You are on page 1of 8

Separation and Purication Technology 64 (2009) 280287

Contents lists available at ScienceDirect


Separation and Purication Technology
j our nal homepage: www. el sevi er . com/ l ocat e/ seppur
Hydrogen adsorption equilibrium and kinetics in metalorganic framework
(MOF-5) synthesized with DEF approach
Dipendu Saha, Zuojun Wei
1
, Shuguang Deng

Department of Chemical Engineering, New Mexico State University, P.O. Box 30001, MSC 3805, Las Cruces, NM 88003, USA
a r t i c l e i n f o
Article history:
Received 28 June 2008
Received in revised form
22 September 2008
Accepted 2 October 2008
Keywords:
MOF-5
Hydrogen
Adsorption
Equilibrium
Kinetics
a b s t r a c t
MOF-5, alsoknownas isoreticular MOF-1(IRMOF-1) was successfullysynthesizedwithdiethyl formamide
(DEF) as a solvent using modied procedures aiming at improving its crystal structure, pore texture and
ultimately the hydrogen adsorption performance. The MOF-5 adsorbent was characterized with nitro-
gen adsorption for pore textural properties, scanning electron microscopy for crystal structure, and XRD
for phase structure. Hydrogen adsorption in MOF-5 was measured at low pressure in a volumetric unit
at 77K, 194.5K and 298K and at hydrogen pressure up to 120bar and 77K in a gravimetric adsorp-
tion unit. The MOF-5 synthesized in this work has ideal pore textural properties with a large Langmuir
(3917m
2
/g) and BET specic surface area (2449m
2
/g), a median pore size of 8.6 and a pore volume of
1.39cm
3
/g. The MOF-5 adsorbent synthesizedinthis work has a hydrogenadsorptioncapacity of 1.46wt.%
at 77K and 1bar, an excess hydrogen adsorption capacity of 6.9wt.% at 77K and 100bar, and an absolute
hydrogen adsorption capacity of 11.8wt.% at 77K and 120bar. Hydrogen diffusivity in MOF-5 estimated
from the adsorption kinetic data measured at low pressure are 2.410
5
cm
2
/s, 5.210
5
cm
2
/s and
6.010
5
cm
2
/s at 77K, 194.5K and 298K, respectively. The activation energy for hydrogen diffusion
and the isosteric heat of adsorption for hydrogen adsorption in MOF-5 are 0.8kJ/mol and 2.22.6kJ/mol,
respectively.
2008 Elsevier B.V. All rights reserved.
1. Introduction
The hydrogen-based fuel cell power supply systems have a great
potential of replacing the fossil fuel-based transportation systems
because the fuel cell technologyis cleanandsustainable if hydrogen
can be produced from a renewable resource. On-board hydrogen
storage is one of the major challenges for the hydrogen fuel cell
powered mobile applications. The successful development of new
materials for hydrogen storage is the key to the success of hydro-
genfuel cell technology because the conventional hydrogenstorage
methods including compressed or cryogenic hydrogen storage sys-
tem cannot meet the current and future needs. Hydrogen can be
stored in a mediumin two ways: chemical adsorption and physical
adsorption. In chemical adsorption, hydrogen is reversibly bonded
to a substance like metallic hydrides [1] or nitrides [2]. The main
issue associated with chemical adsorption of hydrogen storage are
high release temperature, limited life time for the storage materials
and incomplete desorption of hydrogen [3]. In physical adsorp-

Corresponding author. Tel.: +1 575 646 4346; fax: +1 575 646 7706.
E-mail address: sdeng@nmsu.edu (S. Deng).
1
Permanent address: Zhejiang University, College of Material Science and Chem-
ical Engineering, Hangzhou, China.
tion, hydrogen gas is adsorbed in micropores and/or mesopores of
porous materials including zeolites [4], carbon nanotubes [5] and
clathrates [6]. Until the present time no current hydrogen storage
technology can meet the DOE target [7,8].
A new type of porous crystalline material, known as metal
organic framework (MOF) has gained signicant interests as
promising adsorbents for hydrogen storage and other gas sepa-
ration and purication [817]. The so-called reticular synthesis of
connectingZn
4
Oclusters withdifferent organic linkers has resulted
in six isoreticular MOFs (IRMOFs). It was also reported that hydro-
gen can be stored in MOFs containing metals other than Zn, such as
manganese [13] or copper [14]. MOF-5 (IRMOF-1), the rst mem-
ber of the MOF family with benzene dicarboxylic acid (BDC) as the
organic linker molecules, is probably the most widely investigated
MOF materials for hydrogen adsorption. Yaghis group reported
a recent progress on MOF-5 with improved hydrogen adsorption
capacity[18]. Applications of hydrogenspillover techniqueonMOF-
5materials have shownsignicant increase inhydrogenadsorption
capacity even at ambient temperature [1923]. Interaction of MOF-
5withhydrogenwas studiedby insituFT-IR[24], direct observation
of hydrogen adsorption sites was made possible by neutron pow-
der diffraction [25], and electronic and vibrational properties of
MOF-5 were measured in a photoluminescence and Raman spec-
troscopic study [26]. MOF-5 can be synthesized with two different
1383-5866/$ see front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.seppur.2008.10.022
D. Saha et al. / Separation and Purication Technology 64 (2009) 280287 281
Nomenclature
a
m
Langmuir equation constant (wt.%)
b Langmuir equation constant (bar
1
)
D
c
intracrystalline diffusivity (cm
2
/s)
D
c0
pre-exponential factor in Eq. (5) (cm
2
/s)
E diffusion activation energy dened in Eq. (5)
(kJ/mol)
k Freundlich equation constant
m
A absolute
absolute adsorbed amount (wt.%)
m
A excess
excess adsorbed amount (wt.%)
m
t
adsorbedamount per unit mass of adsorbent at time
t (wt.%)
m
max
maximumadsorbed amount per unit mass of adsor-
bent at t =(wt.%)
n Freundlich equation constant
p, P pressure (bar)
P
0
gas saturation pressure at temperature T (bar)
q adsorbate concentration in the adsorbent (wt.%)
r
c
radius of single crystal of MOF-5 (cm)
t time (s)
T absolute temperature (K)
H isosteric heat of adsorption (kJ/mol)
solvents: dimethyl formamide (DMF) and diethyl formamide (DEF)
[3,2729]. It is believed that MOF-5 synthesized by DEF approach
has a higher specic surface area and a higher hydrogen adsorption
capacity that the MOF-5 synthesized with DMF approach.
The objective of this work is to synthesize MOF-5 with DEF as
solvent to increase the specic surface area and hydrogen adsorp-
tion capacity of MOF-5. The MOF-5 samples synthesized in this
work were characterized for their physical properties with nitro-
genadsorption, scanningelectronmicroscopyandX-raydiffraction.
Hydrogen adsorption equilibriumand kinetics in MOF-5 were then
measured at 77K, 194.5K and 298K at hydrogen pressure up to
1.2bar, and at 77K and hydrogen pressure up to 120bar.
2. Materials and methods
2.1. Synthesis of MOF-5
Synthesis of MOF-5 were performedfollowing the reportedpro-
cedures [28,29] witha fewmodications. All the chemicals usedfor
MOF-5 synthesis in this work were purchased from Fisher Scien-
tic; theyareof commerciallyavailablehighest purity(99+%, except
zinc nitrate hexahydrate of 98% purity). 0.832g of zinc nitrate hex-
ahydrate and 0.176g of benzene dicarboxylic acid were dissolved
in 20ml of DEF under constant agitation at atmospheric condi-
tions. The resulting mixture was rst degassed thrice using the
freezepumpthawmethod, and then lled up 20ml reaction vials
for crystallization. The capped vials were immediately put in an
oven at 8590

C for crystallization for about 24h. At the end of


the crystallization step clear golden crystals of MOF-5 emerged
from the wall and base of the vials. The MOF-5 crystals were sep-
arated from the reaction solution, washed with DEF to remove the
unreacted zinc nitrate, followed with purication in chloroform.
The chloroform purication was performed by adding chloroform
into 20ml vials containing the raw MOF-5 crystals. The vials were
capped tightly and put back to the oven at 70

C for another 3
days. Solvent in vials was replenished with fresh chloroform every
day. After the chloroform treatment, the MOF-5 crystals changed
from golden color to transparent. Because MOF-5 crystals are very
susceptible to moisture and air [30], they have to be stored and
transported in a Schlenk ask under a vacuumor in vials lled with
chloroform.
2.2. Characterization of MOF-5
The MOF-5 samples were characterized for their pore textu-
ral properties with nitrogen adsorption and desorption at liquid
nitrogen temperature in a Micromeritics

ASAP 2020 adsorption


apparatus. The pore textural properties including specic Langmuir
and BET surface areas, pore volume and pore size were obtained by
analyzing the nitrogen adsorption and desorption isotherms with
the Micromeritics

ASAP2020built-insoftware. Before startingthe


nitrogen adsorption measurements, each sample was activated by
degassing in situ at 373K for 12h to remove any guest molecule
(i.e. chloroform) from the pores of the MOF-5 crystals. To exam-
ine the sample crystal structure and size, the MOF-5 samples were
analyzed with scanning electron microscopy in a Hitachi TM-1000
table-top device without prior gold deposition. The MOF-5 samples
were alsoinvestigatedfor its crystal phase structure andcrystal size
by power A Rigaku

Miniex-II X-ray diffractometer with Cu K


emission, 30kV/15mA current and a K-lter. The X-ray scanning
speed was set at 2

/min and a step size of 0.02

in 2. A Jade 8
+
XRD
pattern processing software (MDI Inc, Livermore, CA) was used to
analyze the XRD data collected on the MOF-5 samples.
2.3. Hydrogen adsorption study
2.3.1. Hydrogen adsorption at low pressure
Hydrogen adsorption equilibrium and kinetics on MOF-5 sam-
ples were measured in the Micromeritics

ASAP 2020 adsorption


apparatus at liquid nitrogen temperature (77K), dry ice temper-
ature (194.5K) and ambient temperature (298K) and hydrogen
pressures ranging from 0bar to about 1.2bar. Ultra-high purity
hydrogen (99.999%) was introduced into a separate gas inlet of
the adsorption unit for the hydrogen adsorption measurements.
The adsorption of hydrogen on MOF-5 was determined by a volu-
metric method. The pressure change after the hydrogen reservoir
was connected to the MOF-5 sample chamber was monitored and
recorded. The hydrogen adsorption amount was calculated ther-
modynamically with an equation of state. The changes of hydrogen
pressure withtime were also recordedandconvertedinto transient
adsorption uptakes or adsorption kinetics, and the nal adsorption
amount at the terminal pressure gives the adsorption equilibrium
amount at a given hydrogen pressure.
2.3.2. Hydrogen adsorption at high pressure
Hydrogen adsorption in MOF-5 at 77K and hydrogen pres-
sure up to 120bar was performed gravimetrically in a Rubotherm
magnetic suspension balance. A process ow diagram for the
Rubothermbalance is showninFig. 1. This balance is equippedwith
an automatic ow gas dosing and pressure control system. Both
adsorptionequilibriumandkinetics canbemeasuredautomatically
inthis unit at temperatures rangingfrom77Kto723K, andpressure
ranging froma vacuumto 500bar. The detail operation procedures
for this balance were described in our previous work [16]. Due to
the buoyancy force involved in the weight measurements at ele-
vated pressures in this study, the hydrogen adsorption amount at
high pressure can be expressed as either excess adsorption amount
(m
A excess
) or absolute adsorption amount (m
A absolute
). The excess
adsorption is determined fromthe experimental measurements by
neglecting the adsorbed phase volume while the absolute adsorp-
tion amount is calculated by considering the adsorbed phase the
volume. The excess adsorption amount is close to the absolute
adsorption amount when the adsorption amount is small or the
gas density is small as compared with the adsorbed phase density.
282 D. Saha et al. / Separation and Purication Technology 64 (2009) 280287
Fig. 1. Process ow diagram of Rubotherm magnetic suspension balance with automatic ow gas dosing and pressure control system.
The excess adsorptiondeviates fromthe absolute adsorptionsignif-
icantly as gas pressure increases. In order to calculate the absolute
adsorption amount, following buoyancy correction was performed
by assuming the adsorbed phase has similar density as a liquid
phase of the adsorbate molecules at saturation temperature and
ambient pressure.
m
A absolute
=
m
A excess
1 (T, P)/
L
(1)
where (T, P) and
L
are densities of gaseous and liquid adsorbate
molecule (H
2
), respectively. Both excess and absolute hydrogen
adsorption isotherms on MOF-5 are reported in this work.
3. Results and discussion
3.1. Physical properties of MOF-5
3.1.1. SEM images
Fig. 2 is the scanning electron microscopy image of MOF-5 crys-
tals synthesized with diethyl formamide as solvent in this work.
Good crystal structure with minor defects was obtained on this
MOF-5 sample; the average crystal size from the SEM image in
Fig. 2. Scanning electron microscopy images of MOF-5 crystals.
D. Saha et al. / Separation and Purication Technology 64 (2009) 280287 283
Table 1
Pore textural properties of MOF-5.
Langmuir specic surface area (m
2
/g) 3917
BET specic surface area (m
2
/g) 2449
HorvathKawazoe median pore diameter () 8.6
HorvathKawazoe pore volume at p/p
0
=0.99 (cm
3
/g) 1.39
Fig. 2 is between 100m and 200m, which is consistent with
the reported data in a recent literature [28].
3.1.2. Pore textural properties
The pore textural properties of the MOF-5 sample are summa-
rized in Table 1. These textural properties were calculated fromthe
nitrogen adsorption and desorption isotherms plotted in Fig. 3(A).
The Langmuir and BET specic surface areas, the median pore size,
and the cumulative pore volume by the HorvarthKwazoe (HK)
method [31] were calculated using the ASAP-2020 built-in soft-
ware. The corresponding pore size distribution that was calculated
from the desorption isotherm using the HK method is displayed
in Fig. 3(B). The nitrogen adsorption uptake at 1 bar on our MOF-5
sample is about 17% higher than the reported value [10]. The Lang-
muir specic surface area of our MOF-5 sample is 3917m
2
/g, which
is about 10% higher than the originally reported value of Langmuir
surface area for MOF-5 [11], and similar to what was reported by
Wong-Foy et al. [12]. The BET surface area of our MOF-5 sample is
2449m
2
/g, which is higher than what was reported by Panella et
al. [28], but lower than the recently reported value of 3800m
2
/g
[18]. The cumulative pore calculated by the HK method [27] is
Fig. 3. (A) Nitrogen adsorption and desorption isotherms in MOF-5 at 77K. (B) Pore
size distribution of MOF-5 calculated by the HK method.
Table 2
Summary of crystallographic properties of MOF-5.
Cell type Space group Lattice
constant ()
Lattice
angle
Cell volume (
3
)
Orthogonal Pbam(55) a =23.097 =90 2596.5
b=14.129 =90
c =7.650 =90
1.39cm
3
/g, which is higher than the result of 1.19cm
3
/g reported
by Yaghis group [10]. The median pore diameter of our MOF-5
calculated by the HK method [27] is 8.6, smaller than 15.2
reported by Yaghis group [10]. Both Langmuir and BET specic sur-
face areas of MOF-5 with the DEF approach in this work are more
than three times larger than those of MOF-5 synthesized with the
DMF approach, and the average pore size of MOF-5 with the DEF
approach is about seven times smaller than those of the MOF-5
samples synthesized with the DMF as solvent in our previous work
[17].
3.1.3. XRD pattern
X-ray powder diffraction pattern for a MOF-5 sample synthe-
sized with DEF approach in this work is shown in Fig. 4. The main
peaks and their corresponding crystal planes are identied. The
overall pattern looks similar to the XRDpattern reported by Panella
et al. [28], and that reported by Huang et al. [32] and Hazovic et
al. [33] The intensities of the rst two peaks (2 of 7.23

, corre-
sponding to a d space of 12.8 and 2 of 9.84

, corresponding to a
d space of 9.1) were reversed, probably due to some alterations of
atomic orientation in crystal planes by solvent and other adsorbate
molecules that ll the micropores of MOF-5 [33]. This is also possi-
bly caused by partial interpenetration of certain crystals in MOF-5.
A sharp peak below at 2 of 9.84

, corresponding to a d space of
9.1 was observed. The crystallographic properties for this MOF-
5 sample were summarized in Table 2. The crystal phase structure
can be indexed as an orthogonal type with a space group Pbam(55),
which is different from the cubic structure reported in the litera-
ture[29,32]. This is probablycausedbytheresidual guest molecules
(chloroform solvent) that alter the crystal phase plane orientation
[33].
3.2. Adsorption equilibrium
The hydrogen adsorption isotherms at 77K, 194.5K and 298K
withhydrogenpressure upto1.2bar are plottedinFig. 5. The hydro-
gen uptake at 77K and 1bar is 1.46wt.%, which is about 7% higher
than the published result of 1.34wt.% at similar conditions [9].
Hydrogen uptakes at two other temperatures are 0.384wt.% and
0.012wt.% at 194.5K and 298K, respectively.
Langmuir and Freundlich isotherm models were used to corre-
late the adsorption isotherms. The Langmuir isotherm is written
as:
q =
a
m
bP
1 +bP
(2)
where q is the adsorbed hydrogen amount on MOF-5 (wt.%), p is
the hydrogen pressure (bar), a
m
(wt.%) and b (mmHg
1
) are the
Langmuir isotherm equation parameters. They can be determined
fromtheslopeandintercept of a linear Langmuir plot of (1/a) versus
(1/p).
Freundlich isotherm is given by:
q = kP
1/n
(3)
where k and n are the Freundlich isotherm equation parameters
that can be determined from the experimental hydrogen adsorp-
tionisotherms. These two isotherms were ttedto eachof the three
284 D. Saha et al. / Separation and Purication Technology 64 (2009) 280287
Fig. 4. Powder X-ray diffraction pattern of MOF-5 crystals.
isotherms. The adsorption isotherm equation parameters for both
Langmuir and Freundlich equations are listed in Table 3 for low-
pressure and high-pressure isotherms.
The high-pressure adsorption isotherm on MOF-5 was deter-
mined gravimetrically by the Rubotherm magnetic suspension
balance. Both excess and absolute adsorption isotherms of hydro-
genonMOF-5at 77Kandpressure upto120bar are plottedinFig. 6.
It is very encouraging to nd out the highest absolute adsorption
amount of hydrogen on MOF-5 synthesized in this work is 11.8wt.%
at 77K and 120bar, and the highest excess adsorption amount
is 6.9wt.% at 77K and 100bar. The excess adsorption amount
obtained in this work is slightly lower than the recently reported
value of 7.1wt.% at similar conditions, but our absolute adsorption
amount at 120bar is higher than all the published results at similar
conditions [912,18,28]. From the trend of the absolute adsorption
isotherm, the MOF-5 adsorbent was not saturated, more hydro-
gen could be adsorbed at higher hydrogen pressures. Although the
hydrogenadsorptioncapacityonthe MOF-5reportedinthis workis
Fig. 5. Hydrogen adsorption isotherms at 77K, 194.5K and 298K and lowpressure.
signicantly higher than the previously reported literature data, it
is still impossible to meet the DOE hydrogen storage requirements
by using MOF-5 at ambient temperature and elevated pressures.
Both Langmuir and Freundlich isotherm models were used to
correlatetheabsoluteadsorptionisothermandFreundlichequation
ts the data better than the Langmuir. The monolayer satura-
tion capacity calculated from Langmuir equation, a
m
, is about
10.98wt.%, more than twice the value reported by Panella et al. [28]
It must be pointed out that the adsorption isothermdescribes only
the absolute adsorption, not the excess adsorption that is signi-
cantly lower than an absolute adsorption amount at high pressure
due to buoyancy correction if the adsorption is determined gravi-
metrically.
3.3. Adsorption kinetics
It should be pointed out that hydrogen adsorption kinetics is as
important as the hydrogen adsorption equilibriumalthough meet-
ing the DOE specications for hydrogen storage capacity is more
Fig. 6. Hydrogen adsorption isotherms at 77K and pressure up to 120bar.
D. Saha et al. / Separation and Purication Technology 64 (2009) 280287 285
Table 3
Summary of adsorption isotherm model parameters for hydrogen in MOF-5.
Isotherm model Parameters T =77K (low pressure) T =77K (high pressure) T =194.5K (low pressure) T =298K (low pressure)
Langmuir am (wt.%) 1.92 10.98 1.1013 0.01689
b (bar
1
) 2.0051 0.0055 0.0005 0.0021
Freundlich k 0.01181 1.0971 0.00058 9.34210
5
n 1.3545 2.0040 1.0258 1.3666
challenging [7,8]. Up to now no hydrogen adsorption kinetic data
on MOF-5 has ever been reported. The hydrogen adsorption kinetic
data were collectedat the same time whenthe low-pressure hydro-
gen adsorption isotherm shown in Fig. 5 was measured in the
Micromeritics

ASAP 2020 adsorption unit. The fractional adsorp-


tion uptake curves at various hydrogen pressures were plotted in
Fig. 7(A)(C). A classical micropore diffusion model described by
Ruthven [34] was applied in order to process the hydrogen kinetic
data to extract the intracrystalline diffusivity for hydrogen in MOF-
5. The fractional adsorption uptake (m
t
/m

) can be correlated with


thediffusiontimebythefollowingequationif thefractional adsorp-
tion uptake is greater than 70% [34].
1
m
t
m
max
=
6

2
exp

2
D
c
t
r
2
c
(4)
The diffusion time constants (D
c
/r
2
c
, s
1
) can be calculated from
theslopeof alinear plot of ln(1(m
t
/m

)) versus t at agivenhydro-
gen pressure. Only data points shown in Fig. 7(A)(C) with (m
t
/m

)
greater than70%andless than99%were usedfor estimating the dif-
fusiontime constants. Anaverage crystal size for MOF-5 is assumed
to be 150m, based on SEM images shown in Fig. 3, the intracrys-
tallinediffusivity, D
c
, was thencalculatedbyfromthediffusiontime
constants (D
c
/r
2
c
). The average diffusivity of hydrogen in MOF-5
is about 2.410
5
cm
2
/s, 5.210
5
cm
2
/s and 6.010
5
cm
2
/s at
77K, 194.5K and 298K, respectively. These values are signicantly
higher than the hydrogen diffusivity in MOF-5 synthesized with
DMF as solvent in our previous study [17]. The estimated diffusiv-
ityis still signicantlylower thanthecalculatedKnudsendiffusivity
(1.210
2
cm
2
/s) for hydrogen owing through straight cylindri-
cal pore of MOF-5 with pore diameter of 20at 77K. The activation
energy for hydrogen diffusion inside the micropores of MOF-5 esti-
mated from the following equation is 0.8kJ/mol, which is in the
same order of magnitude of the activation energy of hydrogen dif-
fusion in MOF-177 [16] and the ordered mesoporous carbon [35].
D
c
= D
c0
e
(E/RT)
(5)
The effect of hydrogen pressure and temperature on hydro-
gen diffusivity time constant is shown in Fig. 8. It was found that
diffusivity increases with temperature, which is expected from a
theoretical point of view. Diffusivity also increases slightly with
hydrogen pressure at a constant temperature. It is believed the
surface diffusion along the MOF-5 internal surface may have con-
tributed to the overall diffusion. The adsorbed phase concentration
alongMOF-5internal surface increases withhydrogenpressure and
therefore provides higher driving force for the hydrogen diffusion
inside the MOF-5 channels.
Fig. 7. Hydrogen adsorption fractional uptakes at 298K (A); 194.5K (B); and 77K (C).
286 D. Saha et al. / Separation and Purication Technology 64 (2009) 280287
Fig. 8. Effect of pressure and temperature on hydrogen pore diffusivity in MOF-5.
The importance of hydrogen adsorption/desorption kinetics in
MOF-5 is well reected in the DOE hydrogen storage require-
ments. For automotive fuel cell applications, the time period for
relling and releasing hydrogen from a hydrogen storage tank is
restricted. The kinetic results from this work have clearly shown
that adsorption of hydrogen in MOF-5 is fast enough that the rell-
ing and release hydrogen from the porous media will not be an
issue.
3.4. Isosteric heat of heat of adsorption
Isosteric heat of adsorption is also an important criterium for
hydrogen storage through physical adsorption. The isosteric heat
of adsorption can be obtained from the vant Hoffs equation:
H
RT
2
=

lnp
T

a
(6)
where His the isosteric heat of adsorption (kJ/mol), T is tempera-
ture, p is pressure (bar), a is the adsorption amount (wt.%), R is the
Fig. 9. Effect of adsorption amount on isosteric heat of adsorption of hydrogen in
MOF-5.
universal gas constant. Integrating equation (6) gives:
ln P =
H
RT
+C (7)
where C is an integration constant.
In the present study, two hydrogen adsorption isotherms at
77K and 194.5K were used to calculate the heat of adsorption.
The hydrogen adsorption isotherms were rst converted to hydro-
gen adsorption isosteres, a plot of p versus T at a given adsorption
amount. The heat of adsorptionwas thencalculatedfromthe slopes
of isosteres according to Eq. (7). The values of heat of adsorption for
hydrogen in MOf-5 obtained in this work are about 2.22.6kJ/mol,
which are smaller than the reported value of 4kJ/mol by Gogotsis
group [36]. As shown in Fig. 9 the heat of adsorption decreases
slightly with the hydrogen adsorption amount.
4. Conclusions
MOF-5 synthesized with diethyl formamide as solvent and
modied procedures has ideal pore textural properties with large
Langmuir (3917m
2
/g) and BET specic surface areas (2449m
2
/g),
an average pore size of 8.6 and large total pore volume of
1.39cm
3
/g. Hydrogen adsorption equilibriumand kinetics in MOF-
5 were measured at 77K, 194.5K and 298K at hydrogen pressure
up to 1.2bar and at 77K and hydrogen pressure up to 120bar. The
MOF-5 adsorbent has a hydrogen adsorption capacity of 1.46wt.%
at 77K and 1bar, excess hydrogen adsorption capacity of 6.9wt.%
at 77K and 100bar, and absolute hydrogen adsorption capacity of
11.8wt.% at 77K and 120bar. The Freundlich adsorption equation
ts both low pressure and high-pressure hydrogen isotherm data
well. Hydrogen diffusivity in MOF-5 estimated from the adsorp-
tion kinetic data measured at hydrogen pressure up to 1.2bar
are 2.410
5
cm
2
/s, 5.210
5
cm
2
/s and 6.010
5
cm
2
/s at 77K,
194.5K and 298K, respectively. The activation energy and isosteric
heat of adsorption for hydrogen adsorption in MOF-5 are 0.8kJ/mol
and 2.22.6kJ/mol, respectively.
Acknowledgement
The authors wish to express their gratitude to US Army Research
Ofce for providing nancial support to this project through grant
W911NF-06-1-0200.
References
[1] A. Zaluska, L. Zaluski, J.O. Strom-Olsen, Method of fabrication of complex alkali
metal hydrates, US Patent 625349B1 (2001).
[2] S.H. Jhe, Y.K. Kwon, Phys. Rev. B 69 (2004) 245407-1245407-6.
[3] B. Panella, M. Hirscher, Adv. Mater. 17 (2005) 538541.
[4] J. Weitkamp, M. Fritz, S. Ernst, Int. J. Hydrogen Energy 20 (1995) 967970.
[5] A.C. Dillon, K.M. Jones, T.A. Bekkedahl, C.H. Kiang, D.S. Bethune, M.J. Heben,
Nature 386 (1997) 377379.
[6] H. Marcel, F. Sluiter, H. Adachi, R.V. Belosludov, V.R. Belosludov, Y. Kawazoe,
Mater. Trans. 45 (2004) 14521454.
[7] Ofce of Science, Basic Research Needs for the Hydrogen Economy, U.S. Depart-
ment of Energy, February 2004.
[8] S. Deng, in: S. Lee (Ed.), Encyclopedia of Chemical Processing, Marcel Dekker,
Inc, New York, NY, 2006, pp. 28252845.
[9] J.L.C. Rowsell, O.M. Yaghi, Micropor. Mesopor. Mater. 73 (2004) 314.
[10] J.L.C. Rowsell, O.M. Yaghi, Angew. Chem. Int. Ed. 44 (2005) 46704679.
[11] J.L.C. Rowsell, A.R. Millward, K.S. Park, O.M. Yaghi, J. Am. Chem. Soc. 126 (2004)
56665667.
[12] A.G. Wong-Foy, A.J. Matzger, O.M. Yaghi, J. Am. Chem. Soc. 128 (2006)
34943495.
[13] M. Dinca, A. Daily, Y. Liu, C.M. Brown, D.A. Neumann, J.R. Long, J. Am. Chem. Soc.
128 (2006) 1687616883.
[14] P. Krawiec, M. Kramer, M. Sabo, R. Kunschke, H. Frode, S. Kaskel, Adv. Eng. Mater.
8 (2006) 293296.
[15] Q.M. Wang, D. Shen, M. Blow, M.L. Lau, S. Deng, F.R. Fitch, N.O. Lemcoff, J.
Samanscin, Micropor. Mesopor. Mater. 55 (2002) 217230.
[16] D. Saha, Z. Wei, S. Deng, Int. J. Hydrogen Energy (in press).
D. Saha et al. / Separation and Purication Technology 64 (2009) 280287 287
[17] S. Saha, S. Deng, Z. Yang, J. Porous Mater. (in press).
[18] S.S. Kaye, A. Daily, O.M. Yaghi, J.R. Long, J. Am. Chem. Soc. 129 (2007)
1417614177.
[19] R.T. Yang, Y.W. Li, G.S. Qi, A.J. Lachawiec, US Patent Applications Serial No. 2006;
11:442898 and Serial No. 2007; 11:820954.
[20] Y.W. Li, R.T. Yang, J. Am. Chem. Soc. 128 (2006) 726727.
[21] Y.W. Li, R.T. Yang, J. Am. Chem. Soc. 128 (2006) 81368137.
[22] Y.W. Li, R.T. Yang, J. Phys. Chem. C 111 (2007) 34053411.
[23] Y.W. Li, R.T. Yang, AIChE J. 54 (2007) 269279.
[24] S. Bordiga, J.G. Vitillo, G. Richiardi, L. Regli, D. Cocina, A. Zecchina, B. Arstad, M.
Bjorgen, J. Hazovic, K.P. Lillerud, J. Phys. Chem. B 109 (2005) 1823718242.
[25] T. Yildirin, M.R. Hartman, Phys. Rev. Lett. 95 (2005) 215504215504-4.
[26] S. Bordiga, C. Lamberti, G. Richiardi, L. Regli, F. Bonino, A. Damin, K.P. Lillerud,
M. Bjorgen, A. Zecchina, Chem. Commun. (2004) 23002301.
[27] H. Li, M. Eddaoudi, M. OKeeffe, O.M. Yaghi, Nature 402 (1999) 276279.
[28] B. Panella, M. Hirscher, H. Putter, U. Muller, Adv. Funct. Mater. 16 (2006)
520524.
[29] J.L.C. Rowsell, Hydrogenstorage inmetalorganic frameworks: aninvestigation
of structureproperty relationships. Ph.D. Dissertation. University of Michigan,
2005, pp. 169172.
[30] D. Saha, S. Deng, Adsorption (in press).
[31] G. Horvath, K.J. Kawazoe, Chem. Eng. Jpn. 16 (1983) 470475.
[32] L. Huang, H. Wang, J. Chen, Z. Wang, J. Sun, D. Zhao, Y. Yan, Micropor. Mesopor.
Mater. 58 (2003) 105114.
[33] J. Hazovic, M. Bjrgen, U. Olsbye, P.D.C. Dietzel, S. Bordiga, C. Prestipino, C.
Lamberti, K.P. Lillerud, The inconsistency in adsorption properties and pow-
der XRD data of MOF-5 is rationalized by framework interpenetration and the
presence of organic and inorganic species in the nanocavities, J. Am. Chem. Soc.
129 (2007) 36123620.
[34] D.M. Ruthven, Principles andAdsorptionandAdsorptionProcesses, WileyInter-
science, 1984, pp. 166205.
[35] D. Saha, Z. Wei, S.H. Valluri, S. Deng, J. Porous Media (in press).
[36] G. Yushin, R. Dash, J. Jagielo, J. Fischer, Y. Gogotsi, Adv. Funct. Mater. 16 (2006)
22882293.

You might also like