Anda di halaman 1dari 8

Chemical Engineering Journal 251 (2014) 3542

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Adsorptive denitrogenation of model fuel with CuCl-loaded


metalorganic frameworks (MOFs)
Imteaz Ahmed, Sung Hwa Jhung
Department of Chemistry and Green-Nano Materials Research Center, Kyungpook National University, Daegu 702-701, Republic of Korea

h i g h l i g h t s

g r a p h i c a l a b s t r a c t

 Cu(I) was introduced on a metal

organic framework at ambient


condition.
 Introduced Cu(I) improved the
performances of adsorptive
denitrogenation.
 Supported Cu(I) increased adsorption
capacity even though decreased
porosity.
 Supported Cu(II) decreased both
porosity and adsorption capacity.

a r t i c l e

i n f o

Article history:
Received 7 March 2014
Received in revised form 8 April 2014
Accepted 10 April 2014
Available online 24 April 2014
Keywords:
Adsorptive denitrogenation
Cuprous ion
Metalorganic framework
p-Complexation

a b s t r a c t
CuCl impregnated MIL-100(Cr) was prepared with a facile method and then utilized for the adsorptive
denitrogenation of model fossil fuels. CuCl was produced from CuCl2 solution at ambient temperature
and pressure by reduction of CuCl2 using sodium sulte as a mild reducing agent. Although the porosity of the adsorbent was reduced after impregnation, the adsorption of nitrogen-containing compounds (NCCs) increased. However, the adsorption of NCCs over CuCl2/MIL-100(Cr), which was
produced in a similar manner but in the absence of Na2SO3, was decreased as a result of the reduced
porosity. The maximum adsorption capacities of CuCl/MIL-100(Cr) as compared to pristine MIL100(Cr) for quinoline (QUI) and indole (IND) were improved by 9% and 15%, respectively, which
may be attributed to the p-complexation effect of the Cu+ sites of CuCl.
2014 Elsevier B.V. All rights reserved.

1. Introduction
With the continuous increase in the worldwide population and
consumption of energy, the optimized utilization of energy and
stable environmental parameters are the primary concerns of the
human race in this century. Because of the scarcity of regular
sources, new and unusual sources of fossil fuels need to be
exploited in the upcoming decades for the high amount of energy
required by society. Therefore, fossil fuels containing high amounts
of contaminants, particularly sulfur- and nitrogen-containing
compounds (SCCs and NCCs), need to be utilized [14]. Acid rain
Corresponding author. Fax: +82 53 950 6330.
E-mail address: sung@knu.ac.kr (S.H. Jhung).
http://dx.doi.org/10.1016/j.cej.2014.04.044
1385-8947/ 2014 Elsevier B.V. All rights reserved.

produced from SCCs and NCCs is also a serious environmental


concern and many governments across the globe are becoming
stricter regarding the emission of these compounds. Therefore,
the removal of SCCs and NCCs from fossil fuel is an important topic
of research within the scientic community.
The removal of SCCs from fossil fuel has been extensively
studied over the past several decades and these compounds are
typically removed by the hydrodesulfurization (HDS) process.
NCCs [5,6] are one of the contaminants that should be removed
before the removal of SCCs because NCCs have an adverse effect
on both catalysts and conventional HDS processes [57]. NCCs
are typically removed by hydrodenitrogenation (HDN) in the
presence of expensive hydrogen at high temperatures and pressures; thus, HDN is an energy-intensive and costly process.

36

I. Ahmed, S.H. Jhung / Chemical Engineering Journal 251 (2014) 3542

Additionally, HDN, compared with HDS, is an inefcient process


because the nitrogen in NCCs can only be removed after the heterocyclic-rings have been hydrogenated. Therefore, a high amount
of hydrogen is consumed during the HDN process and this process,
compared with HDS, is a kinetically slow process [5,6,8]. Therefore,
an alternative method to remove NCCs from fuels is very important, and adsorptive-denitrogenation (ADN) is one of the choices
that can be used to remove NCCs because this procedure does
not require expensive hydrogen, high temperatures or high pressures and is therefore more cost effective than other processes. A
wide variety of adsorbents have been used for ADN thus far including activated carbons [5,6,913], Cu(I)-Y zeolites [7,14], HCl-loaded
silicaaluminas [15], ion exchange resins [16], meso-silicas [17,18]
including Ti-HMSs [19], microporous carbons [20], activated aluminas [12], Ni-based adsorbents [12], and NiMoS [21].
As the eld of materials chemistry continues to grow, many
functional materials are rapidly emerging and porous materials
such as metalorganic frameworks (MOFs) are one of the fastest
growing elds of research [2231]. Interest in MOF materials is
a result of the relatively simple tunability of their pore size and
shape from the microporous to the mesoporous scale upon altering the connectivity of the inorganic moieties and the nature of
the organic linkers. MOFs can be used for a plethora applications
which contain gas adsorption/storage such as adsorptive desulfurization (ADS) [3238] and ADN [3943], separation, catalysis, drug
delivery, luminescence, electrode materials, carriers for nanomaterials, magnetism, polymerization, imaging [4447]. MOFs are
promising materials for adsorption-related applications because
their pore surfaces can be easily modied which leads to the
selective adsorption of specic guest molecules containing particular functional groups. Maes et al. [41] conrmed that MOFs could
be used in ADN and/or ADS processes if they are selected and
reduced appropriately. Although MOFs alone exhibit many promising physical and chemical properties in various aspects, their
properties can be further improved by modifying or specifying
its structure or chemical nature in various ways. The synthesis
kinetics, morphology, physicochemical properties, and potential
applications of MOFs can be vastly improved by modifying the
MOFs with suitable materials [4852]. MOFs grafted or impregnated with acidic and basic materials can affect the adsorption
behavior of SSCs and NCCs due to acidbase interactions
[39,42,52]. Several studies have been also carried out using MOFs
impregnated with suitable functional materials and a large
portion of them were capable of the adsorptive removal of certain
contaminants [5356].
Porous adsorbents containing transition metal ions such as Cu+,
Ag+, Pd2+ or Pt2+ have shown that they can effectively remove SCCs
and NCCs from commercial fuels through p-complexation which is
a special bond between specic metal ions and the p-orbital of
olens and aromatics [34,35,5763]. In particular, materials containing Cu+ have been widely investigated and applied because of
their promising p-complexing properties, easy availability and
low cost. Although, adsorptive removal of SCCs is quite common
with these materials but there have been very few reports on their
use in ADN [7,14]. Additionally, NCCs can be easily removed compared to SCCs due to the selectivity of the adsorption of NCCs over
SCCs [39,4143]. However, p-complexation adsorbents have typically been prepared through the high temperature calcination of
metal-ion-exchanged materials which requires excessive energy
and subsequently causes the process to be quite expensive [59].
Therefore, this process would be more feasible if the materials
can be prepared or impregnated under ambient temperature and
pressure. For example, loading specic metal salts onto porous
support materials through re-crystallization could be an effective
method to incorporate p-complexing metal ions. Furthermore,
and most importantly, porous materials with low thermal and

chemical stabilities cannot be used as a support material for the


loading of active metal components because of the requirement
of high temperature calcination or harsh chemical treatment. For
example, porous materials such as MOFs, because of their low
thermal and chemical stability [64], cannot be used as support
materials for these types of active metal components through high
temperature calcination [30,31,65]. In order to utilize their promising adsorptive properties, MOFs need to be exploited by mild,
easy and cost effective way suitable for them.
In this report, we demonstrated for the rst time the loading of
CuCl onto a porous MOF, MIL-100(Cr) through the room temperature reduction of CuCl2 in the presence of sodium sulte (Na2SO3)
as the mild reducing agent. MIL-100(Cr) is one of the highly porous
MOFs having composition of Cr3F(H2O)3O[C6H3-(CO2)3]2nH2O
(n 28) [66] with various applications. The loading procedure
was very simple to apply and did not require much energy. After
the preparation, it was successfully utilized for the selective
removal of NCCs such as quinoline (QUI) and indole (IND) by
adsorption from a model fuel containing NCCs, SCCs and aromatics.

2. Experimental
2.1. Chemicals and synthesis of adsorbents
p-Xylene and n-octane were purchased from Junsei chemical
company. Hydrochloric acid (HCl, 35%) and hydrouoric acid (HF,
48.0%) were obtained from OCI Company Ltd. Benzothiophene
(BT), quinoline (QUI), indole (IND), trimesic acid (H3BTC) and
metallic chromium (Cr) were obtained from SigmaAldrich Co.
All the chemicals in this study were used without further
purication. The syntheses were carried out solvothermally under
microwave irradiation [67]. Detailed synthetic procedures for
MIL-100(Cr) are listed below. Metallic chromium, HF, H3BTC and
H2O were mixed at a molar ratio of 1:2:0.67:265. The mixture
was then transferred to a Teon-lined autoclave and heated in a
microwave oven (Mars-5, CEM) for 2 h at 220 C. After synthesis,
the MOF was ltered, washed, and puried by stirring at rst with
water then with ethanol and nally dried in a drying oven at
100 C.
An exact amount (0.1 g) of MIL-100(Cr) was transferred to glass
vials containing 0.5, 1.0 or 2.0 mL of a 0.2 M CuCl2 aqueous solution
and stirred magnetically for 30 min at room temperature. An equal
amount (0.5, 1.0, or 2.0 mL) of a 0.1 M Na2SO3 aqueous solution
was then added dropwise while stirring to the CuCl2 solution containing MIL-100(Cr). The resulting mixture was allowed to stir for
30 min and then 1 mL of a sulfurous acid (H2SO3) solution was
added. The sulfurous acid solution was prepared by mixing
25.0 mL of 0.25 mM Na2SO3 and 3.0 mL of 0.2 M HCl. The mixture
was then ltered and washed with sulfurous acid solution. Finally,
the wet solids were transferred to a vacuum oven (to avoid air contact) and dried overnight at 100 C. The adsorbents were named
CuCl (x)/MIL-100(Cr) where x corresponds to the volume of the
CuCl2 (or Na2SO3) solution used in the preparation.

2.2. Characterization
The X-ray powder diffraction patterns were obtained using a
diffractometer D2 Phaser (Bruker, with Cu Ka radiation). The nitrogen adsorption experiments of the adsorbents were carried out at
196 C with a surface area and porosity analyzer (Micromeritics,
Tristar II 3020) after evacuation at 150 C for 12 h. XPS analyses
were carried out using a Quantera SXM X-ray photoelectron
spectrometer (ULVAC-PHI) equipped with a dual beam charge
neutralizer.

37

I. Ahmed, S.H. Jhung / Chemical Engineering Journal 251 (2014) 3542

Stock solutions (each at a concentration of 10,000 ppm) of the


model fuel were prepared separately for the three adsorbates
(BT, QUI, and IND) by dissolving them in a mixture of 75 vol%
n-octane and 25 vol% p-xylene. Combined solutions of varying
concentrations were prepared by successive dilution and mixing
of the three solutions. A solution containing a xed concentration
of NCCs and SCCs (400 ppm QUI, 400 ppm IND, and 800 ppm BT)
was used to determine the adsorption capacity at various adsorption times. To calculate the adsorption capacities of the individual
adsorbates, a solution of QUI and IND was prepared separately at
varying concentrations (3001200 ppm) in the same solvent mixture. Prior to adsorption, the adsorbents were dried in a vacuum
oven at 150 C for 12 h and stored in a desiccator. For each adsorption experiment, an exact amount of the adsorbents (5.0 mg) was
added to the model fuel solution (5.0 mL) and stirred magnetically for a predetermined amount of time (2120 min) while the
adsorption temperature was maintained at 25 C. After adsorption,
the solution was separated from the solid using a syringe lter
(PTFE, hydrophobic, 0.5 lm) and analyzed with a GC (DS Science,
IGC 7200) equipped with an FID.
All the adsorption capacities (mg/g) were calculated from the
difference between nal concentration and initial concentration
of an adsorbate by using following equation:

C i  C f V
qt
m

(a)

Intensity (a.u.)

MIL-100 (Cr)
CuCl (1.0)/MIl-100 (Cr)
CuCl
CuCl2

15

30

where qt is the adsorbed amount in time t (mg/g), Ci the initial


concentration of the adsorbate (mg/ml), Cf the nal concentration
after adsorption (mg/ml), V the volume of the solution subjected
to a single adsorption (ml) and m is the mass of the adsorbent taken
during a single adsorption (g).
The maximum adsorption capacity (Qo) was calculated with the
Langmuir adsorption isotherm. The adsorption isotherms under
various conditions have been plotted to follow the Langmuir
equation [68,69]:

Ce Ce
1

qe Q o Q o b
where Ce is the equilibrium concentration of the adsorbate (mg/L),
qe the amount adsorbed at equilibrium (measured similarly as qt,
mg/g), Qo the Langmuir constant (maximum adsorption capacity,
mg/g) and b is the Langmuir constant (L/mg).
Therefore, the maximum adsorption capacity Qo can be obtained
from the reciprocal of the slope of a plot of Ce/qe against Ce.
3. Results and discussion
3.1. Textural properties and characteristics of CuCl/MIL-100(Cr)
Fig. 1(a) displays the XRD patterns of MIL-100(Cr) and CuCl
(1.0)/MIL-100(Cr), along with the XRD patterns of CuCl and CuCl2,
which exhibited a similar structure of the virgin and impregnated

(b)

CuCl (1.0)/MIL-100 (Cr)


CuCl
+
CuCl2
Cu

Cu

Cu

Intensity (a.u)

2.3. General procedures for the adsorption experiments

45

960

2 theta (deg)

940

BE (eV)

Fig. 1. (a) XRD patterns and (b) nitrogen XPS spectra of MIL-100(Cr) and CuCl/MIL-100(Cr). Pure CuCl and CuCl2 are also included in both gures for reference.

600

(b)
6

450

300

MIL-100 (Cr)
CuCl (1.0)/MIl-100 (Cr)

dV/dlog (D)

Quantity adsorbed (cm3/g)

(a)

MIL-100 (Cr)
CuCl (1.0)/MIl-100 (Cr)
4

150

0
0.0

0.2

0.4

0.6

0.8

Relative pressure (P/Po)

1.0

15

30

45

Pore Diameter ()

Fig. 2. (a) N2 adsorption isotherms and (b) pore size distributions of MIL-100(Cr) and CuCl (1.0)/MIL-100(Cr).

60

38

I. Ahmed, S.H. Jhung / Chemical Engineering Journal 251 (2014) 3542

Table 1
Textural properties of MIL-100(Cr) and CuCl (1.0)/MIL-100(Cr) and Langmuir parameters of QUI and IND adsorption over the two adsorbents.
Adsorbent

BET surface area (m2/g)

Total pore volume (cm3/g)

Micropore volume (cm3/g)

Adsorbate

Qo (mg/g)

b-Value (L/mg)

MIL-100(Cr)

1510

0.985

0.402

CuCl (1.0)/MIL-100(Cr)

1310

0.635

0.399

QUI
IND
QUI
IND

420
149
457
171

7.06  103
3.14  103
1.03  102
3.22  103

30

250

(a)

(b)

MIL-100 (Cr)
CuCl (0.5)/MIl-100 (Cr)
CuCl (1.0)/MIl-100 (Cr)
CuCl (2.0)/MIl-100 (Cr)

qt (mg/g)

qt (mg/g)

20

200

150

100

10
50

0
0

Time (hr)

Time (hr)
60

(c)

qt (mg/g)

45

30

15

0
0

Time (hr)
Fig. 3. Adsorption of (a) BT, (b) QUI and (c) IND with time over MIL-100(Cr) and CuCl/MIL-100(Cr)s.

materials. CuCl shows peaks at angles 24.3 and 43.1 while CuCl2
shows peaks mainly at 16, 26.3, 31.2, 38.5 and 49.2. No
diffraction patterns for CuCl2 were observed because of the reduction of CuCl2 into CuCl or well-dispersed CuCl2. The absence of CuCl
diffraction patterns (the presence of CuCl was conrmed by XPS,
see below) indicated that a uniform, well-dispersed, and small
amount of CuCl was present within the pores of MIL-100(Cr). The
nitrogen adsorption analyses also supported the XRD data because
there was small effect of impregnation procedure on the surface
area or pore volume of the MOF, as shown in Fig. 2. Table 1 and
Fig. 2 show that the surface area and the pore volume of CuCl
(1.0)/MIL-100(Cr) were slightly smaller than those of the virgin
MIL-100(Cr). Moreover, the loading procedure affected the pore
structure of the MOF a bit because the pore volume (particularly
the large pore) decreased for the loaded MIL-100(Cr). The impregnated salt was in the Cu+ state which was conrmed by the XPS
analysis (Fig. 1(b)). The XPS patterns of impregnated adsorbent
was compared to those of standard CuCl and CuCl2 [70] and only
the peaks corresponding to CuCl were found in the impregnated
adsorbent.

3.2. Adsorption results


As mentioned earlier, NCCs should be removed prior to the
removal of SCCs. Therefore, in order to investigate the selectivity
of the adsorption, a solution containing high concentration of BT
as an SCC was prepared (along with QUI and IND) because SCCs
are usually present at a high concentration in fuels relative to
NCCs. Additionally, commercial fuels usually contain approximately 25% aromatics; therefore, p-xylene was used as an aromatic
compound to mimic the composition of the commercial fuels. The
adsorption experiments were carried out at 25 C over a period of
6 h. Fig. 3 shows the adsorption results of different adsorbates over
the various adsorbents. It was observed that both QUI and IND
were selectively adsorbed by both the virgin and impregnated
MIL-100(Cr) to a much greater extent than BT in the combined system despite the BT concentration being much higher than QUI or
IND. Therefore, it was very difcult to observe any trend of BT
adsorption. The usual preference for the adsorption of fossil fuel
components is NCCs > SCCs > aromatic hydrocarbons. The selective
adsorption of NCCs over SCCs could be explained by much more

39

I. Ahmed, S.H. Jhung / Chemical Engineering Journal 251 (2014) 3542

250

60

(a)

(b)

200

qt (mg/g)

qt (mg/g)

40
150

100

0.5 h
2h
4h
6h

50

20

0
0.0

0.5

1.0

1.5

2.0

0.0

CuCl2 Loading amount (ml)

0.5

1.0

1.5

2.0

CuCl2 Loading amount (ml)

Fig. 4. Adsorbed amounts of (a) QUI and (b) IND over CuCl/MIL-100(Cr)s having different amount of CuCl in various adsorption times.

450

150

(a)

(b)

300

qe (mg/g)

qe (mg/g)

100

MIL-100 (Cr)
CuCl (1.0)/MIl-100 (Cr)

150

50

0
0

300

600

900

1200

300

600

Ce (ppm)

900

1200

Ce (ppm)

Fig. 5. Adsorption isotherms for (a) QUI and (b) IND over MIL-100(Cr) and CuCl (1.0)/MIL-100(Cr).

3.0

(b)

(a)
9

2.5

R = 0.984

R = 0.995

1.5

ce/qe

ce/qe

2.0

R = 0.997

6
2

R = 0.995

1.0
3
0.5

MIL-100 (Cr)
CuCl (1.0)/MIl-100 (Cr)

0.0

0
0

400

800

1200

Ce (ppm)

400

800

1200

Ce (ppm)

Fig. 6. Langmuir plots obtained from adsorption isotherms for (a) QUI and (b) IND over MIL-100(Cr) and CuCl (1.0)/MIL-100(Cr). Correlation factors (R2) are shown alongside
the respective plots.

negative adsorption enthalpies of NCCs compared with SCCs


according to a previous report [7]. Acidbase interaction also
favors the adsorption of NCCs over SCCs [39,41,42]. Therefore, it
can be concluded that the selectivity of NCCs for the adsorbents
was much higher compared to the SCCs. The kinetics between different adsorbents did not vary much due to the relatively large

pore sizes (1.5 nm, as shown in Fig. 2) of the adsorbents. However, the overall kinetics of the adsorption was remarkably fast
which conrms the favorable adsorption of NCCs over the adsorbents. Fig. 4 displays the relationship between the adsorbed
amount of NCCs and the CuCl2 content (used in impregnation),
which was adopted from Fig. 3. The adsorbed amounts of both

40

I. Ahmed, S.H. Jhung / Chemical Engineering Journal 251 (2014) 3542

300

75

(a)

(b)

1 hr
6 hr

225

qt (mg/g)

qt (mg/g)

50
150

25
75

0
MIL-100 (Cr)

CuCl/MIL-100 (Cr) CuCl2 /MIL-100 (Cr)

MIL-100 (Cr)

Adsorbent

CuCl/MIL-100 (Cr) CuCl2 /MIL-100 (Cr)

Adsorbent

Fig. 7. Comparison of adsorptions of (a) QUI and (b) IND over MIL-100(Cr), CuCl (1.0)/MIL-100(Cr) and CuCl2 (1.0)/MIL-100(Cr).

QUI and IND generally increased as the loading of CuCl2 increased


up to a certain point, and then decreased as the amount of CuCl2
further increased. In the case of short adsorption time of 0.5 h,
the adsorbed amount of QUI decreased with increasing the CuCl2
content probably caused by partial blocking of the pores with the
impregnated CuCl species. Therefore, the optimized loading
amount was determined to be 1.0 mL of the 0.2 M CuCl2 solution
(0.027 g of CuCl2) per 0.1 g of MOF; and thus CuCl (1.0)/MIL100(Cr) and virgin MIL-100(Cr) were used for further experiments.
To obtain the maximum adsorption capacities of MIL-100(Cr)
and CuCl (1.0)/MIL-100(Cr), solutions of varying concentrations
of the adsorbates QUI and IND were prepared separately and the
adsorption amount was measured. Each adsorption was carried
out for 6 h which was considered to be a sufcient amount of time
to reach the adsorption equilibrium. The isotherms which are
shown in Fig. 5 were obtained by plotting the adsorbed amount
against the equilibrium concentration of the adsorbates. The maximum adsorption capacities (Qo) were calculated using the
Langmuir plots shown in Fig. 6.
The maximum adsorption capacities along with the physicochemical properties were summarized in Table 1. For CuCl (1.0)/
MIL-100(Cr), the Qo value of QUI increased from 420 mg/g to
457 mg/g and that of IND increased from 149 mg/g to 171 mg/g
relative to the virgin MIL-100(Cr). Therefore, compared to the virgin MIL-100(Cr), the increase in the adsorption of QUI and IND was
approximately 9% and 15%, respectively, over CuCl (1.0)/MIL100(Cr). Although the porosity of the loaded MOF slightly
decreased (Table 1), the adsorption capacities were enhanced.
The b-values of the corresponding Langmuir isotherms also show
favorable adsorptions over CuCl (1.0)/MIL-100(Cr). As shown in
Table 1, the b-value for QUI adsorption over CuCl (1.0)/MIL100(Cr) was around 1.5 times of that of pristine MIL-100(Cr), and
the value was slightly increased for CuCl (1.0)/MIL-100(Cr) in the
case of IND. These improvements could be attributed to the selective interaction between the NCCs and the loaded CuCl materials
through p-complexation. The p-complexation occurred as a result
of the interaction between the p-electrons of NCCs and Cu+ ions
with d-electrons and empty s-orbital [7]. Recently, adsorbents
using p-complexation have been extensively investigated for
adsorptive desulfurization processes since they allow higher selectivities and capacities for organosulfur compounds compared to
normal physisorption processes [5763,71]. It has been well
reported that Cu+ sites impregnated into suitable materials can signicantly improve the adsorption of SCCs as well as NCCs
[35,43,57,60].
There have been previous reports detailing the preparation of
p-complexing materials by impregnating CuCl with the help of

strong mineral acids [57,60], the thermal reduction of loaded CuCl2


to CuCl [72] or a mild reducing procedure (from CuCl2 to CuCl) with
Na2SO3 [33]. However, using p-complexing materials (obtained
through impregnation) for the removal of NCCs is not commonly
practiced and previous studies typically utilized modied zeolites
by ion exchange (with precursors for p-complexing metal sites)
[7,14]. In this report we were able to show the ADN process by
using p-complexing materials in a new way which was carried
out under mild operation conditions. In this study, the CuCl particles were produced through the facile reduction of the loaded
CuCl2 in the presence of sodium sulte under ambient conditions
in which the sodium sulte was used as a mild reducing agent
for the redox process shown below.

2CuCl2 aq Na2 SO3 aq H2 O


! 2CuCl s Na2 SO4 aq 2HCl aq

2

or 2Cu2 SO2
3 H2 O ! 2Cu SO4 2H

The +1 oxidation state of the loaded copper species was the active
component for ADN which was demonstrated for the favorable
interaction between Cu+ and SCCs [33] or NCCs [43] through
p-complex formation.
In order to conrm the effects of p-complexation in the adsorption procedure, an adsorption experiment was also carried out
using the CuCl2 loaded onto MIL-100(Cr) without reducing the
CuCl2 with Na2SO3. In this case a combined solution of BT, QUI,
and IND was used under the same conditions as the previous
experiments. Fig. 7 shows the adsorption results of QUI and IND
over the virgin MIL-100(Cr), CuCl (1.0)/MIL-100(Cr), and CuCl2
(1.0)/MIL-100(Cr). It was found that the adsorbed amounts of
QUI and IND followed the trend of CuCl (1.0)/MIL-100(Cr) > virgin
MIL-100(Cr) > and CuCl2 (1.0)/MIL-100(Cr). The minimal adsorption of QUI and IND over CuCl2 (1.0)/MIL-100(Cr) was likely a result
of the reduced porosity and the lack of interactions between CuCl2
and the adsorbates. CuCl2 is known to not exhibit p-complexion
which makes CuCl2 loaded MIL-100(Cr) less capable of adsorption
compared to CuCl loaded MIL-100(Cr) or the virgin MIL-100(Cr).
4. Conclusion
The impregnation of the Cu+ sites onto a MOF (MIL-100(Cr))
was carried out for the rst time by reducing CuCl2 salt with
sodium sulte, which is a very mild, facile and energy efcient procedure. The impregnated material showed an improved adsorption
capacity of QUI and IND relative to the virgin MIL-100(Cr) as a
result of p-complexation. The adsorption of QUI and IND in this

I. Ahmed, S.H. Jhung / Chemical Engineering Journal 251 (2014) 3542

study was found to be improved by 9% and 15%, respectively, even


though the porosity of adsorbent decreased upon CuCl loading.
Through this study it could be conrmed that the selective
p-complexing sites like Cu+ can be impregnated in porous materials in a mild and energy efcient way, which is particularly useful
for porous materials that are not stable in harsh conditions e.g.
high temperature, etc. Moreover, this procedure can be applied
to a variety of adsorption and catalytic applications which are
dependent on the presence of Cu+ sites.
Acknowledgement
This research was supported by Basic Science Research Program
through the National Research Foundation of Korea (NRF) grant
funded by the Korea government (MSIP) (grant number:
2013R1A2A2A01007176).
References
[1] V.C. Srivastava, An evaluation of desulfurization technologies for sulfur
removal from liquid fuels, RSC Adv. 2 (2012) 759783.
[2] B. Pawelec, R.M. Navarro, J.M. Campos-Martin, J.L.G. Fierro, Towards near
zerosulfur liquid fuels: a perspective review, Catal. Sci. Technol. 1 (2011) 23
42.
[3] A. Stanislaus, A. Mara, M.S. Rana, Recent advances in the science and
technology of ultra low sulfur diesel (ULSD) production, Catal. Today 153
(2010) 168.
[4] A. Samokhvalov, B.J. Tatarchuk, Review of experimental characterization of
active sites and determination of molecular mechanisms of adsorption,
desorption and regeneration of the deep and ultradeep desulfurization
sorbents for liquid fuels, Catal. Rev.: Sci. Eng. 52 (2010) 381410.
[5] M. Almarri, X. Ma, C. Song, Role of surface oxygen-containing functional groups
in liquid-phase adsorption of nitrogen compounds on carbon-based
adsorbents, Energy Fuels 23 (2009) 39403947.
[6] G.C. Laredo, P.M. Vega-Merino, F. Trejo-Zrraga, J. Castillo, Denitrogenation of
middle distillates using adsorbent materials toward ULSD production: a
review, Fuel Process. Technol. 103 (2013) 2132.
[7] A.J. Hernndez-Maldonado, R.T. Yang, Denitrogenation of transportation fuels
by zeolites at ambient temperature and pressure, Angew. Chem. Int. Ed. 43
(2004) 10041006.
[8] S. Eijsbouts, V.H.J. De Beer, R. Prins, Hydrodenitrogenation of quinoline over
carbon-supported transition metal suldes, J. Catal. 127 (1991) 619630.
[9] Y. Sano, K.-H. Choi, Y. Korai, I. Mochida, Selection and further activation of
activated carbons for removal of nitrogen species in gas oil as a pretreatment
for its deep hydrodesulfurization, Energy Fuels 18 (2004) 644651.
[10] N. Li, M. Almarri, X.-L. Ma, Q.-F. Zha, The role of surface oxygen-containing
functional groups in liquid-phase adsorptive denitrogenation by activated
carbon, New Carbon Mater. 26 (2011) 470478.
[11] J. Wen, X. Han, H. Lin, Y. Zheng, W. Chu, A critical study on the adsorption of
heterocyclic sulfur and nitrogen compounds by activated carbon: equilibrium,
kinetics and thermodynamics, Chem. Eng. J. 164 (2010) 2936.
[12] J.H. Kim, X. Ma, A. Zhou, C. Song, Ultra-deep desulfurization and
denitrogenation of diesel fuel by selective adsorption over three different
adsorbents: a study on adsorptive selectivity and mechanism, Catal. Today 111
(2006) 7483.
[13] Y. Sano, K.-H. Choi, Y. Korai, I. Mochida, Adsorptive removal of sulfur and
nitrogen species from a straight run gas oil over activated carbons for its deep
hydrodesulfurization, Appl. Catal. B: Environ. 49 (2004) 219225.
[14] D. Liu, J. Gui, Z. Sun, Adsorption structures of heterocyclic nitrogen compounds
over Cu(I)Y zeolite: a rst principle study on mechanism of the
denitrogenation and the effect of nitrogen compounds on adsorptive
desulfurization, J. Mol. Catal. A: Chem. 291 (2008) 1721.
[15] C.A. Audeh, Removal of nitrogen compounds from lubricating oils, Ind. Eng.
Chem. Prod. Res. Dev. 22 (1983) 276279.
[16] L.-L. Xie, A. Favre-Reguillon, X.-X. Wang, X. Fu, M. Lemaire, Selective
adsorption of neutral nitrogen compounds from fuel using ion-exchange
resins, J. Chem. Eng. Data 55 (2010) 48494853.
[17] A. Koriakin, K.M. Ponvel, C.-H. Lee, Denitrogenation of raw diesel fuel by
lithium-modied mesoporous silica, Chem. Eng. J. 162 (2010) 649655.
[18] J.-M. Kwon, J.-H. Moon, Y.-S. Bae, D.-G. Lee, H.-C. Sohn, C.-H. Lee, Adsorptive
desulfurization and denitrogenation of renery fuels using mesoporous silica
adsorbents, ChemSusChem 1 (2008) 307309.
[19] H. Zhang, G. Li, Y. Jia, H. Liu, Adsorptive removal of nitrogen-containing
compounds from fuel, J. Chem. Eng. Data 55 (2010) 173177.
[20] B. Sun, G. Li, X. Wang, Facile synthesis of microporous carbon through a soft
template pathway and its performance in desulfurization and denitrogenation,
J. Nat. Gas Chem. 19 (2010) 471476.
[21] M. Sun, A.E. Nelson, J. Adjaye, First principles study of heavy oil organonitrogen
adsorption on NiMoS hydrotreating catalysts, Catal. Today 109 (2005) 4953.

41

[22] Z. Li, J.C. Barnes, A. Bosoy, J.F. Stoddart, J.I. Zink, Mesoporous silica
nanoparticles in biomedical applications, Chem. Soc. Rev. 41 (2012) 2590
2605.
[23] F. Tang, L. Li, D. Chen, Mesoporous silica nanoparticles: synthesis,
biocompatibility and drug delivery, Adv. Mater. 24 (2012) 15041534.
[24] K. Ariga, A. Vinu, Y. Yamauchi, Q. Ji, J.P. Hill, Nanoarchitectonics for
mesoporous materials, Bull. Chem. Soc. Jpn. 85 (2012) 132.
[25] A. Vinu, M. Miyahara, T. Mori, K. Ariga, Carbon nanocage: a large-pore cagetype
mesoporous carbon material as an adsorbent for biomolecules, J. Porous
Mater. 13 (2006) 379383.
[26] A. Vinu, K. Ariga, New ideas for mesoporous materials, Adv. Porous Mater. 1
(2013) 3671.
[27] G. Frey, Hybrid porous solids: past, present, future, Chem. Soc. Rev. 37 (2008)
191214.
[28] Q. Yang, D. Liu, C. Zhong, J.R. Li, Development of computational methodologies
for metalorganic frameworks and their application in gas separations, Chem.
Rev. 113 (2013) 82618323.
[29] H. Furukawa, K.E. Cordova, M. OKeeffe, O.M. Yaghi, The chemistry and
applications of metalorganic frameworks, Science 341 (2013) 1230444.
[30] N.A. Khan, Z. Hasan, S.H. Jhung, Adsorptive removal of hazardous materials
using metalorganic frameworks (MOFs): a review, J. Hazard. Mater. 244245
(2013) 444456.
[31] S.H. Jhung, N.A. Khan, Z. Hasan, Analogous porous metalorganic frameworks:
synthesis, stability and application in adsorption, CrystEngComm 14 (2012)
70997109.
[32] N.A. Khan, J.W. Jun, J.H. Jeong, S.H. Jhung, Remarkable adsorptive performance
of a metalorganic framework, vanadium-benzenedicarboxylate (MIL-47), for
benzothiophene, Chem. Commun. 47 (2011) 13061308.
[33] N.A. Khan, Z. Hasan, K.S. Min, S.-M. Paek, S.H. Jhung, Facile introduction of Cu+
on activated carbon at ambient conditions and adsorption of benzothiophene
over Cu+/activated carbon, Fuel Process. Technol. 116 (2013) 265270.
[34] N.A. Khan, S.H. Jhung, Low-temperature loading of Cu+ species over porous
metalorganic frameworks (MOFs) and adsorptive desulfurization with Cu+loaded MOFs, J. Hazard. Mater. 237238 (2012) 180185.
[35] N.A. Khan, S.H. Jhung, Remarkable adsorption capacity of CuCl2-loaded porous
vanadium benzenedicarboxylate for benzothiophene, Angew. Chem. Int. Ed. 51
(2012) 11981201.
[36] K.A. Cychosz, A.G. Wong-Foy, A.J. zMatzger, Enabling cleaner fuels:
desulfurization by adsorption to microporous coordination polymers, J. Am.
Chem. Soc. 131 (2009) 1453814543.
[37] H.-X. Zhang, H.-L. Huang, C.-X. Li, H. Meng, Y.-Z. Lu, C.-L. Zhong, D.-H. Liu, Q.-Y.
Yang, Adsorption behavior of metalorganic frameworks for thiophenic sulfur
from diesel oil, Ind. Eng. Chem. Res. 51 (2012) 1244912455.
[38] G. Blanco-Brieva, J.M. Campos-Martin, S.M. Al-Zahrani, J.L.G. Fierro,
Effectiveness of metalorganic frameworks for removal of refractory organosulfur compound present in liquid fuels, Fuel 90 (2011) 190197.
[39] I. Ahmed, Z. Hasan, N.A. Khan, S.H. Jhung, Adsorptive denitrogenation of model
fuels with porous metalorganic frameworks (MOFs): effect of acidity and
basicity of MOFs, Appl. Catal. B: Environ. 129 (2013) 123129.
[40] A.L. Nuzhdin, K.A. Kovalenko, D.N. Dybtsev, G.A. Bukhtiyarova, Removal of
nitrogen compounds from liquid hydrocarbon streams by selective sorption on
metalorganic framework MIL-101, Mendeleev Commun. 20 (2010)
5758.
[41] M. Maes, M. Trekels, M. Boulhout, S. Schouteden, F. Vermoortele, L. Alaerts, D.
Heurtaux, Y.-K. Seo, Y.K. Hwang, J.-S. Chang, I. Beurroies, R. Denoyel, K. Temst,
A. Vantomme, P. Horcajada, C. Serre, D.E. De Vos, Selective removal of N
heterocyclic aromatic contaminants from fuels by Lewis acidic metalorganic
frameworks, Angew. Chem. Int. Ed. 50 (2011) 42104214.
[42] I. Ahmed, N.A. Khan, Z. Hasan, S.H. Jhung, Adsorptive denitrogenation of model
fuels with porous metalorganic framework (MOF) MIL-101 impregnated with
phosphotungstic acid: effect of acid site inclusion, J. Hazard. Mater. 250251
(2013) 3744.
[43] I. Ahmed, N.A. Khan, S.H. Jhung, Graphite oxide/metalorganic framework
(MIL-101): remarkable performance in the adsorptive denitrogenation of
model fuels, Inorg. Chem. 52 (2013) 1415514161.
[44] A.U. Czaja, N. Trukhan, U. Mller, Industrial applications of metalorganic
frameworks, Chem. Soc. Rev. 38 (2009) 12841293.
[45] H. Wu, Q. Gong, D.H. Olson, J. Li, Commensurate adsorption of hydrocarbons
and alcohols in microporous metal organic frameworks, Chem. Rev. 112
(2012) 836868.
[46] N. Stock, S. Biswas, Synthesis of metalorganic frameworks (MOFs): routes to
various MOF topologies, morphologies, and composites, Chem. Rev. 112 (2012)
933969.
[47] P. Horcajada, R. Gref, T. Baati, P.K. Allan, G. Maurin, P. Couvreur, G. Frey, R.E.
Morris, C. Serre, Metalorganic frameworks in biomedicine, Chem. Rev. 112
(2012) 12321238.
[48] L.D. ONeill, H. Zhang, D. Bradshaw, Macro-/microporous MOF composite
beads, J. Mater. Chem. 20 (2010) 57205727.
[49] R. Ostermann, J. Cravillon, C. Weidmann, M. Wiebcke, B.M. Smarsly, Metal
organic framework nanobers via electrospinning, Chem. Commun. 47 (2011)
442444.
[50] M. Jahan, Q. Bao, J.-X. Yang, K.P. Loh, Structure-directing role of graphene in the
synthesis of metalorganic framework nanowire, J. Am. Chem. Soc. 132 (2010)
1448714495.
[51] S.-T. Zheng, T. Wu, C. Chou, A. Fuhr, P. Feng, X. Bu, Development of composite
inorganic building blocks for MOFs, J. Am. Chem. Soc. 134 (2012) 45174520.

42

I. Ahmed, S.H. Jhung / Chemical Engineering Journal 251 (2014) 3542

[52] N.A. Khan, S.H. Jhung, Adsorptive removal of benzothiophene using porous
copper-benzenetricarboxylate loaded with phosphotungstic acid, Fuel Process.
Technol. 100 (2012) 4954.
[53] N.A. Khan, Z. Hasan, S.H. Jhung, Ionic liquids supported on metalorganic
frameworks: remarkable adsorbents for adsorptive desulfurization, Chem. Eur.
J. 20 (2014) 376380.
[54] G.W. Peterson, J.A. Rossin, J.B. DeCoste, K.L. Killops, M. Browe, E. Valdes, P.
Jones, Zirconium hydroxidemetalorganic framework composites for toxic
chemical removal, Ind. Eng. Chem. Res. 52 (2013) 54625469.
[55] S.-H. Huo, X.-P. Yan, Facile magnetization of metalorganic framework MIL101 for magnetic solid-phase extraction of polycyclic aromatic hydrocarbons
in environmental water samples, Analyst 137 (2012) 34453451.
[56] C.-Y. Sun, S.-X. Liu, D.-D. Liang, K.-Z. Shao, Y.-H. Ren, Z.-M. Su, Highly stable
crystalline catalysts based on a microporous metalorganic framework and
polyoxometalates, J. Am. Chem. Soc. 131 (2009) 18831888.
[57] Y. Wang, R.T. Yang, Desulfurization of liquid fuels by adsorption on carbonbased sorbents and ultrasound-assisted sorbent regeneration, Langmuir 23
(2007) 38253831.
[58] X.-L. Song, L.-B. Sun, G.-S. He, X.-Q. Liu, Isolated Cu(I) sites supported on bcyclodextrin: an efcient p-complexation adsorbent for thiophene capture,
Chem. Commun. 47 (2011) 650652.
[59] R.T. Yang, A.J. Hernandez-Maldonaldo, F.H. Yang, Desulfurization of
transportation fuels with zeolites under ambient conditions, Science 301
(2003) 7981.
[60] Z.Y. Zhang, T.B. Shi, C.Z. Jia, W.J. Ji, Y. Chen, M.Y. He, Adsorptive removal of
aromatic organosulfur compounds over the modied Na-Y zeolites, Appl.
Catal. B: Environ. 82 (2008) 110.
[61] W.-H. Tian, L.-B. Sun, X.-L. Song, X.-Q. Liu, Y. Yin, G.-S. He, Adsorptive
desulfurization by copper species within conned space, Langmuir 26 (2010)
1739817404.
[62] W.-J. Jiang, Y. Yin, X.-Q. Liu, X.-Q. Yin, Y.-Q. Shi, L.-B. Sun, Fabrication of
supported cuprous sites at low temperatures: an efcient, controllable strategy
using vapor-induced reduction, J. Am. Chem. Soc. 135 (2013) 81378140.

[63] A.J. Hernndez-Maldonado, G. Qi, R.T. Yang, Photocatalytic degradation of


organic pollutants on surface anionized TiO2: common effect of anions for high
hole-availability by water, Appl. Catal. B: Environ. 61 (2005) 212218.
[64] I.J. Kang, N.A. Khan, E. Haque, S.H. Jhung, Chemical and thermal stability of
isotypic metalorganic frameworks: effect of metal ions, Chem. Eur. J. 17
(2011) 64376442.
[65] N.A. Khan, Z. Hasan, S.H. Jhung, Adsorption and removal of sulfur or nitrogencontaining compounds with metalorganic frameworks (MOFs), Adv. Porous
Mater. 1 (2013) 91102.
[66] G. Ferey, C. Serre, C. Mellot-Draznieks, F. Millange, S. Surble, J. Dutour, I.
Margiolaki, A hybrid solid with giant pores prepared by a combination of
targeted chemistry simulation and powder diffraction, Angew. Chem. Int. Ed.
(43) (2004) 62966301.
[67] S.H. Jhung, J.-H. Lee, J.-S. Chang, Microwave synthesis of a nanoporous
hybrid material, chromium trimesate, Bull. Korean Chem. Soc. 26 (2005)
880881.
[68] B.H. Hameed, A.A. Rahman, Removal of phenol from aqueous solutions by
adsorption onto activated carbon prepared from biomass material, J. Hazard.
Mater. 160 (2008) 576581.
[69] S.-H. Lin, R.-S. Juang, Adsorption of phenol and its derivatives from water using
synthetic resins and low-cost natural adsorbents: a review, J. Environ. Manage.
90 (2009) 13361349.
[70] Handbook of X-ray photoelectron spectroscopy, in: C.D. Wanger, W. M. Riggs,
L.E. Davis, J.F. Moulder, G.E. Muilenberg (Eds.), PerkinElmer Corporation,
Minnesota, 1978.
[71] E. Vilarrasa-Garca, A. Infantes-Molina, R. Moreno-Tost, E. Rodrguez-Castelln,
A. Jimnez-Lpez, C.L. Cavalcante Jr., D.C.S. Azevedo, Thiophene adsorption on
microporous activated carbons impregnated with PdCl2, Energy Fuels 24
(2010) 34363442.
[72] G.-X. Yu, J.-B. Li, X.-L. Zhou, C.-L. Li, L.-F. Chen, J.-A. Wang, Adsorption of
dibenzothiophene on transition metals loaded activated carbon, Adv. Mater.
Res. 132 (2010) 141148.

Anda mungkin juga menyukai