98     and broadened resonance.\cite{Hodak:2000rb} The SPR for pure silver
99     occurs at 400 nm and for copper at 570 nm. On Al$_2$O$_3$ films,
100     these resonances move to 424 nm and 572 nm for the pure metals. For
101     bimetallic nanoparticles with 40\% Ag an absorption peak is seen
102     between 400-550 nm. With increasing Ag content, the SPR shifts
103     towards the blue, with the peaks nearly coincident at a composition of
104     57\% Ag. The authors (WHO?) cited the existence of a single broad
105     resonance peak as evidence of a mixed alloy particle rather than a
106     phase segregated system. Unfortunately, they were unable to determine
107     whether the mixed nanoparticles were an amorphous phase or a
108     supersaturated crystalline phase. One consequence of embedding the
109     Ag-Cu nanoparticles in a glass matrix is that the SPR can be shifted
110     because of the nanoparticle-glass matrix
111     interaction.\cite{De:1996ta,Roy:2003dy}
112 chuckv 3218
113 gezelter 3219 Characterization of glassy behavior by molecular dynamics simulations
114     is typically done using dynamic measurements such as the mean squared
115     displacement, $\langle r^2(t) \rangle$. Liquids exhibit a mean squared
116     displacement that is linear in time (at long times). Glassy materials
117     deviate significantly from this linear behavior at intermediate times,
118     entering a sub-linear regime with a return to linear behavior in the
119 chuckv 3220 infinite time limit.\cite{Kob:1999fk} However, diffusion in nanoparticles
120 gezelter 3219 differs significantly from the bulk in that atoms are confined to a
121     roughly spherical volume and cannot explore any region larger than the
122     particle radius ($R$). In these confined geometries, $\langle r^2(t)
123 chuckv 3220 \rangle$ approaches a limiting value of $3R^2/40$.\cite{ShibataT._ja026764r} This limits the
124 gezelter 3219 utility of dynamical measures of glass formation when studying
125     nanoparticles.
126 chuckv 3218
127 gezelter 3219 However, glassy materials exhibit strong icosahedral ordering among
128     nearest-neghbors in contrast to crystalline or liquid structures.
129     Local icosahedral structures are the three-dimensional equivalent of
130     covering a two-dimensional plane with 5-sided tiles; they cannot be
131     used to tile space in a periodic fashion, and are therefore an
132     indicator of non-periodic packing in amorphous solids. Steinhart {\it
133     et al.} defined an orientational bond order parameter that is
134     sensitive to icosahedral ordering.\cite{Steinhardt:1983mo} This bond
135     order parameter can therefore be used to characterize glass formation
136 chuckv 3220 in liquid and solid solutions.\cite{wolde:9932}
137 chuckv 3218
138 gezelter 3219 Theoretical molecular dynamics studies have been performed on the
139     formation of amorphous single component nanoclusters of either
140     gold,\cite{Chen:2004ec,Cleveland:1997jb,Cleveland:1997gu} or
141     nickel,\cite{Gafner:2004bg,Qi:2001nn} by rapid cooling($\thicksim
142     10^{12}-10^{13}$ K/s) from a liquid state. All of these studies found
143     icosahedral ordering in the resulting structures produced by this
144     rapid cooling which can be evidence of the formation of a amorphous
145 gezelter 3221 structure.\cite{Strandburg:1992qy} The nearest neighbor information was
146 gezelter 3219 obtained from pair correlation functions, common neighbor analysis and
147     bond order parameters.\cite{Steinhardt:1983mo} It should be noted that
148     these studies used single component systems with cooling rates that
149     are only obtainable in computer simulations and particle sizes less
150     than 20\AA. Single component systems are known to form amorphous
151     states in small clusters,\cite{Breaux:rz} but do not generally form
152     amorphous structures in bulk materials. Icosahedral structures have
153     also been reported in nanoparticles, particularly multiply twinned
154     particles.\cite{Ascencio:2000qy}
155 chuckv 3218
156 gezelter 3219 Since the nanoscale Ag-Cu alloy has been largely unexplored, many
157     interesting questions remain about the formation and properties of
158     such a system. Does the large surface to volume ratio aid Ag-Cu
159     nanoparticles in rapid cooling and formation of an amorphous state?
160     Would a predisposition to isosahedral ordering in nanoparticles also
161     allow for easier formation of an amorphous state and what is the
162     preferred ordering in a amorphous nanoparticle? Nanoparticles have
163     been shown to have size dependent melting
164     transition,\cite{Buffat:1976yq} and we would expect a similar trend
165     to follow for the glass transition temperature.
166 chuckv 3218
167 gezelter 3219 In the sections below, we describe our modeling of the laser
168     excitation and subsequent cooling of the particles in silico to mimic
169     real experimental conditions. The simulation parameters have been
170     tuned to the degree possible to match experimental conditions, and we
171     discusss both the icosahedral ordering in the system, as well as the
172     clustering of icosahedral centers that we observed.
173 chuckv 3218