Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

1991 International Conference on Coal Science Proceedings: Proceedings of the International Conference on Coal Science, 16–20 September 1991, University of Newcastle-Upon-Tyne, United Kingdom
1991 International Conference on Coal Science Proceedings: Proceedings of the International Conference on Coal Science, 16–20 September 1991, University of Newcastle-Upon-Tyne, United Kingdom
1991 International Conference on Coal Science Proceedings: Proceedings of the International Conference on Coal Science, 16–20 September 1991, University of Newcastle-Upon-Tyne, United Kingdom
Ebook3,238 pages25 hours

1991 International Conference on Coal Science Proceedings: Proceedings of the International Conference on Coal Science, 16–20 September 1991, University of Newcastle-Upon-Tyne, United Kingdom

Rating: 0 out of 5 stars

()

Read preview

About this ebook

1991 International Conference on Coal Science Proceedings
LanguageEnglish
Release dateSep 17, 2013
ISBN9781483142142
1991 International Conference on Coal Science Proceedings: Proceedings of the International Conference on Coal Science, 16–20 September 1991, University of Newcastle-Upon-Tyne, United Kingdom
Author

Sam Stuart

Dr. Sam Stuart is a physiotherapist and a research Fellow within the Balance Disorders Laboratory, OHSU. His work focuses on vision, cognition and gait in neurological disorders, examining how technology-based interventions influence these factors. He has published extensively in world leading clinical and engineering journals focusing on a broad range of activities such as real-world data analytics, algorithm development for wearable technology and provided expert opinion on technology for concussion assessment for robust player management. He is currently a guest editor for special issues (sports medicine and transcranial direct current stimulation for motor rehabilitation) within Physiological Measurement and Journal of NeuroEngineering and Rehabilitation, respectively.

Read more from Sam Stuart

Related to 1991 International Conference on Coal Science Proceedings

Related ebooks

Chemical Engineering For You

View More

Related articles

Reviews for 1991 International Conference on Coal Science Proceedings

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    1991 International Conference on Coal Science Proceedings - Sam Stuart

    Concentrates

    Electron Transfer in Coals

    John W. Larsen, Robert A. Flowers, II, Peter Hall, Bernard G. Silbernagel and Layce A. Gebhard,     Chemistry Dept., Lehigh University/Exxon Res. & Engr. Co.

    Publisher Summary

    This chapter discusses electron transfer in coals. Most coals containing tetracyanoquinodimethane (TCNQ) at a level of one TCNQ per aromatic unit in the coal are bright blue. Coals containing similar amounts of tetracyanoethylene (TCNE) are bronze colored. The only exceptions discovered are low-volatile bituminous coals and anthracites. The chapter presents evidence that these striking and unusual colors are because of the formation of valence bands in coals resulting from the addition of these two good electron acceptors. A significant amount of electron density is transferred to TCNQ and TCNE in coals. Both of these materials are good organic oxidants. It has been demonstrated that there is a linear relationship between the amount of electron density transferred from a donor to TCNQ and the shift in the TCNQ CN infrared stretching frequency that occurs upon electron transfer. Electrons added to the TCNQ must enter its lowest unoccupied molecular orbital.

    Most coals containing tetracyanoquinodimethane (TCNQ) at a level of one TCNQ per aromatic unit in the coal are bright blue. Coals containing similar amounts of tetracyanoethylene (TCNE) are bronze colored. The only exceptions we have discovered are low volatile bituminous coals and anthracites. We will present evidence here that these striking and unusual colors are due to the formation of valence bands in coals resulting from the addition of these two good electron acceptors.

    A significant amount of electron density is transferred to TCNQ and TCNE in coals. Both of these materials are good organic oxidants. It has been demonstrated that there is a linear relationship between the amount of electron density transferred from a donor to TCNQ and the shift in the TCNQ CN infrared stretching frequency which occurs upon electron transfer[1]. Electrons added to the TCNQ must enter its lowest unoccupied molecular orbital (LUMO). This is an anti-bonding orbital and its occupation weakens bonds throughout the structure. It is this bond weakening that results in a shift in the CN stretching frequency. This shift is 44 cm−1 for transfer of one electron. Transfer of one electron to TCNE gives a 47 cm−1 shift. It is thus possible to use infrared spectroscopy to determine the extent of electron transfer from coals to these acceptors. TCNQ in the following coals at several molar levels corresponding to between 3% and 100% of the concentration of aromatic structures gives the full 44 cm−1 shift: Beulah-Zap Lignite (74%), Wyodak Sub-bituminous (76%C), Illinois No. 6 (81%C), Pittsburgh No. 8 (85%C), Upper Freeport (88%C). The concentrations of aromatics were calculated from the data of Solum[2]. This shift indicates that the LUMO of TCNQ contains one electron and is thus half occupied. Similar behavior was observed with TCNE. If there also exists a linear relationship between extent of electron transfer and the CN shift for TCNE, the observed 36 cm−1 shift corresponds to a LUMO which is 30% occupied. It is striking and important that all of the TCNQ or TCNE added to these coals is shifted. There is no evidence for a CN absorption band at its normal position and there is no CN having an intermediate value for the stretching frequency.

    ESR results demonstrate that the TCNQ radical anion is not formed. The ESR spectra do show that there is a small increase in the spin population in the doped coals. These data are presented in Table 1 together with the calculated spin population if all the added TCNQ formed the radical anion. The spectra themselves do not show any of the lines expected for the radical anion.

    Table 1

    EPR Spin Densities (Measured and Calculated) of Coals Containing TCNQ

    * Spin Density EPR Measurements were obtained by Dr. Bernard Silbernagel and Layce Gebhard, Exxon Research & Engineering Company

    Aromatic molecules of the types found in the coal are not individually capable of transferring significant electron density to coal. This is in accord with theory and is demonstrated by data which appear in Table 2. All of the aromatic systems examined, a sampling of which appears in Table 2, bring about shifts in the CN stretch between 3 cm−1 and 5 cm−1. This demonstrates conclusively that individually these molecules are not capable of bringing about the behavior observed with coals.

    Table 2

    Changes in CN Stretching Frequency in Formation of Aromatic Complexes with TCNQ in CHCI3

    Our experiments have established that the TCNQ LUMO is half occupied in many coals, that this occupation does not result from the formation of the radical anion, and that individual aromatic structures of the types found in coals cannot populate the LUMO producing the large IR shifts observed. Similar experiments with coal extracts demonstrate that the observed effects are not due to the presence of mineral matter. The most reasonable explanation for these data is that the addition of the electron acceptors to coal results in the formation of extended electronic valence bands. These are well established in solid mixtures of good electron donors and acceptors and there exists an extensive literature dealing with them[3]. Their existence in an amorphous system such as coal is surprising and noteworthy. The data are consistent with the formation of such bands by the extended overlap of the aromatic II systems with those of the added oxidants. The LUMO of the oxidants must form part of this band structure which is half occupied in the case of TCNQ and 30% occupied in the case of TCNE. The fact that electronic transitions occur at infrared frequencies for anthracites and other high rank coal, demonstrates that extended II structures occurs naturally in these materials[4]. This is not the case for the lower rank coals, though we seemed to have induced even more extensive II structures by the addition of the oxidants.

    Our explorations of the electronic and physical structure of these interesting materials continues and we are also exploring the role of internal electron transfers in the chemistry of coals.

    ACKNOWLEDGEMENTS

    We thank the U. S. Department of Energy for support of the Lehigh University portion of this work.

    REFERENCES

    1. Chappel, J. S., Bloch, A. N., Boyden, W. A., Maxfield, M., Poehler, T. O., Cowan, D. O. J. Am. Chem. Soc.. 1981; 103:2442.

    2. Solum, M. S., Pugmire, R. J., Grant, D. M. Energy & Fuels. 1989; 3:187.

    3. Soos, Z. G., Klein, D. J. Charge Transfer in Solid State Complexes. In: Foster R., ed. Molecular Association. New York: Academic Press; 1981:1–109.

    4. Brown, J. R. J. Chem. Soc. (London). 1955; 52–744.

    EXAMINATION OF MACROMOLECULAR NETWORK OF COAL BY DIFFERENTIAL SCANNING CALORIMETRY

    Kouichi Miura, Kazuhiro Mae, Shoji Takebe and Kenji Hashimoto*,     Research Laboratory of Carbonaceous Resources Conversion Technology,

    Publisher Summary

    This chapter explores the macromolecular network of coal by differential scanning calorimetry. It describes the properties of five coals used in a study described in the chapter. These coals were ground into the particles of less than 74 μm in diameter and were then dried in vacuo at 110°C for 24 h before use. Four solvents, namely, tetralin (Tet.), 1-buthanol (1-BuOH), ethanol (EtOH), and quinoline (Q) were used for the swelling. Tetralin was mainly used because of its nonpolarity. The swelling of coal was performed by simply mixing coal and solvent by the ratio of 1 to 0.6 by weight in a closed tube. It was performed at 30°C for 24 h for the polar solvent and was performed at 100 to 220°C for 1 h under 1 MPa of N2 for the nonpolar solvent. The chapter presents the swelling ratio of Taiheiyo (TC) coal treated with tetralin and that of the vacuum-dried coal against the swelling temperature. It was found that TC began to swell at around 70°C and seemed to reach an equilibrium swelling ratio of ca. 1.35 over 150°C.

    1 INTRODUCTION

    Several attempts[1–4] have been made to estimate the non-covalent bond in coal from calorimetric measurements during the solvent swelling, in which the heat of wettability of the strong polar solvent was mainly measured using the micro calorimeter. However, the information on the non-covalent bond could not be successfully extracted from these studies because of the strong interaction between the polar solvent and the coal. Non-polar or weak polar solvents should be used to study carefully the macromolecular structure of coal through the solvent swelling.

    From this viewpoint we used the non-polar solvent for the swelling, but the swelling was performed at the temperatures as high as 220 °C. Then we tried to estimate the heat required to break the non-covalent bonds in coal by measuring the DSC and TG profiles of this swollen coal.

    2 EXPERIMENTAL

    The properties of five coals used are given in Table 1. These coals were ground into the particles of less than 74 μm in diameter, then dried in vacuo at 110 °C for 24 h before use. Four solvents, tetralin (Tet.), 1-buthanol (1-BuOH), ethanol (EtOH) and quinoline (Q) were used for the swelling. Tetralin was mainly used because of its non-polarity. The swelling of coal was performed by simply mixing coal and solvent by the ratio of 1 to 0.6 by weight in a closed tube. It was performed at 30 °C for 24 h for the polar solvent, and was performed at 100 to 220 °C for 1 h under IMPa of N2 for the non-polar solvent. The swollen coal (STC) was evacuated at 70 °C for 24 h to completely remove the solvent from the STC. The swelling ratio of the STC and thus prepared vacuum-dried coal (VDC) was measured by the volumetric technique [5].

    Table 1

    Properties of Coal

    To examine the change of non-covalent bond of the coal during the swelling by calorimetrically, the DSC profile and the TG curve of the raw coal, the STC, and the solvent were measured under a constant heating rate of 5 °C/min by use of a differential scanning calorimeter (Shimadzu Co., DSC 50) and a thermobalance (Shimadzu Co., TGA 50), respectively.

    3 RESULTS AND DISCUSSION

    The swelling ratio of TC treated with tetralin and that of the VDC are shown against the swelling temperature. TC began to swell at around 70 °C, and seemed to reach an equilibrium swelling ratio of ca.1.35 over 150 °C. The swelling ratio of the VDC, ηVDC, is shown by the closed key in Fig. 1. At the swelling temperatures lower than 100 °C the ηVDC value is exactly unity. This means the swollen then vacuum dried coal returned to the raw coal as far as the volume is concerned at these temperatures. On the other hand, the ηVDC value does not return to unity at the swelling temperatures of 150 and 220 °C, and interestingly ηVDC is very close to ηSTC at 220 °C. This indicates that the swelling at 220 °C is almost irreversible. This is also the case for JR as shown in Table 1. The irreversibly swollen VDC is utilized later to estimate solvent-coal interaction solely.

    Fig. 1 Changes of the swelling ratios of the STC and its VDC

    We have clarified that the swelling by tetralin affects solely to the non-covalent bonding of coal as far as the swelling is performed at a temperature lower than a critical temperature, Tc. The critical temperature, Tc, which varied with coal type was estimated to be around 230 °C for TC, for example. In the solvent swollen coal some non-covalent bonding are broken, and the macromolecular network is expected to be altered from the raw coal. In this work we have tried to estimate the energy required to break the non-covalent bonding through the TG and DSC measurements during the heat up of raw coal, solvent, and the STC from 25 to 220 °C.

    Figure 2 shows schematically the enthalpy levels of the raw coal heated to 220 °C (state A), the STC at 25 °C (state B), and the STC heated to 220 °C (state C) on the basis of the raw coal and the solvent at 25 °C. The enthalpy at state A, HC′, is easily measured. In the STC at state B some non-covalent bonding are broken, and the solvent-coal interaction exists. Therefore, the enthalpy at state B, Hi, is the sum of the energy to break the non-covalent bonding, HNC, and that deriving from the solvent-coal interaction, HS-C. The value of HNC is expected to be positive, whereas that of HS-C is negative. The enthalpy at state C, Hf, consists of the enthalpy of the solvent, HS, and the coal, HC. HS is easily obtained from the DSC measurement of the solvent. The enthalpy difference ΔH (=Hf – Hi) can be obtained from the DSC and TG measurements of the STC. Then we can extract the value of Hi from the measurable quantities if we assume that the coal at state C is the same as that at state A, namely Hc=Hc′, by

    Fig. 2 Schematic enthalpy level of various state of coal and STC

    (1)

    If we could further estimate the value of HS-C, we can determine the enthalpy required to break the non-covalent bonding by

    (2)

    It was possible to determine the value of HS-C for the coal swollen by tetralin by measuring the DSC and TG profiles during the desorption of the tetralin from the irreversibly swollen VDC which adsorbed tetralin. Figure 3 shows the TG and DSC curves for tetralin and the TC swollen at 100 °C and 220 °C. The endothermic heats integrated over 25 to 220 °C correspond to HS for tetralin, HC for the coal and ΔH for the STC, respectively.

    Fig. 3 TG and DSC profiles of STCs and tetralin

    Figure 4 shows the values of Hi of TC estimated using several solvents. Hi decreased with the increase in the polarity of solvent, and it was large negative value in the case of the swelling by quinoline as mentioned in other earlier works [1–4]. This is because the polar solvent interacts with the functional groups in the coal and the strong coal-solvent hydrogen bonding is formed in place of the coal-coal hydrogen bonding during the swelling, leading to |HS-C|>HNC.

    Fig. 4 Effect of solvent type on Hi

    Figure 5 shows the changes of the values of HS-C and HNC against the swelling ratio. All the values were estimated by utilizing tetralin as the solvent. HS-C decreased with the increase of the swelling ratio. This means that the solvent interacts strongly with the stronger non-covalent site in the coal at higher swelling ratio. HNC increased with the increase in the swelling ratio, and reached to 110 kJ/kg-coal at the swelling ratio of 1.3. Since HS-C is much smaller than HNC, we can safely say that the enthalpy to break the non-covalent bonding was successfully measured by use of tetralin as a solvent.

    Fig. 5 Changes of HNC and HS-C through the solvent swelling

    Figure 6 shows the effect of coal type on the value of HNC. The values of HS-C were estimated from the data in Fig. 5. HNC increased with the decrease in coal rank, because the lower rank coal contained more oxygen and many non-covalent bondings. If we further assume that the dissociation energy of all the non-covalent bonding are 28 kJ/mol, we can roughly estimate that about 35 % of oxygen of coal are broken through the swelling for these three coals. This is just a rough estimation, but we are expecting that the method proposed here will give a clue for studying the non-covalent bonding, namely the macromolecular network of coal.

    Fig. 6 Effect of coal type on HNC

    4 CONCLUSIONS

    We presented a method to estimate the heat required to break the non-covalent bonding in coal using a differential scanning calorimetry (DSC) and a thermogravimetry (TG). Utilizing DSC and TG profiles of three samples: the coal swollen by a solvent, the raw coal, and the solvent measured by heating them up to 220 °C, we could successfully estimate the enthalpy which is related to the strength and the amount of non-covalent bonds in coal.

    5. REFERENCES

    1. Tempy, G. K., et al. Energy & Fuels. 1988; 2:787–793.

    2. Fowkes, F. W., Jones, K. L., Li, G., Loyd, T. B. Energy & Fuels. 1989; 3:97–105.

    3. Groszek, A. J., Templer, C. E. Fuel. 1988; 67:1658–1661.

    4. Hollenhead, J. B., et al. Energy & Fuels. 1988; 2:121–124.

    5. Green, T. K., Kovac, J., Larsen, J. W. Fuel. 1984; 63:935–938.


    *)Departmant of Chemical Engineering, Kyoto University, Kyoto, 606 JAPAN

    COAL-WATER MOLECULAR INTERACTIONS

    Leo J. Lynch, Wesley A. Barton and David S. Wesbter,     CSIRO Division of Coal and Energy Technology, PO Box 136, North Ryde 2113, Australia

    Publisher Summary

    This chapter presents an investigation of coal–water molecular interactions. In a study described in the chapter, crushed samples of Australian coal were prepared with equilibrium water content between 0 and 25% (wt/wt). ¹H NMR transverse relaxation signals of each specimen were measured using solid echo pulse sequence at regular intervals during a temperature cycle from 310 to 210 K and back to 310 K. The adsorption/desorption isotherm was determined at 298 K by allowing samples to equilibrate at seven different relative humidities for one month. Also, duplicate predried samples were monitored for four weeks by regular weighing for their uptake of water in an atmosphere of 100% rh at 293 K. The equilibrium adsorption/desorption isotherm showed little hysteresis, and the approach to saturation was not asymptotic so that an estimate of 12.4% for the saturation water content (SWC) was obtained with good precision. The uptake at 100% rh was rapid at first, but at least 200 min were required for equilibrium to be closely approximated. There was a significant difference between the two samples tested, and the SWC estimates obtained were 14.5% and 13.0%—both greater than the isotherm estimate.

    1 INTRODUCTION

    Water sorbed or otherwise intimately associated with solid-like materials such as coals has properties which differ somewhat from those of the normal thermodynamic states of bulk water. These differences can be appreciated and understood best at the molecular level and, in particular, if the microdynamic processes or molecular dynamics of the water are considered. An experimental technique which allows such interpretations is proton nuclear magnetic resonance (¹H NMR) spectroscopy. There are well developed theories relating ¹H NMR relaxation data to molecular dynamic models of the system under study [1]. In particular, the measured ¹H NMR transverse relaxation often allows distinction of rigid from mobile molecular structures/lattices on the basis of whether the molecular reorientation rates are below or above ∼10⁵ Hz respectively. Modifications to the water properties result from molecular interactions between water and substrate and thus to an extent are determined by the nature of the substrate. For coals the extent of polar interactions is greatest for species rich in oxygen and other electronegative atoms. Hydrophobic effects related to apolar regions of the coal, are also likely to contribute to perturbation of the water. That vicinal water is perturbed with respect to bulk water leads to the concept of interacting as distinct from non-interacting water and allows definition of the maximum capacity of a material to interact with water, i.e., its saturation water content (swc). The swc is expected to be equivalent to the equilibrium water content (ewc) at 100% relative humidity (rh).

    The substrate also is likely to be affected by interaction with water, but most models of sorbed water systems imply that the substrate is inert. Clearly this is not so for brown coals which are altered irreversibly when dried from the bed moist condition [2]. Higher rank coals are less affected, but the extent can still be important for understanding these coal-water systems. Water vapour sorption isotherms often are used to characterise water – substrate systems – the isotherm shape can be related to models of the system [3] and the swc estimated as the rh approaches 100%. However, the accuracy of this determination is limited by instability of the system near 100% rh. In this paper a refinement of a previously described [4] ¹H NMR method to characterise and estimate the swc of coal-water systems will be discussed and applied to a high inertinite, high moisture content coal.

    2 EXPERIMENTAL

    Crushed (-212 μm) samples of the Australian coal studied (82.6%C, 4.5%H, 10.7%O (diff) (daf); 7.5% ash (ad); 28% vitrinite, 69% inertinite (vol. % mmf)) were prepared with ewc’s between 0 and 25% (wt/wt). ¹H NMR transverse relaxation signals i(t) of each specimen were measured using the solid echo pulse sequence [5] at regular intervals during a temperature cycle from 310 to 210 K and back to 310 K. The adsorption/desorption isotherm was determined at 298 K by allowing samples to equilibrate at seven different relative humidities for one month. Also duplicate predried samples were monitored for four weeks by regular weighing for their uptake of water in an atmosphere of 100% rh at 293 K.

    3 RESULTS AND DISCUSSIONS

    The equilibrium adsorption/desorption isotherm showed little hysteresis and the approach to saturation was not ‘asymptotic’ so that an estimate of 12.4% for the swc was obtained with good precision. The uptake at 100% rh was rapid at first but at least 200 minutes were required for equilibrium to be closely approximated. There was a significant difference between the two samples tested and the swc estimates obtained were 14.5% and 13.0% – both greater than the isotherm estimate.

    The typical ¹H NMR transverse relaxation signal i(t) shown in Fig.1 can be approximated by a slowly decaying exponential function im(t) = im0 exp (-t/T2) and a residual signal ir(t) = i(t)-im(t) which is Gaussian-like. It is reasonable to attribute these component signals to hydrogen contained in mobile and rigid-like fractions of the molecular structures respectively. The ‘mobile’ signal amplitude can be contributed to by unfrozen water molecules and parts of the coal structure inherently mobile at the temperature of measurement or ‘plasticised’ by interaction with water. T2 is the spin-spin relaxation time constant of the mobile species and the initial amplitudes im0 and ir0 are proportional to the hydrogen contents of the two fractions [1]. Thus a hydrogen-weighted measure of the fraction of the specimen’s molecular structure which is mobile is given by I = im0/(im0 + ir0). Plots the parameter I against measurement temperature during heating are shown in Fig. 2 for the eight samples of ewc between 0 and 25%. Clearly I has a discontinuity near the ice fusion temperature for specimens with ewc of 12.5% and higher but not for those of lower ewc. In Fig. 3 a comparison of I versus ewc at 260 and 300 K shows that below the fusion transition I is insensitive to ewc for these higher ewc specimens in contrast to its great sensitivity to ewc for specimens which do not exhibit a discontinuity in I near 273 K. This general behaviour is consistent with the presence of free or non-interacting water in the higher water content specimens co-existing with a fixed quantity of interacting water contained at a saturation level equivalent to some value between 10 and 12.5% ewc.

    Figure 1 ¹H NMR transverse relaxation signal i(t) approximated by mobile imt and rigid ir(t) components.

    Figure 2 Variation with increasing temperature of percentage of ‘mobile’ ¹H NMR signal of coal-water samples of various water content.

    25%

    Figure 3 Variations with ewc of I below (260 K) and above (300 K) the ice fusion temperature.

    The trend for I to increase gradually with temperature over the entire range for all levels of ewc reflects the thermal enhancement of molecular mobility which gradually transfonns molecular species from the rigid-lattice to the mobile condition according to the ¹H NMR criterion. This gradual effect can be quantified in terms of a continuous distribution of states model for the interacting water [6].

    The criterion distinguishing non-interacting and interacting water is the occurrence of the critical fusion transition characteristic of bulk water at 273 K. Where this occurs for the higher water content specimens a step change ΔI in the parameter I near 273 K can be measured. ΔI is, in effect, a direct measure of the non-interacting water. If the total ewc (mass of water as a % of the mass of dry coal) and the elemental hydrogen concentration (k) of the coal specimen are known, the quantity of interacting Wi (and hence non-interacting) water on a weight % of dry coal can be calculated from Wi = ([H]w – ΔI)·ewc / [H]w where [H]w = (k.ewc)/(k.ewc + 100 K) and K = 11.19 is the weight percent hydrogen content of water. The plot of Wi values against ewc in Fig. 4 provides an accurate estimate of 11.8% for the swc of the high inertinite coal studied.

    Figure 4 Interacting water content Wi (wt/wt%) and fraction of water ‘plasticised’ coal structure Cp (H/H%) plotted against ewc.

    A similar procedure can be used to estimate the extent Cp to which coal structure is ‘plasticised’ by its interaction with water. This requires knowledge of the hydrogen fractions of the water [H]w and the coal [H]c = 1-[H]W and the measured ‘mobile’ signal amplitude parameter I. It is necessary to presume that all the water contributes to the ‘mobile’ signal which is clearly not the case for high water content specimens below 273 K but is likely to be a good approximation near 300 K. Thus the hydrogen weighted contribution of the coal to the mobile fraction is [H]mc = I – [H]w and the estimate of the coal plasticised by water on a hydrogen weighted basis is Cp = ([H]mc – [H]mco) / [H]c. Here [H]mco is an estimate of the proportion of mobile hydrogen (structure) in the dry coal. The steady increase in Cp at 300 K with ewc up to ∼10% ewc (Fig. 4) can be attributed to the gradual penetration and mobilisation of parts of the coal structure stabilised by polar bonding and which would be regions of affinity for the interacting water. The trend for Cp to ‘level off above the swc (∼11.8% ewc) is consistent with this idea. However, the increased value of Cp at 25% ewc is unexpected and probably due to experimental error. Whereas these results are of limited precision, they clearly demonstrate that a significant part of the coal structure – perhaps at least 8% H/H – is destabilised by its exposure to and interaction with water. These findings emphasise the fact that coals generally are not inert substrates where uptake of water can be explained only in terms of macroscopic phenomena such as porosity [7].

    4. REFERENCES

    1. Lynch, L. J.Cohen J.S., ed. Magnetic Resonance and Biology. J. Wiley: N.Y., 1983:248–304. [Chapter 5].

    2. Murray, J. B., Evans, D. G. Fuel. 1972; 51:290.

    3. Brunauer, S., Emmett, P. H., Teller, E. J. Am. Chem. Soc.. 1938; 60:309.

    4. Lynch, L.J. and Webster, D.S. Fuel, 1982, 61, 271–275.

    5. Powles, J. G., Mansfield. P. Phys. Lett.. 1962; 2:58.

    6. Lynch, L. J., Marsden, K. H., George, E. P. J. Chem. Phys.. 1969; 51:80–5673.

    7. Mraw, S. C., O’Rourke, D. F. Science. 1979; 130:901.

    CHARGE TRANSFER COMPLEXES BETWEEN COAL AND ELECTRON ACCEPTORS: NON COVALENT INTERACTION IN COAL STRUCTURE

    M. Sasaki, H. Kumagai and Y. Sanada,     Metals Reserch Institute, Hokkaido University, N-13, W-8, Kita-ku, Sapporo 060, Japan

    Publisher Summary

    This chapter describes charge transfer complexes between coal and electron acceptors. In a study described in the chapter, coal samples tested are mostly chosen from Argonne Premium Coal Samples. Iodine and 7,7,8,8-tetracyanoquinodimethane (TCNQ) as electron acceptors were selected and used without further purification. A known amount of the guest substance dissolved in chloroform was added to coal. The suspensions were placed in an ultrasonic bath for 10 min and stirred further with a magnetic stirrer for 1 h at room temperature. Solvent then evaporated off, and the coals were kept at 40°C in a vacuum oven over night. Sample was degassed at 10−4 Pa and sealed in an ESR glass tube. Spectra were obtained using a Varian model E-109 spectrometer. The value of spin concentration was calibrated by DPPH and was reproducible within 2%. The nature of charge transfer complexes of coal/TCNQ and coal/iodine system are quite different. TCNQ interacts with oxygen containing functional groups which are able to form hydrogen bonding, and iodine interacts with aromatic rings in the structure of coal.

    1 INTRODUCTION

    The inter- and intramolecular associations have long been acknowlegde as being of fundamental importance to overall physical properties of coal. There are several types of non-covalent interactions such as van der Waals forces and their relative population changes with coal rank. It is well known that three different types of specific non covalent interactions, i.e., hydrogen bonding, charge transfer interactions and aromaic π – π interactions. These inter- and intramolecular associations are expected to contribute to the secondary structure in coal. These are believed to affect greatly on the solubulity of coal and coal-derived liquid in various organic solvents and the viscosity of coal-derived liquids[1]. Hydrogen bonding in coal and model compounds has been investigated by means of infrared[2] and NMR spectroscopies[3]. Calorimetric studies have also been tried[4].

    In a previous paper, we have investigated that coal molecules are interacting with selected guest molecules such as iodine and TCNQ. The interaction between coal and the guest molecules gives us an information of non covalent bonding[5]. This paper is to get deep understanding of non covalent bonding in coal by means of ESR measurement for electron acceptor doped coal.

    2 Experimental

    Samples: Coal samples tested are mostly chosen from Argonne Premium Coal Samples kindly supplied from Dr. Vorres, Argonne National Laboratory. The Analytical data of the samples are shown in elsewhere[6].

    Electron acceptor as a guest substance: Iodine and 7,7,8,8-tetracyanoquinodimethane(TCNQ) as electron acceptors are selected and used without further purification.

    Preparation: A known amount of the guest substance dissolved in chloroform was added to coal. The suspensions were placed in an ultrasonic bath for 10 min. and stirred further with a magnetic stirrer for 1 h. at room temperature. Solvent then evaporated off and the coals were kept at 40°C in a vacuum oven over night.

    ESR measurement: Sample was degassed at 10−4 Pa and sealed in an ESR glass tube. Spectra were obtained using a Varian model E–109 spectrometer. The value of spin concentration is calibrated by DPPH and was reproducible within 2 %.

    3 Spin concentration in coal and electron acceptor doped coal

    Strong interaction between iodine and coal molecules results formation of charge-transfer complexes[7]. We have established a good correlation between the spin concentration of polynuclear aromatic hydrocarbon-iodine complexes and ring compactness factor. For coals, the difference between the values of spin concentration for parent coals, Nso, and electron acceptor(iodine) doped coals, Ns, rise monotonically with increasing coal rank up to approximately 91 %C[5, 8], Here, we have studied the non covalent interaction regarding to spin concentration of charge transfer complexes between the coal and iodine and TCNQ as the guest molecules.

    Figure 1 shows the value of Ns-Nso for iodine doped coal as a function of H/C atomic ratio. Good correlation is obtained between spin concentration and H/C. A remarkable increase of Ns values for the iodine doped coal with decrease H/C. The decrease of H/C means development of aromatization of coal. The interpretation of relationship shown in Figure 1 is that Ns-Nso for iodine may reflect information on the development aromatic structure in coal.

    Fig. 1 Relation between Ns-Nso and H/C for coal/iodin system,

    On the other hand, a remarkable decrease of Ns-Nso values for TCNQ doped coal are recognized with increase of coal rank, although TCNQ is a molecule of an electron acceptor, same as iodine[5, 8]. TCNQ as a guest plays paticularly important role for low rank coal. No relationship was found between the value of Ns-Nso and H/C value of parent coal. The value of Ns for aromatic compound with OH functional group, such as anthraflavic acid, is higher than that with bare aromatic ring hydrocarbons by the addition of TCNQ. Then, the relation between the value of Ns-Nso and 0/C atomic ratio of parent coal can be depicted as Figure 2. From the Figure 2, an approximately linear relation exists between the value of Ns-Nso and 0/C. Oxygen in coal have been found to occur predominantly such functional groups as phenolic groups. This is able to form hydrogen bonding. Therefore, TCNQ molecule associates to the sites of functional groups which are able to form hydrogen bonding. Interaction of TCNQ with these particular sites in coal structure may result the change of spin concentration of coal.

    Fig. 2 Relation between Ns-Nso and O/C for coal/TCNQ system.

    4 ESR spectral shape

    Figure 3 illustrates spectra obtained for Beulah-Zap lignite, Blind Canyon and Pittsburgh #8 high volatile bitumious coals and their TCNQ doped coals. The spectra of Beulah-Zap and Blind Canyon coal consist of one broad line. While, Pittsburgh #8 coal consists of two distinct lines, one is narrow and the other is broad. The structual model of bituminous coal recently proposed by Kovac and Larsen[9, 10] is very well adapted to the interpretation of the ESR signals. Their model involves a small molecular(M) phase and a macromolecular(MM) phase. The MM phase forms a three-dimensional skeleton, consisting of macromalecular fragments, connected by cross-bonds. The M phase is distributed in the pores of the MM phase or on its edges. The analysis of relaxation times led to the result that the paramagnetic center related with the narrow line are attacked to the latticed macromolecular phase MM, whereas the centers related with the broad line are distributed throughout the molecular phase M[11].

    Fig. 3 ESR spectra of coal and TCNQ doped coal

    The peak to peak linewidth, Δ Hpp, of coal incerases by addition of TCNQ or iodine molecule, except for Beulah-Zap/TCNQ system. A broad line is due to increase in M phase in coal. The electron acceptor molecule such as iodine and TCNQ attacks non covalent bonding in coal. The intermolecular association of non covalent bonding, which is apparently acting as covelent bonding, becomes broken by electron acceptor molecule. The interaction between coal molecule and electron acceptor molecule results probably in the increase in M phase in coal.

    To make sure that the narrow line is due to MM phase in coal, ESR was measured at low temperature. For Pocahontas #3 low volatail bituminous coal with two distinct lines itself, no changes of narrow line was observed with temperature decrease. Figure 4 shows change of ESR spectra for Beulah-Zap/TCNQ system in various temperature. From this figure, it was proved that the narrow line decrases with temperature decrease. Finally, the narrow line have almost disappeared at −160°C. Appearance of a narrow component for Beulah-Zap/TCNQ system might be due to differnt interaction such as single electron transfer from coal to TCNQ molecule.

    Fig. 4 Change of ESR spectra of Beulah-Zap/TCNQ system with temperature.

    5 CONCLUSIONS

    In conclusion, the nature of charge transfer complexes of coal/TCNQ and coal/iodine system are quite different. TCNQ interacts with oxygen containing functional groups which are able to form hydrogen bonding, and iodine interacts with aromatic rings in the structure of coal. The guest and host interaction using ESR technique will give us another insight into the non covalent interaction in coal molecular structure.

    6. REFERENCES

    1. Stenberg, V. I., Baltisbuger, R. J., Patal, K. M., Raman, K., Woolsey, N. F.Gorbaty, Larsen, Wender. Coal Science; vol.2. Academic Press, 1983:125.

    2. Ignasiki, T. Fuel. 1977; 56:359.

    3. Schweighardt, F. K., Friedel, R. A., Retcofsky, H. L. Appl. Spectrosc.. 1976; 30:291.

    4. Tewari, K. C., Wang, J. T., Li, N. C., Yeh, J. J.C. Fuel. 1979; 58:371.

    5. Sasaki. M, Yokono, T. and Sanada, Y., Australian Coal Sci. Conf., Brisbane, 1990, Dec. 3–5

    6. Vorres, S. User Book for Argonne Premium Coal Sample Program 1989, Oct. 1;

    7. Yokono, T., Takahashi, N., Sanada, Y. Energy & Fuels. 1987; 1:227.

    8. Kaneko, T., Sasaki, M., Yokono, T. and Sanada, Y., 1989 Intern. Chem. Congress of Pacific Basin Soc. 51,536, Hawaii, Dec. 17–22.

    9. Kovac, J., Laesen, J. W. Am. Chem. Soc. Div. Fuel Chem.. 1977; 22:181.

    10. Larsen, J.W. and Kovac, J., Organic Chemistry of Coal (Ed. Larsen, J.W.), Am. Chem. Soc. Symposium Series, 1978, 71, 142.

    11. Duber, S., Wieckoski, A. B. Fuel. 1982; 61:431.

    Solid State C-13 NMR Investigations about the Reactivity of Coals and Maceral Concentrates

    I. Wieschenkämper¹, M. Steller¹ and W. Kalkreuth²,     ¹DMT-Gesellschaft für Forschung und Prüfung mbH, Essen (Germany F. R.); ²Geologisches Institut, Universität Köln (Germany F. R.)

    Publisher Summary

    This chapter focuses on solid-state C-13 NMR investigations on the reactivity of coals and maceral concentrates. In an investigation described in the chapter, the ¹³C cross-polarization/magic angle spinning (CP/MAS) spectra were obtained on a Bruker MSL 200 spectrometer operating at 50.3 MHz. The spectra were recorded with the TOSS sequence at 0.6 ms contact time, 4 s pulse repetition time, 50 ms data acquisition, 50 kHz XH decoupling r.f. field, and 3600 Hz spinning rate. About 300 mg of the ground material were filled into cylindrical double bearing rotors made of ZrO2. The Hartmann Hahn match was achieved using a sample of adamantane. Chemical shift scale was determined measuring an acetone probe. The aromaticities of the feedcoals were found to correlate well with coal rank. The hydrogenation residues are more aromatic than the feedcoals. Aromaticity and relative amounts of carbons in different structural groups are nearly the same for all hydrogenation residues independent of rank and composition of the feedcoals.

    Introduction

    Solid state ¹³C nmr has become an established method of coal analysis and yields important information on coal structure and about changes of the organic matter. The aromaticity fa (fraction of aromatic carbon with respect to total carbon) is of particular importance to characterize coal. Refering to the numerous literature it is generally accepted that a gradual increase of aromaticity is part of coalification process. Moreover it is wellknown that the aromaticity of the main macerals of coal increase following the sequence exinite < vitrinite < inertinite. Apart from measuring aromaticity, solid state ¹³C nmr can yield valuable information on the different types of carbons present (methyl, methine, methylene, quaternary)¹,². In this study the reactivity of Canadian and German coals and of German maceral concentrates was investigated by means of solid state ¹³C nmr. The reactivity was tested by hydrogenation and combustion experiments. The coals within a rank range from lignite to anthracite (0,23–3,71 % Rrandom) were obtained from major deposites of Canada and Germany. Maceral concentrates were produced by density gradient centrifugation of five German coals. The original coals and maceral concentrates as well as the hydrogenation and combustion residues were measured to characterize changes of the chemical structure. The aromaticity fa and the relative amount of carbon atoms in different structural groups (CH3-, CH2/CHaliphatic-, CHaromatic-, Cnon-protonated) determined and correlated with coal rank.

    Experimental

    Solid state nmr spectroscopy

    The ¹³C CP/MAS spectra were obtained on a Bruker MSL 200 spectrometer operating at 50,3 MHz. The spectra were recorded with the TOSS sequence³ [combination of cross polarization, magic angle spinning (CP/MAS) and spinning sideband suppression] using the following conditions: 0,6 ms contact time, 4 s pulse repetition time, 50 ms data acquisition, 50 kHz ¹H decoupling r.f. field, 3600 Hz spinning rate. About 300 mg of the ground material (< 0,1 mm particle size) were filled into cylindrical double bearing rotors made of ZrO2. The Hartmann Hahn match was achieved using a sample of adamantane. Chemical shift scale was determined measuring an acetone probe (chemical shift of the methyl carbon 29,8 ppm relative to TMS). In general 1600 transients were accumulated to record the spectra of the coals and maceral concentrates. Up to 4000 transients were necessary to obtain adequate S/N ratios investigating hydrogenation and combustion residues. Additionally the combustion residues were measured applying single pulse excitation. The ¹³C−90° pulse was determined to 3,5 μs. A pulse repetition time of 30 s was used and up to 6000 transients were accumulated performing the latter measurements. The aromaticity fa was calculated from the signal intensities of protonated and nonprotonated aromatic carbon atoms relative to total carbon signal intensities. The procedure used to determine the different types of carbons [CH3, CH2/CHaliphatic′ CHaromatic, non-protonated C (Cq)] was published elsewhere²,⁴.

    Coals und maceral concentrates

    Coal rank was established by vitrinite random reflectance Rr. Basic compositional and petrographical data of the coals and maceral concentrates are refered to elsewhere⁵,⁶.

    Hydrogenation

    The hydrogenation experiments were performed by a stirred bench scale autoclave of 75 ml volume at a temperature of 708 K, pressure 150 bar, time 2 h.

    Combustion

    The combustion tests were realized with a drop-tube reactor at a temperature of 1173 K, pressure 1 bar, atmosphere 21 % O2 / 79 % N2, droping length 640 mm, time 2,1 s.

    Results and Discussion

    Hydrogenation

    The aromaticities of the feedcoals were found to correlate well with coal rank⁶. The present study shows that the hydrogenation residues are more aromatic than the feedcoals (figure 1). Aromaticity and relative amounts of carbons in different structural groups are nearly the same for all hydrogenation residues independent of rank and composition of the feedcoals. The aromaticities of the hydrogenation residues were determined to 0.82 (lower limit) up to 0.93 (upper limit) whereas fa values of the corresponding coals range from 0.52 up to 0.93. Hydrogenation residues of lignites and subbituminous coals make up the upper limit (fa around 0.9). Hydrogenation residues of bituminous coals have somewhat lower fa values (0.82 up to 0.88). The changes of aromaticity and relative carbon type distribution caused by hydrogenation decrease with increasing coal rank (figure 1). Large changes in the structure of the organic matter as determined by ¹³C nmr are accompanied by high conversion degree. Substantial alterations were observed for carbon atoms in CHaro and CH2/CHali groups. Lignites and subbituminous coals are most reactive whereas anthracites seemed to be nearly unchanged. The relative amount of CHaro increases significantly during hydrogenation. A distinct decrease was noticed for CH2/CHali Alcoholic, etheric, esteric, anhydride and acid structures (¹³C signals at 60–90 ppm e.g. > 150 ppm) present in lignites and subbituminous coals were decomposed during hydrogenation. Further information about the individual components of the whole coal was obtained by hydrogenation of maceral concentrates from five German coals⁵. As expected the main macerals (exinite, vitrinite, inertinite) become more aromatic after hydrogenation. The fa values of the hydrogenation residues of high volatile bituminous coal exinites are lower (fa=0.61 resp. 0.69) than for hydrogenation residues of parent coals and of vitrinites, inertinites and higher rank exinites. According to the differences between the nmr structural parameters the exinite concentrates are more reactive than vitrinite concentrates. Inertinites were less reactive (figure 2). The increase of aromaticity as well as the changes of relative carbon type distribution decrease with increasing rank of the parent coal.

    Figure 1 Aromaticities of feedcoals and hydrogenation residues

    Figure 2 Differences of aromaticities between main macerals and hydrogenation residues. Plot of fa values versus increasing rank (coal rank of the parent coal: 1 − 0,76 % Rr, 2 − 0,78 % Rr, 3 − 0,79 % Rr, 4 − 1,23 %Rr, 5 − 1,46%)

    Combustion

    ¹³C CP/MAS spectra of selected feedcoals from Canada and Germany and of their combustion residues are shown in figure 3. The combustion residues can be divided into three groups:

    Figure 3 ¹³C CP/MAS spectra of feedcoals (A) and combustion residues (B)

    1. combustion residues with fa values higher than 0.9 up to 1.0. The high aromaticity may be caused by the organic matter of the residual coal or possibly may be due to condensation of polycyclic aromatic hydrocarbons formed during combustion. This group comprises the residues from bituminous coals with one exception (1,23 % Rr, 0,85 fa) caused by many macrospores present.

    2. combustion residues with fa values < 0.9 (e.g. 0.67, 0.84). This group comprises the residues from some of the lignites and subbituminous coals as well as from some maceral concentrates. For the latter the low fa values are associated with many microscopically detectable macrospores being more resistent to combustion.

    3. combustion residues that could not be measured by ¹³C CP/MAS nmr. Because cross polarization did not work (very small hydrogen contents possible) direct excitation measurements were carried out. But there were still some combustion residues that could not be measured. These residues behave like conductors due to formation of graphitized structures. Six of the fourteen combustion residues of the maceral concentrates were related this group.

    Conclusions

    The reactivity of coal can be characterized by the differences of fa values and of the relative carbon type distribution between the feedstock and its residual organic matter after reaction.

    Hydrogenation: Lignites and subbituminous coals as well as exinites of high volatile bituminous coals (indicated by low aromaticities) are most reactive. The reactivity decreases with increasing coal rank up to the semi-anthracite stage. Anthracites are least reactive.

    Combustion: During combustion the organic matter has changed considerably and is characterized by high aromaticity. There is only little relationship between nmr behaviour of the combustion residues and coal rank.

    Acknowledgement

    The study was carried out with financial support in part by NATO (International Scientific Exchange Program) and by ECEC (Commission of European Community).

    References

    1. Alemany, L. B., Grant, D. M., Pugmire, R. J., Alger, T. D., Zilm, K. W. J. Am. Chem. Soc.. 1983; 105:2133–2147.

    2. Kasüschke, I., Gerhards, R. Erdöl und Kohle – Erdgas – Petrochemie. 1989; 42:209–212.

    3. Dixon, W. T., Schaefer, J., Sefcik, E., Stejskal, E. O., McKay, R. A. J. Magn. Res.. 1982; 49:341–345.

    4. Opella, S. J., Frey, M. H. J. Am. Chem. Soc.. 1979; 101:5854–5856.

    5. Steller, M., Kasüschke, I., Riepe, W. Erdöl und Kohle – Erdgas – Petrochemie. 1991; 44:34–42.

    6. Kalkreuth, W., Steller, M., Wieschenkämper, I., Ganz, S., FUEL, in press

    Coal as a molecular solid (II)

    Outline

    Chapter 6: SOLVENT SWELLING OF COAL IN BINARY SOLVENT SYSTEM

    Chapter 7: COMPUTER STUDIES OF COAL MOLECULAR STRUCTURE

    Chapter 8: EFFECT OF VARIOUS ADDITIVES ON NON-COVALENT INTERACTIONS IN COALS

    Chapter 9: MOLECULAR STRUCTURE OF BITUMINOUS COALS - NEW INSIGHTS FROM BINARY SOLVENT STUDIES

    Chapter 10: THE ROLE OF MOISTURE IN THE MACROMOLECULAR STRUCTURE OF COALS

    SOLVENT SWELLING OF COAL IN BINARY SOLVENT SYSTEM

    M Yoshihara* and T Yonezawa**,     *Department of Applied Chemistry, Faculty of Science and Technology Kinki University, Kowakae, Nigashi-Osaka, Osaka 577, Japan; **Institute of Science and Technology, Kinki University Kowakae, Nigashi-Osaka, Osaka 577, Japan

    T Aida, Y Shimoura, N Yamawaki and M Fujii,     Department of Industrial Chemistry, Faculty of Engineering, Kinki University in Kyushu, Iizuka, Fukuoka 820, Japan

    Publisher Summary

    This chapter reviews solvent swelling of coal in binary solvent system. Illinois No. 6 coal from the Ames Laboratory Coal Library was used as a coal sample, which was ground, seized, dried at 110°C under vacuum overnight, and stored under a dry nitrogen atmosphere. The cross-linked synthetic polymers, styrene-divinylbenzene co-polymers, were commercial products of BIORAD Laboratory. All solvents used were purified by ordinary procedures, except Aldrich’s Gold Label Grade reagents, which were used without further purification. The solvent-induced swelling measurements were carried out in a modified version of the instrument. The chapter presents the swelling parameters of Illinois No. 6 coal in various solvents. A wide variety of the equilibrium swelling values (Q) and the initial swelling rates (V) were observed. These parameters seem to be controlled by the solvent’s affinity with coal and the steric hindrance of the solvent molecule. The chapter discusses the relation between the equilibrium swelling value (Q) and the ratio of the components of the mixed solvent, methanol/triethylamine and N,N′-dimethylformamide/hexamethylphophoramide (HMPA). Large synergistic effects were observed in both systems.

    1 INTRODUCTION

    Coal swells in various solvents. This fundamental nature of coal is thought to be one of the clear evidence that coal is a macromolecular solid cross-linked with chemical bonds, such as covalent, hydrogen, and charge-transfer bonds. The cross-linking structure of coal causes steric requirement for penetrating molecules. A bulky molecule penetrates into solid coal more slowly than a less bulky one. It was found that in the case of Illinios #6 coal, the diffusion rate of penetrant molecule is decreased by almost thousand folds simply by changing the alkyl substitutent from normal to tertiary. ¹ There are a lot of studies on the chemical transformation of coal such as gasification and liquifaction. Among them, the importance of the accessibilities of chemical reagents and/or solvents to the site buried within solid coal are often ignored. It is obvious that the fundamental understanding of the diffusion mechanism of penetrant molecules into coal is vital to develop an efficient chemical transformation process of coal. A binary solvent system, such as alcohol-benzene, has long been well known to be efficient for the extraction of subbituminous and brown coals,² and to induce a synergistic effect on the extractabilty.³ Some years ago, a powerfull mixture, carbon disulfide and N-methyl-2-pyrrolidinone, for the coal extraction has been developed by Iino et al.⁴ The mixture of trimethylamine (or dimethylaniline) and methanol was once used for examining a solubility parameter of coal based on the synergistic effect on the coal swelling.⁵ Larsen et al.⁶ pointed out the complication of coal swelling in binary solvent system, particularly in the case of use of a base which is able to cleave a hydrogen bonding in macromolecular network of coal. It is obvious that all of these phenomena must be considered on the basis of the interaction between coal and the individual solvent molecules. On this view point, there seem to be some confusions to rationalize the experimental observations in previous literatures. In general, the use of equiliblium swelling values (Q) makes difficult to evaluate the net contribution from the individual component to observing phenomenon. We have examined the swelling behaviors of coals and cross-linked synthetic polymers, styrene-divinylbenzene co-polymer, in binary solvent systems with both equiliblium swelling value(Q) and swelling kinetics including the initial swelling rate (V) which could not easily be measuered by earlier tachniques, and it has found that a synergistic effect on coal swelling was induced in the binary solvent mixture composed of a sterically bulky solvent and a less bulky one.

    2 EXPERIMENTAL

    Illinois #6 coal from the Ames Laboratoy Coal Library was used as a coal sample, which was ground, seized, dried at 110 °C under vacuum overnight, and stored under a dry nitrogen atmosphere. The cross-linked synthetic polymers, styrene-divinylbenzene co-polymers, were commercial products of BIO·RAD Laboratory(Bio-Beads SX series; 200–400 US mesh). All solvents used were purified by ordinary procedures, except Aldrich’s Gold Label Grade reagents, which were used without further purification. The solvent induced swelling measurement were carried out in a modified version of the instrument previously described.⁷ The equiliblium swelling value (Q) is difined as the apparent volume of the swollen coal at the equliblium divided by the apparent volume of the unswollen coal. The initial swelling rate (V) is obtained from the dynamic swelling curve by the method previously reported.⁷

    3 RESULTS AND DISCUSSION

    3.1 Solvent swelling of coal

    Table 1 summarizes the swelling parameters of Illinois #6 coal in various solvents. A wide variety of the equiliblium swelling values (Q) and the initial swelling rates(V) are observed. These parameters seem to be controled by the solvent’s affinity with coal and the steric hindrance of the solvent molecule.

    Table 1

    Swelling Parameters of Illinois #6 Coal in Various Solvents

    a100–200 US mesh coal; 20.0±0.5 °C

    bMeasured after 10 days swelling

    Based on these data, the coal swelling in the mixture of methanol and triethylamine (Hombach’s system) is expected the discriminating penetration of solvent molecules, because the diffusion rate of methanol is almost ten thousand fold higher than that of triethylamine. Thus, methanol will preferentially penetrate into the solid coal at the initial stage of swelling, and form a methanol-swollen coal, so-called methanol-coal gel. Such a gel formation is expected to induce a significant relaxation of the steric requirement of coal initially existed, expanding the cross-linking structure.

    Fig 1 shows the relation between the equiliblium swelling value (Q) and the ratio of the components of the mixed solvent, methanol/triethylamine and DMF/HMPA. Large synergistic effects are observed in both systems. Meanwhile, Fig 2 shows the results which the coal swelling was carried out in the mixture of methanol and n-butylamine. There is not significant difference between each solvent’s diffusion rates. Obviously, a synergistic effect is induced in the case of the binary mixture in which the steric bulk of one component is larger than another. These observations are consisted with the relaxation of the steric requirment of coal by the initial formation of coal-gel.

    FIG. 1

    FIG. 2

    3 Solvent swelling of synthetic polymer

    In order to generalize the concept introduced by the experiments described above, The simulation experiment was carried out by using cross-linked synthetic polymers, styrene-divinylbenzene co-polymer (Bio-Beads SX series, BIO·RAD Laboratory). Table 2 lists some of the swelling parameters of Bio-Bead SX-4(divinylbenzene 4%) in various solvents. A large steric effect on the swelling is observed as well as coal, It is more evident on the swelling rate (V) than the equiliblium swelling value (Q). The high diffusion rate of benzene compared to those of butylbenzene derivatives is thought to be involved the steric effect and the affinity with the polymer.

    Table 2

    Swelling Parameters of Bio-Beads SX-4a in Various Solvents

    aStyrene-divinylbenzene (4%) copolymer; 200–400 US mesh

    bMeasured after 12 hours swelling at 20. 0±0. 5 °C

    Fig 3 and 4 demonstrate the correlation between the equiliblium swelling value (Q) and the component ratio of the mixed solvent, n-butylbenzene/benzene and tert-butylbenzene/benzene mixed systems, respectively.

    FIG. 3

    FIG. 4

    The synerigistic effect on solvent swelling is observed in the binary mixture, which is composed of a sterically bulky solvent and a less bulky one. However, in this case the drastic increase of the Q-value is appeared at the very low concentration of less bulky solvent. This means that the relaxation of the steric requirment in synthetic polymer takes place quite easily, compared to coal, of which macromolecular network is thought to be composed of not only covalent bonding, but also non-covalent bondings such as hydrogen, and charge-transfer bondings.

    4 CONCLUSIONS

    One of the synergistic effect induced in the binary solvent swelling is rationalized by the relaxation of steric requirement of coal through the initial solvent-coal gel formation. This phenomena used to observed in the mixed solvent system composed with a sterically bulky solvent and a less bulky one.

    5. REFERENCES

    1. Aida, T, Fuku, K., Fujii, M., Yoshihara, M., Squires, T. G. Energy & Fuels. 1991; 5:79–83.

    2. Wielopolski, A., Witt, I. Koks. Smola. Gaz.. 1977; 22:47. [CA, 1977, 87, 138440w].

    3. Iino, M., Matsuda, M. Fuel. 1983; 62:744.

    4. Iino, M., Takanohashi, T., Ohsuga, H., Toda, K. Fuel. 1988; 67:1639–1647.

    5. Hombach, P. H. Fuel. 1980; 59:465–470.

    6. Larsen, J. W., Green, T. K. Fuel. 1984; 63:465.

    7. Aida, T., Squires, T. G. Prepr. Pap. -Am. Chem. Soc., Div. Fuel Chem.. 1985; 30:95–98.

    COMPUTER STUDIES OF COAL MOLECULAR STRUCTURE*

    G.A. Carlson,     Fuel Science Division, Sandia National Laboratories, Albuquerque, N. M. 87185

    Publisher Summary

    This chapter discusses computer modeling of molecular structures based on computer graphics techniques and force field methods, which have begun to be applied to the area or fuel science, including work on kerogen macromolecules and catalyst systems. These computer modeling techniques, often identified as computer-aided molecular design (CAMD) techniques, are also providing new insights into coal structure and the importance of structural features in conversion processes. In initial CAMD work on bituminous coal, molecular models proposed by Given, Wiser, Solomon, and Shinn were constructed and minimum-energy structures were determined. Based on a more detailed study of the energetics, van der Waals and hydrogen-bonding interactions were both found to be very important in defining the three-dimensional minimum-energy structures. The chapter reviews the structural features of coal when viewed as a molecular solid. It also highlights computer modeling studies of coal structure incorporating both molecular and network characteristics.

    1 INTRODUCTION

    The reactivity of coal when utilized as a fuel or as a feedstock for the production of liquid or gaseous products is intimately related to its molecular and structural properties. In the past, bituminous coal has been defined on the basis of average molecular structure (1–4) or as a three-dimensionally cross-linked (network) solid (5–8). The molecular structures cited contain from 200 to 1300 atoms, and are based on data from pyrolysis and liquefaction reactions, solvent interactions, and analytical techniques. In contrast, the network models are constructed from molecular clusters (generally fused aromatic ring systems defined by an average molecular weight but including no molecular details), joined together in a three-dimensional structure by covalent bonds. The behavior of the network models is

    Enjoying the preview?
    Page 1 of 1