1. Trang chủ
  2. » Giáo án - Bài giảng

Tailoring metal-impregnated biochars for selective removal of natural organic matter and dissolved phosphorus from the aqueous phase

9 1 0

Đang tải... (xem toàn văn)

THÔNG TIN TÀI LIỆU

Nội dung

This study aimed to investigate how the production process of metal impregnated biochars (MIBs) affects their selectivity in the simultaneous adsorption of organic matter and dissolved phosphorus from the aqueous phase.

Microporous and Mesoporous Materials 328 (2021) 111499 Contents lists available at ScienceDirect Microporous and Mesoporous Materials journal homepage: www.elsevier.com/locate/micromeso Tailoring metal-impregnated biochars for selective removal of natural organic matter and dissolved phosphorus from the aqueous phase Oleksii Tomin a, *, Riku Vahala a, Maryam Roza Yazdani b, ** a b Department of Built Environment, School of Engineering, Aalto University, P.O Box 15200, FI-00076, Aalto, Finland Department of Mechanical Engineering, School of Engineering, Aalto University, P.O Box 14400, FI-00076, Aalto, Finland A R T I C L E I N F O A B S T R A C T Keywords: Adsorption Biochar Metal impregnation Natural organic matter Selectivity This study aimed to investigate how the production process of metal impregnated biochars (MIBs) affects their selectivity in the simultaneous adsorption of organic matter and dissolved phosphorus from the aqueous phase MIBs were produced via a two-step pyrolysis procedure including impregnation of metal oxides in the structure of the softwood-derived biochars, resulting in copper-impregnated biochar (Cu-MIB) and iron-impregnated biochar (Fe-MIB) The tailoring process was conducted by optimization of pyrolysis temperature during the biochars production stage The MIBs were characterized via advanced characterization analyses to acquire structural, elemental, and morphological properties of the adsorbent The surface area of MIB (99 m2/g and 92 m2/g for Cu-MIB and Fe-MIB respectively) decreased compared to pristine biochar (571 m2/g), indicating a successful impregnation of metal oxide particles within the porous carbon structure The effect of operational parameters on adsorption as well as selectivity tests were examined in the batch mode The optimum doses for NOM removal were g/l for Fe-MIB (96%) and 0.5 g/l for Cu-MIB (87%) For phosphorus removal, optimum doses were g/l for Fe-MIB (95%) and g/l for Cu-MIB (93%) The lower pH values favored adsorption for both MIBs In the binary solution of NOM and phosphorus, the NOM was selectively adsorbed by the Cu-MIB, whereas phosphorus was selectively removed by the Fe-MIB The results provide a deeper understanding of the tailoring process of biochars for producing new biochars as selective adsorbents for specific target pollutants Introduction Natural organic matter (NOM) is a complex matrix of organic com­ pounds with a wide range of molecular masses present in natural water sources NOM raises aesthetic issues including unpleasant taste, odor, and color of the water It significantly influences drinking water pro­ duction by, for example, contributing to the membrane fouling, competing with the removal of other pollutants, increasing process costs, and causing microbial regrowth in the distribution system [1–3] Phosphorus acts as a fundamental yet finite element for the growth of living organisms and many industries Nevertheless, a vast industrial utilization has led to high concentrations of phosphorus in the discharge waters to the environment The release of large phosphorus content into the natural water bodies causes major environmental concerns including eutrophication Phosphorus and dissolved organic matters also often coexist in wastewater [4] The simultaneous presence of phosphorus and organic matter in such water can disturb the removal efficiency of the treatment plant due to competing effects, for example during the chemical treatment Therefore, exploring the simultaneous removal of phosphorus and organic matter is of high importance for the optimiza­ tion of the treatment process Coagulation and adsorption are the commonly used methods to remove NOM and phosphorus [5,6] However, often conventional coagulation fails to reach high removal percentages of NOM [5] Thus, the adsorption process is attractive, as a tertiary treatment step, due to enabling efficient removal of various pollutants and the possibility of adsorbent regeneration/reuse When adsorption is applied in the water treatment process, the activated carbon (AC) is usually used as an adsorbent The major reasons which limit AC application in NOM removal are the high cost and high environmental impact of AC during production and transportation Therefore, there is an urgent need to develop such efficient products from locally-globally available and eco-friendly sustainable resources to lower the carbon footprint of AC implementation in water treatment utilities * Corresponding author ** Corresponding author E-mail addresses: oleksii.tomin@aalto.fi (O Tomin), roza.yazdani@aalto.fi (M.R Yazdani) https://doi.org/10.1016/j.micromeso.2021.111499 Received September 2021; Received in revised form October 2021; Accepted 11 October 2021 Available online 16 October 2021 1387-1811/© 2021 The Authors Published by Elsevier Inc This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/) O Tomin et al Microporous and Mesoporous Materials 328 (2021) 111499 A promising adsorbent that can substitute imported AC is biochar, which is produced through thermal conversion of biomass under limited oxygen conditions Unlike AC, biochars are produced only from biobased resources The production of biochars from locally available biomass such as forestry products, algae, and agricultural wastes in­ creases the accessibility of this bio-product worldwide Recent research in biochar development and application for environmental remedies is a call for such environmental solutions with more advanced features for real-world applicability [7,8] As a reusable material, biochar can be used for resource recovery or act as a fertilizer after water treatment use [9,10] Biochar properties can be tailored via modification and changing production conditions [11] The tailoring process of the biochars is a search for the most suitable of production conditions, which would allow producing the most efficient material, depending on the purpose [12,13] We previously developed highly mesoporous biochars from pinecone-forestry byproduct [1,14] The research showed that process condition and modification play a key role in tailoring the porous structure and functionality of biochars for the efficient removal of NOM from lake water With a proper tailoring process, targeted for specific pollutants, we can maximize biochar’s performance with better results compared to that of commercial AC [14] Biochars showed a very good performance on the simultaneous adsorption of multiple pollutants, e.g metals [15] On the other hand, a major challenge associated with traditional AC and biochars is a lack of selectivity [16] The inability of biochars to selective adsorption causes a rapid loss of adsorption ca­ pacity for the target pollutant due to the adsorption of competing compounds This challenge can be addressed by tailoring selective bio­ chars At the moment selective carbon adsorbents show a knowledge gap and lack of sufficient understanding as it is not well researched Our study attempts to decrease this gap As such, we aimed to tailor selective biochars from forest-based biomass for the simultaneous removal of NOM and phosphorus from lake water Spruce softwood was selected due to its large availability as a local wood species and highly porous structure that benefits the devel­ opment of a porous adsorbent The tailoring process included the impregnation of transition metals such as copper (Cu) and iron (Fe) into the porous carbonous structure, which enables a selective functionality for NOM or phosphorus The selection of these metals for tailoring was done due to the comparison of a well-studied metal impregnated adsorbent (Fe-MIB) and a more novel adsorbent (Cu-MIB), that, as far as we know, has not been studied before for water treatment purposes To the best of our knowledge, while recent literature reports plenty of in­ formation about the iron-impregnated biochar [17–21], there is a scarcity of knowledge on developing selective biochars through chem­ ical activation utilizing copper salt as an activator Besides that, based on the previous research it was decided to pro­ duce biochars via two-step pyrolysis as the most efficient method of production [1] Pyrolysis temperature plays a key role in the final properties of the developed biochars [22] However, this temperature has been rarely optimized for the two-step pyrolysis method in the literature Thus, this research aims (i) to perform chemical modification and compare common Fe-based activator and newer Cu-based activator (ii) to find the most suitable operational parameters for the production of selective adsorbents and (iii) to test the tailored biochars in selective removal of target pollutants from the lake water samples, contaminated with phosphorus and organic matter Table The naming of MIBs with different activators at three different temperatures of first and second step pyrolysis Activator 1st step T, ◦ C 2nd step T, ◦ C 600 700 800 FeCl3 200 250 300 200 250 300 F11 F21 F31 C11 C21 C31 F12 F22 F32 C12 C22 C32 F13 F23 F33 C13 C33 C33 CuCl2 the preparation of synthetic water solution were KH2PO4 (Merck) and humic acid sodium salt powder (Alfa Aeser) Lake water samples were ¨ija ¨nne in Asikkala, Finland To determine the collected from Lake Pa chemical oxygen demand (COD) calibration curve, the used chemicals included sulfuric acid, potassium permanganate, potassium iodide, starch indicator, and sodium thiosulfate 2.2 Tailoring of biochar Eighteen different types of metal-impregnated biochars (MIB) were tailored via two-step pyrolysis with the chemical activation process, reported in our previous research [1,14], but with different activators and pyrolysis temperature Firstly, sawdust was pyrolyzed at three different temperatures: 200, 250, and 300 ◦ C for 15 under nitrogen atmosphere in Naber N60/HR furnace Then, biochars were modified with CuCl2 and FeCl3 The first-step pyrolyzed biochars were mixed with chemical solutions in a 1/2 ratio of biochar/activator for h in room temperature conditions After that, they were dried for 24 h under 105 ◦ C and finally pyrolyzed again under the nitrogen atmosphere at high temperatures 600, 700, and 800 ◦ C resulting in MIBs Additionally, two samples of reference biochars (RBC) were produced for comparison with tailored biochars The lowest observed production temperatures were chosen for comparison as commonly slow pyrolysis temperatures rarely reach 600 ◦ C [22] For the low pyrolysis step temperature 200 ◦ C (R200) was used and for the high pyrolysis step 600 ◦ C (R600) without the chemical activation After preparation, all products were rinsed with 0.1 M HCl and reverse osmosis water (RO-water) until a neutral pH was obtained and then dried at 105 ◦ C overnight All samples were stored at room temperature for further characterization and adsorption steps Generally, MIBs are divided into iron-impregnated biochars (Fe-MIB) and copper-impregnated biochars (Cu-MIB) The naming of the MIB samples is tabulated in Table The letters F for iron impregnation and C for copper impregnation are connected with numbers that refer to three different temperatures in the first step of pyrolysis, and three different temperatures in the second step of pyrolysis For instance, Fe-MIB pro­ duced at 300 ◦ C first step pyrolysis and 700 ◦ C second step pyrolysis is named as F32 and for Cu-MIB at the same temperatures is named as C32 2.3 Characterization of biochar To study the functional properties of the MIBs such as morphology, composition, and porosity, the samples were characterized via scanning electron microscopy (SEM) accompanied with energy dispersive X-ray (EDX) analysis, Brunauer, Emmett and Teller (BET) specific surface area/porosity, Fourier-transform infrared spectroscopy (FTIR) and X-ray diffraction (XRD) The SEM-EDX analysis was performed on the JEOL JSM-7500FA analytical field Emission scanning electron microscope using 10 μA probe current and 10 kV acceleration voltage, to explore the surface morphology and elemental content of the samples To prepare the sample, MIB was fixed on a metal stub with carbon tape The coating was not performed The FTIR analysis was performed on PerkinElmer Spectrum Two FT- Experimental 2.1 Materials Spruce sawdust was ordered from the Swedish University of Agri­ cultural Sciences, Umea, Sweden, where it was prepared and sieved through a mm sieve The sawdust was stored in closed packages in a cold room before pyrolysis Modification chemicals included FeCl3*6H2O (Merck) and CuCl2*2H2O (Sigma Aldrich) The reagents for O Tomin et al Microporous and Mesoporous Materials 328 (2021) 111499 Fig SEM images of a) RBC R200, b) RBC R600, c) Fe-MIB (sample F13), d) Cu-MIB (sample C11); and EDX analyses of e) Fe-MIB (sample F13), f) Cu-MIB (sample C11) IR Spectrometer The spectra were recorded at room temperature within the range of 500–4000 cm− under four repetitious scans with cm− resolution The XRD measurement was performed on a Rigaku SmartLab X-ray diffractometer to determine the crystalline structure of the MIBs The collected patterns were analyzed, and the peaks were identified using Malvern Panalytical HighScore Plus software Before BET measurement samples were dried overnight and were put in Micromeritics FlowPrep 060 sample preparation system for degas­ ification under 200 ◦ C for h with flowing N2 gas The BET measure­ ment was done in the Micromeritics TriStar II 3020 automated gas adsorption analyzer The pH drift method was employed to determine the point of zero charge (pHPZC) of the MIBs [23] The prepared series of NaCl solutions (0.1 M) pH were adjusted using NaOH (0.1 M) and HCl (0.1 M) within the range of 2–10 A known amount of the MIBs was added to the so­ lutions and allowed to equilibrate for 24 h The final pH of solutions was then measured and compared with the initial pH values The pHPZC was noted as the point where the final pH and initial pH were equal parts: (i) separate adsorption of NOM, (ii) separate adsorption of phos­ phorus, and (iii) simultaneous adsorption of NOM and phosphorus The collected lake water had a very low concentration of NOM (CODMn = 6.5 mg/l) and phosphorus (less than 0.005 mg/l) (see Table S1 in the Supplementary Information) Thus, to model the pollu­ tion, the solutions for NOM adsorption and simultaneous adsorption were prepared by adding a certain amount of humic acid sodium salt to the lake water A stock solution of phosphorus with 1000 mg/l con­ centration was prepared by weighing an accurate amount of potassium dihydrogen phosphate (KH2PO4) and dissolving it in RO-water Different dilutions were prepared daily before each phosphorus adsorption set During the simultaneous adsorption test, to keep a fairly constant amount of NOM in the solution, humic acid sodium salt powder was added to the lake water and then different concentrations of KH2PO4 were added to the solution A certain amount of the MIB was used for the adsorption batch ex­ periments in a 50 ml volume of the model solution Mixing was con­ ducted on a shaker at 180 rpm and room temperature for h until the balance was achieved according to our previous research [1] After the h contact time, the solutions were filtered through 0.45 μm syringe filters for further measurement The removal efficiency was tested with different MIBs in the solution of a constant amount of humic acid sodium salt with 5, 10, and 20 mg/l of phosphorus The adsorbent dosage was optimized within the 0.1–2 g/l range The effect of pH was investigated by adjusting the solution pH at values 2, 4, 6, and using HCl and NaOH 2.4 Batch experiments 2.4.1 Adsorption process Adsorption batch experiments were conducted to study the target pollutant removal by the MIBs The experiments were divided into three O Tomin et al Microporous and Mesoporous Materials 328 (2021) 111499 Table EDS thin-film standardless quantitative analysis (Oxide) of the MIBs Sample F13 Sample C11 Element Mass% Compound Mass% Element Mass% Compound Mass% C O Fe 47.05 11.79 41.16 C 47.05 37.47 52.96 Total 100 37.47 11.84 47.03 3.66 100 C FeO C O Cu Cl CuO Cl 58.87 3.66 100 100 Fig a) FTIR spectra of the RBC (R200), Fe-MIB (sample F13), and Cu-MIB (sample C11); XRD patterns of b) Cu-MIB (C11) and c) Fe-MIB (F13) The batch experiments were conducted in two replicates calculated via the equation: 2.4.2 Analytical measurements The concentration of NOM was estimated via UV absorbance mea­ surement at 254 nm wavelength, using a UV-VIS spectrophotometer (Shimadzu UV-1201) The samples were filtered before the measure­ ment through 0.45 μm syringe filters The absorbance was converted to concentration using the calibration curve of chemical oxygen demand (CODMn) To prepare the calibration curve, a known amount of humic acid sodium salt powder was added to the RO-water Different dilution samples were acidified with M H₂SO₄ Then, KMnO₄ was added to the samples with further boiling for 20 After the oxidizing matter in the samples reduced part of the permanganate, the unreduced portion of permanganate was measured by the iodometric titration and collected data was used for CODMn calculation The phosphorus concentration in the filtrate after adsorption was measured with Discrete analyzer Skalar BlueVision, using method PO4low: 5–500 μg/l P and method PO4high: 0.5–5 mg/l P, in compliance with ISO 15923-1 The percent of pollutant removal efficiency from the solution was % removal = ((C0-C1)/C0)*100 (1) where C0 is the initial concentration of the solution, C1 is the final concentration after adsorption Results and discussion 3.1 Material characterization 3.1.1 Morphology and composition The surface morphology of biochars after the first step of pyrolysis R200 (Fig 1a), the second step of pyrolysis R600 (Fig 1b), and MIBs (Fig 1c and d) were studied via scanning electron microscopy Comparing Fig 1a and b, it is seen that biochar produced at higher temperatures has a more diverse and structured surface structure than the one prepared at low temperatures The surface morphology of the MIBs, showed an enhanced heterogeneous and porous structure with crystalline particles of ferric oxide (Fig 1c) and copper oxide (Fig 1d) densely covering the surface of the pores These crystals prove the suc­ cessive impregnation of metals on the surface of biochar Successive O Tomin et al Microporous and Mesoporous Materials 328 (2021) 111499 fabrication and phenolic-aromatic structures were cracked to leave carbon solids [25] Similar results have been reported on pinewood biochar [24] Table BET surface area, pore size, and pore volume parameters for reference and tailored biochars Sample Surface area, m2/g Pore volume cm3/g Pore size, nm R600 F13 C11 571.2 98.9 92.4 0.38 0.31 0.09 2.7 12.4 4.1 3.1.3 XRD crystallinity To investigate the phase transformation of metal salts during the activation and the formation of metal oxide crystals, the XRD mea­ surements were performed on the MIBs Fig 2b and c shows the XRD patterns of Cu-MIB and Fe-MIB respectively In Fig 2b peaks match with patterns of Cu2O located at 36.2◦ , 42.2◦ 2ϴ, CuCl at 28.4◦ , 47.3◦ , 56.1◦ 2ϴ and Cu at 43.2◦ , 50.3◦ 2ϴ As clearly seen in Fig 2c, strong diffraction peaks are located at 26.3◦ , 43.2◦ , 44.6◦ 2ϴ corresponding to carbon, and 30.3◦ , 35.6◦ , 57.2◦ 2ϴ characteristic peaks matching well with the diffraction patterns of Fe3O4 and/or Fe2O3 The detailed peak analyses obtained from HighScore Plus software can be found in Figures S1, S2, and Tables S2, S3 in the Supplementary Information impregnation of metal oxides also was proven by the EDS analysis, presented in Fig 1e and f, and in Table From the EDS analysis, it can be seen that Cl is not present in the sample, which was impregnated with FeCl3 so iron completely transformed in oxide form, but for CuCl2 impregnated sample it has mainly transformed to CuO (59%) However, there is still some small fraction of Cl (3.6%) untransformed in the product 3.1.2 Functional groups The FT-IR spectra of the RBC (R200) and MIBs showed several sig­ nificant bands, illustrated in Fig 2a The bands at 3500–3300 cm− correspond to O–H stretching, at 2980 cm− and 2880 cm− indicate asymmetric and symmetric C–H, at 2150 cm− show C≡C, at 1640 cm− – C stretching, at 1260 cm− are for C–O stretching, and are related to C– − at 1020 cm indicate C–H out-of-plane bending (e.g aromatic structure of lignin) [24,25] The peak at 620 cm− verifies EDS results by showing the C–Cl band for Cu-MIB Additionally, the numerous peaks in the range 1720-1260 cm− for different aromatic compounds, which are reduced or disappeared after the activation process and thermal treat­ ment, confirming the gasification and conversion to the graphitic structure The disappearance of O–H stretching vibration bands for both chars suggests the oxygen in the initial materials was removed during 3.1.4 BET surface area and porosity Table compiles the BET surface area, pore volume, and pore size of the RBC (R600) and MIBs The RBC showed the highest surface area 571 m2/g, while Fe-MIB and Cu-MIB indicated 99 and 92 m2/g surface area values respectively The decreased surface area of the MIBs confirms a successful impregnation process filling up the pores present in the car­ bon structure Surface area and high porosity play a key role in the metal impregnation process The porous structure of the biochar acts as a host for the metal oxide particles which further contribute to complexation with specific pollutants A decreased surface area after iron impregna­ tion was also reported previously on wood-derived biochars [19,26,27] Therefore, the surface area around 100 m2/g is acceptable compared to many noncarbon low-cost adsorbents, such as montmorillonite [28] or bentonite [29] However, surface area might not play a key role Fig Phosphorus (P) adsorption removal versus the concentration (dosage g/l, contact time h, no pH adjustment) a) with Fe-MIB, b) with Cu-MIB; Response surface of phosphorus removal (initial concentration 10 mg/l, T1 – first step pyrolysis temperature, T2 – second step pyrolysis temperature) c) with Fe-MIB, d) with Cu-MIB O Tomin et al Microporous and Mesoporous Materials 328 (2021) 111499 Fig a) Phosphorus adsorption related to pH (initial phosphorus concentration mg/l, adsorbent dose g/l) (lines) and MIBs point of zero charge (markers); Removal percentage versus the adsorbent dosage b) for NOM (initial COD = 14 mg/l) and c) for phosphorus adsorption (initial P concentration = mg/l) regarding the adsorption of the target pollutants when compared to the role of active functional sites in the adsorption [17,30] Analogously to the surface area, pore volume decreased from 0.38 cm3/g to 0.31 and 0.09 cm3/g for Fe-MIB and Cu-MIB respectively Moreover, we can tell that copper crystals occupy much more volume than ferric ones The pore size, on the other hand, increased from 2.7 nm to 12.4 and 4.1 nm for Fe-MIB and Cu-MIB respectively Even though all studied biochars are considered mesoporous adsorbents (2–50 nm), the noticeable dif­ ference in pore size tells us that metal particles change the pore morphology of pristine biochar closer to the macro scale The graphs of pore size distribution can be found in Figures S3-S5 in the Supplemen­ tary Information explained in a way that during the second step of pyrolysis, the tem­ perature around 600 ◦ C is suitable for lignin conversion [32], and impregnated copper salt under nitrogen gas atmosphere can bind with lignin derivatives, consequently producing complexes on the surface of the biochar With the increase of pyrolysis temperature, these complexes start to disintegrate, which causes a reduction in the number of func­ tional groups on the biochar surface Thus, the optimal parameters for MIB preparation were determined as 200 ◦ C first step pyrolysis, 800 ◦ C second step pyrolysis for Fe-MIB, and 200 ◦ C first step pyrolysis, 600 ◦ C second step pyrolysis for CuMIB Therefore, to save time and resources, these two best-performing compositions were selected to produce biochars for further character­ ization and analysis The chosen MIBs F13 and C11, with approximately similar adsorption ability, were characterized via numerous character­ ization methods reported in the previous section 3.2 Adsorption studies 3.2.1 Determination of best-performing MIB All MIB samples showed high phosphorus removal efficiency, which is seen in Fig 3a and b Sample F13, Fe-MIB in Table 1, showed the highest phosphorus removal (92% removal at lower concentration and 74% removal at higher concentration respectively) Cu-MIB C11 removed 100% of phosphorus at low concentration and 64% at high concentration For the identification of optimal parameters for adsorption, the response surface modeling was performed based on the obtained adsorption data, which is illustrated in Fig 3c and d It is seen that activation with different metals affects adsorption in a very different way at different temperatures For Fe-MIB, the best optimal temperature appears low temperature for the first step pyrolysis and high tempera­ ture for the second step pyrolysis The reason for that can be an active carbonyl and carboxyl functional group formation during the decom­ position of wood extractives, lignin, and cellulose, which is left un­ touched after the first step of pyrolysis [31] The Cu-MIB shows the best performance at low temperatures during the first step and low temper­ atures during the second step of pyrolysis, while the adsorption ability decreases with the temperature increase Such phenomenon can be 3.2.2 Effect of pH The effect of pH on the adsorption of phosphorus was tested within the 2–8 pH range The pH of the solution can influence the surface charge of the MIB as well as the natural state of existence of phosphorus in an aqueous medium At lower acidic pH values, e.g 2, phosphorus exists mainly as H3PO4 and H2PO4− forms At pH 6, H2PO4− is the major phosphate species and as the initial pH of the solution increases to 9, phosphate in solution mainly exists as HPO42− and PO43- [33] The binding of phosphate oxyanions to the adsorbent occurred more efficiently in an acidic medium, as shown in Fig 4a The removal of phosphorus with both MIBs generally decreased upon increasing pH Complete removal (100%) was achieved with both F13 and C11 MIBs at pH 2, while this amount decreased to 24% and 59% of removal respectively at pH This change of the removal percentage is consistent with previous studies of phosphorus adsorption on metal-functionalized bio-sorbents [34] Even though both MIBs showed great performance in removing phosphorus at low pH values, the dependency on acidic pH for a higher removal was observed to be stronger for Fe-MIB compared to that of Cu-MIB This suggests that Cu-MIB is less sensitive to the pH of O Tomin et al Microporous and Mesoporous Materials 328 (2021) 111499 Fig Simultaneous removal of phosphorus and NOM from lake water at varying phosphorus concentrations with MIBs and RBC (initial COD = 16 mg/l) a) NOM removal, b) phosphorus removal; a demonstration on selectivity the solution which is also confirmed by the pHpzc At higher pH values (6–8) the surface of Fe-MIB gets negatively charged, while the Cu-MIB remains neutral and consequently can attract negatively charged molecules The enhanced adsorption at lower pH can be caused by the proton­ ation of the surface functional groups of MIBs, resulting in higher phosphorus uptake due to the electrostatic attraction The decrease in the phosphorus adsorption onto MIB by shifting pH from acidic to basic conditions can be attributed to the decrease in surface protonation of the adsorbent as well as the hydroxyl ions competition with phosphate ions for adsorptive sites [33,34] The point of zero charge (pHPZC) was identified around pH for both MIB samples, resulting in a neutral surface of the adsorbent (Fig 4a) Below pHPZC, at lower pH, the surface is positively charged, partly due to the donor/acceptor interactions be­ tween the MIB structure and the hydronium ions At pH 5–6, the surface of the MIB is neutral, the anions of phosphate can move to the surface and chelate with the Fe(III) and Cu(II) active sites As pH increases from to 8, the competitive adsorption of hydroxyl ions with phosphate ions in solution causes a visible decrease in uptake of phosphorus on the F13 MIB surface The repelling effect on the negatively charged phosphate ions is noticed to be less for the case of C11 MIB which results in its higher removal capacity at these pH values 3.2.4 Simultaneous adsorption Although NOM is a complex network of molecules with different sizes and structures, the exact composition cannot be specified [35], thus its concentration is measured with a collective parameter, such as COD, representing NOM as one pollutant Fig 5a and b depict simul­ taneous removal of NOM and phosphorus, which is the variation of adsorbed amounts (mg/g) versus initial concentrations (mg/l) of phos­ phorus As it is seen, for the simultaneous removal of the co-existing compounds in the water, the MIBs show different selectivity In the case of NOM (Fig 5a), C11 shows stable removal within the 80–90% range, which is not affected by the increasing co-existing phosphorus concentration in the water On the other hand, the NOM removal ca­ pacity of F13 decreases dramatically from 90% to 7%, when the con­ centration of phosphorus increase to more than mg/l The removal of phosphorus (Fig 5b) shows almost the opposite results The F13 shows higher removal compared to C11 in all concentrations when NOM is present in the solution When no NOM is added to the solution, C11 shows complete removal of phosphorus, but after the addition of NOM, the removal decreases to 85% With a further increase of phosphorus concentration to 20 mg/l, the phosphorus removal decreases to 36% The F13 similarly shows a decreasing phosphorus uptake with the in­ crease of phosphorus concentration Yet, the removal decreases from 92% at mg/l without NOM to 66% at 20 mg/l of phosphorus with NOM in the solution The RBC (R600) shows the NOM removal from 29% to 37% at and 20 mg/l concentrations of co-existing phosphorus, respectively The adsorption happens due to a relatively high BET sur­ face (570 m2/g) enabling the NOM to reach the pores following the pore-filling mechanism However, the RBC fails in phosphorus adsorp­ tion in simultaneous adsorption of NOM and phosphorus, as no func­ tional groups are present on the biochar surface and no bonds could form in the competitive media As can be seen from Fig 5, the RBC (having a much higher BET surface area) is unable to compete with the MIBs in the removal of the target pollutants This supports our hypothesis that surface area is not a key factor in the adsorption performance of the target pollutants by the MIBs when compared to the importance of surface functionality To achieve valuable and selective adsorption performance for a specific pollutant, the adsorbent needs to have suitable functional groups for the complexation of the pollutants on the surface of the adsorbent 3.2.3 Optimization of absorbent dosage The effect of MIB dosages on the removal of NOM and phosphorus is illustrated in Fig 4b and c respectively Both types of MIBs showed no removal for NOM at 0.1 g/l dosage An increase in dosage from 0.1 g/l to 0.5 g/l did not affect NOM removal by MIB F13, while C11 showed a remarkable increase in removal ability to 87% Further increase in C11 dose to g/l showed a little influence on the removal F13 also reached an increase in removal performance from 15% at g/l to 97% at g/l These results indicate the higher affinity of Cu-MIB for NOM requiring a lower amount of the adsorbent compared to that of Fe-MIB Regarding phosphorus removal, which is presented in Fig 4c, the MIB C11 showed a steady increase in removal percentage with the increase of the dose Low removal (5%) at 0.1 g/l, gradually increased to 26% at 0.5 g/l, and significantly raised to 93% at g/l F13 showed an increase in adsorp­ tion with the increase of the dose until g/l With 0.1 g/l, F13 showed a 23% removal, while with g/l, the removal percentage of F13 raised to 95% With g/l the removal slightly decreased to 92% This indicates that the optimal F13 dosage for phosphorus removal is g/l and a further increase is not required The gradual improvement of adsorption by dosage is due to the access number of the exchangeable adsorptive sites on the biochar surface The highest removal (92%) at g/l by the F13 sample was the same as for C11 (93%) at g/l This suggests a better phosphorus removal performance for Fe-MIB compared to that of CuMIB 3.2.5 Mechanism of removal The identification of the underlying mechanisms for the adsorption process is needed for evaluating the removal efficiency of the contami­ nants by the MIBs The adsorption behavior of MIB for different con­ taminants is different and well correlated with the properties of contaminants The adsorption mechanism also depends on surface functional groups, specific surface area, porous structure, and material composition The usual mechanisms involved in NOM adsorption on biochar are O Tomin et al Microporous and Mesoporous Materials 328 (2021) 111499 Fig Schematic illustration of metal impregnation into biochar structure following by the removal mechanism of phosphorus and NOM from water pore filling, π–π interactions, polar/electrostatic interactions, hydro­ phobic effect, and hydrogen bonding [1,7,36,37] The BET analysis confirmed that metal impregnation consumes the surface area and fills the pores of the MIB with metal oxide particles, which resulted in a relatively smaller surface area of the MIB compared to that of the RBC Therefore, the pore-filling may not be the dominant mechanism of NOM/phosphorus adsorption on the developed adsorbents, yet the porosity of the material serves as available surface sites for functional group implementation (Fig 6) After metal impregnation, target pol­ lutants are attracted to the certain functional groups entrenched in the MIB structure, affecting the selectivity of the material Thus, the NOM molecules undergo ligand complexation with positively charged copper sites, which are the active functional groups provided by the Cu-MIB Phosphorus, on the other hand, has an affinity towards iron and un­ dergoes chemisorption with iron-based active sites on the Fe-MIB sur­ face, showing selective phosphorus adsorption The iron modification introduces functional sites on carbon structure, which can provide the chemical co-precipitation of Fe3+/Fe2+, expressed as follows [38]: Fe 2+ + 2Fe 3+ − + 8OH →Fe3 O4 + 4H2 O specialized applications such as selective removal of specific pollutants from the water phase Funding This work was supported by the Aalto University [Grant number D/ 23/00.01.02.00/2019] and Maa ja vesitekniikan tuki ry [Grant number 388823] Declaration of competing interest The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper Appendix A Supplementary data Supplementary data to this article can be found online at https://doi org/10.1016/j.micromeso.2021.111499 (2) In this process, the biochar surface promotes the nucleation of iron oxide precipitation Crystalline Fe3O4 and Fe2O3 particles form within the Fe-MIB porous structure, as was confirmed by XRD analysis Therefore, adsorption capacity is due to the iron oxide-containing groups that exist in the Fe-MIB References [1] M.R Yazdani, N Duimovich, A Tiraferri, P Laurell, M Borghei, J.B Zimmerman, R Vahala, Tailored mesoporous biochar sorbents from pinecone biomass for the adsorption of natural organic matter from lake water, J Mol Liq 291 (2019) 111248, https://doi.org/10.1016/j.molliq.2019.111248 [2] S Velten, D.R.U Knappe, J Traber, H.-P Kaiser, U von Gunten, M Boller, S Meylan, Characterization of natural organic matter adsorption in granular activated carbon adsorbers, Water Res 45 (2011) 3951–3959, https://doi.org/ 10.1016/j.watres.2011.04.047 [3] A Matilainen, M Vepsă ală ainen, M Sillanpă aă a, Natural organic matter removal by coagulation during drinking water treatment: a review, Adv Colloid Interface Sci 159 (2010) 189–197, https://doi.org/10.1016/j.cis.2010.06.007 [4] A.I Omoike, G.W vanLoon, Removal of phosphorus and organic matter removal by alum during wastewater treatment, Water Res 33 (1999) 3617–3627, https:// doi.org/10.1016/S0043-1354(99)00075-5 [5] X.-H Guan, C Shang, G.-H Chen, Competitive adsorption of organic matter with phosphate on aluminum hydroxide, J Colloid Interface Sci 296 (2006) 51–58, https://doi.org/10.1016/j.jcis.2005.08.050 [6] M Yan, D Wang, J Ni, J Qu, C.W.K Chow, H Liu, Mechanism of natural organic matter removal by polyaluminum chloride: effect of coagulant particle size and Conclusions The tailoring process of MIB was optimized by comparing the in­ fluence of pyrolysis temperature and two different activators during the production stage The NOM was selectively adsorbed from the binary solution by the Cu-MIB and phosphorus was selectively removed by the Fe-MIB Adsorption results, along with the characterization, confirm that the metal oxide contents within the porous structure of MIB play the role of active sites for the selective removal of target pollutants This study points towards the tailoring process of biochars for more O Tomin et al [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] Microporous and Mesoporous Materials 328 (2021) 111499 [23] M.R Yazdani, T Tuutijă arvi, A Bhatnagar, R Vahala, Adsorptive removal of arsenic (V) from aqueous phase by feldspars: kinetics, mechanism, and thermodynamic aspects of adsorption, J Mol Liq 214 (2016) 149–156, https://doi.org/10.1016/j molliq.2015.12.002 [24] M Taheran, M Naghdi, S.K Brar, E.J Knystautas, M Verma, A.A Ramirez, R Y Surampalli, J.R Valero, Adsorption study of environmentally relevant concentrations of chlortetracycline on pinewood biochar, Sci Total Environ 571 (2016) 772–777, https://doi.org/10.1016/j.scitotenv.2016.07.050 [25] E Apaydın-Varol, A.E Pütün, Preparation and characterization of pyrolytic chars from different biomass samples, J Anal Appl Pyrolysis 98 (2012) 29–36, https:// doi.org/10.1016/j.jaap.2012.07.001 [26] R Li, J.J Wang, L.A Gaston, B Zhou, M Li, R Xiao, Q Wang, Z Zhang, H Huang, W Liang, H Huang, X Zhang, An overview of carbothermal synthesis of metal–biochar composites for the removal of oxyanion contaminants from aqueous solution, Carbon 129 (2018) 674–687, https://doi.org/10.1016/j carbon.2017.12.070 [27] R He, X Yuan, Z Huang, H Wang, L Jiang, J Huang, M Tan, H Li, Activated biochar with iron-loading and its application in removing Cr (VI) from aqueous solution, Colloids Surf Physicochem Eng Asp 579 (2019) 123642, https://doi org/10.1016/j.colsurfa.2019.123642 [28] J.-Q Jiang, C Cooper, S Ouki, Comparison of modified montmorillonite adsorbents: Part I: preparation, characterization and phenol adsorption, Chemosphere 47 (2002) 711–716, https://doi.org/10.1016/S0045-6535(02) 00011-5 [29] Y Yu, X Li, J Cheng, A comparison study of mechanism: Cu2+ adsorption on different adsorbents and their surface-modified adsorbents, J Chem 2016 (2016), e7936258, https://doi.org/10.1155/2016/7936258 [30] M Galamboˇs, M Daˇ no, E Viglaˇsov´ a, L Krivosudský, O Rosskopfov´ a, I Nov´ ak, D Berek, P Rajec, Effect of competing anions on pertechnetate adsorption by activated carbon, J Radioanal Nucl Chem 304 (2015) 1219–1224, https://doi org/10.1007/s10967-015-3953-4 [31] C.L Waters, R.R Janupala, R.G Mallinson, L.L Lobban, Staged thermal fractionation for segregation of lignin and cellulose pyrolysis products: an experimental study of residence time and temperature effects, J Anal Appl Pyrolysis 126 (2017) 380–389, https://doi.org/10.1016/j.jaap.2017.05.008 [32] M Brebu, C Vasile, Thermal degradation of lignin – a Review, Cellul Chem Technol 44 (2010) 353–363 [33] L Ding, C Wu, H Deng, X Zhang, Adsorptive characteristics of phosphate from aqueous solutions by MIEX resin, J Colloid Interface Sci 376 (2012) 224–232, https://doi.org/10.1016/j.jcis.2012.03.002 [34] M.R Yazdani, E Virolainen, K Conley, R Vahala, Chitosan–Zinc(II) Complexes as a Bio-Sorbent for the Adsorptive Abatement of Phosphate: Mechanism of Complexation and Assessment of Adsorption Performance, 2018, p 19 [35] E Worch, Adsorption Technology in Water Treatment: Fundamentals, Processes, and Modelling, De Gruyter, Inc., Berlin/Boston, GERMANY, 2012 http://ebookce ntral.proquest.com/lib/aalto-ebooks/detail.action?docID=893772 (Accessed 13 October 2020) accessed [36] X Tan, Y Liu, G Zeng, X Wang, X Hu, Y Gu, Z Yang, Application of biochar for the removal of pollutants from aqueous solutions, Chemosphere 125 (2015) 70–85, https://doi.org/10.1016/j.chemosphere.2014.12.058 [37] W Gwenzi, N Chaukura, C Noubactep, F.N.D Mukome, Biochar-based water treatment systems as a potential low-cost and sustainable technology for clean water provision, J Environ Manag 197 (2017) 732–749, https://doi.org/ 10.1016/j.jenvman.2017.03.087 [38] Q Yang, X Wang, W Luo, J Sun, Q Xu, F Chen, J Zhao, S Wang, F Yao, D Wang, X Li, G Zeng, Effectiveness and mechanisms of phosphate adsorption on iron-modified biochars derived from waste activated sludge, Bioresour Technol 247 (2018) 537–544, https://doi.org/10.1016/j.biortech.2017.09.136 hydrolysis kinetics, Water Res 42 (2008) 3361–3370, https://doi.org/10.1016/j watres.2008.04.017 M Ahmad, A.U Rajapaksha, J.E Lim, M Zhang, N Bolan, D Mohan, M Vithanage, S.S Lee, Y.S Ok, Biochar as a sorbent for contaminant management in soil and water: a review, Chemosphere 99 (2014) 19–33, https://doi.org/ 10.1016/j.chemosphere.2013.10.071 A.U Rajapaksha, S.S Chen, D.C.W Tsang, M Zhang, M Vithanage, S Mandal, B Gao, N.S Bolan, Y.S Ok, Engineered/designer biochar for contaminant removal/immobilization from soil and water: potential and implication of biochar modification, Chemosphere 148 (2016) 276–291, https://doi.org/10.1016/j chemosphere.2016.01.043 H Yang, S Ye, Z Zeng, G Zeng, X Tan, R Xiao, J Wang, B Song, L Du, M Qin, Y Yang, F Xu, Utilization of biochar for resource recovery from water: a review, Chem Eng J 397 (2020) 125502, https://doi.org/10.1016/j.cej.2020.125502 Y Zheng, B Wang, A.E Wester, J Chen, F He, H Chen, B Gao, Reclaiming phosphorus from secondary treated municipal wastewater with engineered biochar, Chem Eng J 362 (2019) 460–468, https://doi.org/10.1016/j cej.2019.01.036 P.R Yaashikaa, P Senthil Kumar, S.J Varjani, A Saravanan, Advances in production and application of biochar from lignocellulosic feedstocks for remediation of environmental pollutants, Bioresour Technol 292 (2019) 122030, https://doi.org/10.1016/j.biortech.2019.122030 F.J Chac´ on, M.A S´ anchez-Monedero, L Lezama, M.L Cayuela, Enhancing biochar redox properties through feedstock selection, metal preloading and post-pyrolysis treatments, Chem Eng J 395 (2020) 125100, https://doi.org/10.1016/j cej.2020.125100 Z Wan, Y Sun, D.C.W Tsang, E Khan, A.C.K Yip, Y.H Ng, J Rinklebe, Y.S Ok, Customised fabrication of nitrogen-doped biochar for environmental and energy applications, Chem Eng J 401 (2020) 126136, https://doi.org/10.1016/j cej.2020.126136 M.R Yazdani, N Duimovich, A Tiraferri, P Laurell, M Borghei, J.B Zimmerman, R Vahala, Dataset for natural organic matter treatment by tailored biochars, Data Brief 25 (2019) 104353, https://doi.org/10.1016/j.dib.2019.104353 S Tuomikoski, H Runtti, H Romar, U Lassi, T Kangas, Multiple heavy metal removal simultaneously by a biomass-based porous carbon, Water Environ Res 93 (8) (2019) 1303–1314, https://doi.org/10.1002/wer.1514 R Mahmudov, C.P Huang, Selective adsorption of oxyanions on activated carbon exemplified by Filtrasorb 400 (F400), Separ Purif Technol 77 (2011) 294–300, https://doi.org/10.1016/j.seppur.2010.12.019 X Hu, Z Ding, A.R Zimmerman, S Wang, B Gao, Batch and column sorption of arsenic onto iron-impregnated biochar synthesized through hydrolysis, Water Res 68 (2015) 206–216, https://doi.org/10.1016/j.watres.2014.10.009 F Meng, Z Li, C Lei, K Yang, D Lin, Removal of trichloroethene by iron-based biochar from anaerobic water: key roles of Fe/C ratio and iron carbides, Chem Eng J (2020) 127391, https://doi.org/10.1016/j.cej.2020.127391 ˇ ska, E Moreno-Jimenez, B Mich´ alekov´ a-Richveisov´ a, V Friˇst´ ak, M Pipíˇska, L Duriˇ G Soja, Iron-impregnated biochars as effective phosphate sorption materials, Environ Sci Pollut Res 24 (2017) 463–475, https://doi.org/10.1007/s11356016-7820-9 Y Qiu, X Xu, Z Xu, J Liang, Y Yu, X Cao, Contribution of different iron species in the iron-biochar composites to sorption and degradation of two dyes with varying properties, Chem Eng J 389 (2020) 124471, https://doi.org/10.1016/j cej.2020.124471 Y Yi, Z Huang, B Lu, J Xian, E.P Tsang, W Cheng, J Fang, Z Fang, Magnetic biochar for environmental remediation: a review, Bioresour Technol 298 (2020) 122468, https://doi.org/10.1016/j.biortech.2019.122468 H.S Kambo, A Dutta, A comparative review of biochar and hydrochar in terms of production, physico-chemical properties and applications, Renew Sustain Energy Rev 45 (2015) 359–378, https://doi.org/10.1016/j.rser.2015.01.050 ... Effect of pH The effect of pH on the adsorption of phosphorus was tested within the 2–8 pH range The pH of the solution can influence the surface charge of the MIB as well as the natural state of. .. parameters for the production of selective adsorbents and (iii) to test the tailored biochars in selective removal of target pollutants from the lake water samples, contaminated with phosphorus and organic. .. investigate the phase transformation of metal salts during the activation and the formation of metal oxide crystals, the XRD mea­ surements were performed on the MIBs Fig 2b and c shows the XRD patterns

Ngày đăng: 20/12/2022, 22:15

TÀI LIỆU CÙNG NGƯỜI DÙNG

TÀI LIỆU LIÊN QUAN

w