Anda di halaman 1dari 15

Materials Science and Engineering R 62 (2008) 175189

Contents lists available at ScienceDirect

Materials Science and Engineering R


journal homepage: www.elsevier.com/locate/mser

Nanostructures for photovoltaics


Loucas Tsakalakos *
General Electric Global Research Center, Niskayuna, NY 12309, USA

A R T I C L E I N F O

A B S T R A C T

Article history: Available online 15 August 2008 Keywords: Photovoltaics Nanostructures Nanocomposites Quantum wells Nanowires Nanoparticles

The use of various nanostructures in new solar cell designs and modes of enhancing conventional solar cells are described. The cell designs and enhancements are categorized by the type of nanostructure utilized. These include: (a) bulk nanostructured materials [3D]; (b) quantum wells [2D]; (c) nanowires [1D]; and (d) quantum dots/nanoparticles [0D]. The methods of fabricating such structures are rst described, followed by examples from the literature of how they have been utilized in a photovoltaic application. Scientic challenges associated with nanostructured photovoltaic devices are also discussed, followed by the prospects for use in real applications. 2008 Elsevier B.V. All rights reserved.

Contents 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1. High efciency conversion concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2. Thin lm solar cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Nanostructure-based concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1. Nanocomposites and nanostructured polycrystalline materials. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2. Quantum well solar cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3. Nanowires and tubes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4. Quantum dots and nanoparticles. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . General technical challenges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Summary and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175 176 178 178 178 179 180 182 187 187 188 188

2.

3. 4.

1. Introduction In recent years there has been a signicant, resurgent interest in renewable energy sources. This has been partially motivated by the increase in oil prices worldwide as a result of geopolitical and economic factors, and the general concern associated with global warming that is exacerbated by the emission of greenhouse gases during the production of primary power by conventional means [1,2]. While many technologies are being considered to supple-

* Tel.: +1 518 387 5715. E-mail address: tsakalakos@research.ge.com. 0927-796X/$ see front matter 2008 Elsevier B.V. All rights reserved. doi:10.1016/j.mser.2008.06.002

ment oil as a primary energy source, renewable energy sources are seen as the key to long-term weaning of industrialized economies from strict reliance on oil, coal, and natural gas. These include wind, fuel cells, solar cells, geothermal, biofuels, etc. Solar energy conversion is perhaps the most appealing of all these solutions, since the energy source is readily available. Indeed, in 1 h the sun radiates upon the earth as much energy as is used in 1 year by humanity [1]. Solar technologies also easily capture the understanding and imagination of the general public, as they are intuitively attractive owing to the abundant nature of the source. The concept of converting light to electricity was rst introduced with the discovery of the photovoltaic (PV) effect by Edmond Becquerel in 1839 [3]. However, the rst commercially

176

L. Tsakalakos / Materials Science and Engineering R 62 (2008) 175189

viable demonstration of a solar cell did not occur until over 100 years later, with the invention of the crystalline Si-based cell rst revealed to the world by researchers at Bell Labs in 1954 [4]. The rst Si cell had an efciency of 4% and since then researchers have been attempting to demonstrate, and companies to commercialize, solar cells with high conversion efciency and low manufacturing cost. In the ensuing 50 years, Si cells have been demonstrated with an efciency of nearly 25% [5], very close to the theoretical limit for a single junction under one sun illumination of 31% [6,7]. Such high efciency bulk Si solar cells are, unfortunately, prohibitively expensive for mass production. Indeed, most of todays commercial Si solar cells packaged in modules are 1417% efcient. There is a realization in the PV technical and customer/user communities that increasing cell efciency while decreasing cost will be critical if PV technology is to be widely utilized for primary or secondary energy needs. Silicon is not the ideal semiconducting material for solar energy conversion. The indirect bandgap of Si makes optical absorption inefcient due to the requirement of phonon emission/scattering with photons in order to conserve crystal momentum. As a result, the thickness of Si required to absorb 90% of sunlight (of all photons of energy above the bandgap) is 125 mm, whereas the thickness requirement for a direct bandgap material such as GaAs is 0.9 mm [8]. The key reason why Si is the leading material used in the PV industry today is the fact that it is the second most abundant element in the earths crust, making it a relatively inexpensive semiconductor (incidentally, it is most likely the most scientically and technologically studied material in the history of science). Furthermore, while silicon was still an emerging semiconductor material when the rst silicon solar cell was demonstrated, today there is a strong technological base for silicon due to the success of the electronics industry that can be directly applied to the mass manufacturing of Si-based solar cell technologies. Even though it is so abundant, it does require a minimum purity level in order to be useful for solar applications [9], though not to the level of purity required for the electronics industry. Until recently, silicon used in the PV industry today was recycled from the high performance, computing electronics industry, which lead to a shortage of suitable Si PV material. This trend continues, though there are now signicant investments occurring in Si wafer production factories for the solar industry [10,11].

The above discussion highlights three key questions facing PV: (1) How can the efciency of solar cells be increased to competitive levels with other energy sources? (2) How can the cost of solar cells be decreased to a level suitable rst for both secondary and primary power generation? (3) How can both of these goals be achieved in a single solar cell device/module technology and related manufacturing process? Fig. 1 shows the well-known plot, as dened by Green [12], of cell efciency vs. cost per unit area for the key general PV technologies, mainly bulk Si, thin lms, and the so-called Generation III concepts, the technologies of which are still being dened. Reference to this gure will be made in the subsequent discussion. The above questions lead to yet another question that is the central theme of this paper: can nanotechnology be used to address the above three questions, and if so, how? Owing to the fact that the $11 billion PV industry (in 2006, and estimated at $17B in 2008) has been growing at a rate of 3040% in the last few years and is expected to continue to do so in the next decade [13], it is of great interest for the PV research and industrial communities to address these questions. A brief summary of leading approaches to the rst two questions will be described, followed by a general description of how proposed approaches employing nanoscale structures are capable of answering the third question. 1.1. High efciency conversion concepts As noted above, the limiting efciency of a single junction is 31% [6]. Interestingly, early papers published shortly after Bell Labs demonstration of the rst Si cell estimated the efciency to be 21.6% [14] until Shockley and Queisser established the rst detailed balance theory showing 31% [6], which has since been modied by Kerr et al. [7] using additional calculations of Si solar cells incorporating Auger recombination to be 29%. If one performs a more generalized calculation of solar energy conversion efciency, it is possible to show that the thermodynamic limit is 85%, though as will be shown below we have not reached that level and will almost certainly never do so. The single bandgap thermodynamic efciency assuming blackbody radiation is 44%, the so-called ultimate efciency. The detailed balance limit efciency of 31% derived by Shockley and Queisser further takes into account the shape and refractive index of the solar cell, a

Fig. 1. Solar cell efciency as a function of cost per unit area showing the present status of the solar cells landscape. Generation I cells are based on bulk single and polycrystalline Si. Generation II cells are based on thin lms of various compositions. Generation III solar cells utilize various proposed energy conversion concepts. Only the multijunction Generation III solar cell has been demonstrated to date, though with a cost that puts it well outside the boundaries of Generation III cells as dened here [after Green [12], Fig. 1.1; with permission; copyright of Springer Science+Business Media, 2003]. (For interpretation of the references to color in the artwork, the reader is referred to the web version of the article.)

L. Tsakalakos / Materials Science and Engineering R 62 (2008) 175189

177

realistic solar spectrum, concentration, and radiative recombination. There are a number of well-described concepts for increasing the efciency of solar cells above the single junction limit (31%) so as to begin to approach the thermodynamic limit. These fall into the general category of Generation III solar cell concepts as dened by Green (see Fig. 1). Bulk Si solar cells are considered to be Generation I technologies, whereas Generation II cells are based on thin lm technologies that allow for the use of thinner absorber materials deposited on lower cost substrates and hence reduced cell cost, often at the expense of efciency (see below). Generation III cells are based on several new band structure and energy conversion concepts (to be described below) that have the potential to achieve limiting efciencies greater than the single junction limit. It is expected that such technologies, which are still emerging, will also be able to achieve cost levels similar to or better than Generation II cell technologies. The main reason why single junction solar cells are limited to 31% is that they do not absorb the signicant fraction (20%) of the photons in the solar spectrum that are below the bandgap in energy. These photons are simply lost. On the other hand, highenergy photons are lost due to thermalization of high-energy charge carriers in the conduction band duo to phonon scattering. This can be shown graphically by plotting the solar ux absorbed by a semiconductor as a function of bandgap energy and overlaying a plot of the work performed by a device that includes radiative recombination, as shown by Henry [15]. Such an analysis also yields the 31% efciency expected for the optimum bandgap of 1.351.45 eV at one sun concentration. If we add to this other device-level loss mechanisms such as reection, non-radiative recombination, it is apparent that the efciency quickly deteriorates from the 31% entitlement value. The multi-junction cell (MJC) allows absorption of a wider range of wavelengths in the solar spectrum by combining solar cells of varying bandgap in a series (tandem) stack. A generalized theory based on 24 and beyond (innite) number of bandgaps shows that the theoretical efciency for four junctions is 71% under maximum concentration [16]. This is the only high efciency solar cell concept that has been shown experimentally to work, with a 3-junction device recently performing at 40.7% under 240 sun illumination [17], the current world record. The typical structure for such cells is a multi-layer epitaxial thin lm stack grown on a Ge substrate with tunnel junctions in between to match the currents between each bandgap cells (Fig. 2) [18]. At present these cells are expensive (>$7 W1) and used in space applications, as well as terrestrial concentrator systems in solar power stations in which small area is necessary. Another leading concept for high efciency is the intermediate band (IB) solar cell [19]. Again, the idea is to absorb more of the solar spectrum, though in this case by introducing states within the bandgap of a semiconductor material such that low-energy photons can be absorbed in a two-photon process that promotes charge carriers to the conduction band (Fig. 3). Such a band structure would have a limiting efciency of 60% and thus has great promise. However, to date very few bulk materials have been shown to possess such intermediate states. One of these is the ZnCdTe system demonstrated by Yu et al. using bulk samples using a combined high-energy oxygen implantation and laser recrystallization process [20]. There are still challenges for such an approach: (1) a way of pn doping to separate charge carriers must be developed; (2) the problem of potentially signicant nonradiative recombination in such materials must be addressed; (3) assuming the above are solved, a manufacturable process would need to be developed. Other materials have been studied theoretically [21], but not demonstrated experimentally. The

Fig. 2. Schematic cross-section of a typical multi-junction (or tandem) solar cell employing three bandgaps, i.e. InGaP, GaAs, and Ge [courtesy of Ref. [18], with permission; copyright 2005, Wiley-Interscience].

use of quantum dot assemblies to form an effective intermediate band structure (mini-bands) is being pursued by Luque and coworkers [22,23], who have shown promising initial results. Technical challenges for this concept include absorption, charge transport, and manufacturability issues that must be addressed by further research (see below). At present, IB solar cell concepts are

Fig. 3. The intermediate band structure showing the valence band (VB), intermediate band (IB), and conduction band (CB), transitions between these bands (A), the quasi-Fermi levels for each band (e), and the chemical potentials dened as the difference between the quasi-Fermi levels associated with each band (m) [courtesy of Ref. [19], with permission; copyright 1997, American Physical Society; http://prola.aps.org/abstract/PRL/v78/i26/p5014_1].

178

L. Tsakalakos / Materials Science and Engineering R 62 (2008) 175189 Table 1 Record performance of various thin lm laboratory solar cells and corresponding modules based on 2005 values [adapted from Ref. [36]; see references therein for specic citations for these values provided, as many values were obtained from company web sites; see the journal Progress in Photovoltaics: Research and Applications for regularly updated PV efciency tables] Technology Record champion cell, h 24.7 21.2 19.9 16.5 10 Record module, h 16.9 13.3 11 7.6 6.4

in the early research stage, though are of great interest for future PV devices. A third, and perhaps longest-term, approach for high efciency solar cells is the concept of carrier multiplication. This implies that for a single absorbed photon, more than one electronhole pair is generated, mainly by an avalanche-type process employing a high local eld. Nozik rst proposed this theoretically [24]; recent photoluminescence experiments on PbSe and PbS nanocrystals have shown that it is indeed possible to multiply the number of excitons to 7 for a single absorbed photon [25]. The practical aspects of this, including charge separation [26], etc. require additional basic research. Finally, there are various efforts in up [27] or down [28] energy conversion of photons into a suitable energy regime for a single bandgap solar cell. This is primarily done by use of another layer integrated above or behind the cell. The layer typically consists of micron or nano-sized phosphor particles that absorb part of the spectrum and convert it to a more suitable energy for the solar cell in the composite system. To date, efforts in down-conversion have been successful using bulk materials and with limited success when using nanostructures, with attempts using Si nanocrystals [29] and quantum dots [30] being the main approaches. Recent work in up-conversion [31] has shown a small effect that is promising. While research continues in developing such ultra-high efciency Generation III solar cells, there are also signicant efforts to develop Generation II-type thin lm, single junction cells in order to decrease cost. 1.2. Thin lm solar cells Thin lms solar cells are widely recognized as a key solution to reducing the manufacturing cost of PV cells in the near to medium term. In 2007, thin lm solar cells comprised 58% of the PV market, and their share is growing. Thin lm solar cells are able to be produced at low cost by removing the bulk active substrate and using an additive deposition process on top of a low cost substrate such as glass, metal foil, plastic, etc. However, by using a thin lm the efciency is generally reduced compared to the bulk for two main reasons: (1) the thinner cell cannot absorb as much of the solar spectrum without more elaborate light trapping schemes or use of exotic compositions; (2) thin lm cells are typically polycrystalline or amorphous and therefore contain a signicantly higher density of non-radiative recombination centers at grain boundaries or in the bulk of the active semiconductor grains. Typical thin lm cells and thin lm cell-based modules, with a few exceptions (see below) [3234], have an efciency of less than 10%. [35,36]. A brief review of the leading technologies is now given. The leading technology is the CdTe system that is typically combined with CdS to form a heterojunction [37]. Such cells have been shown in the lab to produce a 16.5% power conversion efciency [32], though manufacturing efciency is expected to be on the order of 1011%. A suitable recycling program is also typically required with CdTe technologies. Another leading thin lm material is Cu(In,Ga)Se2 (CIGS). This materials system, which sometimes includes substitution of sulfur for selenium and has now been studied for over 20 years [38], has been shown to produce a lab efciency of 19.9% [39] and module efciencies on the order of 812%. CIGS is able to absorb >90% of the solar spectrum above its bandgap with a thickness of only 2 3 mm [40], and reasonable efciencies are also achievable with a 1mm thick lm [41]. The high cost and low abundance of In is a concern with CIGS in the long-term, as is the relatively high degradation observed under long-term exposure to the atmosphere [42] that requires high quality encapsulation to minimize.

Float-zone single crystal Si Multi/poly-crystalline Si CIGS thin lm CdTe thin lm a-Si (single junction) thin lm

Si-based thin lm cells have been explored in various formats, including polycrystalline and amorphous materials. The efciency of such cells has been 8% or less, making them suitable for specic PV applications, though hydrogenated amorphous silicon (a-Si:H) thin lms cells make up a large fraction of the thin lm PV market share. Efforts to grow either large grain lms or well-oriented (crystallographically) lms have been made, with limited success. Recently, tandem structures based on amorphous Si-based materials have yielded maximum efciencies of 15.5% in laboratory cells, with stabilized values of 13% [43]. It is noted that a-Si:H cells are subject to the so-called StaeblerWronski effect, in which the power conversion efciency of the cells degrades after initial exposure (1000 h) to sunlight until a stable lower efciency value is obtained [44]. This is related to the creation of additional point defects upon excitation of light that act as non-radiative recombination centers in the a-Si:H thin lm material. A summary of best results achieved to date for thin lm cells [36] and modules is given in Table 1. 2. Nanostructure-based concepts Many types of nanostructures have been applied to solar cells. The nanostructures to be discussed will be classied into four types: (a) nanocomposites [3D], (b) quantum wells [2D], (c) nanowire and nanotubes [(quasi) 1D], and (d) nanoparticles and quantum dots [(quasi) 0D]. These structures have been employed in various functions and for various performance/energy conversion enhancement strategies. What follows is a brief review of the historical development for the nanostructure class under consideration, the main applications beyond PV, and the major synthetic methods, followed by a critical review of the major works that have utilized the particular nanostructure type in a solar cell or in which fundamental measurements of key parameter(s) of interest to PV have been performed. Both the potential advantages of each nanostructure approach, as well as the disadvantages will be discussed, with an emphasis on possible future areas of research interest. 2.1. Nanocomposites and nanostructured polycrystalline materials Nanostructured materials and composites have been studied for many years, primarily for mechanical/load-bearing applications [45]. Metallic, ceramic, polymer, and hybrid systems have been explored. For example, it is well known that the strength of a bulk material is inversely proportional to the average grain size, i.e. the so-called HallPetch relationship. However, as the grain size is decreased below a critical material-dependent, value, the strength decreases due to grain boundary sliding [46]. Therefore, signicant work has been undertaken to pin grain boundaries and stabilize the nanoscale grain structure by adding secondary nanoparticle phases [47]. Metal nanomaterials are typically processed after

L. Tsakalakos / Materials Science and Engineering R 62 (2008) 175189

179

forging in a cold working processing such as twist extrusion which helps to rene the grain size by a factor of 210 per pass, or in powder form using cryomilling [48], followed by sintering. Ceramic nanomaterials may be formed by rst synthesizing nanopowders using a solgel process, solution-based chemical processes, laser pyrolysis of precursor gases, and other related methods [49]. The powder is typically cold-pressed followed by a high temperature sintering process, often in a hot isostatic press (HIP) [50]. It is also possible to directly deposit nanostructured ceramic/composite layers using plasma spray processes [51,52]. Similarly, polymer and hybrid organicinorganic systems have been demonstrated using various methods and strategies [53,54]. One of the rst examples of using 3D nanostructured materials for photovoltaic applications is the dye-sensitized tzel and ORegan solar cell (DSC), originally developed by Gra [55]. This cell utilizes organic dyes to absorb photons and then via electrochemical charge transfer processes separate electrons to a titanium dioxide conducting layer and positive charge within an electrolyte solution typically containing an iodine complex. The original work on this type of novel solar cell involved sintered titania colloidal lms, however, more recently this has been extended to nanostructured titania lm produced by sputtering [56] as well as electrodeposition [57]. Such structures, in combination with new absorbing molecules, have helped to improve the performance of DSCs (Fig. 4) [58,59] Since the invention of the DSC in 1990, efciency has improved from 2% to 11%, a value that has led some companies to attempt scale-up to the module level. To date the best DSC modules are on the order of 6.3% in power conversion efciency [60]. Furthermore, the presence of a liquid electrolyte in DSCs has stimulated efforts to develop polymer-based solar cells [61,62] (also see below). Biologically inspired processes have also been shown to produce inorganic nanostructured materials for hybrid solar cell applications [63]. tzel cell, more recent work has attempted In addition to the Gra to use all-inorganic nanocomposite structures. Nanu et al. [64] spray coated a composite of CuInS2 (CIS) and TiO2 nanoparticles onto a graphite/nanocrystalline titania electrode structure (Fig. 5), and completed the structure with a dense titania lm followed by a transparent conducting oxide (TCO). It is noteworthy that such a simple nanocomposite structure, with strong potential for manufacturability, achieved a power conversion efciency of 5%. The authors showed that the act of creating a nanocomposite of these two materials resulted in the observed efciency value, i.e. planar control samples with titania deposited on CIS resulted in poor performance. The detailed operation of this cell was described in subsequent work and showed that the efciency is reduced with decreasing titania particle size and increasing titania layer thickness [65]. Another form of bulk nanostructured solar cells is the use of nanocrystalline silicon (nc-Si) to replace amorphous Si or combine it with a-Si for tandem structures [66]. nc-Si is typically applied as a thin lm using plasma-enhanced chemical vapor deposition (PECVD). The major technical challenge with such nanocrystalline solar cells is the fact that the surface area of grain boundaries is very high, signicantly increasing the density of recombination centers as well as the probability of recombination due to charge carriers having to traverse so many boundaries. More work is required in understanding the nature of grain boundaries in nanostructured materials and strategies for effectively passivating them are necessary. The deposition rate of these materials is also relatively low, typically in the range of /s, making development of a high rate manufacturing 1 A process a challenging task.

tzel (dye-sensitized) solar cell with nano titania electron conductor: (a) Fig. 4. Gra typical microstructure (image width is 210 nm) of a nanocrystalline titania layer, and (b) typical current densityvoltage characteristics in dark and under AM1.5 conditions showing an efciency as high as 8% depending on whether there is only a sensitizer molecule present (device A) or if an additional hexadecylmalonic acid absorber molecule is co-grafted to the nano-titania particles [from Ref. [59], with permission; copyright 2003, American Chemical Society]. (For interpretation of the references to color in the artwork, the reader is referred to the web version of the article.)

2.2. Quantum well solar cells Quantum wells have been studied for several decades now for application in electronics and photonics. Their key feature is connement of charge carriers in 1D, creating a sheet of carriers with well-dened energy levels and high mobility owing to modication of the band structure [67]. These structures also have potential for high absorption due to a higher density of states at the band edge. In the past decade there has been a concerted effort to fabricate solar cells that incorporate (multi) quantum wells (MQW) in the active region of the device. These are primarily based on IIIV materials, mainly GaAs and related alloys such as AlGaAs and InGaAs [68]. The main expected benets of such cells are high short circuit current due to the enhanced absorption. Quantum wells are typically fabricated using metal organic chemical vapor deposition (MOCVD) or molecular beam epitaxy (MBE). These deposition processes allow precise, atomic level control of the thickness of specic alloy layer/composition. For example, it is possible to directly grow alternating layers of InAlAs/ ngstrom level precision (Fig. 6), such that the InGaAs with A

180

L. Tsakalakos / Materials Science and Engineering R 62 (2008) 175189

a MQW, the power conversion efciency is enhanced at high concentration though it is equal to or reduced compared to planar cells under AM1.5 conditions [58]. More recent work has shown [7476], however, that by incorporating a distributed Bragg reector (DBR) behind the MQW (Fig. 7b), it is possible to obtain an enhanced Voc compared to a planar control. This is due to photon recycling that allows collection of photons that have passed through the structure to be absorbed on multiple passes. While this effect only begins to manifest itself are concentrations greater than 50 suns, this is nonetheless an interesting possible direction of future development of MQW-based cells. 2.3. Nanowires and tubes The next level of conned nanostructures we explore is in two dimensions, i.e. nanowire and nanotubes, providing quasi or true 1D structures. Nanowires have been extensively studied in the last decade, though their roots can be traced to early work on whisker growth performed at Bell Labs by Wagner and Ellis [77]. In the 1970s and 1980s there was much work performed on synthesis of various whisker materials [78], though literature on the top down fabrication and low temperature testing of quantum wires also began to appear [79]. In the 1990s there was pioneering research performed by Haraguchi et al. at Hitachi on InAs/GaAs-based nanowires that culminated in the demonstration of a light emitting diode (LED) based on an array of such structures [80]. Concurrently, initial synthesis of Ge nanowires was also demonstrated at IBM by Heath and LeGoues [81]. In 1998, Morales and Lieber published a seminal paper on synthesis of various nanowire materials using pulsed laser ablation [82], which re-ignited strong interest in the eld of nanowires. By 2000 both Lieber and coworkers [83] and Heath and co-workers [84] demonstrated eld effect transistors using silicon nanowires, and since there has been an explosion of research in this eld. Carbon nanotubes (CNTs) were rst discovered by Iijima and Ichihashi [85] in 1991 and are perhaps the prototypical nanomaterial. They are essentially graphene sheets that have been rolled into the form of a tube [86]. Single-walled nanotubes (SWNT) or multi-walled nanotubes (MWNT) may be formed. They may be synthesized by arc discharge methods [87], chemical vapor deposition [88], chemical [89], and laser ablation [90] processes. Nanotubes are typically characterized by their chirality, which is dened by the 2D lattice vectors that relate the location where the graphene sheet is bonded [91]. CNTs are being explored for numerous applications, including conventional electronics [92], sensors [93], RF electronics [94], eld emitters [95], cancer therapy [96], and many others. More recently, nanotubes of inorganic materials such as GaN [97], Pb(Zr,Ti)O3 [98], titania [99], and alumina [100] have been demonstrated. Nanowires (NW) are elongated solid nanostructures. They may be synthesized by many methods, including laser ablation [82] (as noted above), chemical vapor deposition (Fig. 8) [77], solutionbased chemical synthesis [101], critical point synthesis [102], electrochemical deposition [103], wet etching [104], and dry etching of nanopatterned features [105]. These methods allow for the synthesis of many classes of functional materials, including group IV, IIIV, and IIVI semiconductors, transition metal carbides [106], metals, dielectrics, ferroelectrics, etc. They may be single crystalline, polycrystalline [107], or contain multilayers. Therefore, many applications of nanowires have been explored. These include electronics (transistors, etc.), biological and chemical sensors, light emitting diodes, laser, thermoelectrics, magnetic storage media, structural applications, and numerous others. The application of nanowires (and nanorods, dened here as elongated nanostructures with an aspect ratio of less than or equal

Fig. 5. (a) Schematic of a 3D all-inorganic CuInS2/TiO2 nanocomposite solar cell fabricated by spray coating and atomic layer deposition and (b) typical device performance with an efciency up to 5% [from Ref. [64], with permission; copyright 2005, American Chemical Society]. (For interpretation of the references to color in the artwork, the reader is referred to the web version of the article.)

potential barrier between each layer creates quantum-conned states [69]. Owing to the relatively high cost nature of the described processes, it is expected that application of quantum well-based solar cells will be limited to space or concentrator PV systems. Nevertheless, such cells constitute a prototypical quantum conned nanostructured system, and as such are of great fundamental interest. Barnham and co-workers were among the rst to study quantum well solar cells (QWSC) [70]. The reason for a potential increase in efciency using quantum wells is that absorption in quantum wells is very high due to the carrier density obtained by quantum connement in the plane of the well. The work by Barnham and co-workers has shown that while it is possible to obtain a short-circuit current density enhancement in such cells, there is often a corresponding reduction in the open-circuit voltage. This is associated with recombination of quantum conned charge carriers as they attempt to traverse the multiple wells, as well as defects (dislocations, etc.) that are typically present within the MQW structure (Figs. 7 and 8). There is indeed some debate regarding the expected efciency enhancement in QWSCs, owing to the difculty in accurately calculating the increase in absorption as well as the increase in recombination (dark current) in these structures. Some workers have argued that one cannot expect an enhanced efciency for QWSCs [71], whereas others have developed detailed recombination models QWSCs [72] and calculated an enhancement of 20% (relative) for AlGaAs/GaAs QWSCs compared to baseline GaAs cells [73]; the main difference in these approaches being assumptions on the quasi-Fermi level spatial prole within the quantum wells. In general, when comparing single junction GaAs solar cells with ones containing

L. Tsakalakos / Materials Science and Engineering R 62 (2008) 175189

181

Fig. 6. Typical structure of InGaAsNSb/GaAs multi-quantum wells (MQW) as viewed by transmission electron microscopy: (a) InGaAs/GaAs, (b) InGaAsSb/GaAs, (c) InGaAsN/ GaAs, and (d) InGaAsNSb/GaAs, MQWs. The data shows that the introduction of Sb suppresses 3D growth, leading to less defects in the MQW stack [from Ref. [68], with permission; copyright 2001, American Institute of Physics].

to 5:1) to solar cells has been attempted in several device congurations and materials systems. The earliest demonstration was from Alivisatos and co-workers in 2002, in which CdSe nanorods were utilized as the electronic conducting layer of a hole conducting polymermatrix solar cell (Fig. 9) [108]. These cells produced an efciency of 1.7% for AM1.5 irradiation. The nanowires act as a direct path for transport of charge to the anode without the presence of grain boundaries, thus leading to an enhanced performance compared to solar cells employing nanostructures with aspect ratios approaching 1:1. Several variants of this cell were subsequently demonstrated, with a power conversion efciency of 1.8% for branched nanostructures [109]. Similar structures have been demonstrated for dye sensitized solar cells (DSC) using smooth [110] and dendritic [111] ZnO nanowires. The arguments for the use of nanowires are similar to that described above; they provide a more direct path for charge transport to the contacts, whereas dendritic nanowire structures also allow for enhance light harvesting. For example, Yang and coworkers [110] have shown a 2 increase in the short circuit current density compared to a nanocrystalline titania-based DSC, with a maximum power conversion efciency under AM1.5 light of 1.5%. Similarly, dendritic ZnO nanowire-based dye sensitized cells have been shown by Aydil and Baxter [111] to produce an improvement of 100 in photocurrent and efciency compared to smooth nanowires, with a power conversion efciency of 0.5% (Fig. 10). In addition to the benet of more efcient charge transport compared to other nanostructured materials congurations, nanowires also offer the potential for enhanced optical absorption

characteristics. Tsakalakos et al. [112] have shown this effect directly in Si nanowires fabricated directly on fused silica substrates. The nanowires were fabricated by rst bonding a single crystal Sih1 0 0i substrate to glass using a thermocompression bond. The Si wafers were then etched down to a thickness of between 1 and 20 mm in a KOH bath to obtain a single crystal Si lm. The lm was textured on the micron scale by the KOH etch. Nanowires were formed in the bulk Si lm using a wet etching chemistry [104] that nucleates Ag nanoparticles on the Si surface that create a nanoscale electrochemical cell which drives an electrical eld along the Sih1 0 0i direction. It was found by directly comparing the optical properties of solid Si lms to nanowire lms (produced in the same sample) that the optical reectance decreases signicantly, as expected due to the better index matching of the nanowire array to air. However, in combination with total transmission measurements it was shown that the optical absorption was enhanced signicantly compared to the solid lm over all the wavelengths of interest for photovoltaics (Fig. 11). Similarly, high broadband optical absorption for chemical vapor deposited (CVD) Si nanowire lms was observed. It was argued this is mainly the result of strong subwavelength scattering from individual nanowires that results in light trapping within the nanowire ensemble lm. This work clearly shows how sub-wavelength optical phenomena may act to enhance the properties of nanostructured solar cells. Furthermore, this author [113,114] and others [115117] proposed to use nanowires in an all-inorganic solar cell structures. This has the potential for producing cells with both improved efciency as well as greater stability compared to organic-based cells. Alivisatos and co-workers [115]

182

L. Tsakalakos / Materials Science and Engineering R 62 (2008) 175189

Fig. 7. (a) Schematic of InGaAs/GaAsP MQWs employed in a pin strain-balanced solar cell conguration [from Ref. [70], with permission; copyright 1999, American Institute of Physics] and (b) demonstration of enhanced efciency as a function of light concentration for quantum well solar cells employing a distributed Bragg reector (DBR) behind the MQW structure. Sample Qt1850U is a 50-well strainbalanced quantum well solar cell with no DBR, whereas sample Qt1850D is a 35well strain-balanced quantum well solar cell with a 20.5 period DBR centered at 920 nm [from Ref. [75], with permission; copyright 2005, Elsevier B.V.].

demonstrated an all-inorganic nanorod solar cell consisting of CdSe nanorods deposited on a layer of CdTe nanorods (Fig. 12). These cells produced a power conversion efciency of 1% and upon further sintering of the nanowire composite lm (to allow for better adhesion/bonding of the nanowire interfaces), yielded cells with an efciency of 3%. Silicon nanowire-based solar cells have been proposed both by this author [113] and by others [116118]. Our own calculations, which assume conventional semiconductor physics (no quantum effects), and those in Ref. [116], have shown that the use of Si nanowire-based solar cells provides an efciency entitlement on the order of 1518% depending on nanowire size (diameter and length) and quality (carrier lifetime). One of the key aspects of such all-inorganic Si nanowire solar cells is that it is possible to form pn junction conformally about the nanowire surface in a high-density array. This has the benet of decoupling the absorption of light from charge transport by allowing lateral diffusion of minority carriers to the pn junction, which is at most 50200 nm away, rather than many microns away as in Si bulk solar cells (Fig. 13). Furthermore, the nanowires may be synthesized using standard techniques such as chemical vapor deposition (CVD), with the possibility of direct growth on exible substrates such as metal foil. Finally, the use of CVD grown nanowire structures may yield solar cells with an improved cost benet ($/W) compared to bulk solar cells owing to the lower materials consumption (only gases are

used for fabricating the active material), yet comparable efciency to bulk crystalline Si. We have demonstrated initial Si nanowire solar cells using the etch technique described above [119]. The wires were etched into a bulk h1 0 0i-oriented p-type Si substrate and the pn junction was formed by depositing n-type amorphous silicon (a-Si) on the nanowire array. Fig. 14 shows currentvoltage characteristics for a Si NW solar cells and a planar Si solar cell fabricated in the same process. The planar cell had an efciency of 13%, whereas the NW cell had an efciency of 1.3%. This loss in efciency can be explained by the quantum efciency curves shown in Fig. 14. It is clear that there is a loss in the short wavelength response for the nanowire arrays. We believe this is due to the presence of surface states that cannot be fully passivated owing to the high-density of the nanowires, which inhibits deposition of the a-Si beyond a few hundred nanometers from the surface of the array. This highlights the need to trade-off nanowire density with surface passivation in these novel cells. It is noted, however, that the superior optical properties of nanowire cells compared to the planar control samples is evident. We have recently also demonstrated large area (11.8 cm2) CVD-grown (650 8C) silicon nanowire cells on metal foil substrates (and hence promise for roll-to-roll type processing) with a promising photocurrent of 1.6 mA/cm2 (conversion efciency of 0.1%) and broadband external quantum efciency response [114]. Recent measurements by Kelzenberg et al. [120] on individual silicon nanowires grown by CVD at relatively high temperature (1050 8C using SiCl4) have shown that the diffusion length in Si NWs can be as 2 mm, which corresponds to a minority carrier lifetime of 15 ns, and may have implication on future design of large area Si NW PV devices. The apparent power conversion efciency for these single wire devices was 0.46%. In summary, nanowires and nanorods show great promise for future solar cell devices. Remaining technical challenges include proper surface passivation, shunting, and high quality contacts. Carbon nanotubes have also been shown to yield a photovoltaic effect. Lee et al. [121] showed that an individual carbon nanotube diode electrostatically doped in a split-gate eld effect transistor conguration is an ideal pn junction, with an ideality factor of 1. It was argued this is due to the fact that there are no surface states on a nanotube since the carbon bonds are well satised in the graphene structure of the CNT. Such a device was then shown to possess a small photovoltaic effect (5% estimated power conversion efciency) [122], and the role of exciton transitions in the photoconductivity spectra of CNT devices, as well as the impact of bandgap renormalization and minority carrier diffusion (n and p), was further elucidated [123]. Practical implementation of CNT-only based large-area PV devices has not been shown to date, though the use of CNTs in polymer matrix PV devices [124] was shown to yield an improvement of efciency of 1000 times to a level of 0.04% [125]. This is hampered by the fact that CNTs grow in chiralities that yield both semiconducting nanotubes of varying bandgap, as well as metallic nanotubes. The problem of separating CNT by their electronic structure remains a challenge to further application of these nanostructures in PV application, though recent work in separation is promising [126]. Carbon nanotubes are also being explored as electrodes for solar cells [127]. 2.4. Quantum dots and nanoparticles The so-called 0D structures have been implemented in various PV applications and enhancement schemes. It is possible to synthesize quantum dots (QD) and nanoparticles (NP) in many compositions, including semiconductors and metals, as well as coat them with dielectrics or additional semiconductors to create coreshell nanoparticles. Quantum dots may be synthesized using

L. Tsakalakos / Materials Science and Engineering R 62 (2008) 175189

183

Fig. 8. Schematic of the vaporliquidsolid (VLS) growth mechanism of nanowire growth, in this case for silicon nanowire synthesis. This is typically implemented in laser ablation, CVD, or MOCVD growth processes. (For interpretation of the references to color in the artwork, the reader is referred to the web version of the article.)

solution-based chemical methods, chemical vapor methods, and physical vapor deposition. Pioneering work in the eld was performed by Bawendi and co-workers [128] and Alivisatos and co-workers [129]. Quantum dots have been used in many biological labeling applications [130], nanoelectronic applications [131], and others. Here, we review the major efforts towards utilization of quantum dots and nanoparticles in solar cells. One of the early efforts in the utilization of quantum dots was in the down-conversion of high-energy photons [132]. In a typical solar cell, high-energy photons are excited well beyond the conduction band edge, where they are mostly lost due to interaction with phonons (thermalization). Down-conversion is aimed at absorbing these high-energy photons and shifting them to lower energies that are matched to the particular absorber material in the solar cell. Photoluminescence-based enhancement (see below) of CdTe thin lm solar cells has been demonstrated using bulk crystal slabs [133]. Several schemes have been proposed to capture and convert them to lower energy photons that are better tuned to the bandgap of the semiconductor using nanostructures. Svrcek et al. attempted to implement such a scheme by forming Si quantum dots in a dielectric layer deposited on top of a standard Si solar cell [134]. They observed a small enhancement in quantum efciency at short wavelengths, however, the overall efciency was not enhanced. This was attributed to losses associated with non-radiative recombination at surface states due to incomplete surface passivation. The basic downconverting properties of silicon nanostructures have also been studied and similar to the above report, show a dependence on passivation [135]. The fundamental mechanism of down-conversion is the absorption of high-energy photon with relaxation into intermediate states within the bandgap, thus emitting two lower energy photons. Therefore, the maximum expected quantum efciency of down-conversion is 200%. To date very few materials have shown a quantum efciency is greater than one [136]. Another way to achieve spectral down-conversion is using quantum dots as photoluminescent converters, as has been shown, for example, with CdSe quantum dots by van Sark et al. [137].

Calculations showed an expected relative efciency improvement of 10% for npp+ multicrystalline Si (mc-Si) solar cells based on experimental absorption and luminscence data of 4.3 nm diameter CdSe quantum dots. Experimental verication of such an improvement was not shown, however. Quantum dots have also been applied to luminescent concentrator systems [138]. In general, improving the performance of a down converter and hence a down-converter composite cell system will require careful engineering of the band structure, associated transition rates, and defects. The opposite of down-conversion is the up-conversion of photons of energy below the bandgap that are typically not absorbed by the semiconductor [139]. There have been recent demonstrations of using micron scale phosphor particles to enhance the quantum efciency of a solar cell [140], and some reports of synthesis of up-converting nanoparticles phosphors [141143]. The typical mechanism of up-conversion involves absorption of sub-bandgap light into an intermediate state, followed by further absorption of a second photon to the conduction band edge. The excited charge carrier then relaxes back to the valence band edge, emitting a single higher energy photon. There are, however, multiple up-conversion mechanisms that have been identied, including various energy transfer schemes from a state associated with a dopant ion to those associated with the host lattice (Fig. 15). The theoretical quantum efciency of such a process is 50%. Recent work on nanoparticlebased up-converters has shown that the luminescence efciency is on the order of 1% and is strongly inuenced by particle size and annealing conditions (Fig. 15). One of the principal issues with such phosphors at the nanoscale is the loss in efciency associated with non-radiative recombination at the surface as the size of the particles is decreased. This eld is still in its infancy and further work in this area is expected in the future. Another novel band structure that can be obtained with quantum dots is the intermediate band (IB), which is indeed a form of up-conversion. This is possible because the states of closely spaced QDs can overlap to form an effective band structure

184

L. Tsakalakos / Materials Science and Engineering R 62 (2008) 175189

Fig. 9. Hybrid organicinorganic nanorod solar cells demonstrated by Alivisatos and co-workers showing (a) the structure of these devices, (b) the typical microstructure of such layers as viewed in a TEM, and (c) a typical IV characteristic under AM1.5 illumination with an efciency of 1.7% [from Ref. [108]; reprinted with permission from AAAS, copyright 2002].

(min-bands) that when nely tuned yields an intermediate band. Prototype cells have been fabricated using quantum dots [144] and a photocurrent attributed to IB was measured. The key challenge with such cells is the relatively low absorption crosssection of the QDs, which to date has limited performance and hindered demonstration of the full potential of intermediate bands in a practical manner. This and other promising IB approaches are expected to be studied in more detail in coming years. Multi-exciton generation (MEG) based solar cells is another promising application of quantum dots. MEG, the generation of more than one electronhole pair per photon, is a process that occurs in all semiconductors, though in the bulk the efciency of the process is very low [145]. However, Nozik and co-workers [146] and Klimov and co-workers [147] recently demonstrated that in semiconductor quantum dots the efciency of MEG can be very high, though MEG has not been shown in other materials systems [148]. This is fundamentally associated with quantum connement that leads to enhanced inverse Auger recombination (impact ionization) [149] and has been proposed to be due to superposition of excitonic states [150,146]. Indeed, it was shown that up to 7 excitons (bound) per photon can be produced in PbSe quantum dots [147]. Such processes fundamentally change the efciency limit, as compared to the ShockleyQueisser limit, in a

Fig. 10. Dye sensitized solar cell employing ZnO nanowires as the electron conducting phase; (a) schematic of the structure and (b) JV response of such cells as well as an inset of the internal quantum efciency [from Ref. [111], with permission; copyright 2005, American Institute of Physics]. (For interpretation of the references to color in the artwork, the reader is referred to the web version of the article.)

manner that the ultimate conversion efciency can increase to 65% from 44% for a bandgap below 1 eV [151]. The effect is typically observed above a specic photon energy, and can in principle be achieved under AM1.5 illumination. Challenges that remain include charge separation in the quantum dot layers and their transport to contacts. Several quantum dot-based solar cell designs have been proposed in the literature for harnessing the above mentioned high efciency mechanisms [26]. One of the promising approaches is the structure in which the quantum dots are embedded in a pi n structure, similar to the approach of Luque and co-workers [144]. Others include quantum dots as a replacement for organic dyes in a tzel-type (dye-sensitized) cell, and in an exciton recycling Gra conguration. Organometallic vapor phase epitaxy (OMVPE) grown InAs quantum dot solar cells on GaAs substrates have also recently been demonstrated by Raffaelle and co-workers [152], though an efciency degradation (associated with reductions in short-circuit current density, open-circuit voltage, and ll factor) from 14.7% to between 3.7% and 6.9% was reported that was minimized by employing a GaP strain-compensating layers to

L. Tsakalakos / Materials Science and Engineering R 62 (2008) 175189

185

Fig. 11. Total absorption from (a) a vertically aligned Si nanowire array formed by wet etching on a glass substrate, (b) a randomly oriented Si nanowire array formed by CVD on glass and (c) a solid Si thin lm on glass of the same thickness as that in (a) demonstrating enhanced broadband absorption for the nanowire array both above and below the bandgap of Si. The inset showed a picture of large area planar (left) and nanowire-based (right) Si PV devices showing enhanced optical appearance in the nanostructured device. [see Refs. [112,114] for more details]. (For interpretation of the references to color in the artwork, the reader is referred to the web version of the article.)

achieve an efciency of 9.5%. The introduction of InAs quantum dots provided an enhanced quantum efciency (QE) in the near IR range above the GaAs bandgap (9251025 nm). However, the QE was reduced in the 400700 nm range. This degradation was associated with introduction of mist dislocations originating at the quantum dot layer and threading to the surface of the cell within the GaAs matrix layer. The same group presented a discussion of single source solution-based synthesis of CdSe, CuInS2, and CuInSe2 QDs [153]; solar cell congurations as well as basic optical properties of quantum dot layers were discussed. In addition to MEG and IB band structures, Green and coworkers have proposed to use multi-layered Si quantum dot arrays to create an all Si-based multi-junction solar cell [154]. Multi-junction solar cells involve stacking of multiple bandgap cells, separated by tunnel junctions, to absorb a broader portion of the solar spectrum [155], as discussed above. The larger bandgap cell is grown on the top of increasingly lower bandgap cells. This is typically implemented in III-arsenide and phosphide systems with 24 bandgaps. The theoretical efciency for an innite number of bandgaps is 65.5% under one sun and 85% under maximum concentration of sunlight [16]. As noted, multi-

Fig. 12. All inorganic CdTe/CdSe nanorod heterostructure solar cells demonstrated by Alivastos and co-workers showing (a) the relevant band structure and a typical SEM image of a nanorod lm, and (b) typical currentvoltage data [Ref. [115]; Reprinted with permission from AAAS, copyright 2005].

junction solar cells are the only high efciency concept to have been successful to date, with a record solar cell efciency of 40.7% (under concentration) having been recently demonstrated by Boeing Spectrolab using a structure containing three bandgaps [17]. Work towards four and even six bandgaps is in progress. The concept described in Ref. [154] is novel in that it uses Si quantum dots of increased diameter with depth into the structure to create a multi-bandgap effect. Work to date has focused on basic synthesis of such structures. One of the challenges with the present implementation is the fact that the quantum dots are embedded in a dielectric medium, making charge transport more difcult, even if the quantum dots were to be well coupled. Nevertheless, the materials growth results are very promising. This work further points to the general issues of charge transport in quantum dot arrays, which will be discussed below.

Fig. 13. Schematic of the decoupling of light absorption and charge transport in a nanowire solar cell. (For interpretation of the references to color in the artwork, the reader is referred to the web version of the article.)

186

L. Tsakalakos / Materials Science and Engineering R 62 (2008) 175189

Fig. 14. Currentvoltage, quantum efciency, and optical reectance data of a typical Si nanowire solar cell fabricated using wet etching of bulk Si substrates (Ref. [119]). (For interpretation of the references to color in the artwork, the reader is referred to the web version of the article.)

The nal mechanism associated with nanoparticles that is to be discussed is that of plasmonics [156,157]. Surface plasmonpolaritons (SPPs) are collective oscillations of the Fermi Sea in metallic materials [158]. In metallic nanostructures/particles these SPPs are resonant at frequencies that are size- and shapedependent [159161]. Furthermore, the resonant SPPs created within the nanoparticles by absorption of light can produce electromagnetic elds with a very high local intensity, thus acting to provide a local concentrator effect. At slightly larger particle sizes (20100 nm), the particles effectively scatter light. Yu and coworkers [162] showed that when Au NPs are placed on top of a standard Si pn junction, the short-circuit current can be improved by a few percent, and peaks at wavelengths associated with SPP resonant frequency (Fig. 16). The effect is associated with the scattering from larger particles rather than from re-radiation of the electric eld, as the length scale over which the eld concentration occurs is on the order of 10s of nanometers. This is too short to be of relevance to standard Si p-n junction, in which the junction typically lies 0.21 mm below the surface. Yu and co-workers further showed enhanced photo-current effects in a-Si-based thin lm cells and that the concentration of nanoparticles plays a strong role in the observed photocurrent (and power conversion) enhancement [163]. There exists a peak about which the enhancement is highest. Below this peak the number of particles is simply too low to make an appreciable difference in the photocurrent due to the enhanced light scattering, and above this

Fig. 15. (a) Various up-conversion mechanisms and their corresponding efciencies (from Ref. [136] with permission) in which the vertical arrows imply a real transition, the curved arrows represent energy transfer between ions, APTE (addition de photon par transferts denergie), ETU (energy transfer up-conversion), GSA (ground state absorption), ESA (excited state absorption), and SHG (second harmonic generation) [from Ref. [136] with permission; copyright 2007, Elsevier B.V]; (b) TEM images of NaYF4:Yb,Er nanoparticles (45 nm diameter, the micron marker is 100 nm) and (c) up-conversion spectra under 980 nm illumination of as prepared nanoparticles (A) and after 400 8C (B), 600 8C (C), 700 8C (D) annealing, as well as large diameter (160 nm) particles (E) [from Ref. [142] with permission; copyright 2004, American Chemical Society].

concentration the particles reect too much light. The polarizability of the NPs further contributes strongly, and shaping the particles into ellipsoid or other shapes may further allow enhancement of certain optical modes for improved coupling into the underlying solar cell. The optimum size distribution for absorption of the solar spectrum by plasmonic nanoparticles has been calculated by Cole and Halas [161]. Silver nanoparticles have also been applied to silicon solar cells and showed an absorption enhancement of between 7 and 16 times in the wavelength range of 10501200 nm [164]. Returning to the issue of charge transport in quantum dots, in recent years there has been some work towards fabrication and testing of quantum dots lms for transistor application that shows promise for PV applications. Murray and Talapin fabricated PbSe QD lms and measured their transistor characteristics [165]. They obtained electron mobilities of up to 0.9 cm2/(V s) by linking the QDs with molecules that allow for efcient electron transfer

L. Tsakalakos / Materials Science and Engineering R 62 (2008) 175189

187

not well understood and require further addressing. This was made evident by the work of Mora-Sero et al. [168], who showed preliminary data that the electrical characteristics of nanostructured solar cells may be fundamentally different from conventional bulk or thin lm pn junction solar cells. The electrical characteristics of one polycrystalline thin lm Si solar cell and three nanostructure organic solar cells where analyzed in detail and found to be signicantly different. Specically, it was found that the frequency dependence of the capacitance often led to a negative capacitance, whereas the silicon-based solar cell remained positive. The detailed mechanisms of this was not elucidated, however, it was hypothesized that because nanostructured solar cells have a much higher surface area, the detailed physics of conversion are modied due to the increased capacitance. However, it is noted that in this work only organicbased nanostructured solar cells were analyzed, and further work on all-inorganic solar cells is required to obtain a more complete picture of the physics involved. This is because organic solar cells are exciton-based and transport occurs via hopping rather than drift and diffusion mechanisms. Nevertheless, this work points to the need for more research on this and related topics to obtain a more complete picture of the conversion and transport physics that can be subsequently applied to practical improvements in nanostructure-based solar cell performance. Another critical issue that must be addressed is defect passivation, since the very high surface area of nanostructure solar cells will lead to a large surface recombination velocity. In addition to fundamental physics issues discussed above, practical issues in fabricating nanostructures must be developed in order to cost effectively manufacture such devices in large volume and with high rates. A leading approach appears to be the use of solution processing to enable roll-to-roll processes. Other processes that hold promise due to their inherent scalability and low cost include chemical vapor deposition, electrochemical deposition, and sputtering. Further work in controlling nanostructure size, shape, and location with high rate processes is critical to enabling the next generation of advanced solar cells with low cost and high efciency. 4. Summary and outlook The major approaches to applying nanostructures to photovoltaics have been discussed, following a review of the status of standard solar cells based on silicon and thin lms, as well as the major high-efciency concepts that have been proposed in the literature. It was shown that 3D nanocomposite cells have already shown promise in a commercial context with dye-sensitized solar cells having shown a champion cell efciency of 11% and modules have been produced with an efciency of 6%. 3D inorganic nanocomposite cells have also shown a promising cell efciency of 5%. Quantum well solar cells have been studied for some time and have proven to be an excellent structure for understanding the fundamentals of applying quantum connement to photovoltaics as well as the effect of photon recycling. Various elongated nanostructures such as nanorods and nanowires have been applied to novel cell designs. Hybrid structure employing inorganic nanorods in an organic matrix have yielded efciencies on the order of 1.7%, whereas solution deposited allinorganic nanorod-based cells have produced efciencies of 3%. Large-area silicon nanowire-based solar cells have been demonstrated by this author with efciencies ranging from 0.01% to 1.3% depending on the structure and processing method employed, and individual silicon nanowire devices recently presented in the literature have yielded an apparent efciency of 3%. Individual carbon nanotube-based devices have been

Fig. 16. Demonstration of the use of plasmonic nanoparticles to enhance the photocurrent in PV devices; (a) schematic of the structures used to demonstrate the concept, (b) normalized photocurrent as a function of wavelength and (c) extinction efciency of nanoparticles of the same size used to show that the peaks directly correlate with plasmon resonance [from reference [162], with permission; copyright 2005, American Institute of Physics].

between dots. Similarly, Yu et al. also fabricated CdSe QD lms and obtained mobilities on the order of 0.01 cm2/(V s) [166]. These values are on the order of typical amorphous silicon thin lms and therefore show the potential for using QDs in advanced solar cell application. Efcient charge separation from quantum dots is also an area of fundamental interest. There has also been interesting work in energy transfer of excitons (not charge carriers) from small to large quantum dots in multi-layer quantum dot composite layers that may also be of interest to PV applications [167]. 3. General technical challenges The above discussion on charge transport in QDs highlights the general problem that the absorption, charge separation/conversion, and transport mechanisms in nanostructured solar cells are

188

L. Tsakalakos / Materials Science and Engineering R 62 (2008) 175189 [19] A. Luque, A. Marti, Phys. Rev. Lett. 78 (1997) 5014. [20] K.M. Yu, W. Walukiewicz, J. Wu, W. Shan, J.W. Beeman, M.A. Scarpulla, O.D. Dubon, P. Becla, Phys. Rev. Lett. 91 (2003) 246403. n, P. Palacios, K. Sa nchez, I. Aguilera, J.C. Conesa, in: Proceedings of the [21] P. Wahno 4th World Conference on Photovoltaic Energy Conversion, vol. 1, Hawaii, (2006), p. 63. `az, E. Can ovas, P.G. [22] A. Mart, E. Antol, C.R. Stanley, C.D. Farmer, N. Lopez, P. D Linares, A. Luque, Phys. Rev. Lett. 97 (2006) 247701. [23] A. Luque, A. Martin, in: Proceedings of the 4th World Conference on Photovoltaic Energy Conversion, vol. 1, Hawaii, (2006), p. 49. [24] A.J. Nozik, Annu. Rev. Phys. Chem. 52 (2001) 193. [25] R.D. Schaller, M. Sykora, J.M. Pietryga, V.I. Klimov, Nano Lett. 6 (2006) 424. [26] A.J. Nozik, Physica E 14 (2002) 115. rfel, J. Appl. Phys. 92 (2002) 4117. [27] T. Trupke, M. Green, P. Wu rfel, J. Appl. Phys. 92 (2002) 1668. [28] T. Trupke, M. Green, P. Wu [29] V. Svrcek, A. Slaoui, J.-C. Muller, Thin Solid Films 451 (2004) 384. [30] W.G.J.H.M. van Sark, A. Meijerink, R.E.I. Schropp, J.A.M. van Roosmalen, E.H. Lysen, Sol. Energy Mater. Sol. Cell 87 (2005) 395.e. [31] A. Shalav, B.S. Richards, M.A. Green, Sol. Energy Mater. Sol. Cell 91 (2007) 829 842. [32] X. Wu, J.C. Keane, R.G. Dhere, C. DeHart, D.S. Albin, A. Duda, T.A. Gessert, S. Asher, D.H. LKevi, P. Sheldon, in: Proceedings of the 17th European Photovoltaic Solar Energy Conference, 2001, p. 995. [33] B. Yan, G. Yue, J.M. Owens, J. Yang, S. Guha, in: Proceedings of the IEEE 4th World Conference on Photovoltaic Energy Conversion, vol. 2, 2006, p. 1477. [34] M.A. Contreras, B. Egass, K. Ramanathan, J. Hiltner, A. Swartzlander, F. Hasoon, R. Nou, Progr. Photovolt.: Res. Appl. 7 (1999) 311. [35] A. Shah, P. Torres, R. Tscharner, N. Wyrsch, H. Keppner, Science 285 (1999) 692. [36] B. von Roedern, K. Zweibel, H.S. Ullal, in: Proceedings of the 31st IEEE Photovoltaic Spec. Conference, 2005, p. 183. [37] B.M. Basol, in: Proceedings of the 21st IEEE Photovoltaic Specialists Conference, vol. 1, 1990, p. 588. [38] U. Rau, H.W. Schock, Appl. Phys. A: Mater. Sci. Process 69 (1999) 131. [39] National Renewable Energy Laboratory, <http://www.nrel.gov/news/press/ 2008/574.html>, March 24, 2008. [40] M.V. Yakushev, A.V. Mudry, V.F. Gremenok, E.P. Zaretskaya, V.B. Zalesski, Y. Feofanov, R.W. Martin, Thin Solid Films 451452 (2004) 133. [41] K. Ramanathan, R. Nou, B. To, D.L. Young, R. Bhattacharya, M.A. Contreras, R.G. Dhere, G. Teeter, in: Proceedings of the IEEE 4th World Conference on Photovoltaic Energy Conversion, vol. 1, 2006, p. 380. [42] M. Powalla, B. Dimmler, Thin Solid Films 387 (2001) 251. [43] J. Yang, A. Banerjee, S. Guha, Appl. Phys. Lett. 70 (1997) 2975. [44] D.L. Staebler, C.R. Wronski, Appl. Phys. Lett. 31 (1977) 292. [45] P. Knauth, J. Schoonman, Nanostructured Materials: Selected Synthesis Methods, Properties, and Applications, Springer, Berlin, 2002. [46] A. Jerusalem, R. Radovitzky, in: Proceedings of the 45th AIAA/ASME/ASCE/AHS/ ASC Structures, Structural Dynamics & Materials Conference, 2004, p. 1702. [47] S. Zhang, D. Sun, Y. Fu, H. Du, Surf. Coat. Technol. 167 (2003) 113. [48] K.H. Chung, J. He, D.H. Shin, J.M. Schoenung, Mater. Sci. Eng. A356 (2003) 23. [49] H.J. Fissan, J. Schoonman, Adv. Mater. 8 (2004) 559. [50] K. Ewsuk, Y. Gogotsi (Eds.), Characterization, Design, and Processing of Nanosize Powders and Nanostructured Materials: Ceramic Transactions, vol. 190, Wiley, New York, 2006. [51] S. Siegmann, M. Leparoux, L. Rohr, Proc. SPIE 5824 (2005) 224. [52] J. Gang, J.-P. Morniroli, T. Grosdidier, Scr. Mater. 48 (2003) 1599. [53] A.J. Crosby, J. Young Lee, Polym. Rev. 47 (2007) 217. [54] C. Sanchez, G.J. de A, A. Soler-Illia, F. Ribot, T. Lalot, C.R. Mayer, V. Cabuil, Chem. Mater. 13 (2001) 3061. tzel, Nature 353 (1991) 737. [55] B. ORegan, M. Gra [56] J.-B. Han, N. Wang, G.-P. Yu, Z.-H. Wei, Z.-G. Zhou, Q.-Q. Wang, Sol. Energy Mater. Sol. Cell 88 (2005) 293. tzel, Adv. Mater. 12 (2000) [57] B. ORegan, D.T. Schwartz, S.M. Zakeeruddin, M. Gra 1263. tzel, Appl. Phys. [58] P. Wang, C. Klein, R. Humphry-Baker, S.M. Zakeeruddin, M. Gra Lett. 86 (2005) 123508. tzel, J. [59] P. Wang, S.M. Zakeeruddin, P. Comte, R. Charvet, R. Humphry-Baker, M. Gra Phys. Chem. B 107 (2003) 14336. [60] L. Han, A. Fukui, N. Fuke, N. Koide, R. Yamanaka, in: Proceedings of the IEEE 4th World Conference on Photovoltaic Energy Conversion, vol. 1, 2006, p. 179. [61] J.Y. Kim, K. Lee, N.E. Coates, D. Moses, T.-Q. Nguyen, M. dante, A.J. Heeger, Science 317 (2007) 222. [62] K.M. Coakley, B.S. Srinivasan, J.M. Ziebarth, C. Goh, Y. Liu, M.D. McGehee, Adv. Funct. Mater. 15 (2005) 1927. [63] B. Schwenzer, K.M. Roth, J.R. Gomm, M. Murr, D.E. Morse, J. Mater. Chem. 16 (2006) 401. [64] M. Nanu, J. Schoonman, A. Goossens, Nano Lett. 5 (2005) 1716. [65] R. OHayre, M. Nanu, J. Schoonman, A. Goossens, Nanotechnology 18 (2007) 055702. [66] S. Guha, J. Yang, J. Non-cryst. Solids 352 (2006) 1917. [67] R.A. Stradling, Phys. Scr. T35 (1991) 237. [68] X. Yang, J.B. Heroux, L.F. Mei, W.I. Wang, Appl. Phys. Lett. 78 (2001) 4068. [69] L.S. Yu, S.S. Li, P. Ho, Proc. SPIE 1675 (1992) 255. [70] N.J. Ekins-Daukes, K.W.J. Barnham, J.P. Connolly, J.S. Roberts, J.C. Clark, G. Hill, M. Mazzer, Appl. Phys. Lett. 75 (1999) 4195. [71] G.L. Araujo, A. Mart, Sol. Energy Mater. Sol. Cell 33 (1994) 213240.

shown to be an ideal pn junction, and a photovoltaic effect with an apparent efciency of 5% was shown. Finally, nanoparticles and quantum dots have been shown to be useful in PV devices in various modes, including as plasmonic structures, multi-exciton generating structures, demonstration of intermediate (mini) bands, spectral converters to be applied to conventional solar cells, and as the active layers in novel solar cell designs. It was also shown that there are a number of high-efciency concepts that have been discussed in the literature for achieving the so-called Generation III solar cell technologies, i.e. structures that can yield both high efciency and low cost. These include intermediate bands, carrier multiplication, up/down-conversion, and multi-junctions. While many of these approaches may be implemented in conventional structures or bulk materials, it is apparent that the use of low-dimensional nanostructures, especially quantum wells, nanowires/tubes, and nanoparticles/quantum dots may ultimately play a key role in implementing these novel band structures and conversion mechanisms in future high efciency solar cells. The eld of nano-photovoltaics is relatively new, and while it was not possible to cover all the developments made in recent years within this review, there are many exciting prospects for applying nanostructures and nanotechnology to one of the most promising renewable energy technologies, i.e. photovoltaics and related applications such as photo-electrolysis. Clearly, many technical challenges remain and much fundamental and technological work must be done to properly apply nanostructures to photovoltaics, though the future of nanostructures for photovoltaics is bright. Acknowledgements The author would like to thank D. Merfeld, E. Buttereld, T. Feist, G. Trant, and M.L. Blohm for their support of this work. Thanks also to B.A. Korevaar, O. Sulima, J. Rand, R. Corderman, J. Balch, J. Fronheiser, and R. Rohling for technical collaboration and useful discussions. References
[1] N. Lewis (California Institute of Technology): <http://nsl.caltech.edu/les/ energy.ppt>. [2] M.I. Hoffert, Ken Caldeira, A.K. Jain, Erik F. Haites, L.D. Danny Harvey, S.D. Potter, M.E. Schlesinger, S.H. Schneider, R.G. Watts, T.M.L. Wigley, D.J. Wuebbles, Nature 395 (1998) 881. [3] E. Becquerel, La lumi_ere: ses causes et ses e_ets, tome second, Paris (1867), p. 122. [4] D.M. Chapin, C.S., Fuller, G.S. Pearson, A new silicon pn junction photocell for converting solar radiation into electrical power, J. Appl. Phys. 25 (1954) 676. [5] J. Zhao, A. Wang, M.A. Green, F. Ferrazza, Appl. Phys. Lett. 73 (1998) 1991. [6] W. Shockley, H.J. Queisser, J. Appl. Phys. 32 (1961) 510. [7] M.J. Kerr, P. Campbetl, A. Cuevas, in: Proceedings of the 29th IEEE Photovoltaic Specialists Conference, New Orleans, (2002), p. 438. [8] J. Singh, Electronics and Optoelectronic Properties of Nanostructures, Cambridge University Press, 2003. [9] Y.S. Tsuo, P. Menna, T.H. Wang, T.F. Ciszek, New opportunities in crystalline silicon R&D, NREL Report #CP-590-25612 (1998): <http://www.osti.gov/bridge/ servlets/purl/6593-aQxa5y/native/6593.pdf>. [10] W.G.J.H.M. van Sark, G.W. Brandsen, M. Fleuster, M.P. Hekkert, Energy Policy 35 (2007) 3121. [11] J. Carey, Whats raining on solars parade, Business Week, February 6, 2006. [12] M.A. Green, Third Generation Photovoltaics: Advanced Solar Energy Conversion, Springer-Verlag, Berlin, 2003. [13] W. Hoffman, Sol. Energy Mater. Sol. Cell 90 (2006) 3285. [14] M.B. Prince, J. Appl. Phys. 26 (1955) 534. [15] C.H. Henry, J. Appl. Phys. 51 (1980) 4494. [16] A. Marti, G.L. Araujo, Sol. Energy Mater. Sol. Cell 43 (1996) 203. [17] R.R. King, D.C. Law, K.M. Edmondson, C.M. Fetzer, G.S. Kinsey, H. Yoon, R.A. Sherif, N.H. Karam, Appl. Phys. Lett. 90 (2007) 183516. [18] H. Yoon, J.E. Granata, P. Hebert, R.R. King, C.M. Fetzer, P.C. Colter, K.M. Edmondson, D. Law, G.S. Kinsey, D.D. Krut, J.H. Ermer, M.S. Gillanders, N.H. Karam, Prog. Photovolt: Res. Appl. 13 (2005) 133.

L. Tsakalakos / Materials Science and Engineering R 62 (2008) 175189 [72] J.P. Connolly, J. Nelson, K.W.J. Barnham, I. Ballard, C. Roberts, J.S. Roberts, C.T. Foxon, in: Proceedings of the Conference Records of the 28th IEEE Photovoltaic Specialists, NJ, (2000), p. 1304. ndez, J.P. Connolly, K.W.J. Barnham, Microelectron. J. 38 [73] J.C. Rimada, L. Herna (2007) 513. [74] T.N.D. Tibbits, I. Ballard, I.K.W.J. Barnham, D.C. Johnson, M. Mazzer, J.S. Roberts, R. Airey, N. Foan, in: Proceedings of the IEEE World Conference on Photovoltaic Energy Conversion, vol. 1, 2006, p. 861. [75] D.C. Johnson, I.M. Ballard, K.W.J. Barnham, D.B. Bishnell, J.P. Connolly, M.C. Lynch, T.N.B. Tibbits, N.J. Ekins-Daukes, M. Mazzer, R. Airey, G. Hill, J.S. Roberts, Sol. Energy Mater. Sol. Cell 87 (2005) 169179. [76] D.C. Johnson, I.M. Ballard, K.W.J. Barnham, M. Mazzer, T.N.D. Tibbits, J.S. Roberts, G. Hill, C. Calder, in: Proceedings of the IEEE World Conference Photovoltaic Energy Conversion, vol. 1, 2006, p. 26. [77] R.S. Wagner, W.C. Ellis, Appl. Phys. Lett. 4 (1964) 89. [78] E.I. Givargizov, in: E. Kaldis (Ed.), Current Topics in Material Science, vol. 1, North-Holland, Amsterdam, 1978, p. 79. [79] P.M. Petroff, A.C. Gossard, R.A. Logan, W. Wiegmann, Appl. Phys. Lett. 41 (1982) 635. [80] K. Haraguchi, T. Katsuyama, K. Hiruma, K. Ogawa, Appl. Phys. Lett. 60 (1992) 745. [81] J.R. Heath, F.K. LeGoues, Chem. Phys. Lett. 208 (1993) 263. [82] A.M. Morales, C.M. Lieber, Science 279 (1998) 208. [83] Y. Cui, X. Duan, J. Hu, C.M. Lieber, J. Phys. Chem. B 104 (2000) 5213. [84] S.W. Chung, J.-Y. Yu, J.R. Heath, Appl. Phys. Lett. 76 (2000) 2068. [85] S. Iijima, T. Ichihashi, Nature 363 (1993) 603. [86] M.S. Dresselhaus, Physical Properties of Carbon Nanotubes, Imperial College Press, London, 2003. [87] Y. Ando, in: H.S. Nalwa (Ed.), Encyclopedia of Nanoscience and Nanotechnology, vol. 1, American Scientic Publishers, Valencia, CA, 2004, p. 603. [88] J. Kong, H.T. Soh, A.M. Cassell, C.F. Quate, H.J. Dai, Nature 395 (1998) 878. [89] Y. Gogotsi, J.A. Libera, M. Yoshimura, J. Mater. Res. 15 (2000) 2591. [90] T. Guo, P. Nikolaev, A.G. Rinzler, D. TomBnek, D.T. Colbert:, R.E. Smalley, J. Phys. Chem. 99 (1995) 10694. [91] M.S. Dresselhaus, G. Dresselhaus, Ph. Avouris (Eds.), Carbon Nanotubes: Synthesis, Structure, Properties, and Applications, Springer, 2001. [92] S.J. Tans, R.M. Verschueren, C. Dekker, Nature 393 (1998) 49. [93] E.S. Snow, F.K. Perkins, E.J. Houser, S.C. Badescu, T.L. Reinecke, Science 307 (2005) 1942. [94] P.J. Burke, S. Li, Z. Yu, IEEE Trans. Nanotechnol. 5 (2006) 314. [95] N.V. Quy, N.D. Hoa, M. An, Y. Cho, D. Kim, Nanotechnology 18 (2007) 345201. [96] N.W.S. Kam, M. OConnell, J.A. Wisdom, H. Dai, Proc. Natl. Acad. Sci. 102 (2005) 11600. [97] J. Goldberger, R. He, Y. Zhang, S. Lee, H. Yan, H.-J. Choi, P. Yang, Nature 422 (2003) 599. [98] H.-S. Min, J.-K. Lee, Ferroelectrics 336 (2006) 231. [99] T. Kasuga, M. Hiramatsu, A. Hoson, T. Sekino, K. Niihara, Adv. Mater. 11 (1999) 1307. [100] Y.F. Mei, X.L. Wu, X.F. Shao, G.G. Siu, X.M. Bao, Europhys. Lett. 62 (2003) 595. [101] B. Liu, H.C. Zeng, Langmuir 20 (2004) 4196. [102] T. Hanrath, B.A. Korgel, Adv. Mater. 15 (2003) 437. [103] M.E. Toimil Molares, V. Buschmann, D. Dobrev, R. Neumann, R. Scholz, I.U. Schuchert, J. Vetter, Adv. Mater. 13 (2001) 62. [104] K.Q. Peng, Z.P. Huang, J. Zhu, Adv. Mater. 16 (2004) 73. [105] N.A. Melosh, A. Boukai, F. Diana, B. Gerardot, A. Badolato, P.M. Petroff, J.R. Heath, Science 300 (2003) 112. [106] L. Tsakalakos, M. Rahmane, M. Larsen, Y. Gao, L. Denault, P. Wilson, J. Balch, J. Appl. Phys. 98 (2005) 044317. [107] L. Tsakalakos, J. Fronheiser, L. Rowland, M. Rahmane, M. Larsen, Y. Gao, Mater. Res. Soc. Symp. Proc. 963 (2007), 0963-Q11-03. [108] W.U. Huynh, J.J. Dittmer, A.P. Alivisatos, Science 295 (2002) 2425. [109] D.J. Milliron, I. Gur, A.P. Alivisatos, MRS Bull. 30 (2005) 41. [110] M. Law, L.E. Greene, J.C. Johnson, R. Saykally, P. Yang, Nat. Mater. 4 (2005) 455. [111] J.B. Baxter, E.S. Aydil, Appl. Phys. Lett. 86 (2005) 053114. [112] L. Tsakalakos, J. Balch, J. Fronheiser, M.-Y. Shih, S.F. LeBoeuf, M. Pietrzykowski, P.J. Codella, B.A. Korevaar, O. Sulima, J. Rand, A.D. Kumar, U. Rapol, J. Nanophoton. 01 (2007) 013552. [113] L. Tsakalakos, R.J. Wojnarowski, A. Srivastava, C.S. Korman, J.U. Lee, S.F. LeBoeuf, A. Ebong, O. Sulima, High efciency inorganic nanorod enhanced photovoltaic devices (140342), US 2006/0207647. [114] L. Tsakalakos, J. Balch, J. Fronheiser, B.A. Korevaar, O. Sulima, J. Rand, Appl. Phys. Lett. 91 (2007) 233117. [115] I. Gur, N.A. Fromer, M.L. Geier, A.P. Alivisatos, Science 310 (2005) 462. [116] B.M. Kayes, N.S. Lewis, H.A. Atwater, J. Appl. Phys. 97 (2005) 114302. [117] K. Peng, Y. Xu, Y. Wu, Y. Yan, S.-T. Lee, J. Zhu, Small 1 (2005) 1062.

189

[118] B. Tian, X. Zheng, T.J. Kempa, Y. Fang, N. Yu, G. Yu, J. Huang, C.M. Lieber, Nature 449 (2005) 885. [119] L. Tsakalakos, B.A. Korevaar, J. Balch, J. Fronheiser, previously unpublished work. [120] M.D. Kelzenberg, D.B. Turner-Evans, B.M. Kayes, M.A. Filler, M.C. Putnam, N.S. Lewis, H.A. Atwater, Nano Lett. 8 (2008) 710. [121] J.U. Lee, P.P. Gipp, C.M. Heller, Appl. Phys. Lett. 85 (2004) 145. [122] J.U. Lee, Appl. Phys. Lett. 87 (2005) 073101. [123] J.U. Lee, Phys. Rev. B 75 (2007) 075409. [124] B.J. Landi, R.P. Raffaelle, S.L. Castro, S.G. Bailey, Progr. Photovolt.: Res. Appl. 13 (2005) 165. [125] E. Kymakis, G.A.J. Amaratunga, Appl. Phys. Lett. 80 (2002) 112. [126] S. Campidelli, M. Meneghetti, M. Prato, Small 3 (2007) 1672. [127] A. Du Pasquier, H.E. Unalan, A. Kanwal, S. Miller, M. Chhowalla, Appl. Phys. Lett. 87 (2005) 203511. [128] C.B. Murray, D.J. Norris, M.G. Bawendi, J. Am. Chem. Soc. 115 (1993) 8706. [129] V.L. Colvin, A.P. Alivisatos, J.G. Tobin, Phys. Rev. Lett. 66 (1991) 2786. [130] A.P. Alivisatos, W. Gu, C. Larabell, Ann. Rev. Biomed. Eng. 7 (2005) 55. [131] Y. Cui, M.T. Bjork, J.A. Liddle, C. Sonnichsen, B. Boussert, A.P. Alivisatos, Nano Lett. 4 (2004) 1093. [132] T. Trupke, M.A. Green, P. Wurfel, J. Appl. Phys. 92 (2002) 1668. [133] B.-C. Hong, K. Kawano, Sol. Energy Mater. Sol. Cell 80 (2003) 417. [134] V. Svrcek, A. Slaoui, J.-C. Muller, Thin Solid Films 451452 (2004) 384. [135] B. Gonzalez-Daz, R. Guerrero-Lemus, P. Haro-Gonzalez, D. Borchert, C. Hernandez-Rodrguez, Thin Solid Films 511512 (2006) 473. [136] C. Strumpel, M. McCann, G. Beaucarne, V. Arkhipov, A. Slaoui, V. Svrcek, C. del Canizo, I. Tobias, Sol. Energy Mater. Sol. Cell 91 (2007) 238. [137] W. van Sark, C. De Mello Donega, C. Harkisoen, R. Kinderman, J. van Roosmalen, R. Schropp, E. Lysen, in: Proceedings of the 19th EPVSEC, Paris, 2004. [138] S.J. Gallagher, B. Norton, P.C. Eames, Solar Energy 81 (2007) 813. [139] T. Trupke, M.A. Green, P. Wurfel, J. Appl. Phys. 92 (2002) 4117. [140] A. Shalav, B.S. Richards, M.A. Green, Sol. Energy Mater. Sol. Cells 91 (2007) 829. [141] F. Wang, D.K. Chatterjee, Z. Li, Y. Zhang, X. Fan, M. Wang, Nanotechnology 17 (2006) 5786. [142] G. Yi, H. Lu, S. Zhao, Y. Ge, W. Yang, D. Chen, L.-H. Guo, Nano Lett. 4 (2004) 2191. guez, Appl. Phys. Lett. 87 [143] E. De la Rosa, P. Salas, H. Desirena, C. Angeles, R.A. Rodr (2005) 241912. [144] A. Mart, E. Antoln, C.R. Stanley, C.D. Farmer, N. Lopez, P. Daz, E. Canovas, P.G. Linares, A. Luque, Phys. Rev. Lett. 97 (2006) 247701. [145] M. Wolf, R. Brendel, J.H. Werner, H.J. Queisser, J. Appl. Phys. 83 (1998) 4213. [146] R.J. Ellingson, M.C. Beard, J.C. Johnson, P. Yu, O.I. Micic, A.J. Nozik, A. Shabaev, A.L. Efros, Nano Lett. 5 (2005) 865. [147] R.D. Schaller, M. Sykora, J.M. Pietryga, V.I. Klimov, Nano Lett. 6 (2006) 242. [148] G. Nair, M.G. Bawendi, Phys. Rev. B 76 (2007) 081304(R). [149] M. Califano, A. Zunger, A. Franceschetti, Appl. Phys. Lett. 84 (2004) 2409. [150] R.D. Schaller, V.M. Agranovich, V.I. Klimov, Nat. Phys. 1 (2005) 189. [151] R.D. Schaller, I.V. Klimov, Phys. Rev. Lett. 92 (2004) 186601. [152] S.M. Hubbard, R. Raffaelle, R. Robinson, C. Bailey, D. Wilt, D. Wolford, W. Maurer, S. Bailey, Mater. Res. Soc. Symp. Proc. 1017 (2007), 1017-DD13-11. [153] R.P. Raffaelle, S.L. Castro, A.F. Hepp, S.G. Bailey, Prog. Photovolt.: Res. Appl. 10 (2002) 433. [154] G. Conibeer, M. Green, R. Corkish, Y. Cho, E.-C. Cho, C.-W. Jiang, T. Fangsuwannarak, E. Pink, Y. Huang, T. Puzzer, T. Trupke, B. Richards, A. Shalav, K.-I. Lin, Thin Solid Films 511512 (2006) 654. [155] J.E. Granata, J.H. Ermer, P. Hebert, M. Haddad, R.R. King, D.D. Krut, M.S. Gillanders, N.H. Karam, B.T. Cavicchi, in: Proceedings of the 29th IEEE Photovoltaic Specialist Conference, IEEE, New York, (2002), p. 824. [156] E. Ozbay, Science 311 (2006) 189. [157] S.A. Maier, H.A. Atwater, Plasmonics:, J. Appl. Phys. 98 (2005) 011101. [158] C. Kittel, Introduction to Solid State Physics, 8th edition, Wiley, New York, 2004. [159] B.-H. Choi, H.-H. Lee, S. Jin, S. Chun, S.-H. Kim, Nanotechnology 18 (2007) 075706. [160] S. Berciaud, L. Cognet, P. Tamarat, B. Lounis, Nano Lett. 5 (2005) 515. [161] J.R. Cole, N.J. Halas, Appl. Phys. Lett. 89 (2006) 153120. [162] D.M. Schaadt, B. Feng, E.T. Yu, Appl. Phys. Lett. 86 (2005) 063106. [163] D. Derkacs, S.H. Lim, P. Matheu, W. Mar, E.T. Yu, Appl. Phys. Lett. 89 (2006) 093103. [164] S. Pillai, K.R. Catchpole, T. Trupke, M.A. Green, J. Appl. Phys. 101 (2007) 093105. [165] D.V. Talapin, C.B. Murray, Science 310 (2005) 86. [166] D. Yu, B.L. Wehrenberg, P. Jha, J. Ma, P. Guyot-Sionnest, J. Appl. Phys. 99 (2006) 104315. [167] T.A. Klar, T. Franzl, A.L. Rogach, J. Feldmann, Adv. Mater. 17 (2005) 769. [168] I. Mora-Sero, J. Bisquert, F. Fabregat-Santiago, G. Garcia-Belmonte, G. Zoppi, K. Durose, Y. Proskuryakov, I. Oja, A. Belaidi, T. Dittrich, R. Tena-Zaera, A. Katty, C. Levy-Clement, V. Barrioz, S.J.C. Irvine, Nano Lett. 6 (2006) 640.

Anda mungkin juga menyukai