Chitrakshi Goel,
Haripada Bhunia and
Pramod K. Bajpai*
Department of Chemical Engineering, Thapar University, Patiala-147004, Punjab, India. E-mail: chitrakshigoel@yahoo.co.in; hbhunia@thapar.edu; pkbajpai@thapar.edu; Fax: +91-175-2393005; Tel: +91-175-2393438
First published on 26th October 2015
Silica templated nanostructured carbons were developed from a resorcinol–formaldehyde polymeric precursor by varying the carbonization temperature from 400 °C to 800 °C. The prepared carbons were characterized thoroughly for their textural, surface and chemical properties followed by dynamic CO2 capture performance at various adsorption temperatures from 30 °C to 100 °C under simulated flue gas conditions. Among the prepared carbons, carbonization at 700 °C resulted in the nanostructured carbon material, as indicated by XRD and TEM results, having the best textural properties i.e. specific surface area and total pore volume around 435 m2 g−1 and 0.22 cm3 g−1, respectively. The sample obtained by carbonization at the most severe conditions (≥800 °C) exhibited textural properties comparable to that of RF-700 but showed lower CO2 adsorption capacity on account of reduction in surface basicity at higher temperatures. On the other hand, preparation of the carbon material by direct carbonization of the polymeric precursor, i.e. without using a template, resulted in a completely non-porous material with very low CO2 adsorption capacity. Moreover, both the textural properties and the surface chemistry had an effect on the CO2 adsorption performance of the prepared carbons. RF-700 exhibited the highest dynamic CO2 adsorption capacity of 0.761 mmol g−1 at 30 °C in a binary mixture of 12.5% CO2 in N2 attributed to a well-developed porous structure and high surface basicity of 1.93 meq g−1. It also demonstrated high selectivity towards CO2 over N2 and stable adsorption capacity over multiple adsorption–desorption cycles. CO2 adsorption on the prepared carbons was well described by a fractional order kinetic model. Fitting of equilibrium data of CO2 adsorption by Temkin isotherm model and variation in isosteric heat of adsorption with surface coverage indicated an energetically heterogeneous adsorbent surface. Thermodynamics of CO2 adsorption on the carbon material suggested an exothermic, random and spontaneous nature of the process. The thermal energy required for desorption of CO2 was also estimated to be around 1.9 MJ per kg CO2.
Adsorption using porous adsorbent materials presents various advantages over other methods like low operating cost and ease of application.6 Adsorbents being used for CO2 capture include porous carbons, zeolites, silica and metal–organic frameworks.7–12 Zeolite and silica adsorbents are also treated with various amine solutions in order to improve their affinity towards CO2 and hence the adsorption capacity.13,14 But the toxic and corrosive amine solutions need high regeneration temperature and degrade over various adsorption cycles.9,11,15 On the other hand, carbon based adsorbents are the versatile adsorbents being widely used in gas and liquid adsorption on account of their properties such as high specific surface area, hydrophobicity, complete regeneration over multiple adsorption cycles, high mechanical and thermal stability.16 They can be produced from a wide range of low cost materials like coal, industrial by-products, wood or biomass sources etc.17,18 by using various methods including sol–gel process, chemical vapour deposition, carbonization of carbon containing precursor and nanocasting. Among these available techniques, nanocasting technique has been used in recent times for the fabrication of carbon materials with controlled pore structure.19 A carbon precursor is generally infiltrated into the pore of a hard template and is treated thermally under a controlled atmosphere. This is followed by template removal by dissolution of template in hydrogen fluoride or sodium hydroxide solution.20 Porous carbons were obtained from nanocasting technique by using sucrose as the carbon source and nano-CaCO3 as hard template.21 Direct pyrolysis of resorcinol and formaldehyde in the presence of lysine as catalyst produced carbons having static CO2 uptake of 3.13 mmol g−1 at 25 °C under 100% CO2 flow.22 Melamine–formaldehyde and urea–formaldehyde based carbons by chemical activation with K2CO3 exhibited CO2 uptake of 1.03 mmol g−1 and 1.8 mmol g−1 at 25 °C.23 Phenolic resin and resorcinol resin based carbons were subjected to chemical activation with KOH in order to improve their textural properties.24 These materials exhibited CO2 uptake of ca. 4.36 mmol g−1 at 25 °C under static conditions.
It can be seen that most of the carbon synthesis is carried out by direct carbonization of various carbon sources and/or followed by post-synthesis treatments such as physical activation or chemical activation. But development of porous carbons by nanocasting technique helps in tailoring the pore structure of the resultant material. Also, activation is done to improve the textural properties of the obtained carbons in spite of the high energy requirement in these processes. Furthermore, these materials are evaluated for CO2 capture under pure CO2 flow at static conditions mainly at 0 °C or 30 °C and adsorption capacities evaluated under static conditions are always higher than the values obtained in dynamic systems. It is worth mentioning that for practical applications in the field of CCS, the developed adsorbents should be evaluated at high temperatures under dynamic conditions. Also, CO2 adsorption by carbon materials have been reported to strongly depend on both the textural properties and surface chemistry of the materials. Thus, it is important to develop porous carbon adsorbents with optimal textural properties in addition to highly basic character in order to achieve higher affinity towards CO2 than N2.
Present study focuses on the development of nanostructured carbons from resorcinol–formaldehyde polymeric precursor by nanocasting technique followed by thorough characterization by using various sophisticated techniques and evaluation for CO2 adsorption by fixed-bed experiments. Effect of silica template on the physico-chemical properties of the prepared carbons was also evaluated by preparing one sample by direct carbonization of RF resin and was compared with nanocasted carbons. Kinetics, isotherm and thermodynamics for CO2 adsorption on these carbons were also studied. In addition, energy duty for desorption of adsorbed CO2 was also calculated.
A series of adsorbents were prepared by carbonizing the templated resin samples at varying carbonization temperatures ranging from 400 °C to 800 °C for 1 h in N2 flow with a heating rate of 10 °C min−1. Silica template from the obtained materials was removed by dissolution in NaOH solution for at least 24 h followed by washing with deionized water and drying at 100 °C for 2 h. Synthesized carbon materials were marked as RF-x, where x stands for temperature of carbonization (400–800 °C). One more carbon sample was synthesized by direct carbonization of RF resin at 700 °C (based on optimized conditions for RF based carbons) in order to investigate the effect of silica template on the physico-chemical properties of the prepared carbons and this sample was denoted as ‘RF-C’.
The X-ray diffraction (XRD) patterns of the synthesized materials were acquired on a PANalytical X'Pert Pro diffractometer operating at 45 kV and 40 mA with Cu-Kα radiation in the scanning range (2θ) of 10 to 80°. Transmission electron microscopy (TEM) images were obtained with a Philips CM200 transmission electron microscope operating at an accelerating voltage of 200 kV. Scanning electron microscopy (SEM) images were recorded on a JEOL JSM-6510 LV scanning electron microscope with an accelerating voltage of 10.0 kV. The samples were coated with a gold film of 50 μm thickness in an automatic sputter coater (Polaron) in order to avoid charging under electron beam.
Thermogravimetric analysis of the samples was carried out using a TA Q500 (TA Instrument, USA) thermogravimetric analyzer in nitrogen at a flow rate of 50 ml min−1. The samples were heated at a heating rate of 10 °C min−1 from 30 °C to 900 °C.
Fourier transform infrared (FTIR) spectra of carbon samples were obtained using a PerkinElmer Spectrum 100 FTIR spectrometer (PerkinElmer, USA). The spectra were recorded from 4000 to 625 cm−1 at a resolution of 4 cm−1. Elemental analysis was carried out using a Thermo Finnigan Flash EA 1112 Series elemental analyzer. X-ray photoelectron spectroscopy (XPS) analysis of porous carbons was performed on a SPECS system operating at 15 kV and 10 mA. The pressure in the analysis chamber was less than 2 × 10−9 torr. The data processing was carried out with XPS peak 4.1 software and the core level spectra were fitted with mixed Gaussian–Lorentzian convoluted function (80/20) and Shirley function was used for background subtraction.
Boehm titration method25 was employed to quantify the surface functional groups of the prepared carbons. About 200 mg of the samples was immersed in 20 ml of 0.1 M solution of NaOH and HCl and stirred for 24 h. The suspension was filtered and the excess base and acid were titrated with HCl and NaOH respectively. Amount of acid and base that reacted with the carbon sample gives the number of basic and acidic sites respectively.
![]() | (1) |
Carbon dioxide temperature programmed desorption (CO2-TPD) study of RF-700 was carried out on a Micromeritics AutoChem II 2920 chemisorption analyzer equipped with a thermal conductivity detector. Around 100 mg of the carbon sample was placed in a quartz reactor and pretreated under flow of pure helium gas at 200 °C. Hereafter the temperature was decreased to 30 °C and CO2 adsorption was carried out by switching to pure CO2 gas for 30 minutes. Desorption experiment was performed by switching back to He gas at a flow rate of 20 ml min−1 and increasing the temperature from 45 °C to 250 °C at a heating rate of 10 °C min−1.
![]() | (2) |
qt = qe(1 − e−k1t) | (3) |
The pseudo-second order model can be written as:28
![]() | (4) |
![]() | (5) |
The fractional order kinetic model assumes adsorption rate to be affected by nth power of driving force and mth power of the adsorption time29,30 and is expressed as:
![]() | (6) |
![]() | (7) |
Based on normal standard deviation, an error function is calculated to ascertain the sufficiency of each model.
![]() | (8) |
![]() | (9) |
Freundlich isotherm explains the non-ideal and reversible adsorption on energetically heterogeneous surface with non-uniform distribution of adsorption heat and affinities over the surface and is not restricted to monolayer adsorption.32 It can be written as:
qe = KFP1/n | (10) |
Temkin isotherm describes the adsorption on heterogeneous surfaces with the assumption that adsorption heat of the molecules in the layer decreases linearly rather than logarithmically with coverage because of adsorbent–adsorbate interactions. Its derivation is characterized by a uniform distribution of binding energy.33 It can be written as:
qe = B![]() | (11) |
Parameters of the above adsorption kinetic and isotherm models were obtained by nonlinear regression of the experimental data with the aid of OriginPro 8 software.
k = Ae−(Ea/RT) | (12) |
The Gibbs free energy change for the adsorption process can be obtained from the adsorption equilibrium constant by using the following equation:
ΔG° = −RT![]() | (13) |
The classical van't Hoff equation relates the equilibrium constant with temperature according to the following equation:
![]() | (14) |
![]() | (15) |
ΔG° = ΔH° − TΔS° | (16) |
From eqn (13) and (16), following relation is derived with the assumption that both ΔH° and ΔS° change slightly with temperature:34
![]() | (17) |
The thermal energy input (Q) required to regenerate the adsorbent is the sum of heat of desorption (Qst) and the sensible heat required to heat the adsorbent from the adsorption temperature up to temperature for desorption.
Q = Qst + sensible heat | (18) |
Sensible heat requirement is a function of adsorbent's heat capacity, difference between adsorption and desorption temperatures and the adsorption capacity of the adsorbent material.35,36
![]() | (19) |
Heat of desorption is assumed to be equal to heat of adsorption.37 The isosteric heat of adsorption can be calculated from CO2 adsorption isotherms obtained at different temperatures by using Clausius–Clapeyron equation:38
![]() | (20) |
![]() | ||
Fig. 2 N2 adsorption (closed symbols) and desorption (open symbols) isotherms at −196 °C of nanostructured carbons. |
As seen in Table 1, BET surface area, total pore volume, micropore volume and mesopore volume tend to increase with increase in carbonization temperature from 400 °C to 700 °C but further increase in carbonization temperature to 800 °C lead to decrease in the textural properties. Micropores contributed largely towards the total pore volume for samples obtained at carbonization temperature ≥600 °C. BET surface area and total pore volume of RF-700 were found to be 435 m2 g−1 and 0.22 cm3 g−1 with ca. 67% contribution from micropores (Vmicro = 0.148 cm3 g−1) respectively. RF-800 sample demonstrated a very small hysteresis loop indicating mainly the presence of micropores in this material and the same can be inferred from very small amount of mesopore volume as compared to micropore volume. RF-400 and RF-500 consisted mainly of mesopores having majority of the pore volume in pores of diameter above 10 nm while rest of the samples showed a well-developed PSD (<10 nm). On the other hand, for the carbon derived from RF resin without nanocasting technique, almost no adsorption of N2 was observed at liquid N2 temperature indicating this carbon to be completely non-porous material. This signifies that the carbons obtained from the same starting material under same conditions can range from completely non-porous material to highly porous material depending on the method employed for their synthesis.
Sample | Textural parameters | Structural parameters | Elemental composition | Surface properties | |||||||
---|---|---|---|---|---|---|---|---|---|---|---|
SBET (m2 g−1) | VP (cm3 g−1) | Vmicro (cm3 g−1) | Vmeso (cm3 g−1) | 2θ | d002 (nm) | C (%) | H (%) | O (%) | Acidity (meq g−1) | Basicity (meq g−1) | |
a SBET: BET surface area; VP: total pore volume obtained at a relative pressure of 0.99; Vmicro: micropore volume obtained from difference between total pore volume and mesopore volume; Vmeso: mesopore volume obtained from BJH method using adsorption branch; 2θ: diffraction angle; d002: interlayer d-spacing of (002) diffraction plane. | |||||||||||
RF-400 | 27 | 0.031 | 0.012 | 0.019 | 23.04 | 0.385 | 68.28 | 3.57 | 28.13 | 1.67 | 0.54 |
RF-500 | 58 | 0.085 | 0.020 | 0.065 | 22.55 | 0.394 | 65.11 | 3.01 | 31.87 | 1.48 | 0.81 |
RF-600 | 369 | 0.180 | 0.128 | 0.052 | 22.77 | 0.390 | 62.02 | 2.11 | 35.87 | 1.24 | 1.75 |
RF-700 | 435 | 0.222 | 0.148 | 0.074 | 23.03 | 0.386 | 66.25 | 1.49 | 32.23 | 0.78 | 1.93 |
RF-800 | 407 | 0.176 | 0.146 | 0.030 | 23.70 | 0.375 | 70.02 | 0.97 | 28.99 | 1.09 | 0.78 |
RF-C | — | — | — | — | 24.12 | 0.368 | 83.66 | 1.49 | 14.83 | 0.85 | 0.88 |
Fig. 3 represents the powder XRD patterns of the synthesized carbon materials. All the samples show similar diffraction patterns with two broad peaks at ca. 23° and 43° corresponding to (002) and (100) diffraction planes respectively of graphitic carbon. Table 1 presents the 2θ values for (002) diffraction plane along with interlayer spacing (d002) values. Observed interlayer d-spacing values are larger than value for ideal graphite (d002 = 0.335 nm)39 indicating the presence of turbostratic structures in the carbon i.e. formation of fully disordered structures.40 Increase in d-spacing values leads to shifting of XRD peaks to lower angles. Carbonization up to 500 °C resulted in decrease in 2θ value and an increase in the d-spacing while further increase in carbonization temperature reduced the d002 values suggesting the formation of short ordered structures and more graphitic character at higher temperatures. The irregularity of the layer structures also increased as indicated by a decrease in the intensity of (002) diffraction plane. RF-C carbon sample exhibited very broad diffraction peak suggesting this carbon to be an amorphous material.
The synthesized carbons are nano-materials as suggested by the broadened diffraction peaks, which is due to very small crystallite size. Moreover, broadened peaks suggest the formation of amorphous carbon materials and these results are in good agreement with TEM results.
Transmission electron micrographs of prepared carbons can be seen in Fig. 4. Development of nanostructures can be clearly observed from the TEM images along with amorphous porous structure. But there is no such development of nanostructures in the carbon sample prepared without nanocasting technique thereby signifying the role of silica template in the development process of carbon materials.
SEM micrographs (Fig. S2†) of obtained carbons show the development of irregular pores on the surface that assist the adsorbate diffusion from bulk phase to the surface. Fine needle like structures are observed for carbons obtained from carbonization at or above 700 °C. However, RF-C sample is found to have very compact surface with very less porous structure.
All the prepared carbon samples show a weight loss of ca. 6–8% (Fig. S3†) up to temperature of 100 °C due to desorption of moisture and other adsorbed gases. There is an increase in the thermal stability of the materials with increase in the carbonization temperature with maximum stability being exhibited by RF-800 (up to 500 °C). On the other hand, among the prepared carbons RF-400 and RF-500 are the least thermally stable materials and show a maximum weight loss of ca. 50%. RF-700 and RF-800 exhibit maximum weight loss of 25% and 15% respectively even up to temperature of 900 °C. Carbon material obtained by direct carbonization of RF resin showed ca. 2 wt% loss up to 100 °C but exhibited similar thermal stability as that of RF-700 carbon.
Fig. S4† depicts the FTIR spectra of the porous carbons derived from RF resin at various carbonization temperatures. Small peak at 1056 cm−1 can be assigned to stretching vibration of alkoxy C–O bond while a peak at 1250 cm−1 corresponds to acyl or phenyl C–O bond. A band around 1540 cm−1 is observed attributing to stretching vibrations of CC bond of aromatic rings of quinone group. Peaks at 1770 cm−1 and 3000 cm−1 correspond to C
O stretching vibration in carboxylic groups, anhydrides, lactones, etc. and C–H stretching vibration respectively. With increase in carbonization temperature, there is a decrease in the intensity of the band around 1540 cm−1 suggesting loss of C
C bonds present in the samples. This may be due to a decrease in the carbon content of the prepared carbons with increase in carbonization temperature. Similar results are obtained from elemental analysis of the carbons as shown in Table 1. There is a decrease in carbon content and an increase in oxygen content of the prepared carbons with increase in temperature of carbonization up to 600 °C. This could be due to cleavage of more carbon–carbon bonds than oxygen functionalities. On the other hand, further increase in carbonization temperature resulted in the opposite trend which is attributed to decomposition of oxygen containing groups at high temperatures of pyrolysis. In case of RF-C carbon, peaks for C–O and C
O bond are very small as compared to other samples, demonstrating lesser oxygen functionalities which is in good accord with the elemental analysis. Majorly a broad band ca. 1540 cm−1 for stretching vibrations of C
C bond is observed. This sample had very high carbon content and low oxygen content as compared to that of RF-700 indicating that presence of silica template effected the chemical composition of the obtained materials.
X-ray photoelectron spectroscopy (XPS) analysis was carried out for all the synthesized carbons to study the evolution of oxygen functional groups during carbonization process. Two typical peaks for carbon (C1s) and oxygen (O1s) are observed, around 285 and 532 eV respectively, in the survey spectra of the synthesized carbons (Fig. 5a). The absence of peak for Si suggests removal of silica via dissolution in NaOH solution.
![]() | ||
Fig. 5 (a) Survey scan spectra of prepared carbons, and deconvoluted C1s spectra of (b) RF-400, (c) RF-500, (d) RF-600, (e) RF-700, (f) RF-800, and (g) RF-C. |
Fig. 5b–g show the C1s XPS spectra of nanostructured carbons and RF-C carbon. The C1s spectra of all the samples were deconvoluted into four well resolved peaks and their corresponding binding energy (B.E.), full width at half maximum (FWHM) and relative area contribution (A%) are reported in Table S1.† Peaks C1 and C2 are assigned for graphitic carbon and carbon in alcohol, phenol or ether groups respectively. Peak C3 corresponds to carbonyl or quinone groups and C4 corresponds to carboxyl, ester and/or lactone linkages.41 With increase in carbonization temperature from 400 °C to 500 °C, there is a drastic decrease in the relative area percentage of peak C1 from ca. 60% to 35.7% indicating decreased graphitic character, which has also been indicated by XRD results. Further increase in carbonization temperature resulted in increase in the relative area percentage of C1 peak and hence improvement in the graphitic character of the porous carbons. For RF-800, C1 peak has small area percentage of ca. 17.7% with the development of a new peak at ca. 283.4 eV for carbidic carbon. After C1 peak, relative area percentage is the maximum for peak C2 for all the prepared carbons suggesting that most of the carbon is linked to one oxygen atom by single bond. For carbons obtained at carbonization temperature ≥700 °C, no peak is observed at ca. 286.3 eV indicating the absence of carbonyl and/or quinone groups in these samples. In case of RF-C carbon, C1 and C2 peaks have similar relative percentage area as that of RF-700. But it contained carbon double bonded to oxygen in form of both carbonyl/quinone (ca. 15.3%) and carboxyl/ester (ca. 7.13%) unlike from RF-700. A considerable change is observed in C1s core level spectra of synthesized samples with carbonization temperature which indicates that condensation and cyclization of the aromatic ring structures take place during the carbonization process leading to the development of carbidic and graphitic carbon structures.
High resolution XPS spectra of O1s region of the carbon adsorbents are shown in Fig. 6. The O1s spectra were deconvoluted into three different peaks and their corresponding binding energy (B.E.), full width at half maximum (FWHM) and relative area contribution (A%) are listed in Table S2.† Peak O1 corresponds to carbonyl, ketone or lactone groups, peak O2 corresponds to oxygen of ether, phenol, alcohol and/or carboxyl oxygen of esters and/or anhydrides and peak O3 is attributed to oxygen in carboxyl groups.41 As seen in Fig. 6, there is an appreciable change in the nature of oxygen functionalities present on the carbon surface with increase in carbonization temperature. Increase in carbonization temperature from 400 °C to 500 °C lead to decrease in carboxylic content of the carbons as suggested by decrease in relative area percentage of O3 peak from ca. 28% to 16.9%. But further increase in temperature of carbonization up to 700 °C lead to increase in carboxylic content. No peak for carboxylic is observed for RF-800 sample indicating the conversion of carboxylic groups to lactone groups as indicated by large increase in relative area percentage of O1 peak. Peak O1 for RF-800 is only due to lactone group because other functional groups (i.e. carbonyl and quinone) attributing to O1 peak are already found to be absent for this carbon sample. As compared to RF-700 carbon, RF-C contained higher carbonyl or lactonic content and lower carboxylic content. Difference in these oxygen functionalities will affect the surface basicity of these carbons. Increase in lactone groups in RF-800 results in increase in surface acidity of this carbon sample as seen in Table 1. Surface acidity is found to decrease from 1.67 to 0.78 milli equivalents per gram (meq g−1) with increase in carbonization temperature from 400 °C to 700 °C. But further increase in temperature to 800 °C resulted in increase in surface acidity to 1.09 meq g−1. Carboxyl, lactone or lactols and phenol groups are responsible for surface acidity of carbon materials while functional groups like carbonyls, ethers, pyrones and chromenes contribute towards the surface basicity. Surface basicity of the prepared carbons tends to increase from 0.54 to 1.93 meq g−1 with increase in carbonization temperature from 400 °C to 700 °C and finally decreases to 0.78 meq g−1 for carbonization at 800 °C. In contrast, carbon obtained from direct carbonization of RF resin exhibited similar surface acidity as that of RF-700 but exhibited very low surface basicity of 0.88 meq g−1. For RF-700 and RF-600, amount of basic functional groups is higher than amount of acidic functional groups thereby indicating primarily basic nature of the carbon materials whereas rest of the samples have mainly acidic character.
![]() | ||
Fig. 6 Deconvoluted O1s spectra of (a) RF-400, (b) RF-500, (c) RF-600, (d) RF-700, (e) RF-800, and (f) RF-C. |
![]() | ||
Fig. 7 (a) CO2 breakthrough profiles, and (b) CO2 adsorption capacity of synthesized carbons at 30 °C and 10% CO2 feed concentration. |
CO2 adsorption breakthrough curves for RF-700 as a function of adsorption temperature for varying CO2 feed concentrations (5–12.5% by volume) are presented in Fig. 8 and S5.† Longer breakthrough time appears at lower temperatures for a fixed feed concentration indicating drop in adsorption capacity with temperature. With increase in adsorption temperature from 30 °C to 100 °C, tb decreased from 2.12 minutes to 0.22 minutes for 5% feed concentration and for 12.5% feed concentration, tb decreased from 1.67 minutes to 0.21 minutes. Moreover, increase in feed concentration also lead to decrease in breakthrough time at a fixed temperature. It decreased from 2.12 minutes to 1.67 minutes with increase in feed concentration from 5% to 12.5%. But this decrease in tb is almost negligible at higher adsorption temperatures like at 100 °C the value is ca. 0.22 minutes at all inlet concentrations.
![]() | ||
Fig. 8 Breakthrough profiles for RF-700 at 5% feed CO2 concentration at different adsorption temperatures. |
Dynamic adsorption capacity of RF-700 is also calculated from the breakthrough curves at all adsorption temperatures and inlet CO2 concentrations. It is observed that qe tends to increase with increase in feed CO2 concentration but decreases with increase in adsorption temperature (Fig. 9). Value of qe increased from 0.405 mmol g−1 to 0.761 mmol g−1 with increase in CO2 concentration from 5% to 12.5% at 30 °C adsorption temperature. This is attributed to increase in the driving force for the adsorption process i.e. the concentration gradient therefore leading to increased CO2 mass transfer rate across the boundary layer and within the adsorbent pores. On the other hand, decrease in qe with temperature at a fixed inlet concentration is due to exothermic nature of the adsorption process and it is favored at lower temperatures. Also with increase in adsorption temperature, both surface energy and CO2 diffusion rate increases thus decreasing the stability of CO2 on adsorbent surface and hence causing CO2 desorption. For 12.5% feed concentration, value of qe decreased from 0.761 mmol g−1 to 0.281 mmol g−1 on increasing temperature from 30 °C to 100 °C.
![]() | ||
Fig. 9 CO2 adsorption capacity of RF-700 at different adsorption temperatures as a function of inlet CO2 concentration. |
Activated carbons obtained from fly ash demonstrated CO2 adsorption capacity of 0.93 mmol g−1 and 0.42 mmol g−1 at 30 °C and 75 °C respectively.42 RF-700 showed comparable adsorption capacity of 0.405 mmol g−1 at 75 °C and 12.5% feed concentration. This value was ∼1.8 times higher than adsorption capacity of commercial activated carbon (0.225 mmol g−1) at 75 °C under pure CO2 flow.23 For CO2 capture application from flue gas, dynamic capacities of the adsorbent materials are more pertinent at higher temperatures. Dynamic adsorption capacity of RF-700 at all temperatures is higher than dynamic capacities of synthetic carbon obtained from carbonization and steam activation of coal tar pitch and furfural.43 This synthetic carbon exhibited dynamic CO2 adsorption capacity of 0.61 mmol g−1 at 30 °C and 0.30 mmol g−1 50 °C under 15% CO2 concentration. Dynamic capacities of RF-700 are in good accord with literature reported values for carbon adsorbents evaluated under similar conditions.44
Also CO2 uptake of RF-700 at 75 °C under 12.5% CO2 concentration (0.405 mmol g−1) is comparable with dynamic adsorption capacity of isopropanol amine modified zeolite 13× (0.52 mmol g−1) and monoethanol amine modified zeolite 13× (0.45 mmol g−1) under 15% CO2 balance in He flow.9 However, monoethanol amine modified β-zeolite exhibited static CO2 uptake of 0.80 mmol g−1 at 30 °C and 1 bar.13 Amine impregnated silica adsorbents are found to exhibit higher CO2 adsorption capacity than the prepared carbons. For instance, polyethyleneimine (PEI) modified SBA-15 silica exhibited CO2 adsorption capacity of ∼2 mmol g−1 at 45 °C and 1 bar under static conditions.45 PEI impregnated monolith silica showed CO2 uptake of 3.75 mmol g−1, determined gravimetrically, at 75 °C and 5% CO2 concentration while tetraethylenepenta amine (TEPA) modified silica was evaluated for CO2 adsorption only under pure CO2 flow. There was a steady decrease in CO2 adsorption capacity during a 5 run adsorption–desorption cycle.46 In another work, TEPA modified MCM-41 silica also suffered from decrease in its CO2 uptake just after 6 adsorption cycles.47 Although amine modified silica materials have been reported to show very high CO2 uptakes as compared to other adsorbents, they suffer from the stability issue with respect to adsorption capacity over repeated cycles. However, carbon adsorbents prepared in the present work have been found to exhibit practically constant CO2 adsorption capacity over repeated cycles (as discussed in the next section).
For an efficient adsorption process, selectivity of CO2 over N2 is another significant factor besides adsorption capacity. Fig. 10a displays breakthrough curves of N2 and CO2 on RF-700 for adsorption at 30 °C and 50 °C under 12.5% feed CO2 concentration. N2 appeared instantly in the exit gas stream after the onset of adsorption process. This indicates very low N2 adsorption capacity of the adsorbent material. In contrast for adsorption at 30 °C, concentration of CO2 in exit gas stream was zero for ca. 2 minutes and increased gradually with time indicating adsorbent's affinity towards CO2. Moreover, concentration of N2 in the exit gas stream surpassed the N2 feed gas concentration. This is due to replacement of adsorbed N2 in the adsorbent by CO2 gas molecules.
It is important for an adsorbent material to maintain CO2 adsorption capacities in multiple adsorption–desorption cycles from practical application point of view. Fig. 10b demonstrates the CO2 adsorption capacity of RF-700 during multiple adsorption–desorption cycles at different adsorption temperatures and 10% feed concentration. Desorption was carried out at 200 °C under pure N2 gas. CO2 adsorption performance of RF-700 is almost unaltered for four cycles suggesting completely reversible adsorption process thereby indicating the reusability of these carbons in long-term operations.
Fig. S7† presents the CO2-TPD plot of RF-700 sample. Desorption peak in the temperature range of 75–175 °C is observed indicating chemisorption of CO2 on carbon materials48 which is attributed to covalent bond formation between the surface basic functionalities of carbon and CO2 gas molecules. Desorption peak is centred around 120 °C suggesting maximum CO2 desorption occurring at this temperature. XPS and Boehm titration results also confirm the presence of basic functional groups on the carbon surface.
![]() | ||
Fig. 11 Comparison of kinetic models for CO2 uptake on RF-700 at 30 °C as a function of feed concentration. |
For fractional order kinetic model, small value of error % and high R2 values indicated good conformity between experimental data and model predicted values. For all CO2 concentrations, adsorption rate constant kn was found to decrease with increase in adsorption temperature because of exothermic nature of adsorption process. At higher temperatures, adsorption rate may become faster but CO2 adsorbed is lesser due to faster desorption of CO2 from carbon surface. On the other hand, kn tends to increase with increase in CO2 concentration due to enhanced concentration gradient leading to faster diffusion of CO2 molecules into the adsorbent surface. Kinetic model parameters n and m demonstrate the effect of driving force and diffusion resistance respectively.49 Parameter n decreased with increase in adsorption temperature and was maximum at 30 °C suggesting maximum driving force at this temperature while it increased with increase in CO2 concentration. This could be attributed to increase in concentration gradient at high CO2 concentrations thereby leading to faster diffusion of CO2. Value of m indicates the fastness of adsorption process. With increase in CO2 concentrations, m tends to increase suggesting the faster adsorption of CO2 on RF-700 at high CO2 concentrations. Equilibrium CO2 adsorption capacities as predicted by fractional order model were also in good accord with the experimental values.
![]() | ||
Fig. 12 (a) CO2 adsorption isotherm on RF-700 at different adsorption temperatures, and (b) isosteric heat of adsorption of CO2 on RF-700. |
Model | Parameters | Temperature ( °C) | |||
---|---|---|---|---|---|
30 | 50 | 75 | 100 | ||
Langmuir | qm (mmol g−1) | 1.79 | 1.49 | 1.47 | 1.11 |
KL (atm−1) | 6.12 | 5.00 | 3.13 | 2.82 | |
R2 | 0.978 | 0.981 | 0.982 | 0.953 | |
Freundlich | KF (mmol g−1 atm−1/n) | 3.13 | 2.51 | 2.19 | 1.61 |
n | 1.50 | 1.41 | 1.25 | 1.22 | |
R2 | 0.961 | 0.974 | 0.972 | 0.940 | |
Temkin | KT (atm−1) | 55.84 | 52.13 | 43.89 | 41.63 |
b (kJ mol−1) | 6.32 | 8.85 | 12.05 | 17.74 | |
R2 | 0.989 | 0.986 | 0.997 | 0.985 |
CO2 concentration (volume %) | A | Ea (kJ mol−1) | R2 |
---|---|---|---|
5 | 0.0933 | −0.7 | 0.990 |
7.5 | 0.0545 | −2.6 | 0.996 |
10 | 0.0483 | −3.0 | 0.988 |
12.5 | 0.0463 | −3.5 | 0.960 |
Table 4 presents the thermodynamic parameters for CO2 adsorption on RF-700 as obtained from eqn (13) and (17). Standard enthalpy change and standard entropy change for the adsorption process were determined to be −4.2 kJ mol−1 and 0.02 kJ mol−1 K−1. The negative values of Gibbs free energy change at each adsorption temperature indicate the feasibility and spontaneity of the adsorption process. The value of ΔH° was also negative indicating the exothermic nature of the adsorption process. The positive value of ΔS° demonstrated the increase in randomness at the adsorbent–adsorbate interface and also adsorbent's affinity towards CO2.52 Small value of ΔS° indicated that there was no significant change in entropy.
Temperature (°C) | ΔG° (kJ mol−1) | ΔH° (kJ mol−1) | ΔS° (kJ mol−1 K−1) |
---|---|---|---|
30 | −10.1 | −4.2 | 0.02 |
50 | −10.6 | ||
75 | −10.9 | ||
100 | −11.6 |
Energy duty for CO2 desorption was calculated for RF-700 sample as it showed the best results for CO2 capture, by using eqn (18). The specific heat capacity of RF-700 was found to be 1.215 J g−1 K−1. Adsorption was carried out at 30 °C under 12.5% CO2 in N2 while desorption was carried out at 200 °C under pure N2 flow. Adsorption capacity of RF-700 under these conditions was found to be 0.76 mmol CO2 g−1. Hence the sensible heat energy required to heat the adsorbent for desorption was calculated to be 271.8 kJ per mol of CO2 captured. It is assumed that around 75% of the sensible heat required for heating the adsorbent can be recovered by direct or indirect heat exchanger.35 Therefore, the net sensible heat required for the system was estimated to be 67.9 kJ per mol of CO2 captured.
To calculate isosteric heat of adsorption, value of P is determined at multiple temperatures from isotherm for CO2 adsorption on RF-700 for a given qe and the slope is determined from the plot of lnP against 1/T. The isosteric heat of adsorption for RF-700 is in the range of 15.1–16.8 kJ mol−1 with an average value of 15.74 kJ mol−1, demonstrating strong interaction between the carbon surface and CO2 gas molecules (Fig. 12b). With increase in surface coverage, there is small decrease in isosteric heat of adsorption which indicates energetically heterogeneous adsorbent surface. But at higher surface coverage, an increase in isosteric heat of adsorption is observed which is due to intermolecular interaction between the adsorbed CO2 molecules. Value of Qst for carbon materials have been reported to be in the range of 11–30 kJ mol−1.12,53–55
Thermal energy requirement for desorption is equal to 83.9 kJ per mol of CO2 captured or 1.9 MJ per kg of CO2 captured i.e. for desorption of 1 kg of CO2, energy equal to 1.9 MJ has to be supplied which is generated from the combustion of fossil fuels. As CO2 adsorption capacity at 30 °C was 0.76 mmol per g adsorbent = 0.03344 kg CO2 per kg adsorbent, energy required to desorb this much quantity of CO2 = 0.064 MJ. Assuming that bituminous coal is used as fossil fuel for energy production amount of CO2 generated is 0.0884 kg per MJ of energy.56 Hence CO2 generated to produce 0.064 MJ of energy for desorption is 0.0056 kg CO2. It can be concluded (on 1 kg adsorbent basis) that energy duty for desorption of 0.03344 kg CO2 requires 0.064 MJ of energy which leads to generation of 0.0056 kg CO2 from combustion of fossil fuel (bituminous coal) (detailed calculations are given in the ESI†).
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra16255f |
This journal is © The Royal Society of Chemistry 2015 |