A Comparative Thermodynamic Study of Equilibrium ... - MDPI

1 downloads 0 Views 2MB Size Report
May 7, 2018 - higher O/C values to the one used in predictions should be applied to avoid ... The presented thermodynamic analyses were carried out for a range .... In the following two sections, the detailed explanation is presented for the ...
energies Article

A Comparative Thermodynamic Study of Equilibrium Conditions for Carbon Deposition from Catalytic C–H–O Reformates Zdzisław Jaworski *

ID

and Paulina Pianko-Oprych

Department of Chemical Technology and Engineering, West Pomeranian University of Technology, Szczecin, 71065 Szczecin, Poland; [email protected] * Correspondence: [email protected]; Tel.: +48-91-449-4020 Received: 3 April 2018; Accepted: 2 May 2018; Published: 7 May 2018

 

Abstract: The importance of carbon deposition occurring during catalytic fuel reforming is briefly described along with former studies on the process. Thermodynamic fundamentals of modeling the critical conditions of the deposition equilibrium are presented. Computational results of ternary C–H–O diagrams with the threshold lines between the carbon deposition and deposition-free regions are discussed for two new pressure levels of 3 and 30 bar and a temperature range from 200 to 1000 ◦ C. The process pressure does not affect the temperature range typical for the type of deposited carbon allotrope; either graphite, multi-walled carbon nanotubes, or single-walled carbon nanotubes in bundles. However, pressure has a profound influence on the location of the threshold lines for carbon deposition. Three reforming processes of two hydrocarbon fuels are analyzed; catalytic partial oxidation, and wet and dry reforming. Chord lines representing varied compositions of process mixtures are introduced to the ternary diagrams. The intersection points of the chord lines with the threshold lines are used in a novel interpretation of the functions of the oxygen-to-carbon critical ratio against temperature and pressure, which can be used in avoiding carbon deposition in catalytic reforming of natural gas and liquefied petroleum gas. Keywords: carbon deposition; hydrocarbon reformates; thermodynamic equilibrium; C–H–O diagrams

1. Introduction The demand for useful energy is systematically growing as the world aims to improve its standard of living. The burning of all fossil fuel forms has serious environmental consequences. A promising alternative of fossil fuels for the future is seen in hydrogen. Energy produced by hydrogen-based fuel cells is characterized as high-quality energy, because of the high conversion efficiency from chemical to electrical energy. The efficiency is almost twice as high compared to internal combustion engines and because of the zero emissions of pollutants and greenhouse gases to the environment, hydrogen fuel cells are currently the promoted form. Hydrogen can be obtained using several technologies: steam reforming (SR) [1]; auto-thermal reforming (ATR) [2]; dry reforming (DR) [3]; or catalytic partial oxidation reforming (CPOX ) [4]. Catalytic steam reforming is the most common process due to its high energy efficiency. This technology has been widely applied in chemical industries for large-scale hydrogen production, which accounts for 50% of the hydrogen generated worldwide by using methane from natural gas (NG) as the main hydrocarbon source [5]. It is a well-known process described recently by Adiya et al. [6] and Tuna et al. [7]. The reactions that occur in this process primarily produce H2 , CO2 , and CO, but there is no set quantity for these compounds. The compounds’ concentrations depend on several factors, such as reagent concentration, temperature, and pressure of the reformer, as well as the physical and

Energies 2018, 11, 1177; doi:10.3390/en11051177

www.mdpi.com/journal/energies

Energies 2018, 11, 1177

2 of 12

chemical characteristics of the chosen catalyst. The overall reaction that occurs in the steam reforming process takes place at temperatures between 650 and 850 ◦ C obtaining H2 yields of 60–70% at a H2 /CO ratio equal to 3.0 [7]. In addition to the main reforming reaction, other reactions can simultaneously occur and modify the equilibrium conversion of CO2 and CH4 . Boudouard’s reaction leads to carbon formation by carbon monoxide decomposition. This reaction is very important for the reforming process since it is responsible for solid carbon deposition in catalysts. It occurs when the decomposition reaction of carbon monoxide, together with the decomposition reaction of methane, are faster than the carbon removal rate [8]. The formation of carbon deposits on catalysts leads to a blockage of catalyst pores by the deposited carbon, separation of the catalyst from its support, and lack of gas flow due to an increase in pressure caused by pore blockage [9]. The carbon formation may take place by three routes that result in different kinds of coke [10]. At low temperatures, less than 500 ◦ C, adsorbed hydrocarbons may accumulate on the nickel surface and slowly polymerize into an encapsulating film, blocking and deactivating the nickel surface. At high temperature, above 600 ◦ C, pyrolytic coke formed by the thermal cracking of hydrocarbons may encapsulate and deactivate the catalyst particle. At temperatures greater than 450 ◦ C, whisker (filamentous) carbon is the principal product of carbon formation via a mechanism involving the diffusion of carbon through nickel crystals, nucleation, and whisker growth with a nickel crystal on the top. The whisker-type carbon does not deactivate the nickel surface, but rather causes a breakdown of the catalyst by pore plugging [10]. Ginsburg et al. [11] calculated the correlation between the CO2 /CH4 ratio and the process of carbon formation on the surface of Ni-based catalysts. According to their model, the reactant ratio should be equal to 2 in order to provide less carbon accumulation. The oxidizing system must have an excess of CO2 as an oxidant. The modeling of carbon deposition from lower hydrocarbon fuels at thermodynamic equilibrium was widely discussed by Jaworski et al. [12,13]. A minimization algorithm of the system’s total Gibbs energy was applied using the commercial software HSC for both dry and wet reforming processes within the range of 200 ◦ C to 1000 ◦ C and for four levels of constant pressure. Five carbon solid allotropes, such as graphite, diamond, amorphous, multi-walled carbon nanotubes (MWCNT), and single-walled nanotubes in bundles (SWCNT), were allowed to appear in the solid phase. However, no indication of deposits of diamond and amorphous carbon were found for the analyzed systems. The computation results of the deposition conditions at thermodynamic equilibrium showed for the wet reforming of NG and LPG that the maximum critical atomic oxygen to carbon (O/C) ratios were found at 2.0, 2.4, 2.7, and 3.2 for pressures of 1, 3, 10 and 30 bar, respectively. It was concluded that higher O/C values to the one used in predictions should be applied to avoid nanotubular carbon depositions [13]. In addition, the threshold temperature of 431 ◦ C or 577 ◦ C for the deposition boundaries between graphite and MWCNT or MWCNT and SWCNT were confirmed regardless of pressure. Another available option for hydrogen-rich fuel production beyond steam or CO2 reforming is catalytic partial oxidation (CPOx ). A thermodynamic analysis of equilibrium conditions of hydrogen production from natural gas and LPG using the CPOx reformer, with reference to carbon deposition phenomena, was performed by Jaworski and Pianko-Oprych [14]. Methane catalytic reforming in the range of 1 < O/C < 4 resulted in the formation of solid carbon deposits. However, only a single form of depositing carbon was predicted at a time if the O/C ratio fell below the critical value. Depending on the temperature, it was either graphite from 200 to 431 ◦ C, or one of the two carbon nanotubes. Multi-walled carbon nanotubes were found most stable in the range from 431 to 577 ◦ C and single-walled nanotubes in bundles prevailed from 577 up to 1000 ◦ C. The maximum critical O/C ratio was about 2.7 or 2.9 for methane with air or pure oxygen as the oxidant, respectively, while for Liquefied petroleum gas (LPG)–oxidant pairs the O/C maxima were lower by about 0.2 [14]. The presented thermodynamic analyses were carried out for a range of fuels indicating the predominant role of temperature in solid carbon formation. The carbon element had different

Energies 2018, 11, 1177

3 of 12

allotropic forms in the gas and solid phases and, therefore, the electrochemical data for both purified single-walled and multi-walled CNTs were critical in predicting the conditions of carbon deposition. The key objective of the paper is to present the development of an integrated approach for understanding the mechanism of carbon deposition from catalytic C–H–O reformates. This is envisaged through a theoretical modeling presented in [12–14] over an extended range of operating conditions. To achieve this objective, a thermodynamic analysis on carbon deposition from C–H–O reformates will be expanded for the total pressure of 3 bar and 30 bar and compared with the previous results [12] obtained for 1 bar and 10 bar. Differences in the deposition boundaries for all solid carbon forms predicted for the catalytic partial oxidation, and dry and wet reforming of the NG and LPG will be explained based on the deposition boundaries in the ternary C–H–O diagrams for different temperatures and pressures. The results of this study will quantify the relationship between carbon formation and reforming system conditions promoting coke deposition. Such a comparative analysis of equilibrium deposition for the CPOx , and dry and wet reforming of lower hydrocarbons in NG and LPG will allow the indication of the critical temperature for the deposition at chosen process pressures and reactant compositions. 2. Results This section presents the computational results of threshold atomic ratios of carbon, oxygen, and hydrogen, which separate regions of single-phase gas mixture concentrations from two-phase regions containing solid carbon deposits under thermodynamic equilibrium. The results are presented in C–H–O ternary diagrams as the isothermal threshold lines for the temperature range from 200 to 1000 ◦ C and two process pressures of 3 and 30 bar. The modeling is carried out separately for the deposition of graphite only and when deposition of carbon nanotubes is also allowed. In addition, the diagrams are supplied with chord lines drawn between points representing compositions of two fuels and three oxidants that represent different compositions of reformates. Jointly with the results of our former study [12] of ternary diagrams for 1 and 10 bar, those diagrams are used for the interpretation and discussion of the critical O/C values for the fuel-oxidant pairs in the full temperature and pressure ranges. 2.1. Ternary Diagrams for 3 and 30 Bar Additional modeling, explained in Section 4, is carried out to construct the critical lines for the two missing ternary diagrams, for 3 and 30 bar. Like in our previous studies, only three allotropes of solid carbon result from the computations, namely graphite up to 431 ◦ C, then multiwalled carbon nanotubes up to 577 ◦ C, and from there to 1000 ◦ C, only single-walled carbon nanotubes were found. Since this study aims to explain differences in the critical O/C ratios for different fuels and oxidants, their composition points are also shown in the diagrams. It can be observed that the computed threshold lines, for temperatures from 200 ◦ C to 1000 ◦ C in Figure 1a,b, exhibit a similar evolution to that shown in corresponding graphs for 1 and 10 bar, published in [12]. Most pronounced differences, for the tested pressure range, are found for the highest pressure of 30 bar (Figure 1b). They are mainly in the region of fuel-oxidant mixtures of low hydrogen content, on the right side of the diagrams, where the temperature effect on the threshold lines is minimal. Two types of fuel: natural gas (containing about 94% of CH4 ) or liquefied petroleum gas (containing mainly C3 H8 and C4 H10 ) are used in this study. Depending on the type of the reforming process, a fuel reacts with one of three types of oxidant; oxygen, O2 , from air, water steam, H2 O, or carbon dioxide, CO2 . The composition of a given mixture of the fuel–oxidant pair is represented in ternary diagrams by points located along the straight line connecting the fuel point at the left triangle edge with the oxidant point located on the opposite side. A crossing point of the fuel-oxidant line with the relevant threshold line determines the critical composition of the mixture in terms of the C, H, and O atomic ratios. Raising the C content in the mixture from the crossing point means entering

Energies 2018, 11, 1177

4 of 12

the carbon deposition region. The location of the fuel-oxidant lines relative to the relevant deposition threshold line strongly influences the course of the critical O/C curves versus temperature. Energies 2018, 11, x

4 of 11 0,0 0,1

0,9

0,2

0,8

0,3

0,7

0,4

H

Carbon deposition region

0,5

H

C

0,5

0,6

0,7

0,8

0,4 CO2 0,3 0,2

No carbon deposition

0,1

1,0

0,0 O 1,0 2

0,9

C

0,5

0,9

0,1 0,3H2O0,4

0,6

CH4 0,8

0,2

No carbon deposition

0,2

0,5

0,7 C3H8+C4H10

1,0 0,1

0,7

Carbon deposition region

0,6

CO2 0,3

0,9

0,8

0,3

0,4

0,7 C3H8+C4H10 0,8 CH4

0,9

0,2

0,4

0,5

1,0

0,1

0,6

0,6

0,0

0,0

200 oC 300 oC 400 oC 500 oC 600 oC 700 oC 800 oC 900 oC 1000 oC

1,0

0,0

0,1

H O 0,3 2 0,4

0,2

0,5

O

0,6

0,7

0,8

0,0 O 1,0 2

0,9

O

(a)

(b)

Figure 1. Ternary C–H–O diagrams for: (a) pressure of 3 bar; and (b) pressure of 30 bar. Figure 1. Ternary C–H–O diagrams for: (a) pressure of 3 bar; and (b) pressure of 30 bar.

In almost almost all all published published results results of of computation computation of of carbon carbon deposition, deposition, graphite graphite was was the the only only solid solid In carbon allotrope allotrope predicted predicted [12]. [12]. It It is, is, therefore, therefore, informative informative to to compare compare the the threshold threshold lines lines for for either either carbon graphite only, or all possible carbon allotropes independent of temperature. The effect of allowing for graphite only, or all possible carbon allotropes independent of temperature. The effect of allowing the filamentous forms of carbon deposits is represented in Figure 2a,b for the process pressure of 3 bar for the filamentous forms of carbon deposits is represented in Figure 2a,b for the process pressure of Figure 3a,b for 30 bar. The threshold lines arelines shown pairs in of pairs constant temperature, with the 3and barinand in Figure 3a,b for 30 bar. The threshold are in shown of constant temperature, solid lines representing all carbon forms and the broken lines ascribed to graphite. Starting from the with the solid lines representing all carbon forms and the broken lines ascribed to graphite. Starting ◦ ◦ temperature of 500 C, of only three temperature lines with 200lines C increments areincrements shown in each from the temperature 500 °C, pairs only of three pairs of temperature with 200 °C are diagram, for the sake of clear presentation. shown in each diagram, for the sake of clear presentation. 0.0 0.1

0.9

0.2

0.8

0.3 0.4

H

0.5

0.4

0.6 0.5

H

C CO2 0.3

No carbon deposition

0.2

H O 0.3 2 0.4

0.5

O

(a)

0.6

0.7

0.8

0.8

0.9

0.7

Carbon deposition region

0.6 0.5

0.0 O 1.0 2

C 0.4 CO2 0.3

No carbon deposition

0.2

0.9

0.1

1.0 0.1

0.9

0.7 C3H8+C4H10 CH4 0.8

0.2

0.9

0.5

600 C All C o 600 C Gra o 800 C All C 800 oC Gra 1000 oC All C 1000 oC Gra

0.6

0.4

0.7 C3H8+C4H10 CH4 0.8

o

0.2

0.7

Carbon deposition region

1.0

0.1

0.3

0.6

0.0

0.0

500 oC All C 500 oC Gra 700 oC All C 700 oC Gra 900 oC All C 900 oC Gra

1.0

0.1

1.0 0.0

0.1

0.2

0.3H2O0.4

0.5

0.6

0.7

0.8

0.9

0.0 O 1.0 2

O

(b)

Figure threshold lines for for the process pressure of 3 bar, for graphite only— Figure 2. 2. Deposition Deposition threshold lines the process pressure of 3separately bar, separately for graphite broken lines, and all carbon allotropes—solid lines: (a) temperatures of 500, 700, and 900 °C; and ◦ C; only—broken lines, and all carbon allotropes—solid lines: (a) temperatures of 500, 700, and 900 (b) temperatures of 600, 800, and 1000 °C. and (b) temperatures of 600, 800, and 1000 ◦ C.

Energies 2018, 11, 1177

5 of 12

Energies 2018, 11, x Energies 2018, 11, x

5 of 11 5 of 11 0.0

1.0

0.1

0.9

0.1

0.9

0.2

0.8

0.2

0.8

0.3

Carbon deposition Carbon region deposition region

H 0.5 H0.6 0.5

0.4

0.5

H 0.5 H0.6 0.5

0.4

0.2 0.2 0.1

0.9

0.1

1.0 0.0

0.1

0.2

H O 0.3 2 0.4 H O 0.3 2 0.4

0.5

0.6

0.7

0.8

0.9

0.5

0.6

0.7

0.8

0.9

O O

0.8 0.7 0.7

Carbon deposition Carbon region deposition

0.4

0.6 0.6

0.0 O2 1.0 0.0 O2 1.0

C C

0.5

region

0.5

0.4

0.6

0.4 CO2 CO0.3 2 0.3

No carbon deposition No carbon deposition

0.2

0.8

0.2

C C

0.5

0.9

0.3

0.6

600 oC All C 600 ooC C All GraC 600 800 ooC C Gra All C 600 o 800 oC C All GraC 800 o o C All C 1000 800 C Gra 1000 ooC C All GraC 1000 1000 oC Gra

0.9

0.1

0.3

0.6

1.0 1.0

0.2

0.4

0.6

0.1

0.1

0.7

0.7 C3H8+C4H10 0.7 C3CH H8+C0.8 H 4 4 10 CH4 0.8 0.9

1.0 0.0

0.0

900 oC Gra

0.7

0.3 0.4

0.0

500 oC All C 500 ooC C All GraC 500 700 ooC C Gra All C 500 o o 700 C C All GraC 700 900 ooC C Gra All C 700 900 ooC C All GraC 900

1.0

0.0

CO2 0.4 CO0.3 2 0.3

0.7 C3H8+C4H10 0.7 C3CH H8+C0.8 H 4 4 10 CH4 0.8 0.9

No carbon deposition No carbon deposition

0.2 0.2 0.1

0.9

0.1

1.0 1.0 0.0

0.1

0.2

0.0

0.1

0.2

H O 0.3 2 0.4 H O 0.3 2 0.4

(a) (a)

0.5

0.6

0.7

0.8

0.9

0.5

0.6

0.7

0.8

0.9

O O

0.0 O2 1.0 0.0 O2 1.0

(b) (b)

Figure 3. Deposition separately for Figure 3. Deposition threshold threshold lines lines for for the the process process pressure pressure of of 30 30 bar, bar, separately for graphite graphite only only Figure 3. Deposition threshold linesallotropes for the process pressure lines: of 30 (a) bar,temperatures separately for graphite only (Gra)—broken lines, and all carbon (All C)—solid of (Gra)—broken lines, and all carbon allotropes (All C)—solid lines: (a) temperatures of 500, 500, 700, 700, and and (Gra)—broken lines, and all carbon allotropes (All C)—solid lines: (a) temperatures of 500, 700, and 900 ◦ C; and ◦ C. 900 °C; and (b) (b) temperatures temperatures of of 600, 600, 800, 800, and and 1000 1000 °C. 900 °C; and (b) temperatures of 600, 800, and 1000 °C.

It follows from Figures 2 and 3 that the maximum effect of the presence of filamentous carbon, It follows from Figures 22and that the maximum effect of filamentous carbon, as from Figures and33in that maximumdeposits, effectofofthe thepresence presence filamentous carbon, as an alternative to solely graphite the the equilibrated was computedoffor 800 °C or 1000 °C ◦ ◦ an alternative to solely graphite in the equilibrated deposits, waswas computed for for 800800 C or 1000 C for as to3solely in the equilibrated deposits, computed °C 1000 °C foran thealternative pressure of or 30 graphite bar, respectively. For the purpose of highlighting that effect, theor threshold the pressure of 3 or 30 bar, respectively. For the purpose of highlighting that effect, the threshold lines for the of 3 or 30 bar,levels respectively. Forin the purpose ofseparately highlighting effect, the threshold lines forpressure the applied pressure are shown Figure 4a,b, forthat the two temperatures. for the pressure levelslevels are shown in Figure 4a,b, separately for thefor two lines forapplied the applied pressure are shown in Figure 4a,b, separately thetemperatures. two temperatures. 0.0 0.0 0.1

0.0

1.0 0.9

0.1

0.9

0.2

0.4

0.1

0.6 0.6 0.5 0.5

C C

H 0.5 H0.6 0.5

0.4

0.2 0.2 0.1

0.9

0.1 0.2

0.0

0.1

0.2

0.3H2O0.4 0.3H2O0.4

0.5

0.6

0.7

0.8

0.9

0.5

0.6

0.7

0.8

0.9

O O

(a) (a)

0.7

0.4

Carbon deposition Carbon region deposition region

0.7 0.6 0.6 0.5 0.5

0.0 O2 1.0 0.0 O2 1.0

C C 0.4

0.6

0.4 CO2 CO0.3 2 0.3

1.0 0.1

0.8

0.3 0.4

No carbon deposition No carbon deposition

1.0 0.0

0.8

0.2

0.6 0.7 C3H8+C4H10 0.7 C3CH H8+C0.8 H 4 4 10 CH4 0.8 0.9

0.9

0.3

0.7

1 bar bar 13 bar bar 310bar 30 bar 10 bar 30 bar

0.9

0.1

0.7

Carbon deposition Carbon region deposition region

1.0 1.0

0.2

0.8

0.3

H 0.5 H0.6 0.5

0.0

30 bar

0.8

0.2 0.3 0.4

1 bar bar 13 bar bar 310bar 30 bar 10 bar

1.0

0.4 CO2 0.3 CO2 0.3

0.7 C3H8+C4H0.7 10 C3CH H8+C 0.8H 4 4 10 CH4 0.8 0.9

No carbon deposition No carbon deposition

0.2 0.2 0.1

0.9

0.1

1.0 1.0 0.0

0.1

0.2

0.0

0.1

0.2

0.3H2O0.4 0.3H2O0.4

0.5

0.6

0.7

0.8

0.9

0.5

0.6

0.7

0.8

0.9

O O

0.0 O2 1.0 0.0 O 1.0 2

(b) (b)

Figure 4. Deposition threshold lines for the border process pressures of 1 and 30 bar: (a) temperature Figure 4. threshold lines for the Figure 4. Deposition Deposition threshold of lines for°C. the border border process process pressures pressures of of 11 and and 30 30 bar: bar: (a) (a) temperature temperature of 800 °C; and (b) temperature 1000 of 800 °C; and (b) temperature of 1000 °C. of 800 ◦ C; and (b) temperature of 1000 ◦ C.

A direct conclusion that follows from the examination of the two diagrams is that rising of the A direct conclusion that temperatures follows from the examination the two diagrams is that rising the reforming pressure for those results in of lowering critical carbon atomicof A direct conclusion that follows from mostly the examination of the twothe diagrams is that rising ofratio the reforming pressure for those temperatures mostly results in lowering the critical carbon atomic ratio of the reformed mixture. However, the opposite can noticedthe forcritical reformed mixtures ofratio a high reforming pressure for those temperatures mostlyeffect results inbe lowering carbon atomic of of the reformed mixture. However, the0.66. opposite effect the can effect be noticed for reformed mixtures of a high hydrogen atomic ratio, However, for about H Therefore, of temperature onmixtures the critical the reformed mixture. the> opposite effect can be noticed for reformed of values a high hydrogen atomic ratio, for about 0.66. Therefore, the effect temperature on the critical values of the carbon, C, ratio, and oxygen, O,HH is>>not wetofof reformates. pressure hydrogen atomic for about 0.66.monotonic Therefore,for thethe effect temperatureThe on the criticaleffects valuesare of of the carbon, C, and oxygen, O, is not monotonic for the wet reformates. The pressure effects are smaller for the lower limit of studied temperature, i.e.,reformates. below 500 °C, theeffects most stable form the carbon, C, and oxygen, O, the is not monotonic for the wet Thewhere pressure are smaller smaller foristhe lower limit of the studied temperature, i.e., below 500 °C, where the most stable form of graphite. forcarbon the lower limit of the studied temperature, i.e., below 500 ◦ C, where the most stable form of carbon of carbon is graphite. In the following two sections, the detailed explanation is presented for the published [13,14] is graphite. In the following two sections, explanation is presented for the This published [13,14] different courses of the critical O/C the ratiodetailed for the analyzed fuel-oxidant mixtures. is considered different of natural the critical for thepetroleum analyzed fuel-oxidant mixtures. is considered separatelycourses for the gasO/C andratio liquefied gas. Different shapes This of the published separately for the natural gas and liquefied petroleum gas. Different shapes of the published

Energies 2018, 11, 1177

6 of 12

In the following two sections, the detailed explanation is presented for the published [13,14] different courses Energies 2018, 11, x of the critical O/C ratio for the analyzed fuel-oxidant mixtures. This is considered 6 of 11 separately for the natural gas and liquefied petroleum gas. Different shapes of the published threshold threshold O/C curves can be by justified by comparing the crossing of thetemperature critical temperature O/C curves can be justified comparing the crossing points ofpoints the critical lines in linesternary in the diagram ternary diagram with arelevant chord relevant to the fuel–oxidant For instance, theline chord the with a chord to the fuel–oxidant pair. Forpair. instance, the chord for line for C 3 H 8 + C 4 H 10 (LPG) – CO 2 runs across the diagram almost horizontally, which means the C3 H8 + 4 10 (LPG) – CO2 runs across the diagram almost horizontally, which means the critical critical ratio, carbon is thenconstant. roughlyHowever, constant.the However, the critical oxygen atomic ratio, O, carbon C, ratio, is thenC, roughly critical oxygen atomic ratio, O, increases with increases with increasing temperature. In that case of the dry LPG reforming, the critical O/C ratio increasing temperature. In that case of the dry LPG reforming, the critical O/C ratio lowers with rising lowers with On rising On the contrary, the chord run linesinfor reformates runtwo in fuel the temperature. the temperature. contrary, the chord lines for wet reformates the wet diagrams from the diagrams the two fuel points (CH 4high or LPG) to “H 2O”,of through high atomic ratios of hydrogen, points (CHfrom or LPG) to “H O”, through atomic ratios hydrogen, H, and are almost parallel to 4 2 ◦ H, and are almost parallel to the threshold lines of low temperature range from 200 to 400 °C. the threshold lines of low temperature range from 200 to 400 C. 2.2. Natural Gas Reformates A cumulative graph is presented in Figure 5a for three reformates of natural gas (NG): from catalytic partial oxidation [13], water steam reforming, and carbon dioxide [14]. Computed data for two border levels of the process pressure, pressure, either either 1 bar (solid lines) or 30 bar (broken lines), are chosen for comparison. 1 bar CPOx 30 bar CPOx 1 bar wet refor. 30 bar wet refor. 1 bar dry refor. 30 bar dry refor.

3

2

1

0 200

GRAPHITE

300

400

MW CNT

500

700

800

o

Temperature [ C]

(a)

3

2

1

SW CNT

600

1 bar CPOx 30 bar CPOx 1 bar wet refor. 30 bar wet refor. 1 bar dry refor. 30 bar dry refor.

4

O/C atomic ratio

O/C atomic ratio

4

900

1000

0 200

GRAPHITE

300

400

MW CNT

500

SW CNT

600

700

800

900

1000

o

Temperature [ C]

(b)

Figure 5. Critical O/C lines for the catalytic partial oxidation reforming (CPOx), wet and dry Figure 5. Critical O/C lines for the catalytic partial oxidation reforming (CPOx), wet and dry reformates reformates of: gas; (a) natural and (b) liquefiedgas. petroleum gas. of: (a) natural and (b)gas; liquefied petroleum

In the range of graphite deposition, pressure has almost no effect on the threshold O/C values In range of graphite deposition, pressure almost effect5a. onThis the threshold O/C values for for the the CPOx or dry reforming of natural gas, ashas shown in no Figure is a consequence of only the CPOx or dry reforming of natural gas, as shown in Figure 5a. This is a consequence of only small small differences in the critical line locations in the ternary diagrams for temperatures below◦500 °C differences in the critical the ternary forpressure. temperatures below 500 isCvalid and and the hydrogen atomicline ratiolocations H < 0.6, in irrespective ofdiagrams the process This conclusion the hydrogen atomic ratio H pressure 0.66), results a considerable influence of chords from “CH ” to “H O” that lie in the range of H > 0.66), results in a considerable influence 4 2 pressure. In the temperature range below 700 °C, the critical O/C values are smaller with the of pressure.pressure. In the temperature range below 700 is◦ C, the critical values smallertypes, with i.e., the increasing However, above 700 °C there a similar trendO/C to the other are reforming ◦ increasing pressure. However, above 700 C there is a similar trend to the other reforming types, i.e., the critical O/C values grow with rising pressure. At the pressure of 1 bar, all tree O/C lines approach the critical O/C values grow with rising pressure. At the pressure of 1 bar, all tree O/C lines approach a constant value of O/C = 1 when the process temperature approaches 1000 °C. a constant value of O/C = 1 when the process temperature approaches 1000 ◦ C. 2.3. Liquified Petroleum Gas Reformates A similar collective graph of equilibrium conditions for three reformates of liquefied petroleum gas (LPG) is presented in Figure 5b. A considerable similarity in the shape of the O/C threshold lines to those of NG reformates (Figure 5a) can be noticed for the CPOx and dry reforming types.

Energies 2018, 11, 1177

7 of 12

2.3. Liquified Petroleum Gas Reformates A similar collective graph of equilibrium conditions for three reformates of liquefied petroleum gas (LPG) is presented in Figure 5b. A considerable similarity in the shape of the O/C threshold lines to those of NG reformates (Figure 5a) can be noticed for the CPOx and dry reforming types. However, the critical lines for the wet reforming of LPG have different shapes than those of NG for the deposition regions of graphite and multi-walled carbon nanotubes, i.e., for temperatures lower than 577 ◦ C. They start at 200 ◦ C from about O/C = 1.9 then decrease to a minimum at roughly 400 ◦ C, and start growing with increasing temperature in the area of MWCNT deposition and the beginning of the SWCNT. In the region of SWCNT deposition from LPG wet reformates, the O/C curves show maxima dependent on the process pressure with a close similarity to their equivalents for NG wet reformates. 3. Discussion In the modeling of carbon deposition, the same computing methodology as that published in former papers [12–14] is applied to predict conditions of the thermodynamic equilibrium of reformates. Unchanged chemical potential data from the literature are also applied for considered forms of solid carbon allotropes. The C–H–O ternary diagrams for two previous levels of process pressure, 1 and 10 bar [12], are now extended to 3 and 30 bar. It follows from the comparative analysis of those diagrams that the flow of the threshold lines, between the deposition and deposition-free regions, continues the trends found earlier [12], both for the process temperature and pressure. In particular, the ranges of deposition temperature for three carbon allotropes; graphite, and multi-walled and single-walled nanotubes, were independent of pressure. The maximum differences between the threshold lines are found at higher temperatures than formerly in [12], i.e., about 800 ◦ C and 1000 ◦ C for the pressure of 3 and 30 bar, respectively. The ternary diagrams are now also equipped with chord lines indicating the compositions of possible fuel-oxidant mixtures. The crossing points of the chord and threshold lines correspond to relevant points of the O/C critical lines for a given fuel-oxidant pair. The points are used in the interpretation and explanation for the critical lines in the function of temperature and pressure in our former results [13,14]. Six types of reformed mixtures of fuel-oxidant are considered using two fuels (NG or LPG) and three oxidants: air in CPOx reforming [13] and water (wet reforming) or carbon dioxide (dry reforming) in [14]. The non-monotonic influence of both process temperature and pressure on the computed critical O/C lines is easily explained by referencing to the distribution of the relevant threshold and chord lines in the ternary diagram. Detail explication of the specific O/C functions of temperature and pressure are presented in Sections 2.2 and 2.3. It is explained in reference to appropriate—for the analyzed fuel-oxidant mixtures—threshold and chord lines in ternary diagrams. 4. Materials and Methods 4.1. Basic Reactions and Energy Effects The reforming reactions depend on the type of fuel and on the reforming reactant. They are usually carried out in the gas phase to obtain hydrogen, H2 , and carbon monoxide, CO. This study concerns lower, i.e., C1 to C4 , saturated hydrocarbons, which are the most popular and still low-cost fossil fuels. Those fuels can be reformed in reactions with an oxygen-carrying reactant, such as water steam (H2 O), carbon dioxide (CO2 ), or oxygen (O2 ). Basic reactions, (a) to (c), of the hydrocarbon fuels with the reforming reactants are presented in the summary forms in Table 1. The number of carbon atoms in the fuel molecules is equal to “n”. The reaction equations are followed by associated standard values of the reaction enthalpy, ∆Hi0 , and chemical potential, ∆µ0i , at the temperature of 298 K. The data were obtained from the recognized database [15]. The wet and dry reforming reactions, respectively (a) and (b), are highly endothermal, whereas the partial oxidation reaction (c) is exothermal.

Energies 2018, 11, 1177

8 of 12

Table 1. Basic reactions and their standard thermodynamic data. Energies 2018, 11, x

∆H0i (298) kJ mol−1

Equation

kJ mol−1



8 of 11

Reaction (i)



Cn H2n+2 + nHdeposits (2ncomputed + 1)H2 + 148.4 n of the reacting n (a) =56.4 =61.6 + 79.6 2 O(g) = nCO carbonaceous can+be from compositions fuel–oxidant mixtures at ∼ ∼ Cn H2n+2 + nCO2 = 2nCO + (n + 1) H2 (b) =56.4 + 189.6 n =61.6 + 108.2 n theCcoking threshold. This was already reviewed in our previous study [12] with the conclusion that ∼ ∼ (c) =56.4 − 93.4 n =61.6 − 149.0 n n H2n+2 + n/2O2 = nCO + (n + 1) H2 the wide scatter chemical potential estimates resulted most probably CH4 =ofCthe + 2H 791.2 721.8 from deriving (d)them for 2 (g) different mixtures crystallographic forms of544.2 the carbonic deposits 551.2 in relevant experiments. The 2CO = Cof + CO (e) 2 (g)the CO + H = C + H O 585.4 579.8 (f) pentagon shown2by the 2 lines in Figure 6 shows the range of estimates of the standard chemical (g) blue + CO2 = C(g) +carbon 2H2 O deposits, as determined 626.5 608.5 (g) potential2H of2 filamentous from the coking threshold compositions in m C(g) = m Cg(g) −716.7 m −671.2 m (h) 0 several reforming reactions. The estimates are presented as their difference, ΔμC ( s ) , from the 0 standard chemical potential of graphite, μCg ( s ) , which is assumed to be 0 for any temperature, Substrates andEnergies products C(g) , 2018, 11,of x the basic reactions can further react to form also vapors of carbon, 8 of 11 according the convention applied in physical chemistry. as shown intoreactions (d) to (g). The thermodynamic effects of the deposition of m atoms of carbon carbonaceous deposits can be computed from compositions of the reacting fuel–oxidant mixtures at vapors to form solid graphite, Cg(s) ,This arewas also included in Table 1. the coking threshold. already reviewed in our previous study [12] with the conclusion that Table 1. Basic reactions and their standard thermodynamic data. Reliable thermodynamic data forchemical carbonpotential nanotubes at the reference temperature of 298them K cannot be the wide scatter of the estimates resulted most probably from deriving for different mixtures of the crystallographic forms ofof thethe carbonic deposits in relevantof experiments. The 0 0 −1 −1 found in the subject literature. However, several values chemical potential the carbonaceous ΔH ( 298 ) kJ mol Δμ ( 298 ) kJ mol Equation Reaction (i) pentagon shown by the blue linesi in Figure 6 shows the range ofi estimates of the standard chemical deposits can be computed from compositions of the reacting fuel–oxidant mixtures at the coking potential of filamentous carbon deposits, as determined from the coking threshold compositions in CnH2n+2 + nH2O(g) = nCO + (2n + 1)H2 (a) ≅56.4 + 148.4 n ≅61.6 + 79.6 n 0 threshold. This was2nCO already reviewed in The ourestimates previous [12] with conclusion that the reforming arestudy their the difference, the wide CnH2n+2 + nCO2 =several + (n + 1) reactions. H2 ≅56.4 + 189.6 npresented as ≅61.6 + 108.2 n ΔμC ( s ) , from (b) 0 scatter of the chemical potential probably from deriving them for different , which is assumed to be 0 for any temperature, standard chemical estimates potential of resulted graphite, μmost Cg ( s ) CnH2n+2 + n/2O2 = nCO + (n + 1) H2 (c) ≅56.4 − 93.4 n ≅61.6 − 149.0 n according to the convention applied in physical chemistry.in relevant experiments. The pentagon mixtures of the crystallographic forms of the carbonic deposits CH4 = C(g) + 2H2 (d) 791.2 721.8 shown by the2CO blue= C lines in Figure 6Table shows the range of estimates of the standard chemical potential 1. Basic reactions and their standard thermodynamic data. (e) (g) + CO2 544.2 551.2 of filamentous carbon deposits, as determined from the coking threshold compositions in several 0 0 CO + H2 = C(g) + H2O Equation 585.4 579.8 ΔH i ( 298 ) kJ mol−1 Δμi 0( 298 ) kJ mol−1 Reaction (i)(f) reforming2H reactions. The estimates are presented as their difference, ∆µC(s) , from the standard (g) chemical 2 + CO2 = C + 2H 2O 626.5 608.5 Cn(g) H2n+2 + nH 2O(g) = nCO + (2n + 1)H2 (a) ≅56.4 + 148.4 n ≅61.6 + 79.6 n 0 potential of graphite, is assumed be≅56.4 0 for any temperature, to the (b) convention CnH +, which nCO2 = 2nCO + (n + 1) H2 to n ≅61.6 + according 108.2 m C(g) = µ (h) m C(2n+2 g(g) −716.7 m+ 189.6 −671.2 mn Cg s) CnH2n+2 + n/2O2 = nCO + (n + 1) H2 (c) ≅56.4 − 93.4 n ≅61.6 − 149.0 n applied in physical chemistry.

791.2 544.2 585.4 626.5 −716.7 m

CH4 = C(g) + 2H2 2CO = C(g) + CO2 10000

CO + H2 = C(g) + H2O 2H2 + CO2 = C(g) + 2H2O m C(g) = m Cg(g)

721.8 551.2 579.8 Data [16] Extrapolation 608.5 Linearized [17] −671.2 m

(d) (e) (f) (g) (h)

5000

Δμ0C(s)

10000

[J/mol]

Data [16] Extrapolation

GRAPHITE Linearized [17]

0 5000

Δμ0C(s)

MWCNT

[J/mol] GRAPHITE

0

-5000

MWCNT

SWCNT

-5000

-10000 400

500 400

SWCNT

600

-10000 500

T [oC]

700 600

700

T [oC]

Figure6.6.Chemical ChemicalFigure potential the formation of carbon deposits relative to graphite. represents represents Figure potential ofofthe formation carbon relative to graphite. the 6. Chemical potential of theof formation ofdeposits carbon deposits relative to graphite. represents 0 0 0 Δ μ for carbon nanotubes (CNTs) [12] estimated from the compositions the range of carbon [12] nanotubes (CNTs) from [12] estimated from the compositions in the in the the of ΔμC ( s ) for range of ∆µC(s) forCcarbon (CNTs) estimated the compositions in the literature at ( s ) rangenanotubes literature at the coking threshold. the coking threshold. literature at the coking threshold. Chemical potential data for the carbon nanotubes, CNT, should be obtained for their purified forms, either MWCNT, or single-walled in bundles, Such data fortheir elevated Chemicalpotential potential datamulti-walled, forthe thecarbon carbon nanotubes, CNT, shouldSWCNT. beobtained obtained for theirpurified purified Chemical data for nanotubes, CNT, should be for temperatures MWCNT, is available from electrochemical measurements published by Gozzi et al. [16,17]. For forms, either multi-walled, or single-walled in bundles, SWCNT. Such data for elevated forms, either multi-walled, MWCNT, or single-walled in bundles, SWCNT. Such data for elevated measurements with the multi-walled CNT and in temperature ranging from 820 K to 920 K, the temperatures isis available electrochemical measurements Gozzi et al. For temperatures from electrochemical measurementspublished publishedbyby Gozzi et [16,17]. al. [16,17]. authors [16]from obtained the following relationship: measurements with the the multi-walled CNT and in temperature ranging from 820 K to 920 K, the For measurements with multi-walled and in temperature ranging from 820 K to 920 K, 0 CNT ΔμMWCNT (T ) = 8250 − 11.72T (J/mol) for T (K). (1) authors [16] obtained the following relationship: the authors [16] obtained the following relationship: This relationship is graphically shown in Figure 6 by the dark blue solid line, which is also 0 0 μC ( s ) = 0T, to=represent its chemical potential equal to that of graphite, which occurs extrapolated0 Δ toμ Δ 8250 − 11.72 T (J/mol) for T (K). MWCNT ∆µ MWCNT ( T ) = 8250 − 11.72T (J/mol) for T (K). at about 431 °C. Results of similar measurements carried out for bundled single-walled carbon

( )

(1) (1)

This relationship is graphically shown in Figure 6 by the dark blue solid line, which is also extrapolated to ΔμC0 ( s ) = 0 , to represent its chemical potential equal to that of graphite, which occurs at about 431 °C. Results of similar measurements carried out for bundled single-walled carbon

Energies 2018, 11, 1177

9 of 12

This relationship is graphically shown in Figure 6 by the dark blue solid line, which is also extrapolated to ∆µ0C(s) = 0, to represent its chemical potential equal to that of graphite, which occurs at about 431 ◦ C. Results of similar measurements carried out for bundled single-walled carbon nanotubes [17], in the temperature range from 750 K to 1015 K, were presented in the form of the 9th degree polynomial: ∆µ0SWCNT ( T ) =

9

∑ an T n .

(2)

n =0

A piecewise approximation of that function is presented in Figure 6 by the red line. The common point of Equations (1) and (2) occurs at a temperature of about 577 ◦ C. Such temperature functions of the chemical potentials of the MWCNT and SWCNT were used in calculations of the equilibrium concentrations of the gas and solid phases with the HSC software [18]. The calculations were based on the principles of the thermodynamic equilibrium, which are briefly described in the following section. 4.2. Thermodynamic Equilibrium The basic postulate of chemical statics, for the thermodynamic equilibrium (*) to occur [19], is the reaction system reaches then a minimum of its total Gibbs energy, G (J):  n o dG T ∗ , P∗ , n∗j =0

(3)

The chemical potential of a species “j”, µj (J/mol), is the partial molar Gibbs energy of the system for constant temperature and pressure: "

∂G ∂n j

#

= µj

(4)

T,P

In a homogeneous mixture of many species, the chemical potential can be expressed as the sum of its standard value (for pure species “j”), µ0j , and a contribution of the ratio of the species activity in the mixture, aj , to its value in the standard state, a0j : µ j = µ0j + RT ln

aj a0j

(5)

When dealing with gas phase mixtures, the species activity is usually expressed as a product of the species partial pressure, Pj , and the fugacity coefficient, ϕj : a j = ϕ j Pj

(6)

The fugacity coefficient practically equals 1.0 [19] when the species partial pressure, Pj , is significantly lower than its critical pressure and also the process temperature is much higher than the species critical temperature. In the state of thermodynamic equilibrium, denoted by “*”, the species chemical potential should be constant in all phases containing the species: solid (s), liquid (l), and gas (g): µ∗j(s) = µ∗j(l ) = µ∗j( g)

(7)

Consequently, in the case of chemical reactions between species in a given phase, the postulate of Equation (3) leads to Equation (8), which is valid for constant temperature and pressure:

∑ νj,i ·µ∗j = 0 j

(8)

Energies 2018, 11, 1177

10 of 12

The stoichiometric coefficients, νj,i , of the j-th reactant, that participates in the i-th reaction, are negative for substrates and positive for products. The standard enthalpy and chemical potential of a chemical reaction “i” are calculated in a similar way, accounting for contributions from substrates and products: ∆Hi0 ( T ) =

∑ νj,i · Hj0 (T )

(9)

∑ νj,i ·µ0j (T )

(10)

j

∆µ0i ( T ) =

j

Their values for relevant reactions are quoted in Table 1. The conditional minimum of the total Gibbs energy, expressed in Equation (3), can be computed in several ways [20]. This leads to the determination of the composition of each phase at the thermodynamic equilibrium for the prescribed initial mole number of substrates and the final phase species, temperature, and pressure. Those computations can be carried out with the help of a commercial package, such as HSC [18], which was used in this study. 4.3. Modeling Procedure and Scope The modeling of equilibrium conditions of C–H–O mixtures was specifically directed to detect the beginning of the formation of solid carbon. Two types of such computations were applied in our former studies [12–14] and they were based on the principles presented in the previous section. Details of the computing procedure were also described there and the interested reader is referred to those open access papers. The relationship (Equation (1)) for MWCNT was directly applied to widen the HSC databank. However, the ∆µ0SWCNT function of temperature, Equation (2), was piecewise linearly approximated for ∆µ0SWCNT < 0 and then introduced to the HSC software database. The approximation lines are presented in Figure 6, with the extrapolation of Equation (1) to the level of ∆µ0MWCNT = 0 shown by the broken line. The border temperatures of 431 and 577 ◦ C were found independent of the process pressure in the equilibrium calculations. They correspond to the crossing points of the pairs of lines in Figure 6: graphite–MWCNT and MWCNT–SWCNT. The depositing type of carbon is characterized in the thermodynamic equilibrium by the lowest value of its standard chemical potential. Our former paper [12] reported details on the critical C, O, and H atomic fractions around which a step change from none to a detectable amount of solid carbon occurs. This allowed for determining critical lines, which divide in ternary C–H–O diagrams the carbon deposition area from the deposition-free area. However, only two levels of the total pressure, i.e., 1 bar and 10 bar, of the equilibrated mixture of carbon-, hydrogen-, and oxygen-containing compounds were applied in [12]. This present study is aimed at a comparison and discussion of the critical deposition lines presented for the CPOx reforming in [13] and for the wet and dry reforming in [14]. Those two papers discussed the critical O/C atomic ratios against temperature for two types of hydrocarbon fuels, NG and LPG, and three types of oxygen-containing reforming oxidants; O2 from air, H2 O, and CO2 . Four levels of constant pressure: 1, 3, 10, and 30 bar, were used in [13,14], while only two of them, i.e., 1 and 10 bar, were applied in [12] Therefore, the former computational results are supplemented in this study by new data for C–H–O ternary diagrams for the pressure of 3 and 30 bar, as presented in Section 2. Author Contributions: Z.J. and P.P.-O. designed the study; Z.J. designed and performed the computations, Z.J. analyzed a majority of the data; P.P.-O. contributed to analyses of the data and wrote the main part of the Introduction; and Z.J. wrote the rest of the paper. Funding: This research was funded by [European Union’s Seventh Framework Programme] grant number [FP7/2007–2013], [Fuel Cells and Hydrogen Joint Undertaking (FCH JU)] grant number [621213] and [Polish research funds] grant number [3126/7.PR/2014/2].

Energies 2018, 11, 1177

11 of 12

Acknowledgments: The research program leading to these results received funding from the European Union’s Seventh Framework Programme (FP7/2007–2013) for the Fuel Cells and Hydrogen Joint Undertaking (FCH JU) under grant agreement no. 621213 with the STAGE-SOFC acronym. Information contained in the paper reflects only the view of the authors. The FCH JU and the Union are not liable for any use that may be made of the information contained therein. The work was also financed from the Polish research funds awarded for project no. 3126/7.PR/2014/2 of international cooperation within STAGE-SOFC in the years 2014–2017. The authors received funds from FCH JU for covering the costs to publish in open access. Conflicts of Interest: The authors declare no conflict of interest.

References 1.

2.

3.

4. 5. 6. 7.

8.

9. 10. 11. 12. 13. 14. 15. 16. 17. 18.

Xie, H.; Yu, Q.; Zhang, Y.; Zhang, J.; Liu, J.; Qin, Q. New process for hydrogen production from raw coke oven gas via sorption-enhanced steam reforming, Thermodynamic analysis. Int. J. Hydrogen Energy 2017, 42, 2914–2923. [CrossRef] Purima, P.; Jayanti, S. A high-efficiency auto-thermal system for on-board hydrogen production for low temperature PEM fuel cells using dual reforming of ethanol. Int. J. Hydrogen Energy 2016, 41, 13800–13810. [CrossRef] Sumrunronnasak, S.; Tantayanon, S.; Kiatgamolchai, S.; Sukonket, T. Improved hydrogen production from dry reforming reaction using a catalytic packed-bed membrane reactor with Ni-based catalyst and dense PdAgCu alloy memberane. Int. J. Hydrogen Energy 2016, 41, 2621–2630. [CrossRef] Araiza, D.G.; Gomez-Cortes, A.; Diaz, G. Partial oxidation of methanol over copper supported on nanoshaped ceria for hydrogen production. Catal. Today 2017, 282, 185–194. [CrossRef] Energy USD-EEAR. Hydrogen, Fuel Cells & Infrastructure Technologies Program. 2005. Available online: http://www1.eere.energy.gov/hydrogenandfuelcells/mypp/ (accessed on 26 March 2018). Adiya, Z.I.S.G.; Dupont, V.; Mahmud, T. Steam reforming of shale gas in a packed bed reactor with and without chemical looping using nickel based oxygen carrier. Int. J. Hydrogen Energy 2018, 43, 6904–6917. [CrossRef] Tuna, C.E.; Silveira, J.L.; da Silva, M.E.; Boloy, R.M.; Braga, L.B.; Perez, N.P. Biogas steam reformer for hydrogen production: Evaluation of the reformer prototype and catalysts. Int. J. Hydrogen Energy 2018, 43, 2108–2120. [CrossRef] Alves, H.J.; Junior, C.B.; Niklevicz, R.R.; Frigo, E.P.; Frigo, M.S.; Cimbra-Araujo, C.H. Overview of hydrogen production technologies from biogas and the applications in fuel cells. Int. J. Hydrogen Energy 2013, 38, 5212–5225. [CrossRef] Loffler, D.G.; Taylor, K.; Mason, D. A light hydrocarbon fuel processor producing high purity hydrogen. J. Power Sources 2003, 117, 84–91. [CrossRef] Rostrup-Nielson, J.R. Catalytic Steam Reforming. Catalysis—Science and Technology; Anderson, J.R., Boudart, M., Eds.; Springer: Berlin/Heidelberg, Germany, 1984; Volume 5, ISBN 978-3-642-93247-2. Ginsburg, J.M.; Pina, J.; Solh, T.E.I.; Lasa, H.I. Coke formation over a nickel catalyst under methane dry reforming conditions: Thermodynamic and kinetic models. Ind. Eng. Chem. Res. 2005, 44, 4846–4854. [CrossRef] Jaworski, Z.; Zakrzewska, B.; Pianko-Oprych, P. On thermodynamic equilibrium of carbon deposition from gaseous C-H-O mixtures: Updating for nanotubes. Rev. Chem. Eng. 2017, 33, 217–235. [CrossRef] Jaworski, Z.; Pianko-Oprych, P. On the deposition equilibrium of carbon nanotubes or graphite in the reforming processes of lower hydrocarbon fuels. Entropy 2017, 19, 650. [CrossRef] Jaworski, Z.; Pianko-Oprych, P. On nanotube carbon deposition at equilibrium in catalytic partial oxidation of selected hydrocarbon fuels. Int. J. Hydrogen Energy 2017, 42, 16920–16931. [CrossRef] Yaws, C.L. Yaws’ Critical Property Data for Chemical Engineers and Chemists. 2012. Available online: http: //app.knovel.com/hotlink/toc/id:kpYCPDCECD/yaws-critical-property (accessed on 4 January 2017). Gozzi, D.; Iervolino, M.; Latini, A. The thermodynamics of the transformation of graphite to multiwalled carbon nanotubes. J. Am. Chem. Soc. 2007, 129, 10269–10275. [CrossRef] [PubMed] Gozzi, D.; Latini, A.; Lazzarini, L. Experimental thermodynamics of high temperature transformations in single-walled carbon nanotube bundles. J. Am. Chem. Soc. 2009, 131, 12474–12482. [CrossRef] [PubMed] Outotec’s HSC Chemistry v.8.0 Software; Outotec (Finland) Oy, Research Center Pori: Pori, Finland, 2014.

Energies 2018, 11, 1177

19. 20.

12 of 12

Smith, J.M.; Van Ness, H.C.; Abbott, M.M. Introduction to Chemical Engineering Thermodynamics, 7th ed.; McGraw Hill Higher Education: New York, NY, USA, 2004; ISBN 0-07-310445-0. Castillo, J.; Grossmann, I.E. Computation of phase and chemical equilibria. Comput. Chem. Eng. 1981, 5, 99–108. [CrossRef] © 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).