Microsensors by Igor V. Minin and Oleg V. Minin - HTML preview

PLEASE NOTE: This is an HTML preview only and some elements such as links or page numbers may be incorrect.
Download the book in PDF, ePub, Kindle for a complete version.

typically observed nanotubular layers) can simply be obtained by controlling the water

content in the electrolyte during the anodization process. It is proposed that the morphology

transition from pores to tubes is based on the rate of preferential etching at the hexagonal

cell triple points in the oxide.

Zhang prepared the highly ordered TiO2 thin films by anodic oxidation followed by

calcination at various temperatures (300, 400, 500 and 600 °C) (Zhang, 2008). The author

investigated the humidity sensing behaviours of prepared samples. The samples calcined at

600 °C showed high sensitivity with nearly two orders change in the resistance and short

response and recovery time (< 190 s) during the relative humidity variation from 11 to 95%.

Another method is the deposition of WO3 thin films on highly ordered nanoporous alumina

template. Nanoporous anodic oxide layers were formed by anodizing aluminum films in

malonic acid electrolyte. Tungsten trioxide sensing films were deposited on the top of

nanoporous alumina layers by rf magnetron sputtering of a metallic target (Fig. 7). The

tungsten oxide gas sensing structures supported by nanoporous alumina templates showed

high responsiveness to toxic gases, especially to NO2 (Gorokh, 2006; Khatko, 2009, 2006;

Vallejos, 2008).

index-167_1.png

index-167_2.png

Chemical Microsensors with Ordered Nanostructures

155

Fig. 7. SEM images of cross-fracture (left) and the surface (right) of alumina films with

sputtered WO3

4. Conclusion

Described non-litographic techniques are based on template-assisted method. The template preparation of thin film with highly ordered pores is a suitable way for nanostructured

material synthesis since they are cheap, fast and easy reproducible. Due to the special

properties arising from their behavior, these highly ordered nanostructures can find various

applications in environmental analysis as well as medicine and pharmacy.

In the case of environmental analysis application, the nanostructures are used to modify

either the sensing elements from the semiconducting materials of vapor and gas sensors or

the electrodes of electrochemical sensors.

Concerning the pharmacy and medicine, quantum dots (QDs) in planar form (so-called lab-

on-chip) deposited on various solid surfaces seems to be a new approach of template-based

method application. The sensor array created from separately deposited QDs, also called

“fluorescence array detector”, can be used for in-vitro large-field imaging. This allows the

easy detection of many different biomolecules at the same time, since each QD can emit the

light at different wavelength. Electrochemical biosensors with functionalized electrodes for

rapid detection and mass screening are very promising in near future in cases of pandemic

and epidemic. Cultivation of cells on gold nanodots has also high impact in biochemistry

research for medicine.

5. Acknowledgment

This work has been supported by Grant Agency of the Academy of Sciencies of the Czech

Republic under the contract GAAV KAN208130801 (NANOSEMED) and by Grant Agency

of the Czech Republic under the contract GACR 102/08/1546 (NANIMEL).

6. References

Alivov, Y. et al. (2009). Titanium nanotubes grown by titanium anodization. Journal of Applied

Physics, Vol. 106, No. 3, pp. 5, ISSN 0021-8979

Alonso-Gonzalez, P. et al. (2006). Ordered InAs QDs using prepatterned substrates by

monolithically integrated porous alumina. Journal of Crystal Growth, Vol. 294, No. 2,

pp. 168-173, ISSN 0022-0248

156

Microsensors

Anitha, V. C. et al. (2010). Electrochemical tuning of titania nanotube morphology in inhibitor

electrolytes. Electrochimica Acta, Vol. 55, No. 11, pp. 3703-3713, ISSN 0013-4686

Berger, S. et al. (2008). Formation of hexagonally ordered nanoporous anodic zirconia.

Electrochemistry Communications, Vol. 10, No. 12, pp. 1916-1919, ISSN 1388-2481

Cao, C. B. et al. (2010). Layer-by-Layer Growth Mechanism of TiO2 Nanotube Arrays. Journal

of the Electrochemical Society, Vol. 158, No. 1, pp. E8-E11, ISSN 0013-4651

Cao, G. Z. & Liu, D. W. (2008). Template-based synthesis of nanorod, nanowire, and

nanotube arrays. Advances in Colloid and Interface Science, Vol. 136, No. 1-2, pp. 45-

64, ISSN 0001-8686

Gorokh, G. et al. (2006). Anodic formation of low-aspect-ratio porous alumina films for metal-

oxide sensor application. Electrochimica Acta, Vol. 52, No. 4, pp. 1771-1780, ISSN 0013-

4686

Graham, A. H. D. et al. (2010). Nanostructured electrodes for biocompatible CMOS integrated

circuits. Sensors and Actuators B-Chemical, Vol. 147, No. 2, pp. 697-706, ISSN 0925-4005

Hahn, R. et al. (2010). Bright visible luminescence of self-organized ZrO2 nanotubes. Journal

of Solid State Electrochemistry, Vol. 14, No. 2, pp. 285-288, ISSN 1432-8488

Hassan, F. M. B. et al. (2009). Formation of Self-Ordered TiO2 Nanotubes by Electrochemical

Anodization of Titanium in 2-Propanol/NH4F. Journal of the Electrochemical Society,

Vol. 156, No. 12, pp. K227-K232, ISSN 0013-4651

Hrdý, R. & Hubálek, J. (2007). Using a Porous Alumina Film as a Mask for Formation of

Ordered Nanostructures by Deposition Technique, Acta Metallurgica Slovaca, Vol.

13, No. 2, pp. 155-158, ISSN 1335-1532

Hrdy, R. & Hubalek, J. (2005). Self ordered Pore Structure of Anodized Alumina Thin Film

on Si Substrate, Proceedings of Electronic Devices and Systems, pp. 300-304, ISBN 80-

214-2990-9, Brno, Czech Republic, September, 2005

Hu, M. Z. et al. (2009). Synthesis and characterization of anodized titanium-oxide nanotube

arrays. Journal of Materials Science, Vol. 44, No. 11, pp. 2820-2827, ISSN 0022-2461

Cha, Y. K. et al. (2004). Nonlithographic SiO2 nanodot arrays via template synthesis

approach. Japanese Journal of Applied Physics Part 1-Regular Papers Short Notes &

Review Papers, Vol. 43, No. 8A, pp. 5657-5659, ISSN 0021-4922

Chen, P. L. et al. (2005). Fabrication and field emission characteristics of highly ordered

titanium oxide nanodot arrays. Electrochemical and Solid State Letters, Vol. 8, No. 10,

pp. H83-H86, ISSN 1099-0062

Chen, P. L. et al. (2004). Preparation and phase transformation of highly ordered TiO2

nanodot arrays on sapphire substrates. Applied Physics Letters, Vol. 84, No. 19, pp.

3888-3890, ISSN 0003-6951

Chen, P. L. et al. (2003). Self-organized titanium oxide nanodot arrays by electrochemical

anodization. Applied Physics Letters, Vol. 82, No. 17, pp. 2796-2798, ISSN 0003-6951

Cho, S. J. et al. (2008). Titanium oxide nanotubes anodized in aqueous and non-aqueous

electrolytes. Journal of Ceramic Processing Research, Vol. 9, No. 5, pp. 449-451, 1229-9162

Chu, S. Z. et al. (2005). Self-organized nanoporous anodic titania films and ordered titania

nanodots/nanorods on glass. Advanced Functional Materials, Vol. 15, No. 8, pp. 1343-

1349, ISSN 1616-301X

Chu, S. Z. et al. (2005). A new electrochemical lithography - Fabrication of self-organized

titania nanostructures on glass by combined anodization. Journal of the

Electrochemical Society, Vol. 152, No. 3, pp. B116-B124, ISSN 0013-4651

Joo, S. et al. (2010). Hydrogen Gas Sensor Using Pt- and Pd-Added Anodic TiO2 Nanotube

Films. Journal of the Electrochemical Society, Vol. 157, No. 6, pp. J221-J226, ISSN

0013-4651

Chemical Microsensors with Ordered Nanostructures

157

Jung, J. S. et al. (2008). Electrodeposited Nickel Nanodots Array on the Silicon Wafer. Bulletin

of the Korean Chemical Society, Vol. 29, No. 11, pp. 2169-2171, ISSN 0253-2964

Jung, M. et al. (2006). Fabrication of the uniform CdTe quantum dot array on GaAs substrate

utilizing nanoporous alumina masks. Current Applied Physics, Vol. 6, No. 6, pp.

1016-1019, 1567-1739

Kemell, M. et al. (2007). Atomic layer deposition of nanostructured TiO2 photocatalysts via

template approach. Chemistry of Materials, Vol. 19, No. 7, pp. 1816-1820, ISSN 0897-4756

Khatko, V. et al. (2006). Tungsten trioxide sensing layers on highly ordered nanoporous

alumina template. Sensors and Actuators B-Chemical, Vol. 118, No. 1-2, pp. 255-262,

0925-4005

Khatko, V. et al. (2009). Micro-machined WO3-based sensors with improved characteristics.

Sensors and Actuators B-Chemical, Vol. 140, No. 2, pp. 356-362, ISSN 0925-4005

Klosová, K. & Hubálek, J. (2008). Advanced electrodes with nanostructured surfaces for

electrochemical microsensors, Physica Status Solidi, Vol. 205, No. 6, pp. 1435-1438,

ISSN 0031-8965

Klosova, K. et al. (2006). New Microelectrodes for Electrochemical Application with

Nanomachined Surface, Proceedings of the International Conference NANO´06, pp.

210-214, ISBN 80-214-3331-0, Brno, Czech Republic, November, 2006

Klosova, K. et al. (2006). New Approach to Electrochemical Sensor Electrodes Construction,

Proceedings of Junior Scientist Conference, pp. 139-140, ISBN 3-902463-05-8, Vienna,

Austria, April, 2006

Kokonou, M. et al. (2007). Few nanometer thick anodic porous alumina films on silicon with

high density of vertical pores. Thin Solid Films, Vol. 515, No. 7-8, pp. 3602-3606,

ISSN 0040-6090

Kouklin, N. et al. Capacitance-voltage spectroscopy of self assembled ordered arrays of

quantum dots. New York: Ieee, 2000.

Li, A. P. et al. (1998). Hexagonal pore arrays with a 50-420 nm interpore distance formed by

self-organization in anodic alumina. Journal of Applied Physics, Vol. 84, No. 11, pp.

6023-6026, ISSN 0021-8979

Li, C. P. et al. (2006). Fabrication and structural characterization of highly ordered sub-100-

nm planar magnetic nanodot arrays over 1 cm(2) coverage area. Journal of Applied

Physics, Vol. 100, No. 7, pp. 7, ISSN 0021-8979

Li, L. L. et al. (2010). Morphologic Characterization of Anodic Titania Nanotube Arrays for

Dye-Sensitized Solar Cells. Journal of the Chinese Chemical Society, Vol. 58, No. 5B,

pp. 1147-1150, ISSN 0009-4536

Liang, J. Y. et al. (2002). Two-dimensional lateral superlattices of nanostructures:

Nonlithographic formation by anodic membrane template. Journal of Applied

Physics, Vol. 91, No. 4, pp. 2544-2546, ISSN 0021-8979

Lim, J. H. et al. (2009). Electrochemical determination of whole blood clotting time by using

nanodot arrays. Electrochemistry Communications, Vol. 11, No. 11, pp. 2141-2144,

ISSN 1388-2481

Liu, Y. B. et al. (2009). Comparison of photoelectrochemical properties of TiO2-nanotube-

array photoanode prepared by anodization in different electrolyte. Environmental

Chemistry Letters, Vol. 7, No. 4, pp. 363-368, ISSN 1610-3653

Mao, R. et al. (2009). In situ preparation of an ultra-thin nanomask on a silicon wafer.

Nanotechnology, Vol. 20, No. 2, pp. 6, ISSN 0957-4484

Masuda, H. & Fukuda, K. (1995). Ordered metal nanohole arrays made by a 2-step

replication of honeycomb structures of anodic alumina. Science, Vol. 268, No. 5216,

pp. 1466-1468, ISSN 0036-8075

158

Microsensors

Masuda, H. et al. (1998). Self-ordering of cell configuration of anodic porous alumina with

large-size pores in phosphoric acid solution. Japanese Journal of Applied Physics Part

2-Letters, Vol. 37, No. 11A, pp. L1340-L1342,

Matefi-Tempfli, S. et al. (2009). Nanowires and nanostructures fabrication using template

methods: a step forward to real devices combining electrochemical synthesis with

lithographic techniques. Journal of Materials Science-Materials in Electronics, Vol. 20,

No., pp. 249-254, ISSN 0957-4522

Montero-Moreno, J. M. et al. (2009). Production of alumina templates suitable for

electrodeposition of nanostructures using stepped techniques. Electrochimica Acta,

Vol. 54, No. 9, pp. 2529-2535, ISSN 0013-4686

Mozalev, A. et al. (2009). Growth of multioxide planar film with the nanoscale inner

structure via anodizing Al/Ta layers on Si. Electrochimica Acta, Vol. 54, No. 3, pp.

935-945, ISSN 0013-4686

Mun, K. et al. (2010). A Stable, Label-free Optical Interferometric Biosensor Based on TiO2

Nanotube Arrays. Acs Nano, Vol. 4, No. 4, pp. 2070-2076, ISSN 1936-0851

Oide, A. et al. (2006). Fabrication of ordered nanostructure on silicon substrate using

localized anodization and chemical etching. Electrochemistry, Vol. 74, No. 5, pp. 379-

384, ISSN 1344-3542

Possin, G. E. (1970). A method for forming very small diameter wires. Review of Scientific

Instruments, Vol. 41, No. 5, pp. 772-&, ISSN 0034-6748

Sennik, E. et al. (2010). Synthesis of highly-ordered TiO2 nanotubes for a hydrogen sensor.

International Journal of Hydrogen Energy, Vol. 35, No. 9, pp. 4420-4427, ISSN 0360-3199

Shingubara, S. (2003). Fabrication of nanomaterials using porous alumina templates . Journal

of Nanoparticle Research, Vol. 5, No. 1-2, pp. 17-30, ISSN 1388-0764

Song, Y. Y. & Schmuki, P. (2010). Modulated TiO2 nanotube stacks and their use in

interference sensors. Electrochemistry Communications, Vol. 12, No. 4, pp. 579-582,

ISSN 1388-2481

Tan, L. K. et al. (2010). Transparent, Well-Aligned TiO2 Nanotube Arrays with Controllable

Dimensions on Glass Substrates for Photocatalytic Applications. Acs Applied

Materials & Interfaces, Vol. 2, No. 2, pp. 498-503, ISSN 1944-8244

Vallejos, S. et al. (2008). Micro-machined WO3-based sensors selective to oxidizing gases.

Sensors and Actuators B-Chemical, Vol. 132, No. 1, pp. 209-215, ISSN 0925-4005

Vorozhtsova, M. et al. (2010). Ta2O5 Nanocrystals Created by Anodization, Proceedings of X.

Workshop of Physical Chemists and Electrochemists, pp. 259 - 261, ISBN 978-80-7375-

396-2, Brno, Czech Republic, June, 2010

Wang, A. W. & White, R. M. (1995). Thin-film anodized aluminum on an acoustic sensor. In:

Ieee Ultrasonics Symposium Proceedings, Vols 1 and 2, Levy, Schneider, McAvoy, pp.

437-440, IEEE, ISBN 1051-0117, New York

Wang, C. C. et al. (2007). Organic nanowire-templated fabrication of alumina nanotubes by

atomic layer deposition. Nano Letters, Vol. 7, No. 6, pp. 1566-1569, ISSN 1530-6984

Wang, H. W. et al. (2006). Standing [111] gold nanotube to nanorod arrays via template

growth. Nanotechnology, Vol. 17, No. 10, pp. 2689-2694, ISSN 0957-4484

Wang, J. G. et al. (2004). Microstructure and interdiffusion of template-synthesized Au/Sn/Au

junction nanowires. Nano Letters, Vol. 4, No. 7, pp. 1313-1318, ISSN 1530-6984

Wang, Q. et al. (2010). Resistive and capacitive response of nitrogen-doped TiO2 nanotubes

film humidity sensor. Nanotechnology, Vol. 22, No. 2, pp. 11, ISSN 0957-4484

Wang, Y. et al. (2005). Synthesis and electrochemical properties of vanadium pentoxide

nanotube arrays. Journal of Physical Chemistry B, Vol. 109, No. 8, pp. 3085-3088, ISSN

1520-6106

Part 3

Optical Microsensors

7

Surface-Enhanced Raman Scattering Sensors

based on Hybrid Nanoparticles

Rafael Contreras-Cáceres, Benjamín Sierra-Martín and

Antonio Fernández-Barbero

Applied Physics Department, University of Almería

Spain

1. Introduction

Surface-enhanced Raman scattering (SERS) is a powerful vibrational spectroscopic

technique that allows ultra-sensitive chemical or biochemical analysis (Kneipp, Kneipp et

al. 1999). It works by increasing the Raman signal of analyte molecules located nearby the

surface of metallic nanostructures that can undergo localized surface plasmon resonance.

Among these nanostructures, gold and silver nanoparticles are the dominant substrates, for

both experimental and theoretical perspectives (Kneipp, Wang et al. 1997; Nie and Emery

1997), since they can support plasmon resonance properties able to increase the Raman

signal up to 14 or 15 orders of magnitude, high enough to detect single molecules (Nie and

Emery 1997; Qian and Nie 2008). Since the first report concerning the enhanced Raman

signal of pyridine molecules adsorbed on a roughened silver electrode (Fleischm, Hendra et

al. 1974), considerable efforts have been made in understanding the SERS mechanisms

(Schatz 1984; Campion and Kambhampati 1998). Nowadays, analytical applications have

centred the attention, and research is devoted to optimize the specific conditions for

detecting each particular analyte (Porter, Lipert et al. 2008). Interestingly, the enhancement

factor is found to depend on the different affinity of the functional groups in the analyte

toward gold or silver surfaces because it is the affinity which determines the analyte

retention (Pearson 1963; Pearson 1966). To improve the surface-analyte interaction, various

approaches have been developed, including the functionalization of nanoparticle surface

(Guerrini, Garcia-Ramos et al. 2006; Guerrini, Garcia-Ramos et al. 2008); however, a problem

inherent to this alternative is that usually the assembled molecules provide strong SERS

signals that overlap and screen those corresponding to the analyte. Another alternative

relies on controlling the surface charge of the nanoparticles to promote the electrostatic

attraction of the analyte onto the particle surface (Alvarez-Puebla, Arceo et al. 2005; Aroca,

Alvarez-Puebla et al. 2005). This approach has been reported to consistently enhances the

signal for acids and amines, but it hardly helps in the case of alcohols, ethers, and other

oxygen containing groups, as well as for non-functionalized molecules. Thereby, there is a

clear need for development of new nanocomposites, based on noble-metals, containing a

sensitive material that enables the physical trapping of a wide variety of analyte molecules.

Herein we present the synthesis and applications of novel core-shell nanocomposites

comprising Au and Au-Ag bimetallic cores, with spherical or rod-shaped morphology,

162

Microsensors

coated with thermally responsive poly-(N-isopropylacrylamide) (pNIPAM) microgel

(Contreras-Caceres, Sanchez-Iglesias et al. 2008). In these systems, whereas the metallic core

provides the necessary enhancing properties, the pNIPAM shell, that can swell or collapse

as a function of temperature, is used to trap the analyte molecules and get them sufficiently

close to the core. These materials present unique optical properties as a consequence of the

thermally responsive surface plasmon resonance, which can be ultimately exploited for

SERS analysis. Although similar systems have been proposed for applications in catalysis

(Lu, Mei et al. 2006), temperature or pH sensing (Kim and Lee 2004), or light-responsive

materials (Gorelikov, Field et al. 2004), we report here that the hybrid nanoparticles can

function as general sensors for detecting different types of analytes. Apart from the SERS

enhancement, these nanocomposites can also be used to modulate the fluorescence intensity

of adsorbed chromophores as a function of temperature. It is important to note, that the

pNIPAM shell not only enhances the colloidal stability of the system in aqueous solutions,

but additionally prevents electromagnetic coupling between metal particles, thus providing

highly reproducible SERS signal and intensity, which is crucial for quantitative applications.

Through a rational choice of model analytes, we report the applications of these

thermoresponsive hybrid materials for Surface Enhanced Raman Scattering and

Fluorescence (SERS and SEF, respectively). The nanocomposites are first tested using 1-

naphthalenethiol (1NAT) as a model analyte with large affinity for gold, and consecutively

against a common dye, Nile Blue A (NBA), whose affinity for gold is lower than of 1NAT. In

addition, we present the SERS analysis of 1-napthol, a substance that had remained elusive

for SERS since it does not easily adsorb onto conventional silver or gold surfaces and whose

detection is decisive because is considered a relevant biomarker (Hansen, Omland et al.

1994; Sun, Shen et al. 2008) and also causes genotoxicity under chronic exposure to humans

(Kozumbo, Agarwal et al. 1992; Grancharov, Engelberg et al. 2001). To conclude the report,

the SERS efficiency of the different hybrid nanocomposites is compared for a couple of

analytes. The wide range of systems investigated, lead us to establish the effect of

parameters, such as particle morphology or core composition, on the detection capabilities.

Interestingly, sensors based on Au-Ag core coated by the pNIPAM shell are found to

provide much higher SERS intensities than their Au-pNIPAM counterparts, not only in the

case of spheres but particularly for nanorods.

2. Plasmon resonance and surface-enhanced Raman scattering

Plasmons are quantized collective oscillations of the free electron gas density that occurs

between any two materials whose dielectric function changes sign across the interface, for

instance metal-dielectric interfaces (Barnes, Dereux et al. 2003). Surface plasmons are those

confined to surfaces; they can strongly couple with photons resulting in surface polaritons,

which are considered quasi-particles that propagate along the metal surface until its energy

decays via absorption into the metal or radiation into the free-space (Zayats, Smolyaninov et

al. 2005). Light or electric fields can excite those plasmons, then resulting in surface and

localized surface plasmon resonance (SPR and LSPR) in the case of planar and nanometric-

sized metallic structures, respectively (Mulvaney 1996). Plasmon oscillation is resonant with

the light at a particular frequency. The electric field intensity, the scattering and the

adsorption cross-sections are then enhanced. Materials exhibiting surface plasmon

properties are used to maximize surface sensitive spectroscopic techniques, such as Raman

scattering or fluorescence (Hutter and Fendler 2004). The resonance frequency strongly

Surface-Enhanced Raman Scattering Sensors based on Hybrid Nanoparticles

163

depends on the size and shape of the metal nanoparticles, as well as, on the metal complex

dielectric function and surrounding medium. Noble metals such as copper, silver, and gold

exhibit strong visible-light plasmon resonance, whereas other transition metals show only a

broad and poorly resolved absorption band in the ultraviolet region (Link and El-Sayed

1999). To understand the optical properties of these metals, it is not only necessary to

account for the effect of free-electrons, responsible for plasmon resonance, but also for the

interband transitions (Wang, Tam et al. 2005). For instance, copper nanoparticles have

strong interband transitions which overlap with the plasmon resonance energies, then

leading to a damping effect that minimizes its optical response. Contrarily, in case of gold

and silver nanoparticles, both effects are well separated in the spectrum. Therefore, electrons

of the conduction band can move freely, showing higher polarizability. This fact, in turn

shifts the plasmon resonance to lower frequencies with sharp bandwidth. Since copper is

also easily oxidized, gold and silver nanoparticles are more attractive for optics-based

applications, specifically silver since it has by