Next Article in Journal
Correction: Paciej-Gołębiowska et al. Twenty-Year Mortality Trends in Patients with Kidney Disease in Poland with the Use of the Years of Life Lost Measure, 2000–2019. Int. J. Environ. Res. Public Health 2022, 19, 2649
Previous Article in Journal
Visual Diagnostics of Dental Caries through Deep Learning of Non-Standardised Photographs Using a Hybrid YOLO Ensemble and Transfer Learning Model
Previous Article in Special Issue
Recent Advances in Thallium Removal from Water Environment by Metal Oxide Material
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Influence of Zinc and Humic Acids on Dye Adsorption from Water by Two Composts

1
CRETUS-Department of Soil Science and Agricultural Chemistry, University of Santiago de Compostela, 15782 Santiago de Compostela, Spain
2
Department of Earth Sciences and Environment, Prince El-Hassan bin Talal Faculty for Natural Resources and Environment, The Hashemite University, Zarqa 13133, Jordan
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2023, 20(7), 5353; https://doi.org/10.3390/ijerph20075353
Submission received: 28 February 2023 / Revised: 13 March 2023 / Accepted: 22 March 2023 / Published: 31 March 2023

Abstract

:
Searching for alternative low-cost biosorbents for the removal of textile dyes from wastewater is currently an important subject of research. In this work, we have investigated how the presence of other contaminants in textile wastewaters can affect dye adsorption by biosorbents. We tested the adsorption of three dyes of different types: Basic Violet 10 (BV10), Acid Blue 113 (AB113) and Direct Blue 71 (DB71) by two different composts—municipal solid waste compost and pine bark compost—in the presence of Zn (5 mg L−1) or dissolved organic matter (100 mg humic acids L−1) in batch experiments. Dye adsorption capacity for both composts followed the following sequence: BV10 > AB113 > DB71. In general, dye sorption at the equilibrium was adequately described by the Freundlich model, but not always by the Langmuir model, which did not allow for the estimation of maximum retention capacities in all cases. In general, these were around 1 mg g−1 for DB71, 2 mg g−1 for AB113, and 40 mg g−1 for BV10. Municipal solid waste compost had slightly higher affinity than pine bark compost for the anionic dyes AB113 and DB71, whereas for the cationic dye BV10, pine bark compost presented a much higher adsorption capacity (41.7 mg g−1 versus 6.8 mg g−1). The presence of Zn or dissolved organic matter in the solutions at typical wastewater concentrations did not decrease the dye adsorption capacity of the composts. This result is positive both for the real application of composts to real textile wastewaters and for the validity of the results of biosorbent performance obtained with single-dye solutions.

1. Introduction

The presence of organic dyes in waters represents a major environmental concern for the textile industry because they are highly visible, which greatly affects the perception of water quality by the public, and some types are toxic for organisms [1]. However, many dyes are not readily degraded under the aerobic conditions of biological treatment plants, and other technologies must be used for their elimination, such as electrocoagulation, biodegradation, and others [2]. Among these technologies, low costs of operation and ease of design have made adsorption as a good technology for their removal from water [3,4]. Although activated carbon would be the ideal adsorbent for a wide range of compounds, its high cost has encouraged research for alternative materials as substitutes. Among these, abundant research is being developed around biosorbents, natural materials of different origins, especially of residuals, which can be available in large amounts [4,5,6,7].
In this line of search for low-cost biosorbents, increasing attention has been addressed in recent years to the use of composted wastes [8,9,10,11,12]. Given that they can be produced using a wide range of organic wastes, which otherwise would be landfilled, incinerated, or dumped, composts present interesting characteristics for this use in comparison to other adsorbents, in particular their low production cost and high availability. This low cost in comparison to other adsorbents means that they do not necessarily need to be regenerated and reused several times, so after use they can be discarded or even composted to remove the adsorbed dyes [13]. In this sense, promising results have been presented in the studies that have dealt with textile dye removal from solution by composts so far [8,12,14,15,16,17]. Previous studies from our team have assessed different aspects of the use of compost biosorbents for dye removal, including competitive and noncompetitive removal of different types of dyes, in batch and continuous flow conditions, as well as the effect of other factors such as particle size, solid:liquid ratio, solution pH, or salinity on dye removal, as well as competition with other dyes or continuous flow versus batch conditions [14,15,16,17].
However, some aspects of this application still need to be studied. In most cases, published works have assessed the capacity of composts for dye removal using aqueous solutions of single dyes. However, the presence of other pollutants in wastewaters could potentially interfere with the performance of biosorbents for dye removal. This question is important not only for assessing results in conditions closer to real wastewaters, but also to know if the results obtained from single-dye solutions are comparable or applicable to the treatment of real complex wastewaters. In this context, here we studied dye adsorption by compost in the presence of zinc or dissolved organic matter, which are often presently in textile wastewaters [18], in order to determine the influence of these factors and to assess the validity of the data obtained in laboratory with single aqueous solutions.

2. Materials and Methods

2.1. Adsorbates

Three textile dyes, each of one class (basic, acid, and direct), were used (Table 1). Basic Violet 10 (BV10 or rhodamine B) was purchased from Panreac (Barcelona, Spain), whereas Acid Blue 113 (AB113) and Direct Blue 71 (DB71) were purchased from Sigma–Aldrich® chemicals (Sigma–Aldrich, St. Louis, MO, USA) with purity 98%.

2.2. Biosorbents

In this work we employed two composts produced in composting facilities located in the region of Galicia (northwestern Spain). The first is a municipal solid waste compost (MSWC), obtained at the Complexo Medioambiental do Barbanza (A Coruña, Spain) by large-scale aerobic treatment of the source-separated organic fraction of municipal solid waste. The second one is a composted pine bark (CPB), supplied by the company Costiña Orgánica (A Coruña, Spain), produced by aerobic composting of pine bark in windrows. Properties of the composts are summarized in Table 2. Selection of these specific composts for potential use as biosorbents is due to their limitations to the first-choice application, which would be organic amendments for agronomic use. In this sense, CPB is very acidic and presents very low nutrient levels, while MSWC presents high salinity and Pb and Cu contents. Regarding their surface and spectral properties, studies of N2 adsorption–desorption have shown that both materials exhibit type-II isotherms, according to IUPAC classification, which indicate the predominant mesoporous-to-macroporous structure of the adsorbents. Specific surface area of adsorbents, obtained on multipoint BET analysis, is 5.8 and 22.4 m2 g−1 for MSWC and CPB, respectively (Table 2). Macroporous structure of both composts is confirmed by pore size analysis, with average diameters of 33 Å for MSWC and 90 Å for CPB [15]. The FTIR spectra of the composts (Supplementary Materials) present wide bands in the regions 1000 cm−1 and 1600 cm−1, corresponding to functional groups potentially involved in dye retention. The peaks at 1050–1030 cm−1 are assigned to C-O stretching of carbohydrates, polysaccharides, and alcohol and sulfoxide groups. The bands around 1500–1600 cm−1 can be attributed to C=C in the aromatic rings of lignin and lignocellulose and amide and carboxylate group. Further information about the properties of these composts can be found in Paradelo et al. [14] and Al-Zawahreh et al. [15].

2.3. Adsorption Experiments

2.3.1. Single-Dye Solutions

For the batch adsorption experiments, we prepared a stock solution of 1000 mg L−1 of each single dye in deionized water; this solution was subsequently diluted to the needed concentrations.
The tests for equilibrium adsorption of single dyes were carried out in batch conditions by shaking 0.50 g of each compost with 10 mL of solution with different initial dye concentrations (in the range 0–1000 mg L−1 for BV10 adsorption on CPB, and 0–100 mg L−1 for the rest of systems) at 130 rpm and 25 °C for 24 h. Afterwards, suspensions were centrifuged for five minutes at 4000× g and the clean supernatant was separated for analysis. The dye concentrations (Ce, mg L−1) were quantified by measuring absorbance using an UV/VIS spectrophotometer (Varian Cary 100, Agilent Technologies, Inc., Santa Clara, CA, USA) at 547 nm for BV10, 584 nm for DB71 and 566 nm for AB113. Blanks without adsorbent were run in parallel in order to test dye adsorption to the tubes. All tests were carried out in triplicates.
For the study of dye adsorption kinetics, 0.50 g of each compost was put in contact with 10 mL of a solution with 200 mg L−1 dye in centrifuge tubes. Suspensions were shaken at room temperature (25 ± 1 °C) on a rotary shaker for different times and afterwards centrifuged at 4000× g for 5 min. Aliquots from the supernatant were analyzed immediately, as explained above.
The dye adsorbed by each compost (qe, mg g−1) was calculated as shown in Equation (1):
q e = ( C o C e ) V m
where V is the volume of solution (in liters) and m is the mass of adsorbent (in grams). Equilibrium adsorption curves were drawn by plotting Ce versus qe.
For the desorption study, 0.50 g of each compost was suspended in 10 mL of solutions containing 100 mg L−1 of each dye individually, shaken, and centrifuged, as explained above. After removing the supernatant, the centrifuged residue was weighed to calculate the amount of dye solution still occluded in the adsorbent. Then, they were resuspended in 10 mL of deionized water, shaken again for 24 h, and centrifuged. Dye concentrations were analyzed in the supernatants as explained above, and the results were expressed as the percentage of desorption of the previously absorbed dye.

2.3.2. Effect of Zn and Dissolved Organic Matter

In order to test the effect of Zn and dissolved organic matter (DOM) on dye removal, equilibrium adsorption curves were built following the general procedure as explained above, but with the addition of a solution of ZnSO4 (Panreac, Barcelona, Spain) or humic acids (Sigma-Aldrich, Munich, Germany). The final concentrations in all dye solutions were 5 mg Zn L−1 or 100 mg humic acids L−1, which were selected as typical concentrations found in real wastewaters [18]. Blanks were run in all cases without dye to determine the absorbance of the Zn or DOM solutions alone. All tests were carried out in triplicates.

2.4. Modeling of Adsorption Curves

Equilibrium adsorption was described using the Freundlich and Langmuir isotherms. The Langmuir equation, assuming one monolayer coverage on equal-energy active sites without interaction of the adsorbed solutes, has the following form:
q e = q m K L C e 1 + K L C e
where qm (mg g−1) is the maximum sorption capacity and KL (L mg−1) is the adsorption constant. The expression of the Freundlich isotherm model defines the heterogeneity of the surface as well as the exponential distribution of the active sites and energies:
q e = K F C e 1 / n
where KF (Ln mg1−n g−1) is the Freundlich constant and can indicate uptake capacity, while 1/n (dimensionless) measures favorability of the process.
Modeling of the adsorption curves was performed using nonlinear regression analysis. The adsorption models were fitted to the experimental data using the R statistical software version 3.6.1 [19] and the R Commander [20] and nlstools packages for R (The R Foundation for Statistical Computing, Vienna, Austria) [21], which were also used to calculate the significance and quality of the model fits.

3. Results and Discussion

The effect of contact time on dye adsorption onto the composts is shown in Figure 1. Equilibrium was more rapidly achieved for AB113 than for the other dyes, in only 4–8 h, and also for the compost MSWC with respect to CPB. The results of modeling are included in the Supplementary Materials.
The dye adsorption curves with and without Zn or DOM are shown in Figure 2, while the results of the adsorption modeling are presented in Table 3. Both composts had higher sorption capacity for the basic dye than for the rest of types, with an overall sequence of adsorption BV10 >> AB113 > DB71. This significantly higher affinity of compost towards basic dyes is a common observation in the literature [11]. For the dye BV10, the compost CPB was clearly superior to MSWC, whereas MSWC showed a slightly higher adsorption than CPB for the other two dyes. Overall, the Freundlich model was more adequate to describe the experimental data than the Langmuir model, so maximum adsorption capacities could only be calculated confidently in some cases. In general, these were around 1 mg g−1 for DB71, 2 mg g−1 for AB113, and 40 mg g−1 for BV10.
When the composts are compared with other materials used as sorbents, their overall performances are different depending on the type of dye considered. The maximum sorption capacity for the basic dye BV10 is in the same range as those obtained for more expensive synthetic adsorbents such as graphene oxides and activated carbon [22,23], and natural adsorbents such as zeolite and kaolinite [22], citric peels [24] or other composts [25]. In the case of anionic dyes (direct and acid), both composts presented lower affinities than other biosorbents in the literature, as, for example, peels from Cucumis sativus [26] or potato [27] for AB113 adsorption. For DB71, the values obtained here are also lower than those reported for other biosorbents such as synthetic alumina/silica oxides [28,29], wheat shells [30], or chitosan/SiO2/carbon nanotubes [31], zeolites [32], or active carbon [33]. These results confirm that composts are less competitive as biosorbents for anionic than for cationic dyes, as suggested in a previous work with an exhaustive study of adsorption of different dye classes [15].
With the objective of assessing the potential recycling of the composts, desorption of dyes was also studied. The percentages of desorption for the three dyes were very low, as shown in Figure 3, without differences observed between the two composts for DB71. In turn, desorption was lower from CPB than from MSWC in the case of BV10 and AB113 (values under 10% in most cases, with no desorption of BV10), which indicates a stronger interaction of these two dyes with CPB. Overall, the results of the desorption experiment agree with the relative dye adsorption capacities of the composts as has been observed in the studies of equilibrium adsorption.
Regarding the influence of other pollutants on adsorption, the comparison of the results in the absence and presence of Zn or DOM have shown the absence of any effect, at least at the concentrations tested here. The presence of Zn in the solution did not modify the dye adsorption curves. Some authors have suggested that Zn could have an antagonistic effect on cationic dye removal [34,35] due to electrostatic interaction competition, since Zn and cationic dyes are positively charged in solution. Given that composts have a high affinity for Zn [36], in theory a high competition with the dyes should have been observed. However, this was not the case, as seen in Figure 4, which shows clearly that the presence of dye at different concentrations did not modify Zn adsorption. A likely explanation would be that the adsorption of Zn and dyes takes place by different mechanisms. Dye adsorption by composted materials is a very complex process, and it is likely that the nature of this interaction cannot by explained by a single mechanism. In this sense, previous works where we investigated the influence of pH and salinity on dye removal, as well as the study of FTIR spectra of dye-loaded composts, have allowed important information to be obtained about the mechanisms of interaction of composts with dyes of different classes [14,15]. These have shown that although electrostatic interaction is an important mechanism of adsorption for both cationic and anionic dyes on compost, it is not the main mechanism, and that a relevant contribution of other mechanisms should be taken into account to explain dye adsorption, for example hydrophobic interactions and dipole–dipole forces [14,15]. If electrostatic interaction is not a major mechanism for adsorption, then Zn is not expected to influence much dye adsorption. Thus, in the experiment presented here, probably hydrophobic interactions are more important for dye adsorption, and Zn does not have an effect in this process. The results obtained in the present work provide additional evidence that electrostatic interaction is not the only mechanism in dye adsorption, as discussed in previous works.
The situation in the case of DOM should be different since the mechanisms for adsorption of the humic acids employed is expected to be similar to the organic dyes. From this point of view, a competition effect would have been expected. Some authors, such as Zermane et al. [37], suggest in turn that the addition of humic acids to the solution should favorably influence the adsorption of basic dyes. However, no effect at all was observed here in most cases, only modifications of the shape of the curve for AB113 with MSWC.
In conclusion, the results of this experiment showed that the presence of other pollutants at typical wastewater concentrations did not significantly modify the dye adsorption capacity of the composts. This does not mean that Zn or DOM cannot influence positively or negatively dye adsorption, but rather that the concentrations at which they are typically found in wastewater, such as those tested here, are too low to have an effect. However, the importance of this fact should not be underestimated: this suggests that the results obtained in laboratory tests using water solutions would be similar to those with real wastewaters, and that the performance of composts for dye removal under real conditions would not be affected by the presence of other pollutants.

4. Conclusions

This work evaluated the influence of the presence of common pollutants in textile wastewaters on the performance of compost for dye removal. A pine bark compost and a municipal solid waste compost were tested for the removal of three types of dyes from aqueous solutions—a basic dye (BV10), an acid dye (AB113), and a direct dye (DB71)—with Zn and DOM as the competing pollutants studied. Dye adsorption capacity followed the sequence BV10 > AB113 > DB71, with a slightly higher affinity of municipal solid waste compost for AB113 and DB71, and a much higher adsorption capacity of pine bark compost for BV10. The results of the experiments do not show a significant negative influence of the presence of Zn or DOM with respect to the results in pure dye solutions, suggesting that the presence of these pollutants at typical wastewater concentrations does not modify the adsorption capacity of the composts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijerph20075353/s1, Figure S1: FTIR spectra of the composts. MSWC: municipal solid waste compost; CPB: composted pine bark. Table S1: Kinetic parameters for pseudo-first and pseudo-second order models for dyes sorption rates by composts. CPB: composted pine bark; MSWC: municipal solid waste compost; qe(model): equilibrium sorption value predicted from the model (mg g−1); k1: pseudo-first order model constant (h−1); k2: pseudo-second order model constant (g mg−1 h−1).

Author Contributions

Conceptualization, R.P. and M.T.B.; methodology, M.T.B., R.P. and K.A.-Z.; validation, K.A.-Z. and R.P.; formal analysis, M.T.B., R.P. and K.A.-Z.; investigation, R.P., A.G. and P.G.; resources, M.T.B.; data curation, R.P.; writing—original draft preparation, R.P.; writing—review and editing, R.P., K.A.-Z. and M.T.B.; visualization, R.P.; supervision, M.T.B. and R.P.; project administration, M.T.B. and R.P.; funding acquisition, M.T.B. and R.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. O’Neill, C.; Hawkes, F.R.; Hawkes, D.L.; Lourenco, N.D.; Pinheiro, H.M.; Delée, W. Colour in textile effluents—Sources, measurement, discharge consents and simulation: A review. J. Chem. Technol. Biotechnol. 1999, 74, 1009–1018. [Google Scholar] [CrossRef]
  2. Slokar, Y.M.; Majcen Le Marechal, A. Methods of decoloration of textile wastewaters. Dye. Pigment. 1997, 37, 335–356. [Google Scholar] [CrossRef]
  3. Jain, A.K.; Gupta, V.K.; Bhatnagar, A.; Suhas. Utilization of industrial waste products as adsorbents for the removal of dyes. J. Hazard. Mater. 2003, 101, 31–42. [Google Scholar] [CrossRef] [PubMed]
  4. Crini, G. Non-conventional low-cost adsorbents for dye removal: A review. Bioresour. Technol. 2006, 97, 1061–1085. [Google Scholar] [CrossRef]
  5. Anastopoulos, I.; Kyzas, G.Z. Agricultural peels for dye adsorption: A review of recent literature. J. Mol. Liq. 2014, 200, 381–389. [Google Scholar] [CrossRef]
  6. Salleh, M.A.M.; Mahmoud, D.K.; Karim, W.A.W.A.; Idris, A. Cationic and anionic dye adsorption by agricultural solid wastes: A comprehensive review. Desalination 2019, 5, 91–108. [Google Scholar] [CrossRef]
  7. Zaidi, N.A.H.M.; Lim, L.B.L.; Usman, A.; Kooh, M.R.R. Efficient adsorption of malachite green dye using Artocarpus odoratissimus leaves with artificial neural network modelling. Desalination Water Treat. 2018, 101, 313–324. [Google Scholar] [CrossRef]
  8. McKay, G.; Hadi, M.; Samadi, M.T.; Rahmani, A.R.; Aminabad, M.S.; Nazemi, F. Adsorption of reactive dye from aqueous solutions by compost. Desalination Water Treat. 2011, 28, 164–173. [Google Scholar] [CrossRef]
  9. Anastopoulos, I.; Kyzas, G.Z. Composts as biosorbents for decontamination of various pollutants: A review. Water Air Soil Pollut. 2015, 226, 61–72. [Google Scholar] [CrossRef]
  10. Paradelo, R.; Moldes, A.B.; Barral, M.T. Treatment of red wine vinasses with non-conventional substrates for removing coloured compounds. Water Sci. Technol. 2009, 58, 1585–1592. [Google Scholar] [CrossRef] [Green Version]
  11. Paradelo, R.; Vecino, X.; Moldes, A.B.; Barral, M.T. Potential use of composts and vermicomposts as low-cost adsorbents for dye removal: An overlooked application. Environ. Sci. Pollut. Res. 2019, 26, 21085–21097. [Google Scholar] [CrossRef]
  12. Pushpa, T.; Vijayaraghavan, J.; Jegan, J. Utilization of Effective Microorganisms based water hyacinth compost as biosorbent for the removal of basic dyes. Desalination Water Treat. 2016, 57, 24368–24377. [Google Scholar] [CrossRef]
  13. Dey, M.D.; Das, S.; Kumar, R.; Doley, R.; Bhattacharya, S.S.; Mukhopadhyay, R. Vermiremoval of methylene blue using Eisenia fetida: A potential strategy for bioremediation of synthetic dye-containing effluents. Ecol. Eng. 2017, 106, 200–208. [Google Scholar] [CrossRef]
  14. Paradelo, R.; Al-Zawahreh, K.; Barral, M.T. Utilization of composts for adsorption of methylene blue from aqueous solutions: Kinetics and equilibrium studies. Materials 2020, 13, 2179. [Google Scholar] [CrossRef]
  15. Al-Zawahreh, K.; Barral, M.T.; Al-Degs, Y.; Paradelo, R. Comparison of the sorption capacity of basic, acid, direct and reactive dyes by compost in batch conditions. J. Environ. Manag. 2021, 294, 113005. [Google Scholar] [CrossRef]
  16. Al-Zawahreh, K.; Al-Degs, Y.; Barral, M.T.; Paradelo, R. Optimization of Direct Blue 71 sorption by organic rich-compost following multilevel multifactor experimental design. Arab. J. Chem. 2022, 15, 103468. [Google Scholar] [CrossRef]
  17. Al-Zawahreh, K.; Barral, M.T.; Al-Degs, Y.; Paradelo, R. Competitive removal of textile dyes from solution by pine bark-compost in batch and fixed bed column experiments. Environ. Technol. Innov. 2022, 27, 102421. [Google Scholar] [CrossRef]
  18. Bisschops, I.; Spanjers, H. Literature review on textile wastewater characterization. Environ. Technol. 2003, 24, 1399–1411. [Google Scholar] [CrossRef]
  19. R Core Team. R: A Language and Environment for Statistical Computing; R Foundation for Statistical Computing: Vienna, Austria, 2018. [Google Scholar]
  20. Fox, J.; Bouchet-Valat, M. Rcmdr: R Commander, R package version 2.8-0; R Foundation for Statistical Computing: Vienna, Austria, 2022. [Google Scholar]
  21. Baty, F.; Ritz, C.; Charles, S.; Brutsche, M.; Flandrois, J.-P.; Delignette-Muller, M.-L. A toolbox for nonlinear regression in R: The package nlstools. J. Stat. Softw. 2015, 66, 1–21. [Google Scholar] [CrossRef] [Green Version]
  22. Yu, Y.; Murthy, B.; Shapter, J.; Constantopoulos, K.; Voelcker, N.; Ellis, A. Benzene carboxylic acid derivatized graphene oxide nanosheets on natural zeolites as effective adsorbents for cationic dye removal. J. Hazard. Mater. 2013, 260, 330–338. [Google Scholar] [CrossRef]
  23. Filipkowska, U.; Jóźwiak, T.; Szymczyk, P.; Kuczajowska-Zadrożna, M. The use of active carbon immobilised on chitosan beads for RB5 and BV10 dye removal from aqueous solutions. Prog. Chem. Appl. Chitin Its Deriv. 2017, 22, 14–26. [Google Scholar] [CrossRef] [Green Version]
  24. Jóźwiak, T.; Filipkowska, U.; Zajko, P. Use of citrus fruit peels (grapefruit, mandarin, orange, and lemon) as sorbents for the removal of Basic Violet 10 and Basic Red 46 from aqueous solutions. Desalination Water Treat. 2019, 163, 385–397. [Google Scholar] [CrossRef]
  25. Jóźwiak, T.; Filipkowska, U.; Rodziewicz, J.; Mielcarek, A.; Owczarkowska, D. Application of compost as a cheap sorbent for dyes removal from aqueous solutions. Rocz. Ochr. Srodowiska 2013, 15, 2398–2411. [Google Scholar]
  26. Lee, L.Y.; Gan, S.; Tan, M.S.Y.; Lim, S.S.; Lee, X.J.; Lam, Y.F. Effective removal of Acid Blue 113 dye using overripe Cucumis sativus peel as an eco-friendly biosorbent from agricultural residue. J. Clean. Prod. 2016, 113, 194–203. [Google Scholar] [CrossRef]
  27. Hoseinzadeh, E.; Samarghandi, M.R.; McKay, G.; Rahimi, N.; Jafari, J. Removal of acid dyes from aqueous solution using potato peel waste biomass: A kinetic and equilibrium study. Desalination Water Treat. 2014, 52, 4999–5006. [Google Scholar] [CrossRef]
  28. Wawrzkiewicz, M.; Nowacka, M.; Klapiszewski, L.; Hubicki, Z. Treatment of wastewaters containing acid, reactive and direct dyes using aminosilane functionalized silica. Open Chem. 2015, 13, 82–95. [Google Scholar] [CrossRef]
  29. Wawrzkiewicz, M.; Wiśniewska, M.; Gun’koc, V.M.; Zarkoc, V.I. Adsorptive removal of acid, reactive and direct dyes from aqueous solutions and wastewater using mixed silica–alumina oxide. Powder Technol. 2015, 278, 306–315. [Google Scholar] [CrossRef]
  30. Buliut, Y.; Gözübenli, N.; Aydin, H. Equilibrium and kinetics studies for adsorption of direct blue 71 from aqueous solution by wheat shells. J. Hazard. Mater. 2007, 144, 300–306. [Google Scholar] [CrossRef]
  31. Abasi, M. Synthesis and characterization of magnetic nanocomposite of chitosan/SiO2/carbon nanotubes and its application for dyes removal. J. Clean. Prod. 2017, 145, 105–113. [Google Scholar] [CrossRef]
  32. Moradi, R. Adsorption of Azo Dye Direct Blue 71 from Aqueous Solution Using NiFe2O4/CP Nanoparticles: Kinetic and Isotherm Model. Russ. J. Phys. Chem. A 2018, 92, 2781–2789. [Google Scholar] [CrossRef]
  33. Baseri, J.R.; Palanisamy, P.N.; Sivakumar, P. Comparative studies of the adsorption of direct dye on activated carbon and conducting polymer composite. J. Chem. 2012, 9, 1122–1134. [Google Scholar] [CrossRef] [Green Version]
  34. Visa, M.; Bogatu, C.; Duta, A. Simultaneous adsorption of dyes and heavy metals from multicomponent solutions using fly ash. Appl. Surf. Sci. 2010, 256, 5486–5491. [Google Scholar] [CrossRef]
  35. Hernández-Montoya, V.; Pérez-Cruz, M.A.; Mendoza-Castillo, D.I.; Moreno-Virgen, M.R.; Bonilla-Petriciolet, A. Competitive adsorption of dyes and heavy metals on zeolitic structures. J. Environ. Manag. 2013, 116, 213–221. [Google Scholar] [CrossRef]
  36. Paradelo, R.; Barral, M.T. Evaluation of the potential capacity as metal biosorbents of two MSW composts with different Cu, Pb and Zn content. Bioresour. Technol. 2012, 104, 810–813. [Google Scholar] [CrossRef]
  37. Zermane, F.; Cheknane, B.; Basly, J.P.; Bouras, O.; Baudu, M. Influence of humic acids on the adsorption of Basic Yellow 28 dye onto an iron organo–inorgano pillared clay and two Hydrous Ferric Oxides. J. Colloid Interface Sci. 2013, 395, 212–216. [Google Scholar] [CrossRef]
Figure 1. Kinetics of dye adsorption by the composts. Initial dye concentration: 200 mg L−1; solid:liquid ratio 1:20MSWC: municipal solid waste compost; CPB: composted pine bark; BV10: Basic Violet 10; AB113: Acid Blue 113; DB71: Direct Blue 71.
Figure 1. Kinetics of dye adsorption by the composts. Initial dye concentration: 200 mg L−1; solid:liquid ratio 1:20MSWC: municipal solid waste compost; CPB: composted pine bark; BV10: Basic Violet 10; AB113: Acid Blue 113; DB71: Direct Blue 71.
Ijerph 20 05353 g001
Figure 2. Sorption isotherms of dyes with and without Zn or dissolved organic matter (DOM). MSWC: municipal solid waste compost; CPB: composted pine bark.
Figure 2. Sorption isotherms of dyes with and without Zn or dissolved organic matter (DOM). MSWC: municipal solid waste compost; CPB: composted pine bark.
Ijerph 20 05353 g002
Figure 3. Desorption of the previously adsorbed dye from compost (mean ± standard deviation, n = 3). CPB: composted pine bark; MSWC: municipal solid waste compost; BV10: Basic Violet 10; AB113: Acid Blue 113; DB71: Direct Blue 71.
Figure 3. Desorption of the previously adsorbed dye from compost (mean ± standard deviation, n = 3). CPB: composted pine bark; MSWC: municipal solid waste compost; BV10: Basic Violet 10; AB113: Acid Blue 113; DB71: Direct Blue 71.
Ijerph 20 05353 g003
Figure 4. Adsorption of Zn by the composts as a function of initial dye concentrations in the solution (mean ± standard deviation, n = 3). CPB: composted pine bark; MSWC: municipal solid waste compost; BV10: Basic Violet 10; AB113: Acid Blue 113; DB71: Direct Blue 71.
Figure 4. Adsorption of Zn by the composts as a function of initial dye concentrations in the solution (mean ± standard deviation, n = 3). CPB: composted pine bark; MSWC: municipal solid waste compost; BV10: Basic Violet 10; AB113: Acid Blue 113; DB71: Direct Blue 71.
Ijerph 20 05353 g004
Table 1. Main properties of the three dyes.
Table 1. Main properties of the three dyes.
DyeMolecular
Formula
StructureMolecular Weight (g mol−1)Water Solubility (g L−1)
Basic Violet 10C28H31ClN2O3Ijerph 20 05353 i001478.58–15
Acid Blue 113C32H21N5Na2O6S2Ijerph 20 05353 i002681.740
Direct Blue 71C40H23N7Na4O13S4Ijerph 20 05353 i0031029.910
Table 2. Physicochemical parameters of the composts. CPB: composted pine bark; MSWC: municipal solid waste compost; pHw: pH in water; pHPZC: pH at point of zero charge; EC: electrical conductivity; OM: organic matter; AEC: anionic exchange capacity; CEC: cation exchange capacity; SSA: specific surface area.
Table 2. Physicochemical parameters of the composts. CPB: composted pine bark; MSWC: municipal solid waste compost; pHw: pH in water; pHPZC: pH at point of zero charge; EC: electrical conductivity; OM: organic matter; AEC: anionic exchange capacity; CEC: cation exchange capacity; SSA: specific surface area.
CPBMSWC
pHW5.38.5
pHPZC4.48.1
EC, dS m−10.47.0
OM, %91.441.8
C, %53.124.3
N, %0.281.6
C/N ratio19416
CEC, cmol(+) kg−125.418.8
AEC, cmol(−) kg−14.63.6
SSA, m2 g−122.45.8
Table 3. Isotherm parameters for dye sorption under different conditions; BV10: Basic Violet 10; AB113: Acid Blue 113; DB71: Direct Blue 71; CPB: composted pine bark; MSWC: municipal solid waste compost; DOM: dissolved organic matter; qm: Langmuir maximum capacity (mg g−1); KL: Langmuir constant (L mg−1); KF: Freundlich constant (Ln mg1−n g−1); 1/n: Freundlich coefficient. Significance of the parameter estimations is indicated as follows: * significant at a p-value of 0.05; ** significant at a p-value of 0.01; *** significant at a p-value of 0.001.
Table 3. Isotherm parameters for dye sorption under different conditions; BV10: Basic Violet 10; AB113: Acid Blue 113; DB71: Direct Blue 71; CPB: composted pine bark; MSWC: municipal solid waste compost; DOM: dissolved organic matter; qm: Langmuir maximum capacity (mg g−1); KL: Langmuir constant (L mg−1); KF: Freundlich constant (Ln mg1−n g−1); 1/n: Freundlich coefficient. Significance of the parameter estimations is indicated as follows: * significant at a p-value of 0.05; ** significant at a p-value of 0.01; *** significant at a p-value of 0.001.
DyeCompostFactorLangmuir Freundlich
KLpqmpR2KFp1/npR2
BV10CPB-0.11 ± 0.050.0841.7 ± 10.2 *0.010.9855.2 ± 0.9 **0.0040.63 ± 0.08 **0.0020.979
Zn0.16 ± 0.090.1439.3 ± 12.1 *0.030.9816.2 ± 1.1 **0.0050.62 ± 0.11 **0.0040.980
DOM-----1.3 ± 0.2 *0.011.20 ± 0.08 ***0.00060.935
MSWC-0.04 ± 0.030.256.8 ± 3.90.160.9920.31 ± 0.04 **0.0010.78 ± 0.06 ***0.00020.993
Zn0.10 ± 0.040.063.4 ± 0.8 *0.010.9850.33 ± 0.04 **0.0010.72 ± 0.06 ***0.00030.988
DOM0.03 ± 0.030.4610.4 ± 10.90.390.9910.30 ± 0.06 **0.0060.83 ± 0.10 **0.0010.990
AB113CPB-0.16 ± 0.03 **0.0073.0 ± 0.3 ***0.00040.9930.45 ± 0.06 **0.0020.61 ± 0.06 ***0.00070.981
Zn0.38 ± 0.150.072.2 ± 0.3 **0.0010.9900.70 ± 0.10 **0.0020.39 ± 0.07 **0.0040.988
DOM-----0.11 ± 0.02 **0.0021.01 ± 0.05 ***0.000050.997
MSWC-0.51 ± 0.200.072.4 ± 0.4 **0.0030.9770.8 ± 0.1 **0.0020.47 ± 0.09 **0.0070.972
Zn4.6 ± 8.50.622.3 ± 0.90.060.8591.8 ± 0.4 *0.010.18 ± 0.270.540.859
DOM11.0 ± 10.30.621.8 ± 0.3 **0.0040.9101.5 ± 0.2 **0.0020.21 ± 0.120.140.937
DB71CPB-0.02 ± 0.030.430.5 ± 0.20.120.7650.04 ± 0.030.260.49 ± 0.190.060.772
Zn0.007 ± 0.0060.291.3 ± 0.70.160.9640.02 ± 0.010.120.78 ± 0.13 **0.0040.962
DOM0.034 ± 0.008 *0.010.32 ± 0.03 ***0.00040.9880.02 ± 0.010.070.53 ± 0.10 **0.0070.945
MSWC-0.05 ± 0.020.091.3 ± 0.2 **0.0050.9640.14 ± 0.04 *0.020.50 ± 0.07 **0.0030.965
Zn0.09 ± 0.02 *0.021.4 ± 0.1 ***0.00050.9930.21 ± 0.03 **0.0010.46 ± 0.04 ***0.00020.994
DOM0.06 ± 0.02 *0.041.5 ± 0.2 **0.0020.9910.17 ± 0.03 **0.0030.50 ± 0.05 ***0.00040.993
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

Paradelo, R.; García, P.; González, A.; Al-Zawahreh, K.; Barral, M.T. Influence of Zinc and Humic Acids on Dye Adsorption from Water by Two Composts. Int. J. Environ. Res. Public Health 2023, 20, 5353. https://doi.org/10.3390/ijerph20075353

AMA Style

Paradelo R, García P, González A, Al-Zawahreh K, Barral MT. Influence of Zinc and Humic Acids on Dye Adsorption from Water by Two Composts. International Journal of Environmental Research and Public Health. 2023; 20(7):5353. https://doi.org/10.3390/ijerph20075353

Chicago/Turabian Style

Paradelo, Remigio, Paula García, Alba González, Khaled Al-Zawahreh, and Maria Teresa Barral. 2023. "Influence of Zinc and Humic Acids on Dye Adsorption from Water by Two Composts" International Journal of Environmental Research and Public Health 20, no. 7: 5353. https://doi.org/10.3390/ijerph20075353

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop