Next Article in Journal
Adsorption Characteristics between Ti Atoms of TiO2(100) and Corrosive Species of CO2-H2S-Cl System in Oil and Gas Fields
Previous Article in Journal
Experimental Research on Dynamic Behavior of Stiffened Plates under Drop-Weight Impacts of a Wedge Impactor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic Biochar Obtained by Chemical Coprecipitation and Pyrolysis of Corn Cob Residues: Characterization and Methylene Blue Adsorption

by
Norma Araceli Guel-Nájar
1,
Jorge Carlos Rios-Hurtado
1,*,
Elia Martha Muzquiz-Ramos
2,
Gloria I. Dávila-Pulido
3,
Adrián A. González-Ibarra
3 and
Aurora M. Pat-Espadas
4
1
Facultad de Metalurgia, Universidad Autónoma de Coahuila, Carretera 57 Km 5, Monclova 25710, Coahuila, Mexico
2
Facultad de Ciencias Químicas, Universidad Autónoma de Coahuila, Blvd. Venustiano Carranza S/N, República, Saltillo 25280, Coahuila, Mexico
3
Escuela Superior de Ingeniería, Universidad Autónoma de Coahuila, Boulevard Adolfo López Mateos S/N, Independencia, Nueva Rosita 26830, Coahuila, Mexico
4
CONACyT, Estación Regional del Noroeste del Instituto de Geología de la UNAM, Luis D Colosio S/N Esquina Madrid, Hermosillo 83200, Sonora, Mexico
*
Author to whom correspondence should be addressed.
Materials 2023, 16(8), 3127; https://doi.org/10.3390/ma16083127
Submission received: 3 March 2023 / Revised: 27 March 2023 / Accepted: 10 April 2023 / Published: 15 April 2023

Abstract

:
Biochar is a carbonaceous and porous material with limited adsorption capacity, which increases by modifying its surface. Many of the biochars modified with magnetic nanoparticles reported previously were obtained in two steps: first, the biomass was pyrolyzed, and then the modification was performed. In this research, a biochar with Fe3O4 particles was obtained during the pyrolysis process. Corn cob residues were used to obtain the biochar (i.e., BCM) and the magnetic one (i.e., BCMFe). The BCMFe biochar was synthesized by a chemical coprecipitation technique prior to the pyrolysis process. The biochars obtained were characterized to determine their physicochemical, surface, and structural properties. The characterization revealed a porous surface with a 1013.52 m2/g area for BCM and 903.67 m2/g for BCMFe. The pores were uniformly distributed, as observed in SEM images. BCMFe showed Fe3O4 particles on the surface with a spherical shape and a uniform distribution. According to FTIR analysis, the functional groups formed on the surface were aliphatic and carbonyl functional groups. Ash content in the biochar was 4.0% in BCM and 8.0% in BCMFe; the difference corresponded to the presence of inorganic elements. The TGA showed that BCM lost 93.8 wt% while BCMFe was more thermally stable due to the inorganic species on the biochar surface, with a weight loss of 78.6%. Both biochars were tested as adsorbent materials for methylene blue. BCM and BCMFe obtained a maximum adsorption capacity (qm) of 23.17 mg/g and 39.66 mg/g, respectively. The obtained biochars are promising materials for the efficient removal of organic pollutants.

1. Introduction

Due to rapid population growth, accelerated urbanization, and industrialization, the volume and diversity of agro-industrial wastes have increased. These wastes are considered a significant source of pollution due to poor management and improper disposal [1,2] (Awogbemi and Kallon, 2022; Karić et al., 2022). A strategy to reduce the disposed residues is to prepare new materials for different applications. Biochar is a solid material obtained from the thermochemical conversion of biomass in an oxygen-limited environment [3]. It is a carbon-rich organic material that is prepared by heating biomass. Depending on the feedstock used to obtain it, the properties of the biochar such as its high specific surface area, microporosity, and ion exchange capacity can be improved [4,5]. Furthermore, corn residues are suitable to obtain biochar due to their lignocellulosic biomass, which favors its formation [6,7].
The use of biochar as an adsorbent material has been widely studied; however, the adsorption efficiency can be improved. Surface activation or modification with chemical reagents, steam and gas, mineral oxides, organic compounds, clay minerals, and microorganisms has been widely reported [8,9,10]. Composite biochar is prepared by impregnating the surface with metal oxides to modify the surface properties and increase the adsorption capacity [10,11].
One of the ideal methods of surface modification is with magnetite (i.e., Fe3O4). This oxide is non-toxic, can improve the adsorption capacity towards various pollutants, and provides strong magnetic properties. Ion exchange, complexation, electrostatic attraction, and metal–π interaction are mechanisms that make adsorption techniques suitable. Due to its magnetic properties, Fe3O4 can be used to recycle biochar, as iron metal oxides allow adsorbent recovery [12,13,14]. This makes it important to obtain magnetic biochars for the removal of water pollutants. Santhosh et al. synthesized a magnetic biochar from sewage sludge and woodchips, for Cr (VI) removal [15]. A magnetic biochar in a single-step method was obtained for the removal of organic pollutants by Chen et al. (2022) [16]. A filamentous fungus, Trichoderma reesei, was used as a template for 3D biochar obtention.
Water pollution is one of the challenges facing society [17]. One of the main sources of pollutants are organic dyes such as methylene blue (MB). This dye is widely used in the textile, food, cosmetic, and pharmaceutical industries; it usually causes negative effects on photosynthetic activity [17,18,19]. For this reason, it is important to remove dyes using adsorbent materials, such as biochars.
To increase the MB adsorption capacity, a wide variety of precursors have been used to obtain biochars with modified surfaces. Mu et al. (2022) obtained a magnetic biochar by slow pyrolysis of tea residue powder with FeCl3 solution impregnation. The authors studied MB adsorption as a dye model and described the adsorption process due to the main interactions of pore filling, hydrogen bonding, π–π interactions, and electrostatic attraction [20]. Güleç et al. (2022) prepared biochars from seaweed and rapeseed, obtaining a qm of 117.6 and 70.9 mg/g, respectively [21]. Zeng et al. (2021) prepared a sludge-based magnetic biochar by pyrolysis and studied MB adsorption [22]. Li et al. (2022) made a biochar from moringa seed shells. In the study, Fe3O4 was added to activate the surface of the biochar, and a high removal of MB was observed [23].
Therefore, in this work, biochars were obtained from corn cob residues by pyrolysis. A pure biochar and a magnetite impregnated with magnetite by the coprecipitation technique prior to the pyrolysis process were obtained. Both biochars were characterized to determine their physicochemical, surface, and structural properties. MB adsorption was tested to understand the behavior of biochars when modified with Fe3O4.

2. Materials and Methods

2.1. Reagents

To obtain biochar, corn cob waste from the municipality of Frontera, Coahuila, was used as a precursor. The chemicals used are of reagent grade: hydrochloric acid (CTR Scientific, Monterrey, MX), ferric chloride hexahydrate (FeCl3·6H2O), ferrous chloride tetrahydrate (FeCl2·4H2O), and sodium hydroxide (Fermont, Monterrey, MX).

2.2. Experimental Procedures

2.2.1. Pretreatment of Samples

The corn cob waste was cut into small pieces of approximately 3 cm in length, then washed with distilled water. A part of the cleaned waste was dried at 100 °C in an ECOSHEL 9053A oven (LABOTECA, Monterrey, MX) for 24 h. The remaining corn cob was passed through the chemical coprecipitation process.

2.2.2. Chemical Coprecipitation of Magnetite in Corn Cob Waste

The coprecipitation method was based on the work of Rodriguez et al. (2019) [24]. A solution was prepared in a beaker with 50 mL of deionized water, FeCl3·6H2O and FeCl2·4H2O at a 2:1 ratio according to Equation (1):
2 Fe 3 + + Fe 2 + + 8 OH Fe 3 O 4 + 4 H 2 O
The solution was heated at 50 °C for 15 min. Corn cobs cleaned of waste were placed in contact with 5 mL of Fe3+:Fe2+ ion solution. Then, 2 mL of concentrated NaOH was added by dripping and subsequently dried at 100 °C for 24 h.

2.2.3. Obtaining Biochar from Corn Cob by Pyrolysis

The research of Lee et al. (2017) and Xie et al. (2021) was used to develop corn cob pyrolysis [25,26]. Dried corn cob residues were placed in a CARBOLITE™ CTF horizontal tube furnace (Fischer Scientific, Monterrey, MX) and heated to 500 °C for 1 h. The heating rate and N2 flow rate were varied. The obtained sample was allowed to cool slowly at room temperature. Biochar was labeled as BCM, and the modified biochar as BCMFe.

2.2.4. Characterization of Biochar

The biochar yield was determined according to the procedure reported by Qin et al. (2020), as well as the ash content using ASTM D2866-94 (ASTM, n.d.) [27,28]. In addition, SEM and EDS analysis were performed on a HITACHI model SU8230 cold cathode SEM machine (Ciudad de México, MX) with backscattered electron and secondary electron detectors; FTIR analysis with a PerkinElmer FTIR Spectrometer Frontier (Monterrey, Mexico) in a range of 4000 to 600 cm−1 using 32 scans and 4 cm−1 resolution; XRD on a Bruker D8 ADVANCE model (Ciudad de México, MX) (CuKa: 1.54 Å, 40 mA and 40 kV); XRF analysis on a Malvern Panalytical Epsilon 1 model (Monterrey, MX); and TGA on a TA Instruments TGA 550 (Ciudad de México, MX), in a temperature range of 30 °C to 700 °C at a heating rate of 10 °C/min, using air atmosphere; and N2 physisorption at 77 K was performed in a BEL BELSORP-miniX surface analyzer (Ciudad de México, MX).

2.2.5. Methylene Blue Adsorption Efficiency and Kinetic Models

MB adsorption efficiency was determined for both biochars. In 15 mL conical tubes, 0.02 g of biochar and 10 mL of an MB solution at different concentrations (i.e., 1, 2, 4, 8, 10, and 25 ppm) were placed in contact. The pH was adjusted to 7 ± 0.2 in an Orion StarA2110 potentiometer (Monterrey, MX). The initial concentration (Co) absorbance of each of the solutions was measured in a Mettler Toledo UV-Vis instrument model UV5 (Ciudad de México, MX) (λ = 664 nm). Then, the tubes were placed in a LabTech LSI-3016A incubator (El Crisol, Monterrey, México) at 25 °C and 150 rpm for 24 h. Once equilibrium was reached, the remaining concentration (Ce) values were measured by UV-Vis. The maximum adsorption capacity was calculated according to Equation (2) and fitted to the Langmuir isotherm model according to Equation (3):
Qe = ( C O C e ) ( V ) m  
q e = q max k L C e 1 + k L C e
where:
q e = adsorption capacity at equilibrium, mg·g−1
C o = initial concentration of MB solution, mg·L−1
C e = final concentration of MB solution, mg·L−1
V = solution volume, L
m = adsorbent mass, g
q max = maximum adsorption capacity, mg·g−1
k L = adsorption energy constant, L·mg−1
For kinetic experiments, 60 mL of a 10 ppm MB solution and 60 mg of the material were placed in a beaker. Sampling was carried out during 0, 1, 3, 3, 5, 10, 20, 40, and 60 min. The concentration was measured by a UV/Vis spectrometer. The data obtained were adjusted to the pseudo-second-order model using the following equation (Equation (4)):
t q t = 1 k 2 ( q e ) 2 + 1 q e t
where:
t = time, min
qt = adsorption capacity at time, mg·g−1
qe = adsorption capacity at equilibrium, mg·g−1
k2 = PSO constant (g·mg−1·g−1)

3. Results and Discussions

3.1. Biochar Yield and Ash Percentage

It is important to dry the corn cob samples prior to pyrolysis in order to calculate the amount of biochar obtained. Besides, subjecting wet feedstock samples to pyrolysis would consume energy, and carbonization would last longer [8]. In addition, a large amount of moisture in the biomass leads to the production of liquid by-products [29]. The pyrolysis experimental conditions of the BCM samples and the yield are shown in Table 1. The N2 flow rate was tested, and for a 50 mL/min rate, the yield was 5.60%, and for a 35 L/min rate, the yield was 28.48%. Moreover, different heating ramps were tested. The results showed that when a lower heating ramp was used (5 °C/min), the biochar yield was 26.91%. However, a higher heating ramp (10 °C/min) produced a biochar yield of 28.48% and reduced the time needed to reach the pyrolysis temperature. In addition, raw and cooked corn cob (180 °C, 1 h) conditions were tested, and no significant difference in biochar yield was observed. Therefore, to obtain a higher biochar yield, an N2 flow rate of 35 mL/min and a heating ramp of 10 °C/min were used, in which an average yield of 28.48 ± 0.098% was obtained.
In Table 2, the ash contents of different samples are shown. It can be observed that corn cob residues contain 0.4%, while BCM shows a significant increase with 4.0%, which is lower compared to what was reported by Zhao et al. (2018); these authors determined 10.79% of ash content when heating at 300 °C and 23.27% at 500 °C [30].
The results shown in Table 2 are also different from those reported by Mohan et al. (2014) and Xie et al. (2021), who used corn as biomass and the same carbonization temperature [26,31]. For BCMFe, an 8.0% ash content was observed, a higher amount compared to BCM, which was attributed to the presence of inorganic particles, such as iron oxides, on the surface of BCMFe, due to the melting point of iron being >1500 °C.
The biochar yield varied according to the type of biomass and the temperature used in the pyrolysis process, according to Amalina et al. (2022) [32]. Table 3 shows a comparison with biochars reported in the literature. For BCM, the results are similar to those reported by Mohan et al. (2014) [31]. However, as reported by Suo et al. (2021), an increase in the percentage can be attributed to a lower carbonization temperature [33].

3.2. Inorganic Composition

In order to determine the ash composition, XRF analysis was performed. The results are shown in Table 4, where it can be observed the presence of elements such as K, P, and Si, which are in a higher proportion in the BCM biochar when compared with corn cobs. The amount of K increases considerably by the pyrolysis process, being the main element in BCM with a concentration greater than 11%, similar to that reported by Gómez-Vásquez et al. (2022) [36]. The high amount of elements was attributed to the fertilizer used for plant growth (Land and Water Division, FAO, 2002) since the main elemental composition is K, P, and Si with concentrations greater than 0.6% [37]. Other elements such as Cl, Fe, and S were also detected, albeit in lower concentrations. For BCMFe, an increase in Fe amount (8.761%) was attributed to the formation of iron oxides by the coprecipitation method.

3.3. Morphological Analysis

SEM analysis was performed to characterize the surface morphologies of the samples. Figure 1 shows the BCM surface, where a highly porous structure, a large surface area (SBET: 1013.52 m2/g), and evenly distributed pores were observed. The pore size range was determined as mesopores and macropores, as micropores are generally more difficult to detect, according to the report by Quillope et al. (2021) [38].
The surface structure (SBET = 903.67 m2/g) of magnetic biochar (BCMFe) is shown in Figure 2. Particles with an intense brightness were observed, indicating the presence of an element with a higher atomic mass such as Fe [39]. Fe3O4 particles generally adopted a spherical shape and formed agglomerates, which were distributed over the biochar surface, according to the reports by Ouyang et al. (2017) and J. Yan et al. (2020) [40,41]. It can be determined from the SEM analysis of Figure 2 using Image-Pro Plus software that the average agglomerate diameter is 5.83 ± 2.23 μm. This is in accordance with the value reported by Navarathna et al. (2019), in which the agglomerated particle size was up to 7 μm [42].
The BCMFe surface elemental composition was obtained by EDS analysis. Four points of the sample (see Figure 3) were analyzed to obtain the mass percentage of each element (Table 5). Biochar BCM consists mainly of C (carbon) and O (oxygen). Other elements were also detected, although in smaller proportions. The results were consistent with those obtained by XRF analysis for Si, K, and Fe. No presence of Cl was observed in the sample, so Fe ions were transformed into iron oxide. In addition, the Fe percentage increased significantly due to the presence of Fe particles in the biochar.

3.4. Structure of Modified Biochar

The crystalline structures of the modified biochar were characterized by X-ray diffraction. The XRD pattern of the BCMFe is presented in Figure 4, which confirms that the precipitated iron ions are mainly magnetite (Fe3O4). A series of diffraction peaks corresponding to Fe3O4 were detected at 2θ values of 30.02°, 35.45°, 43.08°, 47.38°, and 62.53°, with Miller indices of (220), (311), (400), (331), and (440), respectively. The observed signals coincided with the diffraction patterns of Fe3O4 reported by Compeán-Jasso et al. (2008) and Peña-Rodríguez et al. (2018), confirming the successful formation of magnetite on the biochar surface [43,44].
In Figure 4, an intense signal at 29.29° due to silicon oxide (SiO2) with crystallographic chart 96-900-0809 can be observed. Since SiO2 is a compound that is prominent in the inorganic components of corn cob [36], or it may belong to a potassium mineral (K) with crystallographic chart 96-901-1978, these elements can be found in biochar due to the inorganic elements present in the biomass that was not carbonized. The signals for amorphous graphite were also observed at signals between 20 and 30° and 40 and 50°.
In order to identify if the material had the characteristic magnetic properties of magnetite, a simple test was carried out with the addition of a magnet to a suspension with 0.1 g of the magnetic biochar, and as can be seen in Figure 5, the material stuck to the magnet due to the magnetic properties.

3.5. Surface Functional Groups

An FTIR analysis was performed to observe surface functional groups on the obtained biochars. The results are presented in Figure 6, where the low intensity of the signals, which are attributed to the temperature at which the pyrolysis process was carried out, can be observed. According to Manoharan et al. (2022), the number of functional groups decreases at higher temperatures due to the release of CO2 [45].
Both biochars showed similar bands, one observed at 1712 cm−1, corresponding to the C=O stretching vibration of the carbonyl group [46], which is formed during the carbonization process, and the double signals between 2800 and 3000 cm−1, corresponding to -C–H aliphatic compounds stretching vibrations [35,47]. The band near 500 cm−1 could be described as a stretching vibration of metallic oxides due to the presence of inorganic compounds in BCM and BCMFe [48].

3.6. Thermal Behavior

Thermal analysis was carried out on both biochars (BCM and BCMFe). The thermogravimetric (TG) and differential thermogravimetric (DTG) curves of Figure 7 show two main weight loss processes for both biochars. In the first curve, BCM showed a 4% loss at 89 °C, while BCMFe showed a 10% weight loss, due to water molecules evaporating from the biochar surface. At higher temperatures, a greater reduction in weight loss was observed. In BCM, a weight loss of 93.8% was observed at 500 °C, while a weight loss of 78.6% was observed in BCMFe at 650 °C. The loss was attributed to the decomposition of lignocellulosic biomass and extracts such as proteins, fats, and sugars, among others [14,49], results observed in FTIR analysis, in addition to the loss of functional groups such as carboxyl and hydroxyl groups [50]. Compared to BCM biochar, BCMFe carbonized at a higher temperature and showed higher thermal stability, which corresponded to a greater amount of inorganic components [51].
The difference between the amount of weight loss for BCMFe and BCM was 15.2% and corresponds to the F3O4 present on the biochar surface according to XRD results. The DTG curve of the impregnated biochar showed two signals between 400 and 650 °C, attributed to the Fe3O4 oxidation to α-Fe2O3 and γ-Fe2O3.

3.7. Methylene Blue Adsorption Capacity

MB adsorption capacity was studied at pH 7 since the PZC of BCM was measured as 7.06. The adsorption capacity was adjusted to the Langmuir model, as can be seen in Figure 8. Non-linear regression was used to adjust the experimental data using Origin® software. BCM biochar showed a maximum adsorption capacity (qm) of 23.17 mg/g and BCMFe biochar of 39.66 mg/g, as shown in Table 6.
The qm of BCMFe biochar is similar to that reported by Zeng et al. (2021), with a qm of 39.35 mg/g at a temperature of 35 °C, using a sewage sludge-based magnetic biochar (Fe3O4) by pyrolysis for the removal of methylene blue in water [22]. Mubarak et al. (2015) prepared a magnetic biochar derived from an empty fruit bunch treated with ferric chloride (FeCl3) to remove methylene blue, obtaining a qm of 31.25 mg/g [52], while Ruthiraan et al. (2017) obtained a qm of 46.30 mg/g in MB removal with a magnetic iron oxide biochar (impregnated with Fe2O3 solution) obtained from mangosteen peel [53].
In addition to qm, Table 6 shows the kL, RL, and R2 parameters. The Langmuir constant (kL) indicates the degree of interaction between the adsorbate and the surface. When the adsorption energy constant (kL) was less than 1, it indicates that there was a strong interaction between the adsorbate and the adsorbent. The Langmuir separation factor (RL) showed values lower than 1 in both biochars, which indicated an appropriate adsorption process. If RL > 1 is not suitable, while if 0 < RL < 1 is appropriate [54].
Further, the linear correlation coefficient (R2) was high, so the adsorption isotherms fit well for the Langmuir model. This model indicates that the adsorbent surface was completely flat and homogeneous and that each surface site can only hold one molecule of the adsorbate, resulting in monolayer-type adsorption [55,56]. The metal–π interaction is one of the mechanisms reported in the literature for organic compounds in modified biochars with metallic oxides [57,58,59,60]. By this, an increase in the MB adsorption capacity for BCMFe was observed, due to the presence of Fe3O4 particles on the surface. Several reports have corroborated that MB adsorption by activated carbons from lignocellulosic materials demonstrates a strong correlation with the Langmuir isotherm equation [61].
Furthermore, experimental data were adjusted to the PFO and PSO order kinetic models; however, both biochars were better fitted to the PSO (R2 > 0.998) (Figure 9). The PSO model describes the adsorption process in a surface monolayer. The results agreed with those observed in adsorption isotherms.

4. Conclusions

The results showed that the yield of BCM was 28.483%, higher than that of BCMFe (i.e., 26.160%). Both biochars show a highly porous surface, a large surface area, and uniformly distributed pores and they are mainly composed of C and O, in addition to Fe, Si, and K. BCMFe biochar contains a higher amount of Fe compared to BCM due to its impregnation with Fe ions, successfully achieving the addition of Fe3O4 particles, and XRD analysis supports this result.
While the ash content of BCM was 4% lower than BCMFe (i.e., 8%), there were 4% more inorganic compounds in BCMFe. Both biochars presented carbonyl and aliphatic groups on the surface. The TGA revealed that BCM carbonizes at 500 °C with a weight loss of 93.8%, while BCMFe carbonizes at 650 °C with a weight loss of 78.6%. The above indicates that BCMFe showed higher thermal stability.
It was possible to obtain a material capable of removing the organic pollutant methylene blue, with a maximum adsorption capacity of 23.17 mg/g and 39.66 mg/g for BCM and BCMFe biochar, respectively. The modified biochar (BCMFe) showed a higher adsorption capacity (i.e., 71.17%) than BCM.

Author Contributions

N.A.G.-N. performed the experimental part, wrote the introduction and methods of the manuscript; J.C.R.-H. designed the experimentation, followed up on the methodology and supervised the discussion of the manuscript; E.M.M.-R. supported in the thermodynamic experiments of methylene blue adsorption and analyzed the results; G.I.D.-P. supported in experimentation and discussion of characterizations; A.A.G.-I. supported in the discussion of the results and language revision; A.M.P.-E. supported in the materials characterization. All authors have read and agreed to the published version of the manuscript.

Funding

Norma Araceli Guel-Nájar received a grant from CONACyT. Grant number 620952; Jorge Carlos Rios-Hurtado received a financial support for the project from COECyT. Project number FONCyT COAH-2021-C15-C013.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The grant (620952) awarded by the Consejo Nacional de Ciencia y Tecnología (CONACYT) is gratefully acknowledged for Guel-Najar, and all authors thank the Fondo de Investigación Científica y Tecnológica (FONCyT COAH-2021-C15-C013) for the financial support of this project.

Conflicts of Interest

We know of no conflicts of interest associated with this publication, and there has been no significant financial support for this work that could have influenced its outcome.

References

  1. Awogbemi, O.; Von Kallon, D.V. Pretreatment techniques for agricultural waste. Case Stud. Chem. Environ. Eng. 2022, 6, 100229. [Google Scholar] [CrossRef]
  2. Karić, N.; Maia, A.S.; Teodorović, A.; Atanasova, N.; Langergraber, G.; Crini, G.; Ribeiro, A.R.L.; Đolić, M. Bio-waste valorisation: Agricultural wastes as biosorbents for removal of (in)organic pollutants in wastewater treatment. Chem. Eng. J. Adv. 2022, 9, 100239. [Google Scholar] [CrossRef]
  3. International Biochar Initiative. Standardized Product Definition and Product Testing Guidelines for Biochar That Is Used in Soil. International Biochar Initiative, November, 1–61. 2015. Available online: https://www.biochar-international.org/wp-content/uploads/2018/04/IBI_Biochar_Standards_V2.1_Final.pdf (accessed on 13 January 2023).
  4. Mehdi, R.; Khoja, A.H.; Naqvi, S.R.; Gao, N.; Amin, N.A.S. A Review on Production and Surface Modifications of Biochar Materials via Biomass Pyrolysis Process for Supercapacitor Applications. Catalysts 2022, 12, 798. [Google Scholar] [CrossRef]
  5. Saravanan, A.; Senthil Kumar, P. Biochar derived carbonaceous material for various environmental applications: Systematic review. Environ. Res. 2022, 214, 113857. [Google Scholar] [CrossRef] [PubMed]
  6. Tehreem, S.; Yousra, M.; Alamer, K.H.; Alsudays, I.M.; Sarwar, S.; Kamal, A.; Naeem, S. Analysis of the role of various biochar in the remediation of heavy metals in contaminated water and its kinetics study. J. Saudi Chem. Soc. 2022, 26, 101518. [Google Scholar] [CrossRef]
  7. Wang, L.; Olsen, M.N.P.; Moni, C.; Dieguez-Alonso, A.; de la Rosa, J.M.; Stenrød, M.; Liu, X.; Mao, L. Comparison of properties of biochar produced from different types of lignocellulosic biomass by slow pyrolysis at 600 °C. Appl. Energy Combust. Sci. 2022, 12, 100090. [Google Scholar] [CrossRef]
  8. Amalina, F.; Syukor Abd Razak, A.; Krishnan, S.; Sulaiman, H.; Zularisam, A.W.; Nasrullah, M. Advanced techniques in the production of biochar from lignocellulosic biomass and environmental applications. Clean. Mater. 2022, 6, 100137. [Google Scholar] [CrossRef]
  9. Li, R.; Wu, Y.; Lou, X.; Li, H.; Cheng, J.; Shen, B.; Qin, L. Porous Biochar Materials for Sustainable Water Treatment: Synthesis, Modification, and Application. Water 2023, 15, 395. [Google Scholar] [CrossRef]
  10. Huang, W.H.; Lee, D.J.; Huang, C. Modification on biochars for applications: A research update. Bioresour. Technol. 2021, 319, 124100. [Google Scholar] [CrossRef]
  11. Roy, H.; Islam, M.S.; Arifin, M.T.; Firoz, S.H. Synthesis, Characterization and Sorption Properties of Biochar, Chitosan and ZnO-Based Binary Composites towards a Cationic Dye. Sustainability 2022, 14, 14571. [Google Scholar] [CrossRef]
  12. Isaac, R.; Siddiqui, S. Sequestration of Ni(II) and Cu(II) using FeSO4 modified Zea mays husk magnetic biochar: Isotherm, kinetics, thermodynamic studies and RSM. J. Hazard. Mater. Adv. 2022, 8, 100162. [Google Scholar] [CrossRef]
  13. Pipíška, M.; Krajcíková, E.K.; Hvostik, M.; Frišták, V.; Duriška, L.; Cernicková, I.; Kanuchová, M.; Conte, P.; Soja, G. Biochar from Wood Chips and Corn Cobs for Adsorption of Thioflavin T and Erythrosine B. Materials 2022, 15, 1492. [Google Scholar] [CrossRef] [PubMed]
  14. Xu, Z.; Hu, Y.; Guo, Z.; Xiao, X.; Peng, C.; Zeng, P. Optimizing pyrolysis temperature of contaminated rice straw biochar: Heavy metal(loid) deportment, properties evolution, and Pb adsorption/immobilization. J. Saudi Chem. Soc. 2022, 26, 101439. [Google Scholar] [CrossRef]
  15. Santhosh, C.; Daneshvar, E.; Tripathi, K.M.; Baltrėnas, P.; Kim, T.; Baltrėnaitė, E.; Bhatnagar, A. Synthesis and characterization of magnetic biochar adsorbents for the removal of Cr(VI) and Acid orange 7 dye from aqueous solution. Environ. Sci. Pollut. Res. 2020, 27, 32874–32887. [Google Scholar] [CrossRef] [PubMed]
  16. Chen, X.; Lin, J.; Su, Y.; Tang, S. One-Step Carbonization Synthesis of Magnetic Biochar with 3D Network Structure and Its Application in Organic Pollutant Control. Int. J. Mol. Sci. 2022, 23, 12579. [Google Scholar] [CrossRef] [PubMed]
  17. Ghanbari, N.; Ghafuri, H. Design and preparation the novel polymeric layered double hydroxide nanocomposite (LDH/Polymer) as an efficient and recyclable adsorbent for the removal of methylene blue dye from water. Environ. Technol. Innov. 2022, 26, 102377. [Google Scholar] [CrossRef]
  18. Oladoye, P.O.; Ajiboye, T.O.; Omotola, E.O.; Oyewola, O.J. Methylene blue dye: Toxicity and potential technologies for elimination from (waste)water. Results Eng. 2022, 16, 100678. [Google Scholar] [CrossRef]
  19. Ramutshatsha-Makhwedzha, D.; Mavhungu, A.; Moropeng, M.L.; Mbaya, R. Activated carbon derived from waste orange and lemon peels for the adsorption of methyl orange and methylene blue dyes from wastewater. Heliyon 2022, 8, e09930. [Google Scholar] [CrossRef]
  20. Mu, Y.; Du, H.; He, W.; Ma, H. Functionalized mesoporous magnetic biochar for methylene blue removal: Performance assessment and mechanism exploration. Diamond Rel. Mat. 2022, 121, 108795. [Google Scholar] [CrossRef]
  21. Güleç, F.; Williams, O.; Kostas, E.T.; Samson, A.; Stevens, L.A.; Lester, E. A comprehensive comparative study on methylene blue removal from aqueous solution using biochars produced from rapeseed, whitewood, and seaweed via different thermal conversion technologies. Fuel 2022, 330, 125428. [Google Scholar] [CrossRef]
  22. Zeng, H.; Qi, W.; Zhai, L.; Wang, F.; Zhang, J.; Li, D. Preparation and characterization of sludge-based magnetic biochar by pyrolysis for methylene blue removal. Nanomaterials 2021, 11, 2473. [Google Scholar] [CrossRef] [PubMed]
  23. Li, M.; Dong, C.; Guo, C.; Yu, L. Magnetic Activated Biochar Fe3O4-MOS Made from Moringa Seed Shells for the Adsorption of Methylene Blue. Processes 2022, 10, 2720. [Google Scholar] [CrossRef]
  24. Rodríguez, J.A.; Ríos, J.C.; Flores, S.F.; Treviño, G.G.; Muzquiz, E.M. Síntesis y caracterización de compositos carbón/óxidos de hierro a partir de caña de azúcar. Theorema 2019, 10, 124–129. [Google Scholar]
  25. Lee, X.J.; Lee, L.Y.; Gan, S.; Thangalazhy-Gopakumar, S.; Ng, H.K. Biochar potential evaluation of palm oil wastes through slow pyrolysis: Thermochemical characterization and pyrolytic kinetic studies. Bioresour. Technol. 2017, 236, 155–163. [Google Scholar] [CrossRef]
  26. Xie, R.; Zhu, Y.; Zhang, H.; Zhang, P.; Han, L. Effects and mechanism of pyrolysis temperature on physicochemical properties of corn stalk pellet biochar based on combined characterization approach of microcomputed tomography and chemical analysis. Bioresour. Technol. 2021, 329, 124907. [Google Scholar] [CrossRef]
  27. Qin, L.; Wu, Y.; Hou, Z.; Jiang, E. Influence of biomass components, temperature and pressure on the pyrolysis behavior and biochar properties of pine nut shells. Bioresour. Technol. 2020, 313, 123682. [Google Scholar] [CrossRef] [PubMed]
  28. ASTM. Método de Prueba Estándar Para Contenido Total de Cenizas de Carbón Activado. Available online: https://www.astm.org/d2866-94r04.html (accessed on 8 November 2022).
  29. Al-Rumaihi, A.; Shahbaz, M.; Mckay, G.; Mackey, H.; Al-Ansari, T. A review of pyrolysis technologies and feedstock: A blending approach for plastic and biomass towards optimum biochar yield. Renew. Sustain. Energy Rev. 2022, 167, 112715. [Google Scholar] [CrossRef]
  30. Zhao, N.; Yin, Z.; Liu, F.; Zhang, M.; Lv, Y.; Hao, Z.; Pan, G.; Zhang, J. Environmentally persistent free radicals mediated removal of Cr(VI) from highly saline water by corn straw biochars. Bioresour. Technol. 2018, 260, 294–301. [Google Scholar] [CrossRef] [Green Version]
  31. Mohan, D.; Kumar, S.; Srivastava, A. Fluoride removal from ground water using magnetic and nonmagnetic corn stover biochars. Ecol. Eng. 2014, 73, 798–808. [Google Scholar] [CrossRef]
  32. Amalina, F.; Razak, A.S.A.; Krishnan, S.; Zularisam, A.W.; Nasrullah, M. A comprehensive assessment of the method for producing biochar, its characterization, stability, and potential applications in regenerative economic sustainability—A review. Clean. Mater. 2022, 3, 100045. [Google Scholar] [CrossRef]
  33. Suo, F.; You, X.; Yin, S.; Wu, H.; Zhang, C.; Yu, X.; Sun, R.; Li, Y. Preparation and characterization of biochar derived from co-pyrolysis of Enteromorpha prolifera and corn straw and its potential as a soil amendment. Sci. Total Environ. 2021, 798, 149167. [Google Scholar] [CrossRef] [PubMed]
  34. Muigai, H.H.; Bordoloi, U.; Hussain, R.; Ravi, K.; Moholkar, V.S.; Kalita, P. A comparative study on synthesis and characterization of biochars derived from lignocellulosic biomass for their candidacy in agronomy and energy applications. Int. J. Energy Res. 2021, 45, 4765–4781. [Google Scholar] [CrossRef]
  35. Sahoo, S.S.; Vijay, V.K.; Chandra, R.; Kumar, H. Production and characterization of biochar produced from slow pyrolysis of pigeon pea stalk and bamboo. Clean. Eng. Technol. 2021, 3, 100101. [Google Scholar] [CrossRef]
  36. Gómez-Vásquez, R.; Fernández-Ballesteros, E.; Camargo-Trillos, D. Biogenic nanoporous oxides recovery from by-products of bioenergy production: Rice husks and corncob biochars. Biomass Bioenergy 2022, 161, 106455. [Google Scholar] [CrossRef]
  37. Land and Water Division-FAO. Fertilizantes, su presentación, calidad y etiquetado. In Los Fertilizantes y Su Uso; IFA: Paris, France, 2002; pp. 33–44. [Google Scholar]
  38. Quillope, J.C.C.; Carpio, R.B.; Gatdula, K.M.; Detras, M.C.M.; Doliente, S.S. Optimization of process parameters of self-purging microwave pyrolysis of corn cob for biochar production. Heliyon 2021, 7, e08417. [Google Scholar] [CrossRef] [PubMed]
  39. Leichtweis, J.; Silvestri, S.; Carissimi, E. New composite of pecan nutshells biochar-ZnO for sequential removal of acid red 97 by adsorption and photocatalysis. Biomass Bioenergy 2020, 140, 105648. [Google Scholar] [CrossRef]
  40. Ouyang, D.; Yan, J.; Qian, L.; Chen, Y.; Han, L.; Su, A.; Zhang, W.; Ni, H.; Chen, M. Degradation of 1,4-dioxane by biochar supported nano magnetite particles activating persulfate. Chemosphere 2017, 184, 609–617. [Google Scholar] [CrossRef]
  41. Yan, J.; Yang, L.; Qian, L.; Han, L.; Chen, M. Nano-magnetite supported by biochar pyrolyzed at different temperatures as hydrogen peroxide activator: Synthesis mechanism and the effects on ethylbenzene removal. Environ. Pollut. 2020, 261, 114020. [Google Scholar] [CrossRef]
  42. Navarathna, C.M.; Karunanayake, A.G.; Gunatilake, S.R.; Pittman, C.U.; Perez, F.; Mohan, D.; Mlsna, T. Removal of Arsenic(III) from water using magnetite precipitated onto Douglas fir biochar. J. Environ. Manag. 2019, 250, 109429. [Google Scholar] [CrossRef]
  43. Compeán-Jasso, M.E.; Ruiz, F.; Martínez, J.R.; Herrera-Gómez, A. Magnetic properties of magnetite nanoparticles synthesized by forced hydrolysis. Mater. Lett. 2008, 62, 4248–4250. [Google Scholar] [CrossRef]
  44. Peña-Rodríguez, G.; Rivera-Suárez, P.A.; González-Gómez, C.H.; Parra-Vargas, C.A.; Garzón-Posada, A.O.; Landínez-Téllez, D.A.; Roa-Rojas, J. Efecto de la concentración de magnetita en la estructura, propiedades eléctricas y magnéticas de un material compuesto a base de resina de poliéster. TecnoLógicas 2018, 21, 13–27. [Google Scholar] [CrossRef]
  45. Manoharan, T.; Ganeshalingam, S.; Nadarajah, K. Mechanisms of emerging contaminants removal by novel neem chip biochar. Environ. Adv. 2022, 7, 100158. [Google Scholar] [CrossRef]
  46. Binh, Q.A.; Nguyen, H.H. Investigation the isotherm and kinetics of adsorption mechanism of herbicide 2,4-dichlorophenoxyacetic acid (2,4-D) on corn cob biochar. Bioresour. Technol. Rep. 2020, 11, 100520. [Google Scholar] [CrossRef]
  47. Bekiaris, G.; Peltre, C.; Jensen, L.S.; Bruun, S. Using FTIR-photoacoustic spectroscopy for phosphorus speciation analysis of biochars. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2016, 168, 29–36. [Google Scholar] [CrossRef] [PubMed]
  48. Ríos Hurtado, J.C.; Martínez-Valdés, A.C.; Rangel Méndez, J.R.; Ballesteros Pacheco, J.C.; Múzquiz Ramos, E.M. Facile synthesis and characterization of MnxZn1-xFe2O4/activated carbon composites for biomedical applications. J. Ceram. Sci. Technol. 2016, 7, 289–294. [Google Scholar] [CrossRef]
  49. Parthasarathy, P.; Fernandez, A.; Singh, D.K.; Al-Ansari, T.; Mackey, H.R.; Rodriguez, R.; Mazza, G.; Tirkey, J.V.; McKay, G. Thermogravimetric analysis of camel dung, date stone, and their blend for pyrolytic, kinetic, and thermodynamic studies. Clean. Chem. Eng. 2022, 4, 100072. [Google Scholar] [CrossRef]
  50. Lu, J.; Yang, Y.; Liu, P.; Li, Y.; Huang, F.; Zeng, L.; Liang, Y.; Li, S.; Hou, B. Iron-montmorillonite treated corn straw biochar: Interfacial chemical behavior and stability. Sci. Total Environ. 2020, 708, 134773. [Google Scholar] [CrossRef]
  51. Neeli, S.T.; Ramsurn, H. Synthesis and formation mechanism of iron nanoparticles in graphitized carbon matrices using biochar from biomass model compounds as a support. Carbon 2018, 134, 480–490. [Google Scholar] [CrossRef]
  52. Mubarak, N.M.; Fo, Y.T.; Al-Salim, H.S.; Sahu, J.N.; Abdullah, E.C.; Nizamuddin, S.; Jayakumar, N.S.; Ganesan, P. Removal of Methylene Blue and Orange-G from Waste Water Using Magnetic Biochar. Int. J. Nanosci. 2015, 14, 1550009. [Google Scholar] [CrossRef]
  53. Ruthiraan, M.; Abdullah, E.C.; Mubarak, N.M.; Noraini, M.N. A promising route of magnetic based materials for removal of cadmium and methylene blue from waste water. J. Environ. Chem. Eng. 2017, 5, 1447–1455. [Google Scholar] [CrossRef]
  54. Baytar, O.; Abdullah Ceyhan, A.; Şahin, O. Production of activated carbon from Elaeagnus angustifolia seeds using H3PO4 activator and methylene blue and malachite green adsorption. Int. J. Phytoremed. 2021, 23, 693–703. [Google Scholar] [CrossRef] [PubMed]
  55. Hanaor, D.A.H.; Ghadiri, M.; Chrzanowski, W.; Gan, Y. Scalable surface area characterization by electrokinetic analysis of complex anion adsorption. Langmuir 2014, 30, 15143–15152. [Google Scholar] [CrossRef] [PubMed]
  56. Van Hoa, N.; Minh, N.C.; Cuong, H.N.; Dat, P.A.; Nam, P.V.; Viet, P.H.T.; Phuong, P.T.D.; Trung, T.S. Highly Porous Hydroxyapatite/Graphene Oxide/Chitosan Beads as an Efficient Adsorbent for Dyes and Heavy Metal Ions Removal. Molecules 2021, 26, 6127. [Google Scholar] [CrossRef]
  57. Lee, X.J.; Lee, L.Y.; Hiew, B.Y.Z.; Gan, S.; Thangalazhy-Gopakumar, S.; Ng, H.K. Valorisation of oil palm wastes into high yield and energy content biochars via slow pyrolysis: Multivariate process optimisation and combustion kinetic studies. Mater. Sci. Energy Technol. 2020, 3, 601–610. [Google Scholar] [CrossRef]
  58. Li, Y.; Zhang, Y.; Zhang, Y.; Wang, G.; Li, S.; Han, R.; Wei, W. Reed biochar supported hydroxyapatite nanocomposite: Characterization and reactivity for methylene blue removal from aqueous media. J. Mol. Liq. 2018, 263, 53–63. [Google Scholar] [CrossRef]
  59. Sun, L.; Wan, S.; Luo, W. Biochars prepared from anaerobic digestion residue, palm bark, and eucalyptus for adsorption of cationic methylene blue dye: Characterization, equilibrium, and kinetic studies. Bioresour. Technol. 2013, 140, 406–413. [Google Scholar] [CrossRef]
  60. Wang, J.; Wang, S. Preparation, modification and environmental application of biochar: A review. J. Clean. Prod. 2019, 227, 1002–1022. [Google Scholar] [CrossRef]
  61. Patawat, C.; Silakate, K.; Chuan-Udom, S.; Supanchaiyamat, N.; Hunt, A.J.; Ngernyen, Y. Preparation of activated carbon from Dipterocarpus alatus fruit and its application for methylene blue adsorption. RSC Adv. 2020, 10, 21082–21091. [Google Scholar] [CrossRef]
Figure 1. SEM images of the BCM structure, (a) 400 μm and (b) 100 μm.
Figure 1. SEM images of the BCM structure, (a) 400 μm and (b) 100 μm.
Materials 16 03127 g001
Figure 2. SEM images of the BCMFe structure, (a) 300 μm and (b) 50 μm. The blue circles represents the Fe particles observed in SEM.
Figure 2. SEM images of the BCMFe structure, (a) 300 μm and (b) 50 μm. The blue circles represents the Fe particles observed in SEM.
Materials 16 03127 g002
Figure 3. EDS image of the 4 BCMFe points analyzed. The numbers represents the point where EDS was done for Table 5.
Figure 3. EDS image of the 4 BCMFe points analyzed. The numbers represents the point where EDS was done for Table 5.
Materials 16 03127 g003
Figure 4. BCMFe XRD pattern.
Figure 4. BCMFe XRD pattern.
Materials 16 03127 g004
Figure 5. Magnetic properties of biochar (BCMFe).
Figure 5. Magnetic properties of biochar (BCMFe).
Materials 16 03127 g005
Figure 6. FTIR analysis of BCM and BCMFe.
Figure 6. FTIR analysis of BCM and BCMFe.
Materials 16 03127 g006
Figure 7. BCM and BCMFe biochar thermograms.
Figure 7. BCM and BCMFe biochar thermograms.
Materials 16 03127 g007
Figure 8. Methylene blue adsorption isotherm adjusted to the Langmuir model for the synthesized samples. pH = 7; T = 25 °C; dosage = 0.02 g/10 mL; methylene blue Co = 1–25 ppm.
Figure 8. Methylene blue adsorption isotherm adjusted to the Langmuir model for the synthesized samples. pH = 7; T = 25 °C; dosage = 0.02 g/10 mL; methylene blue Co = 1–25 ppm.
Materials 16 03127 g008
Figure 9. Kinetic PSO model adjust for BCM (a) and BCMFe (b).
Figure 9. Kinetic PSO model adjust for BCM (a) and BCMFe (b).
Materials 16 03127 g009
Table 1. Experimental pyrolysis conditions and performance of BCM.
Table 1. Experimental pyrolysis conditions and performance of BCM.
Sample ConditionN2 Flow Rate (mL/min)Heating Rate (°C/min)Yield (%)Average Yield (%)Error (%)
Raw50104.905.600.117
Raw50106.20
Raw50105.70
Raw351025.3028.480.098
Raw351029.61
Raw351030.54
Cooked35529.2426.910.081
Cooked35526.59
Cooked35524.91
Cooked351022.2722.450.218
Cooked351027.43
Cooked351017.64
Table 2. Ash content of corn cob and BCM and BCMFe biochars.
Table 2. Ash content of corn cob and BCM and BCMFe biochars.
MaterialAsh Content (%)
Corn cob0.4 (±0.04)
BCM4.0 (±0.01)
BCMFe8.0 (±0.02)
Table 3. Comparison of yield and ash percentage for different biochars.
Table 3. Comparison of yield and ash percentage for different biochars.
BiomassTemperature (°C)Yield (%)Ash (%)Reference
Water hyacinth35034.2510.42[34]
Corn straw40031.60NR[33]
Pigeon pea stalk40029.803.08[35]
Corn stalk pellet50033.4120.86[26]
Corn stover50028.216.60[31]
Sugarcane bagasse55021.151.16[34]
Bamboo60027.004.65[35]
Note: Not reported.
Table 4. Samples composition obtained by XRF.
Table 4. Samples composition obtained by XRF.
ElementCorn CobBCMBCMFe
Concentration (%)Error (%)Concentration (%)Error (%)Concentration (%)Error (%)
K1.5480.00311.2690.1409.7180.098
P0.1380.0110.6620.0750.9120.005
Si0.7860.0330.6040.3090.4780.361
Cl0.1740.0030.5550.4141.2360.047
Fe0.5290.0020.2080.0278.7610.152
Table 5. Elemental composition of BCMFe.
Table 5. Elemental composition of BCMFe.
Element% by Mass
Point 1Point 2Point 3Point 4
C28.8444.3355.9155.78
O8.0827.7522.6731.85
Fe46.067.503.764.78
Si14.5815.5312.176.78
K2.444.895.491.80
Total100100100100
Table 6. Langmuir model parameters.
Table 6. Langmuir model parameters.
Materialqm (mg/g)kL (L/mg)RLR2 (%)
BCM23.170.140.22–0.8899.44%
BCMFe39.660.080.33–0.9299.61%
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

Guel-Nájar, N.A.; Rios-Hurtado, J.C.; Muzquiz-Ramos, E.M.; Dávila-Pulido, G.I.; González-Ibarra, A.A.; Pat-Espadas, A.M. Magnetic Biochar Obtained by Chemical Coprecipitation and Pyrolysis of Corn Cob Residues: Characterization and Methylene Blue Adsorption. Materials 2023, 16, 3127. https://doi.org/10.3390/ma16083127

AMA Style

Guel-Nájar NA, Rios-Hurtado JC, Muzquiz-Ramos EM, Dávila-Pulido GI, González-Ibarra AA, Pat-Espadas AM. Magnetic Biochar Obtained by Chemical Coprecipitation and Pyrolysis of Corn Cob Residues: Characterization and Methylene Blue Adsorption. Materials. 2023; 16(8):3127. https://doi.org/10.3390/ma16083127

Chicago/Turabian Style

Guel-Nájar, Norma Araceli, Jorge Carlos Rios-Hurtado, Elia Martha Muzquiz-Ramos, Gloria I. Dávila-Pulido, Adrián A. González-Ibarra, and Aurora M. Pat-Espadas. 2023. "Magnetic Biochar Obtained by Chemical Coprecipitation and Pyrolysis of Corn Cob Residues: Characterization and Methylene Blue Adsorption" Materials 16, no. 8: 3127. https://doi.org/10.3390/ma16083127

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