Sunteți pe pagina 1din 28

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/303091601

Research on the theory and application of


adsorbed natural gas used in new energy
vehicles: A review
Article in Frontiers of Mechanical Engineering May 2016
DOI: 10.1007/s11465-016-0381-2

READS

40

3 authors, including:
Zhengwei Nie

Yuyi Lin

University of Missouri

University of Missouri

6 PUBLICATIONS 0 CITATIONS

17 PUBLICATIONS 67 CITATIONS

SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

SEE PROFILE

Available from: Zhengwei Nie


Retrieved on: 12 August 2016

Research on the theory and application of adsorbed natural gas used in new
energy vehicles: A review
Zhengwei Nie1, Yuyi Lin1*, Xiaoyi Jin2
1

Department of Mechanical & Aerospace Engineering, University of Missouri, Columbia, MO


65211 USA

College of Mechanical Engineering, Shanghai University of Engineering Science, Shanghai,


Shanghai 201620 China
*Corresponding author. Tel.: +1 573 882 7505 Fax: +1 573 884 5090
Email address: LinY@missouri.edu (Yuyi Lin)

Abstract
Natural gas (NG), its primary constituent is methane, has been considered a convincing alternative
for the growth of the energy supply worldwide and intensified interest by recent developments.
Adsorbed natural gas (ANG), the most promising methane storage method, has been an active field
of study in the past decade. ANG constitutes a safe and low-cost way to store methane for natural
gas vehicles at an acceptable energy density while working at substantially low pressures (3.5-4.0
MPa), allowing for conformable store tank. This work serves to review the state-of-the-art
development reported in the scientific literature on adsorbents, adsorption theories, ANG
conformable tanks, and related technologies on ANG vehicles. Patent literature has also been
searched and discussed. The review aims at illustrating both achievements and problems of the
ANG technologies based vehicles, as well as forecasting the development trends and critical issues
to be resolved of these technologies.

Keywords: adsorbed natural gas (ANG); adsorbent; adsorption theory; conformable tank; natural
gas vehicles (NGVs)

1. Introduction
With the development and progress of human society, the fossil fuels such as coal and petroleum
are consumed drastically. Exploiting new and effective alternative energy sources is urgently
required. In addition, emission control calls for using clean fuels for vehicles. Accordingly, people
has been searching for suitable alternatives without pollution since last century. The abundant
reserves of natural gas (methane is the primary constituent), estimated to be about 24% of the total
energy resource [1], has been considered a convincing alternative for the growth of the energy
supply worldwide and intensified interest by recent developments. Figure 1 shows that the natural
gas production from shale formations has increased by 37% from 2002 to 2012 in United States
[2]. The largely available natural gas in shale formations has increased its supplies and decreased
its prices. With regard to environmental benefits, natural gas has a relatively complete clean
combustion compared to other fossil fuels. It produces about 29% and 44% less carbon dioxide
1

per joule delivered than oil and coal respectively, and lower amount of pollutants than other
hydrocarbon fuels [3]. In view of this, natural gas vehicles (NGVs) have received significant
interest since 1990s. In recent years the natural gas vehicles represent a cost-competitive, loweremission alternative to the gasoline-fueled vehicles all over the world.

Figure 1 U.S. natural gas production by source, 1990-2040 (This figure was created using the data
provided by the annual energy outlook 2014 of U.S. Energy Information Administration [2].)

However, it is difficult to store natural gas based vehicle because of its low density and low
volumetric energy density. At normal temperature and pressure the volumetric energy density, and
as well as the mileage per unit volume of the fuel tank, of natural gas is only 0.12% of that of
gasoline [4]. Thus, the application of natural gas on vehicle depends on its storage method and
storage amount in given on-board fuel tank. Three different methods are currently applied for onboard natural gas storage: liquefied natural gas (LNG), compressed natural gas (CNG), and
adsorbed natural gas (ANG). In the case of LNG, the storage is under a low temperature of
approximately -165.5 C. The disadvantages of LNG, such as special insulated vessels requiring
periodic venting, the danger of refueling cryogenic liquids, and the high cost of liquefaction
storage and the refueling facilities, make it only suitable for use on a large scale.
CNG is natural gas stored as a compressed supercritical fluid at room temperature and in the realm
of about 200-250 bar (~2900-3600 psi). The volumetric energy density of natural gas can reach
26% of that of gasoline when compressed at 250 bar, which leads to a lower driving mileage per
unit volume of the fuel tank [5]. Since CNG method is a high pressure application, it requires an
expensive multi-stage compressor that accounts for nearly 50% of the capital investment of
refueling station [4]. Moreover, such a multi-stage compressor consumes substantial energy per
unit of gas compressed, which increases the cost of single refueling. Further, bulky and heavy fuel
tank occupies trunk or cargo space and reduces passenger space. Hence, CNG has its own
limitations for commercialization at a large scale worldwide, especially private cars.
With the process of studying the LNG and CNG technologies, a different method based on
adsorbed natural gas (ANG) has been an active field of study in the past decade. In ANG system,
natural gas is stored in porous materials (adsorbents) as an adsorbed phase at a pressure 35-40 bar
(~500-580 psi). The energy density provided ANG under 500 psi can be equivalent to that of CNG
under 2400 psi [6, 7]. 80% of volumetric energy density of CNG under 3000 psi can be obtained
by ANG under 500 psi. Thus, the maximum volumetric energy density as well as a mileage per
2

unit fuel tank of ANG has achieved about 20% of that of gasoline through some of the best porous
materials currently [4]. The lower operation pressure of ANG means the lower cost of refueling
station and also less of a safety hazard than the higher pressures used for CNG. Moreover, under
low pressure on-board fuel tank could be designed into a conformable shape in order to use the
available space of vehicle without interference to passenger space.
In recent years, the search for a more efficient porous adsorbent in order to improve ANG storage
efficiency and lower the cost to the end user of ANG vehicle has become a hotspot of research. In
addition, some other technologies related to the commercial applications of ANG vehicle, such as
adsorbent compacting technology and machine, and ANG home-fueling compressor, have
attracted much attention than before. Therefore, the researches of adsorbed natural gas used in new
energy vehicles, including adsorbents, adsorption theories, ANG on-board tanks, and homefueling system, represents the technical and engineering design frontiers. These researches would
improve home-fueling ANG vehicles commercialization at a large scale.
This paper serves to review the state-of-the-art development reported in the scientific literature on
adsorbent, adsorption theory, ANG conformable tank, and related technologies on ANG vehicles.
Patent literature has also been searched and discussed. The review aims at illustrating both
achievements and problems of the ANG technologies based vehicles, as well as forecasting the
development trends and critical issues to be resolved of these technologies.
2. Adsorbents
Developing a dedicated and efficient adsorbent is the key of ANG technology. It is widely accepted
that the excellent gas adsorbent should have the following characteristics: (1) Adsorbent should
have a larger specific surface area and appropriate pore structure. The specific surface area of
common adsorbents is about 1,000 m2/g, but that of high quality adsorbents ranges from 2,000
m2/g to 3,000 m2/g [8-11]. (2) Pore size distribution should be centralized with pore sizes ranging
from 1.0 nm to 2.0 nm [8]. (3) Micropore volumes should account for more than 85% of total pore
volumes [12]. (4) The storage capacity of adsorbent for natural gas should be more than 100 V/V
(i.e. liters of gas stored per liter of storage vessel internal volume under NTP conditions) for
immobile storage and more than 150 V/V for mobile storage at 3.5 MPa [13]. Specifically, the
United States Department of Energy (DOE) established a new target for volumetric methane
adsorption from old value of 180 V/V to the much ambitious value of 263 V/V [14, 15]. (5) The
rate of adsorption and desorption should be high under normal conditions. (6) Adsorbent has a
long service life and can be regenerated. Its preparation process is simple and low cost. Several
adsorbents have been developed since 1950s, such as activated carbon, carbon nanotubes,
graphene, and metal organic frameworks.
2.1 Activated carbon
Activated carbon (AC) is an artificial and renewable carbon products with highly developed porous
structure and huge internal surface area. It is made of charcoal, wood, coconut shell, some kinds
of stone pulp slurry and other agricultural and forestry products, coal and heavy oil through the
process of carbonization and activation. Adsorption is one of the most significant characteristics
of the activated carbon, which can adsorb a variety of substances from the gas phase or liquid
3

phase. Moreover, activated carbon possesses a large adsorption capacity, whose pore volume is up
to 0.2-1.0 cm3/g and surface area is from 400 m2/g to 3000 m2/g [12]. More importantly, the
multiple functional groups on the surface of activated carbon make its performance stable and
make it usable in conditions of different temperatures and pH values. In order to meet the needs of
high performance adsorbents, a range of efficient adsorbents with different properties have been
produced. The methane capacity of various activated carbons (granular powder) at 500 psi and 298
K reported in the literature are tabulated in Table 1 [16-20].
Table 1 Methane adsorption capacity of activated carbons (granular powder) at 298 K and 500 psi
reported in the literature
Institution

Raw Materials

Activator

Specific
Surface Area

Petroleum coke,

Harry Mash, UK

Charcoal

KOH

(m2/g)
2700

Amoco, US[16]

Activated carbon

KOH

3000

University of Missouri-Columbia, US[18-20]

Activated carbon

KOH

3500

Kawasaki Steel, Japan[17]

Carbon microspheres

KOH

2145

University of Petroleum of China, China

Lignin, Petroleum coke

2912, 2399

Shanxi Institute of Coal Chemistry, China

Petroleum coke

3882

South China University of Technology, China

PVC

3191

Activator

2966

Beijing University of Chemical Technology,


China

As shown in Table 1, most of the granular carbon powders have specific surface areas more than
3,000 m2/g. Shanxi Institute of Coal Chemistry, China has developed an activated carbon with
3,882 m2/g specific surface area.
Recently, Mirian Elizabeth Casco et al. proved experimentally that properly designed activated
carbon materials can really realize the new DOE value for methane adsorption [15]. They
estimated the storage capacity for different activated carbons using excess adsorption isotherms.
The results are shown in Figure 2, which also compares the storage adsorption for the different
activated carbons and for HKUST-1 (HKUST-1 is the most promising MOF; MOF is discussed in
next section.). As shown in Figure 2, both LMA726 and LMA738, activated carbons prepared by
chemical activation with anhydrous KOH, can achieve the new DOE value of 263 V/V, although
at a slightly higher pressure than that of HKUST-1. At the same time, they also predicted the
methane storage capacity for the most promising activated carbons (LMA405 (activated carbons
4

prepared by olive stones) and LMA738) at 298 K and up to 20 MPa as shown in Figure 3 [15].
Under high pressures, the additional incorporation of exceptional adsorbents to the storage tank
could be a tremendous advantage in terms of storage capacity with a 50% increase (storage
capacity at 20 MPa for sample LMA738 as high as 328 V/V) compared to conventional CNG.

Figure 2 Methane storage capacity for different activated carbons at 298 K and up to 10 MPa [15]

Figure 3 Methane storage capacity for LMA405 and LMA738 at 298 K and up to 20 MPa [15]

2.2 Carbon nanotubes and graphene


Carbon nanotubes (CNTs) are relatively new adsorbents and hold interesting positions in carbonbased nanoparticles. CNTs have been attracting considerable interest as natural gas storage media
because of their unique properties, including uniform porosity, electrical conductivity, high tensile
strength and relative inertness for physical adsorption. Figure 4 shows the structures of singlewalled carbon nanotubes (SWCNTs) and multiwalled carbon nanotubes (MWCNTs) [21].
Bekyarova et al. produced nanostructured disordered carbon, which exhibits a high methane
storage capacity, reaching 160 V/V of nanocarbon at 3.5 MPa and 303 K [22]. H. Tanaka et al.
calculated the methane adsorption on isolated SWCNTs at 303 K over a range of pore sizes and
pressures by the nonlocal density functional theory (DFT). They found that, on a per weight basis,
total excess adsorption of an isolated SWNT exceeds that of the idealized carbon slit pore with the
same size [23]. Cao et al. investigated the methane adsorption on triangular arrays of SWCNTs by
the grand canonical Monte Carlo (GCMC) method at room temperature. It is reported that the total
volumetric of methane on the SWCNT arrays with the optimal parameters reach 216 V/V, which
greatly exceeds the old target (180 V/V) of the DOE (USA) [24]. Masoud D.G. et al. indicated that
the methane molecule is preferentially adsorbed onto the CNT with a binding energy of -2.84 kcal
5

/mol compared to boron nitride nanotubes (BNNTs). They also found that methane adsorptive
capacity increases for wider CNTs and decreases for wider BNNTs [25]. Jae-Wook Lee et al.
measured the adsorption equilibria of methane on MWCNT and made a comparison with two
zeolites: DAY and HSZ-320. The isotherm data were attained by a static volumetric method at
303.15 K-323.15 K and 3 MPa as shown in Figure 5 [26].

Figure 4 (Super) structure representations of (a) a MWCNT and (b) a SWCNT [21]

Figure 5 (a) Validation of the hybrid isotherm model for the adsorption of methane on MWCNT at
303.15 K; (b) Adsorption isotherms of methane on MWCNT: circle, 303.15 K; square, 313.15 K; triangle,
323.15 K; solid line, hybrid Langmuir-Sips isotherm [26]

Graphene, a one-atom-thick planar sheet of sp2-bonded carbon atoms densely packed in a


honeycomb crystal lattice, has also received significant attention in recent years due to its
exceptional properties, such as high surface area and good gas adsorption property. The four
different high-symmetry adsorption sites of methane on graphene are shown in Figure 6 [27]. The
methane molecule may take up one of the four different sites shown with numbers in Figure 6,
with the hydrogen tripod either directed to the surface or to the vacuum. I. Carrillo et al. reported
that the adsorption of methane at 300 K and atmospheric pressure through DFT and molecular
dynamics and found that the adsorption on graphene can be improved by modifying graphene layer
with titanium [28]. C. Thierfelder et al. studied the methane-graphene interaction using DFT. The
6

adsorption energy is 0.17 eV and the molecular distance is 3.28 [27]. Brandon C. Wood et al.
indicated that polar groups, such as COOH, NH2, NO2, and H2PO3, are promising candidates for
improving methane capacity by strengthening adsorption and activating exposed edges and
terraces to produce additional adsorption sites [29].

Figure 6 Four different high-symmetric adsorption sites of methane on graphene. All sites may be taken
up with either the H tripod oriented towards the substrate or towards the vacuum [27]

Hee K. Chae et al. presented a route to high surface area for graphene [30]. As for a graphene sheet
shown in Figure 7(a), exposing the latent edges of the six-membered rings results in significant
enhancement of specific surface area. Thus, when calculating two sides the surface area of a single
infinite sheet is 2,965 m2/g. For units consisting of infinite chains of poly-p-linked six-membered
rings shown in Figure 7(b), the surface area is 5,683 m2/g. Alternatively, if the graphene sheet is
divided into units of three six-membered rings that are 1,3,5-linked to a central ring shown in
Figure 7(c), the surface area reaches to 6,200 m2/g. Lastly, exposing all latent edges to give isolated
six-membered rings as shown in Figure 7(d) obtains an upper Flimit surface area of 7,745 m2/g.
As previously stated, the surface area of most of current activated carbons is around 3,000 m2/g.
Therefore, in this basis, the specific surface area of the carbon materials still has a huge room for
improvement. It is reported that the activated carbons prepared by the University of Missouri have
contained the similar structure to the aforementioned graphene sheet as shown in Figure 8 [31].

Figure 7 The surface area of a graphene sheet, (a) a graphene sheet has a Connolly surface area of 2,965
m2/g; (b) a series of poly-p-linked six-membered rings has the surface area of 5,683 m2/g; (c) Excision of
7

Six-membered rings 1,3,5-linked to a central ring increases the surface are to 6,200 m2/g; (d) the surface
area reaches a maximum of 7,745 m2/g when the graphene sheet is fully decomposed into isolated sixmembered rings [30]

Figure 8 The microstructure of activated carbons prepared by University of Missouri [31]

2.3 Metal-Organic Frameworks (MOFs)


MOFs are porous, crystalline materials that comprise metal or metal oxide nodes connected by
organic linker compounds [32]. Figure 9 shows some of the most studied MOFs [33]. Large
internal surface areas of MOFs make them attractive for methane adsorption. More importantly,
MOFs can be prepared in a modular fashion and are thus tunable: one can assemble different
combinations of organic linkers and metal nodes to obtain a large variety of unique materials with
different affinities for relevant gases. MOFs with surface area of more than 6,000 m2/g have been
produced and have also performed gas adsorption [34, 35].

Figure 9 Three common metal-organic frameworks studied for their gas storage properties: IRMOF-1
(left), MOF-177 (middle), and HKUST-1 (right) [33]

For the adsorption simulation of methane on MOFs, it is often reasonable to assume that the
methane molecules and MOF structure are completely rigid. Under this assumption, the only
intermolecular interactions between nonbonded atoms are modeled by a Lennar-Jones (LJ) +
Coulomb potential [36-40].
8

12 6
qq
ij
ij
+ i j 2
pij (rij )= 4eij
-

rij rij 4 0 rij

(1)

where rij is the distance between atoms i and j with charges of qi and q j , respectively. and
are two Lennard-Jones parameters.
There have been many computational studies on exploring desired characteristics of an optimal
adsorbent for methane adsorption. Dren et al. studied methane adsorption in 18 different materials,
including isoreticular MOFs (IRMOFs) and carbon nanotubes. They indicated new MOF materials
by replacing or adding atoms in the linker molecules of IRMOF-1 and predicted isotherm for
IRMOF-992 indeed showed a higher volumetric methane adsorption under the whole pressure
range with an average isosteric heat of adsorption. Moreover, IRMOF-993 was even predicted to
exceed the DOE old target of 180 V/V [41]. Dren and Snurr investigated methane adsorption in
a series of IRMOFs (IRMOF-1, -8, -10, -14, and -16) to study the effect of the organic linker and
found that the methane adsorption increases with increasing number of carbon atoms on the linker
when pressure is less than 40 bar [42].
Gallo and Glossman-Mitnik calculated adsorption isotherms for methane in IRMOF-1 and MOF177 (two large-surface-area MOFs), Zn-MOF-74 (with open metal sites), IRMOF-11 and MOF14 (two catenated MOFs) at pressures up to 80 bar at 298 K. The results proved that the two largesurface area MOFs showed smaller gravimetric methane adsorption than the two catenated MOFs
at low pressures (around 10 bar). At higher pressures, however, the two large-surface-area MOFs
showed larger gravimetric methane adsorption than the catenated MOFs. Although Zn-MOF-74
had small gravimetric methane adsorption because of its high crystal density, it showed the highest
volumetric methane adsorption of 170 V/V at 35 bar and 298 K, which is close to the old target
180 V/V of DOE for practical methane storage on vehicles [43].
As discussed in Sec. 2.1, HKUST-1, the most promising MOF, is able to achieve both the DOEs
old target for methane adsorption at 3.5 MPa, with a storage capacity of 225 V/V, and the DOEs
new target at 6.5 MPa, with a storage capacity of 276 V/V. However, the thermal stability of MOFs
is poor. The crystal structure of most MOFs would collapse under heat treatment with more than
350 . Also, MOFs have a low mechanical stability under pressure since their structure is
susceptible to environmental factors such as water vapor, preservation method and time. Moreover,
the synthesis of MOFs needs hydrothermal method with lots of organic solvents and subsequent
processing. Thus, the cost of MOFs is relatively high [44]. More importantly, according to the
annual report 2014 of US MOVE project, the adsorption capacity of MOFs decreased by 50% after
20 adsorption-desorption cycles [45]. These deficiencies greatly limit the application of MOFs on
the methane adsorption.

2.4 Other adsorbent


Early work on porous materials for ANG storage systems started with zeolites. Zeolites are
crystalline aluminosilicate, whose both positive ions and negatively charged silicon alumina
9

skeleton have polarity. Positive ions give a strong positive field attracting the negative centers of
polar molecules, or polarizing molecules by electrostatic induction polarization. The stronger
polarity of molecules are, the more easily they are adsorbed by zeolite. Adsorption of methane on
zeolites were considered as a potential means for increasing the on-board fuel storage capacity of
vehicles before the high surface area carbon materials were produced. Currently, it has been
realized that a major constraint associated with zeolites over carbons is that the specific surface
area (accessible to gas/vapor adsorbates) more than 1000 m2/g is unattainable even with synthetic
zeolites due to its structural limitations. In addition, zeolites are extremely hydrophilic and can
lose their adsorption capacity for methane with time because of preferential moisture adsorption.
Moreover, significant macropore volumes in the form of interparticle voids is associated with
zeolite packing. This can be avoided only if packing is done with a single crystal which is
impossible. Increasing packing density would lower micropore surface areas, e.g. the specific
surface area of 5A CaNaA Zeolite is 394 m2/g [46]. The methane capacity of various small and
large pore zeolites at 298 K and 35 bar compiled from the adsorption isotherms reported in the
literature is tabulated in Table 2 [46-50].
Table 2 Methane adsorption capacity of zeolites at 298 K and 35 bar
Type
5A
13X
Na-ZSM-5
Silicalite
NaX
MgX
CaX
SrX
BaX
NaY
MgY
CaY
SrY
BaY

Methane capacity (mg/g) Reference


50
[47]
53
[47]
31
[48]
44
[49]
66
[46]
66
[46]
82
[46]
62
[46]
67
[46]
56
[50]
62
[50]
63
[50]
61
[50]
54
[50]

2.5 Adsorbents monolith


From the foregoing, adsorbents of ANG storage technology must have high storage capacity for
methane. It is reported that the highest storage capacity for ANG storage is carbon materials.
However, most researchers believe that the limits of the maximum surface area for carbon
materials have already been reached and any further improvements by another order of magnitude
seems impossible. Thus, the current research interest is how to increase the monolith density and
retain the high surface area, especially in the activated carbon area. On the other hand, the
volumetric methane adsorption of carbons decreases due to low monolith density. Such
relationship is illustrated in Figure 10 [4]. Furthermore, it is very necessary to increase the
monolith density and then increase ANG storage capacity since the storage volume available on
ANG vehicle is limited.

10

Figure 10 Experimental volumetric methane adsorption capacity of several adsorbents at 298 K and 35
bar, versus gravimetric surface area [4]

The studies on increasing the adsorbent monolith density so far involve efforts towards optimizing
the particle sizes, bulk density, and monolith preparation procedure. Commonly used carbon
compacting methods include compacting with binder, compacting without binder, and extrusion
etc. Significant developments have been made in improving monolith density by modifying the
process of its fabrication. Quinn et al. patented a carbonization/activation procedure for the
preparation of carbon monolith from polymer precursor with minimum marcopore or void volumes.
The methane adsorption of reported carbons is 178 V/V at 298 K and 35 bar (surface areas upto
1415 m2/cc and density upto 0.86 g/cc) [51]. Figure 11 shows a comparison of achievements in
different ANG projects all over the world [52]. As shown in Figure 11, the University of Missouri
has increased the methane storage capacity of activated carbon to 202 V/V, which exceeds the old
target 180 V/V of DOE. The methane storage capacity of activated carbon monolith, however
decreased by 20.3%, only 161 V/V. Therefore, developing optimum compacting procedure for
carbon monolith remains the focus of future research.

11

Figure 11 Comparison of achievements in different ANG projects all over the world [52]

3. Adsorption Theory
The word adsorption is defined as the enrichment of one or more components in an interfacial
layer. Adsorption occurs whenever a solid surface is exposed to a gas (gas-solid interaction) or
liquid (liquid-solid interaction) [53]. There are two types of adsorption: (I) chemisorption
adsorption involves chemical reaction between the surface and the adsorbate. New chemical bonds
are generated at the adsorbent surface. The strong interaction between the adsorbate and the
adsorbent surface creates new type of electronic bonds. Binding energy (BE) typically of several
eV. The forces involved are valence forces of the same kind as those operating in the formation of
chemical compound. (II) physisorption (also known as Van der Waals adsorption) adsorption
occurs without chemical bonding. The electronic structure of atoms or molecules remain partially
unchanged. Typical BE are 10-100 meV. The force involves is Van der Waal attractive force
between adsorbate and adsorbent, and the force of attraction between the adsorbate and adsorbent
are very weak, which means this type of adsorption can be easily reversed by heating or by
decreasing the pressure [54-56]. Therefore, it is suitable for ANG purpose.
In the matter of theory of adsorption, in 1918 Langmuir derived for the first time a clear concept
of monolayer adsorption, formed on energetically homogeneous solid surfaces kinetic studies. The
statement proposed by Langmuir applied to chemisorption and with some restrictions, to
physisorption [57]. The milestone towards development of adsorption was the multilayer isotherm
equation proposed by Brunauer, Emmett and Teller in 1938. This theory, which is called BET
theory, for the first time, was successful in determining (by means of isotherm adsorption of six
different gases) the surface area of an iron synthetic ammonia catalyst. They also introduced the
point B method [58-60]. In 1940, Brunauer, Deming, Deming and Teller proposed a four
12

adjustable parameter equation, where the forces of capillary condensation were taken into account.
The so-called BDDT equation, contrarily to BET isotherm, can be applied over a wider range of
relative pressures. They also calculated out 5 principal types of adsorption isotherms for gases and
vapours, which are called BDDT adsorption isotherm [61, 62]. Based on the BDDT classification,
the International Union of Pure and Applied Chemistry (IUPAC) proposed an IUPAC
classification with six types of adsorption isotherms in 1985 as shown in Figure 12 [63-65].

Figure 12 The IUPAC classification of adsorption isotherms [65]

The reversible type I isotherm is a monolayer adsorption, which is mainly obtained with
microporous adsorbents. The reversible type II isotherm is mainly obtained with non-porous or
macroporous adsorbents. The reversible type III isotherm is a multilayer adsorption that mainly
depends on adsorbate-adsorbate interactions and has no flat portion in the curve, which indicates
the absence of monolayer formation. The type IV isotherm is mainly obtained from mesoporous
adsorbents, which shows the formation of a monolayer followed by a multilayer adsorption. The
Type V isotherm is uncommon; it is related to the Type III isotherm in that the adsorbent-adsorbate
interaction is weak, but it is obtained with certain porous adsorbents. The Type VI isotherm is a
multilayer adsorption on a uniform non-porous surface [54-56, 66]. The supercritical adsorption
isotherms presented here are type I isotherms.
Figure 13 shows the methane adsorption as a function of various pressures at two different
temperatures [13]. As presented in Figure 13, the behavior obtained is very similar, and the same
trend exists at the two temperatures. Based on above discussion, the adsorption is an exothermic
process, an increase in temperature results in a lower methane capacity under dynamic conditions.
Thus, an increase in temperature can reduce the amount of methane adsorption.

13

Figure 13 Methane adsorption as a function of pressure at two different temperatures [13]

As for simulation studies, Tan et al. used a combination of grand canonical Monte Carlo computer
simulation (GCMC) and nonlocal density functional theory to calculate methane adsorption in
model porous carbons for a wide range of pore sizes [67]. Theoretical studies on carbons by
Matranga et al. and Cracknell et al showed that an optimal micropore would have a wall separation
of about three to four methane diameters [68, 69]. Stella et al. indicated that a highly idealized
carbon composed of single graphitic layer 11.4 apart would have a methane adsorption of 209
V/V at 35 bar [70]. Myers et al. reported that the theoretical maximum capacity of carbons in
monolith form is 220 V/V. They also pointed out that the maximum capacity would decrease to
150 V/V in the case of pelletized carbon. Pelletized carbons provide a gaseous path for the
transport of gas to the surface of the pellets while compacting monoliths inside a storage tank lead
to a slow diffusion through the solid disks [71]. Furthermore, Basumatary et al. simulated the heat
and fluid flow inside the adsorption bed made of activated carbon using a volume averaging
technique and Darcy-Brinkman formulation and calculated the thermal conductivity of the
activated carbon-methane system using the modified Luikov Model. The thermal conductivity as
a function of temperature is shown in Figure 14 [72].

Figure 14 Thermal conductivity of adsorption bed as a function of temperature, nitrogen (star) and
methane (circle) [72]
14

The maximum methane adsorption capacity of carbons with different pore structures at 274 or 298
K and 35 bar obtained by simulation are listed in Table 3. It has to be kept in mind that the
experimental methane adsorption of compacted carbons havent attained the theoretical maximum
adsorption (220 V/V) indicated by simulation results.
Table 3 Maximum methane adsorption capacity of carbons at 35 bar and different temperatures
obtained by simulation
Adsorbent

Pore shape Theoretical methane capacity at 35 bar Pore size Reference


Slit
221 mg/g
12.6
Carbon (AX-21)
[70]
Triangular
157 mg/g
Carbon (274 K)
Slit
187 mg/g
11.43
[73]
Carbon (298 K)
Slit
139 mg/g
11.43
[74]
Slit
168 mg/g
Carbon (300 K)
[75]
Triangular
85 mg/g
Carbon (274 K)
Slit
166 mg/g
11.43
[68]

The common consensus based on simulation studies is that the ideal adsorbent with maximum
adsorption for at 298 K and 35 bar would have only micropores with size and shape according to
the optimal volumetric storage of two layers of methane molecules (7.6 ). However, the pore size
and shape should be in accord with the optimal volumetric storage of at least three layers of
methane molecules (11.4 ) when deliverability is also taken into account, which can maximize
deliverability by minimizing methane retention in the micropores at ambient pressure. Moreover,
the wall thickness of micropores in adsorbents should correspond to only one layer of the adsorbent
atoms (i.e., each adsorbent atom should be accessible to the adsorbate molecules). Macroporosity
or mesoporosity should be avoided due to the presence of non-utilized pore volumes in the
adsorbent by adsorbate. Such non-utilized pore volumes conduce to the presence of compressed
gas, also reduce the adsorbents bulk density and thus resulting in inefficient volumetric adsorption
capacity [4].
Using COMSOL Multiphysics to study gas adsorption is a new research method for methane
adsorption process in ANG system. Humble et al. developed a complex model combining gas
collection, concentration, and purification using COMSOL Multiphysics and optimized physical
parameters. Their numerical simulations have been a valuable tool for assessing the capability of
proposed processes and optimizing process parameters [76]. Jun-zhen et al. solved the non-steady
mobile model of the coalbed methane through COMSOL Multiphysics and reported that the
obtained distribution curve can predict the mobilization and distribution regularity in coalben
methane [77]. Sahoo et al. studied thermal effects associated with dynamic charge of methane in
an ANG cylinder filled with activated carbon theoretically and experimentally. They presented a
2D model coupling hydrodynamics, heat transfer, and adsorption phenomena and reported that the
temperature profiles obtained from experiments agrees well with the prediction given by
simulation. The surface temperature profiles of adsorbent bed with various time intervals during a
slow charging rate at 1.0 l /min and fast charging rate at 30.0 l /min have been depicted in Figure
15 [78, 79]. As a point of reference, using COMSOL Multiphysics to study the adsorption process
of multicomponent gas is expected to have a good application prospect.
15

(a)

(b)
Figure 15 Temperature maps in the adsorbent bed at different times during charging of ANG cylinder at
a flow rate of (a) 1.0 l /min; (b) 30.0 l/min [78, 79]

4. Conformable Tank
On-board methane storage presents unique challenges for the commercialization of ANG vehicles.
Vehicle range, cost, weight of storage system, compatibility of component materials, and
durability are all key issues. Due to low volumetric energy density of methane storage, a larger
volume of storage is required for acceptable vehicle range. With increased volume requirements,
however packaging of the fuel storage system becomes a problem. Liquid fuels such as diesel and
gasoline can stored in tanks that closely conform to the available space on the vehicle without
decreasing passenger capacity. For the methane, the pressurized storage constrains the geometry
of the fuel tank. Cylinders are being used in practice currently since cylindrical tanks have nearoptimum pressure vessel structural efficiency. However, the cylinders often dont lend themselves
to efficient use of the substantially rectangular volumes available on the vehicle. For instance, in
a rectangular envelop with an aspect ratio (width/height) equal to an integer, cylinders provide less
than 75% of the available storage volume. For non-integer aspect ratios, this factor can be as low
as 50%. Thus, the combination of energy density and packaging efficiency results in a significant
range penalty for ANG vehicles.
The problem of maximizing on-board fuel storage is being addressed through the development of
pressurized conformable tanks. Thiokol Propulsion supported from the US Department of Energy
is in the midst of a program to apply and extend its conformable composite pressure vessel
capabilities to the on-board methane storage [80]. According to the physical principle that
16

cylinders efficiently contain internal pressure by membrane response, the fundamental concept for
the conformable tank consists of adjoining cylindrical segments with internal web reinforcements.
In the DOEs call for proposals, conformability factor for a conformable tank is defined as the
outer tank volume divided by the smallest enclosing cuboid volume and simply gives the
packaging efficiency of a tank within a box. It can be explained using Figure 16 [81]. It is known
that a spherical pressure vessel or tank has the best use of tank material property because this
creates the most uniform distribution of stress. However, a spherical tank would have a poor
utilization of external space. Using a spherical tank with a diameter of D as example, in a D unit
cuboid space, only 0.5236 (or /6) of this space is used for the tank, and the remaining 0.4764 of
this space is very difficult to use and is likely wasted. In the case of a cylindrical tank, which is
conventionally and commonly used as pressure vessel for specialty gas storage, if the tank has
completely flat ends at each side, it has a conformity factor of 0.785, or 78.5%.

(a)

(b)
Figure 16 (a) Cylinders in rectangular vs. (b) a conformable tank in a rectangular envelope [81]

In early 2012, the US Department of Energy ARPA-E program announced research and
development funding of 30 million dollars for Methane Opportunities for Vehicular Energy
(MOVE) projects. ARPA-E has set the technical target for conformity at greater than 90%. In the
case of stacking up cylinders with non-flat tank heads, as shown in Figure 17 [82], the conformity
factor would have difficulties in meeting this target. In fact, a well-known earlier effort in
developing conformable and flat-panel ANG tank as shown in Figure 18 has a very good
conformity factor [52].
Ginzburg pointed out an important and inherent performance issue in ANG tanks. During filling
up the tank, the tank would be heated up due to the release of adsorption heat. For natural gas, the
released heat is about 15 KJ/mol. A direct consequence of heat release is that if the tank is charged
17

to a rated pressure of 3.5 MPa (500 psi), due to the rising temperature, the tank is no longer filled
after the tank has been cooled off. An example is that for a 40 liter tank filled in an assumed
adiabatic process to the full volumetric capacity, as shown in Figure 19 [19], the temperature could
raise by 90 , and the pressure could go up as high as 670 psi, almost 1/3 higher than the rated
working pressure. In the tank discharging process, depending on the desorption rate, the tank
temperature can drop so much that it affects the discharging rate. A number of technical papers
studied this problem by theoretical modeling and actual experiments. It is not easy to just cool the
tank or chill the incoming natural gas during charging the tank, then heat up the tank during
discharging, or employ phase change material to store the adsorption heat then later use it for
desorption needs. Any material equipment added for this thermal management purpose can
increase the system cost and vehicle weight. Higher vehicle weight then means lower mileage for
the same amount of fuel (which is called thermal energy equivalent, or gallon gasoline equivalent).

Figure 17 Non-conformable end caps [82]

Figure 18 Conformable and flat-panel ANG tank developed by AGLARG [52]

18

Figure 19 The first generation conformable tank designed by University of Missouri

The composite conformal tank design patented by Blair et al. as shown in Figure 20 [81], or a
commercial non-conformal metal-composite tank as shown in Figure 21, also have the same
problem of poor heat conduction due to the wrapping of the outside composite layer. Lin et al.
patented a conformal tank, which includes a body having substantially flat top and bottom wall
and a pair of opposing sidewalls that join the top and bottom wall to form an open-ended cuboid,
as shown in Figure 22 [83]. This tank is structured and operable for adsorbent storage and
dispensing of compressed gas and has a conformity index of approximately 0.9.
One of the features of ANG tanks in contrast to CNG tanks is that without additional cooling and
heating equipment. A good conduction for the tank wall is necessary for better volumetric
efficiency and performance (e.g., better charging rate, and faster delivery/desorption rate).

Figure 20 Composite conformable tank [81]

19

Figure 21 A Chinese Type II (steel + composite) CNG tank

Figure 22 Conformal tank for adsorbed natural gas storage [83]

5. Related Technologies
5.1 Adsorbent Compacting Machine
In light of the above discussion, adsorbents need to be compacted to monolith in order to obtain
high volumetric methane adsorption. This requires a machine which can provide proper pressure
and temperature to adsorbents. The manufacturing time also is an important issue to consider.
Therefore, developing a continuous production process of adsorbent monolith and a special
equipment for this process should be one of the research directions on ANG technology in the
future.
5.2 Home-fueling Compressor
The estimated number of factory-produced, natural gas vehicles currently in the US is fewer than
200,000. One of the possible reasons for this low number is the lack of infrastructure. There are
too few charging stations available for convenient refueling. Ginzburg labeled this infrastructure
and demand problem a chicken and egg problem. This is because a commercial CNG charging
station may costs an estimated 2-5 million dollars to construct and this initial cost needs a very
large customer base to justify. ANG technology can alleviate this problem in two ways. Due to
20

much lower pressure (recommended standard of 500 psi instead of 3600 psi), the cost of
construction of each charging station for ANG is just a fraction of CNG stations. Also due to low
charging pressure, safe and reliable fueling equipment can be affordable for home fueling. To
overcome this obstacle and provide ease of refueling for millions of potential users of NGVs, one
of the MOVE objectives is to create a home fueling compressor and ANG tank with a combined
cost less than $2,000. Considering the fact that the well-known Phill FuelMaker (it is for CNG
home fueling and can charge to 2900 or 3600 psi), as shown in Figure 23 [84], has a price tag at
about $4,000, a home fueling compressor for lower pressured ANG tanks (charged at about 500
psi) based on traditional compressor technology, to achieve $500 price for mass production seems
not a difficult objective. There is another difficulty that is not obviously seen. Unless a dedicated
power supply is setup, in a regular household the power supply is usually set at 1.5 kW limit.
FuelMaker has actual or calculated power rating at 0.8-1.0 kW. At its specified rate, it would take
more than 20 hours to fill a 10 GGE (Gasoline Gallon Equivalent) CNG tank. It is highly desirable
to charge a 5 GGE ANG tank for about 5 hours (or 1 GGE/hr, this is also a MOVE objective for
home fueling compressor).

Figure 23 Home-fueling CNG compressor produced by Phill [84]

Although a multi-million single project has been funded for a new technology home fueling
compressor by MOVE, no results have been reported so far. In the authors opinion, to achieve a
$500 per unit price in mass production, even with traditional piston multi-stage compressor
technology, is not a very difficult mechanical engineering problem. Based on the current
developments, a safe, low cost, and reliable home fueling ANG compressor and tank will still take
considerable effort to develop, even though it is technically not difficult. There is also an energy
consumption target for charging, which is <1.7 kWh/GGE. Hence, the mechanical design problem
is well defined: to utilize limited power so that there is no need to install a new power supply line;
the charging time for 5 GGE ANG tank should be about 5 hours while FuelMaker may take more
than 10 hours to charge a tank at this size; the total cost of compressor and tank should be less than
$2000.

6. Summary and conclusions


ANG constitutes a safer and low-cost way to store methane for natural gas vehicles at an acceptable
energy density while working at substantially low pressures (3.5-4.0 MPa), allowing for
21

conformable store tank. The funded R&D projects are in four areas. One is a volumetrically more
efficient adsorbent for natural gas storage. The second one is adsorption theory for a certain
adsorbent and the process of adsorption and desorption. The third one a conformable tank for ANG
vehicles that must be light weight and economical. The last one is a low cost and efficient home
fueling compressor.
In the subject area of improving the adsorbent, the aim is to obtain such adsorbents with high
storage capacity, long service life, and less-expensive. Although the basic understanding of the
adsorbent behavior is a physical-chemistry problem, engineering implementation of improved
adsorbents again are full of mechanical engineering challenges. Two kinds of adsorbents attracting
more attentions currently are metal organic frameworks and carbon materials, including activated
carbon, carbon nanotubes, and graphene. Once better adsorbents are developed from the several
large grants that are in adsorbent development, making adsorbent pellets economically and
efficiently, or protecting the adsorbents from long term damage by moisture and harmful gas
molecules in the supplied natural gas using mechanical means are not as obvious mechanical
engineering problem. However, the MOVE objectives proposed by US DOE cannot be
accomplished in this area without a good mechanical engineering implementation.
In the subject area of adsorption theory and adsorption-desorption process, the adsorption behavior
for a certain adsorbent and the adsorption theory of multicomponent gas are invested a lot of R&D
resources from different fields. Numerical simulations on adsorption model through finite element
software have promoted the research progress on adsorption theory. Improvements on this area
would facilitate the research of new adsorbent and the application of ANG technology on natural
gas vehicles.
In the subject area of conformable tank, on-board methane tank presents unique challenges for the
commercialization of ANG vehicles. MOVE projects have set the technical target for conformity
at greater than 90%. Although weight and conformity objectives as set by MOVE are not too
difficult to achieve, greatly lowering the cost and using thermal management to help improve the
volumetric efficiency is more difficult to accomplish. This is because a conformable shape
inherently is not the most efficient way to use tank material. Well-known high strength material
such as carbon fiber composite is not the best for use in an ANG tank. The second challenge is the
compromise between performance and cost. Too much effort on thermal management will increase
the manufacturing cost and vehicle weight. Again, both of these more difficult challenges are
mechanical engineering topics for more R&D efforts.
In the subject area of home fueling compressors, existing technology is reliable. The cost can be
further lowered when the demand for ANG home fueling increases. Unless there is a technological
breakthrough, achieving faster charging rate with limited home power presents a more difficult
challenge. Research is most needed in gas thermo-dynamics, compressor technology and
mechanical design. It also is an obvious mechanical engineering challenge.

Acknowledgements

22

The authors wish to thank Southern California Gas Company for funding the project that is related
to this study. The authors also thank Professor Peter Pfeifer, Chairman of the Department of
Physics and Astronomy of University of Missouri, for permitting to use some data generated from
the ANG research project.

References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.

18.
19.

Wang, L., H. Gardeler, and J. Gmehling, Model and experimental data research of natural gas
storage for vehicular usage. Separation and purification technology, 1997. 12(1): p. 35-41.
EIA, U.S. Annual Energy Outlook 2014. 2014; Available from:
http://www.eia.gov/forecasts/archive/aeo14/index.cfm.
NaturalGas.org. Natural Gas and the Environment. 2011; Available from:
http://naturalgas.org/environment/naturalgas/.
Menon, V.C. and S. Komarneni, Porous Adsorbents for Vehicular Natural Gas Storage: A Review.
Journal of Porous Materials, 1998. 5(1): p. 43-58.
Mason, J.A., M. Veenstra, and J.R. Long, Evaluating metalorganic frameworks for natural gas
storage. Chem. Sci., 2014. 5(1): p. 32-51.
Wegrzyn, J., H. Wiesmann, and T. Lee. Low pressure storage of natural gas on activated carbon.
1992.
Quinn, D.F., J.A. MacDonald, and K. Sosin, Microporous carbons as adsorbents for methane
storage. 1994: U.S. Department of Energy Technical Report.
Biloe, S., V. Goetz, and S. Mauran, Characterization of adsorbent composite blocks for methane
storage. Carbon, 2001. 39(11): p. 1653-1662.
Pupier, O., V. Goetz, and R. Fiscal, Effect of cycling operations on an adsorbed natural gas
storage. Chemical Engineering and Processing: Process Intensification, 2005. 44(1): p. 71-79.
Yang, X.D., et al., Experimental studies of the performance of adsorbed natural gas storage
system during discharge. Applied Thermal Engineering, 2005. 25(4): p. 591-601.
Celzard, A., et al., The effect of wetting on pore texture and methane storage ability of NaOH
activated anthracite. Fuel, 2007. 86(1-2): p. 287-293.
Molina-Sabio, M., C. Almansa, and F. Rodr guez-Reinoso, Phosphoric acid activated carbon discs
for methane adsorption. Carbon, 2003. 41(11): p. 2113-2119.
Bagheri, N. and J. Abedi, Adsorption of methane on corn cobs based activated carbon. Chemical
Engineering Research and Design, 2011. 89(10): p. 2038-2043.
ARPA-E. U.S. ARPA-E's MOVE Projects. 2015; Available from: http://arpa-e.energy.gov/?q=arpae-programs/move.
Casco, M.E., et al., High-Pressure Methane Storage in Porous Materials: Are Carbon Materials in
the Pole Position? Chemistry of Materials, 2015.
Matranga, K.R., A.L. Myers, and E.D. Glandt, Storage of natural gas by adsorption on activated
carbon. Chemical engineering science, 1992. 47(7): p. 1569-1579.
Nitta, T., M. Nozawa, and S. Kida, Gas-Phase Adsorption Characteristics of High-Surface Area
Carbons Activated from Meso-Carbon Micro-Beads. Journal of chemical engineering of Japan,
1992. 25(2): p. 176-182.
Pfeifer, P., et al., Nearly Space-Filling Fractal Networks of Carbon Nanopores. Physical Review
Letters, 2002. 88(11).
Pfeifer, P., et al. High-surface-area biocarbons for reversible on-board storage of natural gas and
hydrogen. in MRS Proceedings. 2007. Cambridge Univ Press.
23

20.
21.
22.
23.
24.
25.

26.
27.
28.
29.
30.
31.
32.

33.
34.
35.
36.
37.
38.
39.
40.

41.

Firlej, L., Sz. Roszak, B. Kuchta, P. Pfeifer, and C. Wexler, Enhanced hydrogen adsorption in boron
substituted carbon nanospaces. J Chem Phys, 2009. 131(16): p. 164702.
Zhao, Y.-L. and J.F. Stoddart, Noncovalent functionalization of single-walled carbon nanotubes.
Accounts of chemical research, 2009. 42(8): p. 1161-1171.
Bekyarova, E., et al., Single-wall nanostructured carbon for methane storage. The Journal of
Physical Chemistry B, 2003. 107(20): p. 4681-4684.
Tanaka, H., et al., Methane adsorption on single-walled carbon nanotube: a density functional
theory model. Chemical physics letters, 2002. 352(5): p. 334-341.
Cao, D., et al., Optimization of single-walled carbon nanotube arrays for methane storage at
room temperature. The Journal of Physical Chemistry B, 2003. 107(48): p. 13286-13292.
Ganji, M.D., A. Mirnejad, and A. Najafi, Theoretical investigation of methane adsorption onto
boron nitride and carbon nanotubes. Science and Technology of Advanced Materials, 2010.
11(4): p. 045001.
Lee, J.-W., et al., Methane adsorption on multi-walled carbon nanotube at (303.15, 313.15, and
323.15) K. Journal of Chemical & Engineering Data, 2006. 51(3): p. 963-967.
Thierfelder, C., et al., Methane adsorption on graphene from first principles including dispersion
interaction. Surface Science, 2011. 605(7): p. 746-749.
Carrillo, I., E. Rangel, and L. Magaa, Adsorption of carbon dioxide and methane on graphene
with a high titanium coverage. Carbon, 2009. 47(11): p. 2758-2760.
Wood, B.C., et al., Methane and carbon dioxide adsorption on edge-functionalized graphene: a
comparative DFT study. The Journal of chemical physics, 2012. 137(5): p. 054702.
Chae, H.K., et al., A route to high surface area, porosity and inclusion of large molecules in
crystals. Nature, 2004. 427(6974): p. 523-527.
Romanos, J., et al., Nanospace engineering of KOH activated carbon. Nanotechnology, 2012.
23(1): p. 015401.
Eddaoudi, M., et al., Modular chemistry: secondary building units as a basis for the design of
highly porous and robust metal-organic carboxylate frameworks. Accounts of Chemical
Research, 2001. 34(4): p. 319-330.
Getman, R.B., et al., Review and analysis of molecular simulations of methane, hydrogen, and
acetylene storage in metalorganic frameworks. Chemical reviews, 2011. 112(2): p. 703-723.
Furukawa, H., et al., Ultrahigh porosity in metal-organic frameworks. Science, 2010. 329(5990):
p. 424-428.
Farha, O.K., et al., Metalorganic framework materials with ultrahigh surface areas: is the sky
the limit? Journal of the American Chemical Society, 2012. 134(36): p. 15016-15021.
Morse, P.M., Diatomic molecules according to the wave mechanics. II. vibrational levels. Physical
Review, 1929. 34(1): p. 57.
Lennard-Jones, J.E., Cohesion. Proceedings of the Physical Society, 1931. 43(5): p. 461.
Allinger, N.L., Conformational analysis. 130. MM2. A hydrocarbon force field utilizing V1 and V2
torsional terms. Journal of the American Chemical Society, 1977. 99(25): p. 8127-8134.
Xu, Q. and C. Zhong, A general approach for estimating framework charges in metal organic
frameworks. The Journal of Physical Chemistry C, 2010. 114(11): p. 5035-5042.
Wilmer, C.E. and R.Q. Snurr, Towards rapid computational screening of metal-organic
frameworks for carbon dioxide capture: Calculation of framework charges via charge
equilibration. Chemical Engineering Journal, 2011. 171(3): p. 775-781.
Frost, H., T. Dren, and R.Q. Snurr, Effects of surface area, free volume, and heat of adsorption
on hydrogen uptake in metal-organic frameworks. The Journal of Physical Chemistry B, 2006.
110(19): p. 9565-9570.

24

42.

43.

44.

45.

46.
47.

48.
49.

50.
51.
52.
53.
54.
55.
56.

57.
58.
59.
60.
61.

Dren, T. and R.Q. Snurr, Assessment of isoreticular metal-organic frameworks for adsorption
separations: a molecular simulation study of methane/n-butane mixtures. The Journal of
Physical Chemistry B, 2004. 108(40): p. 15703-15708.
Gallo, M. and D. Glossman-Mitnik, Fuel gas storage and separations by metal organic
frameworks: Simulated adsorption isotherms for H2 and CH4 and their equimolar mixture. The
Journal of Physical Chemistry C, 2009. 113(16): p. 6634-6642.
Kaye, S.S., et al., Impact of preparation and handling on the hydrogen storage properties of Zn4O
(1, 4-benzenedicarboxylate) 3 (MOF-5). Journal of the American Chemical Society, 2007. 129(46):
p. 14176-14177.
ARPA-E. MOVE Annual Meeting/NGVTF 2014 Fall Meeting Onboard Storage Projects. 2014;
Available from:
http://www1.eere.energy.gov/cleancities/natural_gas_forum_meeting_oct2014.html.
Zhang, S.Y., O. Talu, and D.T. Hayhurst, High-pressure adsorption of methane in zeolites NaX,
MgX, CaX, SrX and BaX. The Journal of Physical Chemistry, 1991. 95(4): p. 1722-1726.
Reich, R., W.T. Ziegler, and K.A. Rogers, Adsorption of methane, ethane, and ethylene gases and
their binary and ternary mixtures and carbon dioxide on activated carbon at 212-301 K and
pressures to 35 atmospheres. Industrial & Engineering Chemistry Process Design and
Development, 1980. 19(3): p. 336-344.
Ding, T.F., et al., A generalized treatment of adsorption of methane onto various zeolites.
Langmuir, 1988. 4(2): p. 392-396.
Abdul-Rehman, H., M. Hasanain, and K. Loughlin, Quaternary, ternary, binary, and pure
component sorption on zeolites. 1. Light alkanes on Linde S-115 silicalite at moderate to high
pressures. Industrial & engineering chemistry research, 1990. 29(7): p. 1525-1535.
Talu, O., S.Y. Zhang, and D.T. Hayhurst, Effect of cations on methane adsorption by NaY, MgY,
CaY, SrY, and BaY zeolites. The Journal of Physical Chemistry, 1993. 97(49): p. 12894-12898.
Quinn, D.F. and J.A. Holland, Carbonaceous material with high micropore and low macropore
volume and process for producing same. 1991, Google Patents.
Pfeifer, P. Advanced Natural Gas Fuel Tank Project. in Natural Gas Vehicle Technology Forum,
San Francisco, October. 2011.
Raff, L., J. Lorenzen, and B. McCoy, Theoretical Investigations of GasSolid Interaction
Phenomena. I. The Journal of Chemical Physics, 1967. 46(11): p. 4265-4274.
Do, D.D., Adsorption analysis. 1998: World Scientific.
Condon, J.B., Surface area and porosity determinations by physisorption: measurements and
theory. 2006: Elsevier.
Rouquerol, F., Jean Rouquerol, Kenneth S.W. Sing, P. Llewellyn, and G. Maurin, Adsorption by
powders and porous solids: principles, methodology and applications. 2nd ed. 2013: Academic
press.
Choudhary, V.R. and S. Mayadevi, Adsorption of methane, ethane, ethylene, and carbon dioxide
on silicalite-l. Zeolites, 1996. 17(5): p. 501-507.
Jensen, H., Chemical Studies on Toad Poisons. VIII. The Dehydrogenation of Cinobufagin. Journal
of the American Chemical Society, 1935. 57(12): p. 2733-2734.
Endo, M., et al., Brunauer S, Emmett PH, Teller E. Adsorption of gases in Dresselhaus MS. Basic
properties and new applications of multimolecular layers. J Am Chem Soc, 1938. 60: p. 309-19.
Kim, S.K. and B.K. Oh, Theory of localized and non-localized adsorption. Thin Solid Films, 1968.
2(5): p. 445-456.
Xi, Y., Z.P. Baant, and H.M. Jennings, Moisture diffusion in cementitious materials Adsorption
isotherms. Advanced Cement Based Materials, 1994. 1(6): p. 248-257.

25

62.
63.
64.
65.
66.

67.
68.

69.
70.
71.
72.
73.
74.

75.

76.
77.

78.

79.

80.
81.

Kruk, M., M. Jaroniec, and Y. Bereznitski, Adsorption study of porous structure development in
carbon blacks. Journal of colloid and interface science, 1996. 182(1): p. 282-288.
Sing, K., et al., Physical and biophysical chemistry division commission on colloid and surface
chemistry including catalysis. Pure Appl. Chem, 1985. 57(4): p. 603-19.
Kapoor, A. and R. Yang, Correlation of equilibrium adsorption data of condensible vapours on
porous adsorbents. Gas Separation & Purification, 1989. 3(4): p. 187-192.
Donohue, M. and G. Aranovich, Classification of Gibbs adsorption isotherms. Advances in colloid
and interface science, 1998. 76: p. 137-152.
Li, H., et al., Establishing microporosity in open metal-organic frameworks: Gas sorption
isotherms for Zn (BDC)(BDC= 1, 4-benzenedicarboxylate). Journal of the American Chemical
Society, 1998. 120(33): p. 8571-8572.
Tan, Z. and K.E. Gubbins, Selective adsorption of simple mixtures in slit pores: a model of
methane-ethane mixtures in carbon. The Journal of Physical Chemistry, 1992. 96(2): p. 845-854.
Cracknell, R.F., P. Gordon, and K.E. Gubbins, Influence of pore geometry on the design of
microporous materials for methane storage. The Journal of Physical Chemistry, 1993. 97(2): p.
494-499.
Mota, J.B., et al., Dynamics of natural gas adsorption storage systems employing activated
carbon. Carbon, 1997. 35(9): p. 1259-1270.
Stella, A. and A. Myers. Adsorption of methane on activated carbon: Comparison of molecular
simulation with experiment. in Proc. Annual AIChE Meeting. 1991.
Sun, J., Todd A. B., Mark J. R., Absorbed Natural Gas Storage with Activated Carbon. 1996.
Basumatary, R., et al., Thermal modeling of activated carbon based adsorptive natural gas
storage system. Carbon, 2005. 43(3): p. 541-549.
Cracknell, R. and K. Gubbins, A Monte Carlo study of methane adsorption in aluminophosphates
and porous carbons. Journal of molecular liquids, 1992. 54(4): p. 239-251.
Ma, Z., et al., Very high surface area microporous carbon with a three-dimensional nano-array
structure: synthesis and its molecular structure. Chemistry of materials, 2001. 13(12): p. 44134415.
Bojan, M.J., R. van Slooten, and W. Steele, Computer simulation studies of the storage of
methane in microporous carbons. Separation science and technology, 1992. 27(14): p. 18371856.
Humble, P.H., R.M. Williams, and J.C. Hayes, Finite Element Modeling of Adsorption Processes for
Gas Separation and Purification. 2009. Medium: X.
Di, J. and J.J. Liu, Numerical simulation based on COMSOL multiphysic to steady and nonequilibrium adsorption-mobilization of coalbed methane. Journal of Wuhan Polytechnic
University, 2007. 3: p. 019.
Sahoo, P.K., Mathew John, Bharat L. Newalkar, N.V. Choudhary, and K.G. Ayappa, Filling
Characteristics for an Activated Carbon Based Adsorbed Natural Gas Storage System. Industrial
& Engineering Chemistry Research, 2011. 50(23): p. 13000-13011.
Sahoo, P.K., Mathew John, Bharat L. Newalkar, N.V. Choudhary, and K.G. Ayappa, Corrections to
Filling Characteristics for an Activated Carbon Based Adsorbed Natural Gas Storage System.
Industrial & Engineering Chemistry Research, 2014. 53(11): p. 4522-4523.
Kunz, R. and R. Golde. High-Pressure Conformable Hydrogen Storage for Fuel Cell Vehicles. in
2000 Hydrogen Program Review. 2000. Alexandria, Virginia.
Xu, H., Yuyi Lin, Optimal design of conformable adsorbed natural gas tank. 2013: University of
Missouri Project Report.

26

82.

83.
84.

APRA-E. Low-Cost Efficient Manufacturing of Pressurized Conformal Compressed Natural Gas


Storage Tanks. 2014; Available from: http://arpa-e.energy.gov/?q=slick-sheet-project/ultralight-conformable-natural-gas-tank.
Lin, Y. and H. Xu, Conformal Tank For Adsorbant Natural Gas Storage. 2013, Google Patents.
FuelMaker, B. Home-fueling CNG compressor. 2015; Available from:
http://www.brcfuelmaker.com/en/phill-domestico-prodotto-brc-fuel-maker.aspx.

27

S-ar putea să vă placă și