Вы находитесь на странице: 1из 7

Journal of Industrial Engineering Research, 1(5) Special 2015, Pages: 1-7

IWNEST PUBLISHER

Journal of Industrial Engineering Research


(ISSN: 2077-4559)

Journal home page: http://www.iwnest.com/AACE/

Hydrogen Adsorption on Agricultural-Based Activated Carbons, Zeolite


Templated-Carbons and Clay-Based Materials; A Review
Nikmah Utami Sawai and Farah W. Harun

Universiti Sains Islam Malaysia, Department of Industrial Chemical Technology, Faculty of Science and Technology, 71800 Nilai, Negeri
Sembilan, Malaysia.

ARTICLE INFO ABSTRACT


Article history: Hydrogen storage has become the main challenge in making hydrogen-fuelled economy
Received 23 March 2014 into reality. Among the approaches is the adsorption on solid-state material. Porous
Accepted 24 April 2015 materials, such as agricultural-based activated carbon, zeolite templated-carbons, and
Available online 28 April 2015 clay-based materials are known to possess a great utility in adsorption of hydrogen.
This paper aims to review and summarize the recent advances of the materials that are
Keywords: currently under investigation to store hydrogen in solid-state materials. In this paper,
Hydrogen adsorption comparison between the achievements of the agricultural-based activated carbon,
Hydrogen storage zeolite templated-carbons and clay-based materials as the sorbent materials has been
Activated carbon discussed. In a nutshell, by tailoring their textural properties, they show a remarkable
Agricultural potential in storing hydrogen.
Zeolite templated-carbons
Clay-based materials
© 2015 IWNEST Publisher All rights reserved.
To Cite This Article: Nikmah Utami Sawai and Farah W. Harun., Hydrogen adsorption on agricultural-based activated carbons, zeolite
templated-carbons and clay-based materials; a review. J. Ind. Eng. Res., 1(5), 1-7, 2015

INTRODUCTION

Recently, hydrogen has been the focus of research as an alternative fuel to replace the traditional fuels such
as oil and natural gas due to the numerous benefits of hydrogen. Hydrogen is known as a clean energy carrier
since the only by-product when it is burned in internal combustion is water and it also possesses higher energy
content per mass unit compared with the widely used gasoline and coal. In addition, the other criteria of
hydrogen is it is not a primary fuel, but it have to be manufactured like electricity [1, 2]. In achieving hydrogen-
fuelled economy into a reality, hydrogen storage issues must be tackled since it is the obstacle that impedes
hydrogen future. Compressed gas, liquefaction, metal hydrides are among the approaches to store hydrogen,
nevertheless they still did not satisfy the criteria of size, efficiency, cost and the safety for on-board [2, 3]. In
addition, among various approaches for storing hydrogen, the sorption of molecular H 2 in solid state porous
materials is attractive due to the fast kinetics, excellent cyclability and high adsorption capacity. The use of
porous materials, in general, will reduce both the gravimetric and the volumetric hydrogen storage densities. In
addition, the increased in surface area and porosity in porous materials will provide additional adsorption sites
on the surface and inside the pores which can increase the hydrogen storage capacity through physisorption. The
potential to store a significant amount of H2 on high surface area porous materials is the main driving force in
the investigation on H2 sorption properties of porous materials, including porous carbons, zeolites, metal–
organic frameworks, porous polymers and many others. There are many studies to date have been conducted for
hydrogen storage on various solid state porous materials such as the carbon-based material, zeolites, and metal-
organic frameworks (MOF). However, there are seldom materials can achieve the claims of DOE (Department
of Energy, USA) criterion for the storage uptake on practical applications [3]. In this present review, the
performance of the agricultural-based activated carbon, zeolite templated-carbons and clay-based materials as
the sorbent materials is summarized.

Agricultural-based activated carbon:


To date, the performance of carbon-based materials such as activated carbons (AC), graphite, carbon
nanofibers (CNF) and carbon nanotubes (CNT) as hydrogen storage materials has been an intense research [4].
Numerous advantages such as light weight, easy availability, fast adsorption desorption kinetics, complete
reversibility and low cost, are among the reasons causing the porous carbon materials as a potential candidate
Corresponding Author: Nikmah Utami Sawai. Universiti Sains Islam Malaysia, Department of Industrial Chemical
Technology, Faculty of Science and Technology, 71800 Nilai. Negeri Sembilan. Malaysia.
Ph: +6067987013/6531. E-mail: nikmahutami@gmail.com
2 Nikmah Utami Sawai and Farah W. Harun, 2015
Journal of Industrial Engineering Research, 1(5) Special 2015, Pages: 1-7

for hydrogen adsorbent [5]. Additionally, activated carbon has captured the attention of the research community
due to their simple preparation. However, their relatively low surface area and broad pore size distribution are
limiting their ability to store high amount of hydrogen, hence physical or chemical activation is required to
enhance the specific surface area of porous carbon [5,6]. In physical activation, carbon materials undergo
gasification in the presence of suitable oxidizing gasifying agents, such as CO2 and steam, while in chemical
activation, the reaction occurred between the chemical agents such as KOH and NaOH and the carbon materials.
Lower processing temperature, shorter reaction time, and higher microporosity are among the benefits possessed
by chemical activation compared to the physical activation. However, the main drawback of chemical activation
is that the impurities generated during the activation process must be thoroughly washed away [6]. In this
present review, we summarized the hydrogen adsorption capacity on chemically-activated porous carbon
fabricated from agricultural. The reported results show the hydrogen storage capacity on agricultural-based
porous carbons is scattered over several orders of magnitude in between 1.0 and 10 wt.% depending on the
storage parameters [5, 7-12].
In previous literature, the hydrogen storage ability of KOH-activated carbon material fabricated from coffee
bean wastes was investigated by Akasaka [7] regarding pore structure. The maximum amount of stored
hydrogen on porous carbon material with 2070 m2/g of specific surface area at 298 and 77 K was 0.6 wt.% and
4.0 wt.%, respectively. They also reported that the two-dimensional storage configuration have changed to
three-dimensional configuration due to these changes in the dependence of the hydrogen storage ability on pore
size. The densities of adsorbed hydrogen in porous carbon materials calculated from these values were 5.7 and
69.6 mg/cm3 at 298 and 77 K, respectively, where the change indicated that the state of stored hydrogen in
porous carbon materials is filled up aggregational state, which is extremely close to the liquid state. The
activation process influenced the specific surface area of the synthesized materials as the specific surface area
increased from 780 to 2070 m2/g with the increased in weight ratio from 1 to 5 of KOH to carbon materials.
Moreover, the increased in equilibrium pressure has increased the hydrogen uptake capacity in all samples;
maximum hydrogen uptake capacity was obtained at 12 MPa. The type of adsorbed hydrogen on these carbon
materials is phsysisorption since the hydrogen adsorption capacity was reduced when hydrogen pressure is
reduced, and almost stored hydrogen is desorbed from these porous carbon materials [7].
Next, Chen et al. [5] investigated the hydrogen adsorption capacity on NaOH-activated rice hull-based
carbon. Rice hull, which undergone carbonization and activation by NaOH, exhibits high specific surface area
over 4000 m2/g. In their study, the rice hull was synthesized by direct carbonization at 400 °C, then activated
with NaOH by varying temperature rising segments from 1 to 4 segments. Rapid temperature increase causing
an intense etching of NaOH, which eventually is causing the pores to collapsed, while the previous pores
became larger. Furthermore, the surface area is reduced with the increase in heating ramp rate. They reported a
high hydrogen uptake capacity of 7.7 wt% at 77 K and 1.2 MPa. The study also found that in all samples, high
surface area and large micropore volume resulted in high hydrogen uptake capacity. The presence of large
micropores volume may contribute more at 1.2 MPa than that at 0.1 MPa.
In other study, Xiao and the co-workers [8] reported the performance of KOH-activated porous carbon
fabricated from melaleuca bark as potential candidate for hydrogen storage. The samples were prepared by
varying the mass ratio of KOH/carbonized sample. As a result, the highest adsorption capacity of 4.08 wt.% at
77 K and 10 bar was obtained on samples with mass ratio of 5 KOH: 1 carbonized sample. The sample also
reported high surface area, which was up to 3170 m2/g. Generally, as reported in other literature, high hydrogen
uptake capacity was contributed by the increased in surface area and micropore volume. However in this study,
they reported the sample which obtained highest hydrogen uptake (5 KOH: 1 carbonized sample) possess lower
total pore volume compared to sample with mass ratio of 6 KOH: 1 carbonized sample (Table 1). High
hydrogen uptake is due to higher micropore volume existed in the sample with mass ratio of 5 KOH: 1
carbonized, since greater proportion of pore volume should come from micropores for a desirable H 2 adsorption.

Table 1: Textural parameters and hydrogen uptake on KOH-activated porous carbon fabricated from melaleuca bark [8].
Sample SBET VT Vmicro H2 uptake (wt%)
(m2/g) (cm3/g) (cm3/g)
5 KOH: 1 carbonized 3170 1.51 1.07 4.08
6 KOH: 1 carbonized 2986 1.63 0.86 3.91

Previously, significant hydrogen uptake capacity of 5.05 wt.% at 77K and 10 bar was reported by Cheng et
al.[9]. They investigated the hydrogen storage performance on hydrolytic lignin from sawdust at 77 K over the
pressure range of 0–10 bar. All samples possess high surface areas exceeding 2000 m2/g and large pore volumes
ranging from 1.11-1.68 cm3/g. Based on the nitrogen adsorption isotherms, they suggested that the synthesized
materials exhibited high volume of micropores along with the absence of macropores or mesopores due to very
high adsorption of N2 in a low relative pressure range, therefore a narrow pore size distribution over the range of
0.77-0.91 nm was obtained on the porous carbons. In addition, by increasing temperature and prolonged
dwelling time, product yield is decreased due to the high amount of carbons being consumed which
3 Nikmah Utami Sawai and Farah W. Harun, 2015
Journal of Industrial Engineering Research, 1(5) Special 2015, Pages: 1-7

consequently causing the increase in specific surface area, pore volume, and average pore size. It shows that the
textures of the synthesized materials were greatly influenced by the reaction parameters. Moreover, the study
shows an increasing trend of hydrogen uptake capacity by increasing hydrogen pressure. The same sample that
obtained 5.05 wt.% at 10 bar, showing an uptake capacity of 2.55 wt.% at 1 bar.
Minoda et al. [10] compared the ability of KOH-activated PAN fiber precursor and KOH-activated rice
husk carbon to adsorb hydrogen at a temperature of 303 K and pressure ranges of 0–10 MPa and 0–35 MPa. The
study found that larger micropore volume is exhibited by KOH-activated PAN fiber precursor than KOH-
activated rice husk carbon. Other than that, the maximum hydrogen adsorption capacity observed for KOH-
activated PAN fiber precursor and KOH-activated rice husk carbon is 1.1 wt.% at 30 MPa and 1.0 wt.% at 27
MPa, respectively. The increase in hydrogen storage capacity show linear relationship with the increase in BET
surface area and micropore volume in almost all the synthesized samples. However, they reported on lower
amount of hydrogen adsorbed on their synthesized sample having the maximum micropore size of 1.5 nm,
which is not expected. They concluded that micropore size of 1.5 nm was too large hydrogen storage in mild
temperature and therefore they suggested that smaller size micropores of around 0.9–1.3 nm are suitable for
hydrogen storage.
Wang [11] studied the adsorption equilibrium and kinetics of H 2, CH4, and CO2 on corncob-derived
activated carbon, which is activated by potassium hydroxide (KOH), at different temperatures. At 0.1 MPa, the
maximum adsorption capacity of H2, CH4, and CO2 on the sample was 12.76 (77 K), 7.66 (213 K), and 3.56
(301 K) mmol/g, respectively. The report have revealed that there is strong interaction between the gases and
activated carbon since the adsorption amount increases rapidly with increasing pressure in the low pressure
range. In the other hand, due to the exothermic feature of gas adsorption, the rise in temperature at the same
pressure causing decreasing pattern of the adsorption amount. The adsorption of gas at low temperature was
more favorable. Under the condition of 101 kPa, the adsorption amount of H 2 reached 12.76 mmol/g at 77 K due
to a high surface area of 2789 cm3/g and high total pore volume of 1.55 cm3/g, which give advantages to H2
adsorption. Based on the SEM photographs of corncob-based, they found that the honeycomb shaped holes are
well arranged and regular, having thick and smooth walls of the holes with clear angle lines and large amount of
irregular granules dispersed in the holes. The existence of these irregular granules contributed to the high
surface area and high total pore volume of the corncob.

Table 2: Surface area and hydrogen uptake capacity of agricultural-based activated carbon.
Activated carbon sample SBET H2 uptake Ref.
(m2/g) (wt %)
Coffee bean 2070 4.0 [5]
Rice hull 3969 7.7 [7]
Malaleuca 3170 4.08 [8]
Saw dust 3100 5.05 [9]

There is a linear relationship between the specific surface area (S BET) with the hydrogen uptake at 77 K as
shown in Table 2. Rice hull activated carbon possesses the highest adsorption capacity compared to the other
samples due to its high surface area which is 3969 m2/g. Large specific surface area is required for desirable
hydrogen adsorption within a certain pressure range. This is due to the mechanism for the adsorption of
hydrogen was due to the van der Waals forces between the huge surface area of porous carbon and hydrogen
molecules [12]. The maximum specific surface area reported was higher than specific surface area of 200-1315
m2/g for carbon nanomaterials such as carbon nanotube [7].
Besides, carbon materials show a temperature-dependence trend since low temperature is more favorable
for hydrogen adsorption; the uptake capacity at 77 K is always greater compared to high adsorption temperature.
This is because, at 77 K, the hydrogen molecule is easily adsorbed to the walls of carbon micropores by van der
Waals force due to the low kinetic energy of hydrogen molecules. It is contributed by the hydrogen molecules
agglomerate inside the micropore due to the greater interaction between hydrogen molecules at that temperature.
In addition, a linear relationship between the uptake capacity and the micropore volume at 77 K was reported
[7]. Besides that, the adsorption pressure also influenced the uptake capacity; by rising the adsorption
temperature, we can see an increasing trend of hydrogen uptake capacity [9]. In conclusion, hydrogen storage
capacity increased with decreasing temperature and increasing pressure, but the amount of hydrogen storage for
porous carbon had a certain limit [5, 7-12].
Other than that, if the porous carbons exhibited high surface area and large pore volume, therefore the
micropore volume and small mesopore volume were more important than total pore volume for the hydrogen
storage [8]. The textural characteristics of activated carbons such as surface area, pore size and pore volume of
porous carbons remarkably depended on preparation parameters [5, 7-8]. Hence, according to the above
literature, the activated carbon derived from agricultural shows a promising ability to store hydrogen. In
summary, in order to achieve high capacity of hydrogen adsorption on porous carbon, they must fulfilled the
4 Nikmah Utami Sawai and Farah W. Harun, 2015
Journal of Industrial Engineering Research, 1(5) Special 2015, Pages: 1-7

essential requirements; large amount of active sites, high surface area, large micropore volume related to
optimum pore size [5, 7-12].

Zeolite templated-carbon:
A lot of studies have been conducted and reported that the presence of zeolite in carbonization process may
greatly improve carbons properties such as the morphology, pore structure, and adsorptive behavior. Selection
of a proper inorganic template, precursor and pyrolytic conditions are important since it will tailor the pore
structure of carbon, and consequently may enhance the hydrogen storage capacity. The preparation of
templated-carbons consists of two main steps; pyrolysis of precursor formed in highly ordered micro-channels
or meso-channels of inorganic materials and the destruction of templates [13].
Previously, two types of activated carbon were synthesized by using ammonium-form zeolite Y as the
template, while for the carbon precursors; one was using the using polyfurfuryl alcohol (PFA) [14] while the
other one was using the sucrose [15]. Both template-carbons presented a specific surface area of 2545 m2/g and
1500 m2/g, respectively. At 77 K and 10 bar, both carbon samples exhibited hydrogen storage capacity of 2.4
wt.% [14, 15].
In other literature, they reported an increment between 60 and 96% of the hydrogen uptake capacity of
zeolite Y templated-carbon is contributed by the KOH activation [16]. Sevilla [16] investigated the effects of
chemical activation using potassium hydroxide (KOH) on textural properties of zeolite-templated carbons
(ZTCs) and also compared the performance of zeolite-templated carbons (ZTCs) and KOH-activated zeolite
template-carbon (activated ZTCs) as hydrogen adsorbent. By tailoring the activation parameters; carbon/ KOH
ratio, impregnation method, activation temperature, heating rate, activation time, and gas flow rate, KOH
activation may generate carbon materials with enhanced porosity and uniform pores. KOH activation mainly
acts by generating large micropores and small mesopores with the actual size of pores generated being
dependent on the level of graphitization of the ZTC; graphitization favored the formation of larger pores.
Therefore, in their study, they found that there was enhancement in both textural properties and the hydrogen
uptake capacity of the synthesized material. The surface area increased from 1400-1650 m2/g to 1850 and 3100
m2/g which shows an increment of 84%, whereas the pore volume of 0. 8-1.1 cm3/g for the ZTCs is increased by
50% to 1.5-1.75 cm3/g for the activated ZTCs. Other than that, at 77 K and 20 bar, the capacity increased from
2.4-3.5 wt.% for ZTCs to between 4.3 and 6.1 wt.% for activated ZTCs [16].
Later, Almasoudi [17] synthesized highly microporous ZIF-templated carbon via ‘‘hard template
carbonization technique’’ by using zeolitic imidazolate framework (ZIF) as template. In this technique, furfuryl
alcohol was first impregnated into the pores of the ZIF before undergoing polymerization and carbonization.
During carbonization process, 90-95% microporosity of microporous carbon has been generated with a surface
area of 900–1100 m2/g and a pore volume of ca. 0.7 cm3/g, by removing the ZIF template. They also
investigated the effect of KOH activation on the material and found that after the activation, the porosity is
greatly increased. The activated ZIF templated-carbon retains significant microporosity and exhibits higher
surface area and larger pore volume of 3200 m2/g and 1.94 cm3/g, respectively. 80–90% of surface area and 60–
70% of pore volume owing to the micropores existed in the materials. This occurs because the activation
process mainly enhances existing porosity rather than creating new larger pores. At 77 K and 20 bar, ZIF
templated-carbon which obtained hydrogen uptake capacity of 2.6-3.1 wt.% is increased to the range of 3.9-6.2
wt.% after the activation; which show a significant enhancement between 25 and 140%. This is because, after
the activation, there is rises in the micropore surface area and micropore volume rather than overall increase in
porosity.
Next, the effects of synthesis parameters on template-carbon synthesized using NH4Y-zeolite as template
and furfuryl alcohol as carbon precursor by carbonization method was investigated by Konwar et al. [4]. They
tailored the carbonization temperature and dwelling time, in range of 650–850 °C and 1–4 h, respectively. The
results show that for all carbonization temperatures, the increased in dwelling time causing the BET and
micropore area increased. However, at 3 h of dwelling time, they reached the maximum reading. At 3 h
dwelling time, 750 °C of carbonization temperature recorded highest BET surface area and pore volume of
template, while slightly higher micropore area was recorded at 650 °C. In addition, the samples have been
prepared via two types of heating, continuous and stepwise, both showing maximum hydrogen uptake (100 °C
and atmospheric pressure) at 3h dwelling time. For continuous heating, hydrogen uptake of 0.15 wt.% was
obtained on carbon synthesized at 650 °C, while for stepwise heating hydrogen uptake of 0.29 wt.% was
obtained on carbon synthesized at 750 °C, which shows an increasing trend. Besides that, higher BET surface
area of 1886 m2/g and micropore area of 1136 m2/g have been recorded on the carbon synthesized at 750 °C
from stepwise heating. They also concluded that the samples exhibited large amount of micropores and
mesopores of size less than 6 nm recorded higher hydrogen uptake capacity.
Masika [18] explored the preparation of ultrahigh surface area carbons using zeolite 13X as the template
and reported high hydrogen uptake of 7.3 wt.% at 20 bar and 77 K. The sample was prepared via two-step
nanocasting technique; liquid impregnation (LI) of furfuryl alcohol and chemical vapour deposition (CVD) of
5 Nikmah Utami Sawai and Farah W. Harun, 2015
Journal of Industrial Engineering Research, 1(5) Special 2015, Pages: 1-7

ethylene. During the first stage, liquid impregnation stage, firstly carbon precursor is filled into the zeolite
channels followed by polymerization of the precursor and partial carbonization as the first stage. Next, chemical
vapour deposition (CVD) took place as the second stage where different carbon precursor were used onto the
zeolite/carbon composite obtained in the first step. In addition, the heating ramp rate was also varied between 5,
10 or 15 °C/min and the resulting materials possess zeolite-like structural ordering. Lowest ramp rate, 5 °C /min,
generated zeolite-templated carbon with highest surface area and pore volume of 3332 m 2/g and 1.66 cm3/g,
respectively.
Later study by Cai [3] showed the use of zeolite USY as the template in the preparation of sucrose-based
nanoporous carbons. Polymerization and carbonization process of sulfuric acid-pretreated sucrose solution was
tailored by adjusting the zeolite/sucrose mass ratios and temperatures. The synthesized material exhibited large
surface area and pore size, while the morphology of carbons is retained, and micropore in the carbons is came
from the sucrose precursor. Hydrogen adsorption study at 77 K and 20 bar revealed carbon with a surface area
of 1200 m2/g exhibits a hydrogen storage capacity of up to 3.65 wt.%. Other than that, they also found that the
hydrogen capacities at higher pressure are found to be much more correlated with surface area and micropore
volume of carbons [3].
Recently, Cai and the team [13] had synthesized sucrose-based porous carbons via template-assisted
method by using zeolite 10X as the template. They reported at higher pyrolysis temperature during the
carbonization process, yield in certain graphitic domains and hollow-shell morphologies, however at around 700
°C, the carbons were nongraphitic. The synthesized carbons exhibit large BET surface areas in 1300−3331 m2/g
range and pore volumes up to 1.94 cm3/g with a pore size centered at 1.2 nm. They also reported that at 77 K
and 20 bar, 6.1 wt.% of hydrogen uptake had been obtained on carbon with the surface area up to 3331 m 2/g.

Table 3: Hydrogen uptake capacity on zeolite templated-carbon.


Zeolite template Carbon precursor Synthesis condition Temperature & H2 uptake Ref.
Pressure (wt.%)
Zeolite Y Polyfurfuryl alcohol Pyrolysis at 1323 K 77 K, 2.4 [14]
(PFA) 10 bar
Zeolite Y Sucrose Pyrolysis at 1100 K 77 K, 2.4 [15]
10 bar
Zeolite Y Acetonitrile Vary CVD temperature 77 K, 4.3 [16]
20 bar
Zeolite Y Acetonitrile Vary CVD temperature, 77 K, 6.1 [16]
KOH-activation 20 bar
Zeolitic imidazolate Furfuryl alcohol Vary carbonization 77 K, 3.1 [17]
framework (ZIF) temperature 20 bar
Zeolitic imidazolate Furfuryl alcohol Vary carbonization 77 K, 6.2 [17]
framework (ZIF) temperature, 20 bar
KOH-activation
NH4Y Furfuryl alcohol Vary carbonization 100 °C, 0.29 [4]
temperature & dwelling time atm pressure
Zeolite 13X Furfuryl alcohol Heating ramp rate 77 K, 20 bar 7.3 [18]
Zeolite USY Sucrose Vary zeolite/sucrose mass 77 K, 20 bar 3.65 [3]
ratios & temperature
Zeolite 10X Sucrose Vary pyrolysis temperature 77 K, 20 bar 6.1 [13]

According to Cai [3] the pore structure of carbons are remarkably influenced by the synthesis conditions.
The chemical activation along with the proper tailoring of the synthesis parameter can enhanced the hydrogen
uptake capacity. Table 3 shows that zeolite template-carbon that undergo chemical activation show a significant
hydrogen uptake up to 6 wt.%, ZIF-templated carbon increase from 3.1 to 6.2 wt.% after the activation and
KOH-activated zeolite Y templated-carbon obtained 6.1 wt.% compared. In addition, hydrogen uptake at 20 bar
show higher uptake compared at 10 bar which prove that hydrogen storage at higher pressure is improved by
elevating hydrogen pressure. Other than that, hydrogen uptake capacity is highly dependent on the surface area
and micropore volume of carbons related to the optimum pore size [3, 13, 15-18]. Furthermore, the presence of
zeolite in carbonization process, which will inhibit graphitic degrees to some extent, along with the increased in
carbonization temperature can enhanced the surface area and total pore volume of carbons and also improve
hydrogen storage capacities. In a nutshell, the above literatures show that the template-assisted method show
promising potential in developing material that will fit the DOE’s requirements.

Clay-based materials:
Clays have been widely used in adsorption and catalytic fields, therefore clays are another interesting
material to investigate their performance in hydrogen adsorption. Clays possess numerous advantages;
chemically inert, resistant to deterioration, low-cost and commercially available in large quantities. Besides that,
clays are highly porous materials whose surface area ranges between 3×10 5 and 7×105 m2/kg, clays operate in a
6 Nikmah Utami Sawai and Farah W. Harun, 2015
Journal of Industrial Engineering Research, 1(5) Special 2015, Pages: 1-7

reversible hydration±dehydration mode and can be separated easily in the reaction system for reuse [19,20]. To
the best of our knowledge, there are very scarce studies on clay-based materials as hydrogen storage candidate.
Previously, Campos [21] reported at 77K and atmospheric pressure, the hydrogen uptake capacity for
untreated montmorillonite clay is 0.12 wt.% while for the clay intercalated with silica nanoparticles is 0.40
wt.%. The insertion of SiO2 nanoparticles is to separate clay layer in order to increase the surface area of the
clay. According to the XRD result, the structure of the clay after the intercalation is changed and the bilayers of
clay are separated from each other by 12 nm, which is the diameter of the SiO 2 particles. The hydrogen
adsorption capacity which is directly proportional to the surface area is due to the mesoporosity that developed.
Gil [22] who studied the hydrogen adsorption capacity of various alumina-pillared clays at 77 K and
pressure of 0.7 bar shows that there is a linear relationship between the hydrogen uptake and the microporous
volume of the pillared clays. In this study, the clay was prepared by intecalating Al-polycations at different
Al/clay ratios and calcined at various temperatures in order to tune the interlayer space. The presence of
ultramicropores is determined as contributing factor to the strong degree of heterogeneity for hydrogen
adsorption.
H2 (g) adsorption on Na synthetic montmorillonite-type clays and Callovo-Oxfordian (COx) clay rock using
gas chromatography was reported by Didier [23]. Synthetic montmorillonites with increasing structural Fe(III)
substitution (0 wt.%, 3.2 wt.%, and 6.4 wt.% Fe) were used. Fe in the synthetic montmorillonites is principally
present as structural Fe(III) ions. Experiments were performed with dry clay samples which were reacted with
hydrogen gas at 90 and 120 °C for 30 to 45 days at a hydrogen partial pressure close to 0.45 bar. The result
indicated that up to 0.11 wt.% of hydrogen is adsorbed on the clays at 90 °C under 0.45 bar of relative pressure
[23].
Mondelli [24] performed the adsorption of hydrogen on sodium montmorillonite (Na-Mt) at 363 K and high
pressures (up to 90 bar). Na-Mt is mainly mesoporous with average pore width of 150 A° and specific areas in
the range 90-120 m2/g. High pressure volumetric measurements showed that hydrogen adsorption at 363 K
saturated between 40 and 60 bar, reaching 0.2 ± 0.02 wt.% at the plateau.
Ruiz-García [25] investigated the adsorption capacities for a carbon-clay nanocomposite which has been
prepared by using two types of clay as supports, montmorillonite and sepiolite, and caramel from sucrose. They
reported, at 298 K and 20 MPa, the clay-graphene nanomaterials achieved over 0.1 wt.% and 0.4 wt.% of
hydrogen adsorption excess related to the total mass of the system and specifically related to the carbon mass,
respectively. The very high isosteric heat for hydrogen sorption determined from adsorption isotherms at
different temperatures (14.5 kJ/mol) fit well with the theoretical values available for hydrogen storage on
materials that show a strong stabilization of the H2 molecule upon adsorption.
Table 4: Hydrogen uptake capacity on clay-based sample.
Clay Temperature & Pressure H2 uptake Ref.
(wt.%)
Montmorillonite (untreated) 77 K, 1 atm 0.12 [21]
Montmorillonite (SiO2-intercalated) 77 K, 1 atm 0.40 [21]
Montmorillonite (Fe3+) 363 K, 0.45 bar 0.11 [23]
Montmorillonite (Na) 363 K, 90 bar 0.20 [24]
Montmorillonite(graphene) 298 K, 20 MPa 0.20 [25]
Sepiolite (graphene) 298 K, 20 MPa 0.40 [25]

Table 4 summarizes the performance of clay-based material on storing hydrogen. Based on the above
literatures, we can see that clay-based material show promising potential candidate for storing hydrogen.
However, the uptake capacity is very low, ranging over 0.1 wt.% and 0.4 wt.% compared to the activated carbon
and zeolite templated-carbon material which could achieved up to 7.0 wt.%. Hence, the textural properties of
clay-based material must be enhanced in order to achieve a remarkable performance of hydrogen uptake. The
intercalation of nanoparticle such as SiO2 could enhance the interlayer space as reported by Campos [25].
Besides that, pillaring of clay may also contribute to the increase between the interlayer space and the presence
of ultramicropore in clay sample as discussed by Gil [26]. The enhancement in the textural properties of clay
sample show higher hydrogen uptake [25, 26].

Conclusion:
Present review discussed on the potential of the agricultural-based activated carbons, zeolite templated-
carbons and clay-based materials as the sorbent material for hydrogen. All the samples synthesized in the
literatures showing a high adsorption capacity of hydrogen on the materials. However, many current studies that
show remarkable hydrogen uptake was investigated at cryogenic condition. Hence, in future, efforts should be
emphasized on storing hydrogen at ambient condition. In addition, eventhough the uptake capacity is
remarkable, nevertheless the materials still did not fit the DOE’s requirement for on-board hydrogen storage
application. Therefore, future studies should exploit the existing data of hydrogen storage material from the
literature as a guideline for future development of ideal hydrogen storage materials.
7 Nikmah Utami Sawai and Farah W. Harun, 2015
Journal of Industrial Engineering Research, 1(5) Special 2015, Pages: 1-7

REFERENCES

[1] Akasaka, H., T. Takahata, I. Toda, H. Ono, S. Ohshio, S. Himeno, T. Kokubu, H. Saitoh, 2011. Hydrogen
storage ability of porous carbon material fabricated from coffee bean wastes. International Journal of
Hydrogen Energy, 36(1): 580-585.
[2] Almasoudi, A., R. Mokaya, 2012. Preparation and hydrogen storage capacity of templated and activated
carbons nanocast from commercially available zeolitic imidazolate framework. Journal of Materials
Chemistry, 22(1), 146. http://doi.org/10.1039/c1jm13314d.
[3] Cai, J., L. Li, X. Lv, C. Yang, X. Zhao, 2014. Large surface area ordered porous carbons via nanocasting
zeolite 10x and high performance for hydrogen storage application. ACS Applied Materials and Interfaces,
6(1): 167-175.
[4] Cai, J., M. Yang, Y. Xing, X. Zhao, 2014. Large surface area sucrose-based carbons via template-assisted
routes: Preparation, microstructure, and hydrogen adsorption properties. Colloids and Surfaces A:
Physicochemical and Engineering Aspects, 444: 240-245.
[5] Campos, F., L. de la Torre, M. Román, A. García, A.A. Elguézabal, 2008. Montmorillonite clay
intercalated with nanoparticles for hydrogen storage. Ceramic Processing Research, 9(5): 482-485.
[6] Chen, H., H. Wang, Z. Xue, L. Yang, Y. Xiao, M. Zheng, B. Lei, Y. Liu, L. Sun, 2012. High hydrogen
storage capacity of rice hull based porous carbon. International Journal of Hydrogen Energy, 37(1): 18888-
18894.
[7] Cheng, F., J. Liang, J. Zhao, Z. Tao, J. Chen, 2008. Biomass waste-derived microporous carbons with
controlled texture and enhanced hydrogen uptake. Chemistry of Materials, 20(5): 1889-1895.
[8] Didier, M., L. Leone, J. Greneche, E. Giffaut, L. Charlet, 2012. Adsorption of Hydrogen Gas and Redox
Processes in Clays. Environmental Science & Technology, 46(1): 3574-3579.
[9] Gil, A., R. Trujillano, M.A. Vicente, S.A. Korili, 2009. Hydrogen adsorption by microporous materials
based on alumina-pillared clays. International Journal of Hydrogen Energy, 34(20): 8611-8615.
[10] Guan, C., K. Wang, C. Yang, X.S. Zhao, 2009. Characterization of a zeolite-templated carbon for H2
storage application. Microporous and Mesoporous Materials, 118: 503-507.
[11] Guan, C., X. Zhang, K. Wang, C. Yang, 2009. Investigation of H2 storage in a templated carbon derived
from zeolite Y and PFA. Separation and Purification Technology, 66(3): 565-569.
[12] Konwar, R.J., M. De, 2013. Effects of synthesis parameters on zeolite templated carbon for hydrogen
storage application. Microporous and Mesoporous Materials, 175: 16-24.
[13] Liu, J., G. Zhang, 2014. Recent advances in synthesis and applications of clay-based photocatalysts : a
review. Physical Chemistry Chemical Physics : PCCP, 16(1): 8178-8192.
[14] Masika, E., R. Mokaya, 2013. Preparation of ultrah igh surface area porous carbons templated using
zeolite 13X for enhanced hydrogen storage. Progress in Natural Science: Materials International, 23(3):
308-316.
[15] Minoda, A., S. Oshima, H. Iki, E. Akiba, 2013. Synthesis of KOH-activated porous carbon materials and
study of hydrogen adsorption. Journal of Alloys and Compounds, 580: S301-S304.
[16] Mondelli, C., F. Bardelli, J.G. Vitillo, M. Didier, J. Brendle, D.R. Cavicchia, L. Charlet, 2015. Hydrogen
adsorption and diffusion in synthetic Na-montmorillonites at high pressures and temperature. International
Journal of Hydrogen Energy, 40(6): 2698-2709.
[17] Ruiz-García, C., J. Pérez-Carvajal, A. Berenguer-Murcia, M. Darder, P. Aranda, D. Cazorla-Amorós,
Ruiz- E. Hitzky, 2013. Clay-supported graphene materials: application to hydrogen storage. Physical
Chemistry Chemical Physics : PCCP, 15(42): 18635-41.
[18] Sadek, O.M., W.K. Mekhamer, 2000. Ca-montmorillonite clay as thermal energy storage material.
Thermochimica Acta, 363(1-2): 47-54.
[19] Sevilla, M., N. Alam, R. Mokaya, 2010. Enhancement of hydrogen storage capacity of zeolite-templated
carbons by chemical activation. Journal of Physical Chemistry C, 114(25): 11314-11319.
[20] Sharma, S., S. Krishna, 2015. Hydrogen the future transportation fuel : From production to applications.
Renewable and Sustainable Energy Reviews, 43: 1151-1158.
[21] Wang, Y.X., B.S. Liu, C. Zheng, 2010. Preparation and adsorption properties of corncob-derived activated
carbon with high surface area. Journal of Chemical and Engineering Data, 55(11): 4669-4676.
[22] Wang, Y.X., B.S. Liu, C. Zheng, 2010. Preparation and adsorption properties of corncob-derived activated
carbon with high surface area. Journal of Chemical and Engineering Data, 55(11): 4669-4676.
[23] Xia, Y., Z. Yang, Y. Zhu, 2013. Porous carbon-based materials for hydrogen storage: advancement and
challenges. Journal of Materials Chemistry A, 1(33): 9365. http://doi.org/10.1039/c3ta10583k.
[24] Xiao, Y., H. Dong, C. Long, M. Zheng, B. Lei, H. Zhang, Y. Liu, 2014. Melaleuca bark based porous
carbons for hydrogen storage. International Journal of Hydrogen Energy, 39: 11661-11667.
[25] Zhou, L., Y. Zhou, Y. Sun, 2004. Enhanced storage of hydrogen at the temperature of liquid nitrogen.
International Journal of Hydrogen Energy, 29(3): 319-322.

Вам также может понравиться