Removal of SO2 and NO from flue gas by wet scrubbing using an aqueous NaClO2 solution

https://doi.org/10.1016/S0304-3894(00)00274-0Get rights and content

Abstract

This study used a NaClO2/NaOH solution as the additive/absorbent to determine the extent of NOx removal in a wet scrubbing system. A combined SOx/NOx removal system was also tested. The experiments were performed in a bench-scale spraying sieve tray wet scrubber in a continuous mode. The operating variables included NO and SO2 concentrations, L/G ratio, molar ratio, and initial pH.

The results of the individual DeNOx experiments show that the maximum DeNOx efficiencies ranged from 3.1 to 12.6%. The results of the combined DeSOx/DeNOx experiments show that the maximum DeNOx and DeSOx efficiencies ranged from 36.6 to 71.9% and from 89.4 to 100.0%, respectively. The major parameters affecting NOx removal efficiencies are the L/G ratio and the dosage of additive. The major parameter influencing DeSOx efficiencies is the L/G ratio.

Introduction

Acid precipitation is a current air pollution problem caused mainly by SO2 and NOx. Conventionally, engineers use wet scrubbing SO2; the most widely used process is flue gas desulfurization (FGD) process. Although the wet FGD system has experienced high SO2 removal efficiencies, that is not so for NOx. The reason for this process failure is that NO, which accounts for more than 90% of NOx in the flue gas, is quite insoluble in water. Since the wet scrubbing system currently dominates the FGD market, a minor adjustment of the system may work for the combined SOx/NOx removal system and should be cost-effective. The DeSOx wet scrubbing SOx removal system is not new to us. It has, however, many unknowns if it is to be used as a DeNOx system. In general, additives have to be added to the scrubbing system to convert relatively water insoluble NO to soluble NO2 which can then be removed by alkaline absorbents.

Possible NO oxidants are ClO2 or O3 [1], [2]. However, these additives are quite expensive as well as very dangerous in equipment operating in the gas phase. Therefore, chemical reagents added to the liquid phase have been widely used lately. There are many reagents such as FeSO4/H2SO4, Fe(II) EDTA, KMnO4/NaOH, NaClO2/NaOH, Na2S/NaOH, Na2S2O4/NaOH, H2O2, Na2SO3, FeSO4/Na2SO3, CO(NH2)2, NaOH, Na2CO3 and P4 (yellow phosphorus) that have been tested for NOx absorption [3], [4], [5], [6]. Of this group, NaClO2 was the most effective reagent [7], [8], [9], [10], [11], [12], [13], [14], [15].

The absorption of NOx in NaClO2 solution was studied by Teramoto et al. [7] and Sada et al. [8], [9], [10] in the late seventies. Teramoto et al. investigated the effect of various operating conditions on the absorption rates in mixed aqueous solutions of NaClO2 and NaOH, using a semi-batch agitated vessel with a flat gas–liquid interface. They found that the absorption of NO proceeded in the region of the fast pseudo mth order reaction and the absorption rate was not affected by liquid-phase mass transfer coefficient. The NO concentration dependence of the absorption rate was more pronounced in the higher concentration range than that in the lower concentration range. The absorption rate increased greatly with NaClO2. However, the absorption rate decreased markedly with an increase in NaOH. The effect of temperature on the absorption rate was found to be much greater in the lower concentration range on NO.

Sada et al. [8] also performed a series of kinetic studies using NaClO2 as the additive in a stirred tank absorber. The concentration of NO in the gas stream ranged from 0.8 to 15%. The rate of reaction was found to be second order in NO and first order in NaClO2. They also did some work on the absorption of dilute NO concentrations (<1%) with the same NaClO2 concentrations as above. The results showed that the order of reaction in NO varied from 2 to 1 because of the lower interfacial concentration of NO [9]. In both cases, they claimed that the oxidation power of NaClO2 increased with a decrease in pH value of the solution. However, the reaction product, NO2, may desorb from the solution in the case with less OH. Therefore, the addition of OH to the solution is required in order to fix NO2 as NO3. That suggested that the reaction proceeds as follows:4NO+3ClO2+4OH→4NO3+3Cl+2H2O

In addition to the above kinetic studies, some researchers published their performance investigations on the absorption of NOx. The experimental equipment, operating conditions and results of these studies [11], [12], [13], [14], [15] are summarized in Table 1. Chan [11] found that the liquid and gas flow rates had little effect on NOx removal, while increasing liquid chlorite concentrations enhanced NOx absorption. Absorption of NOx tended to be totally gas-film controlled at all NOx input levels. The experimental results of Yang and co-workers [12], [13] showed that NO can be quantitatively oxidized by NaClO2 in an aqueous solution. During scrubbing, NO was oxidized to NO3 and ClO2 was converted to Cl. Due to the production of HNO3, the pH value decreased sharply from 10 to 3 within minutes of operation. Low concentrations of NaOH increased the effectiveness of NO absorption in the NaClO2 aqueous scrubbing solution by 7%, while higher NaOH concentrations decreased or inhibited the absorption. They also suggested that the reactions proceed under acidic condition as follows:4NO+3NaClO2+2H2O4HNO3+3NaCl5NO+4HCl→4ClO2+5NaCl+2H2O5NO+3ClO2+4H2O5HNO3+NaCl

Brogren et al. [14] found that the pH value of the absorbing liquid had a significant impact on the absorption efficiency. The major fraction of the nitrogen oxides was absorbed via the hydrolysis of N2O3 and N2O4. Hsu et al. [15] performed the NO absorption study using a packed column. Their results indicated that the NO oxidation efficiency and removal efficiency could reach 98.8 and 61.5%, respectively.

So far, no one has attempted to determine the removal efficiencies of NO and SO2 simultaneously by using a NaClO2 solution in a spraying sieve tray wet scrubbing system. In this work, we used a low concentration NaClO2 solution under acidic condition in a bench-scale spraying sieve tray column to investigate the effect of various parameters on SO2 and NOx absorption efficiency.

Section snippets

Experimental setup

The experimental system used in this work included a simulated flue gas production system, a scrubber, and a sampling and analysis system as shown in Fig. 1.

Measurement of spray droplet size and specific surface area

By using laser droplet size distribution analyzer (Malvern 2600) to analyze the size of the sprayed droplets and the specific surface area, the mean droplet size was determined to range from 595 to 1085 μm at an L/G ratio of 4–10 l/m3. As shown in Fig. 2, the higher the L/G ratios were, the greater the gas–liquid contacting surface area was.

Preliminary tests

The results of preliminary tests are shown in Fig. 3, Fig. 4, and Table 2, Table 3. It was found that the NOx removal efficiency was satisfactory at an L/G

Conclusions

For the combined SOx/NOx removal system employed in this study, the maximum NOx and SO2 removal efficiencies are in the range of 36–72 and 88–100%, respectively. These results indicate that this process could be utilized for as FGD system as well as being developed for the simultaneous removal of SO2 and NOx from flue gas. The DeNOx efficiency is mainly affected by the L/G ratio and the concentration of NaClO2 in the solution.

Further work needs to be done on the kinetics of absorption of lean SO

Acknowledgements

This study was funded by National Science Council, Republic of China (Contract no. NSC 83–0410-E-006-075).

References (16)

  • Tennessee Valley Authority, State-of-the-art review for simultaneous removal of nitrogen oxides and sulfur oxides from...
  • S.S., Novoselov, et al., Thermal Eng. 33 (1986)...
  • H. Kobayashi, et al., Environ. Sci. Technol. 11 (1977)...
  • J.B. Joshi, et al., Chem. Eng. Commun. 33 (1988)...
  • K.R. Jethani, et al., Gas Sep. Purif. 4 (1990)...
  • G.C Lee et al.

    Environ. Prog.

    (1992)
  • M. Teramoto, et al., Jpn. Chem. Eng. 2 (1976) 637–640 (in...
  • E. Sada, et al., Chem. Eng. Sci. 33 (1978)...
There are more references available in the full text version of this article.

Cited by (198)

View all citing articles on Scopus
View full text