Skip to main content
Log in

Optimization of Nanoparticle Collection by a Pilot-Scale Spray Scrubber Operated Under Waste Incineration Conditions: Using Box–Behnken Design

  • Original Paper
  • Published:
Waste and Biomass Valorization Aims and scope Submit manuscript

Abstract

The growing production and application of engineered nanomaterials (ENM) in everyday products make their environmental discharge increasingly likely. This risk is particularly elevated at the end of the useful life of ENM. Given the lack of regulations targeting the safe disposal of ENM at end-of-life, nanowaste (NW) currently follows conventional waste management pathways even though their potential adverse effects on humans and the natural environment are well documented in the literature. Presently, preference exists for treating NW via centralized waste-to-energy systems such as incineration, given the potentially hazardous nature of NW. In this work, a Design of experiment (DOE) methodology—Box Behnken design was employed to optimize three independent operating parameters of a pilot-scale scrubber concerning the collection of nanoparticles under waste incineration conditions. The pilot-scale scrubber was designed and operated to be representative of a full-scale spray scrubber found in a hazardous waste incineration plant. The independent parameters studied are; the gas flowrate, the liquid flowrate and the droplet diameter. Six responses were chosen corresponding to the nanoparticle collection efficiencies at particle sizes in the diffusion-dominant, intermediate and impaction-dominant regions. The experimental results were analyzed for statistical significance using Design Expert software (V13). The DOE model by the software was then used to predict optimum operating conditions with maximum nanoparticle collection. A maximum nanoparticle collection efficiency of 41–60% was predicted by the DOE model under optimal conditions of a gas flowrate of 46.1 Nm3 h−1, a liquid flowrate of 4.8 L min−1, and a droplet diameter of 60 µm. These optimal conditions were determined by the highest liquid flowrate, the smallest droplet diameter, and a gas flowrate near the middle of the range studied. The results of the confirmatory experiments (43–62%) at the optimum combination agreed with the predicted data. The experimental results were also found to be in good agreement when compared to a mechanistic model based on impaction, Brownian diffusion and interception collection mechanisms.

Graphical Abstract

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

Data Availability

The data generated or analyzed during this work are available from the corresponding author upon reasonable request.

References

  1. Wennersten, R., Fidler, J., Spitsyna, A.: Nanotechnology: a new technological revolution in the 21st century. In: Handbook of Performability Engineering, pp. 943–952. Springer London, London (2008)

  2. Khan, I.I., Saeed, K., Khan, I.I.: Nanoparticles: properties, applications and toxicities. Arab. J. Chem. 12, 908–931 (2019). https://doi.org/10.1016/j.arabjc.2017.05.011

    Article  Google Scholar 

  3. Global Industry Analysts: Global Nanotechnology Industry, p. 793. Report ID.: 326269 (2022)

  4. Keller, A.A., Lazareva, A.: Predicted releases of engineered nanomaterials: from global to regional to local. Environ. Sci. Technol. Lett. 1, 65–70 (2014). https://doi.org/10.1021/ez400106t

    Article  Google Scholar 

  5. Nemmar, A., Hoet, P.H.M., Vanquickenborne, B., Dinsdale, D., Thomeer, M., Hoylaerts, M.F., Vanbilloen, H., Mortelmans, L., Nemery, B.: Passage of inhaled particles into the blood circulation in humans. Circulation 105, 411–414 (2002). https://doi.org/10.1161/hc0402.104118

    Article  Google Scholar 

  6. Maher, B.A., Ahmed, I.A.M., Karloukovski, V., MacLaren, D.A., Foulds, P.G., Allsop, D., Mann, D.M.A., Torres-Jardón, R., Calderon-Garciduenas, L.: Magnetite pollution nanoparticles in the human brain. Proc. Natl. Acad. Sci. USA 113, 10797–10801 (2016). https://doi.org/10.1073/pnas.1605941113

    Article  Google Scholar 

  7. Nowack, B., Bucheli, T.D.: Occurrence, behavior and effects of nanoparticles in the environment. Environ. Pollut. 150, 5–22 (2007). https://doi.org/10.1016/j.envpol.2007.06.006

    Article  Google Scholar 

  8. Thomas, S., Al Mutairi, E., De, S.: Impact of nanomaterials on health and environment. Arab. J. Sci. Eng. 38, 457–477 (2013). https://doi.org/10.1007/s13369-012-0324-0

    Article  Google Scholar 

  9. Campos, A., López, I.: Current status and perspectives in nanowaste management. In: Handbook of Environmental Materials Management, pp. 2287–2314. Springer (2019)

  10. Allan, J., Belz, S., Hoeveler, A., Hugas, M., Okuda, H., Patri, A., Rauscher, H., Silva, P., Slikker, W., Sokull-Kluettgen, B., Tong, W., Anklam, E.: Regulatory landscape of nanotechnology and nanoplastics from a global perspective. Regul. Toxicol. Pharmacol. 122, 104885 (2021). https://doi.org/10.1016/j.yrtph.2021.104885

    Article  Google Scholar 

  11. Mostafa, S.A., EL-Deeb, M.M., Farghali, A.A., Faried, A.S.: Evaluation of the nano silica and nano waste materials on the corrosion protection of high strength steel embedded in ultra-high performance concrete. Sci. Rep. 11, 1–16 (2021). https://doi.org/10.1038/s41598-021-82322-0

    Article  Google Scholar 

  12. Walser, T., Limbach, L.K., Brogioli, R., Erismann, E., Flamigni, L., Hattendorf, B., Juchli, M., Krumeich, F., Ludwig, C., Prikopsky, K., Rossier, M., Saner, D., Sigg, A., Hellweg, S., Günther, D., Stark, W.J.: Persistence of engineered nanoparticles in a municipal solid-waste incineration plant. Nat. Nanotechnol. 7, 520–524 (2012). https://doi.org/10.1038/nnano.2012.64

    Article  Google Scholar 

  13. Vejerano, E.P., Leon, E.C., Holder, A.L., Marr, L.C.: Characterization of particle emissions and fate of nanomaterials during incineration. Environ. Sci. Nano 1, 133–143 (2014). https://doi.org/10.1039/c3en00080j

    Article  Google Scholar 

  14. Baran, P., Quicker, P.: Fate and behavior of nanoparticles in waste incineration. Osterr. Wasser- und Abfallwirtschaft. 69, 51–65 (2016). https://doi.org/10.1007/s00506-016-0362-z

    Article  Google Scholar 

  15. Part, F., Berge, N., Baran, P., Stringfellow, A., Sun, W., Bartelt-Hunt, S., Mitrano, D., Li, L., Hennebert, P., Quicker, P., Bolyard, S.C., Huber-Humer, M.: A review of the fate of engineered nanomaterials in municipal solid waste streams. Waste Manag. 75, 427–449 (2018). https://doi.org/10.1016/j.wasman.2018.02.012

    Article  Google Scholar 

  16. Kim, H.T., Jung, C.H., Oh, S.N., Lee, K.W.: Particle removal efficiency of gravitational wet scrubber considering diffusion, interception, and impaction. Environ. Eng. Sci. 18, 125–136 (2001). https://doi.org/10.1089/10928750151132357

    Article  Google Scholar 

  17. Adah, E., Joubert, A., Boudhan, R., Henry, M., Durécu, S., Le Coq, L.: Spray scrubber for nanoparticle removal from incineration fumes from the incineration of waste containing nanomaterials: theoretical and experimental investigations. Aerosol Sci. Technol. 56, 75–91 (2021). https://doi.org/10.1080/02786826.2021.1974332

    Article  Google Scholar 

  18. Di Natale, F., Carotenuto, C., D’Addio, L., Jaworek, A., Krupa, A., Szudyga, M., Lancia, A.: Capture of fine and ultrafine particles in a wet electrostatic scrubber. J. Environ. Chem. Eng. 3, 349–356 (2015). https://doi.org/10.1016/j.jece.2014.11.007

    Article  Google Scholar 

  19. Huang, C.-H., Tsai, C.-J., Wang, Y.-M.: Control efficiency of submicron particles by an efficient venturi scrubber system. J. Environ. Eng. 133, 454–461 (2007). https://doi.org/10.1061/(ASCE)0733-9372(2007)133:4(454)

    Article  Google Scholar 

  20. Tsai, C.-J., Lin, C.-H., Wang, Y.-M., Hunag, C.-H., Li, S.-N., Wu, Z.-X., Wang, F.-C.: An efficient venturi scrubber system to remove submicron particles in exhaust gas. J. Air Waste Manag. Assoc. 55, 319–325 (2005). https://doi.org/10.1080/10473289.2005.10464622

    Article  Google Scholar 

  21. Yang, L., Bao, J., Yan, J., Liu, J., Song, S., Fan, F.: Removal of fine particles in wet flue gas desulfurization system by heterogeneous condensation. Chem. Eng. J. 156, 25–32 (2010). https://doi.org/10.1016/j.cej.2009.09.026

    Article  Google Scholar 

  22. Vasudevan, T.V., Gokhale, A.J., Mahalingam, R.: Phoretic phenomena in tar vapor-particulate mixture separation from fuel gas streams. Can. J. Chem. Eng. 63, 903–910 (1985). https://doi.org/10.1002/cjce.5450630606

    Article  Google Scholar 

  23. Commission, E., Centre, J.R., Pinasseau, A., Zerger, B., Roth, J., Canova, M., Roudier, S.: Best available techniques (BAT) reference document for waste treatment: Industrial Emissions Directive 2010/75/EU (integrated pollution prevention and control). Publications Office of the European Union (2018)

  24. Chen, T.M., Tsai, C.J., Yan, S.Y., Li, S.N.: An efficient wet electrostatic precipitator for removing nanoparticles, submicron and micron-sized particles. Sep. Purif. Technol. 136, 27–35 (2014). https://doi.org/10.1016/j.seppur.2014.08.032

    Article  Google Scholar 

  25. Dunuweera, S.P., Rajapakse, R.M.G.: Cement types, composition, uses and advantages of nanocement, environmental impact on cement production, and possible solutions. Adv. Mater. Sci. Eng. 2018, 4158682 (2018). https://doi.org/10.1155/2018/4158682

    Article  Google Scholar 

  26. Jung, C.H., Park, H.S., Kim, Y.P.: Analytic solution to estimate overall collection efficiency and the size distribution change of polydispersed aerosol for cyclone separator. Sep. Purif. Technol. 68, 428–432 (2009). https://doi.org/10.1016/j.seppur.2009.05.023

    Article  Google Scholar 

  27. Lim, K.S., Lee, S.H., Park, H.S.: Prediction for particle removal efficiency of a reverse jet scrubber. J. Aerosol Sci. 37, 1826–1839 (2006). https://doi.org/10.1016/j.jaerosci.2006.06.010

    Article  Google Scholar 

  28. Wu, Q., Gu, M., Du, Y., Zeng, H.: Synergistic removal of dust using the wet flue gas desulfurization systems. R. Soc. Open Sci. 6, 181696 (2019). https://doi.org/10.1098/rsos.181696

    Article  Google Scholar 

  29. Taheri, M., Haines, G.F.: Optimization of factors affecting scrubber performance. J. Air Pollut. Control Assoc. 19, 427–431 (1969). https://doi.org/10.1080/00022470.1969.10466508

    Article  Google Scholar 

  30. Ravi, G., Gupta, S.K., Viswanathan, S., Ray, M.B.: Optimization of venturi scrubbers using genetic algorithm. Ind. Eng. Chem. Res. 41, 2988–3002 (2002). https://doi.org/10.1021/ie010531b

    Article  Google Scholar 

  31. Viswanathan, S., Ananthanarayanan, N., Azzopardi, B.: Venturi scrubber modelling and optimization. Can. J. Chem. Eng. 83, 194–203 (2008). https://doi.org/10.1002/cjce.5450830206

    Article  Google Scholar 

  32. Mukherjee, S., Verma, A., Biswas, S., Bal, M., Meikap, B.: Removal of cement dust particulates via fully submerged self-primed venturi scrubber. Clean: Soil, Air, Water 49, 2000241 (2021). https://doi.org/10.1002/clen.202000241

    Article  Google Scholar 

  33. Ferreira, S.L.C., Bruns, R.E., Ferreira, H.S., Matos, G.D., David, J.M., Brandão, G.C., da Silva, E.G.P., Portugal, L.A., dos Reis, P.S., Souza, A.S., dos Santos, W.N.L.: Box–Behnken design: an alternative for the optimization of analytical methods. Anal. Chim. Acta 597, 179–186 (2007). https://doi.org/10.1016/j.aca.2007.07.011

    Article  Google Scholar 

  34. Kotsilkov, S., Ivanov, E., Vitanov, N.K.: Release of Graphene and Carbon Nanotubes from Biodegradable Poly(Lactic Acid) Films During Degradation and Combustion: Risk Associated with the End-of-Life of Nanocomposite Food Packaging Materials (2018)

  35. Bouillard, J.X., R’Mili, B., Moranviller, D., Vignes, A., Le Bihan, O., Ustache, A., Bomfim, J.A.S.S., Frejafon, E., Fleury, D., R’Mili, B., Moranviller, D., Vignes, A., Le Bihan, O., Ustache, A., Bomfim, J.A.S.S., Frejafon, E., Fleury, D.: Nanosafety by design: risks from nanocomposite/nanowaste combustion. J. Nanopart. Res. 15, 1519 (2013). https://doi.org/10.1007/s11051-013-1519-3

    Article  Google Scholar 

  36. Box, G.E.P., Behnken, D.W.: Some new three level designs for the study of quantitative variables. Technometrics 2, 455–475 (1960). https://doi.org/10.1080/00401706.1960.10489912

    Article  MathSciNet  Google Scholar 

  37. Jung, C.H., Lee, K.W.: Filtration of fine particles by multiple liquid droplet and gas bubble systems. Aerosol Sci. Technol. 29, 389–401 (1998). https://doi.org/10.1080/02786829808965578

    Article  Google Scholar 

  38. Raj Mohan, B., Jain, R.K., Meikap, B.C.: Comprehensive analysis for prediction of dust removal efficiency using twin-fluid atomization in a spray scrubber. Sep. Purif. Technol. 63, 269–277 (2008). https://doi.org/10.1016/j.seppur.2008.05.006

    Article  Google Scholar 

  39. Nguyen, T.H., Nguyen, N.T., Nguyen, T.T., Doan, N.T., Tran, L.A., Nguyen, L.P., Bui, T.T.: Optimizing the conditions of cationic polyacrylamide inverse emulsion synthesis reaction to obtain high–molecular–weight polymers. Polymers 14, 2866 (2022)

    Article  Google Scholar 

  40. Greenfield, S.: Rain scavenging of radioactive particulate matter from the atmosphere. J. Meteorol. 14, 115–125 (1957). https://doi.org/10.1175/1520-0469(1957)014%3c0115:RSORPM%3e2.0.CO;2

    Article  Google Scholar 

  41. Derringer, G., Suich, R.: Simultaneous optimization of several response variables. J. Qual. Technol. 12, 214–219 (2018). https://doi.org/10.1080/00224065.1980.11980968

    Article  Google Scholar 

  42. Mohan, B.R., Biswas, S., Meikap, B.C.: Performance characteristics of the particulates scrubbing in a counter-current spray-column. Sep. Purif. Technol. 61, 96–102 (2008). https://doi.org/10.1016/j.seppur.2007.09.018

    Article  Google Scholar 

  43. Ali, M., Yan, C., Sun, Z., Gu, H., Mehboob, K.: Dust particle removal efficiency of a venturi scrubber. Ann. Nucl. Energy 54, 178–183 (2013). https://doi.org/10.1016/j.anucene.2012.11.005

    Article  Google Scholar 

  44. Pak, S.I., Chang, K.S.: Performance estimation of a Venturi scrubber using a computational model for capturing dust particles with liquid spray. J. Hazard. Mater. 138, 560–573 (2006). https://doi.org/10.1016/J.JHAZMAT.2006.05.105

    Article  Google Scholar 

  45. Mayinger, F., Lehner, M.: Operating results and aerosol deposition of a venturi scrubber in self-priming operation. Chem. Eng. Process. Process Intensif. 34, 283–288 (1995). https://doi.org/10.1016/0255-2701(94)04015-X

    Article  Google Scholar 

  46. Meikap, B.C., Biswas, M.N.: Fly-ash removal efficiency in a modified multi-stage bubble column scrubber. Sep. Purif. Technol. 36, 177–190 (2004). https://doi.org/10.1016/S1383-5866(03)00213-2

    Article  Google Scholar 

  47. Lee, B.-K., Mohan, B.R., Byeon, S.-H., Lim, K.-S., Hong, E.-P.: Evaluating the performance of a turbulent wet scrubber for scrubbing particulate matter. J. Air Waste Manag. Assoc. 63, 499–506 (2013). https://doi.org/10.1080/10962247.2012.738626

    Article  Google Scholar 

  48. Darbandi, T., Risberg, M., Westerlund, L.: CFD modeling of the forces in the wet scrubber acting on particulate matter released from biomass combustion. Therm. Sci. Eng. Prog. 25, 100997 (2021). https://doi.org/10.1016/J.TSEP.2021.100997

    Article  Google Scholar 

  49. Zhao, W., Fu, Y., Chen, Y., Sun, J., Tang, Q., Jiang, B., Kong, H., Xu, F., Ji, Q.: Effective fenton catalyst from controllable framework doping of Fe in porous silica spheres. Microporous Mesoporous Mater. 312, 110704 (2021). https://doi.org/10.1016/j.micromeso.2020.110704

    Article  Google Scholar 

  50. Schauer, P.J.: Removal of submicron aerosol particles from moving gas stream. Ind. Eng. Chem. 43, 1532–1538 (1951). https://doi.org/10.1021/ie50499a022

    Article  Google Scholar 

  51. Lancaster, B.W., Strauss, W.: A study of steam injection into wet scrubbers. Ind. Eng. Chem. Fundam. 10, 362–369 (1971)

    Article  Google Scholar 

  52. Yoshida, T., Kousaka, Y., Okuyama, K.: Growth of aerosol particles by condensation. Ind. Eng. Chem. Fundam. 15, 37–41 (1976). https://doi.org/10.1021/i160057a007

    Article  Google Scholar 

  53. Bao, J., Yang, L., Yan, J., Xiong, G., Shen, X.: Experimental investigation on improving the removal effect of WFGD system on fine particles by heterogeneous condensation. In: Qi, H., Zhao, B. (eds.) Cleaner Combustion and Sustainable World, pp. 667–675. Springer (2013). https://doi.org/10.1007/978-3-642-30445-3_92

    Chapter  Google Scholar 

  54. Fan, F., Yang, L., Yuan, Z., Hu, X.: Removal and condensation growth of inhalable particles in spray scrubber. Huagong Xuebao/CIESC J. 61, 2708–2713 (2010)

    Google Scholar 

  55. Fan, F., Yang, L., Yan, J., Bao, J., Shen, X.: Experimental investigation on removal of coal-fired fine particles by a condensation scrubber. Chem. Eng. Process. Process Intensif. 48, 1353–1360 (2009). https://doi.org/10.1016/J.CEP.2009.06.011

    Article  Google Scholar 

  56. Lehner, M.: Aerosol separation efficiency of a venturi scrubber working in self-priming mode. Aerosol Sci. Technol. 28, 389–402 (1998). https://doi.org/10.1080/02786829808965533

    Article  Google Scholar 

  57. Calvert, S.: Venturi and other atomizing scrubbers efficiency and pressure drop. AIChE J. 16, 392–396 (1970). https://doi.org/10.1002/AIC.690160315

    Article  Google Scholar 

  58. Danzomo, B.A., Salami, M.J.E., Jabirin, S., Khan, M.R., Nor, I.M.: Performance evaluation of wet scrubber system for industrial air pollution control (2012)

  59. Wang, S., Wang, J., Song, C., Wen, J.: Numerical investigation on urea particle removal in a spray scrubber using particle capture theory. Chem. Eng. Res. Des. 145, 150–158 (2019). https://doi.org/10.1016/J.CHERD.2019.03.011

    Article  Google Scholar 

  60. Rafidi, N., Brogaard, F., Chen, L., Håkansson, R., Tabikh, A.: CFD and experimental studies on capture of fine particles by liquid droplets in open spray towers. Sustain. Environ. Res. 28, 382–388 (2018). https://doi.org/10.1016/j.serj.2018.08.003

    Article  Google Scholar 

  61. Carotenuto, C., Di Natale, F., Lancia, A.: Wet electrostatic scrubbers for the abatement of submicronic particulate. Chem. Eng. J. 165, 35–45 (2010). https://doi.org/10.1016/j.cej.2010.08.049

    Article  Google Scholar 

  62. Mi, T., Yu, X.M.: Dust removal and desulphurization in a novel venturi scrubber. Chem. Eng. Process. Process Intensif. 62, 159–167 (2012). https://doi.org/10.1016/j.cep.2012.07.010

    Article  Google Scholar 

  63. Keshavarz, P., Bozorgi, Y., Fathikalajahi, J., Taheri, M.: Prediction of the spray scrubbers’ performance in the gaseous and particulate scrubbing processes. Chem. Eng. J. 140, 22–31 (2008). https://doi.org/10.1016/j.cej.2007.08.034

    Article  Google Scholar 

  64. Pranesha, T.S., Kamra, A.K.: Scavenging of aerosol particles by large water drops 1. Neutral case. J. Geophys. Res. Atmos. 101, 23373–23380 (1996). https://doi.org/10.1029/96jd01311

    Article  Google Scholar 

  65. Pilat, M.J., Jaasund, S.A., Sparks, L.E.: Collection of aerosol particles by electrostatic droplet spray scrubbers. Environ. Sci. Technol. 8, 360–362 (1974). https://doi.org/10.1021/es60089a006

    Article  Google Scholar 

  66. Zhao, H., Zheng, C.G.: Modeling of gravitational wet scrubbers with electrostatic enhancement. Chem. Eng. Technol. 31, 1824–1837 (2008). https://doi.org/10.1002/ceat.200800360

    Article  Google Scholar 

  67. Ardon-Dryer, K., Huang, Y.-W., Cziczo, D.J.: Laboratory studies of collection efficiency of sub-micrometer aerosol particles by cloud droplets on a single-droplet basis. Atmos. Chem. Phys. 15, 9159–9171 (2015). https://doi.org/10.5194/acp-15-9159-2015

    Article  Google Scholar 

  68. Lee, J.: Black carbon and elemental carbon concentrations of spark-generated carbon particles. In: ETH-Conference on combustion generated nanoparticles, vol. 25 (2013001650004), pp. 1–37. Zurich, Switzerland (2008)

Download references

Funding

This work was carried out under the collaborative agreement for the supervision of doctoral research N° ADEME: TEZ19-002 and was supported by the French Agency for Ecological Transition (ADEME), the Région Pays de la Loire and the enterprise Séché Environnement.

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed to the research conception and design. Full-scale scrubber data were measured and supplied by SD and MH. EA performed material preparation, experiments, and data collection. EA, AJ, and LLC performed data analysis and interpretation. EA wrote the first draft of the manuscript and all authors critically reviewed all versions. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Aurélie Joubert.

Ethics declarations

Conflict of interest

The authors declare that they have no known competing financial or non-financial interests that could have influenced the work reported in this paper.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix: Supplementary Information

Gas Flowrate

The gas flowrate in a scrubber could (dis)favor the collection of particles by a given mechanism, thus affecting the overall performance of the wet scrubber [28, 42]. In their work on the modeling of PM (1–3 µm) collection efficiency by a spray scrubber with twin-fluid air-assist atomizers, Mohan et al. [38] revealed that at a constant liquid flowrate of 16.67 × 10–6 m3 s−1, an increase in the gas flowrate (from 1 × 10–6 to 5 × 10–6 m3 s−1) led to an improvement in the particle collection efficiency of the scrubber. Similar conclusions were reached by Lim et al. [27] in their study to predict the particle collection efficiency of a reverse jet scrubber. Lim et al. [27] argued that the increase in the collection efficiency as the gas flowrate increased as a result of an increase in the relative velocities between the droplets and the gas (particles). In a spray scrubber, Wu et al. [28] obtained an increment of the dust (20 µm in diameter) removal efficiency from 90.7 to 95.6% when they studied the impact of varying the gas flowrate from 79 to 129 m3 h−1. Wu et al. [28] also reported an increase in the dust removal efficiency from 80.9 to 94.8% when the gas flowrate of a sieve-tray spray scrubber was increased from 83 to 136 m3 h−1. Wu et al. [28] associated this improvement to improve inertia collision with higher gas flowrate. In their research on the study of the removal efficiency of dust in a self-priming venturi scrubber, Ali et al. [43] also concluded that a higher gas flow led to a higher removal efficiency due to improved relative velocity between gas and droplets in the throat section. Generally, as the gas flowrate in a scrubber increases, the particle collection is improved [44, 45]. This is not always the case. Meikap et al. [46] conducted experimental studies on the scrubbing of fly ash (2–15 µm) in a novel wet scrubber using water as the scrubbing medium. Beyond a gas flow rate of 4.2 × 10−3 Nm3 s−1, the collection efficiency did not improve with an increase in the gas flowrate but was instead limited due to coalescence which led to reduced bubble reformation. Lee et al. [47] reported that higher gas flowrates do indeed favored the collection of sub-micrometer particles smaller than 1 µm. Lee et al. [47] argues that the gas–liquid contact for effective particle collection is well established for larger particles even at low gas flow rates.

Gas temperature

A change in the flue gas temperature in a wet scrubber could play a significant role in the collection of particles by the scrubber. The gas temperature could also modify other influent parameters such as the droplet diameter, droplet evaporation rate and gas flowrate. Darbandi et al. [48] conducted CFD modeling to investigate the forces responsible for the collection of nanoparticles in a wet scrubber. The authors varied the flue gas temperature from 100 to 300 °C increasing the collection efficiency. The temperature gradient positively impacted the collection of particles due to the diffusiophoresis mechanism. Fan et al. [49] presented in their paper the experimental collection of coal-fired particles (PM2.5) by a condensation scrubber. For the particle size 1.0–3.0 m, the removal efficiency was higher (15–35%) when the inlet gas temperature was set at 95 °C than it was when the gas temperature was at 35 °C (8–35%). The authors attributed this to heterogeneous nucleation and particle collection due to the diffusiophoresis mechanism as a result of the temperature gradient.

Gas Humidity Rate

The collection of particles in a flue gas in a wet scrubber could be improved if the particles are grown into larger ones by a preconditioning technique such as heterogeneous condensation of water vapor with the particles acting as nuclei [21, 50,51,52,53]. The grown particles can then be efficiently collected by an inertial mechanism or a mist separator. Heterogeneous condensation of water vapor is most suitable for gas streams with significant humidity content. Steam is injected in the gas inlet or above the scrubbing liquid inlet to saturate the wet scrubbers with water vapor which leads the latter to condense on the fine particles and to grow their volume. Several authors have investigated this effect on the removal of fine particles in a scrubber. The influence of humidity on the removal of a fine particle by a wet scrubber was studied by Yang et al. [54]. The authors found that an increase in the gas inlet steam led to an increase in the amount of condensable water on the particle surface. This resulted in the improvement of the collection efficiency from 50 to 85%. Huang et al. [19] attained a significantly enhanced collection efficiency of particles with diameters less than 150 nm by mixing saturated steam at 100 °C with a normal temperature waste stream permitting submicron particles to grow into micron sizes. At a liquid-to-gas ratio of 2.5 L m−3, the removal efficiency of particles less than 100 nm by the venturi scrubber was greater than 90%. Fan et al. [55] conducted experimental studies on the removal of coal-fired fine particles (PM2.5) by a condensation scrubber. The authors added steam to the gas inlet. This resulted in the improvement of the removal of fine particles as the amount of condensable water vapor increased leading to the condensation growth of particles.

Liquid Flowrate

The liquid flow rate is another very important operating parameter in a wet scrubber. Lehner et al. [56] investigated the aerosol (PM3.0) collection efficiency of a venturi scrubber working in self-priming mode (i.e., operating without an external pump). The authors found that, as the liquid flowrate increased from 1.5 to 3.0 L m−3 the collection efficiency for the entire particle range also increased. Lehner et al. [56] attributed this to the increase in the interfacial surface for particle collection at higher liquid flow rates. Meikap et al. [46] reported that at a constant gas flow rate of 3.0 × 10−3 Nm3 s−1 in a multi-stage bubble column scrubber, the fly-ash (2–15 µm) collection efficiency increases as the liquid flow rate is increased. Meikap et al. [46] argued that this was due to an increase in the dispersed phase hold-up and interfacial contact area. Meikap et al. [46] conclusion was supported by the investigation made by Calvert et al. [57]. Likewise, as the liquid flowrate increases, the number of collectors also increases leading to a better droplet-particle collision. A similar conclusion was reached by Wu et al. [28] when they developed a mathematical model to investigate the dust removal efficiency of wet flue gas desulfurization systems. The overall spray scrubber collection efficiency increased as the liquid flowrate increased from 1.32 to 2.26 m3 h−1. Wu et al. [28] determined that more intense moving droplets increased the chances of particle capture. Vasudevan et al. [22] conducted lab-scale studies using a spray scrubber to treat the flue gas stream from a fixed bed gasifier. Vasudevan et al. [22] reported that, as the spray liquid flow rate increased from 25 to 30 mL min−1, collection efficiency improved from 50.3 to 60%. Vasudevan et al. [22] attributed this to an increase in the number of droplets due to higher liquid flowrate.

Droplet Diameter

The droplet diameter plays an important role in the collection of particles by a wet scrubber. Danzomo et al. [58] developed a model for the collection efficiency of particles in a counter-current scrubber at different droplet sizes of 500 µm, 1000 µm, 1500 µm and 2000 µm. For 5 µm and 10 µm particle sizes, the collection efficiencies were found to be 89.7% and 98.2% at 500 µm droplet diameter and for a liquid-to-gas ratio of 0.7 L m−3. As the droplet diameter increased to 2000 µm, the collection efficiency at 5 µm and 10 µm decreased to 43.9% and 58.8% respectively. Simin et al. [59] investigated the scrubbing of urea duct particles in a spray scrubber using CFD simulations. As the droplet diameter increased from 200 to 1000 µm, the collection efficiency at particle size 40 µm decreased. Simin et al. [59] concluded that the droplet diameter is negatively correlated to the collection efficiency thus, the surface area for particle collection increases with decreasing droplet diameter. Rafidi et al. [60] developed a model to predict the removal of particles in a spray scrubber and compared the results with experiments from a full-scale spray scrubber in a coal-fired power plant. Rafidi et al. [60] reported that the collection efficiency increased as the droplets became smaller. The authors argued that smaller droplets have higher velocities, thus increasing relative velocities between the droplets and the particles, resulting in a higher Stokes number and hence higher collection efficiency by impaction. Claudia et al. [61] developed a mathematical model to evaluate particle capture efficiency in a wet electrostatic scrubber. The authors found that the model predicted a decrease in the capture efficiency for particle sizes 100 nm and 1 µm with an increase in droplet sizes. As droplet sizes increases, both the collisional efficiency and the droplet–particle contact probability are increased and an improvement of removal efficiency is obtained. Larger particles (5 µm) did not witness an increase in the capture efficiency even when the droplet diameters were smaller. Claudia et al. [61] concluded that the effect of a lower collisional efficiency competed with the increase of droplet–particle contact probability leading to lower capture efficiency.

Particle Diameter

Although not an operating parameter, the influence of the particle diameter has been investigated by several authors. The diameter of the particle in a flue gas is arguably the most important parameter in determining its collection by a wet scrubber [22, 59]. Lee et al. [47] evaluated the performance of a turbulent wet scrubber to effectively remove dust particles arising from a coal-powered thermal power plant. Lee et al. [47] disclosed that an increase in the fly ash particle diameter from 0.65 to 5.0 µm led to an increase in the removal efficiency from 43 to 100%. A similar tendency was observed in a venturi scrubber by Mi et al. [62], where the overall removal efficiency improved from 92.7 to 95.1% as the particle size increased from 10 to 45 µm. In a study involving a sieve-tray spray scrubber by Wu et al. [28], the dust collection efficiency improved from 14.8 to 99% as the particle size increased from 1 to 20 µm. In a spray scrubber, Keshavarz et al. [63] achieved a significant increase in the particle collection efficiency from 60% at a particle size of 0.5 µm to almost 90% at a particle size of 4.5 µm. A laboratory experiment to measure the collection efficiency of aerosol particles by water drops was studied by Pranesha and Kamra [64]. The authors reported that the collection efficiency increased with an increase in particle size. Pranesha and Kamra [64] also suggested that with an increase in particle diameter, the collection efficiency may attain an upper limit which may be less than one. Simin et al. [59] observed a similar tendency at a particle size of 10 µm with a maximum collection efficiency of 97% in spray scrubber. Another point to consider is that several authors [16, 40, 60, 64,65,66,67,68] have reported a tendency of the particle collection efficiency to increase with particle diameter, then dip in the intermediate region around 0.1–0.6 µm and then increase till it reaches 100% as the particle diameter increases. As particle diameter increases, the collection due to the impaction mechanism improves leading to the overall performance of the scrubber. Pilat et al. [65] confirmed this when they calculated the particle collection efficiencies of single droplets considering impaction, diffusion, diffusiophoresis and thermophoresis. The particle collection efficiency improved (30–88%) with increasing particle size (0.25–4 µm) predominantly due to the inertial impaction mechanism.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Adah, E., Joubert, A., Henry, M. et al. Optimization of Nanoparticle Collection by a Pilot-Scale Spray Scrubber Operated Under Waste Incineration Conditions: Using Box–Behnken Design. Waste Biomass Valor 14, 3455–3474 (2023). https://doi.org/10.1007/s12649-023-02097-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12649-023-02097-5

Keywords

Navigation