Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Refractive index susceptibility of palladium nanoplates with plasmonic resonance in the visible region

Open Access Open Access

Abstract

We synthesized anisotropic Pd nanoplates for use as a novel refractive index (RI) sensing material in the visible region. The nanoplates showed an extinction peak that was attributed to the excitation of the localized surface plasmon resonance at ~620 nm in the visible region. It was found that the peak was red-shifted with increasing the RI of the surrounding medium. The susceptibility was calculated to be 250 nm per RI unit, comparable to some anisotropic Au nanoparticles that are excellent RI sensing materials.

© 2016 Optical Society of America

1. Introduction

The localized surface plasmon phenomenon which is unique to metal nanoparticles generates a definite optical extinction peak at a specific wavelength due to the resonance with the incident light field [1–8]. This is called the localized surface plasmon resonance: LSPR. While it is well known that this phenomenon leads to the generation of strong local electromagnetic fields as compared with the incident light fields which are applicable to surface-enhanced Raman scattering (SERS), etc [9–12], the position of the peak wavelength sensitively responds to the change in the refractive index (RI) at the nanospace around the nanoparticles [13–15]. This phenomenon holds promise for applications in label-free and cost-effective interfacial biosensing, including the cancer biomarkers [16–18], DNA detection [19,20], immunoassay [21], etc.. The sensing detection limit mainly depends on the magnitude of the peak shift against the RI change, as well as the resonance linewidth as indicated by the figure of merit (FoM), which is defined as the ratio of the RI susceptibility to the linewidth [15]. Although metasurfaces consisting of sophisticated subwavelength periodic structures showed an excellent RI susceptibility as reported recently [22], special and expensive apparatuses are needed to fabricate these structures. Therefore, improving the intrinsic RI susceptibility of plasmonic nanoparticles is an important technical objective for the development of highly sensitive label-free biosensing materials. In fact, various plasmonic anisotropic nanoparticles displaying excellent RI susceptibilities have been developed, including Au nanobranches [23], Au nanorods [23–25], Au nanobipyramids [23], Au nanostars [26], Au pyramids [27], Au nanorattles [28], Au nanoprisms [29], Ag nanoprisms [30–32], Ag nanocubes [33], Au nanoframes [34], and Au nanorings [35]. However, all of these nanoparticles consist of the well-known plasmonic elements Au and Ag. The preoccupation with these elements may have been an obstacle to finding novel highly sensitive plasmonic elements.

Recently, we have found that the RI susceptibility of Au-core/Pd-shell type nanospheres (Au/PdNSs), which generate the peak of Pd LSPR, is substantially higher than those of Au and Ag nanospheres with similar diameters [36]. We also theoretically verified, within a quasi-static (QS) approximation framework, that the higher susceptibility of the Pd LSPR originates from the smaller dispersion of the real part of its dielectric function around the resonant wavelength, compared to those of Au and Ag LSPR. However, since the LSPR peak of the Au/PdNSs was generated at a near-ultraviolet region, bioanalytes could be damaged by irradiation of the near-ultraviolet light in the application biosensing [37–39]. Herein, we report the large RI susceptibility of Pd nanoplates which generate the LSPR peak at a visible region.

2. Experimental section

Synthesis of Pd nanoplates

The Pd nanoplates were synthesized according to a previously reported procedure [40]. Firstly, 12.5 mg of Pd(II) acetylacetonate, 40 mg of poly(vinylpyrrolidone) (PVP, MW = 29,000) and 13 mg of NaBr were dissolved in a mixed solvent of 2.5 mL of N,N-dimethylpropionamide and 0.5 mL of water. The transparent yellow solution was transferred to a 10-mL glass pressure vessel. After the vessel was then charged with CO (pressure: 1 bar), the solution was vigorously shaken. After repeating the manipulation (charging and shaking) thrice, the CO-charged solution was heated at 100 °C for 3.0 h. The dark blue solution in which the nanoplates were dispersed was cooled at room temperature. After 0.1 mL of acetone was added, the solution was centrifuged at 10,000 rpm for 10 min to remove the unreacted precursors. After the obtained precipitates were dispersed in anhydrous ethanol (0.7 mL) to which 0.3 mL acetone was added, the solution was centrifuged at 10,000 rpm for 10 min. Finally, the resultant precipitates was dispersed in a mixed solvent of ethanol/glycerol described in the next section.

Investigation of refractive index susceptibility

The precipitate obtained by centrifugation was dispersed in ethanol solutions containing 0, 10, 20, 30, 40, and 50 vol% glycerol. The LSPR peak position was determined by measuring the extinction spectrum of these colloidal Pd nanoplate solutions.

Theoretical calculation of the extinction spectra of the Pd nanoplate

The extinction spectra of an isolated Pd nanoplate were calculated theoretically by the boundary element method within the QS approximation (QS-BEM) framework with the MNPBEM14 program [41] on MATLAB® R2013a. One Pd nanoplate was bare, and the other was coated with the 0.9 nm-thick PVP. The dielectric functions were taken from the literature for Pd [42] and PVP [43]. Note that the dielectric function of PVP is wavelength dependent and the refractive index of PVP around the LSPR wavelength is approximately 1.52. The in-plane dipole resonance mode in the plate was excited for the extinction spectra. The geometry of the nanoplate with circular shape was approximated by a regular polygon with 40 sides. The meshgrids were in which the thickness was divided into 12 divisions. The geometry of the Pd nanoplate is shown in Fig. 1(e).

 figure: Fig. 1

Fig. 1 Pd nanoplates synthesized in this study. (a)-(c) BF-STEM and (d) HAADF-STEM images and (e) Schematic illustration and model structure approximated by a regular polygon with 40 sides. E and k are the electric field vector and the wave vector of the inciendent light.

Download Full Size | PDF

Measurements

UV-vis spectral measurements were carried out using a JASCO V-630 spectrophotometer. Bright-field (BF) and high-angle annular dark-field (HAADF) scanning transmission electron microscope (STEM) images were taken by using STEM (HD-2300C, Hitachi Ltd.) at 200 keV. The STEM sample was prepared by dipping the carbon grid into the colloidal solution of Pd nanoplates in ethanol.

3. Results and discussion

Since the typical Pd nanospheres show a dipole LSPR peak in the middle-to-near ultraviolet regions, we synthesized Pd nanoplates with an anisotropic shape, which show a definite extinction peak in the visible region. To prepare the Pd nanoplates, we employed the CO-confinement growth method. It has been reported that the strong adsorption of CO molecules on the (111) planes of the Pd nanosheets prevents the growth along the [111] direction, resulting the formation of the plate-like structures [40]. The generating wavelength of the LSPR of the nanoplates mainly depends on the aspect ratio, i.e. the ratio of the diameter of the front face to its thickness [44]. The STEM images of the Pd nanoplates synthesized in this study are shown in Figs. 1(a)–1(d). Two typical arrangements i.e., extended lamellar structures consisting of stacked Pd nanoplates and flat-lying nanoplates (indicated by a red, dashed circle, Figs. 1(c) and 1(d)) were observed. The average thickness of the nanoplates was estimated to be 1.5 ( ± 0.2) nm which is almost the same as described in a previous report [40] for some lamellar structures(Fig. 1(b)). The average diameter of the nanoplate was estimated to be 16.3 ( ± 2.0) nm. Thus, the aspect ratio of the nanoplates was approximately evaluated to be 11 (Fig. 1(e)).

The extinction spectrum of the colloidal ethanol solution of the Pd nanoplates is shown in Fig. 2. The solution showed a broad LSPR peak at approximately 620 nm in the visible region.

 figure: Fig. 2

Fig. 2 Extinction spectrum of the Pd nanoplates. (a) Observed spectrum for a colloidal ethanol solution. (b) Calculated spectrum (RI of the surrounding medium: 1.361) of the Pd nanoplate (Fig. 1(e)).

Download Full Size | PDF

The generating wavelength of the LSPR, as well as its spectral shape, agreed well with that calculated by QS-BEM based on the nanoplate shape estimated using the STEM images.

Although the experimental LSPR peak was slightly shorter than that calculated, this could be due to a slight variation in the nanoplate shapes.

Next, we investigated the change in an LSPR wavelength (λmax) of the Pd nanoplate upon the change in RI of the surrounding medium. The RI of the surrounding medium was controlled by the mixing ratio of ethanol (n = 1.361) and glycerol (n = 1.474), and the values were obtained by the following empirical polynomial expression [45].

n(mixed  solvent)=0.0104(0.0171a0.0171a+0.0137b)30.0362(0.0171a0.0171a+0.0137b)2                               0.0658(0.0171a0.0171a+0.0137b)+1.4742
where a and b are the volume fractions of ethanol (a) and glycerol (b) in the mixed solvent (a + b = 1). In the experiment (Fig. 3(a)), the LSPR peak from the colloidal solution of the Pd nanoplates was red-shifted upon increasing the glycerol concentration (10, 20, 30, 40, and 50%) corresponding to an increase in the RI of the surrounding medium (nbulk). As shown in Fig. 3(b), the theoretically-obtained LSPR peak was also red-shifted upon increasing the RI. As summarized in Fig. 3(c), the LSPR peak shifted for both experiments and theory; Δλmax = (λmax in 20–50% glycerol solution) – (λmax in 10% glycerol solution) varied linearly with the RI (1.375–1.425) of the surrounding medium. The RI susceptibility, defined as Sexp = Δλmaxnbulk was obtained from the linear slope to be 250 nm RIU−1, where RIU is the refractive index unit. This value of the RI susceptibility for the Pd nanoplates is comparable to those of some anisotropic Au nanoparticles which have been suggested to be excellent plasmonic RI sensing nanomaterials, such as Au nanorods (198–288 nm RIU−1) [23–25], Au nanopyramids (130–221 nm RIU−1) [27], Au nanobars (219 nm RIU−1) [46], and Au nanotube arrays (250 nm RIU−1) [47]. In addition, while the synthetic routes of these anisotropic gold nanoparticles are quitecomplicated because two or more steps or nanofabrication apparatuses are required, the Pd nanoplates can be simply synthesized by one step. It has thus been demonstrated that the Pd nanoplates can function as an excellent RI sensing material. On the other hand, the susceptibility was substantially lower than that the theory predicted (541 nm RIU−1, Fig. 3(c)). This was possibly because the PVP, which coated the nanoplates as a protective layer, reduced the RI change in the surrounding medium in the vicinity of the nanoplates. From the STEM image of the lamellar structures of the stacked Pd nanoplates (Fig. 1(b)), the distance between the neighboring Pd nanoplates was estimated to be 1.8 nm, giving a thickness of 0.9 nm for the thin PVP films coating the nanoplates, even under dry conditions. To evaluate the contribution of 0.9 nm-thick PVP films to the RI susceptibility, the QS-BEM calculation of PVP-coated Pd nanoplate predicted a much lower RI susceptibility (352 nm RIU−1) compared to that of the bare Pd nanoplate (Fig. 4). Therefore, it is pointed out that the lower experimental RI susceptibility compared to the theoretical prediction was attributed to the existence of the PVP protecting layer. However, the experimentally-obtained RI susceptibility of the PVP-coated Pd nanoplates is still lower than the theoretically-obtained value. This is possibly caused by limited accuracy of the PVP thickness because this was estimated by the STEM images under a dry condition. In order to evaluate our hypothesis, we calculated the dependence of the RI susceptibility of the PVP-coated Pd nanoplates on the PVP thickness.

 figure: Fig. 3

Fig. 3 Shifts of extinction peak of the Pd nanoplates dispersed in media with different RI. Spectra are normalized to highlight the spectral shift. (a) Observed extinction spectra of the Pd nanoplates dispersed in different solvent mixtures (10, 20, 30, 40, and 50 vol. % glycerol ethanol solutions). (b) Calculated extinction spectra of the Pd nanoplates in different RI solvents (n = 1.375, 1.388, 1.401, 1.413, and 1.425 corresponding to ethanol containing 10, 20, 30, 40, and 50 vol.% glycerol, respectively). (c) Dependences of the LSPR peak shifts upon RI changes from (A) experimental and (B) theoretical results.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 (a) Schematic illustration of PVP-coated Pd nanoplate. (b) Theoretical dependences of the LSPR peak shifts upon RI changes of (A) the bare Pd nanoplate (same as (B) in Fig. 3(c)) and (B) the PVP-coated Pd nanoplate and experimental dependence of the peak shifts of (C) the nanoplates (same as (A) in Fig. 3(c)).

Download Full Size | PDF

As shown in Fig. 5, it was found that the RI susceptibility sensitively changes with a small change in the thickness. For example, the susceptibility was significantly decreased to 195 nm RIU−1 with increasing the thickness of PVP to 2.4 nm. It is therefore strongly suggested that the Pd nanoplates will show a higher RI susceptibility than some excellent plasmonic RI sensing nanomaterials as described above [23–25, 27, 46, 47].

 figure: Fig. 5

Fig. 5 (a) Theoretical dependences of the LSPR peak shifts upon RI changes of the Pd nanoplates with the PVP of (A) 0, (B) 0.6, (C) 0.9, (D) 1.2, (E) 1.8, and (F) 2.4 nm in thicknesses. (b) PVP thickness dependence of the RI susceptibility of the Pd nanoplates.

Download Full Size | PDF

4. Conclusion

We investigated the RI susceptibility of plasmonic Pd nanoplates as a novel RI-sensing material. The LSPR peak attributed to the in-plane dipole mode of the nanoplates showed a comparable susceptibility to those exerted by some excellent anisotropic Au nanoparticles. On the otherhand, the susceptibility was substantially lower than the theoretically calculated. This is possibly due to the effect of the PVP thin film coating the nanoplates. The Pd nanoplates may even show a higher susceptibility by removing the PVP protective agent, or replacing it with thinner protective agents, beneficial for the development of highly sensitive label-free and cost-effective interfacial biosensing platforms. Research in this direction is currently underway in our laboratory.

Acknowledgment

This work was supported by Grant-in-Aid for Young Scientists B (Grant No. 26810102) from JSPS KAKENHI, Applied Research Grant from Nihon University, and the Strategic Foundation at Private Universities, Mext and Nihon University.

References and links

1. C. F. Bohren and D. R. Huffman, Absorption and Scattering of Light by Small Particles (Wiley Science Paperback Series, 1998).

2. K. L. Kelly, E. Coronado, L. L. Zhao, and G. C. Schatz, “The optical properties of metal nanoparticles: the influence of size, shape, and dielectric environment,” J. Phys. Chem. B 107(3), 668–677 (2003). [CrossRef]  

3. J. Zhao, A. O. Pinchuk, J. M. McMahon, S. Li, L. K. Ausman, A. L. Atkinson, and G. C. Schatz, “Methods for describing the electromagnetic properties of silver and gold nanoparticles,” Acc. Chem. Res. 41(12), 1710–1720 (2008). [CrossRef]   [PubMed]  

4. V. Giannini, A. I. Fernández-Domínguez, S. C. Heck, and S. A. Maier, “Plasmonic nanoantennas: fundamentals and their use in controlling the radiative properties of nanoemitters,” Chem. Rev. 111(6), 3888–3912 (2011). [CrossRef]   [PubMed]  

5. M. Thämer, A. Kartouzian, P. Heister, T. Lünskens, S. Gerlach, and U. Heiz, “Small supported plasmonic silver clusters,” Small 10(12), 2340–2344 (2014). [CrossRef]   [PubMed]  

6. T. Lünskens, P. Heister, M. Thämer, C. A. Walenta, A. Kartouzian, and U. Heiz, “Plasmons in supported size-selected silver nanoclusters,” Phys. Chem. Chem. Phys. 17(27), 17541–17544 (2015). [CrossRef]   [PubMed]  

7. J. Tiggesbaeumker, L. Koeller, K. H. Meiwes-Broer, and A. Liebsch, “Blue shift of the Mie plasma frequency in silver clusters and particles,” Phys. Rev. A: At. Mol. Opt. Phys. 48(3), R1749–R1752 (1993). [CrossRef]  

8. S. Fedrigo, W. Harbich, and J. Buttet, “Collective dipole oscillations in small silver clusters embedded in rare-gas matrices,” Phys. Rev. B Condens. Matter Mater. Phys. 47(16), 10706–10715 (1993). [CrossRef]  

9. J. J. Mock, S. M. Norton, S.-Y. Chen, A. A. Lazarides, and D. R. Smith, “Electromagnetic Enhancement Effect Caused by Aggregation on SERS-Active Gold Nanoparticles,” Plasmonics 6(1), 113–124 (2011). [CrossRef]  

10. S. M. Adams, S. Campione, J. D. Caldwell, F. J. Bezares, J. C. Culbertson, F. Capolino, and R. Ragan, “Non-lithographic SERS substrates: tailoring surface chemistry for Au nanoparticle cluster assembly,” Small 8(14), 2239–2249 (2012). [CrossRef]   [PubMed]  

11. S. Campione, S. M. Adams, R. Ragan, and F. Capolino, “Comparison of electric field enhancements: linear and triangular oligomers versus hexagonal arrays of plasmonic nanospheres,” Opt. Express 21(7), 7957–7973 (2013). [CrossRef]   [PubMed]  

12. G. M. Akselrod, C. Argyropoulos, T. B. Hoang, C. Ciracì, C. Fang, J. Huang, D. R. Smith, and M. H. Mikkelsen, “Probing the mechanisms of large Purcell enhancement in plasmonic nanoantennas,” Nat. Photonics 8(11), 835–840 (2014). [CrossRef]  

13. S. Underwood and P. Mulvaney, “Effect of the solution refractive index on the color of gold colloids,” Langmuir 10(10), 3427–3430 (1994). [CrossRef]  

14. J. N. Anker, W. P. Hall, O. Lyandres, N. C. Shah, J. Zhao, and R. P. Van Duyne, “Biosensing with plasmonic nanosensors,” Nat. Mater. 7(6), 442–453 (2008). [CrossRef]   [PubMed]  

15. K. M. Mayer and J. H. Hafner, “Localized surface plasmon resonance sensors,” Chem. Rev. 111(6), 3828–3857 (2011). [CrossRef]   [PubMed]  

16. L. Tian, J. J. Morrissey, R. Kattumenu, N. Gandra, E. D. Kharasch, and S. Singamaneni, “Bioplasmonic paper as a platform for detection of kidney cancer biomarkers,” Anal. Chem. 84(22), 9928–9934 (2012). [CrossRef]   [PubMed]  

17. Y.-F. Chang, S.-H. Hung, Y.-J. Lee, R.-C. Chen, L.-C. Su, C.-S. Lai, and C. Chou, “Discrimination of breast cancer by measuring prostate-specific antigen levels in women’s serum,” Anal. Chem. 83(13), 5324–5328 (2011). [CrossRef]   [PubMed]  

18. P. L. Truong, B. W. Kim, and S. J. Sim, “Rational aspect ratio and suitable antibody coverage of gold nanorod for ultra-sensitive detection of a cancer biomarker,” Lab Chip 12(6), 1102–1109 (2012). [CrossRef]   [PubMed]  

19. C. Sönnichsen, B. M. Reinhard, J. Liphardt, and A. P. Alivisatos, “A molecular ruler based on plasmon coupling of single gold and silver nanoparticles,” Nat. Biotechnol. 23(6), 741–745 (2005). [CrossRef]   [PubMed]  

20. T. Sannomiya, C. Hafner, and J. Voros, “In situ sensing of single binding events by localized surface plasmon resonance,” Nano Lett. 8(10), 3450–3455 (2008). [CrossRef]   [PubMed]  

21. K. M. Mayer, S. Lee, H. Liao, B. C. Rostro, A. Fuentes, P. T. Scully, C. L. Nehl, and J. H. Hafner, “A label-free immunoassay based upon localized surface plasmon resonance of gold nanorods,” ACS Nano 2(4), 687–692 (2008). [CrossRef]   [PubMed]  

22. Y. Yang, I. I. Kravchenko, D. P. Briggs, and J. Valentine, “All-dielectric metasurface analogue of electromagnetically induced transparency,” Nat. Commun. 5, 5753 (2014). [CrossRef]   [PubMed]  

23. H. Chen, X. Kou, Z. Yang, W. Ni, and J. Wang, “Shape- and size-dependent refractive index sensitivity of gold nanoparticles,” Langmuir 24(10), 5233–5237 (2008). [CrossRef]   [PubMed]  

24. N. Omura, I. Uechi, and S. Yamada, “Comparison of plasmonic sensing between polymer- and silica-coated gold nanorods,” Anal. Sci. 25(2), 255–259 (2009). [CrossRef]   [PubMed]  

25. P. K. Jain, X. Huang, I. H. El-Sayed, and M. A. El-Sayed, “Noble metals on the nanoscale: optical and photothermal properties and some applications in imaging, sensing, biology, and medicine,” Acc. Chem. Res. 41(12), 1578–1586 (2008). [CrossRef]   [PubMed]  

26. C. L. Nehl, H. Liao, and J. H. Hafner, “Optical properties of star-shaped gold nanoparticles,” Nano Lett. 6(4), 683–688 (2006). [CrossRef]   [PubMed]  

27. J. Lee, W. Hasan, and T. W. Odom, “Tuning the thickness and orientation of single Au pyramids for improved refractive index sensitivities,” J Phys Chem C Nanomater Interfaces 113(6), 2205–2207 (2009). [CrossRef]   [PubMed]  

28. Y. Khalavka, J. Becker, and C. Sönnichsen, “Synthesis of rod-shaped gold nanorattles with improved plasmon sensitivity and catalytic activity,” J. Am. Chem. Soc. 131(5), 1871–1875 (2009). [CrossRef]   [PubMed]  

29. M. J. Banholzer, N. Harris, J. E. Millstone, G. C. Schatz, and C. A. Mirkin, “abnormally large plasmonic shifts in silica-protected gold triangular nanoprisms,” J. Phys. Chem. C 114(16), 7521–7526 (2010). [CrossRef]  

30. D. E. Charles, D. Aherne, M. Gara, D. M. Ledwith, Y. K. Gun’ko, J. M. Kelly, W. J. Blau, and M. E. Brennan-Fournet, “Versatile solution phase triangular silver nanoplates for highly sensitive plasmon resonance sensing,” ACS Nano 4(1), 55–64 (2010). [CrossRef]   [PubMed]  

31. A. J. Haes and R. P. Van Duyne, “A Nanoscale Optical Biosensor: Sensitivity and selectivity of an approach based on the localized surface plasmon resonance spectroscopy of triangular silver nanoparticles,” J. Am. Chem. Soc. 124(35), 10596–10604 (2002). [CrossRef]   [PubMed]  

32. K. Sugawa, D. Tanaka, T. Ichikawa, and N. Takeshima, “Development of plasmon resonance sensing based on alkylthiol-coated triangular silver nanoplates on glass plates,” Jpn. J. Appl. Phys. 52(4), 04CK06 (2013). [CrossRef]  

33. L. J. Sherry, S.-H. Chang, G. C. Schatz, R. P. Van Duyne, B. J. Wiley, and Y. Xia, “Localized surface plasmon resonance spectroscopy of single silver nanocubes,” Nano Lett. 5(10), 2034–2038 (2005). [CrossRef]   [PubMed]  

34. M. A. Mahmoud and M. A. El-Sayed, “Gold nanoframes: very high surface plasmon fields and excellent near-infrared sensors,” J. Am. Chem. Soc. 132(36), 12704–12710 (2010). [CrossRef]   [PubMed]  

35. E. M. Larsson, J. Alegret, M. Käll, and D. S. Sutherland, “Sensing characteristics of NIR localized surface plasmon resonances in gold nanorings for application as ultrasensitive biosensors,” Nano Lett. 7(5), 1256–1263 (2007). [CrossRef]   [PubMed]  

36. K. Sugawa, H. Tahara, A. Yamashita, J. Otsuki, T. Sagara, T. Harumoto, and S. Yanagida, “Refractive index susceptibility of the plasmonic palladium nanoparticle: potential as the third plasmonic sensing material,” ACS Nano 9(2), 1895–1904 (2015). [CrossRef]   [PubMed]  

37. J. Cadet and J. R. Wagner, “Oxidatively generated base damage to cellular DNA by hydroxyl radical and one-electron oxidants: similarities and differences,” Arch. Biochem. Biophys. 557(1), 47–54 (2014). [CrossRef]   [PubMed]  

38. A. M. Pisoschi and A. Pop, “The role of antioxidants in the chemistry of oxidative stress: A review,” Eur. J. Med. Chem. 97(5), 55–74 (2015). [CrossRef]   [PubMed]  

39. K. Sugasawa, “Regulation of damage recognition in mammalian global genomic nucleotide excision repair,” Mutat. Res. 685(1-2), 29–37 (2010). [CrossRef]   [PubMed]  

40. X. Huang, S. Tang, X. Mu, Y. Dai, G. Chen, Z. Zhou, F. Ruan, Z. Yang, and N. Zheng, “Freestanding palladium nanosheets with plasmonic and catalytic properties,” Nat. Nanotechnol. 6(1), 28–32 (2011). [CrossRef]   [PubMed]  

41. J. Waxenegger, A. Trügler, and U. Hohenester, “Plasmonics simulations with the MNPBEM toolbox: Consideration of substrates and layer structures,” Comput. Phys. Commun. 193, 138–150 (2015). [CrossRef]  

42. T. Pakizeh, C. Langhammer, I. Zorić, P. Apell, and M. Käll, “Intrinsic Fano Interference of Localized Plasmons in Pd Nanoparticles,” Nano Lett. 9(2), 882–886 (2009). [CrossRef]   [PubMed]  

43. T. A. F. König, P. A. Ledin, J. Kerszulis, M. A. Mahmoud, M. A. El-Sayed, J. R. Reynolds, and V. V. Tsukruk, “Electrically tunable plasmonic behavior of nanocube-polymer nanomaterials induced by a redox-active electrochromic polymer,” ACS Nano 8(6), 6182–6192 (2014). [CrossRef]   [PubMed]  

44. C. Langhammer, Z. Yuan, I. Zorić, and B. Kasemo, “Plasmonic properties of supported Pt and Pd nanostructures,” Nano Lett. 6(4), 833–838 (2006). [CrossRef]   [PubMed]  

45. A. S. Alkindi, Y. M. Al-Wahaibi, and A. H. Muggeridge, “Physical properties (density, excess molar volume, viscosity, surface tension, and refractive index) of ethanol + glycerol,” J. Chem. Eng. Data 53(12), 2793–2796 (2008). [CrossRef]  

46. Y. H. Lee, H. Chen, Q.-H. Xu, and J. Wang, “Refractive index sensitivities of noble metal nanocrystals: the effects of multipolar plasmon resonances and the metal type,” J. Phys. Chem. C 115(16), 7997–8004 (2011). [CrossRef]  

47. J. McPhillips, A. Murphy, M. P. Jonsson, W. R. Hendren, R. Atkinson, F. Höök, A. V. Zayats, and R. J. Pollard, “High-performance biosensing using arrays of plasmonic nanotubes,” ACS Nano 4(4), 2210–2216 (2010). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 Pd nanoplates synthesized in this study. (a)-(c) BF-STEM and (d) HAADF-STEM images and (e) Schematic illustration and model structure approximated by a regular polygon with 40 sides. E and k are the electric field vector and the wave vector of the inciendent light.
Fig. 2
Fig. 2 Extinction spectrum of the Pd nanoplates. (a) Observed spectrum for a colloidal ethanol solution. (b) Calculated spectrum (RI of the surrounding medium: 1.361) of the Pd nanoplate (Fig. 1(e)).
Fig. 3
Fig. 3 Shifts of extinction peak of the Pd nanoplates dispersed in media with different RI. Spectra are normalized to highlight the spectral shift. (a) Observed extinction spectra of the Pd nanoplates dispersed in different solvent mixtures (10, 20, 30, 40, and 50 vol. % glycerol ethanol solutions). (b) Calculated extinction spectra of the Pd nanoplates in different RI solvents (n = 1.375, 1.388, 1.401, 1.413, and 1.425 corresponding to ethanol containing 10, 20, 30, 40, and 50 vol.% glycerol, respectively). (c) Dependences of the LSPR peak shifts upon RI changes from (A) experimental and (B) theoretical results.
Fig. 4
Fig. 4 (a) Schematic illustration of PVP-coated Pd nanoplate. (b) Theoretical dependences of the LSPR peak shifts upon RI changes of (A) the bare Pd nanoplate (same as (B) in Fig. 3(c)) and (B) the PVP-coated Pd nanoplate and experimental dependence of the peak shifts of (C) the nanoplates (same as (A) in Fig. 3(c)).
Fig. 5
Fig. 5 (a) Theoretical dependences of the LSPR peak shifts upon RI changes of the Pd nanoplates with the PVP of (A) 0, (B) 0.6, (C) 0.9, (D) 1.2, (E) 1.8, and (F) 2.4 nm in thicknesses. (b) PVP thickness dependence of the RI susceptibility of the Pd nanoplates.

Equations (1)

Equations on this page are rendered with MathJax. Learn more.

n( mixed  solvent )=0.0104 ( 0.0171a 0.0171a+0.0137b ) 3 0.0362 ( 0.0171a 0.0171a+0.0137b ) 2                                0.0658( 0.0171a 0.0171a+0.0137b )+1.4742
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.