Sei sulla pagina 1di 81

MOx/SBA-15 ()

NSC 97-2221-E-168-013 97 08 01 98 07 31

- -

98 09 28


MOx/SBA-15

NSC 97-2221-E-168-01397 08 01 98 07 31 - ()

98
1

28

Abstract Visible-light-driven silver vanadate photocatalysts were successfully synthesized using a low-temperature hydrothermal synthesis method. Under various hydrothermal conditions, the structures of silver vanadates were tuned by changing the hydrothermal time and with the assistance of a surfactant. X-ray diffraction (XRD) results reveal that the hydrothermal synthesis powders consisted of three kinds of phase: pure Ag4V2O7 or pure -Ag3VO4 or mixed phases of Ag4V2O7 and -Ag3VO4. UV-vis spectroscopy indicated that silver vanadate particles had strong visible light absorption with associated band gaps in the range of 2.2-2.5 eV. The powder synthesized at 140 for 4 h (HT4) exhibited the highest photocatalytic activity among all samples. The reactivity of HT4 (surface area, 2.04 m2 g-1) on isopropanol and benzene was 16.6 times and 16.2 times higher than those of P25 (surface area, 49.04 m2 g-1) under visible light irradiation. The enhanced photocatalytic activity of HT4 is attributed to mixed crystalline phases of Ag4V2O7 and -Ag3VO4, where -Ag3VO4 is the major component. In addition, the highest large amount of hydroxyl groups on the surface of -Ag3VO4, detected by the in-situ FT-IR diffuse reflectance (DRIFT) technique, was considered as the other factor. Detailed DRIFT experiments also revealed that the stability of Brnsted acid sites of HT4 was excellent, as confirmed by NH3 adsorption. Studies of crystalline phase and surface property showed that the photocatalytic activities of the hydrothermal synthesis of silver vanadates are strongly affected by hydrothermal time. Besides, the amount of surface hydroxyl groups and the crystallinity of -Ag3VO4 phase were verified to be the key factors influencing photocatalytic activity.

Keywords: Ag3VO4; Ag4V2O7; Visible-light driven photocatalyst; DRIFT technique; Surface hydroxyl group


(120~180) (silver vanadate)120(Ag3VO4Ag4V2O7) (Ag3VO4)180Ag3VO4 2.2 ~2.5 eV140/4h IPAP25 16.616.2(DRIFT)140 Brnsted acid() BiVO4/SBA-15 BiVO4(51020%) BiVO4 monoclinic2.44 eV298~435 m2/g 10 W154 ppm BiVO410%BiVO4 10 BiVO4SBA-15SBA-15BiVO4 5%BiVO4240 10 ppmSBA-15 BiVO4

1.

Introduction Semiconductor photocatalysts have been studied extensively for environmental

purification and for splitting water into H2 and O2. Among semiconductor photocatalysts, titanium dioxide is the most widely employed. However, titanium dioxide suffers from low efficiency under visible light illumination since it has a large band-gap energy of 3.2 eV. Therefore, numerous attempts have been made to improve TiO2 as a visible-light active photocatalyst, including anion doping, metal doping, and oxygen deficiency generation. Although the light absorption was extended to visible light region, these processes often suffered from thermal instability or involved an increase in the carrier-recombination centers [1-4]. These methods are not appropriate for developing highly efficient visible-light active photocatalysts. Recently, a band structure control method has been applied to design non-TiO2-based visible-light active photocatalysts, without any dopants. For example, monoclinic and perovskite materials, such as InMO4 (M = V, Nb, Ta) [5], BiVO4 [6], AgTaO3 [7], AgNbO3 [7], and Ag3VO4 [8], have been proven to be good visible-light responsive photocatalysts, especially in the field of photocatalytic splitting of water into H2 and O2. The valence bands of AgTaO3, AgNbO3, and Ag3VO4 consist of hybridized Ag 4d10 orbital and O 2p6 orbitals, and their conduction bands consist of hybridized Ag 5s orbitals with Ta 5d orbitals, Nb 4d orbitals, and V 3d orbitals, respectively. For these Ag-containing photocatalysts, the hybridization of the O 2p6 orbitals with the completely filled 4d10 orbitals of silver ions could form a valence band at a more positive energy level than that of O 2p6, resulting in a narrowed band gap. The most general method for synthesizing monoclinic and perovskite semiconductors involves a solid-state reaction of high purity metal oxide powders, which need to be mechanically mixed and reacted at 850 (or higher) for at least 12 h. To reduce energy

consumption, exploiting low cost metal oxide catalysts with high visible-light photocatalytic activity is desirable. In recent years, the hydrothermal synthesis method has been regarded as a convenient and practical method to prepare various metal oxides, such as ZnO [9], Fe2O3 [10], NiFe2O4 [11], and BaWO4 [12]. The hydrothermal synthesis method offers many advantages, such as controllable particle size, high degree of crystallinity, and high purity while using milder synthesis temperatures and simpler process configurations. Until now, there have been no studies on the preparation of silver vanadates using the hydrothermal synthesis method in the field of photodecomposition of organic compounds. Hence, the synthesis of silver vanadates using a one-step hydrothermal synthesis method is proposed in this study. Isopropanol (IPA) and benzene vapors are the main components of in-door air pollution and are dangerous to human health; they were thus chosen as the model pollutants to evaluate the visible-light photocatalytic activities of the as-prepared powders. The structural, textural, and photocatalytic properties of silver vanadates obtained from both hydrothermal synthesis method and the conventional calcination method were compared with those of commercial TiO2 (P-25). In particular, the surface hydroxyl groups and surface acidic properties of the hydrothermal synthesis of silver vanadates were determined using the techniques of DRIFT and temperature programmed desorption of NH3. They were correlated with the photocatalytic activities of the hydrothermal synthesis samples for IPA and benzene. 2. Experimental

2.1. Catalysts preparation The silver nitrate (AgNO3) and ammonium vanadate (NH4VO3) was used as the silver and vanadium precursors, respectively, to prepare silver vanadium oxide (SVO). In a typical procedure, 0.1 M AgNO3 (60 ml) was added dropwise to 0.1 M NH4VO3

(20 ml) under constant stirring for 30 min. The pH of the mixture was adjusted to 7 using ammonia solution. After the mixture was maintained at room temperature for 24 h, the suspended solution was poured into a 100 ml Teflon-lined stainless autoclave and heated at 140 for different hydrothermal times. These samples were denoted as HT2, HT4, HT6, and HT8, indicating the hydrothermal times of 2, 4, 6, and 8 h, respectively. After cooling to room temperature, the obtained light orange slurry was filtered, washed several times using distilled water to remove NO3 and NH4+ residues, and dried at 110 for 6 h. To prepare various morphologies of SVO, we used cetyltrimethylammonium bromide (CTAB, C19H49BrN) as the structure-directing agent. At a ratio of CTAB to silver of 0.05:1, AgNO3 and CTAB were dissolved in 180 ml of de-ionized water and stirred for 1 h, with the AgNO3 concentration kept at 0.03 M. Then, the transparent solution was added dropwise to 60 ml of 0.03 M NH4VO3 under continuous stirring, which was followed by the hydrothermal synthesis procedures. The resulting products were washed, first with water and then with absolute ethanol in

centrifugation/dispersion cycles 3 times to remove the remaining CTAB, NO3, and NH4+ residues. Finally, the material was dried in an oven at 100 for 6 h. The samples were denoted as CTAB-SVO. The other batch of SVO samples, prepared using a precipitation method and calcined at 380 for 4 h in air, was designated as PC. 2.2. Sample characterization The X-ray diffraction (XRD) patterns of the powders were measured using an X-ray diffractometer (Panalytical XPert PRO) with Cu radiation (= 0.15418 nm) in the 2-range from 10 to 60. The surface morphology was examined using a Field Emission Scanning Electron Microscope (Hitachi S-3000N). Chemical properties of the catalyst surface were determined using a diffuse reflectance infrared Fourier transform

spectroscopy (DRIFTS) cell with a FTIR spectrophotometer (PerkinElmer spectrum GX). The DRIFTS cell was equipped with a heating cartridge controlled by a PID controller (REX-P200). During the experiment, 120 scans were accumulated; these were recorded at a resolution of 4 cm1 in the region of 40001000 cm1; an MCT detector was used [13]. The spectra were presented in the KubelkaMunk mode and some of the selected IR peaks were deconvoluted into several Gaussian curves to obtain individual intensities of the specific peaks. The samples were degassed in situ under purging nitrogen flow (10 ml/min) at 250 for 30 min prior to the IR measurements. UVvis spectra were collected by a spectrophotometer (JASCO V-500) equipped with an integrated sphere assembly over the range of 350 to 700 nm. The specific surface areas of as-prepared powders were measured by nitrogen adsorption/desorption using the Barrett-Emmett-Teller (BET) method (Micromeritics ASAP 2020). 2.3. Photocatalytic activity evaluation Each photoreaction was conducted in a continuous-flow annular photoreactor [14] containing 0.05 g of photocatalyst evenly spread on a Pyrex tube. Prior to the experiments, the samples were pre-treated in situ using a dry N2 flow (100 ml/min) at 110 for 30 min. The reactant mixture (20 ml/min) was prepared by introducing oxygen gas, continuously bubbling, through the flask with organic compounds at ambient temperature, yielding a saturation concentration of ca. 160 ppmv. After the reactant mixture was purged through the photoreactor in the dark for 60 min to achieve gas-solid adsorption equilibrium, a white daylight lamp (TFC, FL10W-EX) was turned on. The outlet gas was sampled and determined using a quadrupole mass spectrometer (SRS QMS300) at regular intervals. In addition, in-situ DRIFT technique coupled with MS at the gas outlet was applied to simultaneously monitor the adsorption of VOC on the catalyst surface and the

composition of the effluent vapors. After degassing, the IPA or benzene vapor (concentration of ca. 160 ppmv) was introduced into the microreactor of the DRIFT at room temperature in the dark for 60 min. Then, the valve used to supply the VOC vapor was closed. For the batch test, another light source, generated by an LED lamp (NICHIA) with a wavelength ranging from 450-470 nm, was used. The visible light was illuminated toward the top of the DRIFTS cell by an optical cable. Its photon intensity was 4 mWcm2 at 2 cm away from the tip of the wand. The spectra were recorded both in the dark and under illumination conditions. A constant amount of VOC vapor was periodically withdrawn from the sampling port and injected into the MS to determine the gas composition. A schematic diagram of an in situ micro-reactor, a diffuse reflectance accessory used to reflect the time evolution of infrared spectrum, and the supply of VOCs, is shown in Fig. 1. 3. Results and discussion 3.1. Characterization Powder XRD patterns of the catalysts prepared under various conditions are shown in Fig. 2. Characteristic peaks are indexed to the standard cards of Ag4V2O7 (JCPDS 77-0097) and -Ag3VO4 (JCPDS 43-0542). As can be seen, the hydrothermal synthesis samples have three kinds of XRD patterns, assigned to the pure Ag4V2O7, to the pure -Ag3VO4, and to the mixed phases of Ag4V2O7 and -Ag3VO4, respectively. The HT2 and HT4 samples had mixed phases of Ag4V2O7 and -Ag3VO4. When the hydrothermal time was increased to 6 h, the degree of -Ag3VO4 crystalline increased, without any detectable Ag4V2O7. When the hydrothermal time was further increased to 8 h, the longer hydrothermal time led to an increase in the diffraction intensities of the -Ag3VO4 crystal. In contrast, the PC samples had fairly weak -Ag3VO4 intensity. Well developed diffraction peaks of Ag4V2O7 are clearly visible for CTAB-SVO, indicating that the addition of CTAB might be favorable for the formation of Ag4V2O7

crystal. In addition, the line width at half maximum height of the HT samples was found in the order of HT2 > HT4 > HT6 > HT8, as shown in Figure 2(b). Based on the Scherrers equation, the corresponding crystal sizes of HT2, HT4, HT6, and HT8 were 44.15, 57.94, 61.80, and 62.64 nm, respectively, listed in Table 1. Apparently, the crystal size of the HT products increased with the hydrothermal time, which implies that long hydrothermal process favored the stage of the grain growth according to the theory of thermodynamic nucleation and growth. Figure 3 shows the corresponding SEM images of as-prepared samples. The morphology changed drastically when CTAB was applied. The shapes of CTAB-SVO samples, with a molar ratio of CTAB/Ag equal to 0.05, were rod-like particles with widths of 0.51 m and lengths of several micrometers (Fig. 3a). When the CTAB concentration became high, the microstructure of CTAB changed from spherical micelles to cylindrical micelles. The transition of the sphere-rod micelles in aqueous solution [15] was initiated by the nucleus growth along the longitudinal direction. Moreover, the aforementioned XRD patterns of CTAB-SVO samples had only the Ag4V2O7 crystal phase, implying that CTAB plays an important role in inhibiting the crystal growth of -Ag3VO4. This might be due to the specific binding of CTAB to a particular crystal face. CTAB also influenced the crystal structure. Irregular ragged particles with sharp edges and clear grain boundaries were observed on pure SVO samples, as shown in Figs. 3b-e. The particle sizes of HT6 and HT8 (2-8 m) were larger than those of HT2 and HT4 (2- 4 m ). PC samples (Fig. 3f) consisted of highly agglomerated particles with some rod-shape assembles. 3.2. Diffuse reflectance spectra (DRS) The DRS of as-prepared silver vanadates are shown in Fig. 4; these were measured with a JASCO V-500 spectrophotometer and converted by the KubelkaMunk equation.

All hydrothermal synthesis samples had an intense absorption in the visible light range. The spectra patterns of HT6 and HT8 were similar; the band gap absorption edge of HT6 was determined to be 550 nm and the value of band gap estimated by the onset point of the absorption curve was 2.2 eV. CTAB/SVO had a steep absorption edge in the visible light region, which was different from the absorbance spectra of HT4, which consisted of mixed structures of Ag4V2O7 and -Ag3VO4. The value of the band gap for CTAB-SVO was found to be 2.5 eV, which is similar to the value reported by Konta et al. [8]. 3.3. FT-IR spectral analyses The FT-IR spectra recorded from 400 to 1200 cm1 at room temperature for hydrothermal synthesis samples are shown in Fig. 5. The absorption bands near 500, 680, 862, 895, 920, and 965 cm1 were observed in all hydrothermal synthesis samples. The absorption bands at 500 and 680 cm1 correspond to the symmetric and asymmetric stretching vibrations of the V-O-V bridge of the V2O74- group [16], respectively. Moreover, the absorption bands at 862, 895, 920, and 965 cm-1 could be assigned to silver vanadates [17]. The bands near 920 and 965 cm-1 were attributed to the symmetric and asymmetric stretching vibrations of VO3 terminal groups of the pyrovanadate ions formed by the condensation of the VO43- tetrahedra [18]. It has been reported that peaks at 2918 and 2855 cm1 correspond to -CH2 vibrations and that the peak at 1461 cm1 belongs to -CH2 bending or the scissor mode of the CTAB tail [19]. In this study, 2918, 2855, and 1461 cm1 peaks were not observed, indicating that there was no CTAB residue in the CTAB-SVO. 3.4. Photocatalytic activity The specific surface areas of all samples are listed in Table 1. The surface area of HT samples decreased with increasing hydrothermal time, which is consistent with the

SEM results. Table 1 shows a summary of the equilibrium adsorption amounts of IPA and benzene for various samples. The equilibrium adsorption amount of VOC in the dark provides a qualitative indication of the adsorption affinity between VOC and the photocatalyst. The adsorption capacities of IPA and benzene showed the trend of HT4 > HT2 > HT6 > HT8 = CTAB-SVO > PC > P25. The adsorption amounts of IPA and benzene of the hydrothermal synthesis samples were similar. However, the hydrothermal synthesis samples, when compared to those of PC and P25, have much higher adsorption values, especially for benzene. In general, a higher surface area of the adsorbent leads to a higher adsorption capacity. However, the adsorption of IPA and benzene showed different trends for surface area. The P25 and PC samples showed lower adsorption capability than those of HT8 and HT6 samples, implying that a large surface area did not favor the adsorption of IPA and benzene. It appears that VOC adsorption is not correlated with the surface area of silver vanadates in this study. Most of the photocatalytic reactions follow the Langmuir-Hinshelwood adsorption model [20-22], in which the reactant is pre-adsorbed on the surface of the photocatalyst prior to the photoreaction. As an irreversible reaction, the L-H model can be simplified to a pseudo-first-order expression: ln(Ce/C) = kt (where Ce and C are the equilibrium concentration of adsorption and the concentration of VOC at the exposure time, t, respectively, and k is the apparent rate constant). Straight lines are obtained using regression fitting techniques; their slopes correspond to the apparent rate constant, kapp. The R2 values are all close to 1.0, implying that the experimental data are consistent with the pseudo first-order kinetic model. Table 1 summarizes the calculated kapp of IPA and benzene. The photodecomposition rates of IPA and benzene decreased in the order: HT4 > HT2 ~ HT6 > HT8 > CTAB-SVO > PC > P25. HT4 (surface area, 2.04 m2 g-1) had the maximum apparent rate constant, which was 16.6 times higher than that of P25

(surface area, 49.04 m2 g-1) for IPA and 16.2 times higher for benzene. The hydrothermal synthesis samples had higher photocatalytic activities than those of PC and P25. Besides, the rate constant of HT6 was 2.4 (IPA) and 2.5 (benzene) times higher than that of CTAB-SVO for IPA and benzene, respectively. In comparison with HT4 and HT6, it was found that the rate constant of the HT4 is larger than that of the HT6. However, HT4 had lower Ag3VO4 crystallinity than that of HT6, indicating that the crystalline phase is not the only key factor that determines the photocatalytic activity. However, Konta et al. [8] reported that among -AgVO3, Ag4V2O7, and Ag3VO4 compounds, only Ag3VO4 showed stronger photocatalytic activity under visible-light illumination, since the holes photogenerated in Ag3VO4 migrate to the reaction sites more easily than those of other silver vanadates. It is well known that the availability of active sites plays a decisive role in photocatalytic activity. Therefore, properties like crystalline structure, density of OH groups, surface acidity, and adsorption/desorption characteristics are crucial to the photodecomposition of organic compounds. To investigate the adsorption/desorption characteristics, the in situ IR technique was used to measure the adsorbates under dark conditions and gaseous VOC during visible light illumination. Figure 6 shows the spectrum of adsorption of IPA and benzene at room temperature on HT4 samples in the dark for 60 min. The spectrum recorded upon adsorption of IPA at room temperature showed three peaks in the low wavenumber region in the range of 1500-1200 cm-1 assigned to the (CH) mode of IPA; in the high wavenumber region, a very intense band at 2978 and a weak band at 2888 cm-1 were observed, corresponding to the stretching (CH) mode of methyl groups of IPA. A band at 3660 cm-1 can be assigned to the hydrogen bonded hydroxyl group of IPA or surface hydrogen bonded hydroxyl groups. Interestingly, the intensities of C-H stretching and C-H bending bands reached a steady

level after 50 min. When the adsorption time increased to 60 min, the (CH) mode reached the highest intensity, indicating the achievement of adsorption equilibrium of IPA for HT4 samples. In the case of benzene adsorption, some significant bands including the 4 parallel band at 684 cm-1 and three perpendicular bands 12 at 3048 cm-1, 13 at 1484 cm-1, 14 at 1038 cm-1, and a few combination bands were observed [23]. Due to measurement limitations, benzene adsorption on HT4 gave rise to 13 and 14 (C-H bending bands), while 13 and 14 bands reached a steady state after 40 min. The DRIFT spectra of gaseous VOCs recorded under visible light irradiation for HT4 samples are shown in Fig. 7. The IR spectra of HT4 equilibrated with IPA and benzene vapors in the dark were taken as the initial state under illumination. The response was a result of photodecomposition and desorption of VOC. These results show that the intensity of the (CH) mode of methyl groups of IPA (peak at 2978 cm-1) started to decrease after 20 min of illumination and that the intensity of the (CH) mode of benzene (peak at 1484 cm-1) disappeared after 15 min of illumination. To further clarify the nature of hydroxyl groups, we heated the hydrothermal synthesis samples to 250 to remove adsorbed water before the collection of infrared spectra on sample surface. Figure 8a shows a broad and strong band ranging from 3700 to 3000 cm-1 that was observed for HT4, while a less intense band in the same range was observed for PC, HT8, and CTAB-SVO samples. The peak in the range of 3700 to 3000 cm-1 may be due to the overlapping of various OH groups. The deconvolution of this peak was conducted using several Gaussian curves based on the IR assignments of OH groups reported from the literature. The deconvolution procedure applied to the HT4 samples resulted in seven individual peaks (Fig. 8b). The peaks located at 3200 and 3694 cm-1 correspond to the OH groups chemically adsorbed on the surface of the catalyst [24, 25]. The peaks at 3425 cm-1, 3634 cm-1, 3663 cm-1, and 3687 cm-1 are

ascribed to hydrogen bonded OH stretching vibration [26], triple coordination [27], (OH) of bridged hydroxyls with Brnsted acidity [28], and linear (OH) [29], respectively. The peak at 3550 cm-1 is assigned to the OH stretching vibration for molecularly adsorbed water since it coexists with the OH bending vibration at 1636 cm-1. The amounts of various OH groups on the as-prepared samples could be obtained by integrating the seven individual curves, as shown in Table 2. The results show that HT4 had the highest integral amount of various OH groups among the as-prepared samples. Especially, HT 4 had a very intense hydrogen bonded OH stretching vibration, high (OH) values of bridged hydroxyls with Brnsted acidity, and many OH groups chemically adsorbed compared to the cases of the other five samples. In addition, Figures 9(a) and (b) show that the photocatalytic activities of as-prepared silver vanadates are dependent on the amount of OH groups. The value of rate constant increases in the order of PC < CTAB-SVO < HT8 < HT6 < HT2 << HT4, which is consistent with the amount of OH groups. This tendency indicates that the intensity of surface hydroxyl groups of as-prepared samples is another significant factor that affects the photocatalytic activity. Comparing the value of rate constant with the number of OH groups, one can find that the value of the rate constant increases in the order of PC < CTAB-SVO < HT8 < HT6 < HT2 << HT4, which is the same order as the amounts of OH groups. This tendency indicates that the surface hydroxyl groups of as-prepared samples are another significant factor affecting photocatalytic activity. Based on the results of XRD and IR spectra of OH groups, the performance of photocatalytic activity of hydrothermal synthesis samples can be divided into two groups, one group being HT2 and HT4 and the other group being HT6, HT8, and CTAB-SVO. As listed in Table 1 and Fig. 2, HT4 consisted of mixed structures of Ag4V2O7 and -Ag3VO4, with -Ag3VO4 as the major phase, and had higher

photocatalytic activity than that of HT6, which had a stronger crystallinity of -Ag3VO4 than that of HT4. Studies have reported that P25 TiO2 is a mixed phase of rutile and anatase (75-25%) and has better photocatalytic activity than the pure anatase phase [30,31]. HT2 had lower crystallinity of -Ag3VO4 and a higher amount of surface hydroxyl groups than those of HT6; however, HT2 had a similar photocatalytic activity to that of HT6. Thus, in the cases of HT4 and HT2, it was confirmed that the intense amounts of surface hydroxyl groups could be beneficial to high photocatalytic activity. In contrast, both HT8 and CTAB-SVO had pure -Ag3VO4 and Ag4V2O7, respectively. Although the crystallinity and the amount of surface hydroxyl groups of CTAB-SVO were higher than those of HT8, the structure of Ag4V2O7 resulted in slightly low photocatalytic activity. This supports the claim that -Ag3VO4 is the active structure. 3.5. DRIFT characterization of NH3 species To further validate the existence of Brnsted acidity in HT4 samples, we carried out temperature programmed desorption of ammonia (NH3) was carried out using DRIFT investigations. Ammonia usually provides the probe molecules in spectroscopic experiments to determine the type of acid sites of heterogeneous catalysis: Brnsted sites or Lewis sites. Prior to the experiments, the samples were heated in situ from room temperature to 250 at 10 /min in N2 flow (30 ml/min), held at 250 for 30 min, and then cooled down to 30 . The samples were saturated at 30 with a gas mixture of 5% NH3 in N2 (30 ml/min) for 30 min. The DRIFT spectra were recorded every 10 min. At the end of the saturation process, the samples were flushed with N2 flow (30 ml/min). The samples were heated again at a heating rate of 10 /min, held at each temperature for 30 min. The DRIFT spectra were collected until the samples heated up to 250 . According to the literature, the region 1300-1700 cm-1 is associated with ammonia adsorbed on Brnsted or Lewis sites. The bands at 1425 and 1670 cm-1 are

correlated with ammonia protonation on Brnsted acid sites, while the bands at 1425 cm-1 and 1670 cm-1 are assigned to the asymmetric and symmetric deformation modes of ammonium ions (NH4+), respectively [32,33]. In addition, the band at 1605 cm-1 is assigned to the asymmetric deformation mode of ammonia species adsorbed on Lewis sites [32]. The spectra of the adsorbed ammonia on HT4 at 30 are shown in Fig. 10a. A band at 1425 cm-1, asymmetric NH4+ bending, was observed after saturation with 5% NH3 in N2 upon HT4 samples for 10 min. With increasing time, the adsorption amount of ammonia became larger. The stability of the adsorbed NH3 on the HT4 surface was further investigated, and the spectra of the desorption of ammonia shown in Fig. 10b. The intensities of the pre-adsorbed ammonia species decreased gradually upon heating. However, the peak of adsorbed ammonia was still observed for HT4 when the sample was heated at 250 for 30 min. This indicates that the stability of the NH3 adsorbed species on HT4 was good. The existence of Brnsted acidity of HT4 samples was confirmed. Generally, the acidity/basicity of a catalyst surface plays a crucial role in determining its catalytic activity. Figure 11 shows in situ DRIFTS spectra of NH3 species adsorbed on as-prepared silver vanadates at 30 and 250 , respectively. An intensive band at 1425 cm-1 was observed exclusively for HT4 samples, indicating that the largest amount of NH3 adsorption occurred on the HT4. However, the PC samples showed the least NH3 adsorption through the Brnsted acid sites. As a result, Figure 11 indicates that the sample HT4 inherited the richest Brnsted acid sites, which verified that the intensity of Brnsted acidity of the catalysts is responsible for the catalytic activity. 4. Conclusions Visible-light active silver vanadate photocatalysts were successfully synthesized

using a direct and simple hydrothermal process. The crystalline phase, surface area, and surface hydroxyl groups of as-prepared samples were correlated with the photocatalytic activities of IPA and benzene degradation. The crystalline structures of silver vanadates could be tuned by changing the hydrothermal time and with the assistance of CTAB. HT2 and HT4 samples had mixed phases of Ag4V2O7 and -Ag3VO4, with -Ag3VO4 as the major phase. HT6 and HT8 samples had a single -Ag3VO4 phase, but the crystallinity decreased with increasing hydrothermal time. The single Ag4V2O7 phase could be obtained for CTAB-SVO samples using the appropriate amount of CTAB. The hydrothermal synthesis samples had higher photocatalytic activities than those of PC and P25. HT2 had the highest surface area of 3.16 m2 g-1 but it had lower photocatalytic activity than that of HT4. This indicates that specific surface areas of samples did not significantly affect the photocatalytic activity. In our experiment, with a lower crystallinity of -Ag3VO4 and a higher number of surface hydroxyl groups, HT2 had similar photocatalytic activity to that of HT6. HT4 had the best photocatalytic activity, probably due to the existence of the major phase of -Ag3VO4 crystallinity and the highest amounts of surface hydroxyl groups. In conclusion, we found that the photocatalytic activities of silver vanadate photocatalysts were closely correlated to the crystallinity of -Ag3VO4 and to the amount of surface hydroxyl groups. Acknowledgments The authors are grateful to the National Science Council of Taiwan, Republic of China, for supporting this study under Contract No. NSC 95-2622-E-168 -020 CC3. References [1] J.M. Herrmann, H. Tahiri, Y. Ait-Ichou, Appl. Catal. B 13 (1997) 219-228. [2] J. Moon, H. Tajagi, Y. Fujishiro, M. Awano, J. Mater. Sci. 36 (2001) 949-955. [3] M. Anpo, M. Takeuchi, J. Catal. 216 (2003) 505-516.

[4] T. Umebayashi, T. Yamaki, S. Yamamoto, A. Miyashita, S. Tanaka, T. Sumita, K. Asai, J. Appl. Phys. 93 (2003) 5156-5160. [5] J. Ye, Z. Zou, H. Arakawa, M. Oshikiri, M. Shimoda, A. Matsushita, T. Shishido, J. Photochem. Photobiol. A: Chem. 148 (2002) 79-83. [6] X. Zhang, Z. Ai, F. Jia, L. Zhang, X. Fan, Z. Zou, Mater. Chem. Phys. 103 (2007) 162-167. [7] H. Kato, H. Kobayashi, A. Kudo, J. Phys. Chem. B 106 (2002) 12441-12447. [8] R. Konta, H. Kato, H. Kobayoshi, A. Kudo, Phys. Chem. Chem. Phys. 5 (2003) 3061-3065. [9] Y.L. Wei, P.C. Chang, J. Phys. Chem. Solids 69 (2008) 688-692. [10] X.M. Liu, S.Y. Fu, H.M. Xiao, C.J. Huang, J. Solid State Chem. 178 (2005) 2798-2803. [11] G.B. Ji, S.L. Tang, S.K. Ren, F.M. Zhang, B.X. Gu, Y.W. Du, J. Cryst. Growth 270 (2004) 156-161. [12] G. Zhou, M. Lu, F. Gu, S. Wang, Z. Xiu, Mater. Lett. 59 (2005) 2706-2709. [13] T.C.K. Yang, S.F. Wang, S.H.Y. Tsai, S.Y. Lin, Appl. Catal. B 30 (2001) 293-301. [14] G.T. Pan, C.M. Huang, L.C. Chen, W.T. Shiu, J. Environ. Eng. Manage. 16(6) (2006) 413-420. [15] J. Xu, G.Z. Li, Colloid Surface 191 (2001) 269-278. [16] K. Nakamoto, Infrared and Raman Spectra of Inorganic and Co-ordination Compounds, 5th Edition, Part A, Wiley, New York, 1997. [17] M. Xue, J. Ge, H. Zhang, J. Shen, Appl. Catal. A 330 (2007) 117-126. [18] S. Murugesan, A. Wijayasinghe, B. Bergman, Solid State Ionics 178 (2007) 779-783. [19] K.D. Dobson, A.D.R. Lanzilotta, A.J. McQuillan, Vibrational Spectroscopy 24

(2000) 287-295. [20] G. Vincent, P.M. Marquaire, O. Zahraa, J. Photochem. Photobiol. A 197 (2008) 177189. [21] M.A. Hasnat, M.M. Uddin, A.J.F. Samed, S.S. Alam, S. Hossain, J. Hazard. Mater. 147 (2007) 471477. [22] N. Barka, A. Assabbane, A. Nounah, Y. At Ichou, J. Hazard. Mater. 152 (2008) 10541059. [23] G.Di Lonardo, L. Fusina, G. Masciarelli, F. Tullini, Spectrochim Acta A 55 (1999)15351544. [24] G. Postole, A. Gervasini, M. Caldararu, B. Bonnetot, A. Auroux, Appl. Catal. A 325 (2007) 227-236. [25] M. Janus, A.W. Morawski, Appl. Catal. B 75 (2007) 118123. [26] N.B. Colthup, L.H. Daly, S.E. Wiberley, Introduction to Infrared and Raman Spectroscopy, 3rd Edition, Academic Press, Boston, 1990. [27] A. Davydov, Molecular Spectroscopy of Oxide Catalyst Surfaces, Wiley, Hoboken, 2003. [28] A. Zecchina, L. Marchese, S. Bordiga, C. Paze`, E. Gianotti, J. Phys. Chem. B. 101 (1997) 10128-10135. [29] J.M. Coronado, S. Kataoka, I. Tejedor-Tejedor, M.A. Anderson, J. Catal. 219 (2003) 219-230. [30] W.G. Zhang, L.L. Zhang, Z.J. Jiang, R.Q. Li, X.J. Yang, X. Wang, L.D. Lu, Mater. Chem. Phys. 105 (2007) 414-418. [31] D. Gumy, S.A. Giraldo, J. Rengifo, C. Pulgarin, Appl. Catal. B 78 (2008) 19-29. [32] L. Lietti, P. Forzatti, G. Ramis, G. Busca, F. Bregani, Appl. Catal. B 3 (1993)

13-35. [33] G. Ramis, L. Yi, G. Busca, Catal. Today 28 (1996) 373-380.

Figures Captions

Fig. 1 Fig. 2

Schematic diagram of an in-situ micro-photoreactor. (a) XRD patterns of as-prepared silver vanadates. (b) A scale-up diagram of the XRD patterns at the peak 30.9 for HT samples.

Fig. 3

SEM micrographs of as-prepared silver vanadates: (a) CTAB-SVO; (b) HT2; (c) HT4; (d) HT6; (e) HT8; (f) PC.

Fig. 4 Fig. 5

Diffuse reflectance spectra of as-prepared silver vanadates. FT-IR spectra of hydrothermal synthesis silver vanadates: (a) CTAB-SVO, (b) HT2, (c) HT4, (d) HT6, and (e) HT8.

Fig. 6

IR spectra of (a) adsorbed IPA and (b) adsorbed benzene on HT4 samples during the 60 min of darkness.

Fig. 7

IR spectra during (a) 20 min of photocatalytic degradation of gaseous IPA and (b) 15 min of photocatalytic degradation of gaseous benzene.

Fig. 8

Infrared spectra of OH groups on (a) hydrothermal-synthesis silver vanadates at 250C and (b) the spectrum (region from 3000 to 3700 cm1) obtained for HT4 taken at 250 with seven peaks located at (1) 3200 cm-1, (2) 3425 cm-1, (3) 3550 cm-1, (4) 3634 cm-1, (5) 3663 cm-1, (6) 3687 cm-1, and (7) 3694 cm-1. The measured peak and the summation of the calculated peaks overlap.

Fig. 9

Relationship between total area of OH groups and the rate constants of (a) IPA decomposition and (b) benzene decomposition for as-prepared silver vanadates. The solid line is for the integral amount of OH groups and the dashed line is for the rate constants of IPA/benzene decomposition.

Fig. 10

DRIFT spectra of NH3 species adsorbed on HT 4 samples: (a) at 30 for a period of 30 min; (b) out-gassing from 30 to 250 .

Fig. 11

DRIFT spectra of NH3 species adsorbed on as-prepared silver vanadates at (a)

30 and (b) 250 .

Table captions Table 1 Summary of surface areas, crystal sizes, adsorption capabilities, and photocatalytic activities of samples. Table 2 Calculated various OH peaks in the IR spectrum (region from 3000 to 3700 cm1) obtained for as-prepared samples at 250 .

Figure 1

Figure 2(a)

Figure 2(b)

Figure 3 (a)
(a)

Figure 3 (b)
(b)

Figure 3 (c)
(c)

Figure 3 (d)
(d)

Figure 3 (e)
(e)

Figure 3 (f)

Figure 4

Figure 5

Kubelka-Munk

Figure 6 (a)
(a)

Figure 6 (b)

(b)

Figure 7 (a)
(a)

Figure 7 (b)

(b)

Figure 8 (a)
(a)

Figure 8 (b)

Figure 9 (a)
220 200 1.5 180 160 140 120 100 80 60 40 20 0 PC HT8 HT6 HT4 HT2 CTAB-SVO 1.0 0.5 0.0

2.0

Figure 9 (b)

Figure 10 (a)

(a)

Figure 10 (b)

(b)

Figure 11 (a)

Figure 11 (b)

Table 1
Sample Surface area (m2 g-1) P25 49.04 PC 2.60 HT8 1.38 HT6 1.65 HT4 2.04 HT2 3.16 Crystal size (nm) 62.64 61.80 57.94 44.15 IPA ads./mass cat. Benzene ads./mass cat. (mol l-1g-1) (mol l-1 g-1) Rate constant of IPA decomposition (mole l-1 min-1) 0.12 0.17 0.59 1.22 1.99 1.21 0.52 Rate constant of benzene decomposition (mole l-1 min-1) 0.08 0.24 0.39 0.94 1.30 0.93 0.38

3.4810-4 4.4810-4 7.6610-4 7.7010-4 7.9010-4 7.8010-4 3.6010-4

7.2810-6 4.2210-5 3.6010-4 3.6210-4 3.8210-4 3.7010-4 7.6610-4

CTAB-SVO 2.15

Table 2
Sample PC HT8 HT6 HT4 HT2 CTAB-SVO Isolated-OH 11.2 12.6 22.2 103.3 26.8 15.3 H2Oad 2.4 5.6 2.2 5.9 2.5 1.1 Total area 13.6 18.2 24.4 109.2 29.3 16.4


1. MOx: 2 (a) C. M. Huang, Guan T. Pan, Lung C. Chen, Thomas C.K. Yang, Wen S. Chang, Effects of Synthesis Conditions on the Crystalline Phases and Photocatalytic Activities of Silver Vanadates via Hydrothermal Method, 2009 Materials Research Society, April ,13-17,San Francisco , USA. (b) G. T. Pan, C. M. Huang, C.K. Yang, An in-situ drift spectroscopic study of the photocatalyst Ag3VO4 by the irradiation of visible-light, The 20th International Symposium on Chemical Reaction Engineering 2008, Oral, Sep. 7-10, Kyoto, JAPAN. SCI2 (a) C.M. Huang, G.T. Pan, K.W. Cheng, W.S. Chang, T.C.K. Yang*, 2009
CTAB-assisted hydrothermal synthesis of silver vanadates and their photocatalytic characterization, Chemical Engineering Science, in press (SCI, Impact Factor: 1.884) (b) C.M. Huang, G.T. Pan, Y.C. M. Li, M.H. Li, T.C.K. Yang*, Crystalline phases and photocatalytic activities of hydrothermal-synthesis Ag3VO4 and Ag4V2O7 under visible light irradiation, Appl. Catal. A 358 (2009) 164-172. (SCI, Impact Factor: 3.190)

2.

MOx/SBA-15: BiVO4/SBA-15

Appl. Catal. B.


98 09 28

MOx/SBA-15


NSC 97-2221-E-168-013

/ /
BiVO4/SBA-15 BiVO4(510 20%) BiVO4 monoclinic 2.44 eV 298~435 m2/g 10 W 154 ppm BiVO4 10% BiVO4 10 BiVO4 SBA-15 SBA-15 BiVO4 5% BiVO4 2 40 10 ppmSBA-15 BiVO4


10

14

98

16

()

Eighth Annual Carbon Capture & Sequestration Conference ( 8 )


98 5 4 98 5 7 4

Pittsburgh, PA, USA ()

U.S. DOE ()

2009 8th Annual Conference on Carbon Capture & Sequestration 98 5 4-7 CCS

CCS CCS


1. .. ()

2. 3. 4. 5.

98 5 18

98 5 4 ~5 7 NSC 97-2221-E-168-013 Sheraton Station Square Pittsburgh, Pennsylvania USA () Eighth Annual Carbon Capture & Sequestration Conference () 8 Exchange Monitor Publications & Forums National Energy Technology Laboratory (NETL) U. S. Department of Energy (DOE, USA) Preliminary Study on the Potential of Geological Sequestration in Taiwan

() () ()() () () ()
1

() 2009 8th Annual Conference on Carbon Capture & Sequestration (CCS) (Carbon Capture) () ()()

05 3

05 4

( ) 8th Annual Carbon Capture & Sequestration Conference

05 4

05 7

/ Exchange Monitor Publications & Forums

05 8

05 10 /

( )

() CCS CCS CCS :


3

1.

Exchange Monitor Publications & Forums 8th Annual Conference on Carbon Capture & Sequestration 98 5 4-7 U.S.DOEShellGE CCS CCS

1 8th Annual Conference on Carbon Capture & Sequestration

2 WorleyParsons CCS

3 Baker Hughes CCS

: (1)2009 Annual Conference on Carbon Capture & Sequestration(2)CCS : (1) 2009 Annual Conference on Carbon Capture & Sequestration (05/05) Opening session Keynote AddressPoster sessionConcurrent Technology session Discussion Session Carl Bauer NREL Director() NETL CCS CCS CCS () CCS

4 Carl Bauer

5 Concurrent Technology session post-combustion /RetrofitsystemCapture-system analysisCCS programs


8

CCS () CCS EOR CCS CCS

6 CCS

( )(Oxy-fuel) CO H2 (catalytic shift converter) CO2 H2 H2 CO2 CO2 CCS CCS (COE) (US 0.012-0.036 kW/h) 12-36 US$ MWh-1 (PC) 40-85%(NGCC) 35-70%(IGCC) 20-55% IEA(2004) CCS US$50-100/tCO2 30-100 % 25-100% 2030 CCS US$25-50/tCO2
10

CCS 1-2(US$ cents) 10-20 % IGCC (IEA, 2009)

7IGCC CCS programs CCS CCS CCSNETLSHELLChervon CO2 Pipeline Studies at NETL(USDOE) 201290 % 99 %

11

8 CO2 Pipeline Studies at NETL

12

Big Sky Carbon Sequestration Partnership

(2) CCS 05/04-05/07 Sheraton Station Square 1 ALSCOM Chilled Ammonia 10-12
13

Gas Technology Institute (gti) /( 13) / /

10 ALSTOM chilled ammonia

14

11 ALSTOM AEPS mountaineer plant

12 ALSTOM chilled ammonia process overview

15

13 gti Carbo-Lock process : A CO2 Capture Technology using Membrane Contactors

(13) Poster Section CCS 14-16

16

14

17

15

16 Charlotte Sullivan, Pacific Northwest National LAB., CCS

18

() CCS ALSCOM Maciej Radosz (Department of Chemical and Petroleum Engineering) CCS CCS 70% ()
19

CCS CCS CCS CCS CCS CCS ( 17)

17 -()

(NSC 97-2221-E-168-013) (NT$ 50,000) 2009 CCS


20

21

22

charming@mail.ksu.edu.tw - 05/19/2009 12:57:20+0800 - Open WebMail (+)

: 682

big5 -- xdVO^z? -24 : Fri, 15 May 2009 07:31:29 -0600 : Maciej Radosz <Radosz@uwyo.edu> : "'Charming'" <charming@mail.ksu.edu.tw> : RE: Charming Huang/TW: P.S. P.S. If you would like to collaborate with us, which we welcome, please keep us informed of your progress and plans. -----Original Message----From: Maciej Radosz Sent: Friday, May 15, 2009 7:27 AM To: 'Charming' Subject: Charming Huang/TW: Thanks for your paper. Dear Prof. Huang, Thank you. I thought you were also working on density functional theory, but I must have gotten confused. I appreciate knowing your papers. Sincerely, ----------------------------------------Maciej Radosz Professor Soft Materials Laboratory Department of Chemical and Petroleum Engineering University of Wyoming Laramie, WY 82071 radosz@uwyo.edu tel 307-766-4926 fax 307-766-6777 http://wwweng.uwyo.edu/economic/sml ----------------------------------------This email and any files transmitted with it are confidential
http://mailtch.ksu.edu.tw/cgi-bin/webmail/o...n=readmessage&headers=simple&attmode=simple 1 / 2 [2009/5/19 12:59:27]

charming@mail.ksu.edu.tw - 05/19/2009 12:57:20+0800 - Open WebMail (+)

and intended solely for the use of the individual or entity to whom they are addressed. -----Original Message----From: Charming [mailto:charming@mail.ksu.edu.tw] Sent: Thursday, May 14, 2009 11:28 PM To: Maciej Radosz Subject: Thanks for your paper. Dear Prof. Radosz: It was nice talking with you. Attached are some of my papers. My current research is on the development and application of photocatalyst. Now, we are interested in the material for CO2 capture. Sincerely yours, Charming Huang PhD Associate Professor Department of Environmental Engineering Kun Shan University No. 949 Da Wan Rd., Yung Kang City, Tainan 710, Taiwan, R.O.C. E-mail: charming@mail.ksu.edu.tw Tel.:+886-6-2050359; fax: +886-6-2050540 24

Open WebMail version 2.32 ?

http://mailtch.ksu.edu.tw/cgi-bin/webmail/o...n=readmessage&headers=simple&attmode=simple 2 / 2 [2009/5/19 12:59:27]

Preliminary Study on the Potential of Geological Sequestration in Taiwan


Tong Lun-Tao, Liao Chi-Wen, Lin Cheng-Kuo, Chang Wen-Sheng Energy and Environment Laboratory Industrial Technology Research Institute, Taiwan Huang Chao-Ming Department of Environment Kun Shan University, Taiwan

EIGHTH ANNUAL CONFERENCE ON CARBON CAPTURE AND SEQUESTRATION - DOE/NETL May 4 7, 2009

ABSTRACT
Total CO2 emission in Taiwan at the year of 2007 had reached 268.8 Mtons, which accounted for around 1% of the world. About 60% of emitted CO2 can be attributed to the energy sector -- primarily from fire power plants burning fossil fuels. Greenhouse gases reduction has become an important issue in energy sector to cope with the emerging crisis of global warming, more specific, the coming shackles from punitive measures such as emission quotas or cap-and-trade mechanisms. Therefore the Bureau of Energy of Taiwan had started to sponsor a 3-years project to build up fundamental techniques of CO2 geological sequestration and to evaluate the preliminary potential of CO2 sequestration in Taiwan since the year of 2006. The goal of this project is to determine whether geological sequestration is a viable option for energy sector to reduce CO2 emission in a significant way. A preliminary assessment of geological sequestration potential for western Taiwan is underway. Complement field surveys including magnetotelluric and seismic reflections are also being conducted in the western Taiwan. Hopefully the integrated results will provide a basis for promoting a pilot-scale injection site for demonstrating geological sequestration in Taiwan. In addition to preliminary assessment of geological sequestration potential, we also invested resources for establishing numerical simulation capabilities including geochemical and transport modeling. Properties of representative caprock and reservoir have been collected from literatures and reports for simulating possible CO2/Rock/Brine interactions. A synthetic case for elaborating optimal sequestration depth in Taiwan had been developed. A simplified geological model for transport modeling was also developed for concept proofing for storing CO2 in aquifers beneath northwestern Taiwan.

EIGHTH ANNUAL CONFERENCE ON CARBON CAPTURE AND SEQUESTRATION - DOE/NETL May 4 7, 2009

INTRODUCTION
Based on published literatures, the APEC Energy Working Group had published an assessment report on CO2 storage prospectivity in the Asia Pacific region (Newlands and Langford, 2005). The prospectivity of geological sequestration in depleted reservoirs was considered as limited due to minor hydrocarbon resources in Taiwan. As for prospectivity of saline aquifer, the option is limited to offshore basins in the western coast. The thickness of Cenozoic sediments in the Taiwan strait ranges from zero to ten kilometers (Lin et al, 2003). The East China Sea basin has come across Taoyuan, Hsinchu and terminated by Peikang high located in Chiayi. To the south is the Taixinan basin which covers Tainan, Kaohsiung, and extended to the offshore area. For geological sequestration in saline aquifers, the target depth shall exceed 800 meters and may go up to two or three thousand meters depending on regional geological settings. This makes two sedimentary basins possible sinks for storing CO2 underground. As for the sources of CO2, major stationary emission sources are distributed along the western coast of Taiwan. The length of Taiwan island is around 400 kilometers, which means all emission sources are within the ranges that CO2 can be transported economically via pipelines.

POTENTIAL ENVIRONMENT FOR GEOLOGICAL SEQUESTRATION


By literature reviews and re-examination of regional seismic survey data, the Kueichulin and Nanchuang formations In northwestern Taiwan are determined as potential reservoirs for geological sequestration. The Kueichulin formation
EIGHTH ANNUAL CONFERENCE ON CARBON CAPTURE AND SEQUESTRATION - DOE/NETL May 4 7, 2009

represents the transition from passive-margin to foreland basin from late-Miocene to early Pliocene (Chiang, 2007) and can be subdivided into three members, that are, Kuandaoshan sandstone, Shiliufen shale, and Yutengping sandstone. The east-

dipping sedimentary sequences of interbedded sandstones and shales will allow injected carbon dioxide to migrate laterally to the west (away from shoreline) and increase the amount of trapped carbon dioxide. Overlaid the Kuechulin formation is a regionally distributed shale formation -- Chinshui shale. The lower bound of the Chinshui shale in northwestern Taiwan is at around 800~1,000 meters, which is favorable for keeping injected carbon dioxide in super-critical status. Therefore the Chinshui shale can act as caprock once the carbon dioxide is injected into the Kueichilin formation. Beneath the Kueichulin formation is the Nanchuang formation which is also a suitable reservoir for geological sequestration. The top member of Nanchuang formation is Shangfuchi sandstone which is mainly composed of thickbedded or massive, medium to coarse-grain white sandstone. Besides, a series of complementary field surveys were conducted in the northwestern Taiwan for characterizing potential storage environment. Several seismic reflection profiles perpendicular and parallel to the coastline have revealed relatively undisturbed and slightly east-dipping formations. Regional distribution of caprock and reservoir matchings were identified using profiles based on magnetotelluric surveys. Combining seismic and magnetotelluric data, the final results have suggested that the area between Yungan and Houhu might be suitable for geological sequestration. The thickness of potential reservoir are in the range of 600 to 1,000 meters and covered by massive Chinshui shale as caprock. To the north the depth of potential reservoir goes deeper which may increase the operation cost. To the south the potential reservoir was disturbed by regional structures that may compromise the safety of injection site. The geological setting of the southwestern Taiwan can be separated into three regions, i.e. Yunlin, Chiayi, and Tainan, separated by the B fault and Yichu fault. The seismicity and fault occurrence is relatively lower in the Yunlin region
EIGHTH ANNUAL CONFERENCE ON CARBON CAPTURE AND SEQUESTRATION - DOE/NETL May 4 7, 2009

compared to Chiayi and Tainan regions . The regional geological formation dips to the north and east. Possible caprock-reservoir matchings are Talu shale and Peiliao formation (green line). Top of the Peiliao formation is at the depth of 1,000~1,500 meters. And the thickness of the Talu shale above the Peiliao formation is around 60~90 meters. The deeper Pachangsi sandstone with high porosities may also be a potential reservoir although it only occurred near Taisi. The Chiayi region is heavily affected by regional faults and therefore not suitable for geological sequestration despite possible caprock-reservoir matchings do exist. As for the Tainan region, the regional geological formations dips to the south. Possible reservoirs are Erchungsi(EC), Liuchungsi(LC), Niaotsui(NT), and Chunglun(CL) formations, and possible caprocks are Liushuang(LS),

Kansialiao(KSL), and Yunshuisi(YS) formations. The seismic reflection profile L-2 has identified the so-called B-fault separating Yunlin and Jiayi region. The L-1, L-A, and L-B profiles have shown relatively undisturbed reservoirs between 1,000 and 1,500 meters above the pre-Miocene basement. The caprock revealed by magnetotelluric surveys is at around 800 meters depth. For overall observation, the area surrounded by Taisi, Kouhu, Shuilin and Peikang is relatively stable and composed of suitable caprock-reservoir matching. To the south the structures are getting more complex and the proportion of muddy materials is getting higher. To the north the depth of suitable reservoirs goes beyond feasible range for geological sequestration.

Preliminary Estimation on Geological Sequestration Potential


A comprehensive estimation on the geological sequestration is not a feasible option due to the limited data. Therefore we adopted the method proposed by USDOE (USDOE, 2007) for preliminarily estimating the regional scale potential of
EIGHTH ANNUAL CONFERENCE ON CARBON CAPTURE AND SEQUESTRATION - DOE/NETL May 4 7, 2009

geological sequestration, GCO2 = AxhxxxE, which considered the area, thickness, porosity, density and efficiency factor. The scenario was to inject carbon dioxide along the coastline. The design of area parameters used for calculation are based on the basin shape and bounded by the territorial waters limit. The thickness of reservoirs were designed based on the thickness data synthesized from literatures and field survey data. Both porosity and density data are derived from literature reviews. The efficiency factor ranged from 0.1 to 0.4. All these parameters are given as simple probability distributions and the probability of resultant geological sequestration potential were obtained. The estimated potential ranged from 9 to 68 billion tons according to the results by Monte Carlo simulation.

Technology R&D for Geological Sequestration


Apart from studying on the potential, ITRI is also developing related technologies essential to the deployment of geological sequestration, such as optimal sequestration depth estimation (Lin, 2008), geochemical modeling using Geochemist's Workbench, transport modeling based on FEHM, and geological modeling based on GOCAD. ITRI is also upgrading its existing capabilities on site investigation and monitoring techniques. An integrated framework will be built to facilitate the promotion and planning of a pilot injection project underway. ITRI has initiated a research framework on CCS technology funded by the Bureau of Energy, Taiwan. Since a promising storage potential of saline aquifer has been found by this study, a pilot injection site for concept proofing on geological storage is being considered for near-term target due to its essentialness to large scale implementation of CCS. The pilot site will be a platform for technology R&D, international collaboration, and public outreach. Related monitoring and assessment technology will also be developed. The pilot site will also play an important role for increasing international visibility of Taiwan. A pilot injection site for concept proofing on geological storage is essential for
EIGHTH ANNUAL CONFERENCE ON CARBON CAPTURE AND SEQUESTRATION - DOE/NETL May 4 7, 2009

near-term target. The pilot site will be a platform for technology R&D, international collaboration, and public outreach. Related monitoring and assessment technology will also be developed. The pilot site will also play an important role for increasing international visibility of Taiwan.

CONCLUSION
There will be increasing pressures on reducing CO2 emissions to the atmosphere, and will soon lead to a new era of carbon economy. It is recognized that CCS can play a more significant role. This ongoing study not only provides prospectivity on geological sequestration in Taiwan, but also tries to setup a technology framework for monitoring and assessment techniques in several aspects. Hopefully this framework can facilitate a pilot scale injection demonstration in Taiwan.

ACKNOWLEDGEMENT
The authors would like to thank the Bureau of Energy, Ministry of Economic Affairs, Taiwan for financial support.

REFERENCES Chiang, S.P. (2007) Early development of the central Taiwan foreland basin revealed from stratigraphic record, Master's Thesis, Institute of Geophysics, National Central University. Lin, A.T., Watts, A.B., and Hesselbo, S.P. (2003) Cenozoic Stratigraphy and Subsidence History of the South China Sea margin in the Taiwan Region.

EIGHTH ANNUAL CONFERENCE ON CARBON CAPTURE AND SEQUESTRATION - DOE/NETL May 4 7, 2009

Basin Research, 15, 453-478. Lin, C.K. (2008) Algorithm for determining optimum sequestration depth of CO2 trapped by residual gas and solubility trapping mechanisms in a deep saline formation, Geofluids, Vol. 8, No. 4., pp. 333-343. Newlands, I.K. and R.P. Langford (2005) Assessment of geological storage potential of carbon dioxide in the APEC region - Phase 1: CO2 storage prospectivity of selected sedimentary basins in the region of China and South East Asia. APEC Energy Working Group EWG Project 06/2003. Innovative Carbon Technologies, Canberra. Newlands, I.K., R.P. Langford and R. Causebook (2006) Assessing the CO2 storage prospectivity of developing economies in APECapplying methodologies developed in GEODISC to selected sedimentary basins in the Eastern Asian region. In: J. Gale, N. Rokke, P. Zweigel and H. Svenson, Editors, Proceedings of the Eighth International Conference on Greenhouse Gas Control Technologies, Elsevier. USDOE (2007) Carbon Sequestration Atlas of the United States and Canada, National Energy Technology Laboratory.

EIGHTH ANNUAL CONFERENCE ON CARBON CAPTURE AND SEQUESTRATION - DOE/NETL May 4 7, 2009

Potrebbero piacerti anche