Sie sind auf Seite 1von 27

Downloaded from rsta.royalsocietypublishing.

org on October 20, 2014

A new approach to gas sensing with


nanotechnology
Swati Sharma and Marc Madou
Phil. Trans. R. Soc. A 2012 370, doi: 10.1098/rsta.2011.0506, published 16
April 2012

References

This article cites 56 articles

Subject collections

Articles on similar topics can be found in the following


collections

http://rsta.royalsocietypublishing.org/content/370/1967/2448.ful
l.html#ref-list-1

biomedical engineering (146 articles)


chemical engineering (12 articles)
materials science (191 articles)
nanotechnology (112 articles)

Email alerting service

Receive free email alerts when new articles cite this article - sign up
in the box at the top right-hand corner of the article or click here

To subscribe to Phil. Trans. R. Soc. A go to:


http://rsta.royalsocietypublishing.org/subscriptions

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

Phil. Trans. R. Soc. A 370, 24482473


doi:10.1098/rsta.2011.0506

REVIEW

A new approach to gas sensing with


nanotechnology
B Y S WATI S HARMA1, *

AND

M ARC M ADOU2

1 School

of Nano-Bioscience and Chemical Engineering, UNIST,


Ulsan 689-798, South Korea
2 Department of Mechanical and Aerospace Engineering, University of
California, Irvine, CA 92697-3975, USA
Nanosized gas sensor elements are potentially faster, require lower power, come with a
lower limit of detection, operate at lower temperatures, obviate the need for expensive
catalysts, are more heat shock resistant and might even come at a lower cost than their
macro-counterparts. In the last two decades, there have been important developments in
two key areas that might make this promise a reality. First is the development of a variety
of very good performing nanostructured metal oxide semiconductors (MOSs), the most
commonly used materials for gas sensing; and second are advances in very low power loss
miniaturized heater elements. Advanced nano- or micronanogas sensors have attracted
much attention owing to a variety of possible applications. In this article, we rst
discuss the mechanism underlying MOS-based gas sensor devices, then we describe the
advances that have been made towards MOS nanostructured materials and the progress
towards low-power nano- and microheaters. Finally, we attempt to design an ideal
nanogas sensor by combining the best nanomaterial strategy with the best heater
implementation. In this regard, we end with a discussion of a suspended carbon nanowirebased gas sensor design and the advantages it might offer compared with other more
conventional gas sensor devices.
Keywords: gas sensor; carbon nanowire; biosensor; nanowire sensor; chemical vapour deposition

1. Introduction
Over the past ve decades, there has been an increasing demand for inexpensive,
accurate, portable and reliable gas sensors that can discriminate between very
low concentrations of analytes. Typically, gases of interest include CO, NO, NO2 ,
NH4 , SO2 , CO2 , CH4 and other hydrocarbons. These gases can be harmful to
human health if present beyond a certain concentration. Historically, gas sensors
were rst primarily used in coal mines where accurate monitoring of hazardous
gases has to be carried out continually. Soon gas sensors were also beginning to
*Author for correspondence (swati_bits@yahoo.com).
One contribution of 10 to a Theme Issue Biosensors: surface structures and materials.

2448

This journal is 2012 The Royal Society

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

Review. Gas sensor

2449

appear in the chemical industry, in environmental pollution monitoring units and


in the human health sector. Some other important applications of gas sensors
include the analysis of organic vapours, such as toluene, methanol and benzene,
for laboratory and industrial safety [1] and, breath analysis for trafc safety [2]
and disease diagnosis [3]. An electronic nose, based on arrays of gas sensors [4]
for testing the smell of food, perfumes and articial fragrances etc., is one of the
more recent embodiments of gas sensing that has received much attention owing
to its enhanced analytical power.
The performance of a gas sensor is foremost measured by its lower limit of
detection (LOD) and by its selectivity. The gas sensor with the lowest LOD is
the one that affords the detection of the lowest level of an analyte and a highly
selective gas sensor enables the detection of very minute amounts of the desired
gas in a mixture of several other gases. Low power consumption, fast response and
recovery time, stability (low drift rate) and sensitivity, i.e. the slope of the sensor
output versus the analyte concentration, are other important sensor performance
characteristics.
Gas sensors can be classied based on the detection mechanisms involved
and the type of materials used in their manufacture. They may be based on
a wide variety of detection mechanisms, such as changes in optical absorption or
uorescence [5] modulation of the eld in eld-effect transistors [4], changes in
surface acoustic wave propagation [6], electrochemical phenomena (e.g. voltage
changes in potentiometry [7,8] or current change in amperometry [9]) and,
resistance changes in metal oxide semiconductors (MOSs) [10]. In terms of gas
sensor materials, metal wires and lms were employed as sensing elements in
some of the very rst gas sensors, but polycrystalline MOSs dominate gas sensor
construction today. Metal catalysts such as Pt and Pd are often mixed with these
semiconductor materials to catalyse gas sensing through the spillover effect that
is explained later in this article.
The measuring principle of the gas sensors we focus on in this contribution is
that of electrical conductivity changes of heated MOS layers as a function of the
composition of gases present in the atmosphere surrounding the sensing element.
The MOS gas-sensing materials are typically n-type semiconductors that under
normal atmospheric conditions and typical working temperatures of 300600 C
develop an electron-depleted surface layer. This electron depletion layer is caused
by atmospheric oxygen that at high temperature adsorbs on the surface and

steals electrons from the MOS lm surface to form reactive O


2 or O species.
The presence of gases in the surrounding atmosphere that react with the reactive
oxygen species at the MOS surface modulates the depletion layer thickness, and
this leads to measurable changes in the resistance of the material. Reducing gases
like CO and H2 react with the highly sensitive MOS surface and remove some of
the chemisorbed oxygen while injecting electrons back into the MOS, which causes
a reduction in the thickness of the depletion layer. Oxidizing gases such as NO2 ,
on the other hand, steal even more electrons from the solid, thereby increasing
the thickness of the depletion region [11]. Any such microscopic changes in the
depletion layer thickness affect the overall resistance of the MOS material, which
is measured using simple and inexpensive electrical circuitry.
To achieve selectivity in these types of gas sensors, the required operating
temperature is usually between 300 C and 600 C; at these temperatures,
chemisorption becomes more important than physisorption and chemisorption
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2450

S. Sharma and M. Madou

is by far the more selective process. In the gas sensor community, we instinctively
know that fact: physisorbed water converts almost every material operating at
low temperature into a humidity sensor [12]. Therefore, the sensing unit must
feature a heating element that provides the desired temperature for adsorption
and desorption of oxygen gas and for reaction of the analyte gases with those
reactive adsorbed oxygen species. The reactions at the heated MOS surface
change the concentration of electrons in the MOS lm depletion layer, and this in
turn changes the conductance of these devices as a function of gas concentration.
These gas-induced resistance changes are inuenced by many material-related
factors such as the type of semiconductor oxide material, oxide lm thickness,
grain size, porosity, material sintering temperature and time, catalyst type and
catalyst particle distribution.
In an optimized gas sensor design, the power consumption should be as low
as possible and the heat losses owing to convection, conduction or radiation by
the heater should thus be minimized. To minimize the power consumption, one
needs to reduce the mass of both the heater and the sensor element and optimize
the design for uniformity of heating. Both the heater and the sensor material
are obviously strongly inuenced by miniaturization. With the advancement of
micro- and nanofabrication techniques, sensors evolved from small catalytic bead
platinum wire detectors and Taguchi gas sensors to micro- and nano- solid-state
gas-sensing devices.
In this contribution, we rst discuss the effect of downscaling on sensing
materials and on heater elements and then suggest possible ways to integrate
optimized nanomaterials and nanoheater elements. The latter will be illustrated
with an example of a suspended carbon nanowire (CNW)-based gas sensor design
conceived in our laboratory. We believe that this newly conceived sensor will
become the basis for many types of new gas sensors for biological and biomedical
applications.

2. Sensing materials: effect of grain size and lm thickness


The most traditionally used materials for solid-state gas sensors are thick lms of
polycrystalline compressed or sintered MOS powders. In the case of compressed
powders, the grain-boundary resistance dominates the response of the sensor
device since the resistance at these intergranular contacts is much larger than
the resistance across a single grain. Compressed powders are not very stable as
the intergranular contact is pressure sensitive. In sintered powders, some of these
grain boundaries disappear and form more stable neck-like structures as the grains
sinter together. These necks often constitute the most sensitive sites of sensing
materials as all the electrons from these narrow necks are potentially tied up with
the adsorbed oxygen, giving rise to very sensitive completely electron-depleted
highly resistive regions.
In gure 1, we illustrate the energy band model for an MOS compressed
powder. A few grains are shown where chemisorbed oxygen has captured electrons
from the material forming a depletion layer (also known as the space charge layer
or the Debye layer). At the intergranular contact, two space charge layers backto-back form a formidable barrier for electrons to cross, and this gives rise to the
higher resistance values at those contact points. The thickness of the depletion
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2451

Review. Gas sensor

O O O

O O

O O O
O O O O O

O O
O

O
O
O

O O

O O O O O

O
O

physical model

O
O O O

absorbed oxygen

depletion (of e)
region
conduction
band electrons
electronic
current

+ +
donors

qVs

O O

+ +

barrier
band model

Figure 1. Three grains of a semiconductor oxide showing how the intergranular contact resistance
comes about. The height of the energy barrier is qVs (see equation (2.1)). (Online version in colour.)

layer or the Debye layer (L) is relatively small compared with the grain diameter
(D) in most traditionally polycrystalline materials. A typical value for L is 100 nm
for MOS lms [12]. Grain-boundary-controlled gas sensing, most important for
relatively large-particle MOS sensors, has been extensively reviewed by Barsan
and Weimar [13,14]. In this contribution, we use these early sensing models to
gain an understanding about what to expect when particle size decreases to the
nanometre range.
When an MOS gas sensor is exposed to reducing gases, such as H2 , CO and
CH4 , at a sufciently high temperature, the chemisorbed oxygen reacts with
the reducing gases, lowering the steady-state surface coverage of the oxygen
while injecting electrons into the MOS crystallites and thus lowering the height
of the potential barriers at the intergranular contacts (qVs in gure 1). The
important reactions involved in this type of gas sensor may thus be summarized
as follows.
Oxygen reactions
MO + O + e MO O
MO + 2O + e MO O
2

(between 150 C and 300 C)


(between 30 C and 150 C).

As a result, electrons are extracted and resistance goes up.


Phil. Trans. R. Soc. A

(R 1)

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2452

S. Sharma and M. Madou

Reducing gas reactions


Reducing gas molecules (R) react with the chemisorbed oxygen at the grain
boundaries and/or necks,
R + MO O MO + RO + e .

(R 2)

As a result, electrons are injected and resistance goes down.


Also, oxidizing gases can be detected. The oxidizing reactions may be
summarized as
(R 3)
MO O + O + e MO O
2.
As a result, electrons are extracted and resistance goes up even further. Therefore,
by measuring the change in the conductivity of the semiconductor oxide lms,
one can detect reducing and oxidizing gases in the atmosphere.
The above analysis of the MOS sensing mechanism, for the sake of clarity and
brevity, is somewhat simplied. Indeed, not only a change in the number of free
electrons (and thus boundary layer thickness modulation) but also changes in
electron mobility may modulate the sensing response, as investigated by Ogawa
et al. [15] and recently reviewed by Tricoli et al. [16]. Furthermore, there is still an
ongoing debate in the MOS gas sensor community about the possibility of another
sensing mechanism altogether. Gurlo & Riedel [17] point out that, although it has
been sought for a long time, there is not yet any convincing spectroscopic evidence
for oxygen ionosorption as shown in gure 1. In their alternative model, there
is no oxygen ionosorption on the MOS surface and the model focuses instead
on oxygen vacancies at the surface, which are considered to be the determining
factor in the chemiresistive behavior [18]. SnO2 , the model gas-sensing material,
is oxygen decient and therefore acts as an n-type semiconductor, with oxygen
vacancies that act as electron donors. Alternate reduction and reoxidation of
the surface by gaseous oxygen control the surface conductivity and therefore the
overall sensing behaviour in this model.
Catalysts such as Pt enhance gas-sensing reactions of MOS via the spillover
effect. Spillover refers to the process in which the catalyst dissociates a gas
molecule and then the atoms spill over onto the surface of the MOS. At the right
temperature, a gas reactant rst adsorbs onto the surface of the catalyst particles
and then migrates to the oxide surface to react there with reactive surface oxygen
species. For this process to be possible, the spilled-over species must be able to
migrate to the interparticle contact (grain boundary or neck). As a consequence,
for a catalyst to be effective, there must be a very good dispersion of the catalyst
particles so that they are available near all contacts. Only then can the catalysts
affect the interparticle contact resistance that controls the gas sensor sensitivity.
Importantly, in the case of nanoparticles, there are several reports which indicate
that the need for metal catalysts is somewhat alleviated [19,20].
Porosity, grain size and lm thickness all play an important role in the LOD
and sensitivity of MOS gas sensors. Lee et al. [21] carried out a comparative
study of thick- and thin-lm SnO2 gas sensors in order to understand the
differences in gas-sensing characteristics as a function of the microstructure of
the gas-sensing materials. Thin lms, in their investigation, were deposited using
metal organic chemical vapour deposition (CVD), which yielded a dense compact
microstructure (gure 2). Thick porous lms were derived from metal organic
decomposition. These thick lms were made up of a network of loosely connected
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2453

Review. Gas sensor


gas

(a)
gas

product

(b)
product
gas

gas
product

product

gas

product

gas
product

porous layer of grains

compact layer of grains

Figure 2. Schematic of a gas-sensing reaction in (a) the compact layer and (b) the porous layer.
(Online version in colour.)

crystallites that also featured several cracks (approx. 1 mm wide, [21]) as well as
small pore channels at the crystallite surfaces that all help in the penetration
of gas molecules throughout the material and thereby enhance the sensitivity
(gure 2). Lee et al. conrmed a much improved gas sensitivity in the case of
porous thick-lm microstructures compared with compact thin layers of SnO2 .
In a porous MOS lm, grain boundaries are present in all directions and more
surface area is available for reactions (R 1)(R 3) to take place. In a compact lm,
on the other hand, only a limited exposed surface area is affected by the presence
of analyte gases. Therefore, porous lms are superior for gas-sensing purposes. In
gure 2, we illustrate the surfaces that are likely to participate in the resistance
change of these two types of MOS gas-sensitive material congurations.
Since the sensitivity of the high-surface-area porous gas sensors resides in the
high resistivity grain-boundary contact points, one wants to make the number of
these contact points as high as possible. So a thicker porous lm is better. One
also wants to make these contact points as stable as possible. A powder that is
not sintered is pressure sensitive and quite unstable. In order to improve this,
the MOS powders are sintered, giving rise to the formation of necks between
the grains in the powder. The sensitivity of the lm depends very much on the
extent of the sintering of the material. One should not sinter so much that the
lm becomes one compact layer, reducing the porosity and thus the sensitivity
of the sensor dramatically. In most sintered MOS lms, both necks and grain
contacts are present. Neckgrain models with different degrees of neck formation
have been investigated by several research groups.
Xu et al. [22] proposed a model with a chain of crystallites connected mostly
by necks and sometimes by grain-boundary contacts. These authors assumed
neck size (X ) 0.8 times the crystallite size (D) and suggested that, when D is
larger than 2L, grain-boundary contacts display higher resistance and govern
the electric gas sensitivity of the chain (grain-boundary control). As D becomes
smaller and comparable to 2L, necks become most resistant and therefore they
start controlling the gas sensitivity (neck control). Finally, when D is smaller than
2L, the resistance of grains dominates the whole resistance of the chain and the gas
sensitivity in this case is controlled by grains themselves (grain control) and yields
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2454

S. Sharma and M. Madou


D >> 2L (grain boundary control)
core region
(low resistance)

D 2L (neck control)

space-charge region
(high resistance)

D < 2L (grain control)

Figure 3. Schematic models for grain-size effects (adapted from [23]).

the largest gas sensor response. Figure 3 demonstrates the three situations. Along
the same line, Rothschild and Komem [19] showed that the gas-induced variations
in the trapped charge density in the MOS are proportional to 1/D, where D is
the average grain size. At the nanoscale, the grain size can be decreased to match
the Debye layer (L) thickness (typical value is 100 nmsee above), or can even
be made smaller, which leads to the extraction of all electrons by the chemisorbed
oxygen present. Such nanostructures can thus be treated as gas sensor materials
with uniformly changing resistance upon exposure to gases (see grain control
in gure 3).
Ma et al. [24] carried out a theoretical study to better understand the
dependence of the grain-boundary potential barrier as a function of the density
of electrons trapped at the surface (nt ) and electron density in the bulk of the
material (nb ). These authors suggest a model for neckgrain boundary control
in order to investigate the combined effect of neck and grain boundaries in a
sintered powder. For a constant distribution of such donors, the height of the
grain-boundary potential barrier (eVs , see gure 1) is
e2n 2
eVs =  t
,
(2.1)
2 0 r nb
 
where e is the electron charge, nt is the surface electron density, 0 r is the
permittivity and nb denotes the free electron density in the grain body.
Assuming that the necks in a sintered ZnO powder are cylindrical,
resistance values were calculated for both neck and grain boundaries employing
equation (2.1). Like Xu et al. [22], these authors concluded that gas sensitivity
increases with decreasing grain size, which results from increased neck control.
From this discussion, we can also appreciate that, in the case of a thin compact
lm, one likes the thickness to be small, perhaps in the range of L or less, so as
to have a completely electron-depleted thin lm.
We can conclude from the discussions so far on grain size, porosity and
thickness that, for the best gas sensor performance, we need a thick lm of lightly
sintered nanocrystalline, porous material. Nanocrystalline metal oxides offer other
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

Review. Gas sensor

2455

advantages as well; they can operate at lower temperatures and might obviate
the need for expensive catalysts [25]. Some words of caution are in order here
with respect to the ultimate MOS grain size and the optimal lm thickness.
Nanomaterials do indeed not come without disadvantages. Especially, thermal
instabilities are worrisome since smaller MOS grains tend to agglomerate at lower
temperatures, and this leads to changing sensor characteristics over time. Thermal
degradation is responsible for temporal instabilities in device parameters and
higher drift rates. According to Korotcenkov & Cho [26], the larger the grain size,
the wider the temperature range in which crystallites retain their size and shape
without changes. Conversely, the smaller the grain size, the lower the temperature
at which structural changes start taking effect; for example, lms with a grain
size of 23 nm start transforming at temperatures of approximately 200 C. The
same authors concluded that very small (approx. 20 nm) particles of SnO2 and
In2 O3 MOS materials are not quite suitable for gas sensing above 500 C, as the
particles melt and agglomerate at much lower temperatures than their respective
bulk melting points. An optimum approach that balances a high sensitivity by
decreasing the particle size with good stability, calibration frequency and lifetime requirements of a commercial gas sensor is therefore recommended. A few
words with respect to optimum lm thickness. Above, we advocated making
porous, lightly sintered nanoparticles into a thick lm for the ideal gas sensor
characteristics. On the other hand, some reports have shown that reducing the
lm thickness does increase the sensors response considerably [27]. This is in
agreement with models of MOS sensor response to several analytes [28] and has
been discussed in detail recently by Tricoli et al. [16]. Without more detailed
studies, this is a hard argument to settle; indeed particle size, degree of sintering
and thus porosity are all intertwined. Depending on the porosity and particle size,
either a thinner or a thicker lm might be required.
MOS nanocatalysis can be considered as a bridge between homogeneous and
heterogeneous catalysis. Because of their nanosize, i.e. high surface area, the
contact between reactants, gas sensors and nanoparticles is huge. Pinna et al.
[29], for example, demonstrated that gas sensors based on tin and indium
oxide nanopowders exhibited high sensitivity and good recovery time at low
temperature. Especially, indium oxide nanoparticles were highly sensitive towards
NO2 with a detection limit of 1 ppb at low temperatures.
Nanobelts (nanowire-like structures but with at cross sections) [30], nanorods
[31] and nanowires [32] have all been successfully implemented as gas sensor
materials and continue to be investigated for their potential advantages over
traditional materials. These nanomaterial-based gas sensor devices are based
either on mass assemblies of nanoparticles (e.g. porous mats) or on one or a
set of individual nanoparticles (e.g. a small number of nanowires).
Sadek et al. [30], for example, developed a conductometric H2 , NO2 and
hydrocarbon gas sensor using a ZnO nanobelt lm as the sensitive layer. In Sadek
et al.s experiments, ZnO nanobelts display a structural morphology (wurtzite
family) characterized by a rectangular cross section and a uniform long structure
along their length. A mat of these nanobelts was deposited by radio frequency
(RF) sputtering on a heater element.
The sensing elements in sensors fabricated by Lee et al. [31] are thin lms
consisting of SnO2 nanorods. The nanorod lms also were equipped with welldispersed Pd catalyst nanoparticles with an average diameter of 3 nm. These
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2456

S. Sharma and M. Madou

thin-lm nanorods were deposited using plasma-enhanced CVD (PECVD) with


post-plasma treatment. The surface of these nanorod thin lms was modied with
Pd nanoparticles, and the sensing properties of the thus fabricated lms were
tested with H2 and ethanol vapours. Yeom et al. [33] employed nanomembranes
for pre-concentration of gases that were detected using a microuidic sensor that
resulted in much enhanced sensitivity.
In another study by Liu et al. [34], vanadium oxide nanobelts coated with
MOS nanoparticles such as Fe2 O3 , TiO2 and SnO2 were employed as hybrid
sensor nanostructures for sensing ethanol vapours. These vanadium pentoxide
(V2 O5 ) nanobelts coated with MOS nanoparticles showed a much improved
sensitivity compared with the nanobelts alone. The authors suggest that the
coated V2 O5 nanobelts exhibit enhanced sensitivity owing to the synergy between
electrical transport through the largely depleted nanobelts and the effective gas
sensing on the high surface area MOS nanoparticles that inject electrons into the
nanobelts. The V2 O5 nanobelts were synthesized using a chemical process based
on a mild hydrothermal reaction of 0.1 M ammonium metavanadate solution in
dilute nitric acid, and the MOS nanoparticle coatings were fabricated from ferric
nitrate, tetrabutyl titanate and tin tetrachloride in ethanol, for depositing Fe2 O3 ,
TiO2 and SnO2 , respectively. The sensing mechanism is claried in gure 4 with
a schematic drawing of a longitudinal cross section of nanobelts coated with
MOS nanoparticles.
Upon exposure to ethanol, the electron-depleted surface layers in both the
nanobelts and the nanoparticles are reduced because of the electron injection that
accompanies the reaction of ethanol with the reactive adsorbed oxygen species.
The authors believe that the boundaries between metal oxide grains on the V2 O5
nanobelts dominate the sensor resistance.
As micromachined heater elements tend to be mechanically less sturdy than
the ceramic substrate-based heaters in traditional gas sensors (3), we also need
to pay close attention to how the MOS nanoparticles are deposited. Techniques
that do not exert mechanical pressure on the brittle micromachined heater
substrates and result in porous nanoparticle lms constitute the more attractive
technologies. RF sputtering, for example, as used by Sadek et al. [30] (see above),
is a possible approach but the process needs to be modied so as to generate
more porous lms.
As another example, Cukrov et al. [20] synthesized tin oxide nanoparticles
(average diameter 24 nm) and tested the O2 sensing properties of porous thin lms
prepared by mechanochemical processing and spin coating. The mechanochemical
process uses a conventional ball mill in which the mechanical energy activates
the necessary chemical reactions and induces structural changes. Figure 5 shows
a transmission electron micrograph of tin oxide nanoparticles fabricated by this
mechanochemical method. These particles were subsequently spin coated in a thin
lm on alumina substrates with interdigitated electrodes (measuring electrodes
on the front side and the heater on the back side). These lms were tested for their
oxygen-sensing potential and were found to be extremely stable and repeatable.
Although the nanomaterial is very interesting because it does not require any
metal catalyst and because of its demonstrated stability, spin coating is not a
method of choice to coat the MOS on fragile microelectromechanical system
(MEMS) or nanoelectromechanical system (NEMS) structures. Screen printing
of electrodes and sensing materials, as often employed in the construction of
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2457

Review. Gas sensor


O

O2

OH

MOx nanoparticles
current

V2O5 nanobelts

current

electron-depleted
surface layers
in air
CH3
CH
H

O
H

O2

current
e
current
electron-depleted
surface layers

Figure 4. Schematic of the ethanol-sensing mechanism of vanadium oxide sensors coated with SnO2 ,
TiO2 and Fe2 O3 [34].

traditional gas sensors, is an even less attractive candidate as signicant pressure


is exerted on the substrate in the silk screen printing [35]. During screen printing,
pressure is exerted on the structures [36] and therefore it may not be a very useful
technique for fabricating fragile nanogas sensors.
Another nanopowder gas sensor, based on the synthesis of Al-doped TiO2
nanoparticles (average diameter 100 nm), was reported by Choi et al. [37]. The
nanopowder in this study was used to make thick-lm gas sensors to measure CO
selectively and sensitively in an O2 environment. A thick lm of this material was
prepared on an alumina substrate with interdigitated Pt-measuring electrodes.
To integrate the material with the measuring electrodes, the nanopowder was
mixed in an appropriate solvent (alpha teripineol) and, drop by drop, was
deposited on the electrodes followed by drying and densifying, at 100 C and
800 C, respectively, on the thus deposited lms. This chemical method as well as
solgel techniques and physical vapour deposition [35] (see RF sputtering used
by Sadek et al. [30]) are all feasible approaches to integrate sensing materials on
brittle MEMS and NEMS structures. An especially attractive approach to deposit
MOS locally on a brittle heater structure is to use the heat of the heater element
itself to initiate the CVD from a metallo-organic precursor. Such an approach
was pioneered by Cavicchi et al. [38], who deposited lms on the prefabricated
gas heater structures employing thermally activated CVD.
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2458

S. Sharma and M. Madou

50 nm

Figure 5. Transmission electron micrograph of tin oxide nanopowder [20].

carrier gas
and reactants

carrier gas, unreacted


reactants and products

gas phase reactions

reactant adsorption

product desorption

surface reactions

substrate surface
Figure 6. Basic outline of the CVD technique [35].

In CVD, a controlled mixture of precursors is brought in contact with a heated


substrate, which causes a chemical reaction between the vapour and the substrate
and leads to the deposition of a layer of the desired compound on the substrate.
A basic outline of the CVD method is illustrated in gure 6.
Commonly employed CVD techniques include atmospheric pressure CVD
(APCVD), aerosol-assisted CVD (AACVD), PECVD and rapid thermal
CVD. An advantage of using CVD for gas sensor purposes is that the thickness
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

Review. Gas sensor

2459

and density of the lm can be optimized by controlling the concentration of


the reagents in the vapour and the time of deposition. A concentrated vapour
precursor yields a compact lm that may grow at a rate of up to 1 mm min1 ; a
raried mixture of reagents on the other hand deposits a more porous lm at a
considerably lower growth rate [35].
Two relevant CVD applications are by Shaw et al. [39] and Ashraf et al. [40].
Shaw et al. [39] employed AACVD for the deposition of tungsten oxide from
W(OPh)6 (the precursor) in toluene at 600 C. After a deposition time of 90 min,
a 5075 nm thick brous lm was obtained. The thus fabricated lms can be made
thicker by using an electrical eld in the CVD process. The authors investigated
the porosity of the tungsten oxide lms, a feature, as we know, of key importance
for the utility of these lms in gas sensing, at different applied voltages.
Ashraf et al. [40] demonstrated that APCVD is an effective technique for
depositing tungsten lms as thick as 3.66.7 mm. They used WCl6 as the precursor
vapour and operated the reactor at 625 C in N2 . The particles in these up to
6.7 mm thick tungsten oxide lms were thin needle-like crystals with a high surface
area and high porosity. Signicantly, Ashraf et al. found that the MOS lms
deposited this way are more sensitive than those fabricated by screen printing.
One of our preferred nanogas sensor designs described in 5 is inspired by the
gas sensor conguration shown in gure 4. In our design, we deposit an MOS on
a suspended CNW that we use as both a heater and a sensing electrode. The
CVD deposition is initiated by ohmic heating of the CNWs (as in the case of the
microhotplates shown in gure 8).
Carbon nanotube (CNT)-based miniaturized gas sensors constitute yet another
innovation in nanogas sensing. A single-walled CNT is a one-molecule-thick
layer of graphite rolled into a cylinder with a diameter of a few nanometres.
Multi-walled CNTs are composed of several such graphite cylinders rolled up
into one. The distribution of electrons on their outer surface only makes CNTs
extremely sensitive to charge transfer and chemical doping effects by surrounding
molecules [41,42]. When electron-withdrawing or electron-donating gas molecules
come in contact with CNTs, it very dramatically changes the surface electron
concentration. CNT-based gas sensors show faster response, higher sensitivity and
lower operating temperatures and detection ability for a wider variety of gases
than any other gas-sensitive material [43]. Because of these extensively perceived
advantages, the use of CNTs, both pristine and functionalized, continues to be
studied extensively. Unfortunately, the difculty in making electrical contact to
CNTs and the tremendous variety of lengths, thicknesses and conductivities they
come in have hampered progress. Our approach to nanogas sensing with suspended CNWs (5) constitutes a viable alternative to CNT-based gas sensing as we
can control contact resistance, thickness and resistivity of the nanowires very well.

3. Miniaturization of heater elements in gas sensors


As we discussed above, most MOS gas sensors operate at temperatures above
300 C and, even though nanopowders enable lower operating temperatures, a
heater remains mandatory. Operating at higher temperatures improves selectivity
and reduces the effect of humidity, or stated more generally: chemisorption
prevails over physisorption at higher temperatures. For this purpose, an MOS lm
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2460

S. Sharma and M. Madou

is typically heated using an integrated heating element that facilitates the desired
exchange of electrons between chemisorbed oxygen and the MOS, and between
the analyte gas and the active oxygen species on the MOS surface (see reactions
R1 and R3). To design a good heater element, one wants to minimize power
consumption, make the heater rugged enough so that a gas-sensitive material
can be readily applied to its surface and avoid temperature gradients that may
lead to thermal cracking of the sensor andin case one relies on a modulated
temperature prole for gas sensingone also wants the sensor response time to
be fast.
Two important dimensionless numbers provide zeroth-order guidance for
designing heaters that meet the above thermal requirements. First is the
dimensionless Biot number, dened as the ratio of the surface heat-transfer
coefcient over the bulk heat conductivity (equation (3.1)). The larger the Biot
number, the higher the accumulated heating stress in the heater element and thus
the more real the danger of the heater element cracking,
lL
,
(3.1)
k
where L is the characteristic length, l is the heat-transfer coefcient at the surface
and k is the bulk heat conductivity.
From equation (3.1), we understand that the Biot number is directly
proportional to the characteristic length L (or also to the volume to surface area
(V /A = L)) of the heater. Values of the Biot number smaller than 0.1 imply that
the heat conduction inside the body is much faster than the heat convection away
from its surface, and temperature gradients are negligible inside of it; in other
words, thermal stresses do not accumulate easily in large surface area to volume
structures (such as small heaters). This constitutes a tremendous advantage for
micromachined heating elements; with a very low Biot number, these devices
can be heated and cooled faster than large heater elements without any risk of
thermal-stress-induced cracking.
A second dimensionless number that correlates heat transfer in a body with
its size is the Fourier number (equation (3.2)). This is the ratio of the heat
conduction rate to the rate of thermal energy storage. It is used in the description
and prediction of the temperature response of materials undergoing transient
conductive heating or cooling,

kt
at

=
Fo =
rCp L2 L2
(3.2)
k

and
a=
,
rCp
Bi = Biot number =

where a is thermal diffusivity, i.e. the bulk heat conductivity divided by the
volumetric heat capacity Cp and the material density r; t is the characteristic time
and L is the characteristic length through which conduction occurs. Calculations
based on the Fourier number suggest that a 10 times reduction in size leads
to 102 = 100 times reduction in time to heat or cool the solid. In other words,
downscaling leads to faster heating or cooling.
The Fourier and Biot numbers can be used in transient conduction problems
in a lumped parameter solution which show that the difference between the
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

Review. Gas sensor

2461

temperature of the system and its surroundings (= DT ) as a function of time


t is given by
DT (t) = DT0 eBiFo = DT0 et/t ,
where
t=

rCp L
rCp V
=
,
hA
h

(3.3)

where h is again the surface heat-transfer coefcient, and L is the characteristic


length (V /A), T (t) is the body temperature at time t and DT0 is the initial
temperature difference at time t = 0. It is observed that a larger thermal time
constant t, caused by larger masses rV and larger heat capacities Cp , leads to
slower changes in temperature, while, with larger surface areas A and with better
heat-transfer coefcients (h), a shorter thermal time constant results in faster
temperature changes. The thermal behaviour of a heater can also be expressed
as a function of just one lumped thermal resistance RT and one lumped capacity
C . This way, we can rewrite the thermal time constant as
t = RT C .

(3.4)

From equation (3.4), the thermal time constant depends linearly on the thermal
resistance and capacity, so that a sensor with a low thermal mass and a small
thermal resistance will exhibit the fastest response. A small thermal mass
unfortunately also increases the power consumption so a trade-off needs to be
made between power consumption and speed of response [11].
From equations (3.1)(3.4), we understand that smaller heaters accumulate
lower thermal stresses and therefore are less prone to cracking, and we also
appreciate now that they heat and cool faster. Micromachined gas sensors have
to be optimized with respect to the overall power budget as well: especially if one
wants to run the gas sensor on batteries, the power requirements should be kept
as low as possible. Finally, a uniform distribution of heat within the sensing layer
is important for reproducible and long-lasting gas sensors [44].
In a traditional macro-Taguchi gas sensor, a coiled metal wire is mounted
inside a small ceramic tube (9.5 mm long and 3 mm in diameter) and is employed
for heating the MOS material that is coated over the electrodes on the ceramic
tube surface. Figure 7a is a representative design of these types of macro-gas
sensors. The cylindrical geometry of the sensor shown ensures an even radial
heating throughout the gas-sensing material.
A ceramic-based cylindrical gas sensor as shown in gure 7a is quite power
hungry (about 800 mW!) and is somewhat cumbersome to mass manufacture. In
a typical ceramic planar gas sensor, a metal oxide gas-sensitive layer is screen
printed on a at alumina substrate (e.g. 250 mm thick). On the bottom side
of the substrate, a platinum heater, which might be used simultaneously as a
temperature sensor, is fabricated. On the top side of the ceramic, an interdigitated
sensing electrode pair is screen printed. Then the gas-sensitive layer is screen
printed over these sensing electrodes and red at a temperature of say 850 C.
A picture of this type of sensor is shown in gure 7b. This device structure is
more amenable to mass manufacturing than the tubular sensor but is still high in
power consumption and has the additional disadvantage that the heating prole
is less symmetric.
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2462

S. Sharma and M. Madou


(a)
sensing film

ceramic tube

electrodes

heating wire
(b)

Figure 7. (a) A macro-cylindrical gas sensor with a metal wire as a heating element [45]. (b) A
typical planar gas sensor with screen-printed heater electrodes on the back [46].

Since heat consumption is proportional to the size of the sensor, miniaturization


plays a very signicant role in reducing the power requirement of gas sensors. In
the past two decades, several efforts have been made to microfabricate small
gas sensors with very small heater elements. Semancik & Cavicchi [47], for
example, investigated microhotplate arrays. In this case, the micromachined
heater is mounted in a thin suspended low thermal mass membrane as illustrated
in gure 8. Owing to the small thermal mass of these microhotplates, one
can achieve very rapid heating and cooling with very low power consumption;
with some hotplate designs, power budgets of 15 mW have been achieved
[48,49]. Microhotplates are manufactured using mass production integrated
circuit techniques, so it is easier to make them into arrays for electronic noses [4],
and their fast heating/cooling behaviour allows for their usage in a temperature
scanning mode [50]. Elmi et al. [51] recently demonstrated an ultra-low-power
microhotplate that ran at a very low power value of 8.9 mW at 400 C; their
device also features an innovative self-insulated layout between the heater and
the sensing layer. The CNW-based heaters that we discuss in 5 are expected to
have a much lower power budget owing to the extremely small mass of these wires.
To project how nanotechnology could potentially benet the gas sensor eld,
we now review the heat loss mechanisms at work in a microhotplate-based
gas sensor of the type shown in gure 8 and we analyse what these losses
mean for our own proposed suspended nanowire gas sensor structure (5).
In a high-temperature suspended heater-based gas sensor, heat is transferred
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2463

Review. Gas sensor

20 kV

16

1mm

45 45

SEI

Figure 8. A planer gas sensor with Pt microheaters on a suspended silicon nitride structure [45].

along the membrane via heat conduction; above and below the heater
membrane it is transferred through heat conduction and convection through
the surrounding atmosphere. Finally, we also need to take into account the heat
leakage via radiation. The different heat-transfer mechanisms are illustrated in
gure 9a.

(a) Conduction, radiation and convection in the air gap of the gas sensor
(i) Conduction
The air gap between the heater and the substrate shown in gure 8 provides
for signicant thermal insulation and helps minimize the required heating power.
The heat transport in this gap can be classied into four categories according to
the magnitude of the Knudsen number (Kn, the ratio of the mean free path,
l, of gas molecules to the characteristic length, L, i.e. the thickness of the
air gap). These four different heat-transfer regimes are the continuum regime
(Kn < 0.001), the temperature jump regime (0.001 < Kn < 0.1), the intermediate
or transition regime (0.1 < Kn < 10) and the free molecule regime (Kn > 10). As
the gas pressure p or the characteristic length L is changed, all four of these
regimes may be encountered for a given system. In the free molecular regime, the
Knudsen number is near to or greater than 1, i.e. the mean free path of a molecule
is comparable to the length scale L of the air gap and the continuum assumption
of uid mechanics is no longer a good approximation. In this case, statistical
methods must be used. Han et al. [52] computed the heat transfer through
this thin air gap in the four different regimes. Curve 1 in gure 9b shows the
variation of the conductive heat ux with gap distance as they calculated it. In the
temperature jump regime (0.001 < Kn < 0.1), q depends on the gap distance d as
q=

K (T1 T2 )/d
,
1 + 4gKn(2 1)/a(1 + l)Pr

(3.5)

where q is the density of the conductive heat ux, d is the air gap width, Pr is the
Prandtl number, Kn is the Knudsen number, g is the adiabatic exponential and
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2464

S. Sharma and M. Madou


Qaconduction

(a)

Qaconvection

Qaradiation

heater

Qmembrane

Qmembrane

Thot

Tamb

Tamb
Qbconduction

Qbconvection

Qbradiation

(b)
1.4 107
1.2 107
q (W m2)

1
1.0 107
2

8.0 106
6.0 106
4.0 106
2.0 106
0
109

108

107

106

105
d (m)

104

103

102

Figure 9. (a) Heat uxes of a micromachined gas sensor [11]. In this schematic, superscripts a and
b represent the edges of the square heater, Q is the heat ux, Tamb is the ambient temperature and
Thot is the temperature of the heater. Qmembrane represents the total heat ux through
the heater

membrane. (b) Heat ux versus air gap size between surfaces at 600 and 300 K ( = 1); q is the
variation of the conductive heat ux and d is the gap distance [52]. Curves 1 and 2 represent heat
ux owing to conduction and radiation, respectively.

a is the thermal accommodation coefcient. As can be seen from this gure, the
conductive heat ux increases fast for air gaps between 105 and 107 m and then
approaches a high plateau as the air gap thickness decreases below 107 m. The
plateau value results from the fact that heat conduction changes from a diffusive
to a ballistic transport mechanism when the gap becomes extremely small. From
this graph, it can also be appreciated that if the air gap is increased to a few
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

Review. Gas sensor

2465

microns (see red dots)a typical gap for surface micromachined structuresthe
heat loss owing to conduction is one order of magnitude lower than in the case
of a nanogap.

(ii) Radiation
The radiative heat transfer between the two surfaces of the air gap, in theory,
could be calculated from the StefanBoltzmann law,
qr = 3s(T14 T24 ),

(3.6)

where qr is the density of the radiative heat ux, 3 is the emissivity and s is the
StefanBoltzmann constant. The radiative heat transfer between two surfaces as
calculated from the StefanBoltzmann law does not take into account the size
effect of the air gap d though. So when the gap distance d is of the order of or
smaller than the maximum radiative wavelength at a specic temperature, the
StefanBoltzmann expression must be amended as demonstrated by Whale &
Cravalho [53]. The expression to use in that regime becomes


0.6806 m(m + 1)
hck 3
,
(3.7)
+
Q12 (k) = 3
4p [exp(hkc/2pkB T1 ) 1] p2 y 2
y2

[P12 (u) + P21 (u)]du
(3.8)
Q=
0

kd
.
(3.9)
p
In these expressions, Q is the net radiative energy ux, h is Plancks constant, c
is the speed of light in a vacuum, k is the wavenumber, d is the air gap thickness,
u is the angular frequency and m is the integer value of y. The subscript 12
means the direction of heat ux is from surface 1 to surface 2, and vice versa.
The gap distance in many MEMS structures is of the order where this amended
StefanBoltzmann must be used rather than the StefanBoltzmann equation.
From gure 9b, it can be concluded that the heat dissipation in air gaps by
thermal radiation (curve 2) begins to play a very signicant role when the gaps are
in the range below 107 m. From the same curve, it is also obvious that heat loss by
radiation from the hot sensor element becomes negligible when the air gap is in the
micrometre range. Radiation losses can be reduced by coating the back side of the
thin suspended membrane with a glossy material like gold. At the heater-radiated
wavelengths, the best reector materials are aluminium, silver, copper and gold.
The four metals are approximately equal in reecting infrared, while aluminium
and silver are superior to the coloured metals in reecting visible energy. Gold is
preferred over aluminium for many reector applications because of the variety
of methods by which it can be applied to many substrate materials.

and

y=

(iii) Convection
Air hardly ows in gaps of less than a dozen microns. The heat generated by
the heater element is thus transferred to the substrate predominantly by heat
conduction of air and thermal radiation between the surfaces. Consequently, it
is usually assumed that there is no signicant contribution of convectional uid
motion because of the small size of the heated structures. The validity of this
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2466

S. Sharma and M. Madou

assumption was supported by FEM simulations from Sberveglieri et al. [54] and
Gotz et al. [55], who showed that the difference in simulated heat loss with or
without convection is only about 5 per cent.
In the above treatise, we detail thermal losses owing to changes in the air
gap thickness of the gas sensor. One must also consider lateral conduction heat
losses from the suspended heater plate through the suspending arms connecting
the heater membrane to the substrate (see gure 9a and four suspending beams
connecting the heater membrane to the Si substrate in gure 8). Making the
heater membrane thin and of low heat conductivity materials and making the
suspending arms that hold the sensor membrane and act as electrical contact
conduits very thin and narrow further minimize power losses. Finally, making the
real estate occupied by the heater small in absolute size and small compared with
the suspended membrane size further reduces the absolute power requirements
[11]. We conclude that the total power consumption of miniaturized gas sensors
can be minimized by:
operating the heater at as low a temperature as possible;
incorporating the heater element in a thin low thermal conductivity
membrane;
designing the heater as small as possiblea fraction of the area of the
suspended membrane only;
suspending the membrane heater elements at least a couple of micrometres
above the substrate; and
reducing radiation losses by coating the back side of the membrane with
a glossy material like gold.

4. Integrating miniaturized heaters with nanomaterials for advanced


gas sensing
In 2 and 3, we reviewed progress towards better nanogas sensor materials
and better micro- and nanogas sensor heater elements. We now look into how
to combine these miniaturized heater and sensor materials in an optimum
way in terms of gas sensitivity, lowest power budget and simplicity and cost
of manufacture. In gure 10, we summarize the evolution of heater designs
(column 1) and sensor materials (column 2) towards our recommended design
shown at the bottom. For example, a microhotplate from column 1 could be
combined with a porous thick lm of MOS nanopowder particles (such as those
described by Lee et al. [21]). The last entry in column 1 is that of a CNW
heater element suspended between two carbon posts. The fabrication process
of this nanoheater element is detailed in 5. The novel gas sensor we propose
here combines a suspended CNW heater, shown in gure 11, with the optimum
MOS material deposition method (bottom columns 1 and 2). Inspired by the
gas sensor conguration in gure 4, where MOS nanoparticles are deposited
on vanadium oxide nanobelts, we wanted to deposit MOS nanoparticles on
CNWs. The big difference from the design in gure 4 is that the nanowire
in our case is also a heater element suspended between two carbon-measuring
electrode posts. As we came to understand from the current review, methods
that are compatible with deposition of MOS nanoparticles on a nanowire include
chemical methods with deposition from solution, as well as solgel techniques
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2467

Review. Gas sensor


size reduction in heater (column 1) and particles of sensing material (column 2)
-

- - O- O O O-O O O- O O O- OOO
OOOO
O - - - - - - - - - O O O O O O OO O O

- - - - -O
O-O-O-O-O-O O- - -O O O O O OOOO O
O- O- O- -O- - O- - O-O
O O-O-O- O-O- OO O O

heated CNW
electrodes

Figure 10. Effect of size reduction on materials and heaters of a gas sensor. In column 1, heater
elements of a macro-Taguchi sensor, a microhotplate thin-lm sensor and a suspended CNW sensor
(see gures 7 and 8) are illustrated. Column 2 represents diagrams of polycrystalline sintered
grains. The third picture in column 2 represents a proposed design for a nanogas sensor that uses
a nanowire with a porous lm of MOS nanopowder coated on it. Over the years, a combination of
the most suitable heater from column 1 and material from column 2 would yield a sensitive and
cost-effective nanogas sensor. (Online version in colour.)

(a)

(b)

(c)

(d )

Figure 11. SEM images displaying the diameter (a) and length (b) of a single suspended CNW. (c)
Low-magnication SEM image displaying the adhesion of CNW onto C-MEMS electrode walls. (d)
Carbon MEMS-CNW structure for gas sensing employing the suspended nanowire. (Online version
in colour.)

and physical vapour deposition and CVD. An especially attractive approach to


deposit an MOS locally on a brittle heater structure such as the CNWs in gure
11 is to use the heat of the heater element itself to initiate the CVD deposition
from a metallo-organic precursor. As we already mentioned in 2, this approach
for local CVD was pioneered by Cavicchi et al. [38].
Another more novel approach is to fabricate the MOS nanopowder together
with the heater itself in a monolithic process described in 5. Importantly, the
proposed suspended nanowire gas sensor, as shown in column 2 of gure 10, is a
very simple one-port-type sensor (with only two contacts). In terms of electrical
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2468

S. Sharma and M. Madou

conguration, gas sensors are either one- (two contacts) or two-port-type (four
contacts) devices [56]. Traditionally, MOS gas sensors are two-port sensors and
have separate heater and sensing electrodes. In one-port gas sensors, the heater
and sensor electrodes are the same and are used for both heating and measuring
the changes in resistance that occur owing to the presence of detectable gases.
The so-called pellistors or catalytic bead sensors [57] work this way. They consist
of a ne powder of catalytic metals supported on an inert porous refractory
ceramic coated onto a Pt wire that is used as both a heater and a sensor [58].
It is important to note that pellistors do not use any semiconductor material for
sensing and therefore differ from a semistor (introduced below). In other words,
pellistors measure the small changes in the conductance of the heater element
(Pt wire) that results from the exothermic oxidation reactions taking place on
the catalyst surface with the heater wire itself. In general, an MOS-based sensor
is more sensitive and versatile since the sensitivity of a pellistor or semistor is
based on temperature change; however, catalytic-type sensors exhibit more linear
and reproducible characteristics [59].
A combination of an MOS-based gas sensor with the one-port operating
mechanism used in a pellistor was proposed by Williams & Coles [58], who called
this new device a semistor (i.e. a semiconductor-based pellistor). A semistor is
prepared by coating a sensitive low-resistance MOS such as SnO2 directly onto
a Pt wire coil. A semistor exploits the sensitivity and resistance modulation
capabilities of an MOS material while maintaining the linearity in behaviour
and ease of manufacture of a pellistor. The nanowire gas sensor we propose is
of the semistor type where resistance changes in the MOS are caused by both
heat of reaction and injection and extraction of electrons. Given the challenge of
patterning two measuring electrodes on an MOS porous lm covering a CNW,
the semistor approach is probably the only practical way to develop nanowirebased gas sensors. Recently, Bissi et al. [60], the same group that introduced the
ultra-low-power microhotplate that ran at an unprecedented low-power value of
8.9 mW (see above), switched to a semistor-like approach. These investigators
do introduce a novel three-terminal device by patterning a single metallization
layer (Pt) and, as in a semistor approach, there is no passivation layer for
electrically insulating the sensing layer from the heater in their design. This
resulted in a simplied, cost-reducing fabrication process that came with even
lower power consumption. Their three-terminal sensor features a two-contact
circular heater and a single measuring electrode. The role of the heater is twofold:
when biased with a suitable voltage, it maintains the sensor at the correct
operating temperature and at the same time it acts as the second electrode for
contacting the sensing layer. When a known voltage difference is applied between
the heater and the third electrode, the current leaking from the heater through
the sensing layer is acquired. This three-terminal approach might be possible to
implement in our proposed suspended nanowire gas sensor as there is no need for
patterning of the third electrode in this case.

5. Proposed fabrication of suspended CNW gas sensors


In our laboratory, we have designed a novel suspended CNW-based gas sensor.
These CNWs are composed of glassy and graphitic carbons and display very
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

Review. Gas sensor

2469

high temperature resistance. This material is very suitable as a miniaturized


heater in a gas sensor owing to the structural stability of carbon at very
high temperatures. The fabrication of these wires involves heating at 900 C
and therefore no further testing for thermal stability is required up to this
temperature. The thus fabricated sensors may have one or many suspended
CNWs connected to carbon microelectrodes without any glue or clip. These
nanowires have an aspect ratio as high as 1000 with sub-50 nm diameters and
approximately 30 mm lengths. The resistance value of these wires falls between
that of glassy carbon and graphite. Structural characterization using X-ray
diffraction, Raman analysis and high-resolution transmission electron microscopy
imaging conrmed that these nanowires contain both amorphous and crystalline
carbon regions.
These suspended CNWs are deposited by electrospinning of a polymer
precursor on MEMS support structures. The MEMS support structures are
fabricated by employing standard SU-8 photolithography and the support
structure and suspended polymer nanobres are carbonized together at 900 C in
a non-oxidizing environment. The details of this C-MEMS/C-NEMS fabrication
process can be found in Sharma et al. [61]. Electrospinning is one of the simplest
and least expensive techniques for the fabrication of ultrathin polymer bres. In
the past several decades, different types of organic polymers and inks have been
successfully processed as nanobres using this technique, with typical examples
including various biopolymers, electrically conductive polymers and uorescent
polymers. Recently, our group was able to modify the electrospinning technique to
put polymer nanobres down in an extremely controlled fashion on both at and
three-dimensional structured surfaces [62]. As a result, after pyrolysis, we can now
produce CNWs of precise length and position on an appropriate high-temperature
substrate. The resulting carbonized structure also comes with integrated contact
pads, and we demonstrated ohmic contact between the carbon wire and the
support structures. In a typical design, the contact pads are 0.5 0.5 cm and
are connected to two SU-8 supporting walls that are 20 mm wide with a 20 mm
gap in between. The gap between the two walls supporting the CNWs becomes
approximately 30 mm as the walls shrink during pyrolysis. The structures are
fabricated on a Si substrate that is coated with an insulating layer of thermal
SiO2 to ensure insulation for electrical characterization of the carbonized bres
and the C-MEMS-supporting structure and contact pads.
Typical SEM images of single suspended CNWs are shown in gure 11ac. In
gure 11a,b, a 42.22 nm thick nanowire with a length of approximately 29 mm is
shown. It can be observed that the CNWs have a very uniform geometry. However,
to account for any experimental or imaging error in determining the diameter
of CNWs, SEM pictures were taken at three or four different points along the
length of the wire. Uniformity of the wires to within 10 per cent was observed. In
gure 11c, a lower magnication SEM image illustrates the adhesion of a typical
CNW onto the carbon walls. The walls are further connected to 0.5 mm wide
electrode pads that can be conveniently connected to an external power supply to
yield a gas sensor like the one shown in gure 10 (column 2, row 3: a combination
of CNW with nanopowder).
For nanogas-sensing applications, the CNWs shown in gure 11 are used as
extremely low-cost, low-power heater elements. In 3 and 4, we detailed various
MOS deposition techniques and singled out local CVD as a most appropriate
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2470

S. Sharma and M. Madou

method to deposit nanopowders on the CNW. These CNWs can be locally heated
by applying a current and, since the resistance of the wire is much higher than
the rest of the structure (such as the nanowire-supporting electrodes), MOS
nanocrystals only deposit on the nanowire and not on the support structure. As
discussed above, to keep the sensor structure as simple as possible, no separate
measuring electrodes are applied over the MOS nanocrystal lm on top of the
carbon nanoheater; instead, the gas sensor is operated as a one-port nano-semistor
where heat changes and electron injection or extraction owing to reactions result
in MOS resistance changes that are measured on the CNW itself. In a slightly
more complex process, we could attempt a three-terminal device, as described
by Bissi et al. [60], and deposit a third sensing electrode on the outside of the
nanoporous lm.
Besides using the heat of the nanowire heater to locally deposit all types
of MOS nanocrystals directly onto the nanoheater, we are experimenting with
another less conventional approach towards integrating MOSs on the CNWs. In
this alternative approach, we rely on the preparation of mixtures of nanopowders
in a polymer solution that can be electrospun. After electrospinning, the entire
structure (integrated nanopowdernanowiresupporting MEMS structure) is then
pyrolyzed together. Along this line, we have already fabricated composite
PAN/CNT [63] nanobres. The right concentration of the MOS nanopowder is
expected to lead to hybrid MOSCNWs suspended on carbon microwalls, which
are in turn connected to carbon macroelectrodes.

6. Conclusion
The evolution in gas-sensitive materials and heater structures for MOS gas sensors
from macro- to nanosize was reviewed. Size reduction of the grains in an MOS
gas-sensitive material leads to lower operating temperatures, lower LOD, higher
sensitivity, faster response time and less reliance on expensive metal catalysts.
Size reduction of heater elements in turn leads to a lower overall power budget,
faster heating and cooling, lower propensity for cracking under heat stress and
mass production amenable to manufacturing techniques. The objective of this
review was to suggest better ways to combine progress in these two areas so as to
enable better gas sensors for the future. In this context, we are suggesting optimal
combinations of nanomaterial deposition methods (e.g. local CVD of MOS
nanocrystalline lms directly on the heater) with optimal fabrication techniques
for the lowest power budget nanoheaters (e.g. a C-NEMS suspended nanowire
heater). A futuristic device we are researching to illustrate this combination
is a nano-semistor. This is a simple one-port gas sensor based on a suspended
CNW made using C-NEMS and C-MEMS coated with nanocrystalline MOSs. The
carbon wire in this nanogas sensor is both the heater and the sensing element.
The authors believe that this device will offer all the sensing advantages one
expects from an optimal combination of nanoheater construction and optimized
nanocrystalline gas-sensing material deposition.
This work was supported by World Class University (WCU) programme R32-2008-000-20054-0 at
UNIST, South Korea, and UC Lab Fees Award 09-LR-09-117362 at UC, Irvine, CA. The authors
also acknowledge the Indo-US Science and Technology Forum, New Delhi, for their support.
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

Review. Gas sensor

2471

References
1 Wang, X., Carey, W. P. & Yee, S. S. 1995 Monolithic thin lm metal oxide gas sensor
arrays with application to monitoring of organic vapors. Sens. Actuat. B Chem. 28, 6370.
(doi:10.1016/0925-4005(94)0 1531-L)
2 Mitsubayashi, K., Matsunaga, H., Nishio, G., Ogawa, M. & Saito, H. 2004 Biochemical gassensor (Bio-sniffer) for breath analysis after drinking. In SICE Annual Conf, Sapporo, Japan,
46 August 2004.
3 McEntegart, C. M., Penrose, W. R., Strathmann, S. & Stetter, J. R. 2000 Detection and
discrimination of coliform bacteria with gas sensor arrays. Sens. Actuat. B Chem. 70, 170176.
(doi:10.1016/S0925-4005(00)00561-X)
4 Arshak, K., Moore, E., Lyons, G. M., Harris, J. & Clifford, S. 2004 A review of gas
sensors employed in electronic nose applications. Sens. Rev. 24, 181198. (doi:10.1108/02602
280410525977)
5 Huang, Y. & Tao, S. 2011 An optical ber sensor probe using a PMMA/CPR coated bent
optical ber as a transducer for monitoring trace ammonia. J. Sens. Technol. 21, 2935.
(doi:10.4236/jst.2011.12005)
6 Grate, J. W. & Zellers, E. T. 2000 The fractional free volume of the sorbed vapor in modeling
the viscoelastic contribution to polymer-coated surface acoustic wave vapor sensor responses.
Anal. Chem. 72, 28612868. (doi:10.1021/ac991192n)
7 Fouletier, J. 1982/83 Gas analysis for potentiometric sensors: a review. Sens. Actuat. 3, 295314.
(doi:10.1016/0250-6874(82)80030-9)
8 Bakker, E. & Pretsch, E. 2005 Potentiometric sensors for trace-level analysis. Trends Anal.
Chem. 24, 199207. (doi:10.1016/j.trac.2005.01.003)
9 Alber, K. S., Cox, J. A. & Kulesza, P. J. 1997 Solid-state amperometric sensors for gas phase
analytes: a review of recent advances. Electroanalysis 9, 97101. (doi:1040-0397/97/0202-097)
10 Basu, S. & Basu, P. K. 2009 Nano crystalline metal oxides for methane sensors: role of noble
metals. J. Sens. 2009, 120. (doi:10.1155/2009/861968)
11 Simon, I., Barsan, N. & Bauer, M. 2001 Micromachined metal oxide gas sensors: opportunities
to improve sensor performance. Sens. Actuat. B 73, 126. (doi:10.1016/S0925-4005(00)00639-0)
12 Madou, M. & Morrison, S. R. 1989 Chemical sensing with solid state devices. New York, NY:
Academic Press.
13 Barsan, N. & Weimar, U. 2001 Conduction model of metal oxide gas sensor. J. Electroceram.
7, 143167. (doi:10.1023/A:1014405811371)
14 Barsan, N. & Weimar, U. 2003 Understanding the fundamental principles of metal oxide based
gas sensors; the example of CO sensing with SnO2 sensors in the presence of humidity. J. Phys.
Condens. Matter 15, R813839. (doi:10.1088/0953-8984/15/20/201)
15 Ogawa, H., Nishikawa, M. & Abe, A. 1982 Hall measurement studies and an electrical
conductive model of tin oxide ultrane particle lms. J. Appl. Phys. 53, 44484454.
(doi:10.1063/1.331230)
16 Tricoli, A., Righettoni, M. & Teleki, A. 2010 Semiconductor gas sensors: dry synthesis and
application. Angew Chem. Int. Ed. 49, 76327659. (doi:10.1002/anie.200903801)
17 Gurlo, A. & Riedel, R. 2007 In situ operando spectroscopy for assessing mechanisms of gas
sensing. Angew Chem. Int. Ed. 46, 38263848. (doi:10.1002/anie.200602597)
18 Zemel, J. 1988 Theoretical description of gaslm interaction on SnOx . Thin Solid Films 163,
189202. (doi:10.1016/0040-6090(88)90424-5)
19 Rothschild, A. & Komem, Y. 2004 The effect of grain size on the sensitivity of nanocrystalline
metal-oxide gas sensors. J. Appl. Phys. 95, 63746380. (doi:10.1063/1.1728314)
20 Cukrov, L. M., McCormick, P. G., Galatsis, K. & Wlodarski, W. 2001 Gas sensing properties of
nanosized tin oxide synthesized by mechanochemical processing. Sens. Actuat. B 77, 491495.
(doi:10.1016/S0925-4005(01)00751-1)
21 Lee, S. W., Tsai, P. P. & Chen, H. 2000 Comparison study of SnO2 thin and thick lm gas
sensors. Sens. Actuat. B Chem. 67, 122127. (doi:10.1016/S0925-4005(00)00390-7)
22 Xu, C., Jun Tamaki, J., Miura, N. & Yamazoe, N. 1991 Grain size effects on gas sensitivity of
porous SnO2 -based elements. Sens. Actuat. B 3, 147155. (doi:10.1002/anie.200903801)
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

2472

S. Sharma and M. Madou

23 Min, Y. 2003 Properties and sensor performance of ZnO thin lms. PhD thesis, Institute of
Technology, MA, USA.
24 Ma, Y., Wang, W. L., Liao, K. J. & Kong, C. Y. 2002 Study on sensitivity of nano-grain ZnO
gas sensors. J. Wide Bandgap Mater. 10, 112120. (doi:10.1177/1524511X02043537)
25 Wang, C., Yin, L., Zhang, L., Xiang, D. & Gao, R. 2010 metal oxide gas sensors: sensitivity
and inuencing factors. Sensors 10, 20882016. (doi:10.3390/s100302088)
26 Korotcenkov, G. & Cho, B. 2010 Grain size effect in structural stability of SnO2 and In2 O3
lms aimed for gas sensing application. In Proc. 2nd Int. Conf. on Computer Research and
Development, Kuala Lumpur, Malaysia, 710 May 2010, pp. 461464.
27 Tricoli, A. & Pratsinis, S. E. 2010 Dispersed nanoelectrode devices. Nat. Nanotech. 5, 5460.
(doi:10.1038/nnano.2009.349)
28 Becker, T., Ahlers, S., Bosch, V., Braunmuhl, C., Muller, G. & Kiesewetter, O. 2001 Gas
sensing properties of thin- and thick-lm tin-oxide materials. Sens. Actuat. B Chem. 77, 5561.
(doi:10.1016/S0925-4005(01)00672-4)
29 Pinna, N., Neri, G., Antonietti, M. & Niederberger, M. 2004 Nonaqueous synthesis of
nanocrystalline semiconducting metal oxides for gas sensing. Angew Chem. Int. Ed. 43,
43454349. (doi:10.1002/anie.200460610).
30 Sadek, A. Z. Choopun, S., Wlodarski, W., Ippolito, S. J., & Kalantar-Zaden, K. 2007
Characterization of ZnO nanobelt-based gas sensor for H2 , NO2 , and hydrocarbon sensing.
IEEE Sens. J. 7, 919924. (doi:10.1109/JSEN.2007.895963)
31 Lee, Y. C., Huang, H., Tan, O. K. & Tse, M. S. 2008 Semiconductor gas sensor based on Pddoped SnO2 nanorod thin lms. Sens. Actuat. B 132, 239242. (doi:10.1016/j.snb.2008.01.028)
32 Wang, W., Huang, H., Li, Z., Zhang, H., Wang, Y., Zheng, W. & Wang, C. 2008 Zinc oxide
nanober gas sensors via electrospinning. J. Am. Ceram. Soc. 91, 38173819. (doi:10.1111/
j.I551-2916.2008.02765.x)
33 Yeom, J., Oh, I., Field, C., Radadia, A., Ni, Z., Bae, B., Han, J., Masel, R. I. & Shannon, M. A.
2008 Enhanced toxic gas detection using a MEMS pre-concentrator coated with metal organic
framework absorber. In Proc. MEMS , Tucson, AZ, 1317 January 2008.
34 Liu, J., Wang, X., Peng, Q., Li, Y. 2006 Preparation and gas sensing properties of
vanadium oxide nanobelts coated with semiconductor oxides. Sens. Actuat. B 115, 481487.
(doi:10.1016/j.snb.2005.10.012)
35 Fine, G. F., Cavanagh, L. M., Afonja, A. & Binions, R. 2010 Metal oxide semiconductor gas
sensors in environmental monitoring. Sensors 10, 54695502. (doi:10.3390/s100605469)
36 Viricelle, J. P., Pijilat, C., Riviere, B., Rotureau, D., Briand, D. & DeRooij, N. F.
2006 Compatibility of screen printing technology with microhotplate for gas sensor
and solid oxide micro fuel cell development. Sens. Actuat. B Chem. 118, 263268.
(doi:10.1016/j.snb.2006.04.031)
37 Choi, Y. J., Seeley, Z., Bandyopadhyay, A., Bose, S. & Akbar, S. A. 2007 Aluminum-doped TiO2
nano-powders for gas sensors. Sens. Actuat. B 124, 111117. (doi:10.1016/j.snb.2006.12.005)
38 Cavicchi, R. E., Poirier, G. E., Semancik, S., Suehle, J. S. & Gaitan, M. 1993 Development of
C-MOS based gas microsensor arrays with materials and kinetic selectivity. In Proc. 40th AVS
Symposium, Orlando, FL, 1519 November 1993.
39 Shaw, G., Parkin I. P., Pratt, K. F. E. & Williams, D. E. 2005 Control of semiconducting oxide
gas-sensor microstructure by application of an electric eld during aerosol-assisted chemical
vapor deposition. J. Mater. Chem. 15, 149154. (doi:10.1039/b411680a)
40 Ashraf, S., Blackman, C. S., Naisbitt, S. C. & Parkin, I. P. 2008 The gas sensing properties
WO3x thin lms deposited via the atmospheric pressure chemical vapor deposition (APCVD)
of WCl6 with ethanol. Meas. Sci. Technol. 19, 025203. (doi:10.1088/0957-0233/19/2/025203)
41 Dresselhaus, M. S., Dresselhaus, G. & Eklund, P. C. 1996 Science of fullerenes and carbon
nanotubes. New York, NY: Academic Press.
42 Zhang, T., Mubeen, S., Myung, N. V. & Deshusses, M. A. 2008 Topical review: recent
progress in carbon nanotube based gas sensors. Nanotechnology 19, 332001. (doi:10.1088/09574484/19/33/332001)
43 Suehiro, J., Zhou, G. & Hara, M. 2003 Fabrication of a carbon nanotube based gas sensor
using dielectrophoresis and its application for ammonia detection be impedance spectroscopy.
J. Phys. D Appl. Phys. 36, L109L114. (doi:10.1088/0022-3727/36/21/L01)
Phil. Trans. R. Soc. A

Downloaded from rsta.royalsocietypublishing.org on October 20, 2014

Review. Gas sensor

2473

44 Gtz, A., Grcia, I., Can, C., Lora-Tamayo, E., Horrillo, M. C., Getino, J., Garca, C. &
Gutirrez, J. 1997 A micromachined solid state integrated gas sensor for the detection
of aromatic hydrocarbons. Sens. Actuat. B Chem. 44, 483487. (doi:10.1016/S0925-4005
(97)00171-8)
45 Flueckiger, J., Ko, F. K. & Cheung, K. C. 2009 Microfabricated formaldehyde gas sensors.
Sensors 9, 91969215. (doi:10.3390/s91109196)
46 Molodovan, C. et al. 2007 Ceramic micro heater technology for gas sensors. Rom. J. Inform.
Sci. Technol. 10, 4352.
47 Semancik, S. & Cavicchi, R. 1998 Kinetically controlled chemical sensing using micromachined
structures. Acc. Chem. Res. 31, 279287. (doi:10.1021/ar970071b)
48 Gardner, A., Pike, N., deRooji, N., Koudelka-Hep, M., Clerk, P., Hierlemann, A. & Gopel, W.
1995 Integrated array sensors for detecting organic solvents. Sens. Actuat. B 2627, 135139.
(doi:10.1016/0925-4005(94)01573-Z)
49 Dusco, C., Vazsonyi, E., Adam, M., Szabo, I., Barsony, I., Gardeniers, J. & van den Berg, A.
1997 Porous silicon hulk micromachining for thermally isolated membrane formation. Sens.
Actuat. A 60, 235239. (doi:10.1016/S0924-4247(97)01384-8)
50 Giberti, A., Guidi, V. & Vincenzi, D. 2011 A study of heat distribution and dissipation in
a micromachined chemoresistive gas sensor. Sens. Actuat. B 153, 409414. (doi:10.1016/j.snb.
2010.11.007)
51 Elmi, I., Zampolli, S., Cozzani, E., Mancarella, F. & Cardinali, G. C. 2008 Development of ultralow-power consumption MOX sensors with ppb level VOC detection capabilities for emerging
applications. Sens. Actuat. B 135, 342351. (doi:10.1016/j.snb.2008.09.002)
52 Han, M. H., Liang, X. G. & Tang, Z. A. 2005 Size effect on heat transfer in micro gas sensors.
Sens. Actuat. A 120, 397402. (doi:10.1016/j.sna.2004.12.031)
53 Whale, M. D. & Cravalho, E. G. 2002 Modeling and performance of microscale
thermophotovoltaic energy conversion devices. IEEE Trans. Energy Convers. 17, 130142.
(doi:10.1109/60.986450)
54 Sberveglieri, G., Hellmich, W. & Muller, G. 1997 Silicon hotplates for metal oxide gas sensor
elements. Microsyst. Technol. 3, 183190. (doi:10.1007/s005420050078)
55 Gotz, A., Gracia, I., Cane, C., Lora-Tamayo, E., Horrillo, M. C., Getino, J., Garcia, C. &
Gutierrez, J. 1997 A micromachined solid state integrated gas sensor for the detection of
aromatic hydrocarbons. Sens. Actuat. B 44, 483487. (doi:10.1016/S0925-4005(97)00171-8)
56 Watson, J. 1984 The tin oxide gas sensor and its applications. Sens. Actuat. 5, 2942.
57 Krawczyk, M. & Namiesnik, J. 2003 Application of a catalytic combustion sensor (Pellistor) for
the monitoring of explosiveness of a hydrogenair mixture in the upper explosive limit range.
J. Autom. Methods Manage. Chem. 25, 115122. (doi:10.1155/S1463924603000208)
58 Williams, G. & Coles, G. S. V. 1999 The semistor: a new concept in selective methane detection.
Sens. Actuat. B 57, 108114. (doi:10.1016/S0925-4005(99)00061-1)
59 Watson, J. 1984 The tin oxide gas sensor and its applications. Sens. Actuat. 5, 2942.
(doi:10.1016/0250-6874(84)87004-3)
60 Bissi, L., Cicioni, M. & Placidi, P. 2011 A programmable interface circuit for an ultralow power
gas sensor. IEEE Trans. Instrument. Meas. 60, 282289. (doi:10.1109/TIM.2010.2049182)
61 Sharma, C. S., Katepalli, H., Sharma, A. & Madou, M. 2010 Fabrication and electrical
conductivity of suspended carbon nanober arrays. Carbon 49, 17271732. (doi:10.1016/
j.carbon.2010.12.058)
62 Bisht, G. B., Canton, G., Mirsepassi, A., Kulinsky, L., Oh, S., Dunn-Rankin, D. & Madou,
M. J. 2011 Controlled continuous patterning of polymeric nanobers on 3D substrates using
low-voltage near-eld electrospinning. Nanoletters 11, 18311837. (doi:10.1021/nl2006164)
63 Maitra, T., Sharma, S., Sharma, A., Cho, Y. K. & Madou, M. 2011 Improved
graphitization and electrical conductivity of suspended carbon nanobers derived from carbon
nanotube/polyacrylonitrile composites by directed electrospinning. Carbon 50, 17531761.
(doi:10.1016/j.carbon.2011.12.021)

Phil. Trans. R. Soc. A

Das könnte Ihnen auch gefallen