Sei sulla pagina 1di 45

A

SEMINAR REPORT ON

ADSORBED NATURAL GAS


PREPARED BY:

BHUVAN SHAH (U13CH064)

B. TECH. IV

CHEMICAL ENGINEERING DEPARTMENT

GUIDED BY:

DR. PARIMAL A. PARIKH

PROFESSOR

DEPARTMENT OF CHEMICAL ENGINEERING

CHEMICAL ENGINEERING DEPARTMENT


SARDAR VALLABHBHAI NATIONAL INSTITUTE OF TECHNOLOGY, SURAT

i
CERTIFICATE

This is to certify that Seminar entitled

“ADSORBED NATURAL GAS”

Submitted by:

BHUVAN SHAH (Roll No. U13CH064)

In partial fulfilment of the requirement for the award of bachelor degree in


Chemical Engineering
From Sardar Vallabhbhai National Institute of Technology, Surat

It is the record of his own work carried out by him under the guidance of

Dr. Parimal A. Parikh (Professor, Chemical Engineering Department, SVNIT Surat).

Dr. PARIMAL A. PARIKH Dr. Z.V.P. MURTHY


(Guide) Head Of Department,
Chemical Engineering Department

ii
Mr. Bhuvan Shah, registered in Chemical Engineering Department of Sardar
Vallabhbhai National Institute of Technology, Surat having Roll No. U13CH064
has successfully presented his seminar (as a part of 7th semester curriculum) on

. The seminar is presented before the following members of the committee .

Signature Date

1) Examiner – 1

2) Examiner – 2

Place: Surat
Date:

iii
ACKNOWLEDGEMENT

I take this opportunity to express my deep sense of gratitude and indebtedness to my Project
Guide ―Dr. Parimal Parikh, Professor in Chemical Engineering Department, S.V.N.I.T, Surat
for his valuable guidance, useful comments and co-operation with kind and encouraging attitude
at all stages of the experimental work for the successful completion of this work.

I am also thankful to S.V.N.I.T, Surat and its staff for providing this opportunity which helped us
in gaining sufficient knowledge and to make this Seminar Report successful.

iv
Abstract

Adsorbed natural gas technology is a low-pressure, single-stage compression substitute for


currently used CNG. The performance of ANG storage vessels rests heavily on the type of
adsorbent used. This seminar report explores mainly three important adsorbents for methane,
viz., activated carbons, zeolites and metal organic frameworks. Activated carbons have been the
most widely studied and developed adsorbents over the years since they are the cheapest options
with naturally available precursor sources. The highest reported methane uptake using activated
carbon is 191.3 V/V of adsorbent. This was achieved with fluorination of electrospun activated
carbon fibers (EsACFs). Zeolites have been used industrially for variety of applications and
find considerable usage for methane storage too with structural modifications. The values of
methane uptake (V/V) for zeolites range from 70-150 V/V of zeolite adsorbent. Metal Organic
Frameworks have proved to be one of the most compatible options for methane storage,
especially with the latest research in flexible MOF structures. Using flexible MOFs,
commercialization could be possible for ANG vehicles and other applications.

v
Table of Contents
ACKNOWLEDGEMENT ................................................................................................................................. iv

ABSTRACT...................................................................................................................................................... v

Chapter 1. INTRODUCTION ........................................................................................................................... 1

1.1 BACKGROUND ..................................................................................................................................... 1

1.2 OUTLINE .............................................................................................................................................. 3

Chapter 2. ADSORBENTS FOR NATURAL GAS STORAGE SYSTEMS ............................................................... 4

2.1 BASICS OF ADSORBENTS ..................................................................................................................... 4

2.2 ACTIVATED CARBON ........................................................................................................................... 5

2.2.1 EXPERIMENTAL ................................................................................................................................ 6

2.2.2 CHARACTERISATION......................................................................................................................... 8

2.2.3 RESULTS............................................................................................................................................ 9

2.4.4 METHANE ADSORPTION ................................................................................................................ 11

CHAPTER 3. ZEOLITES .................................................................................................................................. 26

3.1 INTRODUCTION ................................................................................................................................. 26

3.2 ZEOLITES AS METHANE ADSORBENTS .............................................................................................. 27

CHAPTER 4. METAL ORGANIC FRAMEWORKS ............................................................................................ 32

4.1 INTRODUCTION ................................................................................................................................. 32

4.2 MOFs AS METHANE ADSORBENTS .................................................................................................... 33

REFERENCES ................................................................................................................................................ 37

vi
Chapter 1
INTRODUCTION

1.1 BACKGROUND

In the early 20th century, investor Henry Ford created the internal combustion engine. This
engine used gasoline to power the first Model T more than 100 years ago. The engine has
changed little since then according to Market Watch, but many changes have occurred in the
world of gasoline. Gasoline use has major implications for the environment, health, dependency
on fossil fuels, and the political stage that might be dissipated if the automobile industry could
use an alternate fuel source.

One alternative to gasoline is natural gas, which consists primarily of methane (85-95%) with
minor amounts of ethane, other higher order hydrocarbons, nitrogen, and carbon dioxide. The
advantage of natural gas is that it is cheaper and widely available than gasoline. Natural gas
burns more cleanly than gasoline, emitting fewer hydrocarbons and 90% less carbon monoxide.

Recent discoveries of large natural gas reserves in the continental United States make it an
attractive alternative to gasoline as a transportation fuel. And, compared to gasoline, natural gas
costs only about 1/4 of the gasoline on equal energy basis. But for passenger service, the natural-
gas vehicle (NGV) should have a 400-mile driving range and its fuel storage system should cost
less than US$2000/vehicle in order to be commercially fit for production. The three classes of
storage systems for the NGV are liquefied, compressed and adsorbed natural-gas types.
Liquefied Natural-Gas (LNG) storage systems generally support it as a fuel for long-haul
vehicles (eg., trains, buses and trucks). For passenger vehicles, CNG has been the storage system
of choice, but recent advances in adsorbent technology have ANG systems now approaching the
storage levels of CNG systems. Gasoline yields 34.80 MJ/L; NG at NTP yields 0.04 MJ/L. LNG
at 1 atm and 113 K yields 600 v/v; CNG at 3000 psig and 298 K yields 230 v/v [6]. The
disadvantage of CNG is that its energy density is 29% of that of gasoline. It is stored at high
pressure (20 MPa) in heavy cylinders. And filling a CNG tank requires an expensive multi-stage
compression facility . Whereas ANG is stored at relatively low pressure (3-4 MPa) at ambient

1
temperature in a lightweight cylinder with a porous sorbent. The attractive feature of ANG is that
the storage tank can be filled with an inexpensive single-stage compressor.

Low pressure adsorbed natural gas (ANG) storage makes conformable tank designs a practical
reality. So, ANG tank design can be integrated into the overall layout of the vehicle in the same
way gasoline tanks are engineered today.

Fig. 1.1. Commercialised ANG Vehicle [7]

A lot of research has been done on finding efficient, economical and physically viable ANG
storage systems for transportation systems. There‟s always a play of balancing maximum storage
capacity versus maximum withdrawal capacity of the adsorbed natural gas. The amount of
adsorbed gas depends on pressure, temperature and sorbent type. Since this adsorption process
is exothermic, an increase of pressure or a decrease of temperature enhances the efficiency of the
process. Changing the sorbent also affects efficiency. Activated carbon is a sorbent with high
surface area that can be used in ANG storage tanks. Currently, researchers are developing new
sorbents with higher adsorption ratio to optimize this process. The affinity or strength of
adsorption of natural gas on sorbent depends on intermolecular potentials which can be
calculated approximately using the theories of adsorption. If the affinity is low, the loading of

2
methane at the filling pressure is low; if the affinity is high, the retention of the methane at
exhaustion is high (ie, the delivery of gas to the engine is difficult). Thus, an optimum affinity
needs to be calculated that maximizes the delivery of methane over an operating cycle [2].

Therefore, a perfect sorbent which can provide the optimum affinity at low pressure and ambient
temperature is the need of the hour.

1.2 OUTLINE

In this seminar report Chapter 1 deals with the types of natural gas vehicles with introduction of
ANG storage systems. Chapter 2 deals with the analysis of various kinds of activated carbon
adsorbents developed over the years. Chapter 3 deals with the use of various types of zeolites for
methane adsorption. Chapter 4 discusses the advancements in the use of metal organic
frameworks for their applications in ANG. Finally some references.

3
CHAPTER 2
Adsorbents for Natural Gas Storage Systems

2.1 BASICS OF ADSORBENTS

Adsorbed natural gas (ANG) uses adsorbents to store natural gas at moderate pressures, 3.5 MPa,
compared to the required high-pressures (20 MPa) for current compressed natural gas (CNG)
technology [2].

Natural gas storage capacity of an adsorbent is currently expressed in terms of the delivered
volume of natural gas per unit of volume containing the adsorbent, measured at standard
conditions (P = 0.101MPa, T = 273 K). To be a commercially attractive system, the final
objective is to reach a delivered quantity equal to 150 (v/v) under real operating conditions [3].

The adsorbent has to be predominantly microporous. Optimal storage capacity will occur when
that fraction of the storage volume that is micropore is maximized, with no void or macropore
volume. In addition to having high surface area, low mesoporosity is desirable. The presence of
low mesoporosity is required to some extent in order to provide easy access for the adsorbate
molecules to/from the micropores. The adsorbent material should also have a high packing
density, with low adsorption heat and high heat capacity. Further, the adsorbent has to be
extremely hydrophobic and inexpensive to the end user [2].

However, we need to simultaneously manage the thermal effects resulting from the adsorption
heat on both charge and discharge processes of the ANG. Since adsorption is an exothermic
process and desorption an endothermic process, both the processes reduce the amount of natural
gas stored and amount delivered, respectively.

This report will discuss mainly about activated carbons, zeolites, and metal–organic frameworks
as they have been investigated extensively for CH4 storage.

4
2.2 Activated Carbon (AC)

The most important ANG storage systems are filled with active carbons made from cellulose
sources (coconut shells, peach pits, etc.) and from non-cellulose polyvinylidenechloride (PVDC)
[4]. The majority of carbon adsorbents are granular, powder, or fibers. They have high surface
area and high pore volume on a mass basis.

Fig. 2.2.1. Types of activated carbon

5
In 2001, S. Biloe´ and co. reported that the most successful AC adsorbent was Maxsorb. The
total stored methane capacity is 101 VV-1 in powder form and 172 VV-1 in monolith form (with
thermoplastic binder), with packing densities 300 and 700 kg/m3, respectively. Matranga et al.
[1] have shown that, for an optimal pore size, the available methane capacity measured over an
operating cycle (3.4-0.14 MPa) and under specific conditions (P=0.1MPa, T=288K) was close to
195 and 137 VV-1 for a monolithic carbon and a pelletized carbon respectively [9]. And,
although monolith carbons exhibit smaller inter-particle space and, consequently, higher
volumetric adsorption capacities, granular carbons involve simpler and cheaper preparation
procedures [14]. Vasiliev et al. investigated different modifications of an activated carbon fiber
“Busofit” produced from pyrolized cellulose activated by water vapour. “Busofit” have some
benefits such as high methane storage capacity near 130 m3/m3 [4].

The total volume V, associated with an activated carbon adsorbent may be split up into its
components:
V = Vc + Vu + Vv + Vvoid
where Vc is the volume of the carbon atoms of which the adsorbent is composed; V_ micropores
volume; V_meso- and macropores volume; Vvoid the space inside the vessel free from adsorbent
bed. The advanced ANG needs to have micropores volume near 50%, solid carbon near 40%
and meso/macropores volume near 10% [4].

2.2.1 Experimental

Conventionally, the precursors to these highly microporous adsorbents have been coal,
hardwood, coconut shell, organic polymer and the likes of it. They are subjected to either
chemical impregnation followed by heating, or are first pyrolysed to carbon and then subjected
to steam treatment. Activating techniques like KOH-activation, phosphoric acid activation, and
the use of steam, CO2 and air activation have been attempted to increase the gas storage capacity
of carbon on a volume basis.

6
Fig. 2.2.1.1. Types of activated carbons from their precursors

A traditional method of preparing the activated carbon from coal was proposed by Jian Sun et
al.: Carbon adsorbent production by both physical and chemical activation techniques was
carried out in a bench scale tubular reactor with a horizontal tube furnace (Lindberg; Type
54232). For physical activation, a three-step process was applied: coal oxidation in air at 225°C
for 2 or 4 h; devolatilization of oxidized coal in nitrogen at 400°C for 1 h; and steam activation
of the resulting char in 50% steam in nitrogen at 800-850°C for 0.5-5.5 h. The air oxidation step
was performed in an auto-programmable ashing furnace (Fisher Scientific, Model 49SA) with
unlimited air supply. About 12 g of sample was used during oxidation and devolatilization while
1 to 2 g of sample was used during steam activation. The gas flow rate during devolatilization

7
and steam activation was 1 L/min. For chemical activation, about 2 g of the coal (-100 mesh)
was mixed with granular KOH (coal/KOH mass ratio 1: I) and ground into a gel-like solid using
a mortar and pestle. This mixture was then activated at 800°C in 100% nitrogen for 0.5 and 1.25
h. After chemical activation, the sample was immediately submerged in deionized water, filtered,
crushed, and then washed again in deionized water to remove KOH derivatives which may have
been on the surface of the particles [16].

2.2.2 Characterisation

In order to characterize the activated carbon absorbent samples, their BET surface areas and t-
plot micropore volumes (micropore volume is defined as the volume of pores < I7 A) are
calculated based on the nitrogen adsorption isotherms (relative pressure P/P0: 0.001-1) measured
with a volumetric adsorption apparatus (eg, Micromentics ASAP2400). Nitrogen isotherms were
determined at 77 K using a micromeretics ASAP 2400 instrument after degassing at 383 K. The
nitrogen isotherms were fitted to both t-plot equations for the determination of the micropore
volume and the BET equation for the relative surface area. The total pore volumes were
determined from a point on the isotherm at a relative pressure above 0.98. The average pore
diameter, (d in nm), was calculated by the following equation: d5 (40003V ) /S where V and S are
the total pore total BET total BET volume and BET surface area, respectively. Methane
adsorption capacity, on a mass basis (g/g), at pressures up to 500 psig is determined with a
pressurized thermogravimetric analyzer (eg, Spectrum Research and Engineering Model TL-
TGA 1900/600 PTGA). Buoyancy correction is performed for coal-derived carbon when
calculating the methane adsorption capacity (g/g).

The Dubinin-Polyani (DP) model is applied to the adsorption data of N2 in order to obtain the
micropore volume, and the micropore surface area for monolayer coverage.

The surface morphologies of activated carbon samples and powder were examined by scanning
electron micro-scope. The true density of all the samples were measured using a helium
pycnometer after degassing 473 K. The true density of a nonporous graphite used as reference
was 3 2.3 g/cm . The bulk densities of all the samples were determined by the weight of the dried

8
activated carbon pellet and its geometrical volume. The destruction pressure was evaluated using
the compressive fracture test [17].

2.2.3 Results

The characterisation results of various types of activated carbons can be generalized when they
are prepared with similar approaches as discussed above. Fig. 6 presents the nitrogen adsorption
isotherms for a particular carbon pellet sample. It can be observed that this sample follows Type
I isotherms characteristic of microporous materials. The degree of activation for these
carbonaceous materials is high and the micropore volumes obtained are large.

Fig. 2.2.3.1. N2 adsorption isotherm for an activated carbon

The pore volume and BET surface area decreased with an increase in the compression pressure
of the raw materials under the same heat-treatment conditions. In fact, the total pore volume, the
micropore volume, and the relative surface area increased at longer activation periods.

The micropore volume (cm /cm) almost increased with increased compression pressure of the
raw material.

9
----
Fig. 7 illustrates the changes that occur in total pore volume and BET surface area with
activation expressed as the percent of weight yield from the raw material. These show the same
trend, that is, as activation increases, increases are observed in the total pore volume and BET
surface area [17].

Fig. 2.2.3.2. Total pore volume and BET surface area with percent
yield. Solid and open circles: pellet samples.

Fig. 2.2.3.3. SEM image of raw activated carbon [18]

10
2.2.4 Methane Adsorption

Over the years, several techniques and models have been developed to understand and improve
the methane adsorption, storage capacity and deliverability. This section describes those
advancements in Activated Carbons for methane adsorption.

In 1991, Matranga et. al. [1] used the “Slit model for perfect affinity” to obtain the ideal curve
for methane adsorption with optimum conditions. The optimum adsorption isotherm is calculated
by varying the affinity parameter „C‟ of Langmuir isotherm, i.e.,

a=

to deliver the maximum amount of methane over a cycle operating between filling pressure P2
and exhaustion pressure P1. In the above equation, „a‟ is the specific adsorption (mmol/g). For
isothermal operation, the maximum delivered methane is obtained for

C=

Then, a comparison of this model was made with the Grand Canonical Monte Carlo (GCMC)
simulations with experimental adsorption isotherm of methane on carbon at 250C which is
depicted in the figure below.

11
Fig. 2.2.3.4. Comparison of perfect affinity model and GCMC simulations.

The following conclusions were drawn. The theoretical maximum storage capacity for methane
at 3.4 MPa is 209 V/V for monolithic carbon and 146 V/V for pelletized carbon. Whereas, the
maximum delivered capacity of ANG is 195 V/V for monolithic carbon and 137 V/V for
pelletized carbon. The maximum storage capacity for ANG lies between the values for GCMC
simulations and perfect affinity.

S. Sircar and co. (1995, [12]) plotted methane adsorption isotherms on PX21 carbon, which was
reported to have the highest surface area carbon at that time, at several temperatures. Table
2.2.3.1 shows the net isothermal deliverable capacity of PX21 (granular form) carbon at 303 K
between pressures of 35.0 (storage pressure) and 1.35 atm (exhaust pressure).

12
Table 2.2.3.1. Isothermal deliverable methane capacity of PX21 carbon at 303 K [12]

The figure below depicts methane adsorption isotherms on PX21 carbon at several temperatures

Fig. 2.2.3.5. Adsorption isotherms of pure methane on PX21 carbon at different temperatures
[12]

J. A. F. MacDonald and D. F. Quinn (1997, [13]) worked on composite polymer carbons with
Saran 159 (polyvinylidene chloride or PVDC), Viclan 828, Daran 8600 and Daran 220 latexes as
binders produced in monolithic form. The best composites produced contained approximately
60% superactive carbon and 40% polymer carbon. The composites had percentages of AX-21
carbon ranging from 20 to 90% by weight.

13
The methane adsorption curves were obtained after fitting with Dubinin-Radushkevich equation
at 298 K for a series of AX-21/Daran 220 composite carbons, viz., 0/100, 30/70, 40/60, 50/
50, 60/40, 70/30, 80/20 and 90/10.

Fig. 2.2.3.6. Methane isotherms at 298 K for a series of AX-21/Daran 220 composite carbons per
unit mass of composite carbon [13]

Also, they predicted the methane delivered for all the carbon composites with different
precursors in the following plot.

Fig. 2.2.3.7. Predicted delivery of methane from 500 psia for all the carbon composites produced.
Dotted line: polynomial fit of second order [13]

14
In 1997, V.C. Menon and co. [2] reviewed various types of microporous adsorbents for vehicular
natural gas storage. The effect of other adsorbent properties such as heat of adsorption and heat
capacity on the natural gas storage capacity was also discussed.

The plot below shows methane adsorbed versus the BET surface area of various types of carbons
at 500 psig and 298 K.

Fig. 2.2.3.8. Variation in the experimental gravimetric methane adosorption capacity at 500 psig
and 298 K (313 K for activated carbon fiber) with the gravimetric surface area, for carbon based
organic adsorbents [2]

For carbon it is generally reported that the heat of methane adsorption is about 4 kcal/mol.
Braslaw and coworkers calculated the heat of adsorption of methane based on a measured
carbon temperaturerise of about 25±C when the filling rate is fast enough to prevent significant
wall cooling effects. The heat of adsorption of methane on unconditioned 9 LXC carbon between
15 and 250 psig is about 2.1 kcal/g-mol of methane adsorbed. However, the measured changes in
the internal tank pressure with temperature at constant loading resulted in a heat of adsorption
about 3 kcal/g-mol of methane adsorbed.

15
Carbons have an unusually low heat capacity of 0.7 J/g.K. A fast fill from room temperature and
atmospheric pressure to 500 psig would result in only about 75 to 80% of the amount of gas
stored as would be stored in a slow overnight fill that allows the heat of adsorption to dissipate.
Parkyns and Quinn have observed that a fast fill decreases the performance by 90% over a slow
fill.

It is also pointed out that fast fill of a cold bed (for e.g., resulting from a quick discharge) could
lead to a delivery capacity that exceeds 90% of the capacity of slow fill rates. The maximum
adsorbent temperature recorded was 80±C for a fast 0.5 minutes fill and, 107±C for a 5 minute
fill. During a quick discharge, the adsorbent temperature fell to 30±C for a fast discharge of the
order of a few seconds, and 37±C in about 120 seconds.

In order to maximize adsorbed natural gas storage capacity for vehicular purposes, the emphasis
should be on synthesizing adsorbents with high surface area on a volumetric basis. The literature
reported maximum volumetric ANG storage capacity (180 v/v) attained in the laboratory at 500
psig is about 80% of that attained using CNG at 3000 psig [2].

Vasiliev et al. prepared a simple model (Fig. 2.2.3.9) of cylindrical ANG vessel with internal
finned heater disposed on its axis. They considered several assumptions in their theoretical
model viz., 1. uniform pressure is assumed within a porous structure during the gas
charge/discharge inside the ANG vessel; 2. the demand rate of gas consumption is constant with
constant pressure drop with the time; 3. instantaneous phase equilibrium is assumed between
adsorbed and gaseous methane; the local temperature of adsorbed bed and free gas volume is the
same due to a high heat transfer intensity between solid and gas; 4. the gas inside macropores is
considered as ideal 5. the energy conversion during the gas expansion, or compression is
negligibly small; 6. there is only radial gas flow through the bed; 7. the resistance of mass
diffusion is small [7].

16
Fig. 2.2.3.9. Schematic of the cylindrical sorbent bed element with a finned tube for heating. 1:
fin, 2: sorbent bed, 3: heater (heat pipe), 4: gas channel, 5: vessel envelope [7]

The dynamic model has five components: (1) the equation of energy; (2) the isosteric heat of
desorption; (3) the equation of continuity; (4) the equation of the kinetic of sorption; (5) the
Dubinin and Radushkevich relationship between the gas volume adsorbed in micropore, aeq, and
adsorption potential.

17
where Qhp is the heat flow used to heat one cylinder of the vessel, Thp is the ANG wall
temperature. The suggested simple model gives us a possibility to obtain the field of temperature
and gas concentrations during discharge procedure of the ANG vessel.

Different modifications of an activated carbon fiber “Busofit” (“Busofit-4-055”, “Busofit-055”,


“Busofit AYTM-055”) were investigated as an adsorbent for methane storage. “Busofit” is
produced from pyrolized cellulose activated by water vapor. “Busofit” has such advantages as:
high rate of adsorption and desorption, uniform surface pore distribution (0.6–1.6 nm), small
number of macropores (100–200 nm), with its specific surface 0.5–2 m2/g, small number of
mesopores with 50 m2/g specific surface. Methane isotherms evolution during the cycle of
adsorption/desorption of “Busofit” is shown in figure below.

Fig. 2.2.3.9. Active carbon fiber “Busofit AYTM-055”. Methane sorption isotherms:
experimental data, points; calculated data (Dubinin–Radushkevich equation), lines [7]

Thus, a new type of microporous material “Busofit”, a product of pyrolized cellulose was
suggested, tested and analyzed as an advanced adsorbent, capable of delivering near 150
volumes of methane per volume of the ANG vessel at pressure 3.5 MPa [7].

18
Activated carbon pellets without a binder from cellulose microcrystals as a raw material were
investigated by Inomata et. al. in 2001 [9]. The photograph of an activated carbon pellet and raw
material pellet is shown in Fig. 2.2.3.10.

Fig. 2.2.3.10. Photo image of the raw material (a) and activated carbon pellet (b). [9]

Six samples of carbon pellets were prepared with varying BET surface area, total pore volume
and micro pore volume. The table enlists these properties for the six samples.

Table 2.2.3.2. Pore structures of activated carbon pellets from cellulose microcrystals [9]

The methane isotherms, expressed in milligrams of methane per gram of carbon, at 303 K up to
pressures of 0.85 MPa are shown in Fig. 2.2.3.11. All the adsorption isotherms are Langmuirian.
In this plot it can be observed that the methane capacities of the pellets were greater than that of
the powder at low pressures ( < 0.4 MPa), indicating different micropore size distributions.

19
Fig. 2.2.3.11. The adsorption isotherms of CH4 on activated carbon samples at 303 K.

The total methane storage capacity at 298 K reached 164 cm in 1 cm volume of activated carbon
pellets at 3.5 MPa. This value clearly surpasses the 150 V/V values, which is considered useful
from a commercial point of view [9].

Some advances in the methane storage capacity with activated carbon fibers were observed by Ji
Sun Im and co. (2009, [13]). The prepared Electrospun activated carbon fibers (EsACFs) as a
methane storage medium. The effect of fluorination surface modification was also investigated.
The functional groups played an important role in guiding methane gas into the carbon silt pores
via the attractive force felt by the electrons in the methane molecules due to the high
electronegativity of fluorine. Eventually, the methane uptake increased up to 18.1 wt.% by the
synergetic effects of the highly developed micropore structure and the guiding of methane to
carbon pores by fluorine.

K2CO3 solutions (3 and 5 M) were used as chemical activation agents. Based on the mole
concentrations 3 and 5 M, respectively of the K2CO3 solution used, EsACFs are called EsACF-3
and EsACF-5. Similarly, the fluorinated samples are labelled F-EsACF-3 and F-EsACF-5.

20
The capacity of methane storage is presented in Fig. 2.2.3.12.

Fig. 2.2.3.12. Methane storage of samples [13].

The methane uptake increased steeply for pressures up to 20 MPa and then increased slightly
beyond that point. This may occur because the filling of micropores with methane gas is
completed at low pressures. Comparing EsACF-3 and EsACF-5, a higher ethane uptake can be
observed in the case of EsACF-5. The cause is surely the more highly developed pore structure.
Based on the investigation of fluorination effects for enhancing methane storage, the methane
uptake of the fluorinated samples (F-EsACF-3 and F-EsACF-5) was studied. In all pressure
ranges, fluorination effects can be observed, showing a higher capacity of methane uptake at
about 20% (F-EsACF-3) and 14% (F-EsACF-5), as compared with EsACF-3 and EsACF-5,
respectively. Overall, the methane uptakes of the samples (EsACF-3, F-EsACF-3, EsACF-5 and
F-EsACF-5) were 9.7, 11.7, 15.9 and 18.1 wt.%. The data, calculated using the packing density,
gave values of 102.5, 123.7, 168.1 and 191.3 V/V, respectively.

Hence, the methane uptake increased up to 18.1 wt.% (191.3 V/V) due to the highly developed
micropore structure and catalytic effect of fluorine [13].

21
Recently in 2011, Narges Bagheri and Jalal Abedi [14] came up with a novel technique of
preparing activated carbon from corn cobs. Corn cob waste is an important agricultural by-
product and can be considered as activated carbon precursors for the following reasons: it is a
renewable source and a low-cost material. Due to their botanical origin, corn cobs can be an
excellent starting material to produce nanoporous carbon for natural gas storage.

The amount of methane adsorbed on the activated carbon was determined by using a volumetric
method. To obtain the total number of methane molecules adsorbed, the following formulas were
employed:

PiV = ZniRT
PEV = ZnERT

where subscripts i and E stands for the initial condition at the methane cell and the equilibrium of
the methane and adsorption cells, respectively; V is the volume of the methane cell; P is the
pressure; T is the temperature, R is the universal gas constant, and Z is the compressibility factor
of methane, which is assumed to be 1.

The number of methane molecules adsorbed on the activated carbon in the adsorption cell can,
therefore, be calculated as follows:
nT = nE − ni

where nT is the total mole of methane introduced into the adsorption cell. The amount of methane
trapped in the dead volume of the adsorption cell can be calculated as:

nDead = PDeadVDeadZRT

where, nDead, PDead, and VDead are the mole numbers of methane trapped in the dead volume of the
adsorption cell, the equilibrium pressure of the adsorption cell, and the dead volume of the
adsorption cell, respectively. Finally, the mole numbers of methane adsorbed on the activated
carbon can be determined as:

22
nabsorbed = nT − NDead

The amount of adsorption (the mass ratio of adsorbed methane to activated carbon) at a defined
pressure is, therefore, calculated by dividing the grams of the adsorbed methane (nadsorbed
×16.034) by the weight of the activated carbons filling the adsorption cell.

As shown in Fig. 2.2.3.13, the natural gas adsorption capacity depends on the BET specific
surface area. It can be seen that the amount of natural gas adsorbed presented a sharp increase at
high BET specific surface area, indicating a higher micropore content. The BET surface areas
can vary between 40 and 1320 m2/g. Increasing the BET specific surface area leads to greater
natural gas adsorption capacity.

Fig. 2.2.3.13. Variation of methane adsorption capacity with BET surface area, at a pressure of
500 psi and temperature of 298 K [14].

The best sample with a high BET surface area and a high adsorption natural gas capacity was
chosen to investigate the effect of pressure and temperature on adsorption capacity. Fig. 2.2.3.14
presents the methane adsorption at various pressures and two different temperatures: the
behaviour obtained is very similar, and the same trend exists at the two temperatures. Since the
adsorption is an exothermic process, an increase in temperature results in a lower stored methane
capacity under dynamic conditions; therefore, a raise in temperature can decrease the amount of
adsorbed gas.

23
Fig. 2.2.3.14. Variation of methane adsorption capacity with pressure at two different
temperatures [14]

Thus, activated carbon with a methane uptake of 150 (v/v) or greater and a high surface area
(900–1300m2/g) can be produced from corn cobs. However, contrary to pressure, stored methane
capacity is sensitive to the heat limitation. When the pressure continually increases from 500 to
1500 psi, the stored methane capacity increased from 120 (v/v) to 160 (v/v), the maximum
increase was about 42%. As temperature increases from 298K to 323 K, the loss reaches about
72%, reducing the stored methane capacity from 120 to 70 (v/v).

Also in the year 2011, a group of researchers viz., Ali Morad Rashidi and co. [15] synthesized
nanoporous carbons using furfuryl alcohol and sucrose as precursors and MCM-41 and
mordenite as nanoporous templates. The adsorption curves revealed that sucrose used as carbon
resource had much better performance in methane storage compared with furfuryl alcohol. The
results show that the nanoporous carbon produced from MCM-41 template partially filled with
sucrose had the highest BET surface area of 824 m2/g and methane adsorption capacity of 112
mL CH4/g among the tested samples.

24
Fig. 2.2.3.15. The methane adsorption isotherm of MCM S900A [16]

25
CHAPTER 3
ZEOLITES

3.1 INTRODUCTION

Zeolites are inorganic crystalline materials with uniform sized pores of molecular dimensions.
Zeolite comprises a significant group of alumina-silicate compounds. Its structure is enclosed
with interconnected cavities and is relatively porous. The fundamental building block of the
zeolite is a tetrahedron of four oxygen atoms surrounding a silicon or aluminium atom.
Generally, there are about 50 known natural zeolite minerals which have been identified and
more than about 150 types synthetic zeolite have been prepared. Many researchers have proved
that zeolite has good adsorption characteristics, which has potential to be used as an adsorbent in
gas separation process and adsorptive gas storage.

The framework structures of several types of commercial synthetic zeolite are presented in
Figure 3.1.

Figure 3.1: The framework structural of synthetic zeolites: a) zeolite Linde A; b) zeolite Beta; c)
Ferrierite; d) Mordenite; e) Zeolite Y; and f) ZSM-5 (International Zeolite Association, 2000).

26
3.2 Zeolites as Methane Adsorbents

Early work on adsorbents for ANG storage systems began with zeolites. At that time the
procedures for synthesizing high surface area carbons were not popular. Adsorption of methane
on high surface area materials, especially zeolites, were looked upon as a potential means for
increasing the on-board fuel storage capacity of vehicles [17].

Currently, it has been realized that a major constraint associated with zeolites over carbons is that
due to its structural limitations, surface area (accessible to gas/vapor adsorbates) greater than
1000 m2/g is not attainable even with synthetic zeolites. In addition, zeolites are extremely
hydrophilic and can lose their adsorption capacity for methane with time due to preferential
moisture adsorption. Furthermore, significant macroporosity in the form of interparticle voids is
associated with zeolite packing. This can be avoided only if the packing is done with a single
crystal which is impossible. Relatively high packing densities (compared to carbons) can be
attained, however at the expense of having very low micropore surface areas (5A CaNaA
Zeolite: 394 m2/g and 1.25 g/cc) [3].

By using a binder-free compacting technique, the density of the CaXcompacts could be increased
upto 1 g/cc. Analysis of the data provided in this patent indicates that by compacting a CaX
zeolite to 0.8 g/cc, the methane storage capacity (at ambient temperature and about 132 psig)
could be improved upto 150 v/v. However, by assuming a density of 0.8 g/cc the methane
storage capacity (at ambient temperature and 500 psig) using the adsorption data of Zhang and
co-worker is only about 98 v/v for a CaX zeolite [3].

27
Fig. 3.2.1. Variation in the experimental gravimetric methane adosorption capacity at 500 psig
and 298 K with the gravimetric surface area, for zeolite adsorbents [3].

The methane capacity of various small and large pore zeolites at 500 psig and 298K compiled
from the adsorption isotherms (Fig. 3.2.1) reported in the literature is presented in Table 3.2.1.

28
Table 3.2.1.. Methane adsorption capacity of zeolite adsorbents at 500 psig and 298 K compiled from the
adsorption data reported in the literature by various researchers [3].

Seidel- Morgenstern (2004) [17] discussed several methods capable of measuring adsorption
isotherms, but the most common method is either gravimetric or volumetric method. In this study
adsorption isotherms of CH4 and CO2 were determined at 298 K in a volumetric apparatus as
described in Section 3.5.1. Samples selected to represent the channel type zeolites are ZSM-5,
and ferrierite, and for cage type structures are NaY, NaX and Na-SZ18. It was found that cage-
type zeolites such as NaX and NaY are better adsorbents than channel-type zeolites.

29
Fig. 3.2.2. The methane adsorption isotherms on channel and cage type zeolites at 298 K [17].

These results also suggest that non-polar molecules such as CH4 have low adsorption affinity
towards zeolites. The adsorption isotherms of CH4 reveal interesting phenomena in which
channel type zeolite adsorbed more than cage type zeolites. However this could be due to the
diffusivity of CH4 molecules through the pore of zeolites. Since the average pore diameter of
channel type zeolites are relatively larger than the cage type zeolites, as non-polar molecules,
CH4 are more easily diffused through the larger pore network system.

However, the most recent publication on Nanovalved Adsorbents for CH4 Storage (2016, [18])
throws light on a novel concept of utilizing nanoporous coatings as effective nanovalves on
microporous adsorbents developed for high capacity natural gas storage at low storage pressure.
For the first time it presents the concept of nanovalved adsorbents capable of sealing high
pressure CH4 inside the adsorbents and storing it at low pressure.

30
Fig. 3.2.3. Diagram of CH4 storage by using nanovalved adsorbents

The nanovalved adsorbents are composed of a nanoporous coating on the outer surfaces of the
adsorbent pellets. The nanopores of the coating layer along with the adsorbate adsorbed in the
nanopores function as a valve that can be opened and closed on demand for natural gas storage.
Fig. 3.2.3, panel a shows the structure of nanovalved adsorbents and their functioning process.
Methane loading occurs at high pressures with the nanovalve open. The nanovalve is then closed
by adsorbing strongly adsorbed sealing molecules in nanopores of the coating when the
adsorbents are fully loaded, which holds high pressure methane within the adsorbent pellet.
Loaded adsorbents can then be stored at low pressure, and the nanovalve can be reopened to
release the stored CH4 when needed. In this study, to demonstrate the concept of nanovalved
adsorbents, 5A zeolite was selected as the model sorbent because it is commercially available
and has a reasonable CH4 saturation amount.

A high quality coating layer is the key to the proposed nanovalved adsorbent. In our study,
MCM-48 (Mobil Composition of Matter number-48) was used because it is a periodic
mesoporous amorphous silica possessing long-range ordered framework with uniform
mesopores. In addition, the amorphous nature of MCM-48 promotes the formation of continuous
layers by overcoming the typical issues found in crystalline layers.

Methane adsorption isotherm was measured for the bare 5A zeolite beads. The amounts of CH4
adsorbed were measured at pressure up to 120 bar. We found that CH4 adsorption was saturated

31
at ∼100 bar with a saturation capacity of 2.67 mmol/g. Thus, the maximum volumetric density
(capacity) for the bare 5A was 73 V/V when using a bulk density of 1.22 g/mL.

Fig. 3.2.4. Adsorption isotherm of uncoated 5A zeolite beads at 20°C.

By using 5A zeolite as a model adsorbent, ∼7.5 μm thick, optimized MCM-48 coating with an
average pore size of 3.25 nm was deposited on the external surface of the 5A beads. The
nanopores of the MCM-48 coating functioned as a valve that can be opened and closed on
demand. CH4 was loaded at high pressure (≥50 bar) with the
nanovalve open. The nanovalve was closed by adsorbing 2,2-Dimethylbutane (DMB) in the
pores of MCM-48 coating when the adsorbent was fully loaded. The loaded adsorbent can be
stored at low pressure (∼1 bar), and the nanovalves can be opened again for releasing the stored
CH4 by desorbing DMB. After 4 h storage at low pressure (1 bar), MCM-48 coated 5A
adsorbent can hold 44.1% (32.2 V/V) of the maximum capacity of bare 5A beads (73 V/V).
After further pore size reduction to ∼1.4 nm by molecular layer deposition (MLD), the CH4
storage capacity increased to 55.8−58.4% (40.7−42.6 V/V) of the maximum capacity of the bare
5A beads, which was about 200% higher than storage capacity of the bare 5A beads at the same
storage pressure.

32
CHAPTER 4
Metal Organic Frameworks

4.1 INTRODUCTION

Metal-organic frameworks (MOFs) are compounds consisting


of metal ions or clusters coordinated to organic ligands to form one-, two-, or three-dimensional
structures. MOFs are produced by BASF and commercialized under the denomination of
Basolite C300, Basolite A100 and Basolite Z1200. The main characteristics of these materials
are the well-arranged pore structure as well as the high pore volume. These features make them
attractive for the storage of gases.

They have sponge-like 3D crystals with an extraordinarily large internal surface area – that
feature flexible gas-adsorbing pores. This flexibility gives these MOFs a high capacity for
storing methane, which in turn has the potential to help make the driving range of an adsorbed-
natural-gas (ANG) car comparable to that of a typical gasoline-powered car.

4.2 MOFs as Methane Adsorbents

Similar to Activated Carbons and Zeolites, MOFs have also been studied widely over the years
for their applictions in methane adsorption and storage for ANGVs.

However, in this report the most recent breakthrough in the field of Adsorbed Natural Gas with
MOF for methane storage is discussed.

33
Long et al. [19] came up with a novel solution to the methane storage problems by implementing
flexible MOFs with intrinsic thermal management.

The metal–organic framework Co(bdp) was selected as a potential responsive adsorbent for
methane storage, owing to its large internal surface area and its previously demonstrated high
degree of flexibility. To investigate the ANG storage potential of Co(bdp), a high-pressure CH4
adsorption isotherm was measured at various temperatures (Fig. 4.2.1). There is minimal CH4
uptake at low pressures and a sharp step in the adsorption isotherm at around 16 bar. Although
there is hysteresis in the desorption isotherm, the hysteresis loop is closed by around 7 bar, such
that there is less than 0.2 mmol g−1 of CH4 adsorbed at pressures below 5.8 bar.

Fig. 4.2.1. Total CH4 adsorption isotherms at various temperatures for Co (bdp) [19]

The step in the CH4 isotherm is fully reproducible over at least 100 adsorption– desorption
cycles, and can be attributed to a reversible structural phase transition between a collapsed, non-
porous framework and an expanded, porous framework at transition pressures that are ideal for
ANG storage.

34
Fig. 4.2.2. Space-filling models of collapsed (left) and CH4-expanded (right) Co(bdp) [19]

Fig. 4.2.3.The crystal structures of the collapsed (0 bar,„low P CH4‟) and CH4-expanded (30
bar,„high P CH4‟) phases of Co(bdp) [19]

Responsive adsorbents, such as Co(bdp), offer the possibility of managing heat intrinsically
within a material, rather than through an external system, by using the enthalpy change of a
phase transition to partially, or perhaps even fully, offset the heats of sorption. For Co(bdp), the
expansion of the framework during adsorption is endothermic, because energy is needed to
overcome the greater thermodynamic stability of the collapsed phase. As a result, some of the
enthalpy of CH4 adsorption should go towards providing the heat needed for the transition to the
expanded phase, lowering the overall amount of heat released compared to adsorption in the
absence of a phase transition. Similarly, the transition to the collapsed phase is exothermic, and
some of the heat released by the framework as it collapses should offset the endothermic
desorption of CH4.

The usable CH4 capacity of Co(bdp) at 25 °C is 155 cm3 STP cm−3 (v/v) for adsorption at 35 bar
and 197 v/v for adsorption at 65 bar, which are the highest values of usable CH4 capacity
reported so far for any adsorbent under these conditions. The Co(bdp) usable capacities reported

35
here are a result of the transition from the expanded to the collapsed phase leading to near
complete CH4 desorption by 5.8 bar. For comparison, the highest previously measured 35-bar
and 65-bar usable capacities for any adsorbent are 143 v/v and 189 v/v, as obtained for the
metal–organic frameworks HKUST-1 and UTSA-76a, respectively.

Similarly, Fe(bdp) undergoes a reversible phase transition between a collapsed and expanded
framework. Although the total CH4 uptake is comparable to that of Co(bdp), the adsorption and
desorption steps occur at the considerably higher pressures of 24 bar and 10 bar, respectively,
suggesting that replacing Co with Fe increases the energy of the phase transition. In spite of its
greater expansion and lower crystallographic density, the usable CH4 capacity of Fe(bdp) is still
higher than all known adsorbents at 150 v/v and 190 v/v for 35 bar and 65 bar adsorption,
respectively [19].

Fig. 4.2.4. Total CH4 adsorption isotherms at various temperatures for Fe(bdp) [19]

36
REFERENCES

1. K. R. Matranga, A. L. Myers and E. D. Glandt, 1991, “Storage of Natural Gas By


Adsorption on Activated Carbon”, Chemical Engineering Department, University of
Pennsylvania, Philadelphia, USA
2. J. Wegrzyn, M. Gurevich, 1996, “Adsorbent Storage of Natural Gas”, Applied Energy,
Vol. 55, No. 2, pp 71-83, USA
3. V.C. Menon, S. Komarneni, 1998, Porous Adsorbents for Vehicular Natural Gas Storage:
A Review, Journal of Porous Materials 5, 43–58 (1998), Pennsylvania, USA
4. S. Biloe´, V. Goetz , A. Guillot, 2001, “Optimal design of an activated carbon for an
adsorbed natural gas storage system”, Carbon 40 (2002) 1295–1308, Perpignan, France
5. O. Pupier, V. Goetz, R. Fiscal, 2004, “Effect of cycling operations on an adsorbed natural
gas storage”, Chemical Engineering and Processing 44 (2005) 71–79, Perpignan, France
6. L.L. Vasiliev, L.E. Kanonchik, D.A. Mishkinis, M.I. Rabetsky, 2000, “Adsorbed natural
gas storage and transportation vessels”, Int. J. Therm. Sci. (2000) 39, 1047–1055, Minsk,
Belarus
7. M. J. Prauchner, F. Rodrı´guez-Reinoso, 2007, “Preparation of granular activated carbons
for adsorption of natural gas”, Microporous and Mesoporous Materials 109 (2008) 581–
584, Brazil, Spain
8. R.W. Judd, D.T.M. Gladding, R.C. Hodrien, D. R. Bates, J. P. Ingram, M. Allen, “The
Use of Adsorbed Natural Gas Technology for Large Scale Storage”, Loughborough.UK
9. J. Sun, T. A. Brady and M. J. Rood, M. R.-Abadi and A. A. Lizzio, “Adsorbed Natural
Gas Storage With Activated Carbon”, Illinois, USA
10. K. Inomata*, K. Kanazawa, Y. Urabe, H. Hosono, T. Araki, 2001, “Natural gas storage in
activated carbon pellets without a binder”, Carbon 40 (2002) 87–93, Kanagawa, Japan
11. S. Sircar, T. C. Golden And M. B. Rao, 1995, “Activated Carbon For Gas Separation
And”, Carbon Vol. 34, No. 1, pp. l-12,1996, Allentown, PA, USA
12. J. A. F. MacDonald and D. F. Quinn, 1997, “Carbon adsorbents for natural gas storage”,
Furl Vol. 77, No. 112. pp. 61-64. 1998, Kingston, Canada

37
13. J. S. Im, M. J. Jung, Y.-S. Lee, 2009, “Effects of fluorination modification on pore size
controlled electrospun activated carbon fibers for high capacity methane storage”, Journal
of Colloid and Interface Science 339 (2009) 31–35, Daejeon, Republic of Korea
14. N. Bagheri, J. Abedi, 2011, “Adsorption of methane on corn cobs based activated
carbon”, Chemical Engineering Research and Design 8 9 ( 2 0 1 1 ) 2038–2043, Calgary,
Canada
15. A. M. Rashidi, R. Lotfi, A. Nouralishahi, M. A. Khodagholi,
16. M. Zare, F. Eslamipour, 2011, “Nanoporous carbons as promising novel methane
adsorbents for natural gas technology”, Journal of Natural Gas Chemistry 20(2011)664–
668, Tehran, Iran
17. K. S. N. Kamarudin, H. B. Mat, H. Hamdan, 1998, “Zeolite As Natural Gas Adsorbents”,
Malaysia
18. Z. Song, A. Nambo, K. L. Tate, A. Bao, M. Zhu, J. B. Jasinski, S. J. Zhou, H. S. Meyer,
M. A. Carreon, S. Li, and M. Yu, 2016, “Nanovalved Adsorbents for CH4 Storage”,
Nano Lett. 2016, 16, 3309−3313, USA
19. J. A. Mason, J. Oktawiec, M. K. Taylor, M. R. Hudson, J. Rodriguez, J. E. Bachman, M.
I. Gonzalez, A. Cervellino, A. Guagliardi, C. M. Brown, P. L. Llewellyn, N. Masciocchi
& J. R. Long, 2015, VO L 5 2 7 | N AT U RE | 3 5 7, USA, Italy

38
39

Potrebbero piacerti anche