Photoinduced Silver Precursor Decomposition for Particle Modification

Apr 8, 2014 - Several European chemical firms are expanding their manufacturing footprints in the heart of the region.... POLICY CONCENTRATES ...
1 downloads 0 Views 2MB Size
Article pubs.acs.org/JPCC

Photoinduced Silver Precursor Decomposition for Particle Modification in Tungsten Oxide−Polymer Matrix Nanocomposites Travis J. DeJournett and James B. Spicer* Department of Materials Science and Engineering, The Johns Hopkins University, 3400 North Charles Street, Baltimore, Maryland 21218, United States ABSTRACT: A processing technique based on optical excitation of a polymer matrix nanocomposite containing a chemical precursor is presented to demonstrate the potential of this process for localized modification of nanoparticles in the bulk of the material. Femtosecond laser irradiation of a nanocomposite consisting of tungsten oxide nanoparticles in a fluoropolymer (tetrafluoroethylene-co-hexafluoropropylene) matrix resulted in decomposition of a silver precursor (vinyltriethylsilane(hexafluoroacetylacetonate)silver(I)). Transmission electron microscopy (TEM) results showed an increase in particle size as a result of photoprocessing with particles displaying core−shell type architectures. Optical absorption spectroscopy measurements indicated that silver was deposited in the bulk of the material and were consistent with absorption cross-section models for tungsten oxide core/silver shell nanoparticles. Estimates for silver shell volumes were found to correlate with the surface areas of the base tungsten oxide particles. Modeled results for photothermal temperature rises suggest that thermal processes are probably not responsible for precursor decomposition and the most likely mechanism is multiphoton, photocatalytic precursor decomposition at the surface of the tungsten oxide nanoparticles because tungsten oxide is a known photocatalyst. Models for precursor diffusion in the matrix after particle excitation are used to identify relevant time scales associated with precursor depletion and to provide additional insight into this in situ deposition process for particle modification.

1. INTRODUCTION Polymer matrix nanocomposites (PMNCs) can provide a range of interesting and potentially useful behaviors that depend largely on the nature of a constituent material having nanoscale dimensions. In many cases, the desired functionalities of PMNCs are dictated primarily by nanoparticles or other nanoscale structures that are integrated into the matrix. In these cases, the matrix primarily serves as a supporting medium, but it must also provide appropriate access to these structures to maintain the desired optical/chemical/electronic behaviors. For example, nanostructured systems designed for artificial photosynthesis,1,2 environmental photocatalysis,3−6 or photochromic7,8 applications have all been investigated with various implementations using polymer matrixes as the underlying support for various functional components. Challenges exist not only for the design and synthesis of these components but also for the processing required to incorporate them into matrixes in ways that maintain the desired functionalities. This can be accomplished using a variety of methods. A common approach is to disperse nanostructured components into uncured polymer and to fix these components in place by curing the polymer.9 This method can be accomplished easily using simple processing tools. Another approach is to create the nanostructured components on surfaces and to encapsulate them using a polymer overlayer.10 This maintains the distribution and orientation of components in the resulting composite film and allows for a high degree of control of component location.11 © 2014 American Chemical Society

The general processing approach reported here is to synthesize tungsten oxide nanoparticles directly in a polymer matrix12,13 and to modify these using a photoinitiated, chemical vapor deposition process. The photochromic response of tungsten oxide to ultraviolet light has been studied extensively, and various efforts have sought to extend this response into the visible region by incorporating additional materials into these systems. Even though details regarding the photochromic response of tungsten oxide are not completely settled, approaches to shifting this response to longer wavelengths include incorporating noble metals, semiconductors, or other metal oxides into the overall material structure.14,15 Methods for doing so vary depending on whether the base tungsten oxide is in the form of a thin film, is composed of nanoparticles (or other nanoscale structures), or is a component in a more complicated system. The base tungsten oxide PMNC used here displays a variety of photochromic behaviors depending on the processing used and, like other tungsten oxide-based systems, methods are needed to extend its spectral response to the visible region.16 To achieve this type of change in spectral response, in situ modification of nanoparticles in polymer matrixes using photonically induced chemical reactions has been pursued. Specifically, the nanocomposite is excited using femtosecond Received: February 12, 2014 Revised: April 5, 2014 Published: April 8, 2014 9820

dx.doi.org/10.1021/jp5015324 | J. Phys. Chem. C 2014, 118, 9820−9831

The Journal of Physical Chemistry C

Article

precursor decomposition and permit nanoparticle formation. These processing steps were repeated a second time to increase particle size. The resulting material consisted primarily of the FEP polymer matrix loaded with discrete tungsten oxide nanoparticles distributed throughout the bulk of the material as indicated by the transmission electron micrograph (TEM) of an area of the film cross section shown in Figure 1a where particles of tungsten oxide appear to be randomly distributed in the matrix. Samples were prepared for TEM imaging using room-temperature diamond microtome methods, yielding sections approximately 100 nm thick, and were mounted on copper or nickel TEM grids. Imaging was performed on a 120 kV Philips EM 420 transmission electron microscope, and ImageJ software was used to analyze the micrographs for nanoparticle polydispersity. Higher magnification micrographs of particles in the film cross section are shown in Figure 1b,c. In Figure 1b, larger particles appear to be composed of collections of smaller particles with varying electron absorption cross sections. Figure 1c shows details of the larger particles more clearly. Variation in the opacity of particle structures is likely related to localized variations in the oxidation state of tungsten oxide because this affects the permittivity of the material. Figure 1c also reveals the presence of a large number density of smaller particles with radii in the 1−2 nm range. These particles do not appear at lower magnifications because they are relatively electron transparent and are also obscured by matrix structures. This directly affects estimates of the particle size distribution. The low volume fraction of tungsten oxide in this material prevents ready examination using X-ray diffraction, and two additional processing cycles were performed to increase particle size and volume fraction. A transmission electron micrograph of an interior cross section of this material is shown in Figure 2a. Again, the larger particles are highly structured and appear to be composed of smaller particles. Also, images of isolated, smaller particles are present in the micrograph but are partially obscured by matrix structure. X-ray diffraction measurements on this material were performed on a Philips X’Pert Pro X-ray diffractometer and are shown offset in Figure 2b along with corresponding measurements on the neat matrix. The broad peak centered near 25° is related to tungsten oxide and is consistent with previously reported measurements of noncrystalline tungsten oxide.20 The peak at 31−32° is related to ordering of CF3 side groups along the polymer chains of the matrix. 2.2. Materials, Optical Processing, and Characterization. Using the tungsten oxide nanocomposite produced using two cycles, subsequent processing was pursued to investigate the possibility of modifying particle characteristics by inducing chemical reactions in the near-particle region using photonic excitation. For this work, decomposition of a silver precursor (vinyltriethylsilane(hexafluoroacetylacetonate)silver(I) (99.9%Ag; STREM Chemicals Inc., Newbury, MA) was chosen because it would yield species distinct from the existing material. Also, incorporation of silver into this tungsten oxide system could alter its photochromic response to visible light; this type of effect has been reported in related systems.21 A variety of laser sources could be used to excite this system; one of the selection criteria used here was to maximize direct interactions with the tungsten oxide particles and minimize those with the matrix and precursor. Optical transmission spectra shown in Figure 3a for the matrix and for the nanocomposite were taken using a Cary 50 3.00 UV/vis

laser pulses to decompose a silver precursor in the near-particle environment to alter local chemical/electronic/optical properties. Both the chemical and optical accessibility of particles in this PMNC system permit this type of processing−diffusion of chemical precursors into regions surrounding tungsten oxide nanoparticles as well as the subsequent optical initiation of chemical reactions leading to silver deposition can be readily carried out. This modifies particle structure and properties in ways that affect overall material functionality. The work described here demonstrates this processing scheme, reports various characteristics of materials produced using this approach, and identifies physical/chemical processes that play important roles in the optical processing of tungsten oxide PMNCs. In particular, the roles of precursor photolysis, photothermal excitation, and photocatalytic decomposition are considered. As part of this analysis, photothermal temperature rises resulting from optical excitation of the nanoparticles are modeled and are used to assess the likelihood of thermally induced precursor decomposition. In addition, precursor diffusion in the matrix after particle excitation is considered to identify relevant time scales associated with precursor depletion and to provide insight into design of optimized processes for particle modification. The resulting nanocomposites are examined using transmission electron microscopy and optical absorption spectroscopy to provide additional insight into the characteristics of this processing approach.

2. EXPERIMENTAL METHODS Nanoparticles in the polymer matrix were synthesized in a twostep process. The first step is a type of thermally activated, in situ chemical vapor deposition process that yields tungsten oxide nanoparticles distributed throughout the bulk of the matrix. The second uses photonic excitation to induce decomposition of a silver precursor in the near-particle environment. These processing steps were carried out sequentially with the aim of producing particles that have tungsten oxide core/silver shell architectures. 2.1. Materials, Thermal Processing, and Characterization. The tungsten oxide nanocomposite was produced using a tungsten hexacarbonyl organometallic (W(CO)6; 99% W, 10RP and D ≈ 10−8 μm2/ns.54 Calculations using the full expression for nearparticle concentrations yield results that follow those obtained using eq 19 for times up to approximately 1 ms. Representative results using RP = 10 nm and ΔR = 1 nm are shown in Figure 10 for various distances from the particle center.

Figure 10. Fractional change in precursor concentration as a function of time for various radial distances outside a nanoparticle with a radius of 10 nm.

These results indicate that equilibrium concentrations in the particle environment are not achieved within 100 μs after excitation. In particular, at the particle surface, precursor concentrations exceed 0.999c0 at 100 μs, but even this small deviation from equilibrium has significant implications. If photoexcitation occurs at intervals less than 100 μs, then precursor does not fully populate the near-particle environment, resulting in altered locations and rates of decomposition even after short times corresponding to a small number of events. Indeed, if the interval between excitation events is 1 μs or less, then precursor decomposes as it reaches the particle surface (or near-particle zone); the excitation appears to be continuous from a diffusion viewpoint, and the distribution of various species in the reaction region will vary accordingly. This

5. CONCLUSIONS This work has reported the modification of nanoparticles in a tungsten oxide−polymer matrix nanocomposite containing a 9829

dx.doi.org/10.1021/jp5015324 | J. Phys. Chem. C 2014, 118, 9820−9831

The Journal of Physical Chemistry C

Article

(10) Eilers, H.; Biswas, A.; Pounds, T. D.; Grant Norton, M.; Elbahri, M. Teflon AF/Ag nanocomposites with tailored optical Properties. J. Mater. Res. 2006, 21, 2168−2171. (11) Rosenberg, S. G.; Barclay, M.; Fairbrother, D. H. Electron induced reactions of surface adsorbed tungsten hexacarbonyl W(CO)6. Phys. Chem. Chem. Phys. 2013, 15, 4002−4015. (12) Koloski, T. S.; Vargo, T. G. United States Patent 5,977,241, 1999. (13) Koloski, T. S.; Vargo, T. G. United States Patent 6,608,129, 2003. (14) Bechinger, C.; Wirth, E.; Leiderer, P. Photochromic coloration of WO3 with visible light. Appl. Phys. Lett. 1996, 68, 2834−2836. (15) He, T.; Yao, J. Photochromic materials based on tungsten oxide. J. Mater. Chem. 2007, 17, 4547−4557. (16) DeJournett, T. J.; Spicer, J. B. The influence of oxygen on the microstructural, optical and photochromic properties of polymermatrix, tungsten-oxide nanocomposite films. Sol. Energy Mater. Sol. Cells 2014, 120, 102−108. (17) Bechinger, C.; Burdis, M. S.; Zhang, J.-G. Comparison between electrochromic and photochromic coloration efficiency of tungsten oxide thin films. Solid State Commun. 1997, 10, 753−756. (18) Tamou, Y.; Tanaka, S.-I. Formation and coalescence of tungsten nanoparticles under electron beam irradiation. Nanostruct. Mater. 1999, 12, 123−126. (19) Dimitrova, Z.; Gogova, D. On the structure, stress and optical properties of CVD tungsten oxide films. Mater. Res. Bull. 2005, 40, 333−340. (20) Wang, H.; Xu, P.; Wang, T. The preparation and properties study of photocatalytic nanocrystalline/nanoporous WO3 thin films. Mater. Des. 2002, 23, 331−336. (21) Yao, J. N.; Yang, Y. A.; Loo, B. H. Enhancement of Photochromism and Electrochromism in the Au/MoO3/SnO2 and Pt/MoO3/SnO2 Thin Films. J. Phys. Chem. B 1998, 102, 1856−1860. (22) von Rottkay, K.; Rubin, M.; Wen, S. J. Optical indices of electrochromic tungsten oxide. Thin Solid Films 1997, 306, 10−16. (23) Bienert, R.; Emmerling, F.; Thünemann, A. F. The size distribution of gold standard nanoparticles. Anal. Bioanal. Chem. 2009, 395, 1651−1660. (24) See, K. C.; Spicer, J. B.; Brupbacher, J.; Zhang, D.; Vargo, T. G. Modeling interband transitions in silver nanoparticle−fluoropolymer composites. J. Phys. Chem. B 2005, 109, 2693−2698. (25) Bagratashvili, V. N.; Rybaltovsky, A. O.; Minaev, N. V.; Firsov, V. V.; Yusupov, V. I. Laser-induced atomic assembling of periodic layered nanostructures of silver nanoparticles in fluoro-polymer film matrix. Laser Phys. Lett. 2010, 7, 401−404. (26) Chahadih, A.; El Hamzaoui, H.; Bernard, R.; Boussekey, L.; Bois, L.; Cristini, O.; Le Parquier, M.; Capoen, B.; Bouazaoui, M. Direct-writing of PbS nanoparticles inside transparent porous silica monoliths using pulsed femtosecond laser irradiation. Nanoscale Res. Lett. 2011, 6, 1−7, DOI: 10.1186/1556-276X-6-542. (27) Wang, J.; Lau, W. M.; Li, Q. Effects of particle size and spacing on the optical properties of gold nanocrystals in alumina. J. Appl. Phys. [Online] 2005, 97, Article 114303. http://dx.doi.org/10.1063/1. 1868052. (28) Raza, S.; Yan, W.; Stenger, N.; Wubs, M.; Mortensen, N. A. Blueshift of the surface plasmon resonance in silver nanoparticles: substrate effects. Opt. Express 2013, 21, 27344−27355. (29) Averitt, R. D.; Sarkar, D.; Halas, N. J. Plasmon resonance shifts of Au-coated Au2S nanoshells: insight into multicomponent nanoparticle growth. Phys. Rev. Lett. 1997, 78, 4217−4220. (30) Eggeman, A. S.; Dobson, P. J.; Petford-Long, A. K. Optical spectroscopy and energy-filtered transmission electron microscopy of surface plasmons in core−shell nanoparticles. J. Appl. Phys. 2007, 101, 024307. (31) DeJournett, T. J.; Spicer, J. B. Laser-induced, in situ, nanoparticle shell synthesis in polymer matrix nanocomposites. Phys. Chem. Chem. Phys. 2013, 15, 19753−19762. (32) Persson, B. N. J. Polarizability of small spherical metal particles: Influence of the matrix environment. Surf. Sci. 1993, 281, 153−162.

silver precursor using femtosecond laser excitation. Optical excitation of the nanocomposite resulted in decomposition of precursor and subsequent deposition of silver in the nearparticle environment. Transmission electron micrographs show increased particle sizes associated with optically processed material. Expected temperature rises in near-particle regions indicate that photothermal processes are likely not responsible for precursor decomposition and that some other mechanism mediates the reaction, possibly a multiphoton, photocatalytic process. Models for precursor diffusion in the matrix indicate that decomposition occurs continuously as these species move into near-particle regions and that precursor depletion for particles in the nanocomposite interior can occur for long exposure times. On the basis of interpretation of micrographs as well as optical absorption spectra, silver deposition results in particles having optical spectra consistent with WO3−Ag core− shell structures. This overall processing approach demonstrates the potential for localized, in situ modification and growth of nanostructures in polymer matrixes and points to more sophisticated processing approaches that can be used to create multifunctional materials containing hierarchical structures over various length scales.



AUTHOR INFORMATION

Corresponding Author

*Telephone: 410 516 8524. E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors thank Dr. Michael McCaffery and the Johns Hopkins Integrated Imaging Center for the invaluable contributions made to this work through TEM sample preparation and imaging.



REFERENCES

(1) Ijeri, V.; Cappelletto, L.; Bianco, S.; Tortello, M.; Spinelli, P.; Tresso, E. Nafion and Carbon Nanotube Nanocomposites for Mixed Proton and Electron conduction. J. Membr. Sci. 2010, 363, 265−270. (2) Tortello, M.; Bianco, S.; Ijeri, V.; Spinelli, P.; Tresso, E. Nafion membranes with vertically-aligned CNTs for mixed proton and electron conduction. J. Membr. Sci. 2012, 415−16, 346−352. (3) Pandikumar, A.; Manonmani, S.; Ramaraj, R. TiO2−Au Nanocomposite Materials Embedded in Polymer Matrices and Their Application in the Photocatalytic Reduction of Nitrite to Ammonia. Catal. Sci. Technol. 2012, 2, 345−353. (4) Thomas, R. T.; Sandhyarani, N. Enhancement in the photocatalytic degradation of low density polyethylene−TiO2 nanocomposite films under solar irradiation. RSC Adv. 2013, 3, 14080. (5) Cai, Y.; Strømme, M.; Welch, K. Photocatalytic Antibacterial Effects Are Maintained on Resin-Based TiO2 Nanocomposites after Cessation of UV Irradiation. PLoS One 2013, 8, e75929−e75929. (6) Haldorai, Y.; Sivakumar, K.; Shim, J.-J. ZnO nanoparticles dispersed poly(aniline-co-o-anthranilic acid) composites: Photocatalytic reduction of Cr(VI) and Ni(II). Polym. Compos. 2013, DOI: 10.1002/pc.22727. (7) Li, S.; El-Shall, M. S. Synthesis and characterization of photochromic molybdenum and tungsten oxide nanoparticles. Nanostructured Materials 1999, 12, 215−219. (8) Sun, M.; Xu, N.; Cao, Y. W.; Yao, J. N.; Wang, E. G. Nanocrystalline tungsten oxide thin film: Preparation, microstructure, and photochromic behavior. J. Mater. Res. 2000, 15, 927−933. (9) Gunawidjaja, R.; Myint, T.; Eilers, H. Synthesis of Silver/Sio2/ Eu:Lu2o3 Core-Shell Nanoparticles and Their Polymer Nanocomposites. Powder Technol. 2011, 210, 157−166. 9830

dx.doi.org/10.1021/jp5015324 | J. Phys. Chem. C 2014, 118, 9820−9831

The Journal of Physical Chemistry C

Article

(33) Vora, K.; Kang, S.-Y.; Shukla, S.; Mazur, E. Fabrication of disconnected three-dimensional silver nanostructures in a polymer matrix. Appl. Phys. Lett. [Online] 2012, 100, Article 063120. http://dx. doi.org/10.1063/1.3684277. (34) Yagci, Y.; Sangermano, M.; Rizza, G. A visible light photochemical route to silver−epoxy nanocomposites by simultaneous polymerization−reduction approach. Polymer 2008, 49, 5195−5198. (35) DeJournett, T. J. The Synthesis, Characterization and Modification of Polymer Matrix Nanocomposites. Ph.D. Dissertation, The Johns Hopkins University, 2013. (36) Chen, G. Nanoscale energy transport and conversion: A parallel treatment of electrons, molecules, phonons, and photons; Oxford University Press: New York, 2005. (37) Hu, M.; Hartland, G. V. Heat Dissipation for Au Particles in Aqueous Solution: Relaxation Time versus Size. J. Phys. Chem. B 2002, 106, 7029−7033. (38) Baffou, G.; Rigneault, H. Femtosecond-pulsed optical heating of gold nanoparticles. Phys. Rev. B 2011, 84, 035415. (39) Chou, C.-H.; Chen, C.-D.; Wang, C. R. C. Highly Efficient, Wavelength-Tunable, Gold Nanoparticle Based Optothermal Nanoconvertors. J. Phys. Chem. B 2005, 109, 11135−11138. (40) Ekici, O.; Harrison, R. K.; Durr, N. J.; Eversole, D. S.; Lee, M.; Ben-Yakar, A. Thermal Analysis of Gold Nanorods Heated with Femtosecond Laser Pulses. J. Phys. D: Appl. Phys. 2008, 41, 185501(11pp). (41) Govorov, A. O.; Zhang, W.; Skeini, T.; Richardson, H.; Lee, J.; Kotov, N. A. Gold nanoparticle ensembles as heaters and actuators: Melting and collective plasmon resonances. Nanoscale Res. Lett. 2006, 1, 84−90. (42) Govorov, A. O.; Richardson, H. H. Generating Heat with Metal Nanoparticles. Nanotoday 2007, 2, 30−38. (43) Carslaw, H. S.; Jaeger, J. C. Conduction of Heat in Solids, 2nd ed.; Oxford University Press: New York, 1986. (44) Goldenberg, H. A problem in radial heat flow. Br. J. Appl. Phys. 1951, 2, 233−237. (45) Goldenberg, H.; Tranter, C. J. Heat flow in an infinite medium heated by a sphere. Br. J. Appl. Phys. 1952, 3, 296−298. (46) Winsemius, P.; van Kampen, F. F.; Lengkeek, H. P.; van Went, C. G. Temperature dependence of the optical properties of Au, Ag and Cu. J. Phys. F: Metal Phys. 1976, 6, 1583−1606. (47) Rashidi Huyeh, M.; Shirdel Havar, M.; Palpant, B. Thermooptical properties of embedded silver nanoparticles. J. Appl. Phys. 2012, 112, 103101. (48) Gao, X.; Su, X.; Yang, C.; Xiao, F.; Wang, J.; Cao, X.; Wang, S.; Zhang, L. Hydrothermal synthesis of WO3 nanoplates as highly sensitive cyclohexene sensor and high-efficiency MB photocatalyst. Sens. Actuators, B 2013, 181, 537−543. (49) Janáky, C.; Rajeshwar, K.; de Tacconi, N. R.; Chanmanee, W.; Huda, M. N. Tungsten-based oxide semiconductors for solar hydrogen generation. Catal. Today 2013, 199, 53−64. (50) Di Valentin, C.; Wang, F.; Pacchioni, G. Tungsten Oxide in Catalysis and Photocatalysis: Hints from DFT. Top. Catal. 2013, 56, 1404−1419. (51) Lines, M. E. Influence of d orbitals on the nonlinear optical response of transparent transition-metal oxides. Phys. Rev. B 1991, 43, 11 978−11 990. (52) Yumashev, K. V.; Malyarevich, A. M.; Posnov, N. N.; Denisov, I. A.; Artem’ev, M. V.; Sviridov, D. V.; Mikhaĭlov, V. P. Nonlinear spectroscopy of photocoloured polytungstic acid nanocomposites. Quantum Electron. 1998, 28, 710−714. (53) Yumashev, K. V.; Malyarevich, A. M.; Posnov, N. N.; Denisov, I. A.; Mikhaĭlov, V. P.; Artem’ev, M. V.; Sviridov, D. V. Optical transient bleaching of photochromic polytungstic acid. Chem. Phys. Lett. 1998, 288, 567−575. (54) George, S. C.; Thomas, S. Transport Phenomena through Polymeric Systems. Prog. Polym. Sci. 2001, 26, 985−1017.

9831

dx.doi.org/10.1021/jp5015324 | J. Phys. Chem. C 2014, 118, 9820−9831