• No results found

University of Groningen Development and design of the in-situ regeneration section of Vitrisol®, a novel, highly selective desulphurization process Wermink, Wouter Nicolaas

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Development and design of the in-situ regeneration section of Vitrisol®, a novel, highly selective desulphurization process Wermink, Wouter Nicolaas"

Copied!
15
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Development and design of the in-situ regeneration section of Vitrisol®, a novel, highly

selective desulphurization process

Wermink, Wouter Nicolaas

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date: 2019

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Wermink, W. N. (2019). Development and design of the in-situ regeneration section of Vitrisol®, a novel, highly selective desulphurization process. Rijksuniversiteit Groningen.

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

33

Chapter 3: The oxidation of Fe(II) in acidic sulphate solutions with air at

elevated pressures. Part 2. Influence of H

2

SO

4

and Fe(III)

Reproduced with permission from Ind. Eng. Chem. Res. 2017, 56, 14, 3789-3796, link. Copyright 2017 American Chemical Society.

Abstract

The oxidation of ferrous ions in acidic sulphate solutions at elevated air pressures was investigated. The effect of the Fe2+ concentration, Fe3+ concentration, H2SO4 concentration and partial oxygen

pressure on the reaction rate were determined at three different temperatures, i.e. T = 90 °C, 70 °C and 50 °C. A second order of reaction in Fe2+ and a first order of reaction in O2 were determined,

respectively. A slight inhibition by Fe3+ on the Fe2+ oxidation reaction was observed. The reaction is first order in Fe3+ in the kinetic term for the inhibition by Fe3+. Concentrations of H2SO4 up to 1 M

result in a fractional negative order of -0.6; concentrations of H2SO4 above 1 M result in a zero order.

One kinetic equation for the oxidation of Fe2+ was postulated, in which the order of reaction in H2SO4

is changed depending on the H2SO4 concentration.

[ ] [ ] [ ] ( [ ]) With c = -0.6 for [H2SO4] < 1 M c = 0 for [H2SO4] > 1 M

The activation energy was determined to be EA = 62.1 kJ/mol. The order c in H2SO4 is either -0.6, or

zero, depending on the H2SO4 concentration.

3.1 Introduction

Hydrogen sulphide (H2S) is a highly toxic and corrosive gas. Removal of H2S from acidic gas streams,

such as natural gas, industrial gas or biogas, is important for safety, health, environmental and economic reasons. Several regenerative and non-regenerative H2S removal processes are readily

available, which are economically viable only for specific gas compositions and gas flow rates. Apart from non-regenerative H2S removal by the use of e.g. adsorbents, all the regenerative aqueous liquid

redox processes (e.g. THIOPAQ, LO-CAT, SulFerox) capture CO2 to varying extents besides H2S.

The novel Vitrisol® desulphurization process is based on the removal of H2S by precipitation with

copper sulphate (CuSO4) in an aqueous, acidic solution.1 Copper sulphide (CuS) and sulphuric acid are

formed in the gas treating process, without coabsorption of CO2:2,3

( ) ( ) (3.1)

The Vitrisol® liquid can be regenerated through oxidation of CuS by ferric sulphate (Fe2(SO4)3).

Copper sulphate, elemental sulphur (So) and ferrous sulphate (FeSO4) are produced in the process:

( ) ( ) ( ) (3.2)

Ferrous sulphate can be re-oxidized to ferric sulphate with O2, according to:

(3)

34

Resulting in the overall net reaction for the removal of H2S:

( ) (3.4)

In Part 1 Wermink and Versteeg4 investigated the oxidation of ferrous ions in acidic sulphate solutions (Reaction 3.3), and proposed a power law kinetic equation for the oxidation of Fe2+ at concentrations of H2SO4 of 1 M and higher. The kinetic equation was derived by fitting initial reaction

rates.

In this study (Part 2) the influence of H2SO4 below and above a concentration of 1 M, and the

influence of Fe3+ on the oxidation of Fe2+, were studied. The kinetic equation was derived by fitting experimentally determined Fe2+ concentration profiles.

As will be discussed in this study, both methods of determining a kinetic equation, i.e. by fitting initial reaction rates (Part 1) and fitting experimentally determined Fe2+ concentration profiles (Part 2), respectively, resulted in the same temperature dependence of the reaction and orders in Fe2+, O2 and

H2SO4 for concentrations of H2SO4 above 1 M.

In Part 1 Wermink and Versteeg4 concluded that interpreting the results in terms of activities or chemical potentials could possibly clarify the influence of H2SO4 on the reaction rate and mechanism,

respectively, because the effect of H2SO4 on the oxidation reaction might be attributed to SO42- ions.

The kinetic equations derived in Part 14 and this study are used for the design of the Vitrisol® process. An empirically derived kinetic equation is sufficient for process design, if the derived kinetic equation can be applied satisfactory for concentrations of compounds encountered in the process. Therefore an extended thermodynamic model is beyond the scope of the present study.

To investigate the effect of Fe(II) complexes in solution and minerals known to significantly enhance the Fe2+ oxidation rate, i.e. e.g. hydrolyzed Fe2+ complexes5,6,7 and Fe(III)(oxyhdr)oxide minerals,8 speciation calculations with an extended thermodynamic model might be required. However, concentrations of these species were negligible due to the low pH in our study (i.e. a pH of 2 and below). King,5 Santana-Casiano et al.,6 Pham and Waite7 and Jones et al.8 studied the oxidation of

Fe2+ at a pH of 4 and higher. Solids formation was not observed during experimentation, which agrees well with the study by Ter Maat et al.2 that Fe(II) hydroxide precipitation starts at a pH of 5.85 and higher for a 1 M FeSO4 solution. If Fe(III)(oxyhydr)oxides were to be present, an increase in

oxidation rate because of Fe2+ adsorbed to Fe(III)(oxyhydr)oxide surfaces would only be expected at a pH of 4.5 and higher as explained by Jones et al.8

For an extended literature review on the oxidation reaction of ferrous ions in acidic sulphate solutions the reader is referred to Wermink and Versteeg.4

A detailed description of the experimental setup, procedures and analytical techniques was provided by Wermink and Versteeg.4

3.2 Results

The kinetic equation for the oxidation of ferrous ions in acidic sulphate solutions was determined by fitting experimentally determined Fe2+ concentration profiles. In the present study, the following kinetic equation was assumed:

[ ] [ ] [ ] ( [ ] ) (3.5)

The orders of reaction a, b, c and d, the constant A as well as the reaction rate constant k, were derived by fitting experimentally determined Fe2+ concentration profiles using the Newton-Raphson

(4)

35 optimization method. The sulphuric acid concentration in Equation 3.5 is considered to be the lumped sulphuric acid concentration, i.e. the dissociation of H2SO4 is not taken into account.

For determining the orders of reaction and the reaction rate constants, 18 experiments at T = 90 °C, 9 experiments at T = 70 °C and 10 experiments at T = 50 °C were used. Experimental conditions are presented in Table 3.1.

Table 3.1: Fe2+ oxidation experiments.

Exp T (°C) PO2 (MPa) [Fe2+] [H2SO4]

6 90 0.0901 0.235 0.237 7 90 0.0450 0.261 0.263 8 90 0.0225 0.257 0.259 9 90 0.0901 1.00 1.03 10 90 0.0450 1.00 1.03 11 90 0.0225 1.00 1.00 12 90 0.0901 0.503 0.524 13 90 0.0450 0.504 0.524 14 90 0.0225 0.503 0.525 16 90 0.0901 0.248 1.01 17 90 0.0901 0.494 1.01 19 90 0.0901 0.248 0.507 21 90 0.0901 0.126 1.03 22 90 0.0901 0.252 2.03 23 90 0.0113 0.500 0.520 24 90 0.0113 1.01 1.03 25 90 0.0113 0.250 0.253 26 70 0.0982 0.253 1.03 27 70 0.0982 0.500 1.02 29 70 0.0982 0.253 0.527 30 70 0.0982 0.251 0.271 31 70 0.0982 0.997 1.02 32 70 0.0901 0.503 0.523 33 70 0.0451 0.503 0.524 34 70 0.0226 0.502 0.523 35 70 0.0982 0.253 2.04 36 50 0.102 1.01 1.03 37 50 0.102 0.501 1.02 38 50 0.102 0.254 1.04 39 50 0.102 0.503 0.525 40 50 0.102 0.501 0.271 41 50 0.102 0.498 2.01 42 50 0.0901 1.01 1.03 43 50 0.0451 1.01 1.03 45 50 0.180 1.01 1.03 46 90 0.180 1.00 1.02 47 50 0.0226 1.01 1.03

(5)

36

3.2.1 Previous paper

In Part 1 by Wermink and Versteeg4 the order of reaction in a reactant was determined through the so-called pseudo-first order approach. By varying the initial concentration of one of the reactants in several experiments, the order in this reactant can be determined from the initial reaction rates of the experiments. The reactants with equal initial concentrations are lumped together with the reaction rate constant.

The power law kinetic equation derived in Part 1 for H2SO4 concentrations above 1 M was of the

form

[ ]

[ ] (3.6)

Table 3.2 summarizes the orders in Fe2+ and O2 from their paper, determined from initial reaction

rates. Experiments with initial concentration of sulphuric acid of 1.0 M and above were used.

Table 3.2: Orders in Fe2+ and O2 at varying temperatures.

Component

Temperature (°C)

50 70 90

Fe2+ 2.08 2.21 2.17

O2 0.94 1.01 0.99

From Table 3.2 it can be concluded that the order of reaction in Fe2+ becomes 2 and the order of reaction in O2 becomes 1.

Table 3.3 summarizes the reaction rate constants from their paper, determined from initial reaction rates.

Table 3.3: Reaction rate constants at varying temperatures. Temperature (°C) k (m3.kmol-1.Pa-1.s-1)

50 1.94 x 10-10

70 7.15 x 10-10

90 2.30 x 10-9

The activation energy was determined to be EA = 60.3 kJ/mol.

3.2.2 This study

As mentioned before (Equation 3.5), the kinetic equation was assumed to be of the form

[ ] [ ] [ ] ( [ ] )

In Table 3.4 the results of the determination of the various reaction orders in Fe2+, O2, H2SO4 and Fe3+,

as derived from the Fe2+ concentration profiles at temperatures of 50 °C, 70 °C and 90 °C, are summarized.

Table 3.4: Orders in Fe2+ and O2 at varying temperatures.

Component Fe2+ O2 H2SO4 (below 1 M) H2SO4 (above 1 M) Fe3+

(6)

37 The constant A in Equation 3.5 is 0.767. As the order in Fe3+ is -1, it is clear that ferric ions inhibit the Fe2+ oxidation reaction.

Table 3.5 summarizes the reaction rate constants determined by curve fitting Fe2+ concentration profiles.

Table 3.5: Reaction rate constants at varying temperatures. Temperature (°C) k (m3.kmol-1.Pa-1.s-1)

50 1.45 x 10-10

70 5.55 x 10-10

90 1.86 x 10-9

By using the Arrhenius equation, the temperature dependence of the reaction rate constant was determined:

{ } (3.7)

A linear curve fit of ln(k) versus T-1 resulted in EA and k0 (see Figure 3.1).

Figure 3.1: ln(k) versus T-1.

Resulting in a value of the activation energy of EA = 62.1 kJ/mol. This is in good agreement with the

value of EA determined in Part 1.4

{ } (3.8)

3.2.3 Concentration profiles

The accuracy of the kinetic equation, derived in this study, in simulating Fe2+ concentration profiles for the oxidation of Fe2+ in acidic sulphate solutions, is shown by comparing experimental and simulated Fe2+ concentration profiles for various experimental settings.

-23.5 -23 -22.5 -22 -21.5 -21 -20.5 -20 0.0027 0.00275 0.0028 0.00285 0.0029 0.00295 0.003 0.00305 0.0031 0.00315

ln

(k)

T

-1

(K

-1

)

(7)

38

The experimental and simulated concentration profiles of several Fe2+ oxidation experiments are shown in Figures 3.2, 3.3 and 3.4. For the simulations the order in H2SO4 has been adjusted in

Equation 3.5 as proposed in Table 3.4. The simulated concentration profiles according to Equations 3.5 and 3.8 are shown as solid lines. The experimental settings are provided in Tables 3.6, 3.7 and 3.8.

Figure 3.2 shows the effect of a change in both concentration of H2SO4, as well as temperature, on

the oxidation rate of Fe2+.

Table 3.6: Experimental settings of experiments in Figure 3.2. Exp T (°C) PO2 (MPa) [Fe2+] [H2SO4]

6 90 0.0901 0.235 0.237

16 90 0.0901 0.248 1.01

22 90 0.0901 0.252 2.03

26 70 0.0982 0.253 1.03

35 70 0.0982 0.253 2.04

Figure 3.2: Experimental Fe2+ concentration profiles of experiments 6, 16, 22, 26 and 35 and simulated Fe2+ concentration profiles of experiments 6, 22 and 35.

From Figure 3.2 it can be concluded that decreasing the H2SO4 concentration, below an initial H2SO4

concentration of 1 M, increases the rate of oxidation of Fe2+. Furthermore, an increase in H2SO4

concentration above a concentration of 1 M does not result in a change in Fe2+ oxidation rate; this observation seems to be temperature independent.

Figure 3.3 shows the effect of the variation of the concentration of Fe2+, as well as temperature,

respectively, on the oxidation rate of Fe2+. It should be noted that there is a slight variation in the oxygen partial pressure with temperature. The experiments were performed at a constant air pressure of 5 bara, and the vapor pressure of water is temperature dependent.

0.00 0.05 0.10 0.15 0.20 0.25 0.30 0 2000 4000 6000 8000 10000 12000 14000 16000 [Fe 2+] (M ) t (s) simulated Exp6 Exp16 Exp22 Exp26 Exp35

(8)

39 Table 3.7: Experimental settings of experiments in Figure 3.3.

Exp T (°C) PO2 (MPa) [Fe2+] [H2SO4]

17 90 0.0900 0.494 1.01

26 70 0.0982 0.253 1.03

27 70 0.0982 0.500 1.02

31 70 0.0982 0.997 1.02

36 50 0.102 1.008 1.03

Figure 3.3: Experimental and simulated Fe2+ concentration profiles of experiments 17, 26, 27, 31 and 36.

From Figure 3.3 it can be concluded that increasing the Fe2+ concentration increases the rate of oxidation of Fe2+ more than proportional.

Figure 3.4 shows the effect of a change in partial pressure of O2 on the oxidation rate of Fe2+.

Table 3.8: Experimental settings of experiments in Figure 3.4. Exp T (°C) PO2 (MPa) [Fe2+] [H2SO4]

45 50 0.180 1.01 1.03

46 90 0.180 1.00 1.02

47 50 0.0226 1.01 1.03

From Figure 3.4 it can be concluded that lowering the partial pressure of O2 decreases the rate of

oxidation of Fe2+ proportionally.

The simulated concentration profiles show good agreement with the experimental data and the determined kinetic expression according to Equations 3.5 and 3.8.

0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 10000 20000 30000 40000 50000 [Fe 2+] (M ) t (s) simulated Exp17 Exp26 Exp27 Exp31 Exp36

(9)

40

Figure 3.4: Experimental and simulated Fe2+ concentration profiles of experiments 45, 46 and 47.

3.2.4 Parity plot

To evaluate the derived kinetic expression according to Equations 3.5 and 3.8, the

experimentally determined concentrations are compared to the calculated concentrations,

respectively.

A parity plot of the comparison of the Fe

2+

concentrations, used for fitting, is shown in Figure

3.5. In total 316 Fe

2+

concentrations were compared.

The parity plot shows good agreement between experimental and simulated Fe2+ concentrations. The maximum deviation between the experimental and simulated Fe2+ concentrations is 28.3%, the average deviation is 1.32%.

3.3 Discussion

In the previous paper by Wermink and Versteeg4 (Part 1) orders of reaction, activation energies and reaction rates of previous studies were discussed. The present study agrees well with the previous paper with respect to orders of reaction in Fe2+ and O2 and the activation energy. In the present

paper the orders of reaction in Fe3+ and H2SO4, the effect of Fe3+ on the reaction rate and a

comparison of the kinetic equations in Part 1 and this study will be discussed.

3.3.1 Orders of reaction

The order in H2SO4 of -0.60 agrees well with the order reported by Wermink and Versteeg4 in Part 1

of -0.62 to -0.72 and the order reported by Iwai et al.9 of -0.6.

0.0 0.2 0.4 0.6 0.8 1.0 0 5000 10000 15000 20000 25000 [Fe 2+] (M ) t (s) simulated Exp45 Exp46 Exp47

(10)

41 Figure 3.5: Parity plot of experimental and calculated Fe2+ concentrations.

An inhibition by ferric ions on the Fe2+ oxidation rate was observed by Iwai et al.9 as well. According to Iwai et al.,9 the oxidation of ferrous ions with oxygen proceeds via two parallel reactions, i.e. one

dependent on sulphate ions in which the reaction rate is proportional to the SO42- concentration, and

one independent of sulphate ions. The Fe(II) sulphate complex was assumed to be the compound susceptible to oxidation with oxygen in the sulphate dependent path. They stated that the retardation of the Fe2+ oxidation reaction may be due to a decrease in sulphate ion activity through the formation of Fe3+ sulphate complexes. The Fe2+ oxidation rate was determined to be inversely proportional to the Fe3+ concentration. However, their kinetic equation only accounts for formation constants of Fe(II) sulphate complexes, and not for formation constants of Fe(III) sulphate complexes.

Huffman and Davidson,10 McKay,11 McKay and Halpern12 and Dreisinger and Peters13 observed no inhibition by Fe3+ on the Fe2+ oxidation rate.

Huffman and Davidson10 studied the effect of Fe3+ on the oxidation of Fe2+ at a temperature of 160 °C with initial concentrations of Fe3+, Fe2+ and H2SO4 of 0.001 M to 0.02 M, 0.001 M to 0.005 M and 1 M,

respectively. It is likely that due to the high sulphuric acid concentration almost no free sulphate ions were present in solution, i.e. almost all the sulphate ions were present as bisulphate ions. Therefore ferric ions could not have affected the oxidation rate of Fe2+. Another possible explanation is the effect of temperature on the formation constants of Fe(III) sulphate complexes, i.e. the formation constants could be substantially low at high temperatures.

McKay11 and McKay and Halpern12 investigated the effect of Fe3+ on the oxidation of Fe2+ at a temperature of 100 °C with initial concentrations of Fe3+, Fe2+ and H2SO4 of 0.054 M, 0.054 M and

0.08 M, respectively. Due to the low concentration of H2SO4, free sulphate ions should be present in

relatively large amounts, resulting in the formation of Fe(III) sulphate complexes. Possibly at a temperature of 100 °C the formation constants of Fe(III) sulphate complexes are relatively low. Dreisinger and Peters13 studied the effect of Fe3+ on the oxidation of Fe2+ at a temperature of 150 °C in the presence of ZnSO4. Initial concentrations of Fe3+, Fe2+, Zn2+ and H2SO4 were 0.1 M, 0.2 M, 1.9 M

and 0.5 M, respectively. Free sulphate ions should be present in relatively large amounts. Possibly at a temperature of 160 °C the formation constants of Fe(III) sulphate complexes are relatively low. This

0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0

[F

e

2+

]

si m

[Fe

2+

]

exp T = 90 °C T = 70 °C T = 50 °C

(11)

42

assumption seems to confirm the statement in the study by Dreisinger and Peters that all equilibria involving Fe3+ ions have been disregarded because of their limited effect on ferrous speciation. They did not clarify this statement with the support of calculations of Fe(III) sulphate complexes.

Huffman and Davidson,10 Iwai et al.9 and Dreisinger and Peters13 agree that sulphate ions form complexes with ferrous ions in solution, which oxidize at a faster rate with oxygen compared to unpaired Fe2+ ions. Wermink and Versteeg4 showed in Part 1 that a decrease in sulphuric acid concentration decreases the H3O+ and bisulphate concentration, but increases the sulphate

concentration, resulting in an increase in Fe2+ oxidation rate. This conclusion was based on speciation calculations accounting only for the dissociation constants of sulphuric acid. Mathews and Robins14 acknowledged that a change in sulphate concentration affects the H3O+ concentration and therefore

influences the Fe2+ oxidation rate. McBain,15 Lamb and Elder,16 Pound,17 McKay,11 McKay and Halpern,12 Mathews and Robins14 and Verbaan and Crundwell18 reported a decrease in Fe2+ oxidation rate with increasing H2SO4 concentration.

It can be concluded that Fe2+ in solutions containing sulphate ions oxidizes at an increased rate with oxygen compared to solutions containing almost no sulphate ions. It seems plausible that Fe(II) sulphate complexes in solution are more susceptible to oxidation with oxygen than unpaired Fe2+ ions, or Fe(II) bisulphate complexes. This statement agrees well with the structure of the kinetic equation of Iwai et al.,9 i.e. two parallel reactions occur, one proceeding independent of sulphate ions, and one proceeding dependent of sulphate ions. To determine such a type of kinetic equation, a correct determination of the speciation is desirable.

3.3.2 Inhibition by ferric ions

The kinetic equation reported in this study is not based on speciation. Only the order in H2SO4

changes according to the H2SO4 concentration. According to Huffman and Davidson,10 Iwai et al.9 and

Dreisinger and Peters13 Fe(II) sulphate complexes oxidize more readily than unpaired Fe2+. Therefore the effect of inhibition by Fe3+ ions should be more pronounced at higher Fe2+ conversions, because increased concentrations of the species Fe3+ and SO42- are present, respectively, at higher Fe2+

concentrations.

This assumption is verified for the kinetic equation developed in the present study compared to a kinetic equation not accounting for the influence of Fe3+. The kinetic parameters of the latter kinetic equation were fitted according to the method described in this study. Fe2+ concentration profiles calculated with and without inhibition by ferric ions for the same experiment are illustrated in Figure 3.6.

A slight improvement in fit of the experimental points is obtained when the kinetic equation accounts for the inhibition by Fe3+. A possible explanation could be that the inhibition by Fe3+ via the formation of Fe(III) sulphate complexes is a combined effect; i.e. the kinetic term for H2SO4 accounts

for a change in sulphate concentration too.

As initial concentrations of H2SO4 were relatively high in the experiments performed in this study, it is

recommended to perform additional experiments at lower initial H2SO4 concentrations to investigate

(12)

43 Figure 3.6: Fe2+ oxidation profiles calculated with and without Fe3+ inhibition. Initial experimental conditions: T = 90 °C, Pair = 0.52 MPa (40 vol.% O2), [FeSO4] = 1.0 M, [H2SO4] = 1.0 M.

3.3.3 Comparison of kinetic equations

Two different methods of interpreting kinetics were performed by Wermink and Versteeg4 in Part 1 and this study (Part 2).

In Part 1 initial reaction rates of experiments were determined by fitting the data points via a polynomial equation. The order of reaction in a reactant was determined through the so-called pseudo-first order approach (for a detailed description the reader is referred to Part 1). Subsequently the initial reaction rates were fitted with only the reaction rate constants as variables.

In Part 2 experimentally determined concentration profiles were fitted with orders of reaction in components, reaction rate constants and the constant A (see Equation 3.5) as variables, respectively.

The experimental and simulated concentration profiles of Fe2+ oxidation experiments 22, 37 and 46 are shown in Figure 3.7. Experiments with concentrations of sulphuric acid of 1 M and higher were compared, because the kinetic equation derived in Part 1 is only valid for this H2SO4 concentration

range. The simulated concentration profiles were calculated from the kinetic equations developed in Part 1 and this study, respectively. The experimental concentration profiles are shown as solid lines. The experimental settings are provided in Table 3.9.

Table 3.9: Experimental settings of experiments in Figure 3.7. Exp T (°C) PO2 (MPa) [Fe2+] [H2SO4]

22 90 0.0901 0.252 2.03 37 50 0.102 0.501 1.02 46 90 0.180 1.00 1.02 0.0 0.2 0.4 0.6 0.8 1.0 1.2 0 2000 4000 6000 8000 10000 12000 14000 [Fe 2+] (M ) t (s) Experimental Simulation - without Fe3+ inhibition Simulation - with Fe3+ inhibition

(13)

44

Figure 3.7: Experimental and simulated Fe2+ concentration profiles of experiments 22, 37 and 46.

From Figure 3.7 it can be concluded that both methods of determining a kinetic equation, i.e. fitting initial reaction rates (Part 1) and fitting experimentally determined concentration profiles (Part 2), respectively, resulted in an accurate simulation of the experimentally determined concentration profiles.

3.4 Conclusion

The oxidation of ferrous ions in acidic sulphate solutions at elevated air pressures was investigated. The effect of the Fe2+ concentration, Fe3+ concentration, H2SO4 concentration and partial oxygen

pressure on the reaction rate were determined at three different temperatures, i.e. T = 90 °C, 70 °C and 50 °C. From experiments performed with varying stirrer speed it was concluded that the reaction was not affected by mass transfer limitations.

A second order of reaction in Fe2+ and a first order of reaction in O2 were determined, respectively. A

slight inhibition by Fe3+ on the Fe2+ oxidation reaction was observed. The reaction is first order in Fe3+ in the kinetic term for the inhibition by Fe3+. Concentrations of H2SO4 up to 1 M result in a fractional

negative order of -0.6, concentrations of H2SO4 above 1 M result in a zero order.

One kinetic equation for the oxidation of Fe2+ was derived, in which the order of reaction in H2SO4 is

changed depending on the H2SO4 concentration.

[ ] [ ] [ ] ( [ ]) With c = -0.6 for [H2SO4] < 1 M c = 0 for [H2SO4] > 1 M 0.00 0.20 0.40 0.60 0.80 1.00 1.20 0 2000 4000 6000 8000 10000 12000 14000 [Fe 2+] (M ) t (s) Simulation -this study Simulation -Part I Experimental

(14)

45 The activation energy was determined to be EA = 62.1 kJ/mol. The order c in H2SO4 is either -0.6, or

zero, depending on the H2SO4 concentration.

It can be concluded that this study is the first study to include both the effect of H2SO4 and Fe3+ in the

kinetic equation of the oxidation of Fe2+ in acidic sulphate solutions without speciation calculations.

3.5 Nomenclature

A constant

EA activation energy [kJ/mol]

k reaction rate constant [m3.kmol-1.Pa-1.s-1]

P pressure [Pa]

R reaction rate [kmol.m-3.s-1]

R universal gas constant [J.mol-1.K-1]

t time [s]

T temperature [°C, K]

Subscripts and superscripts

a, b, c, d orders sim simulated exp experimental Fe2+ ferrous ion O2 oxygen

3.6 References

[1] Versteeg, G.F. and Ter Maat, H. Method and system for selective removal of contamination from gas flows. WO patent 1998055209 A1, assigned to Procede Twente B.V., priority date June 2, 1997.

[2] Ter Maat, H., Hogendoorn, J.A. and Versteeg, G.F. The removal of hydrogen sulfide from gas streams using an aqueous metal sulfate absorbent. Part I. The absorption of hydrogen sulfide in metal sulfate solutions. Sep. Purif. Technol. 2005, 43 (3), 183─197.

[3] Ter Maat, H., Al-Tarazi, M., Hogendoorn, J.A., Niederer, J.P.M. and Versteeg, G.F. Theoretical and experimental study of the absorption rate of H2S in CuSO4 solutions. The effect of enhancement

of mass transfer by a precipitation reaction. Chem. Eng. Res. Des. 2007, 85 (1), 100─108.

[4] Wermink, W.N. and Versteeg, G.F. The oxidation of Fe(II) in acidic sulphate solutions with air at elevated pressures. Part 1. Kinetics above 1 M H2SO4. Ind. Eng. Chem. Res. 2017, 56 (14), 3775–

3788.

[5] King, D.W. Role of carbonate speciation on the oxidation rate of Fe(II) in aquatic systems. Environ. Sci. Technol. 1998, 32 (19), 2997–3003.

[6] Santana-Casiano, J., Gonzalez-Davila, M. and Millero, F.J. The oxidation of Fe(II) in NaCl-HCO3

-and seawater solutions in the presence of phthalate -and salicylate ions: a kinetic model. Mar. Chem. 2004, 85 (1─2), 27─40.

[7] Pham, A.N. and Waite, T.D. Oxygenation of Fe(II) in natural waters revisited: Kinetics modeling approaches, rate constant estimation and the importance of various reaction pathways. Geochim. Cosmochim. Acta 2008, 72 (15), 3616–3630.

[8] Jones, A.M., Griffin, P.J., Collins, R.N. and Waite, T.D. Ferrous iron oxidation under acidic conditions – The effect of ferric oxide surfaces. Geochim. Cosmochim. Acta 2014, 145, 1─12. [9] Iwai, M., Majima, H. and Awakura, Y. Oxidation of Fe(II) in sulfuric acid solutions with dissolved

molecular oxygen. Metall. Trans. B 1982, 13B, 311─318.

[10] Huffman, R.E. and Davidson, N. Kinetics of the ferrous iron-oxygen reaction in sulfuric acid solution. J. Am. Chem. Soc. 1956, 78, 4836─4842.

(15)

46

[11] McKay, D.R. A kinetic study of the oxidation of pyrite in aqueous suspension. Thesis, University of British Columbia, Vancouver, 1957.

[12] McKay, D.R. and Halpern, J. A kinetic study of the oxidation of pyrite in aqueous suspension. Trans. Metall. Soc. AIME 1958, 212, 301─309.

[13] Dreisinger, D.B. and Peters, E. The oxidation of ferrous sulphate by molecular oxygen under zinc pressure-leach conditions. Hydrometallurgy 1989, 22, 101─119.

[14] Mathews, C.T. and Robins, R.G. The oxidation of aqueous ferrous sulphate solutions by molecular oxygen. Proc. Aust. Inst. Min. Met. 1972, 242, 47─57.

[15] McBain, J.W. Oxidation of ferrous solutions by free oxygen. J. Phys. Chem. 1901, 5, 623.

[16] Lamb, A.B. and Elder, L.W. The electromotive activation of oxygen. J. Am. Chem. Soc. 1931, 53, 137.

[17] Pound, J.R. The oxidation of solutions of ferrous salts. J. Phys. Chem. 1939, 43, 955.

[18] Verbaan, B. and Crundwell, F.K. An electrochemical model for the leaching of a sphalerite concentrate. Hydrometallurgy 1986, 16, 345.

Referenties

GERELATEERDE DOCUMENTEN

For small quantities of sulphur to be removed (up to 50 kg/day of sulphur), technologies are preferred with relatively low capital costs like e.g. For increasing H 2 S

The proposed rate law by Verbaan and Crundwell 19 is only valid in the ferric ion concentration range of 0 M to 0.20 M, because in this concentration range a constant order

Ferrous sulphate oxidation experiments in the presence of copper sulphate were performed at a temperature of 150 °C and at an oxygen partial pressure of 0.138 MPa.. It was not

Wermink and Versteeg 4,7 studied the oxidation of ferrous ions in acidic sulphate solutions (Reaction 5.3), and proposed kinetic equations derived by both fitting the

First, the influence of stirring (or equilibrating) time of the solution before taking a sample, the use of a filter and settling time were investigated. For all the used stirring

3) The gas phase mass transfer coefficient was calculated using the approach of Colombet, 19 however, from this work it is not evident how the effect of process

In some implementations, there is provided a method for selectively removing sulphur and/or sulphur containing contaminant compounds from a gas flow, comprising:

In 2011 he became project leader of the ISPT project OG-10-05 ‘Removal and recovery of sulphur components from natural gas’ which resulted in the present dissertation entitled