Journal of Alloys and Compounds

Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Journal of Alloys and Compounds 539 (2012) 205209

Contents lists available at SciVerse ScienceDirect

Journal of Alloys and Compounds


journal homepage: www.elsevier.com/locate/jalcom

Quantum-sized ZnO nanoparticles synthesized in aqueous medium for toxic


gases detection
Dianqing Li a, Jingwei Hu a,c, Faying Fan a, Shouli Bai a,, Ruixian Luo ,a, Aifan Chen a, Chung Chiun Liu b
a

State Key Laboratory of Chemical Resource Engineering, Beijing University of Chemical Technology, Beijing 100029, China
Department of Chemical Engineering, Case Western Reserve University, Cleveland, OH 44106, USA
c
Beijing Institute of Aerospace Testing Technology, Beijing 100074, China
b

a r t i c l e

i n f o

Article history:
Received 5 April 2012
Received in revised form 23 May 2012
Accepted 25 May 2012
Available online 7 June 2012
Keywords:
Quantum-size ZnO
Aqueous medium
Toxic gases

a b s t r a c t
Quantum-sized ZnO nanoparticles have been prepared by the solgel method in aqueous medium using
sodium dodecyl sulfate (SDS) as surface modier to control the growth of ZnO particles and Ostwald ripening, because the majority of bioanalyses require water-stable materials. The crystal structure, grain
size and optical property of the ZnO nanoparticles have been characterized using X-ray diffraction
(XRD), high resolution transmission electron microscopy (HRTEM) and X-ray photoelectron spectroscopic
(XPS). The obtained quantum-sized ZnO nanoparticles as a novel sensing material have been used to
detect NO2 in environment. The sensing tests indicate that the ZnO sensing material not only has high
response to NO2 but also exhibits high selectivity to CO and CH4 at optimum operating temperature of
290 C. The sensing response depends on the calcination temperature of the ZnO precursor and operating
temperature of the sensor. The intrinsic sensing property has also been exploited by photoluminescence
(PL) and Raman spectra.
2012 Elsevier B.V. All rights reserved.

1. Introduction
ZnO is an important semiconductor material with extensive
application in electronic, sensor and photoelectronic device owing
to its wide band gap of 3.37 eV, large exciton binding energy of
60 meV at room temperature, and excellent chemical and thermal
stability. Recently, one-dimensional ZnO nanostructures such as
nanorods, nanobelts, nanowires, etc. have been widely investigated
and applied [14]. The last year, many works have been directed to
investigate the optical and electrical properties of quantum-sized
ZnO nanoparticles with large specic surface and quantum connement effect. Thus, various synthesis methods of quantum-sized ZnO
have been reported, such as vapor condensation, thermal evaporation, spray pyrolysis, magnetron sputtering, metal organic chemical
vapor deposition, and solgel, etc. [59]. Among them, the solgel
route has attracted much attention due to their low treatment temperature, low cost and high efciency. However, the quantum-sized
ZnO is usually synthesized in organic solvent, such as methanol and
ethanol, while the majority of bioanalysis require water stable
material, therefore, the development of synthesizing quantumsized ZnO with narrow size distribution and stable in aqueous dispersion is still a challenge and far from successful [10]. Meanwhile,

Corresponding author. Tel.: +86 10 64444903.


E-mail address: slbai@mail.buct.edu.cn (S. Bai).
0925-8388/$ - see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jallcom.2012.05.106

colloidal ZnO nanoparticles thus obtained tend to aggregate or undergo Ostwald ripening owing to their high surface energy [11]. To
stabilize ZnO nanoparticles in aqueous medium, various capping
agents have been employed, such as polyvinylpyrolidone (PVP),
amines, 3-aminopropyl trimethoxysilane, long chain aliphaticthiols, etc. [1215]. In this study, quantum-size ZnO has been successfully prepared by the controlled hydrolysis of Zn(Ac)22H2O under
room temperature, sodium dodecyl sulfate (SDS) is used as a capping agent to control the particle growth of ZnO. The obtained
ZnO as a novel sensing material is used in gas sensors to detect
the toxic gases such as NO2, CO and CH4, and to investigate the effect of ZnO particle size on the sensing property. It is well known
that air pollution has been more and more serious with the development of industry and the increase of automobile. Especially,
nitrogen dioxide (NO2) is toxic itself, and furthermore, is a precursor of acid rain and photochemical smog [16]. In order to detect
such hazardous gas, there have been lots of efforts in developing
NO2 gas sensor, while the application of quantum size ZnO in gas
sensor is rarely reported in literatures. Recently, Nath et al. synthesized the ZnO quantum dot via quenching technique and investigated the sensing property for acetone [17]. Another article is
reported by Forleo et al., they obtained ZnO quantum dots by alkaline-activated hydrolysis and condensation of zinc acetate solutions, and is used in investigating of sensing properties for
acetone and NO2 [18].

206

D. Li et al. / Journal of Alloys and Compounds 539 (2012) 205209

2. Experimental
2.1. Synthesis of quantum-sized ZnO nanoparticles
All chemical reagents used in the experiment were analytical grade without further purication. Aqueous solution of ammonium bicarbonate (1 mol/L) and 12.5 lL
of aqueous solution of SDS (1 mol/L) were rst mixed together in one beaker. Under
constant magnetic stirring, 25 mL aqueous solution of zinc chloride (1 mol/L) was
added slowly (dropwise for 10 min) into the resulting solution. During the dropping
process, the white precipitation was gradually formed in the solution. After 2 h
reaction, the product was ltered and washed with anhydrous ethanol several
times to remove the impurities, followed by drying at 50 C for 4 h in air. Finally,
ZnO powder with a light yellow color was obtained after annealing the precursor
in air at 200 C for 1 h in the chamber furnace.

2.2. Characterization of quantum-sized ZnO nanoparticles


The crystal structure of the ZnO powder was analyzed by X-ray diffraction with
Rigaku D/max-Ultima III X-ray powder diffractometer, operated at 45 kV and
40 mA, with a scan speed of 10 min1 for 2h in a range from 20 to 80, using Ni-ltered Cu Ka radiation, k = 0.15406 nm. Diffraction peaks of the crystalline phase
were compared with those of the standard compounds reported in the JCPDS data
les. The average crystallite size was estimated by Deby-Scherrer equation. Meanwhile, high resolution transmission electron microscopy (HR-TEM, JEM-3010) was
used to determine the morphology and size of ZnO. For HR-TEM experiment, the
specimen was prepared by depositing a drop dilute solution of the sample in ethanol on a carbon coated copper grid and drying at room temperature. Fourier transform infrared (FT-IR) spectra of quantum size ZnO were recorded using Nicolet 380
spectrometer from 4000 to 400 cm1 at room temperature. Surface elemental analysis was performed using an ESCALAB250 X-ray photoelectron spectroscopy (XPS).
UVVis spectrometer (JASCO V-570) was used to record the UVVis absorption
spectra. HeCd laser with 325 nm emission of RF-5301PC spectrometer was used
for photoluminescence (PL) excitation.

2.3. Gas sensing measurement of quantum-sized ZnO nanoparticles


In order to measure the gas sensing properties for quantum size ZnO nanoparticles, the sensing elements based on ZnO powder calcined at 200800 C were fabricated in a conventional way [19]. The calcined powder was pressed into pellet
under a pressure of 8 MPa and gold leads were mounted on both sides of the pellet
to form the sensing element. The element was enclosed in a quartz tube which was
placed in a temperature-controlled tubular furnace to heat the element. The element was heated to 300 C for several hours to remove the adsorbates from the element surface before testing element performance. To evaluate the dependence of
element sensing property on the temperature, the element resistances were measured as a function of operating temperatures in range of 100400 C. The target
gas and air ows were then introduced through gas ow meters to control their
ow rate. A constant current was applied across the element. The steady-state resistance of the element was measured in air (Ra) and in air-gas mixture (Rg), respectively. The response of the oxidizing gases, such as NO2, is dened as the ratio of the
sensor resistance in airNO2 mixture to that in an air (Rg/Ra), while the response of
reducing gases, such as CO is dened as the ratio of the sensor resistance in air to
that in an airCO mixture (Ra/Rg).

3. Results and discussion


3.1. Structure and size of quantum-sized ZnO nanoparticles
Fig. 1 shows a typical HR-TEM micrograph of the ZnO nanoparticles calcined at 200 C for 3 h, these nanoparticles are essentially
little aggregate but appear to be spherical in shape individually.
The average particle size is observed to be 6 nm. The SDS as a surface modier may have modied the Ostwald ripening kinetics so
that the growth rate of ZnO nanoparticles is decreased. This will
narrow the size distribution of ZnO nanoparticles effectively. The
inset shows a Fourier transform diffraction pattern from the
[0 0 1] zone axis of the single ZnO crystalline particle marked as
an arrow in Fig. 1. The interplanar spacing is 0.28 nm as illustrated
in image of Fig. 1, which corresponds to the {1 0 0} plane of hexagonal ZnO nanocrystals.
Fig. 2 demonstrates the XRD patterns of the ZnO nanoparticles
calcined at different temperatures range from 200 to 800 C for
3 h. The diffraction peaks and their relative intensities of spectra
are all in agreement with JCPDS card No. 361451, so that the

observed patterns can be unambiguously attributed to the presence


of hexagonal wurtzite crystallites with cell constants of a = 3.251
and c = 5.208 . No excess peaks are detected, which indicates that
no other complex products are formed. It should be noted here that
the full width at half maximum (FWHM) of the diffraction peaks decrease with increasing the calcination temperature of ZnO precursor due to aggregation of part crystallites. The peak broadening in
the diffraction patterns mainly consists of four factors: microstrains
(deformations of the lattice), faulting (extended defects) and crystalline domain size as well as domain size distribution. If the strains
and faulting can be ignored, the peak broadening only consists of
the crystalline domain size and distribution. Then the average crystallite size (D) is estimated to be 4.5, 7.4, 9.5 and 11.3 nm by the DebyeScherer formula for ZnO nanoparticles calcined at
temperatures of 200, 400, 600, and 800 C, respectively. The statistical result is consistent with the observation result from HR-TEM.
FT-IR analysis is a useful way to study the surface chemical structure of the sample. Fig. 3 shows the FT-IR spectrum of ZnO nanoparticles, the peaks at around 2920, 2852, 1468 and 1083 cm1 are due
to CH stretching and bending. A broadening peak at around
3440 cm1 is proposed to be due to HOH stretching, the peak at
470 cm1 is a characteristic peak of ZnO stretching. The stretching
mode of vibration corresponding to C@C is obtained at 1509 cm1
[20]. The band at around 1200 cm1 is assigned to an S@O stretching
vibration of SO4 from the SDS molecule [21], and is not observed in
the spectrum of non-modied ZnO, which demonstrates the successful modication of SDS surfactant on the surface of ZnO.
XPS analysis is performed in order to illuminate the surface
composition and the chemical state of the elements existed in
the sample. Fig. 4a shows the XPS deconvoluted spectra of Zn 2p
3/2 core level for the sample. In all cases the peak is deconvoluted
into two peaks, the dominant peak at 1021.1 eV is associated with
the Zn2+ in ZnO wurtzite structure and the weaker peak at
1023.3 eV is associated with zinc hydroxide specie [22]. The deconvolution of XPS spectra for the O1s core level lines shows the presence of three different peaks (Fig. 4b). The main peak centred at
530.7 eV is attributed to the O1s level in the ZnO wurtzite structure
surrounded by the Zn atoms with their full complement of nearestneighbor O2 ions. The peak located 531.6 eV is associated with the
O1s binding energy in the zinc hydroxide specie, which is in agreement with the O1s binding energy in other metal hydroxides and
oxyhydroxides. The peak at 532.5 eV is associated with adsorbed
or chemisorbed oxygen species such as Zn(OH)2, O2, H2O due to
surface hydroxyl groups. This suggests that besides chemisorbed
hydroxyl species Zn(OH)2 is also presented on the surface of ZnO
polar face. However, the Zn(OH)2 is presented in low concentration
and only on the surface of ZnO, since no this compound is observed
by XRD analysis. In addition, the relative abundance of Zn(OH)2 obtained from the Zn 2p 3/2 spectrum, using the integrated peak area
and atomic sensitive factors, indicating that the as-synthesized
ZnO nanoparticles contains a low concentration of Zn(OH)2 and
the atomic ratio of Zn and O is nearly 1:1.
Raman spectrum of the as-synthesized ZnO is illustrated in
Fig. 5. ZnO has wurtzite symmetry with C6m or 6 mm symmetry.
There are four atoms per unit cell leading to 12 phonon branches,
nine optical and three acoustic [23]. As shown in Fig. 5, the peak at
439 cm1 is attributed to ZnO nonpolar optical phonons E2 mode,
which is typical one of Raman active branches. The peak at
349 cm1 corresponds to A1 mode; the peak at 581 cm1 is attributed to ZnO E1 (LO) mode. Clearly, it also demonstrates that the
sample is composed of ZnO with hexagonal structure.
3.2. Gas sensing properties of quantum-sized ZnO nanoparticles
The sensing material has to be calcined at an appropriate temperature to achieve the stable crystal structure and gas sensing

207

D. Li et al. / Journal of Alloys and Compounds 539 (2012) 205209

30

Frequency (%)

25

20

15

10

0
4

Particle size (nm)


Fig. 1. HR-TEM image and particle size distribution of the as-synthesized ZnO. The left upper inset is a fast fourier transform (FFT) pattern, the 0.28 nm as illustrated
corresponds with the d-spacing of ZnO.

Fig. 2. XRD patterns of ZnO calcined at different temperatures: (a) 200, (b) 400, (c)
600 and (d) 800 C.

Fig. 3. FT-IR spectra of the quantum-sized ZnO nanoparticleswith and without SDS
modied.

property. Fig. 7 shows the operating temperature dependence of


the gas response to 40 ppm NO2 in air for ZnO nanoparticles calcined at 200, 400, 600 and 800 C for 2 h, respectively. From
Fig. 7, when the operating temperature is under 290 C, the response increase with increasing the operating temperature can
be attributed to the fact that the thermal energy is high enough
to overcome the activation energy barrier to the reaction and a signicant decrease in electron concentration result from the sensing
reaction. When the operating temperature increases to 290 C, the
response reaches a maximum value of 220 and the temperature at
which the response exhibits a maximum value is termed as optimum operating temperatures. If the temperature is further enhanced above 290 C, the amount of adsorbed gas on material
surface will decrease with temperature, leading to the response decrease. Fig. 7 also indicates the effect of calcination temperatures
on gas response at operation temperature of 290 C. The results
indicate that the sensing response of the sample calcined at
400 C is exceptionally higher than the samples calcined at 200,
600 and 800 C. It is not only attributed to become better of ZnO
crystalline (the XRD proles become stronger and sharper with
increasing calcination temperature), but also attributed to the increase of ZnO stoichiometric defects. The electron concentration
of metal oxide semiconductor is determined mainly by the concentration of stoichiometric defects such as oxygen vacancies, the
annealing in air render ZnO more oxygen vacancies. The non-stoichiometry of the ZnO nanoparticles has been conrmed by Raman
and PL analysis. As observed in Fig. 6, the PL intensity of ZnO calcined at 400 C is lower than that of calcined at 600 C, the lower
PL intensity indicates a lower recombination rate of photo-generated electrons and holes, that is, the ZnO calcined at 400 C presents a higher photonic efciency than that of calcined at 600 C.
Because the PL emission originate from the recombination of excited electrons and holes, the higher photonic efciency result
from the improvement of the separation between electrons and
holes effectively, which effectively promote the sensing response
and photocatalysis of ZnO nanoparticals. However, when calcination temperature is lower than 400 C, the ZnO has a small crystallite size but it may not satisfy the requirement of crystallinity for
the sensing material, so, the lower response is observed. When calcination temperature is too high, the crystallite size of ZnO become
larger and therefore the specic surface become smaller which will
lead to the decrease of the gas response. Although the quantum-

208

D. Li et al. / Journal of Alloys and Compounds 539 (2012) 205209

Fig. 4. High-resolution (a) Zn 2p 3/2 and (b) O 1s XPS spectra for the quantum-sized ZnO nanoparticles.

Fig. 5. Raman spectrum of as-synthesized quantum-sized ZnO nanoparticles.

temperature is unfavorable. The reason is that the target gas molecules inltrated into the agglomerates and the active sites of surface adsorption are decreased in the case of the agglomerate
structures, lead to response decrease [24]. According, calcination
temperature of 400 C is found to be the appropriate calcination
temperature for ZnO nanoparticles, it ensures ZnO nanoparticles
has a high stoichiometric defect and crystallinity, and without signicant specic surface decrease, thereby achieving a maximum
response, which is in good agreement with the observed gas-sensing results.
The selectivity is one of important gas sensing properties for
sensing materials. The cross sensing between NO2 and reducing
gases such as CH4 and CO is still one of the major problem for practical NO2 sensor. Fig. 8 shows the selectivity of NO2 to CO and CH4
in same gas concentration for the quantum-sized ZnO nanoparticles. The highest response reaches 32.0 and 28.9 to CO and CH4,
respectively, the highest selectivity reaches 6.9 and 7.6 to
40 ppm CO and CH4, respectively, at operating temperature of
290 C. It indicates that the quantum-sized ZnO not only has high
response to NO2 but also exhibits high selectivity to reducing
gases, such as CO and CH4. So, the quantum-sized ZnO synthesized
by SDS modied solgel process is a promising sensing material for
NO2 detection.

Fig. 6. Room-temperature PL spectrum of quantum-sized ZnO nanoparticles


calcined at different temperatures: (a) 200, (b) 400 and (c) 600 C.

sized ZnO nanoparticles has large specic surface, which can enhance the response of sensing material, annealing them at high

Fig. 7. Effect of calcination temperature on gas response to 40 ppm NO2: (a) 200, (b)
400, (c) 600 and (d) 800 C.

D. Li et al. / Journal of Alloys and Compounds 539 (2012) 205209

209

3.3. Gas sensing mechanism of quantum-sized ZnO sensing material

4. Conclusion

It is well accepted that the sensing response of semiconductor


metal oxide is attributed to the chemisorption of oxygen on the
oxide surface and the subsequent reaction between adsorbed oxygen and tested gas, which causes the resistance change. The adsorption and reaction process can be described as follows: When ZnO
sensing material is exposed to air, oxygen molecules adsorb on the
surface of the material to form adsorbed oxygen ions (O2, O,
O2) by capturing electrons from the conductance band of material.

The solgel method in aqueous solution is used to prepare


quantum-sized ZnO nanoparticles. The technique is simple and
inexpensive, and it may be useful for the production of other quantum-sized metal oxides. XRD analysis conrms that the ZnO has
hexagonal structure and the average crystallite size estimated by
Scherer formula is 4.5 nm. The HR-TEM image shows roughly
spherical particles with average crystallite size of 6 nm for ZnO.
The UVVis absorption spectrum illustrates that the band gap energy has a little blueshift compared to bulk ZnO due to quantum
connement effect, while the Raman and PL spectra illustrate that
the intrinsic gas sensing arises from non-stoichiometry of ZnO
semiconductor. The quantum-sized ZnO nanoparticles not only
exhibits high response to NO2 but also exhibits high selectivity to
CO and CH4 in same gas concentration at optimum operating temperature of 290 C. From gas sensing measurement, the increase in
the NO2 response with increasing the operating temperature can
be attributed to the fact that the thermal energy is high enough
to overcome the activation energy for the chemisorption and reaction between oxygen and test gas, while the effect of the calcination temperature on response can be attributed to the effect of
temperature on grain size and stoichiometric defect.

O2 gas ! O2 ads
O2 ads e ! O2 ads
O2 ads e ! 2O ads
O ads e ! O2 ads
When the ZnO sensing material is exposed in oxidizing gases (e.g.
NO2), the gas capture electrons from semiconductor (due to its
higher electrophilic property) and react with the adsorbed oxygen
species leading to the formation of adsorbed NO2, the process of
the reaction can be described as follows:

NO2 g e ! NO2 ads;


NO2 ads O2 ads e ! NO2 ads 2O ads;
NO2 ads 2O ads e ! NO2 g 2O2 ads

Acknowledgments

These series of reactions result in the conductance decrease of the


material (that is resistance increase), which result from the electron
concentration decrease on the material surface.
The NO2 detection is performed when the cycling reaction is
continued,

NO2 g e ! NO2 ads


When ZnO sensing material is exposed to reducing gases (e.g. CO or
CH4), the O species adsorbed on the surface of sensing material react with the test gas and release electrons to the conduction band of
sensing material, this leads to a decrease in resistance. The relevant
oxidizing reactions occur on surface of the material are shown as
follows:

CO O ! CO2 e
CH4 g 2O2 ! CO2 2H2 O 4e
These reactions can release electrons to reduce the resistance of the
material and thus the gas response is detected.

221.7

129.6

100

5.9 34.6
2.9
3.6

3.6 28.9

9.6 32

tin

40

50

em

6.5

pe

30

ure

5.6

25

20

(C
)

CH

rat

CO

gT

6.7

2
NO

era

50

38.2
9.4

Op

100

Response

200

Fig. 8. Response of quantum-sized ZnO calcined at 400 C to different tested gases.

This research work was supported by the National Natural Science Foundation (Nos. 21177007 and 51072014) and Beijing Natural Science Foundation (Nos. 8091003, 8112022 and 8102028).
References
[1] M. Arshad, A. Azam, A.S. Ahmed, S. Mollah, A.H. Naqvi, J. Alloys Compd. 509
(2011) 83788381.
[2] L.Y. Wen, K.M. Wong, Y.G. Fang, M.H. Wu, Y. Lei, J. Mater. Chem. 21 (2011)
70907097.
[3] A. Sapkota, A.J. Anceno, S. Baruah, O.V. Shipin, J. Dutta, Nanotechnology 22
(2011) 215703215710.
[4] J. Park, T. Yao, J. Alloys Compd. 513 (2012) 180183.
[5] S.L. Bai, J.W. Hu, D.Q. Li, R.X. Luo, A.F. Chen, C.C. Liu, J. Mater. Chem. 21 (2011)
1228812294.
[6] S.D. Shinde, G.E. Patil, D.D. Kajale, V.B. Gaikwad, G.H. Jain, J. Alloys Compd. 528
(2012) 109114.
[7] S. Sakohara, M. Ishida, M. Anderson, J. Phys. Chem. B 102 (1998) 1016910175.
[8] H.L. Cao, X.F. Qian, Q. Gong, W.M. Du, X.D. Ma, Z.K. Zhu, Nanotechnology 17
(2006) 36323636.
[9] D. Ma, T. Veres, L. Clime, F. Normandin, J. Guan, D. Kingston, B. Simard, J. Phys.
Chem. C 111 (2007) 19992007.
[10] D. Vollath, D.V. Szabo, S. Schlabach, J. Nanopart. Res. 6 (2004) 181191.
[11] M.K. Patra, M. Manoth, K. SinghV, G.S. Gowd, V.S. Choudhry, S.R. Vadera, N.
Kumar, J. Lumin. 129 (2009) 320324.
[12] C.L. Yang, J.N. Wang, W.K. Ge, L. Guo, S.H. Yang, D.Z. Shen, J. Appl. Phys. 90
(2001) 44894493.
[13] H. Zhang, Z. Cui, Y. Wang, K. Zhang, X. Ji, C. Lu, B. Yang, M. Gao, Adv. Mater. 15
(2003) 777780.
[14] R. Sakthi Sudar Saravanan, D. Pukazhselvan, C.K. Mahadevan, J. Alloys Compd.
517 (2012) 139148.
[15] C. Li, N. Murase, Langmuir 20 (2004) 14.
[16] K. Wetchakun, T. Samerjai, N. Tamaekong, C. Liewhiran, C. Siriwong, V. Kruefu,
A. Wisitsoraat, A. Tuantranont, S. Phanichphant, Sens. Actuators B: Chem. 160
(2011) 580591.
[17] S.S. Nath, M. Choudhury, D Chakdar, G Gope, R.K. Nath, Sens. Actuators B:
Chem. 148 (2010) 353357.
[18] A. Forleo, L. Francioso, S. Capone, P. Siciliano, P. Lommens, Z. Hens, Sens.
Actuators B: Chem. 146 (2010) 111115.
[19] S.L. Bai, L.Y. Chen, J.W. Hu, R.X. Luo, A.F. Chen, C.C. Liu, Sens. Actuators B: Chem.
159 (2011) 97102.
[20] P.S. Cale, A.O. Ribeiro, A.M. Pires, O.A. Serra, J. Alloys Compd. 344 (2002) 285
289.
[21] R. Ullah, J. Dutta, J. Hazard. Mater. 156 (2008) 194200.
[22] L. Guo, S.H. Yang, C. Yang, P. Yu, J. Wang, W. Ge, G.K.L. Wong, Appl. Phys. Lett.
76 (2000) 29012903.
[23] J.H. Yang, X.Y. Liu, L.L. Yang, Y.X. Wang, Y.J. Zhang, J.H. Lang, M. Gao, B. Feng, J.
Alloys Compd. 477 (2009) 632635.
[24] S.L. Bai, X. Liu, D.Q. Li, S. Chen, R.X. Luo, A.F. Chen, Sens. Actuators B: Chem. 153
(2011) 110116.

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy