You are on page 1of 29

Accepted Manuscript

Ultrafast Room Temperature Synthesis of GrO@HKUST-1 Composites with


High CO2 Adsorption Capacity and CO2/N2 Adsorption Selectivity
Feng Xu, Ying Yu, Jian Yan, Qibin Xia, Haihui Wang, Jing Li, Zhong Li
PII:
DOI:
Reference:

S1385-8947(16)30800-2
http://dx.doi.org/10.1016/j.cej.2016.05.143
CEJ 15303

To appear in:

Chemical Engineering Journal

Received Date:
Revised Date:
Accepted Date:

20 April 2016
28 May 2016
30 May 2016

Please cite this article as: F. Xu, Y. Yu, J. Yan, Q. Xia, H. Wang, J. Li, Z. Li, Ultrafast Room Temperature Synthesis
of GrO@HKUST-1 Composites with High CO2 Adsorption Capacity and CO2/N2 Adsorption Selectivity, Chemical
Engineering Journal (2016), doi: http://dx.doi.org/10.1016/j.cej.2016.05.143

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Ultrafast Room Temperature Synthesis of GrO@HKUST-1


Composites with High CO2 Adsorption Capacity and CO2/N2
Adsorption Selectivity

Feng Xua,b, Ying Yua, Jian Yana, Qibin Xiaa, Haihui Wanga, Jing Lic,*, Zhong Lia,*
a

School of Chemistry and Chemical Engineering, South China University of Technology,


Guangzhou 510640, PR China
b

School of Environment and Chemical Engineering, Foshan University,


Foshan 528000, PR China

Department of Chemistry and Chemical Biology, Rutgers University, Piscataway,


New Jersey, 08854, USA

ABSTRACT
An ultrafast synthesis method is developed for the preparation of composites of
graphite oxide and HKUST-1, GrO@HKUST-1. Fast synthesis of GrO@HKUST-1
composites can be quickly achieved at room temperature within 1 min. The
synthesized composites were characterized by XRD, SEM, N2 adsorption, FTIR and
TGA. The isotherms of CO2 and N2 on the as-synthesized materials were measured
and the isosteric heats of CO2 adsorption were estimated. The CO2/N2 adsorption
selectivities were predicted by means of ideal adsorbed solution theory (IAST).
Results show that the GrO@HKUST-1 composites have higher BET surface area and
pore volume than the parent HKUST-1. The CO2 adsorption capacity of
2GrO@HKUST-1 (2% of GrO) is up to 9.02 mmol/g at 1 bar and 273 K, giving an
increase of 32% in comparison of the parent HKUST-1. The isosteric heat of CO2
1

adsorption on 2GrO@HKUST-1 is slightly higher than that of HKUST-1, suggesting


the stronger interaction between CO2 molecules and 2GrO@HKUST-1. The CO2/N2
adsorption selectivity of 2GrO@HKUST-1 is significantly enhanced over pristine
HKUST-1. At 1 bar, the CO2/N2 adsorption selectivity of 2GrO@HKUST-1 is up to
186, while that of parent HKUST-1 is only 103. This rapid room temperature
synthesis route is promising for new MOF-based composites.
Keywords: Fast synthesis, Graphite oxide, HKUST-1, GrO@HKUST-1 composite,
CO2/N2 separation
1. Introduction
With a rapidly depletion of fossil fuels since the industrial revolution, extensive
amount of CO2 has been released into the environment, causing destructive
consequences such as dramatic climate change and melting of polar glaciers [1].
Consequently, developing suitable and effective techniques towards abating the
emission of the CO2 into the atmosphere has becoming an urgent subject and topic.
Currently, membrane separation [2], chemical absorption [3] and adsorption [4, 5]
have been proposed for CO2 capture. Among them, physisorption-based processes has
been regarded as one of the most low-energy and cost-effective technologies [6]. In
the adsorption process, selection of the efficient adsorbent is crucial.
In the recent years, metal organic frameworks (MOFs) assembled with metal
ions and organic linkers are attracting a great deal of attention [7]. MOFs are being
considered vastly for their applications in a great variety of fields, ranging from gas
adsorption and separation [8-16] to catalysis [17], luminescence [18-20], electronics
2

[21], magnetism [22], drug delivery [23], sensing [24-26] and health science [27]. Up
to now, MOFs have been developed as promising materials for CO2 adsorption owing
to ultrahigh surface area and porosity, diverse tunable pores and metal sites. Lin et al.
[28] investigated the adsorption of CO2 on MIL-101, and reported that the adsorption
capacity of CO2 was 1.6 mmol/g at 298 K and 1 bar. Yan et al. [29] measured the
isotherm of CO2 on Cu-BTC, and found that the CO2 adsorption capacity on this
MOFs was up to 6.49 mmol/g at 273 K and 1 atm. Caskey et al. [30] reported that the
CO2 adsorption capacity of Mg-MOF-74 was up to 10.3 mmol/g at 298 K and 1 atm.
Moreover, it was reported that the incorporation of graphite oxide into MOFs can
enhance the adsorption performance of the resulting composite. Bandosz et al. [31]
reported two types of MOF/graphite oxide (GO) hybrid materials, MOF-5/GO and
HKUST-1/GO. It was found that these composites exhibited similar features in
crystalline structure and porosity to the parent MOFs, and their ammonia adsorption
capacities became higher compared to their parent MOFs due to creation of new pores
between the two phases MOF units and GO. A few novel GrO@MOFs composites
with higher adsorption capacities than their parental MOFs were recently synthesized.
Zhou et al. [32] reported a novel GrO@MIL-101 composite with a CO2 adsorption
capacity of 3.6 mmol/g (298 K, 1 bar), significantly higher than that of MIL-101 (1.6
mmol/g). Ram et al. [33] demonstrated a simple procedure to prepare hybrid
GO@ZIF-8 nanocomposites. Their results showed that the composite (ZG-20)
exhibited remarkable CO2 storage capacity (72 wt%) compared to the parent ZIF-8
(27.2 wt%). Liu et al. [34] prepared the composites based on Cu-BTC and grapheme
3

layers and reported that the materials obtained exhibited about a 30% increase in CO2
storage capacity (from 6.39 mmol g1 of Cu-BTC to 8.26 mmol g1 of CG-9 at 273 K
and 1 atm).
Nevertheless, most of the MOF composites were prepared by traditional
solvothermal methods, which need high temperature and long reaction time. Rapid
room temperature synthesis of MOF composites is highly desired for industrial
implementation and commercialization. Li et al. [35] proposed a solvent-free
mechanochemical method to rapidly synthesize composites (Cu-BTC@)GO) of
Cu-BTC and graphite oxide within 30 min, and reported a toluene adsorption capacity
of Cu-BTC@GOs up to 9.1 mmol/g at 298 K, having an increase of 47% in
comparison with Cu-BTC. While mechanochemical methods can tackle the challenge
and reduce the reaction time (e.g. 30 min) and synthesis temperature (e.g. 25 oC), but
vast external energy input was required to capacitate the chemical reactions.
Therefore, it is desirable to seek for a more gentle and efficient method for the
industrial implementation of MOF composites.
Recently, some metal oxides and hydroxides have been reported to act as
nucleating agents or sources of cations for rapid synthesis of MOFs [36, 37]. Zhao et
al. [38] reported an ultrafast room temperature synthesis of MOFs using (Zn, Cu)
hydroxyl double salt intermediates. They found that a (Zn, Cu) hydroxy double salt
(HDS) intermediate formed in situ from ZnO particles enabled rapid growth (<1 min)
of HKUST-1(Cu 3(BTC)2) at room temperature. Synthesis of HKUST-1 was
4

completed within 1 min at room temperature, showing feasibility of fast room


temperature synthesis for bulk MOF powders and significant promise for scale-up
processing. However, to the best of our knowledge there is no report about using
ultrafast synthesis to prepare composites of MOF and graphite oxide, in short,
GrO@MOFs.
Herein, we reported an ultrafast synthesis of GrO@HKUST-1 at room
temperature for the first time, and investigated the CO2 and N2 adsorption of the
resulting composites. The synthesized samples were characterized by XRD, SEM, N2
adsorption, FTIR and TGA. The CO2 adsorption isotherms of the composites at
different temperature were measured by a volumetric method. The isosteric heat of
CO2 adsorption on the composites was calculated based on the single component of
CO2. The CO2/N2 selectivities of the samples were predicted by using ideal adsorbed
solution theory (IAST). Comparison between GrO@HKUST-1 and the parent
HKUST-1 for adsorption of CO2/N2 was made herein.

2. Experimental section
2.1. Materials
All chemicals were obtained from commercial sources and used as received
without further purification. Starting materials include copper nitrate hydrate
(Cu(NO3)23H2O, Alfa, 99%), 1,3,5-benzenetricarboxylic acid (C9H6O6, Alfa, 99%),
zinc oxide (ZnO, Alfa, 99%), N,N-dimethylformamide (C3H7NO, Guanghua, 99.8%),
ethyl alcohol absolute (C2H5OH, Fuyu, 99.7%), graphite powder (C, Aladdin, size30
30
m),
m), sodium
sodium nitrate
nitrate (NaNO
(NaNO3, Guanghua, 99%), concentrated sulfuric acid (H2SO4,
5

Guanghua, 98%), potassium permanganate (KMnO4, Kaixin, 99.5%) and hydrogen


peroxide (H2O2, Guanghua, 30.0
30.0 %).
%).
2.2. Solvothermal synthesis of HKUST-1
HKUST-1 was solvothermally synthesized following the modified procedure
reported previously [39]. A solution of copper nitrate hydrate (0.725 g) in deionized
water (12 mL) was added to a solution of 1,3,5-benzenetricarboxylic acid (0.42 g) in
deionized water (12 mL) and ethyl alcohol absolute (24 mL). The mixture was stirred
for 10 min and transferred to a 100 mL Teflon-lined autoclave. The oven was heated
at 110 oC for 24 h. The resultant blue crystals were filtered, washed with ethanol and
chloroform.
2.3. Synthesis of graphite oxide (GrO)
Graphite oxide was synthesized following the modified Hummers procedure
reported previously [40]. Graphite powder (4 g) and sodium nitrate (4 g) were added
to concentrated sulfuric acid (220 mL) and stirred in an ice bath. Then potassium
permanganate (24 g) was slowly added to the mixture (below 283 K). The suspension
was stirred for 30 min under ice bath, then stirred under room temperature for 48 h.
Followed, 368 mL of deionized water was added to the beaker. The mixture was
stirred for extra 15 min, and further diluted with 60 oC warm water (1120 mL).
Subsequently, hydrogen peroxide (100 mL) was added and the suspension was aged
for 12 h. The mixture was separated by centrifugation and washed with water. The
remaining particles were transferred to a dialysis tubes. The deionized water should be

refreshed frequently until no BaSO4 precipitated by adding BaCl2. Finally, the


suspension was spray-dried and fine brown graphite oxide were obtained.
2.4. Ultrafast synthesis of GrO@HKUST-1
The ultrafast synthesis of GrO@HKUST-1 composites was performed following
the reported procedures [38] with a few modifications. First, ZnO powder (0.293 g)
was dispersed in deionized water (8 mL) and sonicated for 10 min to form nanoslurry.
Second, a certain amount of GrO was dispersed in deionized water (8 mL). The
resultant solution was sonicated for 20 min. Then, 1.74 g of Cu(NO3)23H2O) was
added in the above solution. At the same time, H3BTC (0.84 g) was dissolved in
ethanol (16 mL). Third, ZnO nanoslurries were added with 16 mL of DMF under
stirring, then GrO/Cu(NO3)2 solution was added and next the H3BTC solution. In one
minute, amount of the particles appeared. Then the product was filtered and washed
with ethanol three times (50 mL). Finally, the product was dried at 120 oC for 6 hours
in vacuum oven to obtain the fine powder of GrO@HKUST-1. The GrO amounts
added in the preparation of the composites were 1, 2, and 3 wt% of the initial weight
of Cu(NO3)23H2O and H3BTC, and thus the resulting composites were labeled
correspondingly as 1GrO@HKUST-1, 2GrO@HKUST-1 and 3GrO@HKUST-1.

2.5. Characterization
Powder X-ray diffraction patterns were recorded on a D8 ADVANCE
diffractometer (Bruker, Germany) using Cu-K radiation (= 1.5406 ). The scan
speed was 2 deg/min and operating power was 40 kV and 40 mA. The morphologies
of the samples were characterized by scanning electron microscope (Philips FEI
7

XL-30). The FTIR spectra were carried out on a Bruker Vector 33 spectrometer.
Thermogravimetric analyses were performed on a Beijing Henven Scientific
Instrument thermal analyzer heating from room temperature to 600 oC at a rate of 10
o

C /min under nitrogen atmosphere (30 mL/min). The elemental analysis of Cu was

carried out on the ICP-AES (Thermo Jarrell Ash Co.) to estimate the GrO content of
the GrO@HKUST-1. Nitrogen adsorption and desorption isotherms at 77 K were
measured to investigate the pore textural properties by using Micromeritics ASAP
2020. The samples were outgassed at 423 K for 8 hours under vacuum.
2.6. CO2 and N2 adsorption measurements
The CO2 and N2 sorption isotherms at different temperature (273 K, 288 K, 298
K, 308 K) were performed on a high-resolution Micromeritics 3Flex adsorption
instrument. The gases (CO2, N2) used in the experiment were of ultrahigh purity
(99.999%). The initial outgassing process for the samples were carried out at 423 K
for 8 h under vacuum.
3. Results and Discussion
3.1. Physical Characteristics
Figure 1 depicts the powder X-ray diffraction pattern of the samples and
simulated HKUST-1. The main diffraction patterns of the samples match well with
the simulated XRD pattern [41], implying that GrO@HKUST-1 composites preserve
the crystalline characters of parental HKUST-1. It is noticed that 2GrO@HKUST-1
shows the highest peak intensities among the composites, indicating the higher
crystallinity.
8

Figure 1. PXRD patterns of HKUST-1 and GrO@HKUST-1 composites.


Figure 2 displays the SEM images of the samples. The HKUST-1 crystals
synthesized by solvothermal method have a typical octahedral morphology with a size
range of 1.1-3 m, as shown in Figure 2 (a). Figure 2(b)-(d) show that the small
octahedral HKUST-1 particles are well embedded into the GrO layers, and these
HKUST-1 particle sizes in GrO@HKUST-1 composites are about 0.80.2 m,
which is more uniform and smaller than that of HKUST-1. The reason might be
ascribed to the existence of additional constraints in the degrees of freedom for the
growth of HKUST-1 crystals during the synthesis of the composites caused by the
presence of GrO [42]. In other words, during the growth of HKUST-1 crystals in the
synthesis, the GrO in the composite may exert distortion force on the crystals,
resulting in diminishment of the crystals.

Figure 2. SEM images of (a) HKUST-1, (b) 1GrO@HKUST-1, (c) 2GrO@HKUST-1,


and (d) 3GrO@HKUST-1.
In order to determine the surface area and porosity of the samples, N2
adsorption-desorption experiments at 77 K were carried out. Figure 3 presents the N2
isotherms on HKUST-1 and GrO@HKUST-1 composites at 77 K. The N2 isotherms
of all the samples exhibit a typical type-I sorption behavior with an intense increase at
low nitrogen relative pressure followed by a plateau, indicating that the samples have
plentiful micro-pores. Higher adsorption capacity is clearly visible from the N2
isotherms of the GrO@HKUST-1 composites in comparison with that of HKUST-1,
demonstrating that GrO@HKUST-1 composites have higher specific surface area
compared to HKUST-1. Furthermore, small H4 type hysteresis loops are observed on
the GrO@HKUST-1 composites, indicating the existence of narrow slit-like pores
which might be generated between HKUST-1 particles and GrO layers [43]. This
10

phenomenon is similar with the case of MOF-GO composites as reported elsewhere


[39].
Table 1 lists the pore structure parameters of the samples. These data show that
the BET surface areas and pore volumes of the composites are higher than those of the
parent HKUST-1, which can be ascribed to the possibility that GrO provides new
crystallization sites for HKUST-1 crystallites and thus results in the enhanced porosity
[32]. In other words, new porosity is formed on the interface between HKUST-1
crystals and graphene layers as indicated in the SEM graphs, despite the fact that GrO
shows no porosity. It is noticed that 2GrO@HKUST-1 exhibits the optimal textural
properties with a highest BET and Langmuir surface area of 1554 and 1968 m2/g, and
a highest total pore volume of 0.711 cm3/g among the samples,

which are

significantly higher than the reported Cu-BTC-graphite oxide composites [44].

Figure 3. Nitrogen adsorptiondesorption isotherms at 77 K on the HKUST-1 and


GrO@HKUST-1 composites.
Table 1. Pore structure parameters of the HKUST-1 and GrO@HKUST-1 composites

11

Surface area (m2/g)

Pore volume (cm3/g)

Materials
SBET

SLangmuir

Vtotal

Vmeso

Vmicro

HKUST-1

1193

1576

0.57

0.06

0.51

1GrO@HKUST-1

1489

1956

0.695

0.096

0.599

2GrO@HKUST-1

1554

1968

0.711

0.094

0.617

3GrO@HKUST-1

1493

1887

0.703

0.118

0.585

Vtotal: total pore volume at P/P0=0.995; Vmicro: micro pore volume calculated by t-plot
analysis and Vmeso: meso pore volume calculated by Vtotal- Vmicro.
Figure 4 shows the TG profiles of the samples. The main turning points of the
TG profiles for the composites are very similar with those for HKUST-1. It shows that
there are three weight losses for the samples. The first weight loss below 423 K is
corresponding to the remove of guest molecules. The second one happens between
423 K and 523 K, corresponding to the remove of bound water. The third one is
corresponding to the decomposition of the structures started at 593 K and ended at
673 K. Hence, the combination of GrO and HKUST-1 basically has no impact on the
thermal stability of the composites.

12

Figure 4. TG curves of HKUST-1 and GrO@HKUST-1 composites.


The FTIR spectra of the HKUST-1 and GrO@HKUST-1 samples are plotted in
Figure 5. They clearly show high similarities, with the main characteristic peak
matching well with the published FTIR spectrum of HKUST-1 [45]. Briefly, the
bands between 700 and 1700 cm1 are assigned to BTC, and the characteristic peak
around 500 cm1 are assigned to Cu-O stretching vibrations [46]. The FTIR results
indicate that the chemical environment of the GrO@HKUST-1 surfaces is similar to
that the HKUST-1.

Figure 5. FTIR spectra of the HKUST-1 and GrO@HKUST-1 composites.


13

The results above confirm that the GrO@HKUST-1 composites have been
successfully synthesized using rapid room temperature synthesis method. This can be
attributed to the crucial role of adding ZnO resulting in production of (Zn, Cu)
hydroxy double salt (HDS) during the synthesis. HDSs are layered compounds
containing cationic sheets linked by inorganic/organic interlamellar anions [47], and
possess excellent anion exchangeability and high anion exchange rate [38]. Generally
speaking, ZnO, CuO and NiO can react with different divalent cations such as Zn2+,
Cu2+, Ni2+, and Co2+ salts to form HDS [38]. In our work, (Zn, Cu) hydroxy nitrate
HDS were produced by reacting ZnO with Cu 2+ in Cu(NO3)23H2O. The resulting
(Zn, Cu) hydroxy nitrate HDS with excellent anion exchangeability and high anion
exchange rate significantly drives the ultrafast formation of GrO@HKUST-1.
3.2. Adsorption isotherms of CO2
Figure 6 presents the CO2 adsorption isotherms at 273 K of HKUST-1 and
GrO@HKUST-1 composites. It is visible that all composites exhibit significantly
higher

CO2

adsorption

capacities

than

HKUST-1.

Among

the

samples,

2GrO@HKUST-1 has the highest CO2 adsorption capacity (9.02 mmol/g) at 1 bar,
around 1.32 times of that of HKUST-1. The remarkable enhancement in CO2
adsorption capacities of GrO@HKUST-1 composites can be attributed to the higher
BET surface area and pore volume than those of HKUST-1, as shown in Table 1. A
similar phenomenon was reported in the case of the GrO@Cu-BTC [39] and
GrO@MIL-101 [32] synthesized by solvothermal methods for CO2/CH4 adsorption.

14

Figure 6. Isotherms of CO2 on HKUST-1 and GrO@HKUST-1 composites at 273 K.


For comparison, Table 2 summarizes CO2 adsorption capacities at 1 bar of
some adsorbents. It indicates that the CO2 adsorption capacity of 2GrO@HKUST-1 is
higher than most of the adsorbents listed except Mg-MOF-74. One should note that
both 2GrO@HKUST-1 and GrO@Cu-BTC [39] are composed of GrO and
HKUST-1, but the synthesis methods are different. In our work, 2GrO@HKUST-1
was rapidly prepared at room temperature using a ultrafast synthesis method, which
can be completed in one minute, while in Huangs work [39], higher temperature (383
K) and longer reaction time (24 h) were required to prepare the composite
GrO@Cu-BTC of GrO and Cu-BTC.
Table 2. Adsorption capacities at 1bar of CO2 on different adsorbents
Adsorbent

Adsorption capacity

Temperature

(mmol/g)

(K)

HKUST-1

6.85

273

This work

2GrO@HKUST-1

9.02

273

This work

15

Reference

GrO@Cu-BTC

8.26

273

[39]

Cu-BTC

6.49

273

[29]

MOF-5

2.1

296

[48]

Mg-MOF-74

10.3

278

[30]

MIL-101

1.6

298

[28]

MIL-53

3.6

273

[49]

Pelletized zeolite 13X

4.66

293

[50]

Zeolite LiX

4.64

293

[51]

AC

2.76

293

[51]

3.3. Isosteric heat of adsorption (H)


To evaluate the affinity between gas molecules and the framework, we applied
single component isotherms collected from three temperatures (Figure S1) to calculate
the isosteric heat H of CO2 adsorption on HKUST-1 and 2GrO@HKUST-1. H can
be estimated by using Clausius-Clapeyron equation [52, 53].

ln( p) =

H
+C
RT

(1)

Where p (bar) is pressure, H (J/mol) is isosteric heat of adsorption at a specific


amount adsorbed of CO2, T (K) is temperature, R is gas constant and C is an
integration constant.
H can be estimated based on the obtained CO2 isotherms at different

temperatures (See Figure S1 in Supporting Information). Firstly, these CO2 isotherms


were converted to CO2 adsorption isosteres, and the lnp was plotted to 1/T at a given
amount adsorbed of CO2 on the basis of equation (1). As a result, a fitting straight line
16

with a slope of - H /R was yielded, as shown in Figure S2 in Supporting Information,


and thus, the H of CO2 adsorption on the samples HKUST-1 and 2GrO@HKUST-1
can be calculated from the slopes - H /R directly.
Figure 7 shows the dependence of H on the amounts adsorbed of CO2 on
HKUST-1 and 2GrO@HKUST-1. It is observed that the isosteric heat of CO2
adsorption decreased slightly with amounts adsorbed of CO2. It is noteworthy that the
H of CO2 adsorption on 2GrO@HKUST-1 is somewhat higher than that on

HKUST-1, indicating the stronger interaction between CO2 and 2GrO@HKUST-1


surfaces, which could be mainly ascribed to the existence of graphite oxide resulting
in enhanced electrostatic interactions and dispersive force toward CO2.

Figure 7. The isosteric heats of CO2 adsorption on HKUST-1 and 2GrO@HKUST-1.


3.4. CO2/N2 adsorption selectivity predicted by ideal adsorbed solution theory (IAST)
The ideal adsorbed solution theory (IAST) developed by Myers and Praunitz [54]
has widely been applied to predict the adsorption selectivity and the adsorption
equilibrium of gas mixtures from the isotherms of the pure components. In this work,
IAST was employed to predict CO2/N2 binary mixture adsorption. In order to describe
17

CO2 and N2 adsorption on the samples clearly, dual-site Langmuir-Freundlich (DSLF)


equation [55] was used to fit a single-component isotherm, which can be expressed as:
q = q1max

b1 p1/ n1
b2 p1/ n2
max
+
q

2
1 + b1 p1/ n1
1 + b2 p1/ n2

(2)

Where q (mmol/g) is the adsorbed amount per mass of adsorbent, p (kpa) is the
pressure of the bulk gas at equilibrium with the adsorbed phase, q1max and q2max
(mmol/g) are the saturated adsorption capacities of sites 1 and 2, b1 and b2 (1/kPa) are
the affinity coefficients of sites 1 and 2, and n1 and n2 are the deviations from an ideal
homogeneous surface.
Figure 8 presents a comparison of DSLF model fits and experimental CO2 and N2
isotherm data on the two samples HKUST-1 and 2GrO@HKUST-1 at 273 K. Table 3
lists the fitting parameters of the DSLF model as well as the regression coefficients.
The high regression coefficient R2 up to 0.9999 declares that the DSLF model can fit
the isotherms extremely well.

Figure 8. Comparison of the DSLF model fits and the experimental CO2 and N2
isotherm data on HKUST-1 (a) and 2GrO@HKUST-1 (b) at 273 K. (Symbols are
experimental data points; solid lines are prediction by DSLF model).

18

Table 3. Equation parameters for the DSLF isotherm model


HKUST-1

2GrO@HKUST-1

CO2

N2

CO2

N2

q1max (mmol/g)

14.0048

0.28661

16.94965

0.00154

b1 (kPa-1)

0.00794

0.00569

0.00564

6.38456E-14

n1

0.96912

0.93488

0.87871

0.09628

q2max (mmol/g)

0.15786

0.30714

3.25

b2 (kPa-1)

3.011709E-7

1.6568E-5

0.22374

5.99405E-4

n2

0.23068

0.59391

1.0208

0.97379

R2

0.99999

0.9999

0.99999

0.99999

Figure 9 presents the IAST-predicted CO2/N2 selectivities for equimolar CO2 and
N2 binary mixture at 273 K as functions of the total pressure. In general, the CO2/N2
adsorption selectivities of the two samples increase as a function of total pressure,
which is likely due to the stronger interaction between the samples and CO2 compared
to N2 in high bulk pressure. More importantly, the CO2/N2 selectivity of
2GrO@HKUST-1 is significantly higher than that of HKUST-1. For example, the
CO2/N2 adsorption selectivity of 2GrO@HKUST-1 is up to 186 at 101 kPa, while that
of HKUST-1 is only 103. The former is almost 1.8 times of that of HKUST-1. In
conclusion, 2GrO@HKUST-1 has not only a high adsorption capacity for CO2, but
also a high CO2/N2 adsorption selectivity, which makes it a promising and efficient
candidate adsorbent for the separation of CO2/N2 mixtures.

19

Figure 9. IAST predicted selectivities for equimolar CO2/N2 mixtures on the


HKUST-1 and 2GrO@HKUST-1 at 273 K.
4. Conclusions

In summary, we developed an ultrafast synthesis method for the preparation of


GrO@HKUST-1 composites at room temperature by using (Zn, Cu) hydroxyl double
salt intermediate. The reactions were completed within 1 min. The resulting
GrO@HKUST-1 composites exhibit higher specific surface area and pore volume
than HKUST-1, and the crystal sizes of HKUST-1 in the composites are smaller than
that of HKUST-1. The adsorption capacity of 2GrO@HKUST-1 for CO2 at 1 bar and
273 K is up to 9.02 mmol/g, which is significantly higher than that of HKUST-1 (6.85
mmol/g), with an increase of 32%. The isosteric heat of CO2 adsorption on the
2GrO@HKUST-1 is higher than that on the HKUST-1, suggesting the interaction
between CO2 molecules and the 2GrO@HKUST-1 is stronger than that of HKUST-1.
The selectivity on the 2GrO@HKUST-1 for the equimolar CO2/N2 is significantly
higher over pristine HKUST-1. At 1 bar, the CO2/N2 adsorption selectivity of
2GrO@HKUST-1 is up to 186, which is almost 1.8 times that of HKUST-1 (103).
20

This ultrafast room temperature strategy is promising for the synthesis of new
MOF-based composites, which may become suitable for the sample production at
industrial scale.
Acknowledgements
This work was supported by National Key Basic Research Program of China (2013CB733506),
Key Program of National Natural Science Foundation of China (No. 21436005), National Natural
Science Foundation of China (No. 21276092), the Guangdong Natural Science Foundation (No.
2014A030312007), the Research Foundation of State Key Lab of Subtropical Building Science of
China (C715023z), and the Fundamental Research Funds for the Central Universities.

Reference
[1] R. Monastersky, A burden beyond bearing, Nature 458 (2009) 1091-1094.
[2] J. Fang, J. Tong, K. Huang, A superior mixed electron and carbonate-ion conducting
metal-carbonate composite membrane for advanced flue-gas carbon capture, J. Membrane Sci. 505
(2016) 225-230.
[3] B. Lv, G. Jing, Y. Qian, Z. Zhou, An efficient absorbent of amine-based amino acid-functionalized
ionic liquids for CO2 capture: High capacity and regeneration ability, Chem. Eng. J. 289 (2016)
212-218.
[4] H. Huang, W. Zhang, F. Yang, B. Wang, Q. Yang, Y. Xie, C. Zhong, J.-R. Li, Enhancing CO2
adsorption and separation ability of Zr (IV)-based metalorganic frameworks through ligand
functionalization under the guidance of the quantitative structureproperty relationship model, Chem.
Eng. J. 289 (2016) 247-253.
[5] F. Martnez, R. Sanz, G. Orcajo, D. Briones, V. Yngez, Amino-impregnated MOF materials for
CO2 capture at post-combustion conditions, Chem. Eng. Sci. 142 (2016) 55-61.
[6] N. Konduru, P. Lindner, N.M. AssafAnid, Curbing the greenhouse effect by carbon dioxide
adsorption with zeolite 13X, AICHE J. 53 (2007) 3137-3143.
[7] M. Eddaoudi, J. Kim, N. Rosi, D. Vodak, J. Wachter, M. O'Keeffe, O.M. Yaghi, Systematic design
of pore size and functionality in isoreticular MOFs and their application in methane storage, Science
295 (2002) 469-472.
[8] J.J. Peng, S.K. Xian, J. Xiao, Y. Huang, Q.B. Xia, H.H. Wang, Z. Li, A supported
Cu(I)@MIL-100(Fe) adsorbent with high CO adsorption capacity and CO/N2 selectivity, Chem. Eng. J.
270 (2015) 282-289.
[9] S. Xian, J. Peng, Z. Zhang, Q. Xia, H. Wang, Z. Li, Highly enhanced and weakened adsorption

21

properties of two MOFs by water vapor for separation of CO2/CH4 and CO2/N2 binary mixtures, Chem.
Eng. J. 270 (2015) 385-392.
[10] H. Wang, K. Yao, Z. Zhang, J. Jagiello, Q. Gong, Y. Han, J. Li, The first example of commensurate
adsorption of atomic gas in a MOF and effective separation of xenon from other noble gases, Chem.
Sci. 5 (2014) 620-624.
[11] D. Bahamon, L.F. Vega, Systematic evaluation of materials for post-combustion CO2 capture in a
Temperature Swing Adsorption process, Chem. Eng. J. 284 (2016) 438-447.
[12] A.K. Adhikari, K.-S. Lin, Improving CO2 adsorption capacities and CO2/N2 separation efficiencies
of MOF-74(Ni, Co) by doping palladium-containing activated carbon, Chem. Eng. J. 284 (2016)
1348-1360.
[13] M.-B. Kim, T.-U. Yoon, D.-Y. Hong, S.-Y. Kim, S.-J. Lee, S.-I. Kim, S.-K. Lee, J.-S. Chang, Y.-S.
Bae, High SF6/N2 selectivity in a hydrothermally stable zirconium-based metalorganic framework,
Chem. Eng. J. 276 (2015) 315-321.
[14] F. Xu, S. Xian, Q. Xia, Y. Li, Z. Li, Effect of textural properties on the adsorption and desorption
of toluene on the metal-organic frameworks HKUST-1 and MIL-101, Adsorpt. Sci. Technol. 31 (2013)
325-340.
[15] Z. Zhang, Y. Zhao, Q. Gong, Z. Li, J. Li, MOFs for CO2 capture and separation from flue gas
mixtures: the effect of multifunctional sites on their adsorption capacity and selectivity, Chem.
Commun. 49 (2013) 653-661.
[16] H. Wu, Q. Gong, D.H. Olson, J. Li, Commensurate adsorption of hydrocarbons and alcohols in
microporous metal organic frameworks, Chem. Rev. 112 (2012) 836-868.
[17] V.T. Nguyen, H.Q. Ngo, D.T. Le, T. Truong, N.T.S. Phan, Iron-catalyzed domino sequences:
One-pot oxidative synthesis of quinazolinones using metalorganic framework Fe3O(BPDC)3 as an
efficient heterogeneous catalyst, Chem. Eng. J. 284 (2016) 778-785.
[18] F. Xu, H. Wang, S.J. Teat, W. Liu, Q. Xia, Z. Li, J. Li, Synthesis, structure and enhanced
photoluminescence properties of two robust, water stable calcium and magnesium coordination
networks, Dalton T. 44 (2015) 20459-20463.
[19] Q. Gong, Z. Hu, B.J. Deibert, T.J. Emge, S.J. Teat, D. Banerjee, B. Mussman, N.D. Rudd, J. Li,
Solution processable MOF yellow phosphor with exceptionally high quantum efficiency, J. Am. Chem.
Soc. 136 (2014) 16724-16727.
[20] Z.C. Hu, G.X. Huang, W.P. Lustig, F.M. Wang, H. Wang, S.J. Teat, D. Banerjee, D.Q. Zhang, J. Li,
Achieving exceptionally high luminescence quantum efficiency by immobilizing an AIE molecular
chromophore into a metal-organic framework, Chem. Commun. 51 (2015) 3045-3048.
[21] H. Fei, X. Liu, Z. Li, Hollow cobalt coordination polymer microspheres: A promising anode
material for lithium-ion batteries with high performance, Chem. Eng. J. 281 (2015) 453-458.
[22] Y. Shao, L. Zhou, C. Basunsuno, J. Ma, M. Liu, F. Wang, Magnetic responsive metalorganic
frameworks nanosphere with coreshell structure for highly efficient removal of methylene blue, Chem.
Eng. J. 283 (2016) 1127-1136.
[23] L.-N. Duan, Q.-Q. Dang, C.-Y. Han, X.-M. Zhang, An interpenetrated bioactive nonlinear optical
MOF containing a coordinated quinolone-like drug and Zn (II) for pH-responsive release, Dalton T. 44
(2015) 1800-1804.
[24] Z. Hu, K. Tan, W.P. Lustig, H. Wang, Y. Zhao, C. Zheng, D. Banerjee, T.J. Emge, Y.J. Chabal, J. Li,
Effective sensing of RDX via instant and selective detection of ketone vapors, Chem. Sci. 5 (2014)
4873-4877.

22

[25] Z. Hu, B.J. Deibert, J. Li, Luminescent metal-organic frameworks for chemical sensing and
explosive detection, Chem. Soc. Rev. 43 (2014) 5815-5840.
[26] Z. Hu, W.P. Lustig, J. Zhang, C. Zheng, H. Wang, S.J. Teat, Q. Gong, N.D. Rudd, J. Li, Effective
detection of mycotoxins by a highly luminescent metal-organic framework, J. Am. Chem. Soc. (2015).
[27] J. Vaughn, H. Wu, B. Efremovska, D.H. Olson, J. Mattai, C. Ortiz, A. Puchalski, J. Li, L. Pan,
Encapsulated recyclable porous materials: an effective moisture-triggered fragrance release system,
Chem. Commun. 49 (2013) 5724-5726.
[28] Y. Lin, Q. Yan, C. Kong, L. Chen, Polyethyleneimine incorporated metal-organic frameworks
adsorbent for highly selective CO2 capture, Sci. Rep.-UK 3 (2013).
[29] X. Yan, S. Komarneni, Z. Zhang, Z. Yan, Extremely enhanced CO2 uptake by HKUST-1
metalorganic framework via a simple chemical treatment, Micropor. Mesopor. Mat. 183 (2014) 69-73.
[30] S.R. Caskey, A.G. Wong-Foy, A.J. Matzger, Dramatic tuning of carbon dioxide uptake via metal
substitution in a coordination polymer with cylindrical pores, J. Am. Chem. Soc. 130 (2008)
10870-10871.
[31] T.J. Bandosz, C. Petit, MOF/graphite oxide hybrid materials: exploring the new concept of
adsorbents and catalysts, Adsorption 17 (2011) 5-16.
[32] X. Zhou, W. Huang, J. Miao, Q. Xia, Z. Zhang, H. Wang, Z. Li, Enhanced separation performance
of a novel composite material GrO@MIL-101 for CO2/CH4 binary mixture, Chem. Eng. J. 266 (2015)
339-344.
[33] R. Kumar, K. Jayaramulu, T.K. Maji, C. Rao, Hybrid nanocomposites of ZIF-8 with graphene
oxide exhibiting tunable morphology, significant CO2 uptake and other novel properties, Chem.
Commun. 49 (2013) 4947-4949.
[34] S. Liu, L. sun, F. Xu, J. Zhang, C. Jiao, F. Li, Z. Li, S. Wang, Z. Wang, X. Jiang, Nanosized
Cu-MOFs induced by graphene oxide and enhanced gas storage capacity, Energ. Environ. Sci. 6 (2013)
818-823.
[35] Y. Li, J. Miao, X. Sun, J. Xiao, Y. Li, H. Wang, Q. Xia, Z. Li, Mechanochemical synthesis of
Cu-BTC@GO with enhanced water stability and toluene adsorption capacity, Chem. Eng. J. 298 (2016)
191-197.
[36] G. Majano, J. PrezRamrez, Scalable room-temperature conversion of copper (II) hydroxide
into HKUST-1 (Cu3(btc)2), Adv. Mater. 25 (2013) 1052-1057.
[37] E. Zanchetta, L. Malfatti, R. Ricco, M.J. Styles, F. Lisi, C.J. Coghlan, C.J. Doonan, A.J. Hill, G.
Brusatin, P. Falcaro, ZnO as an efficient nucleating agent for rapid, room temperature synthesis and
patterning of Zn-based metalorganic frameworks, Chem. Mater. 27 (2014) 690-699.
[38] J. Zhao, W.T. Nunn, P.C. Lemaire, Y. Lin, M.D. Dickey, C.J. Oldham, H.J. Walls, G.W. Peterson,
M.D. Losego, G.N. Parsons, Facile conversion of hydroxy double salts to metalorganic frameworks
using metal oxide particles and atomic layer deposition thin-film templates, J. Am. Chem. Soc. 137
(2015) 13756-13759.
[39] W. Huang, X. Zhou, Q. Xia, J. Peng, H. Wang, Z. Li, Preparation and adsorption performance of
GrO@Cu-BTC for separation of CO2/CH4, Ind. Eng. Chem. Res. 53 (2014) 11176-11184.
[40] W.S. Hummers Jr, R.E. Offeman, Preparation of graphitic oxide, J. Am. Chem. Soc. 80 (1958)
1339-1339.
[41] A.A. Yakovenko, J.H. Reibenspies, N. Bhuvanesh, H.-C. Zhou, Generation and applications of
structure envelopes for porous metalorganic frameworks, J. Appl. Crystallogr. 46 (2013) 346-353.
[42] X. Zhou, W. Huang, J. Shi, Z. Zhao, Q. Xia, Y. Li, H. Wang, Z. Li, A novel MOF/graphene oxide

23

composite GrO@MIL-101 with high adsorption capacity for acetone, J. Mater. Chem. A 2 (2014)
4722-4730.
[43] C. Petit, J. Burress, T.J. Bandosz, The synthesis and characterization of copper-based
metalorganic framework/graphite oxide composites, Carbon 49 (2011) 563-572.
[44] C. Petit, T.J. Bandosz, Exploring the coordination chemistry of MOF-graphite oxide composites
and their applications as adsorbents, Dalton T. 41 (2012) 4027-4035.
[45] S. Loera-Serna, M.A. Oliver-Tolentino, M. de Lourdes Lpez-Nez, A. Santana-Cruz, A.
Guzmn-Vargas, R. Cabrera-Sierra, H.I. Beltrn, J. Flores, Electrochemical behavior of [Cu3(BTC)2]
metalorganic framework: The effect of the method of synthesis, J. Alloy. Compd. 540 (2012) 113-120.
[46] E. Borfecchia, S. Maurelli, D. Gianolio, E. Groppo, M. Chiesa, F. Bonino, C. Lamberti, Insights
into adsorption of NH3 on HKUST-1 metalorganic framework: a multitechnique approach, J. Phys.
Chem. C 116 (2012) 19839-19850.
[47] M. Meyn, K. Beneke, G. Lagaly, Anion-exchange reactions of hydroxy double salts, Inorg. Chem.
32 (1993) 1209-1215.
[48] Z. Zhao, Z. Li, Y. Lin, Adsorption and diffusion of carbon dioxide on metal-organic framework
(MOF-5), Ind. Eng. Chem. Res. 48 (2009) 10015-10020.
[49] X. Si, J. Zhang, F. Li, C. Jiao, S. Wang, S. Liu, Z. Li, H. Zhou, L. Sun, F. Xu, Adjustable structure
transition and improved gases (H2, CO2) adsorption property of metal-organic framework MIL-53 by
encapsulation of BNHx, Dalton T. 41 (2012) 3119-3122.
[50] Y. Park, Y. Ju, D. Park, C.-H. Lee, Adsorption equilibria and kinetics of six pure gases on
pelletized zeolite 13X up to 1.0 MPa: CO2, CO, N2, CH4, Ar and H2, Chem. Eng. J. 292 (2016)
348-365.
[51] Y. Park, D.-K. Moon, Y.-H. Kim, H. Ahn, C.-H. Lee, Adsorption isotherms of CO2, CO, N2, CH4,
Ar and H2 on activated carbon and zeolite LiX up to 1.0 MPa, Adsorption 20 (2014) 631-647.
[52] D. Shen, M. Blow, F. Siperstein, M. Engelhard, A.L. Myers, Comparison of experimental
techniques for measuring isosteric heat of adsorption, Adsorption 6 (2000) 275-286.
[53] A. Anson, M.A. Callejas, A.M. Benito, W.K. Maser, M. Izquierdo, B. Rubio, J. Jagiello, M.
Thommes, J. Parra, M. Martnez, Hydrogen adsorption studies on single wall carbon nanotubes,
Carbon 42 (2004) 1243-1248.
[54] A. Myers, J.M. Prausnitz, Thermodynamics of mixed-gas adsorption, AICHE J. 11 (1965)
121-127.
[55] Z. Zhang, S. Xian, Q. Xia, H. Wang, Z. Li, J. Li, Enhancement of CO2 Adsorption and CO2/N2
Selectivity on ZIF-8 via Postsynthetic Modification, AICHE J. 59 (2013) 2195-2206.

24

Figure Captions

Figure 1. XRD patterns of HKUST-1 and GrO@HKUST-1 composites.


Figure 2. SEM images of (a) HKUST-1, (b) 1GrO@HKUST-1, (c) 2GrO@HKUST-1 and (d) 3GrO@HKUST-1.
Figure 3. Nitrogen adsorptiondesorption isotherms at 77 K on the HKUST-1 and GrO@HKUST-1 composites.
Figure 4. TG curves of HKUST-1 and GrO@HKUST-1 composites.
Figure 5. FTIR spectra of the HKUST-1 and GrO@HKUST-1 composites.
Figure 6. Isotherms of CO2 on HKUST-1 and GrO@HKUST-1 composites at 273 K.
Figure 7. The isosteric heats of CO2 adsorption on HKUST-1 and 2GrO@HKUST-1.
Figure 8. Comparison of the DSLF model fits and the experimental CO2 and N2 isotherm data on HKUST-1 (a) and
2GrO@HKUST-1 (b) at 273 K. (Symbols are experimental data points; solid lines are prediction by DSLF model).
Figure 9. IAST predicted selectivities for equimolar CO2/N2 mixtures on the HKUST-1 and 2GrO@HKUST-1 at
273 K.

25

Here is the information of corresponding authors.


* Corresponding authors. Tel.: +1 732 4453758 (J. Li), Tel.: +86 20 87110608 (Z. Li).
E-mail addresses: Jingli@rutgers.edu (J. Li), cezhli@scut.edu.cn (Z. Li).

26

1. The GrO@HKUST-1 composites can be rapidly prepared at room temperature


within 1 min.
2. The GrO@HKUST-1 composites have a higher surface area compared to its parent
HKUST-1.
3. Its CO2 adsorption capacity reached 9.02 mmol/g at 1 bar and 273 K, with an
increase of 32 % compared to HKUST-1.
4. Its CO2/N2 adsorption selectivity reached 186 at 1 bar, 1.8 times that of HKUST-1.

27

You might also like