Next Article in Journal
Influence of Oil-Soluble Catalyst on Composition and Structure of Heavy Oil from Samara Region Field
Previous Article in Journal
Development of Novel Monolithic Catalyst for BTEX Catalytic Oxidation Using 3D Printing Technology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Engineering Calculations for Catalytic Hydrolysis of CF4

by
Robert Barat
Department of Chemical and Materials Engineering, New Jersey Institute of Technology, Newark, NJ 07102, USA
ChemEngineering 2025, 9(1), 10; https://doi.org/10.3390/chemengineering9010010
Submission received: 20 November 2024 / Revised: 27 December 2024 / Accepted: 16 January 2025 / Published: 20 January 2025

Abstract

:
Tetrafluoromethane (CF4) is the simplest perfluorocarbon, a class of compounds with very high greenhouse gas potential. Catalytic hydrolysis offers an opportunity to convert these compounds to manageable CO2 and HF. Recently published data showed the effectiveness of Ga-doping to overcome the fluorine poisoning of various Al2O3 catalysts at relatively modest temperatures. This prior work offered a partial catalytic mechanism together with kinetic and conversion data. The current paper completes the catalytic mechanism, and then analyzes it using the Langmuir–Hinshelwood algorithm for both the initial CF4 conversion, and the catalyst site regeneration. The resulting derived rate expression, together with a catalyst activity coefficient expression, are then used in flow reactor configurations to simulate both relatively short exposure time runs with little loss of activity, as well as longer runs with severe activity loss. The reasonable agreement with the published laboratory data suggests that these expressions can be used for a larger-scale practical reactor design.

1. Introduction

Perfluorocarbons (PFCs) contain only carbon and fluorine atoms. These compounds are non-flammable, chemically inert, and stable. They are even used in medical theranostics [1]. Despite their unique properties, PFCs are under growing scrutiny as “forever chemicals”—species that, due to their chemical stability, tend to linger and accumulate in the environment and the body. The US Environmental Protection Agency recently issued a drinking water national standard [2] for polyfluoroalkyl substances.
A potentially more concerning environmental consequence is the high greenhouse gas potential of PFCs, especially CF4 [3]. The symmetry and high C-F bond strength of CF4 preclude easy decomposition. The key is the activation of the strong C-F bond [4]. Catalytic hydrolysis, especially using alumina-based catalysts, offers a viable pathway to the conversion of CF4 to stable products that can be captured easily. The overall desired hydrolysis reaction is C F 4 + 2 H 2 O C O 2 + 4 H F , H r o = −70.06 kJ/mole.
Thermal decomposition of CF4 has been demonstrated with many different catalysts [5,6]. Transition metal-doped alumina is widely considered to be a useful catalyst for organic reactions [7] and the degradation and adsorption of organic dyes [8]. The often-cited catalyst for the hydrolysis of CF4 is a Lewis acid-based metal oxide, for example, alumina [9]. The active site for CF4 bond activation has been identified as the Al site [10]. The efficiency and resilience of θ- A l 2 O 3 over other alumina forms has been established [11].
However, even at temperatures as high as 973 K, these catalysts suffer rapid F-poisoning [3]. Alumina with a metal dopant offers poisoning resistance [3,12]. Doping θ- A l 2 O 3 with Ga resulted in a strong, sustained activity for CF4 hydrolysis at a reduced temperature (773 K) compared to the undoped case [3]. Switching to ZrO2 doped with -HSO4 is another option [13]. Similar results have been obtained with H2SO4-treated ZnAl2O4 [4].
In this paper, the mechanism and kinetic work in [3] is expanded into an engineering application. Additional details for the initial CF4 conversion steps within the catalytic cycle are proposed to complete the catalytic cycle. A rate expression for CF4 hydration based on the whole cycle is developed with the Langmuir–Hinshelwood algorithm. The rate expression includes the catalyst activity, which is modeled quantitatively. This kinetic model is tested against CF4 conversion data [3] for θ- A l 2 O 3 and γ- A l 2 O 3 catalysts that poison relatively easily. The rate expression is then tested against CF4 conversion data [3] for a Ga/θ- A l 2 O 3 catalyst. The purpose of these derivations and calibrations is to facilitate a future engineering scaleup design.

2. Development of Kinetic Rate Expressions

This section shows the development of a working kinetic rate expression that can be used for a future CF4 catalytic hydrolysis reactor design. Including the activity offers the chance to guide important reactor planning such as parallel trains and catalyst regeneration. Subsections cover an expanded mechanism cycle, the derivation of a working rate expression including catalyst activity, a facilitated unsteady packed bed reactor description, and finally preparation for the activity calibration.

2.1. Mechanism

Figure 5 in Reference [3] proposes a catalytic cycle for the CF4 hydrolysis on Al2O3. The cycle illustrates the relevant catalytic sites, and can be divided into two portions:
Portion   I :   C F 4 + r e g e n e r a t e d   s i t e C O 2 + 2 H F + p o i s o n e d ( f l u o r i n a t e d ) s i t e
Portion   II :   p o i s o n e d   s i t e + 2 H 2 O 2 H F + r e g e n e r a t e d   s i t e
The analysis below builds on the partial catalytic cycle shown in [3]. While much detail is provided in [3] for Portion II, no details are shown for Portion I. Below is a proposed detailed mechanism for each portion, including likely transition states. For clarity, the mechanism is presented in a linear (not cyclic) way. Stable catalyst sites are S*, while transition states are TS*. The mechanism is shown with a Ga replacing an Al but is valid for the non-doped case. Approaching reactant and departing product molecules are also shown. Bond breaking and formation are shown as dashed lines. Portion I reactions are graphically shown in Figure 1 and Figure 2. Portion II reactions are shown in Figure 3 and Figure 4.

2.1.1. Portion I Mechanism

The elementary reactions (steps) in Portion 1 are the following:
1 .             C F 4 + S 1 = H F + S 2
2 .             S 2 = H F + S 3
3 .             S 3 = C O 2 + S 4
The CF4 conversion cycle begins with the removal of the first F by the reaction of a H from the OH group on either Ga or Al. The adsorption energy is greater for F than for OH on Al, while the bond length is shorter. These data indicate the stronger bond of F to Al, thus resulting in the poisoning [3]. The case is less severe for a Ga site than an Al site.
Zhang et al. [3] show conclusively that the poisoned S4 can be more easily regenerated due to the reduced Lewis acidity of the Ga dopant. In the absence of Ga-doping, the rate of removal of the F atoms from S4 in Portion 2 is very much diminished due to the stronger adsorption of F atoms to the Al site (higher Lewis acidity). The Ga-doped site has a weaker adsorption to F atoms (lower Lewis acidity), thus making the F atom removal by subsequent hydrolysis (Step 6 below) more favorable.

2.1.2. Portion II Mechanism

The elementary reactions (steps) in Portion II are the following:
  4 .         S 4 + H 2 O = S 5 + H F
5 .             S 5 = H F + S 6
6 .             S 6 + H 2 O = S 1
The Ga-F bond is weaker than the Al-F bond [3]. This facilitates the start of the poisoned site regeneration in Portion II.

2.2. Langmuir–Hinshelwood Analysis

The classical Langmuir–Hinshelwood (L-H) analysis of the above mechanism begins with the assumption of a slow or rate-determining-step (RDS). The other steps are assumed to be quasi-equilibrated (pseudo-equilibrium). The validity of the overall kinetic rate expression, derived from the subsequent algebraic analysis, is judged by comparison to the observed rates or conversions obtained under specific experimental conditions. If the derived rate expression is not consistent with observed data, then another step should be selected as the RDS, and the derivation is repeated.
The L-H analysis here assumes that Step 1 is the RDS. The six steps can now be presented as follows:
1 .       C F 4 + S 1 H F + S 2
2 . S 2 H F + S 3
3 . S 3 C O 2 + S 4
4 .       S 4 + H 2 O S 5 + H F
5 . S 5 H F + S 6
6 .   S 6 + H 2 O S 1
The overall rate expression is:
r r 1 = k 1 y C F 4 C S 1
where ri = rate of Step i, ki = forward rate constant for Step i, yj = species j mole fraction, and CSi = surface concentration of catalyst site i. To estimate CS1, total site balance is used:
C t = C S 1 + C S 2 + C S 3 + C S 4 + C S 5 + C S 6
where Ct = total surface concentration of all catalytic sites, which is taken as constant. The remaining mechanism steps are now used to estimate the remaining site concentrations.
For CS2, as an example for other site concentrations, consider the net rate of Step 2:
r 2 = k 2 C S 2 k 2 y H F C S 3
where k−i = reverse rate constant for Step i. Rearrangement of Step 3 yields:
r 2 / k 2 = C S 2 y H F C S 3 / k 2 / k 2
Step 2 is elementary, so its rate constant ratio is equal to its equilibrium constant:
k 2 / k 2 = K 2
where Ki = equilibrium constant for Step i. Step 2 is assumed to be in a quasi-equilibrium condition, which is represented by:
0 r 2 / k 2 = C S 2 y H F C S 3 / K 2
The rate r2 is finite non-zero, while k2 is assumed to be large (fast). CS2 is estimated as:
C S 2 y H F C S 3 / K 2
The pattern used in Equations (3)–(7) is repeated for estimates of CS3 through CS6:
C S 3 y C O 2 C S 4 / K 3             C S 4 y H F C S 5 / K 4 y H 2 O
C S 5 y H F C S 6 / K 5             C S 6 C S 1 / K 6 y H 2 O
Successive substitutions result in all expressions for CSi in terms of CS1. All of these are, in turn, substituted into the Equation (2) site balance. The result in an estimate for CS1:
C S 1 = C t 1 + y H F 3 y C O 2 K 2 K 3 K 4 K 5 K 6 y H 2 O 2 + y H F 2 y C O 2 K 3 K 4 K 5 K 6 y H 2 O 2 + y H F 2 K 4 K 5 K 6 y H 2 O 2 + y H F K 5 K 6 y H 2 O + 1 K 6 y H 2 O
Combining Equations (1) and (10) yields an expression for the overall rate:
r = k 1 C t y C F 4 1 + y H F 3 y C O 2 K 2 K 3 K 4 K 5 K 6 y H 2 O 2 + y H F 2 y C O 2 K 3 K 4 K 5 K 6 y H 2 O 2 + y H F 2 K 4 K 5 K 6 y H 2 O 2 + y H F K 5 K 6 y H 2 O + 1 K 6 y H 2 O    

2.3. Reconciliation of Derived L-H Rate Expression with Experimental Conditions

Figure 1c in Reference [3] presents an Arrhenius plot of experimental rate constants k obtained for three catalysts: Ga-doped θ- A l 2 O 3 , θ- A l 2 O 3 , and γ- A l 2 O 3 . Direct communication with the corresponding author of Reference [3] confirmed that the assumed rate expression used for their kinetic analysis was first order in the mole fraction of CF4.
To simplify the derived rate expression (Equation (11)) to first order in CF4, consider the reactor experimental conditions reported in Figure 1 of Reference [3]: feed gas (vapor) was 0.96 mole% CF4, 19.68% H2O, and a balance of Ar. Under these conditions, it is easily argued that y C F 4 y H 2 O , y C O 2 y H 2 O , y H F y H 2 O , and y H 2 O constant. Under these conditions, the non-unity terms in the Equations (10) and (11) denominators fall out. The derived rate expression can be simplified to:
r k 1 C S 1 y C F 4 k 1 C t y C F 4 k y C F 4
where k k 1 C t .

2.4. Impact of F-Poisoning on Rate Expression

Equation (12) suggests that the impacts of F-poisoning and Ga-doping might be expressed through the total site concentration Ct. Figure 1b in Reference [3] shows that the Ga-doped catalyst activity is effectively exposure-time independent. With Ga-doping and excess H2O, the first-order dependence also suggests that Steps 2–6 are favorable; i.e., K2 through K6 are “large”. This all results in C t C S 1 . However, the authors show that the regeneration of fluorinated site S4 effectively stalls if Ga-doping is not used. Therefore, site concentration CS4 builds up since Step 4 becomes unfavorable, i.e., K4 is “small”. All terms containing K4 in the denominator of Equation (10) are now “large”, thus causing C S 1 C t , with a subsequent large decline in CF4 conversion.
The rapid termination of CF4 decomposition using the undoped catalysts suggests a time (exposure)-dependent expression for Ct. This time dependence for Ct can be indirectly expressed through the catalytic activity a:
a r o b s / r o
where ro = catalytic rate without poisoning (with Ga-doping or a fresh catalyst) and robs = observed rate. Combining Equations (12) and (13) gives the working rate expression:
r o b s = a k 1 C t y C F 4 a k y C F 4
For a given catalyst, k is taken as a constant corresponding to no poisoning. The activity accounts for the poisoning ultimately due to exposure to CF4.
There are several expressions for activity a, depending on the situation [14]. A potential candidate here is a first-order dependence:
d a / d t = k p a y p
where kp = rate constant for catalyst poisoning, and yp = mole fraction of the poisoning agent. Since CF4 is the source of the poison (i.e., F atoms), we can take y p y C F 4 .

2.5. Catalytic Reactor Modeling

The catalytic reactor used in [3] is taken to be a packed bed reactor (PBR), with observed CF4 conversions exceeding the 10% cutoff for a differential reactor, as confirmed by the corresponding author. The steady-state PBR model, with CF4 given as A, is:
d F A / d W = r A
where FA = molar flow rate of A within the reactor, W = catalyst mass, and r A = molar reaction rate based on catalyst mass. Applying the same composition argument used for Equation (12) above, and including Equation (14), Equation (16) can be written as:
d y A / d W = k y A / F T
where the total molar flow rate F T constant. Equation (17) is good for either very little or no poisoning (i.e., a 1 ). For poisoning, Equation (17) must include the activity a. But, poisoning here is a time-dependent phenomenon. The poisoned PBR model is:
y A / W = k a y A / F T a / t = k p a y A
The Equation (18) partial differential equation (PDE) problem can be handled more easily by approximating the PBR as several unsteady, equal-sized, continuous stirred tank reactors (CSTRs)-in-series [14], as seen in Figure 5. For this problem, five CSTRs are used. The total catalyst mass W is divided equally (Wi) between the five. This approach changes the PDE problem to a simultaneous ordinary differential equations (ODEs) problem.
For this application, the unsteady ith CSTR species balance is:
F A , i 1 F A , i + r A , i W i = d N A , i / d t
where NA,i = moles of A within the catalyst void space, and is approximated by:
N A , i = y A , i N T , i = y A , i C T , i V g , i y A , i P / R T ϕ V c , i
where NT,i = total moles of gas within reactor i, CT,I = total molar gas concentration, P = pressure, T = temperature, ϕ = catalyst “bed” porosity, and Vc,i = catalyst “bed” volume.
The PDE problem (Equation (18)) is now approximated as:
d y A , i / d t = R T / P ϕ V c , i F T y A , i 1 y A , i k a i y A , i W i d a i / d t = k p a i y A , i
The two ODEs for each CSTR (total of ten) are executed simultaneously using the Polymath® software (v. 6.10) package [15]. The initial conditions (at t = 0) for the activities ai are 1.0, and 0.0 for the mole fractions yA,i. For the first CSTR, yA,o = CF4 feed mole fraction. This model is comparable to a reactor startup. For no-poisoning cases (ai = 1 constant since kp = 0), the transient solution rapidly reaches a steady-state which will be shown to be comparable to the algebraic solution for five CSTRs-in-series. For poisoning cases k p 0 , results below show that the CSTR-in-series reaches a relatively quick solution then shows a slower degradation over time.

2.6. Preparation of Data for Simulations

2.6.1. Pre-Exponentials for Rate Constants

Figure 1c of Reference [3] offers an Arrhenius plot for the CF4 hydration using the three catalysts. Activation energies (Eact) are provided, but no pre-exponentials (Af). For each catalyst, the Af values were estimated at four temperatures covering the presented range and then averaged. The results are shown in Table 1. It is assumed that the k value data shown in Table 1 of Reference [3] were collected before any significant catalyst poisoning occurred. The rising values of the Af, Ea pairs is consistent with the increased difficulty of overcoming the impact of F-poisoning due to the increases in Lewis acidity.

2.6.2. Estimation of ϕ and Vc,i

Solving Equation (21) requires values for the catalyst “bed” porosity ϕ and the volume of catalyst “bed” Vc,i. Figure 1 of [3] reports a total gas feed of 41.46 cm3/min (STP) with a gas hourly space velocity (GHSV) of 1000/h. This gives an estimate of the total catalyst bed volume at 2.49 cm3. With five CSTRs-in-series, each bulk volume of the catalyst “bed” Vc,i is 0.498 cm3. With a total catalyst mass of 2 g, the bulk density of the catalyst bed is 0.803 g/cm3. The bed porosity ϕ values were then estimated using the catalyst mass densities ρcat. The ρcat value for the Ga/θ-Al2O3 was interpolated from the values for Al2O3 and Ga2O3. The catalyst mass in each of the five CSTRs-in-series Wi is 0.4 g. The values for ρcat and ϕ are presented in Table 2.

3. Results and Discussion

3.1. Short-Term Catalytic Runs

Figure 1a of Reference [3] presents CF4 conversion data for each of the catalysts at various temperatures. The exposure times are fairly short so that catalyst activities do not appear to be affected, though there is a fair amount of fluctuation. These runs were simulated using both a steady-state PBR model (Equation (17)) and five CSTRs-in-series. The results are shown in Figure 6. Uncertainty bars have been estimated and included.
Figure 6 shows that the predicted XA values from the five CSTRs-in-series model are consistently slightly lower than the PBR model, which is no surprise since an infinite number of CSTRs-in-series is required to exactly [14] model a plug flow reactor (PBR in this case). This difference is less than any model deviations from the observed data. It is noted that the transient CSTRs-in-series model showed the exact same numerical results as the steady-state version. Therefore, the use of the transient five CSTRs-in-series model with time-dependent activities (Equation (21)) over the PDE problem is justified.
Figure 6 shows that the PBR model fits to the experimental data trend are reasonable. The deviations might be attributed to experimental uncertainty (not quantified in Reference [3]) and non-PBR behavior in the reactor. Insufficient detail was provided to judge whether or not plug flow conditions existed in the experimental reactors used in Reference [3]. Since their conversions exceeded 10%, a differential reactor assumption was not justified. Nonetheless, the agreements are good enough to justify using Equation (21) to simulate the long-term transient runs in Figure 1b of Reference [3].

3.2. Long-Term Catalytic Runs

With the Arrhenius parameters for the CF4 hydration for each catalyst (Table 1) and the unsteady CSTRs-in-series with the activities reactor model (Equation (21)) in hand, the long exposure time experiments shown in Figure 1b of Reference [3] were modeled. The values used for the poisoning rate constant kp are shown in Table 3. They are adjusted to provide a reasonable fit to the data trends. The results are shown in Figure 7.
Figure 6 shows that the five CSTRs-in-series model (Equation (21)) slightly underpredicts the PBR model. A correction for this underprediction has been applied to the model curves shown in Figure 7. In addition, estimates were made for the uncertainty in the experimental data as suggested by a close examination of Figure 1b of Reference [3]. The resulting uncertainty bars are applied in Figure 7.
The model curves shown in Figure 7 begin a short time past zero. This is more of an artifact of the solution method. The solution of the five CSTRs-in-series (Equation (21)) very quickly reaches an early pseudo steady-state before any significant poisoning occurs. The declining conversion trends after the initial drops are due to real activity losses over time observed for the θ-Al2O3 and γ-Al2O3 catalysts. The model fit to the γ-Al2O3 data is remarkably good. The model fits for the θ-Al2O3 and Ga/θ-Al2O3 cases fall within the estimated experimental XA uncertainties. These results demonstrate the ability of the reactor model to acceptably predict CF4 conversion over time with or without poisoning.

4. Conclusions

The mechanism cycle proposed by Zhang et al. [3] on the alumina-catalyzed hydration of CF4 has been expanded to show details for the hydration up to the point of potential F-poisoning. The expanded cycle is analyzed with the Langmuir–Hinshelwood algorithm to give a kinetic rate expression. Consideration of the reported experimental conditions in [3] results is a first-order expression for which Arrhenius parameters are completed based on data in [3]. The rate expression is combined with an unsteady catalyst activity expression, and then applied in simulations of experimental reactor runs shown in [3]. The results show reasonable simulations of the observed constant activity of Ga/θ-Al2O3 and the declining activities for the θ-Al2O3 and γ-Al2O3 catalysts due to F-poisoning. The kinetic rate expression and accompanying activity coefficient offer the opportunity to design future larger-scale hydration reactors.

Funding

This research received no external funding.

Data Availability Statement

Dataset available on request from the author.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Kakaei, N.; Amirian, R.; Azadi, M.; Mohammadi, G.; Izadi, Z. Perfluorocarbons: A perspective of theranostic applications and challenges. Front. Bioeng. Biotechnol. 2023, 11, 1115254. Available online: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC10436007/ (accessed on 20 December 2024). [CrossRef] [PubMed]
  2. Biden-Harris Administration Finalizes First-Ever National Drinking Water Standard to Protect 100M People from PFAS Polution. Available online: https://www.epa.gov/newsreleases/biden-harris-administration-finalizes-first-ever-national-drinking-water-standard (accessed on 1 October 2024).
  3. Zhang, H.; Luo, T.; Chen, Y.; Liu, K.; Li, H.; Pensa, E.; Fu, J.; Lin, Z.; Chai, L.; Cortes, E.; et al. Highly efficient decomposition of perfluorocarbons for over 1000 hours via active site regeneration. Angew. Chem. Int. Ed. 2023, 62, e202305651. [Google Scholar] [CrossRef] [PubMed]
  4. Chen, Y.; Qu, W.; Luo, T.; Zhang, H.; Fu, J.; Li, H.; Liu, C.; Zhang, D.; Liu, M. Promoting C-F bond activation via proton donor for CF4 decomposition. Proc. Natl. Acad. Sci. USA 2023, 120, e2312480120. [Google Scholar] [CrossRef] [PubMed]
  5. Anus, A.; Sheraz, M.; Jeong, S.; Kim, E.; Kim, S. Catalytic thermal decomposition of tetrafluoromethane (CF4): A review. J. Anal. Appl. Pyrolysis 2021, 156, 105126. [Google Scholar] [CrossRef]
  6. Wang, J.; Lin, Z.; He, X.; Song, M.; Westerhoff, P.; Doudrick, K.; Hanigan, D. Critical Review of Thermal Decomposition of Per- and Polyfluoroalkyl Substances: Mechanisms and Implications for Thermal Treatment Processes. Environ. Sci. Technol. 2022, 56, 5355–5370. [Google Scholar] [CrossRef] [PubMed]
  7. Li, X.; Zhong, F.; Li, P.; Xiao, J.; Xi, J. Transition metal modified Al2O3 mesoporous nanospheres for catalysis of organic reactions. Appl. Surf. Sci. 2024, 653, 159355. [Google Scholar] [CrossRef]
  8. Rajendran, S.; Palani, G.; Shanmugam, V.; Trilaksanna, H.; Kannan, K.; Nykiel, M.; Korniejenko, K.; Marimuthu, U. A Review of Synthesis and Applications of Al2O3 for Organic Dye Degradation/Adsorption. Molecules 2023, 28, 7922. [Google Scholar] [CrossRef] [PubMed]
  9. Wang, X.; Zhang, H.; Chen, Y.; Zheng, J.; Chen, H.; Liu, K.; Fu, J.; Lin, Z.; Chai, L.; Liu, M. Promoted CF4 decomposition via enhanced tricoordinated Al active sites. ACS EST Eng. 2024, 4, 1142–1148. [Google Scholar] [CrossRef]
  10. Zhang, T.; Luo, Y.; Long, Y.; Chen, J.; Fu, J.; Liu, H.; Hu, J.; Lin, Z.; Chai, M. Indentification of the active site during CF4 hydrolytic decomposition over g-Al2O3. Environ. Sci. Nano 2022, 9, 954–963. [Google Scholar] [CrossRef]
  11. Zhang, H.; Liu, K.; Chen, Y.; Wang, X.; Li, H.; Fu, J.; Chai, L.; Lin, Z.; Liu, M. Efficient and stable CF4 deocmposition over θ-Al2O3 with extraordinary resistance to HF. Environ. Sci. Nano 2023, 10, 3149. [Google Scholar] [CrossRef]
  12. El-Bahy, Z.; Ohnishi, R.; Ichikawa, M. Hydrolysis of CF4 over alumina-based binary metal oxide catalysts. Appl. Catal. B Environ. 2003, 40, 81–91. [Google Scholar] [CrossRef]
  13. Chen, Y.; Kao, C.-W.; Luo, T.; Zhang, H.; Long, Y.; Fu, J.; Lin, Z.; Chai, L.; Chan, T.-S.; Liu, M. Enhanced surface Lewis acidity of ZrO2 by -HSO4 for efficient CF4 decomposition. Environ. Sci. Nano 2024, 11, 881–888. [Google Scholar] [CrossRef]
  14. Fogler, H.S. Elements of Chemical Reaction Engineering, 6th ed.; Pearson: New York, NY, USA, 2020. [Google Scholar]
  15. Available online: https://polymathplus.org (accessed on 1 October 2024).
Figure 1. Portion I—Stable sites S1 and S2; transition states TS1 and TS2. S1 regenerated (fresh).
Figure 1. Portion I—Stable sites S1 and S2; transition states TS1 and TS2. S1 regenerated (fresh).
Chemengineering 09 00010 g001
Figure 2. Portion I—Stable sites S3, S4 and transition state TS3. S4 is a “poisoned” site.
Figure 2. Portion I—Stable sites S3, S4 and transition state TS3. S4 is a “poisoned” site.
Chemengineering 09 00010 g002
Figure 3. Portion II—Stable site S5 and transition states TS4 and TS5.
Figure 3. Portion II—Stable site S5 and transition states TS4 and TS5.
Chemengineering 09 00010 g003
Figure 4. Portion II—Stable sites S6, S1 and transition state TS6.
Figure 4. Portion II—Stable sites S6, S1 and transition state TS6.
Chemengineering 09 00010 g004
Figure 5. Approximation of PBR by five CSTRs-in-series.
Figure 5. Approximation of PBR by five CSTRs-in-series.
Chemengineering 09 00010 g005
Figure 6. Observed and predicted CF4 conversion for the conditions in Figure 1a of Reference [3].
Figure 6. Observed and predicted CF4 conversion for the conditions in Figure 1a of Reference [3].
Chemengineering 09 00010 g006
Figure 7. Observed and predicted CF4 conversion for the conditions in Figure 1b of Reference [3]. Temperature = 873 K, P = 1 atm.
Figure 7. Observed and predicted CF4 conversion for the conditions in Figure 1b of Reference [3]. Temperature = 873 K, P = 1 atm.
Chemengineering 09 00010 g007
Table 1. Arrhenius parameters for CF4 hydration.
Table 1. Arrhenius parameters for CF4 hydration.
CatalystAf (Mole/g-min) *Ea (Joule/Mole) #
Ga/θ-Al2O323183,900
θ-Al2O332,613122,000
γ-Al2O3779,907144,700
* This work, # Reference [3].
Table 2. Physical properties of catalyst and catalyst bed.
Table 2. Physical properties of catalyst and catalyst bed.
Catalystρcat (g/cm3)ϕ (est)
Ga/θ-Al2O34.23 (est)0.811
θ-Al2O33.690.779
γ-Al2O33.650.783
Table 3. Activity loss rate constant kp as used in the Figure 7 simulations.
Table 3. Activity loss rate constant kp as used in the Figure 7 simulations.
Catalystkp (min−1)
Ga/θ-Al2O30
θ-Al2O30.035
γ-Al2O30.080
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Barat, R. Engineering Calculations for Catalytic Hydrolysis of CF4. ChemEngineering 2025, 9, 10. https://doi.org/10.3390/chemengineering9010010

AMA Style

Barat R. Engineering Calculations for Catalytic Hydrolysis of CF4. ChemEngineering. 2025; 9(1):10. https://doi.org/10.3390/chemengineering9010010

Chicago/Turabian Style

Barat, Robert. 2025. "Engineering Calculations for Catalytic Hydrolysis of CF4" ChemEngineering 9, no. 1: 10. https://doi.org/10.3390/chemengineering9010010

APA Style

Barat, R. (2025). Engineering Calculations for Catalytic Hydrolysis of CF4. ChemEngineering, 9(1), 10. https://doi.org/10.3390/chemengineering9010010

Article Metrics

Back to TopTop