Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

895

Ethyl Acetate Adsorption onto Activated Carbon


Peter Branton1*, Koki Urita2† and Katsumi Kaneko2† (1) Group Research & Development, British
American Tobacco, Regents Park Road, Millbrook, Southampton SO15 8TL, U.K. (2) Graduate School of Science, Chiba
University, 1-33 Yayoi-cho, Inage-ku, Chiba-shi 263-8522, Japan.

(Received 14 March 2010; revised form accepted 12 January 2011)

ABSTRACT: The adsorption and desorption of ethyl acetate on a coconut-based


activated carbon has been studied using static and flow adsorption methods at
different temperatures from 295 K to 325 K. The adsorbed state of ethyl acetate
was studied using nitrogen adsorption at 77 K after the pre-adsorption of ethyl
acetate. The adsorption isotherms of ethyl acetate were of Type I in the IUPAC
classification, suggesting micropore filling by ethyl acetate molecules. The
micropore volume of the activated carbon as evaluated from ethyl acetate
adsorption using the liquid density of ethyl acetate agreed well with that from
nitrogen adsorption. Nitrogen adsorption after pre-adsorption of ethyl acetate
indicated considerably packed molecular states of ethyl acetate molecules within
the micropores. Thus, ethyl acetate molecules can migrate to the deeper parts of
micropores in the case of static adsorption measurements. However, the specific
surface area evaluated from ethyl acetate adsorption was nearly 40% smaller than
that from nitrogen adsorption under the assumption of an isotropic structure. The
assumption of a highly orientated adsorption structure for ethyl acetate molecules
on the pore wall led to a good correlation even as far as the surface area was
concerned. In addition, 26 kJ/mol of the excess stabilization energy determined
from the temperature dependence of the ethyl acetate adsorption isotherm
supports the favourable molecular alignment of ethyl acetate molecules which
increases the molecule–pore wall interactions.

INTRODUCTION

Activated carbon with a large surface area (> 1000 m2/g) has many applications in industry.
Standard activated carbon contains many slit-shaped micropores which can be used for the
removal and concentration of molecules and ions. In the case of adsorption from solution, even
large planar molecules such as Methylene Blue can be efficiently adsorbed in the micropores
(Lei et al. 2006; Wang et al. 2005; Costas and Snoeyink 2000). In contrast, however, the
adsorption of vapour molecules with a long carbon chain, e.g. nonane, is less efficient since the
molecules cannot diffuse into the deep parts of the micropores (Gregg and Langford 1969).
Reduction of the intra-pore diffusion restriction would thus improve the adsorption characteristics
of activated carbon towards many organic solvent molecules with a relatively large molecular
*
Author to whom all correspondence should be addressed. E-mail: peter_branton@bat.com.
†Present addresses:
Katsumi Kaneko: Research Center for Exotic Nanocarbon, Shinshu University, Wakasato 4-17-1, Nagano-shi, Nagano,
Japan 380–8553.
Koki Urita: Applied Physical Chemistry Laboratory, Department of Applied Chemistry, Faculty of Engineering,
Nagasaki University, 1–14, Bunkyo-machi, Nagasaki-shi, Nagasaki, Japan 852-8521.
896 P. Branton et al./Adsorption Science & Technology Vol. 28 No. 10 2010

size. Consequently, a fundamental study on the adsorption of large molecules in micropores would
be of value.
Activated carbon can be used in cigarette filters, where it exhibits excellent adsorption
properties towards many vapour-phase compounds in cigarette smoke despite the complex
environment involved (Davis and Nielsen 1999). The type of carbon typically used in cigarette
filters is a microporous coconut-based material. The micropore volume is an important parameter
in maximizing the adsorption of selected smoke vapour toxicants in the smoke (Mola et al. 2008).
Hence, it is important to maintain as high an effective surface area as possible. Recent studies have
shown that the addition of mesopores and macropores enhances pore diffusion at the high flow
rates encountered during the smoking process (> 1000 cm3/min) (Branton et al. 2009).
The work described herein examines the adsorption and desorption of a large organic
molecule, ethyl acetate, CH3COOC2H5 [CAS No. 141-78-6] as an example, on microporous
coconut-based activated carbon under a range of conditions. Ethyl acetate is a sweet smelling,
widely used solvent with a boiling point of 340 K and a vapour pressure of ca. 10.7 mPa at 295 K
(Lide 2008). Modelling of ethyl acetate adsorption has previously been conducted by Manjare
and Ghoshall on carbon (Manjare and Ghoshal 2006) and on molecular sieves (Manjare and
Ghoshal 2008).

EXPERIMENTAL

Coconut-based activated carbon (Ecosorb CX) was sourced from Jacobi Carbon. This carbon was
pre-treated (2 h at 393 K at a vacuum of 11 mPa) to remove any water adsorbed in the pores. The
dried material was then weighed at ambient temperature. Liquid ethyl acetate was added dropwise
onto the dry carbon under a nitrogen (i.e. dry) atmosphere to the desired loading, the weight of the
carbon being monitored over a period of several days. The amounts of ethyl acetate desorbed were
also determined using Thermal Gravimetric Analysis (TGA) under nitrogen at a flow rate of
100 cm3/min. As a control, activated carbon in the absence of ethyl acetate was also analyzed.
Ecosorb CX activated carbon (base carbon) was saturated with ethyl acetate under the
following conditions:
1. The as-received carbon was refluxed with ethyl acetate (1:40 ratio by weight) for 24 h at
373 K without stirring.
2. The as-received carbon was pre-treated at 393 K for 2 h under a vacuum of 0.7 mPa. This
was then immersed in liquid ethyl acetate at 295 K for 24 h (1:40 ratio by weight) without
stirring.
3. The as-received carbon was immersed in liquid ethyl acetate at 295 K for 24 h (1:40 ratio
by weight) without stirring.
4. The as-received carbon was immersed in liquid ethyl acetate at 295 K for 24 h (1:40 ratio
by weight) with stirring.
Nitrogen adsorption isotherms were measured volumetrically at 77 K using a Quantachrome
Autosorb I instrument, with the carbon sample being outgassed at 393 K and < 0.8 mPa for 2 h to
remove any water in the pores prior to isotherm measurements. To investigate any carbon surface
area still available following ethyl acetate sorption, the carbon was exposed to nitrogen gas at
77 K and evacuated to < 0.8 mPa at 77 K (thus avoiding any desorption of ethyl acetate), and
nitrogen sorption subsequently measured at 77 K. To remove the adsorbed ethyl acetate from the
Ethyl Acetate Adsorption onto Activated Carbon 897

carbon and to re-measure the surface area, the carbon was outgassed at 573 K and < 0.8 mPa for
2 h prior to nitrogen sorption at 77 K.
TGA measurements were made using a Shimadzu DTG-60AHY instrument. Desorption of
ethyl acetate from carbon was measured by heating from 303 K to 873 K under nitrogen gas at a
flow rate of 100 cm3/min and a heating rate of 3 K/min.
Ethyl acetate sorption isotherms were measured using a static gravimetric system (Hiden
Isochema IGA-002) at temperatures of 295, 310 and 327 K, respectively. Carbon outgassing was
carried out at a temperature of 523 K and a pressure < 0.01 mPa for a minimum of 4 h until a
stable weight loss was achieved.
Ethyl acetate sorption isotherms were measured using a dynamic gravimetric system (SMS
DVS-1) at temperatures of 288 K and 295 K, respectively. The carbon was dried in flowing air to
constant weight prior to isotherm determinations.

RESULTS AND DISCUSSION

The nitrogen adsorption isotherms at 77 K on the base activated carbon were of Type I in the
IUPAC classification. The pore structural parameters measured are listed in Table 1.

TABLE 1. Characteristics of Ecosorb


CX Activated Carbon

Characteristic Value

Vp (total) 0.40 cm3/g


Vp (meso) 0.02 cm3/g
Vp (meso)b 0.03 cm3/g
Vp (micro)a 0.37 cm3/g
Vp (micro)b 0.37 cm3/g
SBET 980 m2/g
a
Obtained from the DR plot. bObtained from
the αS-plot.

Liquid ethyl acetate was added to the dry carbon sample as described previously. Desorption of
the ethyl acetate resulted in a constant weight (equilibrium achieved) after ca. 7 d irrespective of
the initial loading, as shown in Figure 1. The amount of ethyl acetate adsorbed at equilibrium was
ca. 4–6% (as determined by weighing the carbon at set time intervals and also by undertaking
TGA on the carbon containing ethyl acetate). As a control, activated carbon in the absence of ethyl
acetate was analyzed, where no measurable uptake was recorded.
The density of ethyl acetate at 295 K is 0.9003 g/cm3. By using the total pore volume listed in
Table 1 and assuming that all the pores are accessible to ethyl acetate, a theoretical maximum
loading of 0.36 g of ethyl acetate/g carbon can be calculated.
Ecosorb CX carbon was saturated with ethyl acetate as previously described. The TGA curves
(using a nitrogen flow rate of 100 cm3/min and a heating rate of 3 K/min) for the carbon treated
samples (a) to (d) are shown in Figure 2 where in (a) the as-received carbon was refluxed with
ethyl acetate at 373 K, in (b) the pre-treated carbon was immersed in liquid ethyl acetate at 295 K
898 P. Branton et al./Adsorption Science & Technology Vol. 28 No. 10 2010

25

Ethyl acetate (wt%)


20

15

10

0
0 5 10 15 20

Time (d)

Figure 1. Desorption of ethyl acetate from activated carbon as a function of time. Data points relate to the following initial
loadings of ethyl acetate: , 15%; , 23%.

100

95

90
TG (wt%)

85

80

75

70
300 500 700 900
Temperature (K)

Figure 2. TG analysis of ethyl acetate adsorbed onto activated carbon under the different saturation conditions (a) to (d)
described in the text. The various traces correspond to the following conditions: , initial base carbon; , condition
(a); , condition (b); , condition (c); , conditions (d).

without stirring, in (c) the as-received carbon was immersed in liquid ethyl acetate at 295 K for
24 h without stirring and in (d) the as-received carbon was immersed in liquid ethyl acetate at
295 K for 24 h with stirring.
The boiling point of ethyl acetate is 340 K. The total ethyl acetate content in all the samples
was ca. 20%. The initial 5% weight loss seen in Figure 2 was due to the loss of excess ethyl acetate
at 303 K (TGA measurements commenced after the balance weight had stabilized in the nitrogen
gas flow). Interestingly, the TG trace for the carbon refluxed with ethyl acetate exhibited two
steps. This trace was repeatable and possibly arose from the presence of ethyl acetate hydrogen-
bonded with surface hydroxy groups on the pore walls, the first and second steps resulting from
physically adsorbed ethyl acetate and ethyl acetate hydrogen-bonded with the surface hydroxy
groups, respectively.
Nitrogen adsorption isotherms at 77 K were measured on (i) the base carbon (following
outgassing at 393 K for 2 h at a vacuum of 0.8 mPa), (ii) the carbon with adsorbed ethyl acetate
(the carbon was outgassed at 393 K for 2 h followed by ethyl acetate addition in vacuum, with
5 cm3 of adsorbate being added to 103 mg of carbon; prior to nitrogen adsorption, the sample was
outgassed at 77 K and a pressure of 0.8 mPa) and (iii) the same carbon where the ethyl acetate had
been completely removed from the pores (via outgassing at 573 K for 2 h at a vacuum of 0.8 mPa).
The corresponding isotherms are depicted in Figure 3.
Ethyl Acetate Adsorption onto Activated Carbon 899

500

Nitrogen uptake (mg/g)


400

300

200

100

0
0 0.2 0.4 0.6 0.8 1

P/P0

Figure 3. Nitrogen adsorption isotherms at 77 K on microporous activated carbon with and without pre-adsorbed ethyl
acetate. The data points depicted correspond to the following: , base carbon; , following ethyl acetate adsorption; ,
following ethyl acetate removal from the pores by heating in vacuum.

Adsorption of ethyl acetate caused a reduction in micropore volume (as determined from the
DR plots) of 0.35 cm3/g, i.e. 95% of the pores were filled with ethyl acetate [(i) → (ii)]. The
micropore volume following (i) ethyl acetate adsorption and (ii) removal via outgassing with heat
and vacuum showed a slight increase in micropore volume (again as determined from the DR
plots) from 0.37 to 0.46 cm3/g [(i) → (iii)].
Micropore swelling is possible in the case of strong interactions between the molecules and the
pores (Kaneko et al. 1989; Suzuki and Kaneko 1990). The average pore width (w0) can be
determined using the Dubinin–Stoeckli (DS) equation (Stoeckli et al. 1993):

10.8
w0 = (nm )
E 0 − 11.4

where E0 is a characteristic adsorption energy determined from the slope of the


Dubinin–Radushkevich profile. Although the micropore volumes of the base carbon and the
carbon previously used for ethyl acetate adsorption were different, both pore widths determined
from the DS equation were 1.0 nm. This activated carbon probably had a bottle-neck structure in
part, which induced an entrance blocking effect towards nitrogen adsorption at 77 K. Adsorption
and desorption treatment of the activated carbon could have widened the partially necked
structures.
Ethyl acetate isotherms were measured using both static and dynamic measurement techniques.
Adsorption isotherms measured using the static method at temperatures of 295, 310 and 325 K,
respectively, are shown in Figure 4. The carbon was outgassed at 523 K at 0.01 mPa to constant
weight prior to isotherm measurements. All the isotherms are of Type I in the IUPAC classification,
with a slight increase in adsorbed weight with decreasing temperature.
Adsorption isotherms measured using a dynamic method at 288 and 295 K are shown in
Figure 5. All the samples studied were dried in flowing air to constant weight prior to isotherm
measurement. However, using the dynamic technique, the vapour pressure of ethyl acetate limited
isotherm measurements to temperatures of 295 K and below.
As can be seen, all the isotherms are Type I in shape indicative of micropore filling, with sharp
“knees” corresponding to strong carbon–ethyl acetate interactions. In both cases, slightly lower
amounts of ethyl acetate were adsorbed as the adsorption temperature increased, indicating that
900 P. Branton et al./Adsorption Science & Technology Vol. 28 No. 10 2010

40
35

Ethyl acetate uptake (%)


30
25

20
15
10
5

0
0 0.2 0.4 0.6 0.8 1

P/P0

Figure 4. Ethyl acetate adsorption onto activated carbon at () 295 K, () 310 K and () 325 K employing the static
measurement technique. Open symbols () denote adsorption; filled symbols ( ) denote desorption. The
addition of shading denotes different runs.

35
Ethyl acetate uptake (%)

30
25
20
15
10
5
0
0 20 40 60 80 100
Relative concentration (%)

Figure 5. Ethyl acetate adsorption onto activated carbon at () 288 K and () 295 K employing the dynamic
measurement technique. Open symbols () denote adsorption; filled symbols () denote desorption.

the adsorption process was physical in nature. The dynamic isotherms were not reversible; the
amount of ethyl acetate that could not be desorbed at both temperatures was ca. 21% by weight
of carbon (or 73% of the total pore volume available for ethyl acetate), again demonstrating strong
interactions between the adsorbate and the adsorbent, and the fact that thermodynamic
equilibrium had not been established. The “knee” of the dynamic isotherm was sharper at lower
temperature, suggesting stronger interactions with the micropore walls.
The BET surface areas (taking the cross-sectional area of the ethyl acetate molecule to be
32 Å2) and the pore volumes (taken at a relative pressure of 0.95) are listed in Table 2. It will be
seen from Table 2 that the pore volumes determined statically using nitrogen and ethyl acetate as
the probe molecules showed good agreement. Evaluations of the pore volumes assumed the
presence of a liquid adsorptive; hence, it may be concluded that, at temperatures between 295 K
and 325 K, ethyl acetate molecules should have a liquid-like structure in the carbon micropores.
However, the surface areas were very different; a value of 32 Å2 was used based on the density of
the adsorbed layer and assuming a spherical shape and hexagonal close-packing (McClellan and
Harnsberger 1967). Ethyl acetate molecules should be orientated on the micropore walls with
sufficient two-dimensional translational motion to give an effective molecular area of 49 Å2. This
Ethyl Acetate Adsorption onto Activated Carbon 901

TABLE 2. Carbon Characteristics Derived from Adsorption Isotherms

Adsorbate Isotherm determination Temperature (K) SBET Vp (total)a


technique (m2/g) (cm3/g)

Nitrogen Static 77 980 0.40


Ethyl acetate Static 295 600 0.40
Ethyl acetate Static 310 630 0.39
Ethyl acetate Static 325 630 0.39a
Ethyl acetate Dynamic 288 420 0.33
Ethyl acetate Dynamic 295 440 0.32
a
Taken at P/P0 = 0.8.

effective molecular area is close to the value obtained by multiplying 32 Å2 by the ratio (980/600)
of the surface areas observed by nitrogen and ethyl acetate adsorption.
The dynamically measured pore volumes and surface areas were smaller, with the BET surface
area of only 420 m2/g being considerably lower than the nitrogen BET surface area of 980 m2/g
at 77 K. Similarly, the total pore volume of 0.33 cm3/g at a relative pressure P/P0 = 0.95 was lower
than that obtained from nitrogen adsorption (0.40 cm3/g). This was probably due, in part, to the
entrance-blocking effect suggested above, since ethyl acetate molecules near the pore mouth
would be highly stabilized and hence block the pore.
By measuring isotherms at two or more temperatures, it is possible to determine the isosteric
heat of adsorption, qst. The isosteric heats of adsorption were determined from the statically
measured isotherms at 295, 310 and 325 K. However, due to the sharp isotherm knees at low
relative pressures, heats could not be accurately determined at low surface coverages, i.e. below
ca. 0.8 of a monolayer. The linearity of the Clausius–Clapeyron plot was poor, giving relatively
large error bars as shown in Figure 6 as a function of the surface coverage, n/nm, where the
monolayer coverage, nm, was determined from the BET analysis.

70

60
qst (kJ/mol)

50

40

30
0.80 0.85 0.90 0.95 1.00 1.05

n/nm

Figure 6. Isosteric heat of adsorption for ethyl acetate onto activated carbon.
902 P. Branton et al./Adsorption Science & Technology Vol. 28 No. 10 2010

Over the observed coverage region, the isosteric heat of adsorption was 58 ± 8 kJ/mol. The heat
of vaporization of ethyl acetate is 36 kJ/mol at 298 K and 32 kJ/mol at 350 K (Lide 2008); hence
the magnitude of the isosteric heat provides clear evidence that the adsorption process was
physical rather than chemical in nature. The additional stabilization energy of
ca. 24 kJ/mol arises from the overlapped interaction potentials from the pore walls. In order to
obtain such a high stabilization energy, the ethyl acetate molecules should be orientated on the
pore wall as suggested above. A classical study of the adsorption of linear alkanes onto graphite
(Clint 1972) showed that even C5-alkane molecules are aligned in a flat orientation. In addition,
according to McClellan and Harnsberger (1967), the cross-sectional area of n-butane (C4H10) is 44
Å2. Hence, the effective molecular area of ethyl acetate (CH3CO2C2H5) of 49 Å2, estimated under
the assumption of a flat orientation on the pore wall, is slightly larger than that of the n-butane
molecule. It would therefore appear that the value of 49 Å2 for the effective molecular area of
ethyl acetate is reasonable.

CONCLUSIONS

The adsorption isotherm of ethyl acetate on coconut-based activated carbon has been shown to be
similar to the nitrogen adsorption isotherm, i.e. Type I in shape, and hence ethyl acetate molecules
are adsorbed via a pore-filling mechanism. This mechanism was also supported by the additional
stabilization energy of ethyl acetate of ca. 24 kJ/mol obtained from measurements of the isosteric
heat of adsorption. In addition, nitrogen adsorption of ethyl acetate pre-adsorbed onto carbon at
77 K also supported the filling mechanism, although the presence of partial bottleneck structures
needs to be considered. A comparative analysis of the effective surface area measurement using
nitrogen at 77 K and ethyl acetate at near-ambient temperature suggested a flat orientation for
ethyl acetate molecules on the micropore walls, giving a value of 49 Å2 for the effective area of
ethyl acetate molecules in the micropores of the activated carbon.

REFERENCES

Branton, P., Lu, A. and Schüth, F. (2009) Carbon 47, 1005.


Clint, H. (1972) J. Chem. Soc., Faraday Trans. I 68, 2239.
Costas, P. and Snoeyink, V.L. (2000) Carbon 38, 1423.
Davis, D.L. and Nielsen, M.T. (1999) Tobacco Production, Chemistry and Technology, Blackwell Science,
Oxford, U.K., Chap. 11B.
Gregg, S.J. and Langford, J.F. (1969) Trans. Faraday Soc. 65, 1394.
Kaneko, K., Fujiwara, Y. and Nishikawa, K. (1989) J. Colloid Interface Sci. 127, 298.
Lei, S., Miyamoto, J., Kanoh, H. and Kaneko, K. (2006) Carbon 44, 1884.
Lide, D.R. (Editor-in-chief) (2008) CRC Handbook of Chemistry and Physics, 89th Edn, CRC Press, Boca
Raton, FL, U.S.A.
Manjare, S.D. and Ghoshal, A.K. (2006) Ind. Eng. Chem., Res. 45, 6563.
Manjare, S.D. and Ghoshal, A.K. (2008) Can. J. Chem. Eng. 83, 232.
McClellan, A.L. and Harnsberger, H.F. (1967) J. Colloid Interface Sci. 23, 577.
Mola, M., Hallum, M. and Branton, P. (2008) Adsorption 14, 335.
Stoeckli, F., Huguenin, D. and Greppi, A. (1993) J. Chem. Soc., Faraday Trans. 89, 2055.
Suzuki, T. and Kaneko, K. (1990) J. Colloid Interface Sci. 138, 590.
Wang, S.B., Zhu, Z.H., Coomes, A., Haghseresht, F. and Lu, G.Q.J. (2005) J. Colloid Interface Sci. 284, 440.

You might also like