Gold-Nanorod-Based Plasmonic Nose for Analysis of Chemical

May 14, 2019 - ... chemical analyte detection for a wide variety of applications including life sciences, environmental monitoring, and homeland secur...
1 downloads 0 Views 1MB Size
Subscriber access provided by NJIT | New Jersey Institute of Technology

Article

Gold Nanorod-based Plasmonic Nose for Analysis of Chemical Mixtures Huzeyfe Yilmaz, Sang Hyun Bae, Sisi Cao, Zheyu Wang, Baranidharan Raman, and Srikanth Singamaneni ACS Appl. Nano Mater., Just Accepted Manuscript • DOI: 10.1021/acsanm.9b00765 • Publication Date (Web): 14 May 2019 Downloaded from http://pubs.acs.org on May 15, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

Gold Nanorod-based Plasmonic Nose for Analysis of Chemical Mixtures Huzeyfe Yilmaz†, Sang Hyun Bae†, Sisi Cao†, Zheyu Wang†, Baranidharan Raman*‡ and Srikanth Singamaneni*† †Department

of Mechanical Engineering and Materials Science, and Institute of Materials

Science and Engineering, ‡Department

of Biomedical Engineering, Washington University in St. Louis, St. Louis, Missouri

63130, United States E-mail: [email protected], [email protected] KEYWORDS: Plasmonic calligraphy, SERS, multiplexed sensing, dimensionality reduction, quantitative SERS

ABSTRACT: We introduce “plasmonic nose” as a novel approach for detection, recognition and quantification of mixtures of chemical species. Using a paper substrate, and a calligraphy-based fabrication approach, we generated an array of surface enhanced Raman scattering (SERS)active sensors with distinct chemical functionalities. Each sensor is comprised of gold nanorods (AuNRs) functionalized with a macromolecule that determined its sensitivity and specificity. We show that the SERS-active sensor array is capable of detecting and discriminating a wide variety of chemical species. To validate this approach, we exposed the sensor array to individual

ACS Paragon Plus Environment

1

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 24

analytes, and their binary/ternary mixtures. We found that each mixture generated a multi-variate fingerprint that varied with identity (vibrational frequency) and intensity. Statistical analysis of SERS spectra from multiple sensors allowed us to not only recognize components of mixtures but also estimate their mixing ratios. In sum, our study presents a highly practical, low-cost sensing approach for quantitative chemical analyte detection for a wide variety of applications including life sciences, environmental monitoring and homeland security.

INTRODUCTION Surface enhanced Raman scattering (SERS) is considered to be a powerful platform for ultrasensitive chemical and biological sensing1–6 and trace detection.7,8 Design and synthesis of plasmonic nanostructures with large enhancement factors allow detection of target analytes at extremely low concentrations9–20. To take advantage of Raman scattering enhancement, it is necessary that molecules of the analyte are in close proximity of the surface of plasmonic nanostructures.9,21 This proximity requirement could be achieved through both physical (i.e. entrapment close to the plasmonic nanostructure surface) or chemical interactions. Analyte specific coatings that utilize antigen-antibody or analyte-receptor interactions, have been successfully employed22–24 to capture molecules from chemical species of interest (i.e. impart selectivity) and take advantage of sensitivity enhancements in plasmonic nanosensors. While having specific chemical targets are often desirable, in most diagnostic or detection problems a unique single marker/analyte is not present. Instead, a number of species of interest coexist, which require the use of arrays of sensors as differential receptors. A crossresponsive SERS-sensor array can detect, identify and quantify species of interest in complex mixtures. Various transduction methods have been utilized with cross-responsive sensing

ACS Paragon Plus Environment

2

Page 3 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

elements as an alternative to “lock-key” recognition based-sensing (i.e., ‘electronic, optical or chemical nose’).25–28 Despite intensive research with cross-reactive array sensors, SERS remains underappreciated as a transduction mechanism.29 In fact, very few generalized quantification methods for have been demonstrated for SERS based sensors so far.30–32 In addition to differential functionalization of transducers, development of arrays of SERS-based sensors would require robust, cost-effective fabrication of substrates to steer this class of sensing technologies into the real world settings.33–35 Amongst many alternatives, a substrate that offers a number of advantages for creating SERS-based sensing arrays is ‘paper’.36–39 It is inexpensive, mechanically flexible, easy to use and compatible with traditional printing techniques. Plasmonic calligraphy or pen-on-paper is a method that has proven to be highly effective in generating uniform patterns of plasmonic nanoparticles on paper substrates,40– 42

transcending immersion method43,44 and circumventing time consuming and costly approaches

such as plasmonic ink-jet printing.45,46 In addition, calligraphy of functional plasmonic nanoparticles can easily be scaled up and automated with robotic arms. Here we introduce “plasmonic nose”, that consists of an array of plasmonic nanostructures functionalized with a variety of polymers. Each polymer offers distinct set of chemical interactions which imparts partial selectivity and allows us to create a cross-responsive sensor. Arrays of such SERS-active sensors are fabricated on paper substrates using a calligraphy-based approach. We validate our plasmonic nose approach on the problem of identifying and quantifying the components of binary and ternary mixtures. In sum, we combine nanoparticle array-based SERS spectroscopy with statistical methods to realize a quantitative sensor for mixture analysis. RESULTS AND DISCUSSION

ACS Paragon Plus Environment

3

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 24

Figure 1. Plasmonic nose based on cross-responsive SERS sensors calligraphed on a paper substrate. (A) Chemical structures of polymers coated on AuNRs and the dominant interactions with target analytes are shown. Note that four different sensors were synthesized by using the following polymeric coatings: two polyelectrolytes (PAH, PSS), PVP and PEG. (B) Zeta potentials of polymer coated AuNRs in aqueous solution. (C) TEM image of AuNRs. (D) Illustration of plasmonic calligraphy showing how the cross-responsive SERS-sensing strips were fabricated on flexible paper-based substrates. (E) Optical image of a ~2 cm test strip with each line corresponding to AuNRs coated with a distinct organic molecule i.e. four distinct sensors. (F) SEM image of a typical sensor exhibiting uniform and dense distribution of AuNRs .

To serve as SERS-active medium, we chose gold nanorods (AuNRs) as plasmonic nanostructures due to the strong electromagnetic fields at their sharp ends and tunable longitudinal localized surface plasmon resonance (LSPR) wavelength.47–50 Dimensions of AuNRs were measured from transmission electron microscopy (TEM) images to be 55 ± 5 nm in length and 18 ± 1 nm in diameter (Figure 1C). Modification with polyelectrolytes was initiated with negatively charged poly(styrene sulfonate) (PSS) through electrostatic interactions since as synthesized AuNRs are capped with positively charged cetyltrimethylammonium bromide

ACS Paragon Plus Environment

4

Page 5 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

(CTAB). Positively charged poly(allylamine hydrochloride) (PAH) was similarly coated on PSSmodified AuNRs. Poly(vinyl pyrrolidinone) and poly(ethylene glycol) coating was obtained via ligand exchange (see experimental details for more information). Likely interactions and chemical structures of polymer coatings are shown in Figure 1A. Functionalization of the AuNRs with different polymers resulted in a relatively small red or blue shift corresponding to an increase or decrease in the effective refractive index of the medium surrounding the AuNRs (Figure S1). It is important to note that the surface functionalization procedures did not result in uncontrolled aggregation of the nanostructures as evidenced by the absence of large red shift or LSPR peak broadening. Further confirmation of the functionalization was obtained through the zeta potential values, -51 mV for PSS-AuNRs, 46 mV for PAH-AuNRs, -20 mV for PEG-AuNRs and -13 mV for PVP-AuNRs. Polyelectrolyte (PAH and PSS) coatings with large (positive and negative) zeta potentials facilitate electrostatic interactions, while PEG and PVP allow noncovalent interactions such as hydrogen bonding and polar interactions (Figure 1B). Polymer-coated AuNRs were concentrated to form plasmonic inks that were calligraphed onto a filter paper using a 0.7 mm ballpoint pen separately to create distinct, cross-responsive plasmonic sensor array as shown in the schematic in Figure 1D. Ballpoint pens are cost-effective and simple tools for conception of fine and uniform structures on ordinary laboratory papers since plasmonic inks prepared from polymer-capped AuNRs can be easily adjusted (by controlling the concentration of the AuNRs) to have suitable vistosities. Each sensor observed in the optical image was obtained by two strokes with the ballpoint pen (Figure 1E). The uniformity of the distribution of AuNRs was confirmed via scanning electron microscopy (SEM) images (Figures 1F and S2). After printing, AuNR sensing domains preserved their optical

ACS Paragon Plus Environment

5

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 24

Figure 2. (A) Chemical structures of the model analytes used in the study. SERS spectra of ternary mixtures of MO, R6G and FITC collected from (A) PAH-AuNRs, (B) PEG-AuNRs, (C) PSS-AuNRs and (D) PVP-AuNRs domains. Each spectrum is an average of ten measurements. Note that each spectrum is obtained for a ternary mixture with a particular mixing ratio. For example [3, 9, 0.6] indicates a mixture of 3 μM MO, 9 μM R6G and 0.6 μM FITC. Also note that the spectra were offset to show multiple spectra in the same plot and only the location of peaks carry analyte specific information.

properties since the extinction spectra collected from all four domains on the paper substrate were similar to that from aqueous solutions except for a blue shift in the LSPR wavelength due

ACS Paragon Plus Environment

6

Page 7 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

to the change in the refractive index of the medium surrounding the nanostructures (Figures S1 and S3). To test the multiplexing capability of the plasmonic sensor array, a set of representative analytes, capable of interacting with the SERS sensors through distinctive interactions were chosen to test the plasmonic nose. Negatively charged methyl orange (MO), positively charged rhodamine 6G (R6G) and fluorescein isothiocyanate (FITC) provided a wide range of interactions that can also be found in many biological systems, presenting a candid challenge for our plasmonic nose. To create binary and ternary mixtures that were not dominated by any one analyte, we characterized the dose-response curves of each analyte. Based on the dose-response curves (FigureS5), analyte concentrations were varied between 3 μM – 12 μM for MO and R6G, 0.3 μM – 1.2 μM for FITC in binary mixtures, and between 3 μM – 18 μM for MO and R6G, 0.3 μM – 1.8 μM for FITC in ternary mixtures. The SERS spectra obtained using for these three target analytes and their binary and ternary mixtures are shown in Figure 2. Note that the SERS spectra obtained using the four differentially functionalized AuNRs were quite different, and that different plasmonic sensors were effective at capturing different analytes. For example, R6G has characteristic Raman bands at 612 cm-1 (C-C-C ring in-plane vibration) and 1363/1509 cm-1 (aromatic C-C stretching vibration)51 that were detected only using PSS-AuNR and PVPAuNRs. MO has bands at 1117 cm-1(Ph-N stretching) and 1143 cm-1 (C-H deformation) that were picked up only in the SERS spectra measured using the oppositely charged PAH-AuNRs52 (Figure 2). FITC Raman bands at 1186 cm-1 (Phenolic OH) and 1604 cm-1 (Xanthene C-C stretching) were noticeable in all four AuNRs with varying SERS intensities (Figures 2 andS6).53

ACS Paragon Plus Environment

7

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 24

Figure 3. Principal component (PC) analysis of SERS spectra based on sensor type. The entire dataset comprised of ten measurements made from twenty-eight mixtures using four different plasmonic sensors (1120 spectra in total). (A) Top three eigenvectors (with largest eigenvalues) of the covariance matrix of the entire dataset. (B) Each highdimensional Raman spectrum after dimensionality reduction is shown as a 3D sphere in the PC space. The measurements are color coded to indicate which type of nanosensor was used to obtain a specific spectrum. (C) Spectra were color coded according to analyte concentration on the respective favorable sensor (darker colors refer to higher concentrations). Three orthogonal directions validate the cross-responsive nature of SERS sensor array. (D) SERS spectra reconstructed (dark colors) from three concentration points in the PC space for each analyte-sensor pair are plotted with corresponding actual SERS spectra (light colors).

The entire dataset included 1120 spectra (twenty-eight analyte combinations, ten measurements each, and four different plasmonic sensors). Each Raman measurement was 1489 wave numbers long (in the range of 400-1800 cm-1 with a resolution of 0.94 cm-1). To qualitatively understand whether different plasmonic sensors were indeed contributing nonredundant information, we performed a principal component analysis (PCA). The top eigenvectors (with three largest eigenvalues; principal components or PCs) of the covariance of

ACS Paragon Plus Environment

8

Page 9 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

the above described data matrix (1120x1489) were computed (Figure 3A). Each 1489dimensional SERS spectrum was projected onto the PCs and visualized as shown in Figure 3B. Note that each SERS spectrum is color-coded to identify which of the four plasmonic sensors was used to obtain the measurement. By visual inspection, it can be observed that each functionality is clustered separately even though for certain scores (PVP) the separation is small. This is expected since a fine concentration gradient was used in the training set (for instance, for FITC, concentration steps were 0.3 μM). Based on the color coded scores in Figure 3C, three orthogonal directions indicate concentration gradients for each analyte. Note that the contribution from the fourth SERS sensor (PVP-AuNR) is not highlighted as it overlaps with PSS-AuNR and PEG-AuNR sensors. Overall, these results demonstrate the cross-responsive nature of the SERS sensor array. PCA of the (1120x1489) entire dataset did not reveal a clear variation between the Raman bands of each of the three analytes in the mixtures, instead captured the variation among sensors (Figure 3B). Nevertheless, PC analysis revealed that each SERS sensor provided separate information. Furthermore, each of the three eigenvectors with largest eigenvalues resembled the spectrum of a separate analyte (i.e., each axis encodes identity of one target analyte). Hence, projection of spectra along axis vectors allowed us to quantify analytes based on their distance from origin. Therefore, both identity and quantity could be determined by our analysis of SERS spectra. To systematically verify that the projections of SERS spectra along PC axes was according to the analyte concentration, three points (12 μM, 9 μM and 3 μM) from each direction were picked and used to reconstruct the rank-3 SERS spectra. Reconstructed spectra (dark colors) were compared to the original SERS spectra (light colors) in Figure 3D and peak

ACS Paragon Plus Environment

9

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 24

Figure 4. SERS data reorganization and PC analysis based on chemical mixture. (A) Average of ten concatenated Raman spectra for each ternary mixture is shown. (B) Top eigenvectors of the covariance matrix corresponding to two largest eigenvalues is shown. Note that only the 130 Raman measurements involving the ternary mixtures are used in this analysis. The first principal component vector (black solid line) has characteristic FITC peaks in negative direction and characteristic MO, R6G peaks in the positive direction. The second principal component vector has characteristic MO peaks in the positive direction and characteristic R6G peaks in the negative direction. (C) PCA scatter plot obtained by multiplying (dot product) the two eigenvectors shown in panel B with each row of the concatenated Raman matrix shown in panel A. Each point is color coded according to the mixture from which the spectra was measured. (D) PCA scatter plot including the blind test measurements. The scatter points of the two test cocktails are shown as hollow squares and hollow triangles.

positions and intensities were closely matched. The mismatching low intensity peaks can be attributed to the fact that only three eigenvectors were used in reconstruction, capturing 81 % of the variation. Next, we sought to examine how separable were the different analytes and their

ACS Paragon Plus Environment

10

Page 11 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

mixtures based on the SERS spectra obtained using all four plasmonic sensors. So, for each analyte mixture, we first concatenated the Raman spectra from the four sensors. This converted the original 1120×1489 Raman spectra matrix into 280×5956 concatenated Raman spectra matrix (Figure 4A). We performed principal component analysis to examine the predominant source of variance in the Raman spectra for each binary or ternary mixture (Figure 4B and Figure S8B, S9B and S10B) and to visualize the Raman spectra generated by each analyte across the plasmonic sensor array (Figure 4C and Figure S8C, S9C and S10C). For the ternary mixtures, the two primary sources of variance in the dataset (the ‘eigenspectra’), i.e. two eigenvectors of the covariance matrix corresponding to largest eigenvalues, are shown in Figure 4B. The first eigenspectra captured 68% of SERS spectra variance, contained all the Raman bands of all analytes. A separation of Raman bands that were specific to MO,R6G and FITC was observed (i.e., MO and R6G Raman bands had positive values where FITC Raman bands had negative values). Projecting the Raman spectra onto this eigenspectra (i.e. computing their dot product) therefore separated FITC from the other two analytes (MO and R6G). This result is expected as FITC interacted with all four sensing domains and contributed the most variance in this dataset. The second eigenspectra captured 18% of the variation. Note that the second eigenspectra does not have any of the FITC bands but clearly separates R6G and MO Raman bands (Figure 4B). When the same analyses were repeated but focusing on each binary mixture (Figure S8-S10), we find that the first eigenvectors always separated the Raman bands of the two components. Next, we visualized each concatenated Raman spectra in two dimensions by projecting the same onto the top two eigenspectra (dot product between the 5956-dimensional concatenated Raman spectra shown in Figure 4A and the eigenspectra shown in Figure 4B). The Raman

ACS Paragon Plus Environment

11

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 24

spectra after this dimensionality reduction was color coded and shown as a scatter plot (Figure 4C; training dataset). Note that exposures of the plasmonic sensor array to the pure analytes generated Raman spectra that mapped onto distinct clusters of points after PCA (blue cluster – FITC, black cluster – MO, and red cluster – R6G). Binary mixtures of two analytes resulted in spectra that were linear combinations of the two component spectra. Therefore, these binary mixture measurements projected onto space between the two component pure analyte clusters (for example, a binary mixture of MO and R6G [9, 9, 0] (shown in dark green), can be spotted between the two component clusters shown in red and black). Qualitatively similar results were obtained when different binary mixtures were individually analyzed (see Figures S8C, S9C, S10C). Ternary mixtures included contributions from all three analytes. The exact location of any mixture depended on the ratios with which the three components were mixed. To determine if the mixing ratios of an unknown mixture could be determined based on the concatenated Raman spectra, we prepared two cocktails not used in the training data (shown in Figure 4D; test data shown as black cluster of points with hollow squares and triangles). Note that the nearest cluster of points to one of unknown ternary mixture (shown in hollow triangles) was the [9, 6, 0.3] mixture. This is indeed compositionally the most similar to the actual mixing ratio of this mixture ([10, 4, 0.4]). Similarly, the other unknown ternary mixture (marked with hollow squares and mixing ratio: [4, 11, 0.3]) was mapped closest to the [6, 9, 0.3] mixture in the training data.

ACS Paragon Plus Environment

12

Page 13 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

To further quantify our predictions, we used a simple linear regression (after dimensionality

ACS Paragon Plus Environment

13

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 24

Figure 5. Results from the linear regression analysis. The three axes correspond to the concentrations of each analyte used in the study (MO, R6G and FITC). The actual composition of each testing-phase mixture used in the study is shown as red cubes. The predicted composition is shown as blue spheres. mixture ([10, 4, 0.4]). Similarly, the other unknown ternary mixture (marked with hollow squares and mixing ratio: [4, 11, 0.3]) was mapped closest to the [6, 9, 0.3] mixture in the training data.

reduction). A total of 8 blind tests were performed with binary and ternary mixtures and average error was calculated to be 11.4 %. Quantitative results of principal component regression from every blind test is shown in Figure 5. These results indicate that not only were the components of

ACS Paragon Plus Environment

14

Page 15 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

various binary and ternary mixtures robustly recognized, but their concentrations could also be calculated with good agreement to the actual values. In addition, we provided evidence for the contribution of each functionality and validated the use of 4 functionalities in our SERS sensor array by measuring the percentage error with only 2 or 3 functionalities (Figure S11). We found that on an average, the error in quantitative blind test measurements were 22.3% and 12.6% with PAH-PSS and PAH-PEG-PSS functionalities, respectively. Since the percentage error varied for each mixture (Figure S11), we also provided a visual observation of the accuracy of the plasmonic nose sensor based on mixtures, as shown in Figure S12. In sum, we have demonstrated differential and quantitative chemical sensing based on arrays of polymer functionalized gold nanorods. We used paper as a substrate to fabricate flexible, disposable and low-cost arrays of plasmonic nanosensors. Chemical diversity of polymers used in functionalization yielded a differential response which was confirmed by statistical analysis. Raman spectra collected from the sensor array exposed to different analytes allowed us to not only recognize the components of various binary and ternary mixtures, but also quantify the mixing ratios. Taken together, our results provide a first demonstration of the plasmonic nose concept where SERS-based measurements are used for chemical analyte detection. We have demonstrated a chemical sensing framework that is practical, selective and capable of quantitative analysis. Our sensor combines multivariate statistical analysis with SERS sensors and has the potential to serve as a reference for researchers in the field. Robust detection, identification and quantification provided by plasmonic nose can be utilized in real-world applications in the future.

MATERIALS AND METHODS

ACS Paragon Plus Environment

15

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 24

Cetyltrimethylammonium bromide (CTAB), chloroauric acid (HAuCl4), ascorbic acid, sodium borohydride (NaBH4), rhodamine 6G (R6G), methyl orange (MO), Fluorescein isothiocyanate (FITC), poly(styrene sulfonate) (Mw 70000g mol-1), poly(vinyl pyrrolidone) (Mw 29000g mol-1) were purchased from Sigma Aldrich. Poly(allyl amine hydrochloride) was purchased from Alfa Aesar, and methoxy PEG thiol (Mw 5000kDa) was purchased from JenKem Technology. Silver nitrate and filter paper (Whatman #1) was purchased from VWR international. All the chemicals have been used as received, with no further purification. Pilot-g1 retractable ballpoint pens (0.7mm) were bought from Amazon. Synthesis of Gold Nanorods (AuNRs) Gold nanorods were synthesized using a seed-mediated approach. Seed solution was prepared by adding 0.75 mL of an ice-cold sodium borohydride solution (10mM) into 10 mL of 0.1 M CTAB and 2.5×10-4 M HAuCl4 solution under 10 minutes of vigorous stirring at 800 rpm at room temperature. Once the sodium borohydride solution was added, the gold seed solution immediately changed from yellow to brown. Seed solution was stirred for five more minutes before use. Growth solution was prepared by mixing 1.8 mL of HAuCl4 (10 mM), 38 mL of CTAB (0.1 M), 0.415 mL of silver nitrate (10 mM), and 0.22 mL of ascorbic acid (0.1 M), in that order. The solution was vortexed after each addition for homogenization. To the resulting solution, 48 μL of freshly prepared seed solution was added, and set in the dark undisturbed for 14 h. The AuNR solution was centrifuged twice at 9000 rpm for 30 minutes to remove excess CTAB, and redispersed in nanopure water (18.2 mΩcm).47,54 Preparation of polymer coated gold nanorods(AuNRs) and corresponding plasmonic ink

ACS Paragon Plus Environment

16

Page 17 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

AuNRs were coated with selected polymers as previously reported.41 40 mL of twicecentrifuged AuNR solution was added dropwise to 40 mL of PSS (0.2% w/v) in NaCl aqueous solution (6 mM) under vigorous stirring of 1000 rpm, followed by stirring for one hour. Then the solution was sonicated for another hour. The above solution was centrifuged to remove excess PSS at 9000 rpm for 30 min, and concentrated to 40 μL solution, after which 10 μL of 2 % PSS was added. 40 mL of PSS encapsulated AuNR solution was added dropwise to 40 mL of PAH (0.2% w/v) in NaCl Aqueous solution (6 mM) under vigorous stirring of 1000 rpm, followed by stirring for one hour. Then the solution was sonicated for another hour. To remove excess PAH, the above solution was centrifuged at 9000 rpm for 30 min and concentrated to 40 μL solution after which 10 μL of 2% PAH was added. 40 mL of twice-centrifuged AuNR was added dropwise to 40 mL of PVP (w/v) in nanopure water, then shaken and sonicated for an hour with the same conditions as PSS and PAH capped AuNR solutions. During the second centrifuge before the dropwise addition, only 80 percent of the supernatant solution was removed. Once the sonication was complete, the PVP capped AuNR solution was centrifuged at 9000 rpm for 30 min and concentrated to 50 μL solution. 8 ml of 1% mPEG-Thiol solution was mixed with 200 μL of 100 mM NaCl. The resulting solution was added to 40ml of twice-centrifuged AuNR, where during the second centrifuge, only 80 percent of the supernatant was removed. The solution was sonicated for 1 hour, then immediately centrifuged at 9000 rpm for 30 minutes to remove excess PEG. The solution was concentrated to 50 μL. First, ball pen cartridges were emptied and washed with ethanol and nanopure water thoroughly via sonication. Then concentrated solutions of each polymer capped AuNR were

ACS Paragon Plus Environment

17

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 24

injected into separate ball point pen cartridges. SERS spectra measurements Whatman #1 laboratory filter paper was calligraphed with each polymer-capped AuNR solution at four distinct regions. Each line was written with a straight edge with two strokes over the same region. After each stroke, the paper was left to dry to prevent damages in the fiber. The width of the pen tip was 0.7 mm, as specified by the manufacturer. The substrates were then immersed and shaken in nanopure water for 10 minutes to release any loose nanorods, then dried naturally in air. Plasmonic inks were calligraphed soon after they were prepared. Extinction spectra of calligraphed plasmonic lines were measured from 10 random points on each polymerAuNR sensor (Figure S3). Uniform distribution of AuNRs on the filter paper suggests stable plasmonic inks and a uniform deposition from the ballpoint pen tip. Since each polymer interacts with the filter paper differently, a small variance between extinction values of AuNR-polymer domains are expected. Ultimately, the viscosity values observed from AuNR-polymer plasmonic inks fell within the compatible viscosity range for ballpoint pens to achieve uniform calligraphy.41 Once dried, they were cut into strips 5 mm across (Fig. 1E). The plasmonic sensors on the paper substrate were subsequently dipped in mixtures of different concentrations of MO, R6G, and FITC aqueous solutions of different concentrations for 1 h. The samples were then shaken in 5 mL of ethanol for 10 minutes, then shaken in 5 mL of nanopure water for 10 minutes and dried with nitrogen for 30 seconds. Neither final washing steps with ethanol and water nor immersion into aqueous analyte solutions discernably affected the adsorption or distribution of AuNRs on the paper substrate as confirmed by apparent Raman background from polymer coatings throughout the synthesis and measurement. (Figure S4)

ACS Paragon Plus Environment

18

Page 19 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

SERS spectra from the substrates were collected using Renishaw InVia Raman microscope and Wire 3.0 software on a dedicated computer, with a 785 nm laser focused using a 20× objective lens with 10 seconds of exposure. Ten spectra were collected from different locations across each substrate. Matlab® was used to perform data processing and statistical analysis. Prior to statistical analysis, background subtractions were performed as shown in Figure S13. Characterization techniques Transmission electron microscopy (TEM) micrographs were recorded on a JEM-2100F (JEOL) field emission instrument. A drop of the solution was dried on a carbon coated grid that was previously made hydrophilic by glow discharge. FEI Nova 2300 Field Emission Scanning Electron Microscope was used to obtain SEM images at an accelerating voltage of 10 kV. Plasmonic paper was sputtered with gold prior to SEM imaging. Zeta potential was measured using a Malvern Zetasizer Nano ZS Dynamic Light Scatterings system. UV-vis extinction spectra of nanorod solutions were collected using a Shimadzu UV-1800 UV-vis spectrophotometer. Extinction spectra of nanorods on paper substrates were collected using a CRAIC microspectrophotometer (QDI 302) coupled to a Leica optical microscope (DM 4000M) with 10x objective in the range of 450-800 nm with 50 accumulations and 0.144 s exposure time in reflection mode.

Acknowledgements We acknowledge support from Office of Naval Research (Award # N00014-16-1-3030). The authors thank Nano Research Facility (NRF) and Institute of Materials Science and Engineering (IMSE) at Washington University for providing access to electron microscopy facilities.

ACS Paragon Plus Environment

19

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 24

REFERENCES (1) Golightly, R. S.; Doering, W. E.; Natan, M. J. Surface-Enhanced Raman Spectroscopy and Homeland Security: A Perfect Match? ACS Nano 2009, 3 (10), 2859–2869. (2) Homola, J. Surface Plasmon Resonance Sensors for Detection of Chemical and Biological Species. Chem. Rev. 2008, 108 (2), 462–493. (3) Fang, X.; Ahmad, S. R. Detection of Explosive Vapour Using Surface-Enhanced Raman Spectroscopy. Appl. Phys. B 2009, 97 (3), 723. (4) Hering, K.; Cialla, D.; Ackermann, K.; Dörfer, T.; Möller, R.; Schneidewind, H.; Mattheis, R.; Fritzsche, W.; Rösch, P.; Popp, J. SERS: A Versatile Tool in Chemical and Biochemical Diagnostics. Anal. Bioanal. Chem. 2008, 390 (1), 113–124. (5) Moskovits, M. Surface-Enhanced Raman Spectroscopy: A Brief Retrospective. J. Raman Spectrosc. 2005, 36 (6–7), 485–496. (6) C. Bantz, K.; F. Meyer, A.; J. Wittenberg, N.; Im, H.; Kurtuluş, Ö.; Hoon Lee, S.; C. Lindquist, N.; Oh, S.-H.; L. Haynes, C. Recent Progress in SERS Biosensing. Phys. Chem. Chem. Phys. 2011, 13 (24), 11551–11567. (7) Campion, A.; Kambhampati, P. Surface-Enhanced Raman Scattering. Chem. Soc. Rev. 1998, 27 (4), 241–250. (8) Kneipp, J.; Kneipp, H.; Kneipp, K. SERS—a Single-Molecule and Nanoscale Tool for Bioanalytics. Chem. Soc. Rev. 2008, 37 (5), 1052–1060. (9) Nie, S.; Emory, S. R. Probing Single Molecules and Single Nanoparticles by SurfaceEnhanced Raman Scattering. Science 1997, 275 (5303), 1102–1106. (10) Kneipp, K.; Kneipp, H.; Itzkan, I.; Dasari, R. R.; Feld, M. S. Surface-Enhanced Non-Linear Raman Scattering at the Single-Molecule Level. Chem. Phys. 1999, 247 (1), 155–162. (11) Doering, W. E.; Nie, S. Single-Molecule and Single-Nanoparticle SERS:  Examining the Roles of Surface Active Sites and Chemical Enhancement. J. Phys. Chem. B 2002, 106 (2), 311–317. (12) Talley, C. E.; Jackson, J. B.; Oubre, C.; Grady, N. K.; Hollars, C. W.; Lane, S. M.; Huser, T. R.; Nordlander, P.; Halas, N. J. Surface-Enhanced Raman Scattering from Individual Au Nanoparticles and Nanoparticle Dimer Substrates. Nano Lett. 2005, 5 (8), 1569–1574. (13) Qin, L.; Zou, S.; Xue, C.; Atkinson, A.; Schatz, G. C.; Mirkin, C. A. Designing, Fabricating, and Imaging Raman Hot Spots. Proc. Natl. Acad. Sci. 2006, 103 (36), 13300–13303. (14) Kattumenu, R.; Lee, C. H.; Tian, L.; McConney, M. E.; Singamaneni, S. Nanorod Decorated Nanowires as Highly Efficient SERS-Active Hybrids. J. Mater. Chem. 2011, 21 (39), 15218–15223. (15) Li, Z.-Y.; Xia, Y. Metal Nanoparticles with Gain toward Single-Molecule Detection by Surface-Enhanced Raman Scattering. Nano Lett. 2010, 10 (1), 243–249. (16) Gandra, N.; Abbas, A.; Tian, L.; Singamaneni, S. Plasmonic Planet–Satellite Analogues: Hierarchical Self-Assembly of Gold Nanostructures. Nano Lett. 2012, 12 (5), 2645–2651. (17) Rycenga, M.; Xia, X.; Moran, C. H.; Zhou, F.; Qin, D.; Li, Z.-Y.; Xia, Y. Generation of Hot Spots with Silver Nanocubes for Single-Molecule Detection by Surface-Enhanced Raman Scattering. Angew. Chem. 123 (24), 5587–5591. (18) Liu, K.-K.; Tadepalli, S.; Wang, Z.; Jiang, Q.; Singamaneni, S. Structure-Dependent SERS Activity of Plasmonic Nanorattles with Built-in Electromagnetic Hotspots. Analyst 2017, 142 (23), 4536–4543.

ACS Paragon Plus Environment

20

Page 21 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

(19) Li, J. F.; Huang, Y. F.; Ding, Y.; Yang, Z. L.; Li, S. B.; Zhou, X. S.; Fan, F. R.; Zhang, W.; Zhou, Z. Y.; Wu, D. Y.; et al. Shell-Isolated Nanoparticle-Enhanced Raman Spectroscopy. Nature 2010, 464 (7287), 392. (20) Kleinman, S. L.; Ringe, E.; Valley, N.; Wustholz, K. L.; Phillips, E.; Scheidt, K. A.; Schatz, G. C.; Van Duyne, R. P. Single-Molecule Surface-Enhanced Raman Spectroscopy of Crystal Violet Isotopologues: Theory and Experiment. J. Am. Chem. Soc. 2011, 133 (11), 4115–4122. (21) Freeman, R. G.; Grabar, K. C.; Allison, K. J.; Bright, R. M.; Davis, J. A.; Guthrie, A. P.; Hommer, M. B.; Jackson, M. A.; Smith, P. C.; Walter, D. G.; et al. Self-Assembled Metal Colloid Monolayers: An Approach to SERS Substrates. Science 1995, 267 (5204), 1629– 1632. (22) Drachev, V. P.; Nashine, V. C.; Thoreson, M. D.; Ben-Amotz, D.; Davisson, V. J.; Shalaev, V. M. Adaptive Silver Films for Detection of Antibody−Antigen Binding. Langmuir 2005, 21 (18), 8368–8373. (23) Dou, X.; Takama, T.; Yamaguchi, Y.; Yamamoto, H.; Ozaki, Y. Enzyme Immunoassay Utilizing Surface-Enhanced Raman Scattering of the Enzyme Reaction Product. Anal. Chem. 1997, 69 (8), 1492–1495. (24) Ryu, K.; Haes, A. J.; Park, H.-Y.; Nah, S.; Kim, J.; Chung, H.; Yoon, M.-Y.; Han, S.-H. Use of Peptide for Selective and Sensitive Detection of an Anthrax Biomarker via Peptide Recognition and Surface-Enhanced Raman Scattering. J. Raman Spectrosc. 41 (2), 121– 124. (25) You, C.-C.; Miranda, O. R.; Gider, B.; Ghosh, P. S.; Kim, I.-B.; Erdogan, B.; Krovi, S. A.; Bunz, U. H. F.; Rotello, V. M. Detection and Identification of Proteins Using Nanoparticle– Fluorescent Polymer ‘Chemical Nose’ Sensors. Nat. Nanotechnol. 2007, 2 (5), 318. (26) Lim, S. H.; Feng, L.; Kemling, J. W.; Musto, C. J.; Suslick, K. S. An Optoelectronic Nose for the Detection of Toxic Gases. Nat. Chem. 2009, 1 (7), 562. (27) Ema, K.; Yokoyama, M.; Nakamoto, T.; Moriizumi, T. Odour-Sensing System Using a Quartz-Resonator Sensor Array and Neural-Network Pattern Recognition. Sens. Actuators 1989, 18 (3), 291–296. (28) Marco, S.; Gutierrez-Galvez, A. Signal and Data Processing for Machine Olfaction and Chemical Sensing: A Review. IEEE Sens. J. 2012, 12 (11), 3189–3214. (29) Pavel Anzenbacher, J.; Lubal, P.; Buček, P.; A. Palacios, M.; E. Kozelkova, M. A Practical Approach to Optical Cross-Reactive Sensor Arrays. Chem. Soc. Rev. 2010, 39 (10), 3954– 3979. (30) Lorén, A.; Engelbrektsson, J.; Eliasson, C.; Josefson, M.; Abrahamsson, J.; Johansson, M.; Abrahamsson, K. Internal Standard in Surface-Enhanced Raman Spectroscopy. Anal. Chem. 2004, 76 (24), 7391–7395. (31) J. Bell, S. E.; S. Sirimuthu, N. M. Quantitative Surface-Enhanced Raman Spectroscopy. Chem. Soc. Rev. 2008, 37 (5), 1012–1024. (32) Chen, H.-Y.; Lin, M.-H.; Wang, C.-Y.; Chang, Y.-M.; Gwo, S. Large-Scale Hot Spot Engineering for Quantitative SERS at the Single-Molecule Scale. J. Am. Chem. Soc. 2015, 137 (42), 13698–13705. (33) Vo-Dinh, T.; Hiromoto, M. Y. K.; Begun, G. M.; Moody, R. L. Surface-Enhanced Raman Spectrometry for Trace Organic Analysis. Anal. Chem. 1984, 56 (9), 1667–1670.

ACS Paragon Plus Environment

21

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 24

(34) He, D.; Hu, B.; Yao, Q.-F.; Wang, K.; Yu, S.-H. Large-Scale Synthesis of Flexible FreeStanding SERS Substrates with High Sensitivity: Electrospun PVA Nanofibers Embedded with Controlled Alignment of Silver Nanoparticles. ACS Nano 2009, 3 (12), 3993–4002. (35) Xu, C.; Wang, X. Fabrication of Flexible Metal-Nanoparticle Films Using Graphene Oxide Sheets as Substrates. Small 2009, 5 (19), 2212–2217. (36) Parolo, C.; Merkoçi, A. Paper-Based Nanobiosensors for Diagnostics. Chem. Soc. Rev. 2012, 42 (2), 450–457. (37) Nie, Z.; Nijhuis, C. A.; Gong, J.; Chen, X.; Kumachev, A.; Martinez, A. W.; Narovlyansky, M.; Whitesides, G. M. Electrochemical Sensing in Paper-Based Microfluidic Devices. Lab. Chip 2010, 10 (4), 477–483. (38) Tian, L.; Morrissey, J. J.; Kattumenu, R.; Gandra, N.; Kharasch, E. D.; Singamaneni, S. Bioplasmonic Paper as a Platform for Detection of Kidney Cancer Biomarkers. Anal. Chem. 2012, 84 (22), 9928–9934. (39) Lee, C. H.; Tian, L.; Singamaneni, S. Paper-Based SERS Swab for Rapid Trace Detection on Real-World Surfaces. ACS Appl. Mater. Interfaces 2010, 2 (12), 3429–3435. (40) Tian, L.; Tadepalli, S.; Hyun Park, S.; Liu, K.-K.; Morrissey, J. J.; Kharasch, E. D.; Naik, R. R.; Singamaneni, S. Bioplasmonic Calligraphy for Multiplexed Label-Free Biodetection. Biosens. Bioelectron. 2014, 59 (Supplement C), 208–215. (41) Tian, L.; Tadepalli, S.; E. Farrell, M.; Liu, K.-K.; Gandra, N.; M. Pellegrino, P.; Singamaneni, S. Multiplexed Charge-Selective Surface Enhanced Raman Scattering Based on Plasmonic Calligraphy. J. Mater. Chem. C 2014, 2 (27), 5438–5446. (42) Polavarapu, L.; Porta, A. L.; Novikov, S. M.; Coronado‐Puchau, M.; Liz‐Marzán, L. M. SERS: Pen‐on‐Paper Approach Toward the Design of Universal Surface Enhanced Raman Scattering Substrates (Small 15/2014). Small 2014, 10 (15), 3064–3064. (43) Lee, C. H.; Hankus, M. E.; Tian, L.; Pellegrino, P. M.; Singamaneni, S. Highly Sensitive Surface Enhanced Raman Scattering Substrates Based on Filter Paper Loaded with Plasmonic Nanostructures. Anal. Chem. 2011, 83 (23), 8953–8958. (44) Abbas, A.; Tian, L.; Morrissey, J. J.; Kharasch, E. D.; Singamaneni, S. Hot Spot-Localized Artificial Antibodies for Label-Free Plasmonic Biosensing. Adv. Funct. Mater. 2013, 23 (14), 1789–1797. (45) Hoppmann, E. P.; Yu, W. W.; White, I. M. Highly Sensitive and Flexible Inkjet Printed SERS Sensors on Paper. Methods 2013, 63 (3), 219–224. (46) Yu, W. W.; White, I. M. Inkjet Printed Surface Enhanced Raman Spectroscopy Array on Cellulose Paper. Anal. Chem. 2010, 82 (23), 9626–9630. (47) Huang, X.; Neretina, S.; El-Sayed, M. A. Gold Nanorods: From Synthesis and Properties to Biological and Biomedical Applications. Adv. Mater. 2009, 21 (48), 4880–4910. (48) Nusz, G. J.; Curry, A. C.; Marinakos, S. M.; Wax, A.; Chilkoti, A. Rational Selection of Gold Nanorod Geometry for Label-Free Plasmonic Biosensors. ACS Nano 2009, 3 (4), 795– 806. (49) Tian, L.; Chen, E.; Gandra, N.; Abbas, A.; Singamaneni, S. Gold Nanorods as Plasmonic Nanotransducers: Distance-Dependent Refractive Index Sensitivity. Langmuir 2012, 28 (50), 17435–17442. (50) Mayer, K. M.; Hafner, J. H. Localized Surface Plasmon Resonance Sensors. Chem. Rev. 2011, 111 (6), 3828–3857.

ACS Paragon Plus Environment

22

Page 23 of 24 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Nano Materials

(51) Vosgröne, T.; Meixner, A. J. Surface- and Resonance-Enhanced Micro-Raman Spectroscopy of Xanthene Dyes: From the Ensemble to Single Molecules. ChemPhysChem 2005, 6 (1), 154–163. (52) Si, M. Z.; Kang, Y. P.; Zhang, Z. G. Surface-Enhanced Raman Scattering (SERS) Spectra of Methyl Orange in Ag Colloids Prepared by Electrolysis Method. Appl. Surf. Sci. 2009, 255 (11), 6007–6010. (53) Zhang, D.; Vangala, K.; Jiang, D.; Zou, S.; Pechan, T. Drop Coating Deposition Raman Spectroscopy of Fluorescein Isothiocyanate Labeled Protein. Appl. Spectrosc. 2010, 64 (10), 1078–1085. (54) Pastoriza-Santos, I.; Pérez-Juste, J.; Liz-Marzán, L. M. Silica-Coating and Hydrophobation of CTAB-Stabilized Gold Nanorods. Chem. Mater. 2006, 18 (10), 2465–2467.

ACS Paragon Plus Environment

23

ACS Applied Nano Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 24

For Table of Contents Only

ACS Paragon Plus Environment

24