You are on page 1of 7

RESEARCH ARTICLE

Reactive adsorption desulfurization over a Ni/ZnO adsorbent


prepared by homogeneous precipitation
Aihua KONG
1
, Yanyu WEI
1
, Yonghong LI ()
1,2
1 Key Laboratory for Green Chemical Technology of Ministry of Education,
School of Chemical Engineering & Technology, Tianjin University, Tianjin 300072, China
2 National Engineering Research Center for Distillation Technology,
School of Chemical Engineering & Technology, Tianjin University, Tianjin 300072, China
Higher Education Press and Springer-Verlag Berlin Heidelberg 2013
Abstract A high-performance Ni/ZnO adsorbent was
prepared by homogeneous precipitation using urea hydro-
lysis and characterized by N
2
adsorption-desorption, X-ray
diffraction (XRD), and scanning electron microscope
(SEM). The adsorbent was applied to the deep desulfur-
ization of gasoline and showed a high breakthrough sulfur
capacity and a remarkably high volume hourly space
velocity. The effects of coexisting olens in gasoline as
well as adsorptive conditions on the adsorptive perfor-
mance were examined. It was found that olens in gasoline
had a slightly inhibiting effect on the desulfurization
performance of the adsorbent. The optimum conditions
were 673 K, 1.0 Mpa with a volume hourly space velocity
of 60 h
1
. Under the optimum conditions, ultralow sulfur
gasoline could be produced and the breakthrough sulfur
capacity of the adsorbent was 360 mg-s/g-sorb for the
model gasoline.
Keywords nickel, reactive adsorption, desulfurization,
thiophene
1 Introduction
Removal of organosulfur compounds from liquid hydro-
carbon fuels has attracted a great deal of attention and
become an important research subject in the environmental
and energy elds. Recently, the fuel specications for
transportation fuels and the requirements for liquid
hydrocarbon fuels for fuel cell applications have become
more and more stringent [1,2]. The U.S. Environmental
Protection Agency announced new regulations that
required the reduction of sulfur levels to 30 ppmw in
gasoline and below 15 ppmw in diesel fuels. For fuel cell
applications, the sulfur concentration in the fuels needs to
be below 0.2 ppmw, because sulfur compounds contam-
inate anode electro-catalysts and degrade their perfor-
mance
1)
[3].
Current hydrodesulfurization (HDS) processes cannot
remove refractory sulfur compounds to the desired level
because HDS catalysts have a low reactivity for thiophenic
compounds. Furthermore, they are not practical for small-
scale residential units and transportation systems. The
more important issue is that HDS systems need to function
at higher temperatures and higher hydrogen pressures.
Recently, adsorptive desulfurization has been reported to
be a complementary and alternative technique for HDS [4].
The sulfur content was reduced from 430 to lower than 0.2
ppmw in a commercial diesel fuel by using metal ion
exchanged Y zeolites [5]. However, their selectivity and
capacity for sulfur compounds varied depending on the
fuel composition. Reactive adsorption desulfurization
(RADS), whose adsorption is assisted by a chemical
reaction between metal particles and adsorbates, has been
implemented in S-Zorb desulfurization technology and
used on an industrial scale [69]. RADS refers to processes
that use metal-based sorbents such as Ni/ZnO-SiO
2
-Al
2
O
3
[10] and it has advantages of both catalytic and adsorption
technologies. RADS can be used to remove relatively
stable compounds like thiophenic compounds to very low
S levels. The active metal (Ni) reacts with the sulfur-
containing compound under an H
2
atmosphere to form
metal suldes. Subsequently, the metal suldes react with
ZnO to form ZnS. This process can achieve deep
desulfurization with a low loss in octane number.
To date, there have been several reports on the
desulfurization of fuels by reactive adsorption. Zhou et
Received August 8, 2012; accepted February 22, 2013
E-mail: yhli@tju.edu.cn
1) USEPA. http://www. epa. gov/sbrefaldocuments/pnl13f. pdf
Front. Chem. Sci. Eng. 2013, 7(2): 170176
DOI 10.1007/s11705-013-1322-9
al. reported that the optimal operating conditions for the
desulfurization of uid catalytic cracking (FCC) gasoline
over Ni/ZnO adsorbents were a total pressure of 2 Mpa, a
temperature of 370C380C and a LHSV of 4 h
1
[11].
Under these conditions, the breakthrough capacity (with 30
ppmw as the breakthrough sulfur level) was 25.4 mg/g and
the loss in octane number was 1.1. Tawara et al. reported
the use of Ni/ZnO as a sorbent for the desulfurization of
kerosene with a LHSV of 0.25 h
1
[12]. Gao et al. also
studied the use of Ni/ZnO as adsorbents with the optimal
operating conditions being a temperature of 370C, a
pressure of 1.5 MPa and a LHSV of 8.5 h
1
[9]. Here a
breakthrough sulfur capacity of 4.3 mg-s/g and a loss in
octane number in the gasoline product of 1.0 unit was
obtained. The reactive adsorption desulfurization used in
industrial production is mainly restricted to low sulfur
capacities and small productivities with low volume hourly
space velocities (0.2510 h
1
). As a consequence, consider-
able attention should be paid to developing an adsorptive
process for deep desulfurization of gasoline under optimal
operating conditions with a high sulfur capacity and
productivity. As is well-known, gasoline not only contains
saturated hydrocarbons but also a large amount of
unsaturated hydrocarbons, such as aromatics and olens
with concentration even up to 20 wt-% and 40 wt-%,
respectively, especially in FCC gasoline. Consequently, the
challenge in the development of an adsorptive process is to
develop an adsorbent that has higher adsorptive selectivity
and capacity for the sulfur compounds. However, the
information about effects of the coexisting species in the
fuel and adsorption conditions on adsorptive performance
is still very limited in the literature. As a consequence, it is
essential to clarify effects of the coexisting species in the
fuel and adsorption conditions on adsorptive performance.
Homogeneous precipitation is an attractive technique for
the synthesis of catalysts due to its low cost and simple
methodology. It leads to homogeneous, high-purity, well-
dispersed and small-sized materials [13]. By comparison,
the coprecipitation technique is pH sensitive and the sol-
gel technique requires expensive precursors. Adsorbents
prepared by homogeneous precipitation via urea hydro-
lysis are high-performance [1417].
In the present study, the performance of Ni/ZnO
adsorbents prepared via homogeneous precipitation was
evaluated in a xed bed system. The structure and activity of
the adsorbent were studied using N
2
adsorption-desorption,
X-ray diffraction (XRD) and scanning electron microscope
(SEM). Furthermore, adsorptive capacity and the effects of
coexisting olens in gasoline as well as adsorptive
conditions on the adsorptive performance were examined.
2 Experiment
2.1 Feedstocks
Thiophene was added to sulfur-free n-heptane at a sulfur
concentration of 100 ppmw and used as a model gasoline
A. Model gasoline B contained 100 ppmw sulfur as
thiophene and 35 wt-% of isopentene as olens, 15 wt-%
toluene as aromatic compounds. The model gasoline B
also contained 10 wt-% of cyclohexane to mimic the
cycloalkane. The thiophene and hydrocarbon compounds
used for preparing the model fuels were purchased from
Sigma-Aldrich and used as received without further
purication.
2.2 Adsorbent
Ni(NO
3
)
2
$6H
2
O, Zn(NO
3
)
2
$2H
2
O and CO(NH
2
)
2
were
purchased from Kermal. Sodium dodecyylsulfate (SDS,
analytical grade) was purchased from Sinopharm. Ni/ZnO
adsorbents were prepared by homogeneous precipitation
described as follows: 5.3 g of Ni(NO
3
)
2
$6H
2
O and 20 g of
Zn(NO
3
)
2
$6H
2
O were dissolved in 80 mL of water and
then 60 g of urea and 24 g of surfactant sodium lauryl
sulfate were added to the obtained light blue solution. The
solution was stirred for 2 h. The precipitate was ltered,
washed thoroughly with water and alcohol, dried in air
overnight at 100C, and nally calcined in a programming
temperature at 1C/min to 250C and maintained for 3 h,
2C/min to 400C and maintained for 3 h. This catalyst
corresponds to 20 wt-% NiO in the NiO/ZnO sample. The
calcined samples (0.25 g) were treated in a hydrogen ow
(1.0 MPa, 50 mL/min) at 400C for 2 h and then used for
desulfurization experiments.
2.3 Analysis of gasoline samples
The hydrocarbon group compositions of model gasoline
were determined by a Fuli GC-9790 gas chromatography
equipped with a ame ionization detector (FID). The total
sulfur was determined with a DL-2B-EE Micro-Coulomb
instrument with a sulfur detection limit of 0.1 ppmw.
2.4 Adsorbent characterization
The BET surface area of the adsorbents was measured by
nitrogen sorption at 196C with a TriStar-3000 gas
absorption analyzer (Micromeritics Instrument Co.). The
samples were degassed at 300C for 3 h prior to the
measurement. X-ray diffraction (XRD) was measured by a
Rigaku D/max 2500v/pc X-ray diffractometer (Cu K, =
0.154 06 nm, 40 kV, 40 mA) in the step scanning mode (6/
min) with a 2 range of 10 to 70. The crystal phase
compositions of the adsorbents were identied by match-
ing the spectra with the ICDD/JCPDS database. The
average crystallite size of the various species was
calculated by the Scherrer formula. The surface shapes
were analyzed using a eld-emission scanning electron
microscope (FEI Quanta 200F scanning electron micro-
scope).
Aihua KONG et al. Deep desulfurization of gasoline over Ni/ZnO adsorbent 171
2.5 Desulfurization experiments
The desulfurization experiments were carried out in a
stainless steel tubular ow microreactor with an internal
diameter of 8.0 mm. 0.25 g of adsorbent (2040 mash) was
used per run. The feedstock was injected through a pump
and mixed with H
2
in a preheating furnace. The mixture
was heated to approximately 400C and injected into the
reactor. The outlet was analyzed per hour. When the sulfur
concentration at the outlet reached 10 ppmw, this time was
dened as the breakthrough time and the sulfur capacity
was dened as the breakthrough sulfur capacity.
3 Results and discussion
3.1 Characterization of adsorbents
3.1.1 X-ray
The XRD patterns of the calcined, reduced and sulfurized
adsorbents are shown in Fig. 1. The diffraction peaks at 2
range of 3040 belong to ZnO. Fig. 1 (a) displays
reection peaks of both the ZnO and NiO phases.
Furthermore, the XRD analysis of the reduced adsorbent
in Fig. 1 (b) shows that all the NiO peaks have disappeared
and diffraction peaks for metallic nickel appear. This result
clearly indicates that all the NiO phase have been
transformed to a Ni
0
phase after the reduction treatment
and no Ni-Zn alloy (2 = 37.12), which was previously
observed by Ryzhikov et al. [18], is detected on the
reduced Ni/ZnO. The DebyeScherrer formula and the full
width at half-maximum of the highest intensity peak of
ZnO (101) are used to estimate the average particle sizes of
ZnO in the samples. The crystallite size of ZnO in both
reduced Ni/ZnO and fresh NiO/ZnO is 18.5 nm and 18.2
nm, respectively, which shows that the reduction treatment
caused little changes in the crystallite size or in the
structure of adsorbents.
The XRD spectrum of the sulfurized adsorbents in Fig. 1
(c) shows that the reections at 2 = 28.58, 47.56, and
56.59 related to ZnS (ICDD/JCPDS card 65-0395)
appear. The weak diffraction peak in the range of 2 =
2931 can be attributed to NiS. The result indicates that
thiophene can react with Ni to form NiS. Then the NiS
reacts with H
2
to form H
2
S. H
2
S reacts rapidly with ZnO to
form ZnS [19]. Most of the ZnO and Ni are transformed to
ZnS and NiS, respectively. Furthermore, a diffraction peak
at 2 = 44.88 can belong to the pattern of a Ni
3
C (ICDD/
JCPDS card 06-0697). Thus, after 80 h
1
, the surface of the
nickel atoms become blocked with carbon deposits, which
is another reason for the deactivation of the adsorbents.
3.1.2 BET surface area
The specic surface area and pore volume of the adsorbents
are summarized in Table 1. The specic surface area and
pore volume of the Ni/ZnO adsorbent are 28 m
2
/g and
0.13 cm
3
/g respectively, which are larger than the values of
adsorbents prepared by incipient impregnation [20].
As shown in Table 1, the specic surface area and pore
volume of the regenerated NiO/ZnO adsorbents are the
same as those of the calcined NiO/ZnO adsorbent, whereas
those of the reduced adsorbents Ni/ZnO are a little larger
than those of the fresh NiO/ZnO adsorbent. The increase of
the specic surface and pore volume in the reduced
adsorbents lead to more contacting area in the adsorbent
and reduces the resistance of thiophene diffusion in the
adsorbent. After suldation, the specic surface area and
pore volume of the adsorbents decreased, which are the
primary reasons for the deactivation of the Ni/ZnO
adsorbent.
3.1.3 Scanning electron microscopy
Scanning electron microscopy(SEM) of the Ni/ZnO are
shown in Fig. 2. The particles of the sample are uniform
with an average particle size 2730 nm with typical round
morphologies. This indicates that the active metal is well-
distributed. SEM results indicate that the homogeneous
precipitation method can enable the preparation of
homogeneous and small-sized Ni/ZnO materials.
3.2 Effect of operating conditions
3.2.1 Effect of reaction temperature
Figures 3 and 4 show the effects of the reaction
temperature on the desulfurization capacity. As shown in
the Fig. 3, when the temperature increases from 200C to
400C, the sulfur concentration at the outlet decreases
Fig. 1 XRD patterns of adsorbents. (a) after calcinations; (b)
after reduction at T = 450C and P = 0.5 MPa for 2 h; (c) after
suldation at T = 400C, P = 1.0 MPa, LHSV = 60 h
1
, and H/O =
0.3 for 80 h
172 Front. Chem. Sci. Eng. 2013, 7(2): 170176
rapidly from 84 ppmw to 0 ppmw in the rst three hours.
When the temperature increases from 400C to 500C, the
sulfur concentration at the outlet rises. Figure 4 shows the
breakthrough curves of the model gasoline on Ni/ZnO at
300C, 400C and 500C. Ni/ZnO does not adsorb the
thiophene effectively at 300C (Fig. 4(a)). Breakthrough
time is 36 h at 400C (Fig 4(b)), and the corresponding
breakthrough sulfur capacity is 360 mg-s/g-sorb. The
desulfurization capability of the adsorbent at 500C is
less than that at 400C as shown in Fig. 4(c). The decreased
desulfurization capability at higher temperatures may be
due to changes in the pore structure of adsorbents. The
result shows that the temperature is a key factor to
inuence the sulfur capacity. The best temperature for the
reactive adsorption is 400C.
3.2.2 Effect of pressure
Figures 5 and 6 show effects of the total pressure on the
desulfurization capability. As shown in Fig. 5, when the
pressure is lower than 0.8 MPa, the sulfur concentration of
the outlet stream decreases with the pressure increases.
When the pressure is higher than 0.8 MPa, no sulfur is
detected at the outlet, indicating that deep desulfurization
capability increases with the pressure increasing. As shown
in Fig. 6, when the pressure increases from 0.5 MPa to 1.0
MPa, the breakthrough time rapidly increases from 11 to
36 h. When the pressure increases from 1.0 MPa to 1.5
MPa, the breakthrough time increases from 36 h to 38 h
because the concentrations of hydrogen and hydrocarbon
on the surface of the adsorbent have reached saturation. In
Table 1 Specic surface area (S
BET
) and pore volume (V
p
) of samples
Sample S
BET
/(m
2
$g
1
) V
p
/(cm
3
$g
1
)
Calcined adsorbent 27 0.13
Regenerated adsorbent 27 0.13
Reduced adsorbent 28 0.14
Sulfurized adsorbent 14 0.05
Fig. 2 SEM micrograph of Ni/ZnO adsorbent
Fig. 3 Sulfur concentration of Ni/ZnO as a function of
adsorption temperature at LHSV = 60 h
1
, H/O = 0.3 and sulfur
content in feed (A) = 100 mg/L
Fig. 4 Breakthrough curves over Ni/ZnO adsorbents at P = 1.0
MPa, LHSV = 60 h
1
and H/O = 0.3. (a) T = 300C ; (b) T = 400C;
(c) T = 500C
Aihua KONG et al. Deep desulfurization of gasoline over Ni/ZnO adsorbent 173
order to realize deep desulfurization and high breakthrough
sulfur capacity, the optimal pressure is 1.0 MPa.
3.2.3 Effect of LHSV
The effect of LHSVon the reactive adsorption desulfuriza-
tion is shown in Fig. 7. When the LHSVincreases from 64
h
1
to 73 h
1
the breakthrough time decreases by 89%.
Although the adsorbent may have a longer breakthrough
time because of longer contact time when the LHSV is
lower than 60 h
1
, the adsorbent has a lower productivity
than that with a LHSV of 60 h
1
. Therefore, the optimal
LHSVis 60 h
1
which has a larger productivity. A LHSVof
60 h
1
is signicantly larger than a LHSVof 8.5 h
1
which
has been recently reported for a similar system [21].
3.2.4 Optimal operating conditions
According to the above study, high temperature and high
pressure are adverse to improve the desulfurization ability
of the adsorbent. In order to realize the deep desulfuriza-
tion and the high capacity, the reactive temperature of
400C, total pressure of 1.0 MPa, LHSV of 60 h
1
are
chosen as the optimal operating conditions. Under the
optimal operating conditions, the breakthrough sulfur
capacity of adsorbents is 360 mg-s/g-sorb which is ream-
arkly higher than what has been reported in the literature
[21].
3.3 Effects of model compounds on adsorption
desulfurization performance
In the commercial gasoline it contains various compounds,
mainly comprising of parafns, cycloalkanes, olens and
aromatic compounds. These compounds, especially the
olens and aromatic compounds, which have similar
electronic structure with thiophene, may affect the
desulfurization performance. And the effects of com-
pounds in model gasoline on the desulfurization capability
over adsorbent Ni/ZnO were examined by using a model
gasoline B. The adsorptive desulfurization of model
gasoline A and model gasoline B over Ni/ZnO was
conducted at 673 K, 1.0 MPa, H/O = 0.3, LHSVof 60 h
1
.
The breakthrough curves are shown in Fig. 8. The
breakthrough time decreased signicantly from 36 h for
model gasoline A to 29 h for model gasoline B and the
corresponding capacity reduced from 360 mg-s/-sorb for
model gasoline A to 322 mg-s/g-sorb for model gasoline
B. This adsorptive breakthrough capacity for model
Fig. 5 Sulfur concentration of Ni/ZnO depending on adsorption
pressure at LHSV= 60 h
1
, H/O = 0.3, and sulfur content in feed
(A) = 100 mg/L
Fig. 7 Sulfur concentration of Ni/ZnO depending on LHSVs at T
= 400C, P = 1.0 MPa, H/O = 0.3, and sulfur content in feed (A) =
100 mg/L
Fig. 6 Breakthrough curves over Ni/ZnO adsorbents at 400C,
LHSV= 60 h
1
and H/O = 0.3. (a) P = 0.5 MPa; (b) P = 1.0 MPa;
(c) P = 1.5 Mpa
174 Front. Chem. Sci. Eng. 2013, 7(2): 170176
gasoline A is much higher than that in the absence of the
additives such as cycloalkane, olen and aromatic
compound for model gasoline B, indicating that the
coexisting additives in model gasoline B decreases the
desulfurization performance over Ni/ZnO.
In order to determine the impact of the various
components of the model gasoline B on desulfurization
performance, We analyzed the change of components
contents with reaction time. Results are shown in Fig. 9.
The gure shows that the concentrations of cycloalkane and
toluene are almost close to the initial value. It implies that
cyclopalkane and toluene do not react on the Ni/ZnO
adsorbent. However, the isopentene content at the outlet
gradually increases over adsorption time in the rst 30
hours, and gradually to a stable value (34%) during the time
from30 h to 65 h. According to the Fig. 9, a minimumolen
saturation with a minimal octane loss is 8.9 wt-% at the
breakthrough sulfur level of 10 ppmw. From the results
discussed above, it shows that the decrease of total olen
content is the main factor affecting the octane number of the
desulfurization products. Also, it indicates that the exsiting
isopentene react over the Ni/ZnO. According to the
literature [18], the rate determining step of reactive
adsorption on Ni/ZnO is the hydrodesulfurization (HDS)
of thiophene molecules on Ni particles, implying that Ni
atoms of Ni/ZnO have hydrogenation activity. Therefore,
part of the isopentene with the C= C double was adsorbed
to the Ni/ZnO adsorbent under H
2
atomasphere and take
part in the hydrogenation reaction. The adsorbed isopentene
take up part of the Ni active sites of the adsorbent, which
leads to decrease of desulfurization performance of Ni/ZnO.
Therefore, the existing olens in the model gasoline B
slightly decrease the total adsorption capacity from 360
322 mg-s/g-sorb, but also resulted in the decrease in the
selectivity for various sulfur compounds over olens, and
slight loss of the octane number. Thus, Ni/ZnO adsorbent
prepared by homogeneous precipitation is an effective
adsorbent for the reactive adsorption desulfurization with
slight loss of the octane number.
4 Conclusions
The Ni/ZnO sorbent prepared by the homogeneous
precipitation method is an effective adsorbent for the
reactive adsorption desulfurization. XRD analysis shows
that the thiophene reacts with Ni to form NiS
,
and most of
the removed sulfur is accumulated in adsorbents as ZnS.
Moreover, the desulfurization capacity of the adsorbent is
weakened signicantly due to the suldation of Ni and
ZnO. The best operating conditions for desulfurization
over Ni/ZnO in a xed bed reactor are 400 C, a LHSVof
60 h
1
and a pressure of 1.0 MPa. Under the optimal
operating conditions, the adsorbent Ni/ZnO has a good
desulfurizating performance for model gasoline. Existing
olens in gasoline have a slightly inhibiting effect on the
desulfurization performance of the Ni/ZnO adsorbent.
Acknowledgements The authors thank the Program for Changjiang
Scholars and Innovative Research Team in University (IRT0936).
References
1. Ma X L, Zhou A, Song C S. A novel method for oxidative
desulfurization of liquid hydrocarbon fuels based on catalytic
oxidation using molecular oxygen coupled with selective adsorp-
tion. Catalysis Today, 2007, 123(14): 276284
2. Ito E, van Veen J A R. On novel processes for removing sulphur
from renery streams. Catalysis Today, 2006, 116(4): 446460
3. Parkinson G. Diesel desulfurization puts reners in a quandary.
Chemical Engineering, 2001, 108(2): 3742
Fig. 8 Breakthrough curves of Ni/ZnO adsorbent at T = 400C,
P = 1.0 MPa, LHSV = 60 h
1
, H/O = 0.3, and sulfur content in two
model gasoline 100 ppmw
Fig. 9 Dependence of hydrocarbon group composition of model
gasoline B product on the adsorption time at T = 573 K, P = 1.0
MPa, LHSV= 60 h
1
, H/O = 0.3
Aihua KONG et al. Deep desulfurization of gasoline over Ni/ZnO adsorbent 175
4. Song C S. An overview of new approaches to deep desulfurization
for ultra-clean gasoline, diesel and jet fuel. Catalysis Today, 2003,
86(14): 211263
5. Yang R T, Hernandez-Maldonado A J, Yang F H. Desulfurization of
transportation fuels with zeolites under ambient conditions. Science,
2003, 301(5629): 7981
6. Babich I V, Moulijn J A. Science and technology of novel process
for deep desulfurization of oil renery streams: a review. Fuel, 2003,
82(6): 607631
7. Bezverkhyy I, Ryzhikov A, Gadacz G, Bellat J P. Kinetics of
thiophene reactive adsorption on Ni/SiO
2
and Ni/ZnO. Catalysis
Today, 2008, 130(1): 199205
8. Khare G P. US Patent, 6274533, 2001-08-14
9. Shi Y H. The desulfurization in petroleum rening. Beijing:
Petrochemical Press, 2009, 187188 (in Chinese)
10. Fan J, Wang G, Sun Y, Xu C, Zhou H, Zhou G, Gao J. Research on
reactive adsorption desulfurization over Ni/ZnO- SiO
2
-Al
2
O
3
adsorbent in a xed-uidized bed reactor. Industrial & Engineering
Chemistry Research, 2010, 49(18): 84508460
11. Xu W Q, Xiong C Q, Zhou G L, Zhou H J. Removal of sulfur from
FCC gasoline by using Ni/ZnO as adsorbent. Acta Petrolei Sinica,
2008, 24(6): 739743 (Petroleum Processing Section)
12. Tawara K, Nishimura T, Iwanami H, Nishimoto T, Hasuike T. New
hydrodesulfurization catalyst for petroleum-fed fuel cell vehicles
and cogenerations. Industrial & Engineering Chemistry Research,
2001, 40(10): 23672370
13. Yan B, Li L L, Shao C H. Studies and progress in preparation
methods of nanometer ZnO desulfurizer used at ambient tempera-
ture. Journal of Natural Science of Heilongjiang University, 2006,
23(2): 163167 (in Chinese)
14. Soler-Illia G J de A A, Candal R J, Regazzoni A E, Blesa M A.
Synthesis of mixed copper-zinc basic carbonates and Zn-doped
tenorite by homogeneous alkalinization. Chemistry of Materials,
1997, 9(1): 184191
15. Candal R J, Regazzoni A E, Blesa M A. Precipitation of copper (II)
hydrous oxides and copper (II) basic salts. Journal of Materials
Chemistry, 1992, 2(6): 657661
16. Huey-Ing C, Hung Y C. Homogeneous precipitation of cerium
dioxide nanoparticles in alcohol/water mixed solvents. Colloids and
Surfaces A: Physicochemical and Engineering Aspects, 2004, 242
(13): 6169
17. Buchanan J S, Nicholas M E. Analysis of olenic gasolines with
multidimensional gas chromatography. Journal of Chromatographic
Science, 1994, 32(5): 199203
18. Ryzhikov A, Bezverkhyy I, Bellat J P. Reactive adsorption of
thiophene on Ni/ZnO: role of hydrogen pretreatment and nature of
the rate determining step. Applied Catalysis B: Environmental,
2008, 84(34): 766772
19. Tawara K, Nishimura T, Iwanami H, Nishimoto T, Hasuike T. New
hydrodesulfurization catalyst for petroleum-fed fuel cell vechicles
and cogenerations. Industrial & Engineering Chemistry Research,
2001, 40(10): 23672370
20. Yang J, Shi Q J, Li B Y. Properties of Ni supported on ZnO-based
mixed oxides for hydrodesulfurization of thiophene. Journal of
Nanchang University (Natural Science), 2009, 33(1): 4245 (in
Chinese)
21. Fan J X, Wang G, Sun Y, Xu C M, Gao J S. Research on reactive
adsorption desulfurization over of Ni/ZnO-SiO
2
-Al
2
O
3
adsorbent in
a Fixedxed-uidized bed reactor. Industrial & Engineering
Chemistry Research, 2010, 49(18): 84578460
176 Front. Chem. Sci. Eng. 2013, 7(2): 170176

You might also like