Sei sulla pagina 1di 13

Catalysis Surveys from Asia Vol. 7, No.

1, April 2003 (# 2003) 63

Oxide semiconductor gas sensors


Noboru Yamazoe, Go Sakai, and Kengo Shimanoe
Faculty of Engineering Sciences, Kyushu University, Kasuga-shi, Fukuoka 816–8580, Japan

Semiconductor gas sensors utilize porous polycrystalline resistors made of semiconducting oxides. The working principle
involves the receptor function played by the surface of each oxide grain and the transducer function played by each grain boundary.
In addition, the utility factor of the sensing body also takes part in determining the gas response. Therefore, the concepts of sensor
design are determined by considering each of these three key factors. The requirements are selection of a base oxide with high
mobility of conduction electrons and satisfactory stability (transducer function), selection of a foreign receptor which enhances
surface reactions or adsorption of target gas (receptor function), and fabrication of a highly porous, thin sensing body (utility
factor). Recent progress in sensor design based on these factors is described.
KEY WORDS: oxide semiconductor; sensor; SnO2; sensor design.

1. Introduction such as gas leak detectors and environmental monitor-


ing. Since the pioneering works reported in 1962 by
A semiconductor gas sensor usually utilizes a porous
Seiyama et al. [1] and Taguchi [2], much technological
sintered block body consisting of polycrystalline parti-
effort has been made in this field, aiming at improve-
cles of a semiconducting oxide such as SnO2 ; WO3 ;
ment in gas response, selectivity, stability, and feasibility
ZnO, and In2 O3 : Recently, thick- and thin-film-type
for practical use. Yet further innovations of gas sensing
sensors have also been investigated for miniaturizing
properties are still in great demand to expand the fields
sensor elements by micro-fabrications. In figure 1
of gas sensor applications. Numerous kinds of gases are
examples of block- or film-type devices and the
emitted from various sources into our living space,
polycrystalline particles involved in the devices are
working space, or outdoors. Many of them are
illustrated. Usually the devices for laboratory tests are
hazardous to human beings and the environment, like
furnished with a heater so that they are heated externally
air pollutants, while some others can be used as
to obtain an optimum operating temperature. Upon
measures to diagnose the state of their sources, like
exposure to a particular gas (target gas) of a low
ethanol and acetone in the breath and flavor compo-
concentration in air, the sensor changes its electrical
nents of foods. Most of these gases are present at very
resistance. When this experiment is carried out in a flow
low concentrations so that extremely good sensing
apparatus, one can obtain response transients like those
characteristics are required for their monitoring. In
shown in figure 2(a). The change in the electrical
order to realize such gas sensors, it seems imperative to
resistance is from a level in air ðRa Þ to a steady level
establish the principles of sensor design on the basis of
ðRg Þ on exposure to the target gas and vice versa on
fundamental understandings of semiconductor gas
cutting off the gas flow. There are two important
sensors. It is the aim of the present article to describe
characteristics of the gas sensor in these transients, i.e.,
the state of the art of sensor design in the hope of further
the gas response (often called sensitivity) defined as
advancement of semiconductor gas sensors.
Ra =Rg and the rates of response and recovery. The gas
response ðRa =Rg Þ is very much dependent on the
operating temperature, going through a maximum for
an inflammable gas as shown in figure 2(b). Naturally 2. Working principles
Rg as well as Ra =Rg depend on the kind and concentra- The process of gas sensing by a semiconductor device
tion of target gas besides other conditions. As illustrated
involves two key functions as illustrated in figure 3, i.e.,
in figure 2(c), Rg and gas concentration are usually
(i) recognition of a target gas through a gas–solid
correlated linearly in a log–log scale. Selectivity is
interaction which induces an electronic change of the
defined by the ratio of the response to a gas to that of
oxide surface (receptor function) and (ii) transduction of
another gas.
the surface phenomenon into an electrical resistance
Semiconductor gas sensors have been investigated
change of the sensor (transducer function). However,
extensively for the purpose of practical applications,
the understanding of these functions is not straight-
forward, because of the complex nature of the porous,
polycrystalline sensing bodies comprising nano-sized
 To whom correspondence should be addressed. grains of semiconducting oxides. It was proposed at a

1384-6574/03/0400–0063/0 # 2003 Plenum Publishing Corporation


64 N. Yamazoe, et al./Oxide semiconductor gas sensors

Figure 1. Schemetic drawings of sensor devices: (a) sintered block


type; (b) thick- or thin-film type. (c) FE-SEM image (top view) of a
thin-film derived from SnO2 sol.

relatively early stage that the receptor function is


provided by the surface chemical properties of oxide
grains to adsorb or react with the target gas. This
understanding has been strengthened by the elucidation
of oxygen adsorptive properties of oxides and the
Figure 2. Typical characteristics of semiconductor gas sensor: (a)
findings of remarkable sensitizing actions played by response transient; (b) temperature dependence of gas response; (c)
various foreign receptors introduced into the grains [3]. dependence of Rg on gas concentration.
On the other hand, findings of grain size effects on the
sensor resistance and response [4] as well as of the bodies (utility factor) can predict how the gas response
remarkable effects of several foreign receptors on the depends on the chemical and physical properties of a
work function of oxides [5] have contributed to establish target gas and the microporous structure of sensing
the concept that a surface space charge layer formed on bodies. Familiar volcano-shaped correlations between
the oxide grains induces gas-dependent barriers for gas response and operating temperature (cf. figure 2(b))
conduction electrons to cross the grain boundaries have become understandable.
(transducer function). The relevance of gas response to These key factors, i.e., receptor function, transducer
the porous structure of gas sensing bodies has been function, and utility factor, can be changed often
analyzed based on a diffusion-reaction equation [6]. The markedly with changes in base semiconducting oxides,
penetration depth of the target gas into the gas sensing foreign materials introduced, and high-order structure
N. Yamazoe, et al./Oxide semiconductor gas sensors 65

Figure 3. Receptor and transducer functions of the semiconductor gas sensor.

of the sensing bodies. The basic aspects of the key without additives. For example, WO3 has been used
factors and related sensor design principles are described preferably for sensing NO2 and NH3 : However, the
in more detail in the following sections. surface properties are of relatively minor importance
because the surface can be modified by depositing
foreign materials.

3. Base semiconductors
3.1. Important bulk properties
3.2. Grain size effects (transducer function)
Semiconductor gas sensors have been developed
Tin oxide powder for sensors is usually prepared by
based on limited kinds of n-type metal oxides such as
the hydrolysis of an aqueous solution of SnCl4 with an
SnO2 ; In2 O3 ; WO3 ; ZnO, and Fe2 O3 : In particular,
aqueous NH3 solution. The resulting precipitate of tin
SnO2 has been widely used. Semiconductor gas sensors
oxide hydrate is washed, dried, and calcined at a
would not have advanced so much if SnO2 were not
prescribed temperature. Tin oxide undergoes crystal
present. This is rather surprising because there are so
growth with a rise in calcination temperature. As shown
many kinds of semiconducting oxides ranging from n-
by the solid line in figure 4, the mean diameter ðdm Þ of
type to p-type, which can interact with various target
gases as widely experienced in catalytic chemistry. It is
suspected that the mobility of main carriers is of
primary importance for base semiconductors, because
it provides the proportionality constant of the change of
electrical conductivity when the number of main carriers
changes as a result of gas–solid interactions. The
mobility of conduction electrons (n-type oxides) is
reportedly 160 ðSnO2 Þ; 200 (ZnO), 100-15 ðIn2 O3 Þ; and
10 cm2 =V  s ðWO3 Þ: On the other hand, the mobility of
positive holes (p-type oxide) is usually much less, as
exemplified by the value of 0:2 cm2 =V  s of NiO. This
appears to be a reason why p-type oxides are not
utilized. The same reason is also applicable to some n-
type oxides. For example, TiO2 ; which has a low
mobility of conduction electrons ð0:4 cm2 =V  sÞ; is not
suited to gas sensors. Instead it has been utilized as a
base semiconductor for automotive air/fuel ratio
sensors.
Another important property for a base semiconduc-
tor is chemical and thermal stability under the sensor
operating conductions. In this respect, SnO2 is the most
stable oxide among the n-type oxides widely used for
semiconductor gas sensors. No doubt, high mobility of
electrons and high stability have led to SnO2 being so
important a base semiconductor for gas sensors. Surface Figure 4. Grain size of SnO2 after calcination at various temperatures
chemical properties are also important in some cases, for several SnO2 -based powders, pure and impregnated with foreign
especially when the base semiconductor is utilized metal oxides (5 wt%).
66 N. Yamazoe, et al./Oxide semiconductor gas sensors

Figure 6. Influence of SnO2 grain size on sensor response to 800 ppm


H2 and 800 ppm CO in air at 300 8C.

Figure 5. Influence of SnO2 grain size on electric resistance in dry air


ðRa Þ and in 800 ppm H2 ðRg Þ at 300 8C.
is determined also by the bulk of each particle and thus
strongly depends on dm :
As a result of such behavior, the gas response ðRa =Rg Þ
SnO2 grains increases with increasing calcination shows characteristic dependence on dm ; decreasing
temperature, to  4 nm at 400 8C and 28 nm at 900 8C sharply with an increase in dm up to  10 nm and then
[7]. The grain growth can be controlled by the becoming almost independent of dm above 15 nm, as
impregnation of foreign metal oxides, as also shown in shown in figure 6. Similar behavior is observed for the
figure 4. The grain growth is suppressed by the addition response to 800 ppm CO in air as shown in the same
of several oxides such as La, Ba, and P, while some figure. Essentially the same correlations hold for H2 and
others such as CoO promote grain growth at high CO even when dm is controlled by the impregnation of
temperature. What is important about the grain size is foreign metal oxides. Such a grain size effect has also
that the electrical resistances of the sensor under been reported for NO2 sensing by WO3 sensors [8].
exposure to air ðRa Þ and a target gas ðRg Þ undergo These results indicate that controlling dm smaller than
very characteristic changes as dm changes. Figure 5 2L is an important guideline for the design of high-
shows an example where sintered-block-type devices sensitivity sensors. However, this is not always practical,
were tested for sensing properties to 800 ppm H2 in air at because such tiny particles tend to grow during sensor
300 8C [4]. An increase in dm to 6 nm results in sharp operation at an elevated temperature, inducing a drift in
decreases of Ra and Rg ; while a further increase in dm Ra and Rg : In commercial SnO2 gas sensors, dm is
causes gradual increases of Ra and Rg : The characteristic usually larger than 15 nm so that the resistance becomes
change in behavior at a critical grain size ðdm ¼ 6 nmÞ is almost independent of such thermal growth even if it
associated with the formation of an electron-depleted occurs. Instead, the gas response is usually promoted by
space charge layer on the constituent particles. Since the the aid of foreign additives, as described below.
thickness ðLÞ of the space charge layer is kept essentially An interesting situation comes about when SnO2 is
constant, the proportion of space charge region in each doped with foreign oxides, because the impurities doped
particle changes relative to a change in dm : It is assumed can act as donors or acceptors of conduction electrons
that the critical dm value corresponds to a point where and thus can change the thickness of space charge layer
dm becomes equal to twice L: The whole region of each ðLÞ: The effects of doping SnO2 with Al2 O3 (1 wt%) or
particle is depleted of electrons for dm < 2L; while the Sb2 O5 (1 wt%) are shown in figure 7 [4]. L is increased
depletion takes place on the surface region only for by 30 times by Al doping, while it is reduced to be very
dm > 2L; as schematically shown in figure 5. In the latter small by Sb doping. As a result, the gas response vs. dm
case, the grain boundary potential (double Schottky correlations become largely different from that of the
barrier) for electron transfer is kept essentially constant undoped SnO2 : It is remarkable that much higher
with a change in dm ; making the resistance dm response to H2 is available in the usual dm region by
independent. In the former case, however, the resistance Al doping. Such improvements of response character-
N. Yamazoe, et al./Oxide semiconductor gas sensors 67

ðO2ÿ Þ may also participate in the reaction under some


conditions, because its energy state is not very different
from that of : The consumption of also results in a
decrease of resistance. It follows that and species
play the role of receptors to the inflammable gas
through their reactivity. For a target gas having a
tendency to adsorb strongly like NO2 or O3 ; the surface
adsorption sites simply act as receptors. For inflamma-
ble gases involving N, O, or functional groups, acidic or
basic oxides such as WO3 or In2 O3 often provide better
sensing characteristics than SnO2 : In most of these
cases, the surface acid–base properties play important
roles. As described later, however, more effective foreign
Figure 7. Sensor response of doped and undoped SnO2 to 800 ppm H2 receptors are often introduced on the oxide surface to
in air as correlated with grain size (300 8C).
replace these native receptors.
istics by a doping method are worth further investiga-
tion.
4. Surface modifiers (foreign receptors)
For the detection of inflammable gases, the gas–solid
3.3. Surface chemical properties (native receptors) interaction involved in the receptor function is the
catalytic oxidation of the target gas over the grain
Generally speaking, n-type oxides have far smaller surface. If the oxide has a bare surface, surface oxygen
amounts of adsorbed oxygen on the surface than p-type
(adsorbed or lattice oxygen) is utilized for the oxidation
oxides. Nevertheless, oxygen adsorption is distinctly
reaction, and the resulting decrease of surface oxygen at
observed for most of the n-type oxides. Temperature-
steady state induces a change in surface space charge
programmed desorption (TPD) chromatograms of
layer, which in turn is transduced into a change in
oxygen from SnO2 (powder) are shown in figure 8 [3].
electrical resistance as mentioned above. In many cases,
Molecular-type adsorbates, O2 ð 1 Þ and Oÿ 2 ð 2 Þ; and a however, the bare oxide surface is not active enough, so
dissociative type one, Oÿ ð Þ; are discerned to desorb
that various kinds of foreign materials are deposited on
depending on the adsorption conditions, in addition to the surface. Actually it was the finding of sensitizing
surface (lattice) oxygen, O2ÿ ð Þ; which desorbs in the
actions by noble metals like Pd that contributed much to
temperature range above 550 8C. Since gas sensors are
establishing inflammable gas sensors. There are also
usually operated at 300 8C and above, the species ðOÿ Þ
other types of sensitization, as described below.
is more important than the other oxygen adsorbates.
Under atmospheric air, SnO2 grains have species
adsorbed on the surface. On contact with an inflam-
mable gas, it is consumed by the reaction, leading to a 4.1. Promoting effects of noble metals
decrease in electrical resistance. Besides, the species The promoting effects of noble metals have been
widely confirmed in semiconductor gas sensors for
inflammable gases. Tin oxide-based sensors can also be
promoted by the addition of small amounts of various
metals such as Pt, Pd, and Ag. The metal additives affect
the sensor properties in a rather complex manner.
Generally speaking, the addition of noble metals results
in a shift of the volcano-shaped correlations between gas
response ðRa =Rg Þ and temperature toward the lower
temperature side, accompanied by an increase in the
height of the volcano top. As an example, the promoting
effect of Pd (actually PdO) on the H2 sensing properties
for SnO2 sensors is illustrated in figure 9 [9]. As shown
later, this behavior results from an increase in the rate of
oxidation of the target gas owing to the catalytic action
of noble metals. To understand the promoting effects,
however, it is necessary to consider both the catalytic
activity of the metals (or metal oxides) and the surface
Figure 8. TPD chromatograms of oxygen from SnO2 powder. properties of the oxide semiconductors.
68 N. Yamazoe, et al./Oxide semiconductor gas sensors

acceptor of electrons from the oxide, inducing a surface


space charge layer which is strongly depletive of
electrons in the oxide near the interface. When the
additive is reduced on contact with the target gas, it
relaxes the space charge layer by giving back electrons to
the oxide. The difference of work function of SnO2
between the oxidized and reduced state is often large,
bringing about a large increase in response ðRa =Rg Þ to
the gas. This type of sensitization has so far been
observed for SnO2 sensors impregnated with Ag, Pd, or
Cu. When exposed to air, all these metals form stable
metal oxides that exert an electronic interaction with
SnO2 : On exposure to H2 or H2 S; however, they are
converted more or less to metals (Ag and Pd) or sulfide
(CuS), respectively, resulting in the disappearance of the
electronic interaction. Such a change in the oxidation
state of the additive is responsible for the promotion of
the gas response.
Direct evidence supporting the electronic interactions
between SnO2 and Ag2 O; PdO, or CuO has been
obtained by X-ray photoelectron spectroscopy (XPS)
Figure 9. Responses to 194 ppm H2 in air at various temperatures for
SnO2 devices, unloaded or loaded with PdO (0.2–0.3 wt%). Loading [10]. The interaction increases the work function of
methods: (1) impregnation of PdCl2 and reduction, (2) adsorption of SnO2 ; causing the XPS binding energies (BEs) of the
colloidal Pd, (3) chemical fixation of ½PdCl4 Š2ÿ and reduction. inner orbitals of Sn and O, typically Sn3d3=2 and O1s1=2 ;
to shift downward. The shifts are as large as 0.3, 0.5, or
0.9 eV for PdO-, Ag2 O-; and CuO-loaded systems,
Two types of interactions between metal additive and
respectively. The shifts disappear when the additives
oxide semiconductor have been conceived, as shown in
are converted to metals (Pd and Ag) or sulfide (CuS) on
table 1. In the first type, an inflammable gas ðH2 Þ is
exposure to H2 or H2 S; respectively. The increases of the
activated by the metal additive and the activated
fragments (H) of the gas are spilt-over to the semi- work function of SnO2 are casually observed from the
increase in electrical resistance ðRa Þ: In all of the
conductor surface to react with the adsorbed oxygen.
promoted systems following electronic sensitization, Ra
The metal additive thus facilitates chemical reaction of
increases often markedly with the addition of promoters.
the gas on the semiconductor. This type of promoting
effect is called chemical sensitization. The spillover
phenomenon is known to occur in the catalysis of Pt-
loaded oxides so that this type sensitization has been
4.2. CuO as a unique receptor for H2 S
assumed for Pt-loaded sensors. In the second type
(electronic sensitization), on the other hand, there is no As indicated above, when dispersed in an SnO2
such mass transfer between the additive and the oxide. sensor, CuO acts as an excellent, very specific receptor to
Instead, the additive in the oxidized state acts as a strong H2 S [11]. This is illustrated in table 2. The SnO2 sensor

Table 1
Mechanism of sensitization by metal or metal oxide additive

Chemical sensitization Electronic sensitization

Model

Role of additive Activation of gas followed Acceptor of electrons


Origin of resistance change by spilling over Change in redox state of
Change in surface additive
oxygen concentration
N. Yamazoe, et al./Oxide semiconductor gas sensors 69

Table 2 acid–base properties is not limited to such cases. The


Response of CuO (5 wt%)–SnO2 sensor to various target gases in air target gas may undergo different reactions depending on
(200 8C)
the acid–base properties. For example, catalytic oxida-
Gas H2S CO i-C4H10 C2H5OH H2 tion of ethanol gas can take two routes as follows:
Conc. (ppm) 50 1000 1000 1000 800 – H2
Response (Ra/Rg) 35 000 1.3 1.2 1.9 1.0
C2 H5 OH ! CH3 CHO ! CO2 ðbasic oxideÞ
C2 H5 OH ! C2 H4 ! CO2 ðacidic oxidesÞ
loaded with CuO (5%) exhibits an extraordinarily high – H2 O
response (35 000) to 50 ppm H2 S in air at 200 8C,
whereas the response to other gases is negligible. An The response to ethanol gas is greatly promoted when
XPS study has shown the presence of a strong electronic SnO2 or In2 O3 is loaded with a basic oxide like CaO [12]
interaction between CuO and SnO2 ; as just mentioned. or La2 O3 [13]. This indicates that the first oxidation
Owing to this interaction, the sensor is maintained at a route via CH3 CHO is more important than the other for
high resistance level in air. On exposure to 50 ppm H2 S the detection of ethanol gas.
in air at 200 8C, CuO changes to CuS, accompanied by
the disappearance of the electronic interaction. It is
concluded that the chemical affinity of CuO to H2 S is 4.4. Other sensors
the origin of the extremely high response to H2 S:
Various gas sensors have been developed based on
combinations between base semiconductors and foreign
receptors. Some examples are listed in table 3 [11,14–32].
4.3. Acid–base modifiers
Among these, two cases are chosen here for further
The acid–base properties of oxide surfaces are some- explanations. Carbon monoxide (CO) is one of the most
times important for semiconductor gas sensors. If the important targets of gas sensing. Various kinds of
target gas is acidic like H2 S; for example, the gas–solid sensors have been devised for CO sensing, because
interaction would be more important when the surface is several metal oxides including SnO2 ; In2 O3 ; ZnO, and
more basic. This tendency has been confirmed for SnO2 Fe2 O3 show significant response to CO, as shown in
sensors modified with basic oxides. The importance of figure 10. However, most of them are also sensitive to

Table 3
Examples of various semiconductor gas sensors developed

Base oxide Additives Target gas Concentration range tested Ref.

SnO2 Ag (3 wt%) H2, C3H8 100–5000 ppm [14]


Pd (0.3–1 wt%)
WO3 NO2 10–800 ppm [15]
Au (0.8 wt%) 5 ppb–50 ppm [16]
WO3 NH3
Pt (0.4 wt%) 0.5–50 ppm [17]
TiO2 Ru (0.5 wt%) (CH3)3N 300 ppm [18]
WO3 Rh (0.4 wt%) 2–100 ppm [19]

WO3 Ru (0.004 wt%) NO 10–200 ppm [20]


SnO2 ZnO (3 at%) H2S, CH3SH 10 ppb–10 ppm [21]
SnO2 CuO (5 wt%) H2S 1–50 ppm [11]
SnO2 La2O3 (5 wt%) C2H5OH 100–1000 ppm [22]
SnO2 S (1 at%) þ (Pd 1 wt%) CH2FCF3 (R-134a) 5–3000 ppm [23]
In2O3 CeO2 (3 at%) 0.05–5 ppm [24]
O3
Fe2O3 (3 at%) 0.008–10 ppm [25]
Pd–SnO2–Sb SiO2 coating H2 100 ppm [26]
SnO2 0.5Pt–Al2O3 coating C3H8 5000 ppm [27]
In2O3 Rb2CO3 (5 wt%) CO 200–4000 ppm [28]
In2O3 Au (0.04 wt%)–Co3O4 (0.5 wt%) CO 200–2000 ppm [29]
Fe2O3 Pr6O11 (5 wt%) CH3SSCH3 5–50 ppm [30]
ZnO MoO3 (5 wt%) CH3COCH3 2–50 ppm [31]
WO3 (5 wt%)
ZnO Er2O3 (5 wt%) C5H11CHO 1–20 ppm [31]
Gd2O3 (5 wt%)
Bi2O3–MoO3 Bi/Mo ¼ 1.0 C3H6 20–8000 ppm [32]
70 N. Yamazoe, et al./Oxide semiconductor gas sensors

Figure 12. Oxidation of CO over In2 O3 based powders loaded with An


and/or Co3 O4 as a function of temperature (fixed-bed gas flow
reactor).
Figure 10. Sensor response of typical n-type oxides to 1000 ppm CO in
wet air at various operating temperatures. The highest response to CO is obtained when the rate of
oxidation is brought to be a moderately high tempera-
H2 so that development of CO-selective sensors is still ture of sensor operation (250 8C). Excessive addition of
required especially for applications at elevated tempera- these receptors therefore leads to an adverse effect.
ture. Such a CO sensor has been realized using In2 O3 There have been several oxide catalysts for the
and foreign receptors (Co3 O4 and Au). The response to selective oxidation of hydrocarbons. It would be of
CO at 250 8C can be promoted effectively by loading interest to know what kind of sensing properties these
In2 O3 with Co3 O4 (0.5 wt%), but further promotion is oxides exhibit to hydrocarbons. Two catalyst systems,
achieved by loading additionally with a small amount of Bi2 O3 –MoO3 ðBi=Mo ¼ 1Þ and Sb2 O4 –Fe2 O3 ðSb=Fe ¼
Au, as shown in figure 11. The response to H2 remains 2Þ; which are known to catalyze the oxidation of
very modest, giving rise to a CO-selective device. In this propylene to acrolein, have been tested for sensing
example, both Co3 O4 and Au enhance the catalytic propylene (1000 ppm) in air. As shown in figure 13, the
activity for the oxidation of CO, as seen from figure 12. Bi2 O3 –MoO3 system shows fairly high response to
propylene over the temperature range of the catalytic

Figure 11. Response vs. gas concentration correlations for devices Figure 13. Propylene sensing properties of mixed oxide catalysts for
using Co(0.5 wt%)–In2 O3 and Au(0.04 wt%)–Co(0.5 wt%)–In2 O3 selective oxidation of propylene at various temperatures. Inset shows
(250 8C). response transients at 400 8C.
N. Yamazoe, et al./Oxide semiconductor gas sensors 71

reaction, whereas Sb2 O4 –Fe2 O3 exhibits only a modest


response. In both cases, lattice oxygen is known to
participate in the oxidation reaction. However, the
resulting redox change of oxide is extended into the
oxide bulk in the former system, while it is limited within
the outer surface in the latter system. This difference
appears to result in the observed difference in gas
response.

5. Microstructure-dependent characteristics (utility


factor)
The above discussion focuses on how oxide grains or
grain boundaries respond to a target gas. Each grain or
grain boundary has been assumed as if it were located in
an open space. However, this is actually not the case
because a large number of oxide grains stack together in
the sensing body. The target gas (inflammable gas)
diffuses in the sensing body through pores, while it is
consumed by the reaction with the surface oxygen.
Upon exposure to an inflammable gas at a fixed
concentration, the target gas concentration decreases
on going inside the sensing body. The grains located
Figure 14. TEM image of SnO2 grains derived from hydrothermally
deep inside may remain intact or inaccessible for the gas treated SnO2 sol.
under certain conditions, leading to reduced gas
response. This means a decrease in utility of the sensing from XRD analysis remains as small as 8 and 13 nm
body or a decrease in accessibility of the target gas. To after calcination at 600 and 900 8C, respectively. The sol
discuss this phenomenon, one needs to know the can be used for the fabrication of SnO2 thin films by a
concentration profile of the target gas inside the sensing spin-coating method. The films after calcination at
body [6]. 600 8C consist of discrete SnO2 grains fairly uniform in
It is known that the mechanism of gas diffusion size ( 10 nm in diameter) and shape, as shown in figure
through a porous material depends on the size of the 1(c). The sensors using these films have shown excellent
pores involved. In the range of pore size between 2 and gas sensing characteristics. Not only is the response
50 nm in radius (conventionally called mesopores), ðRa =Rg Þ to H2 very high and sharp even without any
Knudsen diffusion prevails, while surface diffusion or
molecular diffusion does in the ranges of pore size below
2 nm (micropores) or above 50 nm (macropores), respec-
tively. The gas sensors conventionally prepared possess
pores of various sizes, making the analysis difficult. For
this purpose, one needs to have gas sensors well defined
in grain size, pore size, and thickness (gas diffusion
depth). These conditions have been met only by the thin-
film sensors fabricated from SnO2 sol.

5.1. Thin-film devices derived from SnO2 sol


An aqueous sol of SnO2 can be obtained by the
hydrolysis of SnCl4 with NH4 HCO3 in aqueous
medium, followed by a hydrothermal treatment of the
resulting precipitate (tin oxide hydrate) at 200 8C in
ammoniacal water [33]. As shown in figure 14, the sol
contains tiny grains (crystallites) of SnO2 which are
fairly uniform in size (6 nm in diameter). Notably the
SnO2 grains collected from the sol are far more resistant
to thermal growth than those prepared conventionally. Figure 15. Thermal growth of SnO2 derived from two origins. (a)
As shown in figure 15, the mean grain size as evaluated SnO2 hydrate gel (untreated); (b) hydrothermally treated sol of SnO2 :
72 N. Yamazoe, et al./Oxide semiconductor gas sensors

promoter added, as shown in figure 16, but also very


stable gas sensing properties are achieved without any
aging treatment. The response to a fixed concentration
(800 ppm) of H2 decreases with increasing thickness (80–
300 nm) of the film as plotted in figure 17. Obviously,
this decrease in response reflects the effects of diffusion
and reaction of the target gas. The responses to CO
(800 ppm) for the same devices are lower than the
responses to H2 , as also shown, although their correla-
tion with film thickness is less obvious due to the scatter
of data.

5.2. Concentration profiles of target gas


Let us consider a porous thin film (thickness L)
exposed to target gas A (figure 18). The target gas
molecules diffuse in the film while they are consumed
continuously for the surface reaction. We assume that
the film has uniform pores with a representative radius Figure 17. Sensing properties of SnO2 sol-derived thin films to H2 or
(r) for which Knudsen diffusion is applicable. If the CO (800 ppm in air) as a function of film thickness (350 8C).
surface reaction follows first-order kinetics, the gas
concentration should fulfill the following diffusion- molecular weight of the gas ðMÞ, and temperature ðTÞ,
reaction equation: being given by DK ¼ 4r=3ð2RT=MÞ1=2 . Strictly speak-
ing, the concentration of oxygen inside the film should
@CA @2 CA decrease due to the surface reaction, but the decrease is
¼ DK ÿ kDK ð1Þ assumed to be negligible because of its abundance (21%)
@t @x2
Here, CA is the concentration of target gas, t the time,
DK the Knudsen diffusion coefficient, x the distance
(depth) from the top surface of the sensing layer, and k
the rate constant. DK is a function of pore size ðrÞ,

Figure 18. Model of a gas sensing film.

Figure 16. Dependence of response to H2 for SnO2 thin-film device


derived from hydrothermally treated SnO2 sol and response transients Figure 19. Gas concentration profiles inside a porous sensing film at
to switching on and Off 800 ppm H2 (350 8C). various values of mðm ¼ Lðk=DK Þ1=2 Þ:
N. Yamazoe, et al./Oxide semiconductor gas sensors 73

compared with the target gas (up to a few thousand that in air, 0 , is linear to the target gas concentration
ppm). The effects of reaction products are also assumed ðCA Þ; i.e.
to be negligible.
At the steady state, @CA =@t ¼ 0. If boundary condi- ðxÞ ¼ 0 ð1 þ aCA Þ ð3Þ
tions are set as CA ¼ CA;s at x ¼ 0 (surface) and
@C=@x ¼ 0 at X ¼ L (bottom), equation (1) is solved Here a is a constant, defined as gas response coefficient.
and the solution is expressed as follows: The gas response ðSÞ is expresed as follows:
sffiffiffiffiffiffiffi
CA coshð1 ÿ x=LÞm k Ra aCA
¼ ; m¼L ð2Þ S¼ ¼1þ tanh m ð4Þ
CA;s coshm DK Rg m

Here CA;s is the target gas concentration outside the Figure 20 illustrates the dependence of S on L up to
film. 300 nm by taking ðk=DK Þ1=2 as a parameter. S decreases
As shown in figure 19, the gas concentration profile with increasing L; and its tendency becomes more
depends on the magnitude of m. When m is small marked as ðk=DK Þ1=2 increases. In other words, the
ðm < 1Þ; CA at the bottom of the film is close to CA;s . highest response is obtained only when L is very small if
However, as m increases beyond unity, CA inside the ðk=DK Þ1=2 takes a significant value; thickening the film
film decreases sharply, and, ultimately at large m; CA easily leads to a decrease in S down to 1 (totally
reaches zero except for a shallow region near the insensitive).
surface. The penetration depth of the gas is reduced in Fitting of these correlations to the observed response
this way as m increases, as is well known in the field of to H2 (800 ppm in air) for the thin-film sensors (figure
heterogeneous catalysis. This means a reduction in the 17) is made in the same figure. The observed data fall
utility of the film for gas sensing, and naturally this leads fairly well on the correlation for ðk=DK Þ1=2 ¼ 10ÿ2 nmÿ1
to a reduction in gas response as described in the next if a is assumed to 1 ppmÿ1 : Remarkably the data for CO,
section. although scattered rather extensively, tend to fall near
the correlation for ðk=DK Þ1=2 ¼ 3  10ÿ2 nm when the
same value of a is assumed. The Knudsen diffusion
coefficient of CO, DK ðCOÞ;
p is about 1/4 of that of
5.3. Thickness dependence of gas response H2 ðDK ðH2 Þ=DK ðCOÞ ¼ 28=2Þ: If the other conditions
are the same, ðk=DK Þ1=2 should be about twice as large
Let us assume that the film is a uniform stack of for CO than for H2 : The value of ðk=DK Þ1=2
infinitesimally thin sheets, with electric conductance, ð3  10ÿ2 nmÞ giving the best fit to the CO response
ðxÞ; where x is depth from the surface. The conduc- data is 3 times larger than that to the H2 response data
tance of the whole film is then given by integrating ðxÞ ð10ÿ2 nmÞ; in fair agreement with the estimation. The
over the whole range of xðx ¼ 0–LÞ: An important molecular weight of the target gas influences gas
assumption is introduced here that the sheet conduc- response through diffusivity in this way. This explains
tance under exposure to the gas, ðxÞ; normalized by

Figure 20. Dependence of sensor response to H2 and CO on film Figure 21. Dependence of sensor response on temperature at various
thickness at various values of ðk=DK Þ1=2 : film thicknesses (simulated under assumptions).
74 N. Yamazoe, et al./Oxide semiconductor gas sensors

qualitatively why the CO response is usually lower than


the H2 response.

5.4. Temperature dependence of gas response


The above discussion on gas response ðSÞ can be
extended, under several assumptions, to the case where
the operating temperature ðTÞ is varied. Figure 21
illustrates the results of such simulations on how S
depends on T at various values of film thickness ðLÞ: It is
seen that, at any value of L; S goes through a maximum
on changing T; while the maximum values of S as well
Figure 22. Schematic drawing for microstructure of thick film or block
as the temperature at the maximum clearly decrease with
device.
increasing L: Roughly speaking, the increase of S (the
left side of the maximum) results from an increase in the
molecules inside secondary particles (utility) becomes far
rate of surface reaction of the target gas, while the
more limited than the case of thin films. Obviously, a
decrease of S (the right side) results from a decrease in
decrease in secondary particle size as well as an increase
the utility of the gas sensing body. At the temperature of
in the size of the mesopores are the way to obtain a
the maximum response, the target gas molecules have
higher gas response for these sensors. Recently highly
optimum reactivity for the diffusion in the whole sensing
sensitive gas sensors have been developed in line with
body (large utility) as well as for exerting sufficiently
this design principle [34,35].
large interaction with the surface (large gas response
coefficient). This explains qualitatively why the correla-
tions between S and T take a volcano shape for actual
6. Concluding remarks
sensors.
Although semiconductor gas sensors have long been
regarded to be ‘‘black-box’’ technology, their funda-
5.5. Extension to non-uniform sensing bodies mentals have now been understood fairly well. In this
article, the concepts of sensor design derived from the
As indicated above, the sensing body should be
fundamentals have been described. These concepts are
porous enough and thin enough in order to obtain a
basically concerned with the gas response. Although
high gas response. This is particularly important for the
these are also applicable in principle to the gas
target gas with large molecular weight or large
selectivity, the selectivity is generally a more difficult
reactivity. As the porosity tends to increase with an
subject in sensor design, because now one has to tackle
increase in grain size, higher response can be expected
the responses to two or more different gases simulta-
with larger oxide grains, except for the range of grain
neously. In case the selectivity to a target gas cannot be
size smaller than twice the thickness of space charge
improved well over a disturbing gas, approaches for
layer. This is just opposite to what has often been
selective inhibitors to the disturbing gas or molecular
assumed in the literature. The method and condition of
sieving materials for the target gas seem to be
oxide powder processing are thus very important for the
necessary. Alternatively, simultaneous sensing of gas-
control of the microporous structure of sensing bodies.
eous mixtures by an array of plural sensors would be
Thick films and sintered blocks possess far more
useful.
complex microporous structure than the spin-coated
thin films described above. Nevertheless, almost the
same conclusions are likely to hold for these cases, if the
film thickness is replaced by the secondary particle size. References
In these sensing bodies, oxide grains (primary particle) [1] T. Seiyama, A. Kato, K. Fujiishi and M. Nagatani, Anal. Chem.
gather together to form secondary particles of 1–10 m 34 (1962) 1502.
in diameter, leaving macropores among them, as shown [2] N. Taguchi, Patent, 45-38200 (1962).
[3] N. Yamazoe, J. Fuchigami, M. Kishikawa and T. Seiyama, Surf.
in figure 22. Gas diffusion through macropores (mole- Sci. 86 (1979) 335.
cular diffusion) is rapid so that the target gas can reach [4] C. Xu, J. Tamaki, N. Miura and N. Yamazoe, Sensors Actuators
the surface of each secondary particle very easily. In B 3 (1991) 147.
addition, gas diffusion into secondary particles is [5] N. Yamazoe, Y. Kurokawa and T. Seiyama, in: Proc. of Int.
Meet. Chemical Sensors (1983) p. 35.
essentially very similar to that into thin films because
[6] G. Sakai, N. Matsunaga, K. Shimanoe and N. Yamazoe, Sensors
the mesopore structure should be almost the same. Since Actuators B 80 (2001) 125.
the sizes of secondary particles are usually far larger [7] C. Xu, J. Tamaki, N. Miura and N. Yamazoe, J. Mater. Sci. 27
than the thicknesses of thin films, the accessibility of gas (1992) 963.
N. Yamazoe, et al./Oxide semiconductor gas sensors 75

[8] J. Tamaki, Z. Zhang, K. Fujimori, M. Akiyama, T. Harada and [21] T. Nakahara, K. Takahata and S. Matsuura, Proc. Symp.
N. Yamazoe, J. Electrochem. Soc. 141 (1994) 2207. Chemical Sensors (1987) 55.
[9] S. Matsushima, T. Maekawa, J. Tamaki, N. Miura and N. [22] T. Maekawa, J. Tamaki, N. Miura and N. Yamazoe, Sensors
Yamazoe, Nippon Kagakukaishi 1991 (1991) 1677. Actuators B 9 (1992) 63.
[10] S. Matsushima, Y. Teraoka, N. Miura and N. Yamazoe, Jpn. J. [23] T. Nomura, Y. Matsuura, K. Takahata and S. Matsuura, Dig.
Appl. Phys. 27 (1988) 1798. 11th Chemical Sensor Symp. (1989) 13.
[11] T. Maekawa, J. Tamaki, N. Miura and N. Yamazoe, Chem. Lett [24] T. Takada and K. Komatsu, Dig. Tech. Papers of Transducers
(1991) 575. (1987) 693.
[12] S. Matsushima, T. Maekawa, J. Tamaki, N. Miura and N. [25] T. Takada, K. Suzuki and M. Nakane, Sensors Actuators B 13–14
Yamazoe, Chem. Lett. (1989) 845. (1993) 404.
[13] J. Tamaki, T. Maekawa, S. Matsushima, N. Miura and N. [26] Y. Okayama, Proc. 6th Sensor Symp. (1986) 101.
Yamazoe, Chem. Lett. (1990) 477. [27] N. Yamazoe, Y. Muto and T. Seiyama, Hyomen Kagaku 5 (1984)
[14] N. Yamazoe, Y. Kurokawa and T. Seiyama, Sensors Actuators 4 55.
(1983) 283. [28] H. Yamaura, J. Tamaki, K. Moriya, N. Miura and N. Yamazoe,
[15] M. Akiyama, J. Tamaki, N. Miura and N. Yamazoe, Chem. Lett. J. Electrochem. Soc. 143 (1996) L36.
(1991) 1611. [29] H. Yamaura, J. Tamaki, K. Moriya, N. Miura and N. Yamazoe,
[16] T. Maekawa, J. Tamaki, N. Miura and N. Yamazoe, Chem. Lett. J. Electrochem. Soc. 144 (1997) L158.
(1992) 639. [30] Y. Anno, J. Tamaki, Y. Asano, K. Hayashi, N. Miura and N.
[17] T. Maekawa, J. Tamaki, N. Miura and N. Yamazoe, in: New Yamazoe, Hyomen Kagaku 16 (1995) 474.
Aspects of Spillover Effect in Catalysis, eds. T. Inui, K. Fujimoto, [31] Y. Anno, T. Maekawa, J. Tamaki, Y. Asano, K. Hayashi, N.
T. Uchijima and M. Masai (1993) p. 421. Miura and N. Yamazoe, Sensors Actuators B 24–25 (1995) 623.
[18] Y. Shimizu, M. Egashira and Y. Takao, J. Electrochem. Soc. 135 [32] T. Jinkawa, Master thesis.
(1988) 2539. [33] N.S. Baik, G. Sakai, N. Miura and N. Yamazoe, J. Am. Ceram.
[19] Y. Anno, T. Maekawa, J. Tamaki, Y. Asano, K. Hayashi, Soc. 83 (2000) 2983.
N. Miura and N. Yamazoe, Sensors Materials 5 (1993) [34] M. Shoyama and N. Hashimoto, Proc. Chem. Sensor Symp. 17,
135. Suppl. B (2001) 10.
[20] M. Akiyama, Z. Zhang, J. Tamaki, N. Miura, N. Yamazoe and T. [35] Y.G. Choi, G. Sakai, K. Shimanoe, N. Miura and N. Yamazoe,
Harada, Sensors Actuators B 13–14 (1993) 619. Sensors Actuators B (in press).

Potrebbero piacerti anche