• Aucun résultat trouvé

Colloidal approach to Au-loaded TiO₂ thin films with optimized optical sensing properties

N/A
N/A
Protected

Academic year: 2021

Partager "Colloidal approach to Au-loaded TiO₂ thin films with optimized optical sensing properties"

Copied!
9
0
0

Texte intégral

(1)

Publisher’s version / Version de l'éditeur:

Journal of Materials Chemistry, 21, 12, pp. 4293-4300, 2011-02-14

READ THESE TERMS AND CONDITIONS CAREFULLY BEFORE USING THIS WEBSITE. https://nrc-publications.canada.ca/eng/copyright

Vous avez des questions? Nous pouvons vous aider. Pour communiquer directement avec un auteur, consultez la première page de la revue dans laquelle son article a été publié afin de trouver ses coordonnées. Si vous n’arrivez pas à les repérer, communiquez avec nous à PublicationsArchive-ArchivesPublications@nrc-cnrc.gc.ca.

Questions? Contact the NRC Publications Archive team at

PublicationsArchive-ArchivesPublications@nrc-cnrc.gc.ca. If you wish to email the authors directly, please see the first page of the publication for their contact information.

NRC Publications Archive

Archives des publications du CNRC

This publication could be one of several versions: author’s original, accepted manuscript or the publisher’s version. / La version de cette publication peut être l’une des suivantes : la version prépublication de l’auteur, la version acceptée du manuscrit ou la version de l’éditeur.

For the publisher’s version, please access the DOI link below./ Pour consulter la version de l’éditeur, utilisez le lien DOI ci-dessous.

https://doi.org/10.1039/c0jm03494k

Access and use of this website and the material on it are subject to the Terms and Conditions set forth at

Colloidal approach to Au-loaded TiO

₂ thin films with optimized optical

sensing properties

Gaspera, Enrico Della; Antonello, Alessandro; Guglielmi, Massimo; Post,

Michael L.; Bello, Valentina; Mattei, Giovanni; Romanato, Filippo

https://publications-cnrc.canada.ca/fra/droits

L’accès à ce site Web et l’utilisation de son contenu sont assujettis aux conditions présentées dans le site LISEZ CES CONDITIONS ATTENTIVEMENT AVANT D’UTILISER CE SITE WEB.

NRC Publications Record / Notice d'Archives des publications de CNRC:

https://nrc-publications.canada.ca/eng/view/object/?id=110a8682-6443-4ef0-a845-e7bedb4153e4

https://publications-cnrc.canada.ca/fra/voir/objet/?id=110a8682-6443-4ef0-a845-e7bedb4153e4

(2)

Colloidal approach to Au-loaded TiO

2

thin films with optimized optical

sensing properties†

Enrico Della Gaspera,‡

a

Alessandro Antonello,‡

b

Massimo Guglielmi,

a

Michael L. Post,

c

Valentina Bello,

d

Giovanni Mattei,

d

Filippo Romanato

e

and Alessandro Martucci*

b

Received 15th October 2010, Accepted 5th January 2011 DOI: 10.1039/c0jm03494k

TiO2and Au nanoparticles are synthesized by the colloidal technique and used for nanocomposite thin

film preparation. The effect of thermal treatment and organic template presence is analyzed in order to tailor film microstructure. The Au-TiO2interfaces as well as the overall porosity of the films are

analyzed combining structural and morphological characterization along with spectroscopic ellipsometry and surface plasmon spectroscopy analyses. An efficient surrounding of TiO2

nanoparticles around Au colloids is obtained, leading to an extensive noble metal–metal oxide interface, improving functional properties of the films, while keeping a porous structure. Gas sensing tests are performed on these nanocomposites films: reversible sensing dynamics for CO detection are observed with high sensitivity and a correlation between response and recovery times and

microstructure is reported.

Introduction

Semiconductor and dielectric matrixes containing noble metal nanoparticles (NPs) have been deeply investigated in the last few years, due to their unique properties and potential applications in different scientific fields, such as catalysis and photocatalysis,1–6

gas sensing,7–10optics and optoelectronics.11–15

In particular, Au NPs have attracted a lot of interest for their strong optical absorption in the visible range due to localized surface plasmon resonance, their great stability in a wide range of temperatures and environments, and their chemical activity.

Materials properties at the nanoscale differ from those in the bulk, and they are strongly affected by particle size and shape. Accurate control over such parameters is important for the

achievement of desired functionalities. Moreover, when dealing with nanocomposite fabrication, also the control of the overall microstructure is of paramount importance. This includes homogeneous dispersion of the particles inside the matrix, maintaining the interface between different phases and proper porosity, all these being required if the material has to interact with the surrounding environment.

Different methods to create dielectric-metal nanocomposites are available, such as sputtering,16,17electrochemical reactions,18

electron beam deposition,19

spray pyrolysis,20 evaporation,21 ion implantation,22 CVD23 and sol–gel.24–26

These techniques usually suffer from poor control over physical and morphological properties and often size tuning and ensuring particle mono-dispersivity is difficult to achieve.

Wet chemical synthesis of metal,27–29 semiconductor30,31 or

oxide32–34

colloids provides a useful tool for obtaining high quality NPs with desired size, shape and dispersion. Colloidal methods also allow the particle surface chemistry to be modified, which greatly extends the processing strategies and stability in the desired host matrix.

In this paper we report a study on Au-TiO2nanocomposite

thin films. Crystallinity of both components and control of microstructure are desired in many applications such as gas sensing, but generally the temperature required to induce a crystal order is high and so proper tailoring of nanocomposite structure is limited or even prevented.

For this reason in the present work, both matrix and dispersed materials are prepared by colloidal techniques obtaining suspensions of crystalline particles, and thin films are prepared by spinning directly a mixture of these suspensions, after appropriate washing and surface modification.

aINSTM and Dipartimento di Ingegneria Meccanica Settore Materiali,

Universita di Padova, Via Marzolo 9, 35131 Padova, Italy. E-mail: alex. martucci@unipd.it

bIOM-CNR, INSTM and Dipartimento di Ingegneria Meccanica Settore

Materiali, Universita di Padova, Via Marzolo 9, 35131 Padova, Italy

cInstitute for Chemical Process and Environmental Technology, National

Research Council of Canada, 1200 Montreal Road, Ottawa, Ontario, K1A 0R6, Canada

dDipartimento di Fisica, Universita di Padova, Via Marzolo 8, 35131

Padova, Italy

eIOM-CNR and Dipartimento di Fisica, Universita di Padova, Via

Marzolo 8, 35131 Padova, Italy

† Electronic supplementary information (ESI) available: detailed description of the experimental setup for ellipsometric analyses in different environments. Refractive index dispersion curves for TG4 and TGP4 samples measured in air and in ethanol. Estimated porosity with different effective medium approximation models. See DOI: 10.1039/c0jm03494k

‡ These authors contributed equally to this work.

Materials Chemistry

Cite this: J. Mater. Chem., 2011, 21, 4293

www.rsc.org/materials

PAPER

Downloaded by University of Melbourne on 04 April 2011

(3)

The film porosity and the structural evolution of the colloidal mixture is addressed and tailored by varying the thermal annealing temperature and using an organic template. Optical and morphological characterizations are then discussed and analyzed in order to understand structural evolution and the particle–matrix interface; these data are finally used to explain gas sensing results.

Experimental

TiO2nanoparticles synthesis

TiO2NPs were synthesized as described previously.35In a typical

synthesis, 3 g of titanium isopropoxide was added drop wise to a previously prepared solution consisting of 1.78 g of hydro-chloric acid, 7.18 mL of methanol and 1.24 g of water. The solution was stirred at ambient temperature for one hour and subsequently heated to 70C for four hours. Particles were then

precipitated with excess acetone and centrifuged at 4000 rpm for two minutes. The obtained precipitate was dispersed in minimum amount of methanol obtaining a clear colloidal sol of anatase TiO2NPs.

Au nanoparticles synthesis

Gold colloids were prepared according to the Turkevich method36by reducing HAuCl

4with trisodium citrate in water.

Briefly, 12 mL of 1% trisodium citrate aqueous solution was added to a 200 mL boiling solution of 0.5 mM HAuCl4. After the

solution turned red-wine colored, it was stirred at boiling point for an additional 15 min and then was cooled down to room temperature; 10.000 g mol 1 average molecular weight

poly-(N-vinylpyrrolidone) (PVP) was dissolved in water to yield a 50 g L 1 concentration, and then this solution was mixed with

aqueous gold colloids under constant stirring according to the ratio gPVP/molAu ¼ 1000. After 2 h the solution was concen-trated in a rotary evaporator and Au NPs were precipitated with acetone, centrifuged at 4000 rpm for 5 min and re-dispersed in ethanol leading to a 30 mM concentrated sol.

Nanocomposite thin films synthesis

The starting solutions were prepared by mixing the methanol suspension of TiO2 NPs with PVP-capped Au NPs in ethanol

(TG samples series) leading to a final TiO2concentration of 33 g

L 1and an Au : Ti molar ratio of about 5%.

200 g mol 1 average molecular weight polyethylene glycol

(PEG) was used as template material for porosity tailoring. In this case, PEG was directly added to the same solutions used for TG samples with a Ti : PEG ¼ 1.5 : 1 molar ratio, (TGP samples series). Au-free samples were prepared for comparison purposes by substituting the Au ethanol suspensions with pure ethanol, obtaining the T (without PEG) and TP (with PEG) sample series. All the film samples were deposited by spin coating at 2000 rpm for 30 s on either SiO2 or Si substrates and annealed in

a muffle furnace at 300 

C, 400 

C or 500

C for 1 h in air. Annealing temperature is indicated with 3, 4, 5 respectively, at the end of the sample name.

Final thicknesses are all in the 35–55 nm range, as measured with spectroscopic ellipsometry.

Characterization techniques

Films were characterized by X-ray diffraction (XRD) by using a Philips diffractometer equipped with glancing-incidence X-ray optics. The analysis was performed at 0.5incidence using

Cu-Ka Ni filtered radiation at 30 kV and 40 mA.

Surface morphology of the samples was investigated with an FEI Nova i600 dual beam scanning electron microscope (SEM). High resolution transmission dlectron microscopy (HR-TEM) cross-section measurements were taken with a field emission FEI TECNAI F20 SuperTwin FEG-(S)TEM microscope operating at 200 kV and equipped with an energy-dispersive X-ray (EDX) spectrometer for compositional analysis and a Gatan 794 Multiple Scan Camera, allowing digital image recording on a 1024  1024 pixel CCD array.

Optical absorption spectra of samples were measured in the 250–900 nm range using a Jasco V-570 standard spectropho-tometer.

Transmittance at normal incidence and ellipsometry quantities Jand D have been measured using a J.A. Woollam V-VASE Spectroscopic Ellipsometer in vertical configuration, at two different angles of incidence (65, 75) in the wavelength range

300–1500 nm. Optical constants n and k and film thickness have been evaluated from J, D and transmittance data using WVASE32 ellipsometry data analysis software, fitting the experimental data with Cauchy dispersion, Gaussian and Tauc-Lorentz oscillators for the non absorbing region, Au surface plasmon resonance (SPR) peak and UV absorption edge, respectively. Open porosity measurements have been performed measuring the optical constants under two different environ-ments (air and ethanol) with a custom built cell. Details are reported in the Supplementary Information section.†

Optical gas sensing tests were performed by making optical absorbance/transmittance measurements over the wavelength range 350 nm < l < 800 nm with sample films mounted on a heater inside a custom-built gas flow cell coupled with a Varian Cary1E spectrophotometer. Films heated at temperatures between 250 C and 350 C were exposed to four different

concentrations (10 ppm, 100 ppm, 1000 ppm and 10000 ppm, i.e. 1% v/v) of CO, all balanced in dry air, at a flow rate of about 0.4 L min 1. Tests were performed at 3 different operating

temperatures (OT ¼ 250

C, 300

C, 350

C). The substrate size was approximately 1 cm  2 cm and the incident spectropho-tometer beam was normal to the film surface and covering a 6 mm 1.5 mm section area.

Results and discussion

Structural and optical properties

TiO2 NPs adopted in this work have been characterized

else-where35 and shown to have the anatase crystal structure with

a mean diameter of 4  1.5 nm, while gold colloids were synthesized with a mean diameter of 14.1  0.9 nm, as shown in the TEM image of Fig. 1.

XRD measurements performed on thin films (Fig. 2) confirm the crystalline anatase phase of TiO2[JCPDS No. 841285] and

show also Au peaks [JCPDS No. 040714] for the TG and TGP series. Thermal treatment affects the XRD peak width, which is related to crystallite size. Mean dimensions evaluated by

Downloaded by University of Melbourne on 04 April 2011

Published on 14 February 2011 on http://pubs.rsc.org | doi:10.1039/C0JM03494K

(4)

applying the Scherrer equation, are reported in Table 1 for both the crystalline phases. TiO2NPs grow in size at higher

temper-atures, due to coarsening of the particles, while estimated dimension at 300 C annealing is comparable with the

as-synthesized colloids. In PEG containing samples, the TiO2

particle diameter has been estimated to be slightly smaller as compared with PEG-free samples at the same annealing temperature. It is likely that the porosity introduced by thermal decomposition of PEG, as will be discussed later, slows down the growth processes of titania particles. From these data, it is also shown that gold peaks do not undergo any relevant change in shape and intensity, indicating good Au NPs stability up to 500

C annealing.

Micrographs of PEG-free samples annealed at 300 C and

500 C are reported in Fig. 3. Au NPs of spherical shape are

homogeneously dispersed on a micron scale, and only very few small aggregates are evident: so almost no migration or major aggregation phenomena occur during the deposition and annealing processes. Moreover the 500C treated sample shows

higher surface roughness, as a result of the higher crystallinity and the larger crystal size.

The presence of Au NPs is also evident in the cross-section TEM image (Fig. 4a) of the TG5 sample, where gold NPs appear as dark features, due to their higher mass-thickness contrast and diffrac-tion contrast, with respect to TiO2. A good dispersion of metal

particles inside the oxide matrix can be recognized, even if the presence of some small aggregates is also evident from this analysis. The HR-TEM picture in Fig. 4b shows the crystallinity of both TiO2 and Au particles and lattice fringes of both phases are

evident. The lattice distances retrieved by fast Fourier transform (FFT) of TEM data are found to be in agreement with the anatase and gold crystal structures, as confirmed by XRD data. EDX compositional analysis performed during TEM measure-ment also confirms the two phases detected and gives an Au/Ti atomic ratio of 0.06, very close to the nominal value.

In Fig. 4b and 4c a close facing of TiO2NPs on Au NPs is also

observed, testifying a well developed Au/TiO2 interface in this

sample.

In Fig. 5, UV–Vis spectra of particles in solution are reported. TiO2NPs show a typical absorption edge in the UV range and

transparency in the visible, while a SPR peak at about 525 nm is shown for gold colloids in ethanol. No scattering features are detected for both solutions, implying good colloidal stability and

Fig. 1 TEM micrograph of PVP-protected gold colloids used for film synthesis. The particle size distribution histogram is shown in the inset.

Fig. 2 XRD spectra for: a) TiO2(black line) and Au-TiO2(grey line) films without PEG. b) TiO2(black line) and Au-TiO2films (grey line) with PEG. Theoretical diffraction lines for TiO2-anatase (black line) and Au (grey line) are reported at the bottom.

Table 1 Crystal size values of TiO2and Au particles evaluated from the Scherrer equation, using the full width at half maximum (FWHM) values of the diffraction peaks at 2q ¼ 25.3for anatase and at 38.2for gold

T/

C

Diameter (nm)

T series TG series TP series TGP series

TiO2 Au TiO2 Au TiO2 Au TiO2 Au

300 5.1 / 4.6 5.6 4.2 / 4 5.3 400 7.4 / 7.3 5.9 6.3 / 7.1 5.8 500 10.1 / 10.2 6.1 8.6 / 9.1 6.1

Downloaded by University of Melbourne on 04 April 2011

(5)

absence of aggregation phenomena, which is fundamental for composite materials to be used in optical applications.

Fig. 6 reports the absorption spectra of the Au-TiO2

nano-composite films with (Fig. 6b) and without (fig.6a) PEG. The SPR peak of gold particles is clearly visible, and a progressive red-shift with increasing annealing temperature can be observed, for both the PEG-free and PEG-containing series.

It is well known that the SPR frequency is strongly related to NPs morphology (size and shape), to the dielectric properties of the material surrounding them and to the relative proximity of the particles.37

From structural analyses no substantial morphology changes of Au particles can be observed, while a very adherent coating of TiO2nanocrystals on Au particles is

evident from TEM data. Thus we mainly attribute the change in plasmon peak position to the refractive index variation of the surrounding matrix. In each sample series, SPR wavelength is progressively red-shifted with increasing temperature, and this result is associated with the increase in refractive index of the material surrounding the gold particles.

This situation is confirmed from the results of the ellipsometry measurements. The refractive index curves of gold-loaded composite films are also shown in Fig. 6 (c and d); Au NPs influence is evident, causing a typical anomalous dispersion curve in the SPR absorption region, because of the Kramers–Kroning relationship involving refractive index and absorption coeffi-cient.

Refractive index values at 1100 nm are shown in Table 2. Values at that wavelength are here chosen as reference because they are sufficiently separated from the SPR absorption band to be modeled with a simple Cauchy dispersion model.

Refractive index is observed to increase for both PEG-free and PEG-modified samples with increasing annealing temperature. These data can be used to extract information about average film densification and porosity, but they do not take into account local structural variation. Differences in refractive index can be interpreted as differences in the amount of porosity by means of the effective medium optical models, where the most densified material has the highest refractive index. For this study the

Fig. 3 SEM images of TG3 (a) and TG5 (b) samples. Bright spots correspond to gold nanoparticles.

Fig. 4 Cross section bright field TEM image of the TG5 sample (a), HRTEM image of one Au particle surrounded by TiO2particles (b) and magnification the Au particle zone of Fig. 4b (c). Au NPs appear as darker spots compared to the lighter oxide matrix. Lattice planes and corresponding lattice parameters for TiO2and Au are shown in Fig. 4b.

Fig. 5 UV–Vis absorption spectra of TiO2 (a) and Au (b) colloidal solutions.

Fig. 6 Effect of annealing temperature on UV–Vis absorption spectra of TG (a) and TGP (b) films, and on refractive index dispersion curves of TG (c) and TGP (d) films.

Downloaded by University of Melbourne on 04 April 2011

Published on 14 February 2011 on http://pubs.rsc.org | doi:10.1039/C0JM03494K

(6)

Bruggemann relationship was adopted,38 and the estimated

porosity evaluated with this model can be written as follows:

Porosity ¼ n2 Tio2 n 2 n2 Tio2þ2 n 2  n2 Tio2 n 2 n2 Tio2þ2 n 2   1 n2 1 þ 2 n2 100%

where nTiO2is the refractive index of the TiO2matrix and n the

experimental refractive index of the film. Table 2 reports the porosity obtained with this formula, where the refractive index of the TiO2 anatase framework has been evaluated by measuring

the optical constants of the same sample in two different media, as reported by Cohen and coworkers39(see the Supplementary

Information for details†): the obtained value at 1100 nm is 2.408, slightly lower compared to the value at the same wavelength reported in the literature (2.44).40,41

The refractive index differ-ence between 400

C and 300

C annealed samples is around 0.04, while the average difference in refractive index between 500

C and 400C is around 0.12; a large increment of n occurs between

400C and 500C, as can be clearly seen from Fig. 7, indicating

that the densification process of TiO2 colloids is activated at

higher temperatures, affecting global porosity of the matrix.35

Moreover, the refractive index of PEG-modified samples is lower as compared with PEG-free samples with a constant difference of 0.14 for the three couples of samples, indicating, as expected, that PEG causes an increase in porosity after its

thermal decomposition (300 

C is sufficient for almost all PEG200 decomposition42). Similar evolution of porosity and

refractive index with temperature is detected in Au-free samples (not reported).

Interestingly, optical data can be used also for a deeper appreciation of the local structure of composite film. SPR frequencies are in fact related to local dielectric properties of the matrix around the Au particles and the theoretical refractive index calculated from the SPR peak using the Mie relationship43

is higher than the experimental refractive index of both the PEG-free and PEG-modified TiO2 matrices. The Au-TiO2

nano-composite materials synthesized from colloidal solutions behave differently in comparison with our previous results on TiO2sol–

gel thin films containing Au NPs,44 where the experimental

refractive index is slightly higher than the theoretical value, maybe because of a higher amount of amorphous matrix or porosity adjacent to the Au crystals, compared to the average value of the sample, that can affect the position of the Au SPR peak.

In this work, the optical results suggest that the material close to the gold surface has a higher refractive index as compared to the average value of the films, and thus a higher degree of densification of titania on gold particles can be inferred by these data.

These observations are summarized in Fig. 7, where SPR peak position of Au-TiO2samples associated with refractive index of

TiO2 film measured at the corresponding SPR wavelength is

reported together with the theoretical refractive index calculated from SPR peaks by means of the Mie relationship.

Moreover, increasing annealing temperature leads to further coverage of Au NPs, because the difference between theoretical and experimental refractive index becomes higher: thus, it is likely that in the TG3 sample, TiO2NPs are in poor contact with

gold, while in TG4 and TG5 the coverage of Au becomes higher, while the global matrix densification is not complete, so leaving residual porosity.

This conclusion is also partially supported by HRTEM anal-ysis of the TG5 sample presented in Fig. 4, where a close facing of TiO2 on Au particles is seen, even if no indication about the

densification away from the Au particles can be retrieved from those images.

Such a mechanism of structural evolution is also suggested from the fact that the observed plasmon peaks broaden with increasing temperature (see Fig. 6). Plasmon peak broadening could be partially associated to Au particles aggregation, even if the observed behavior can be better explained by the interaction between Au NPs and anatase surfaces.45,46 This effect would

agree with the increasing peak broadening with annealing temperature in TG samples, since the interface area gets larger. In the TGP series samples, the structural evolution is delayed with respect to thermal treatment. Comparing TGP3 and TGP4 samples, the SPR wavelength only slightly increases while no broadening is observed, suggesting that at 400C the contact

between Au and TiO2is not well developed. At 500C instead,

a more pronounced red shift of the plasmon frequency, together with broadening of the peak is observed.

To further understand the role of PEG, TG3 and TGP4 samples or alternatively TG4 and TGP5 samples can be compared: these couples of samples show similar SPR

Table 2 Refractive index and porosity of Au-loaded samples

Sample Refractive index @ 1100 nm Porosity (%) TG3 2.117 20 TGP3 1.973 30 TG4 2.154 18 TGP4 2.015 27 TG5 2.277 9 TGP5 2.137 19

Fig. 7 SPR peak position wavelengths versus refractive index of TiO2 film measured at the same wavelength for TGP (squares), and TG (triangles) series. Theoretical refractive index values associated to the plasmon peak position according to the Mie relationship are also reported (circles).

Downloaded by University of Melbourne on 04 April 2011

(7)

wavelengths, but refractive index and annealing temperature are different. In PEG-containing films, a higher thermal treatment is necessary in order to have the same SPR frequencies than PEG-free samples. Thus, following the arguments previously explained, a similar Au coverage by TiO2NPs is suggested for the

two couples of samples mentioned before, while if the annealing temperature is kept constant, the extent of TiO2/Au interface is

limited in PEG containing samples. So, the presence of PEG not only introduces porosity but also affects the evolution of the colloids in the mixture interphase region.

In summary, the extent by which gold contacts titania is not strictly dependent on the overall degree of densification of the matrix, since samples with analogous average refractive index show very different SPR positions, and films with similar plas-mon resonance frequencies have different average refractive index values. Furthermore, all the samples show an experimental refractive index much lower than the theoretical value predicted by Mie theory. So, efficient coverage of TiO2 on gold is here

addressed as the key theme for experimental data interpretation: a combination of annealing temperature and template presence is thus responsible for the amount of titania facing the gold surface.

Gas sensing

Au-TiO2nanocomposite films were tested as optical gas sensors

for CO detection. Samples annealed at 300C were tested only at

250

C operating temperature (OT) to avoid structural changes in the film during gas sensing measurements, but they show unstable baseline. This may arise from residual carbonaceous compounds still present inside the film or from incomplete stabilization after the 300C annealing treatment. In fact

ther-mogravimetric analyses performed on TiO2 colloids35 and on

PEG20042

show that the weight loss is not totally completed at 300 

C. Thus, only the results of 400 

C and 500 

C treated samples will be presented.

All samples gave a response to CO, with a typical behavior of semiconductor thin films containing Au NPs,47

which has a wavelength dependence of optical sensitivity, allowing the sensor response tuning. The Au SPR peak during gas exposure undergoes a blue-shift. This effect can be related to the interac-tion of reducing gases with the n-type semiconductor TiO2

matrix, promoting a catalytic oxidation of the gas and an increase in the amount of conducting electrons, which in turn leads to a shift of the plasmon band to higher frequencies. However, a shift of Au SPR band can be also due to adsorption of the target gases, leading to a local change in the refractive index. A CO adsorption would lead to an increase of average refractive index and hence to a red shift of the Au SPR peak, differently from what we observed. Moreover previous studies on electrical conductivity of similar samples48

showed a decrease in resistivity during reducing gases exposure, so the observed SPR blue shift has to be related to the oxidation of CO, resulting in an injection of electrons inside the Au-TiO2 nanocomposite thin

films. It is noteworthy that the catalytic oxidation of CO over the Au-TiO2interface resulting in the formation of CO2is a

well-known topic in literature, even if the great majority of the reports deal with CO oxidation at low temperature (below 400 K).49–54

By plotting the difference in optical absorption between air and target gas exposure versus wavelength, an absolute

maximum and minimum can be determined (details have been reported previously48), with wavelengths l

max and lmin

corre-sponding to absolute maximum and minimum, respectively. All the samples were tested dynamically at both lmaxand lmin, for

each of the three temperatures: Table 3 lists these two wave-lengths for all the tested samples. Fig. 8 shows the temporal response of the four tested samples during 1% CO exposure. The dynamic performance of the sensors is quite good, with an almost square output signal, especially for PEG-free samples. This behavior represents an improvement with respect to our previous results48

where the optical sensing tests of sol–gel TiO2

matrix loaded with Au NPs show a reversible but sometimes incomplete response, and also a drift in the baseline.

The wavelength influence is also evident from the data reported in Fig. 8: operating near lmin causes an increase in

absorption during gas exposure, while operating near lmaxresults

in a decrease of absorbance during gas exposure.

Although these Au-TiO2nanocomposite films are very thin,

they reveal the capability to detect very low amount of the target gas with high sensitivity, as highlighted in Fig. 9: tests with CO concentration ranging from 10 ppm to 10000 ppm were per-formed on the TGP4 sample at 300

C OT. The sample also gives a response to 10 ppm CO, with response intensity decreasing linearly with the order of magnitude of target gas concentration, thus showing higher sensitivity at very low concentrations. Similar responses and dynamic behaviors have been observed in all the other samples. A useful parameter to analyze the sensi-tivity of the sample is the Response Intensity (RI) that can be defined as follows: RI ¼ 1 Absco AbsAir

Fig. 9b shows the RI plot for TGP4 sample as a function of the CO concentration, in a linear-logarithmic scale: as can be seen the trend is linear, allowing to set easily the calibration curves for the sensor performances. Moreover the linear fit also permits the estimation of the sensitivity threshold, being around 2–3 ppm, considering the noise in optical absorption evaluated from the dynamic test reported in Fig. 9a as the lower limit of detection. The relationship between the nanocomposite structure (porosity and interface between Au and TiO2crystals) and its

optical gas sensing is addressed by analyzing the dynamic behavior in terms of response and recovery times, mainly focusing on recovery performance. The times to reach 90% response and 90% recovery have been evaluated from the three dynamic tests reported in Fig. 9. Each result has been averaged

Table 3 List of the wavelengths corresponding to absolute maximum and minimum of OAC curves

Sample lmax (nm) lmin (nm) TG4 600 700 TG5 620 720 TGP4 585 670 TGP5 600 690

Downloaded by University of Melbourne on 04 April 2011

Published on 14 February 2011 on http://pubs.rsc.org | doi:10.1039/C0JM03494K

(8)

on three different measures to prevent anomalous or singular data; the error is about 8%, calculated as signal to noise ratio. Response time is short for every sample at every temperature,

ranging from 20 to 35 s. The comparison of such data between the different samples is quite difficult, being the values very similar. However, more useful information can be gained comparing the average recovery times, since values are higher and the error in the measurement becomes less significant. Fig. 10 shows the 90% recovery times for the 4 samples: TG5 is the best sample while the two PEG-containing samples show the longest recovery times. Thus, it is evident that for optical gas sensing purposes, it is better to have a completely crystalline TiO2matrix

with an extensive gold-anatase contact surface (TG5), rather than a more porous sample with a lower amount of Au-TiO2

interface (PEG series). This behavior can be due to the effect of PEG in slowing the growth of TiO2crystals and limiting the

Au-TiO2interface, as previously discussed when comparing XRD

data and optical characterization. In addition, baseline recovery for PEG-modified samples is sometimes incomplete, suggesting the presence of residual carbon due to incomplete removal of the organic compound, which can react with the target gas affecting negatively the measurement.

Conclusions

An all-colloidal approach has been used to synthesize Au-TiO2

nanocomposite thin films: both TiO2-anatase and gold NPs were

synthesized with colloidal techniques, purified, concentrated and subsequently used to prepare thin films. Structural and morphological characterizations after different annealing temperatures have been presented, confirming the crystallinity of the nanocomposites and uniform dispersion of Au NPs inside the oxide matrix. Combining ellipsometry and surface plasmon spectroscopy measurements, a low average refractive index together with highly red shifted SPR have been observed, sug-gesting an efficient coverage of TiO2crystals onto Au particles,

which is also supported by TEM investigation. Optical gas sensing tests reveal very high sensitivity towards carbon monoxide down to a few ppm. Nanocomposite films with an extensive gold-anatase contact surface showed better dynamic gas sensing behavior than nanocomposite film with higher porosity but with a lower amount of Au-TiO2interface.

Fig. 8 Dynamic responses of the four Au-TiO2nanocomposite films under exposure to Air–1% CO–Air cycle: a) test performed at lminand at 250C OT. b) Test performed at l

minand at 300C OT. c) Test per-formed at lmaxand at 350C OT.

Fig. 9 a) Dynamic response of the TGP4 sample at 585 nm and 300

C OT, during exposure to different CO concentrations (expressed in ppm). b) Sensitivity plot for the TGP4 sample measured at 585 nm: the dashed curve is the linear fit of the experimental data.

Fig. 10 Graph showing 90% average recovery times after 1% v/v CO exposure for the four samples.

Downloaded by University of Melbourne on 04 April 2011

(9)

Acknowledgements

This work has been supported trough Progetto Strategico PLATFORMS of Padova University and 7FP EU Project – ORION #222517. E.D.G. thanks Fondazione CARIPARO for financial support. This contribution has been assigned National Research Council of Canada number NRCC# 52225.

References

1 L. Gang, B. J. Anderson, J. Van Grondelle and R. A. Van Santen,

Appl. Catal., B, 2003, 40, 101.

2 W. S. Ju, M. Matsuoka, K. Iino, H. Yamashita and M. Anpo, J. Phys.

Chem. B, 2004, 108, 2128.

3 M. A. Aramendıa, A. Marinas, J. M. Marinas, J. M. Moreno and F. J. Urbano, Catal. Today, 2005, 101, 187.

4 A. Hagfeldt and M. Gratzel, Acc. Chem. Res., 2000, 33, 269. 5 J. He, I. Ichinose, T. Kunitake, A. Nakao, Y. Shiraishi and

N. Toshima, J. Am. Chem. Soc., 2003, 125, 11034.

6 X. Wang, D. R. G. Mitchell, K. Prince, A. J. Atanacio and R. A. Caruso, Chem. Mater., 2008, 20, 3917.

7 M. Ando, T. Kobayashi and M. Haruta, Catal. Today, 1997, 36, 135. 8 D. Kohl, Sens. Actuators, B, 1990, 1, 158.

9 A. M. Ruiz, A. Cornet, K. Shimanoe, J. R. Morante and N. Yamazoe,

Sens. Actuators, B, 2005, 108, 34.

10 A. Martucci, M. Pasquale, M. Guglielmi, J. C. Pivin and M. L. Post,

J. Am. Ceram. Soc., 2003, 86, 1638.

11 M. Hirasawa, H. Shirakawa, H. Hamamura, Y. Egashira and H. Komiyama, J. Appl. Phys., 1997, 82, 1404.

12 G. De, A. Licciulli, C. Massaro, L. Tapfer, M. Catalano, G. Battaglin, C. Meneghini and P. Mazzoldi, J. Non-Cryst. Solids, 1996, 194, 225. 13 S. Deki, Y. Aoi, H. Yanagimoto, K. Ishii, K. Akamatsu,

M. Mizuhata and A. Kajinami, J. Mater. Chem., 1996, 6, 1879. 14 T. Ung, L. M. Liz-Marzan and P. Mulvaney, J. Phys. Chem. B, 2001,

105, 3441.

15 H. B. Liao, W. J. Wen and J. K. L. Wong, J. Appl. Phys., 2003, 93, 4485.

16 M. Ando, H. Steffes, R. Chabicovsky, M. Haruta and G. Stangl,

IEEE Sens. J., 2004, 4, 232.

17 U. Pal, C. Garcia-Serrano, G. Casarubbias-Segura, N. Koshizaki, T. Sasaki and S. Terahuchi, Sol. Energy Mater. Sol. Cells, 2004, 81, 339. 18 P. Liu, S. Lee, C. E. Tracy, J. A. Turner, J. R. Pitts and S. K. Deb,

Solid State Ionics, 2003, 165, 223.

19 O. Berger, W. J. Fischer and V. Melev, J. Mater. Sci., 2004, 15, 463. 20 L. Madler, A. Roessler, S. E. Pratsinis, T. Sahm, A. Gurlo,

N. A. Barsan and U. Weimar, Sens. Actuators, B, 2006, 114, 283. 21 C. H. Liu, L. Zhang and Y. J. He, Thin Solid Films, 1997, 304, 13. 22 R. H. Magruder III, L. Yang, R. F. Haglund Jr., C. W. White,

L. Yang, R. Dorsinville and R. R. Alfano, Appl. Phys. Lett., 1993, 62, 1730.

23 M. Okumura, S. Nakamura, S. Tsubota, T. Nakamura, M. Azuma and M. Haruta, Catal. Lett., 1998, 51, 53.

24 J. Zhuou, L. Li, Z. Gui, X. Zhang and D. J. Barber, Nanostruct.

Mater., 1997, 8(3), 321.

25 H. Kozuka, G. Zhao and S. Sakka, Bull. Inst. Chem. Res. Kyoto Univ., 1994, 72, 209.

26 G. Zhao, H. Kozuka and S. Sakka, J. Sol-Gel Sci. Technol., 1995, 4, 37.

27 Y. Sun and Y. Xia, Science, 2002, 298, 2176.

28 Y. Sun, B. Gates, B. Mayers and Y. Xia, Nano Lett., 2002, 2, 165. 29 L. M. Liz-Marzan, Langmuir, 2006, 22, 32.

30 C. B. Murray, D. J. Norris and M. G. Bawendi, J. Am. Chem. Soc., 1993, 115, 8706.

31 J. Van Embden, J. Jasieniak and P. Mulvaney, J. Am. Chem. Soc., 2009, 131, 14299.

32 M. Niederberger, M. H. Bartl and G. D. Stucky, J. Am. Chem. Soc., 2002, 124, 13642.

33 N. R. Jana, Y. Chen and X. Peng, Chem. Mater., 2004, 16, 3931. 34 N. Niederberger and N. Pinna, Angew. Chem., Int. Ed., 2008, 47,

5292.

35 A. Antonello, G. Brusatin, M. Guglielmi, V. Bello, G. Mattei, G. Zacco and A. Martucci, J. Nanopart. Res., DOI: 10.1007/s11051-010-9923-4.

36 B. V. Enustun and Turkevich, J. Am. Chem. Soc., 1963, 85, 3317. 37 K. L. Kelly, E. Coronado, L. L. Zhao and G. C. J. Schatz, J. Phys.

Chem. B, 2003, 107, 668.

38 D. A. G. Bruggeman, Ann. Phys. (Leipzig), 1935, 24, 636.

39 D. Lee, M. F. Rubner and R. E. Cohen, Nano Lett., 2006, 6, 2305– 2312.

40 G. E. Jellison Jr, L. A. Boatner, J. D. Budai, B. S. Jeong and D. P. Norton, J. Appl. Phys., 2003, 93(12), 9537.

41 S. Tanemura, L. Miao, P. Jin, K. Kaneko, A. Terai and N. Nabatova-Gabain, Appl. Surf. Sci., 2003, 212–213, 654.

42 S. Grandi, A. Magistris, P. Mustarelli, E. Quartarone, C. Tomasi and L. Meda, J. Non-Cryst. Solids, 2006, 352, 273.

43 C. F. Bohren; D. R. Huffman. Absorption and Scattering of Light by

Small Particles, Wiley (New York), 1998.

44 D. Buso, J. Pacifico, A. Martucci and P. Mulvaney, Adv. Funct.

Mater., 2007, 17, 347.

45 P. Alemany, R. S. Borse, J. M. Burlitch and R. Hoffmann, J. Phys.

Chem., 1993, 97, 8464.

46 M. Lee, L. Chae and K. C. Lee, Nanostruct. Mater., 1999, 11, 195. 47 T. Kobayashi, M. Haruta and M. Ando, Sens. Actuators, B, 1993, 13–

14, 545.

48 D. Buso, M. L. Post, C. Cantalini, P. Mulvaney and A. Martucci,

Adv. Funct. Mater., 2008, 18, 3843.

49 J. Steyn, G. Pattrick, M. S. Scurrell, D. Hildebrandtd, M. C. Raphulub and E. Van Der Lingenb, Catal. Today, 2007, 122, 254.

50 M. M. Schubert, V. Plzak, J. Garche and R. J. Behm, Catal. Lett., 2001, 76, 143.

51 W. Y. Yu, C. P. Yang, J. N. Lin, C. N. Kuo and B. Z. Wan, Chem.

Commun., 2005, 354.

52 Y. Iizuka, T. Tode, T. Takao, K. Yatsu, T. Takeuchi, S. Tsubota and M. Haruta, J. Catal., 1999, 187, 50.

53 B. K. Chang, B. W. Jang, S. Dai and S. H. Overbury, J. Catal., 2005, 236, 392.

54 Y. Denkwitz, B. Schumacher, G. Kucerova and R. J. Behm, J. Catal., 2009, 267, 78.

Downloaded by University of Melbourne on 04 April 2011

Published on 14 February 2011 on http://pubs.rsc.org | doi:10.1039/C0JM03494K

Figure

Fig. 2 XRD spectra for: a) TiO 2 (black line) and Au-TiO 2 (grey line) films without PEG
Fig. 6 reports the absorption spectra of the Au-TiO 2 nano- nano-composite films with (Fig
Fig. 7 SPR peak position wavelengths versus refractive index of TiO 2
Fig. 9b shows the RI plot for TGP4 sample as a function of the CO concentration, in a linear-logarithmic scale: as can be seen the trend is linear, allowing to set easily the calibration curves for the sensor performances
+2

Références

Documents relatifs

We report a rapid and efficient procedure to functionalize SWNT where free radicals generated at room temperature by a redox reaction between reduced SWNT and diacyl

SnO 2 thin films were electrodeposited on fluorine tin oxide substrate in nitric acid solution. The films were found uniform, adherent to the substrate and

Our results show that the structure and the grains size of our layers change when one varies the temperature of treatment as well as the number of steeping and the concentration

XRD diagrams illustrate that undoped and erbium doped TiO 2 crystallize in anatase phase only and indicate that the crystallite size decreases from 24.06 to 21.19 nm as a function

To determine the nature of the limiting stage and the reaction mechanism on the electrode, we plot the polarization curves of pure zinc as a function of the scan rates fig

Characterizations of the samples were performed by X-ray diffraction (XRD), Atomic Force Microscopy (AFM) and the diffuse reflectance of the films was measured using

The scanning electron micrographs show that annealing under selenium pressure (group B sam- ples) introduces a change in films morphology. While in situ

Vibrating sample magnetometer results of un-doped TiO 2 thin films reveal a ferromagnetic behavior at room temperature with M s = 0.67 10-4 emu and low content of nickel doping