Next Article in Journal
In-Season Major Crop-Type Identification for US Cropland from Landsat Images Using Crop-Rotation Pattern and Progressive Data Classification
Previous Article in Journal
Implementation of Mitigation Measures to Reduce Phosphorus Losses: The Vestre Vansjø Pilot Catchment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Biomass-Derived Carbonaceous Adsorbents for Trapping Ammonia

1
Centre for Integrated Energy Research, School of Chemical and Process Engineering, University of Leeds, Leeds LS2 9JT, UK
2
School of Civil Engineering, University of Leeds, Leeds LS2 9JT, UK
*
Author to whom correspondence should be addressed.
Submission received: 16 November 2018 / Revised: 3 January 2019 / Accepted: 5 January 2019 / Published: 9 January 2019

Abstract

:
The preparation of low-cost carbonaceous adsorbents for nitrogen recovery is of interest from agricultural and waste management perspectives. In this study, the gaseous ammonia (NH3) and aqueous ammonium (NH4+) sorption capacities have been measured for different types of carbonaceous chars produced under different conditions. The study includes a comparison of an oak-based hydrochar produced from hydrothermal carbonisation (HTC) at 250 °C with two biochars produced from slow pyrolysis at 450 °C and 650 °C, respectively. The chars were also chemically modified with H2SO4, H3PO4, H2O2, and KOH to investigate the potential for sorption enhancement. The highest sorption capacities for NH3 were observed for the hydrochars with typical uptake capacities ranging from 18–28 mg g−1 NH3. Sorption capacity for oak biochars is significantly lower and ranges from 4–8 mg g−1 for biochars produced at 450 °C and 650 °C, respectively. Hydrochar showed a substantially higher sorption capacity for NH3 despite its lower surface area. The CaCl2 extractable NH4+ following ammonia adsorption is incomplete. Typically, only 30–40% of the N is released upon washing with CaCl2 in form of NH4+. Post chemical modification of the chars resulted in only limited enhancement of char NH3 and NH4+ sorption. H3PO4 treatment showed the greatest potential for increasing NH3/NH4+ sorption in biochars, while KOH and H2O2 treatment increased NH3 sorption in the hydrochar. As only marginal increases to char surface area were observed following char treatment, these findings suggest that char surface functionality is more influential than surface area in terms of char NH3/NH4+ sorption.

1. Introduction

Nitrogen species are present in several agricultural and industrial wastewater effluents, where excessive quantities contribute to eutrophication. Furthermore, NH3 emission into the atmosphere can result in nutrient deposition in nutrient sensitive ecosystems, formation of light–scattering aerosols resulting in haze, and visibility impairment and formation of inhalable aerosol particles which have health concerns [1]. Activated carbon has been the standard adsorbent used for recovering a variety of species owing to its well-developed pore structure and high surface area [2,3,4]. However, due to the high costs associated with traditional activated carbon, alternative waste-derived feedstocks have been considered, ranging from agricultural by-products and industrial waste materials [2,3]. Subsequent thermal and/or chemical activation of such feedstocks via incorporation of metal salts or oxidizing agents result in more cost-effective adsorbents [3,5]. To be considered suitable however, such alternative feedstocks are required to be abundant, carbon-rich, and possess appreciable pore development [3].
Hydrochars and biochars show potential as suitable alternatives to traditional activated carbon as they are carbon-rich and can be produced from a wide range of feed-stock although they are likely to require modification to enhance their nutrient sorption capacities [6]. Studies have demonstrated that adsorption using biochar is roughly consistent with Langmuir and Freundlich adsorption isotherm models, although uncertainty in adsorption mechanisms exist due to different pyrolysis reaction conditions and the fact that the presence of Brønsted and/or Lewis acid sites on the adsorbent surfaces enhance NH3 sorption [7,8]. In the former case, protonation of NH3 occurs as NH3 dissociates in water or via acid-base neutralization reactions with carbonyl and phenolic OH groups to form NH4+ complexes [4,5,9,10,11]. Conversely, the donation of lone pair electrons from N atoms to cationic adsorbent sites such as metal species can occur at Lewis acid sites [9,10,12,13]. NH3 adsorption capacities have been studied for various biochars and have reported no direct correlation of pyrolysis reaction temperature or significance of surface area once activated [14,15]. Moreover, NH4+ adsorption capacities have been reported to be greater in biochars produced at low temperatures between 300 to 400 °C, possibly due to their higher proportion of acid functional groups [16,17] including high pH [18]. However, a study also found biochars with higher NH4+ adsorption capacities with increasing pyrolysis temperatures [19]. Though, different feedstock materials were used in the preparation of the biochars examined.
These studies have shown that biochar sorption mechanisms still require further investigation and suggest that hydrochars from the hydrothermal carbonisation (HTC) process are likely to possess greater ability for NH3/NH4+ sorption compared to biochars, given their higher proportion of acidic functional groups. On the other hand, the higher surface areas of biochars may compensate for their lower acidic functionalities since some studies have reported that high adsorbent surface areas and pore volumes increase NH3/NH4+ removal efficiencies [10,19,20]. Moreover, the high surface areas of some biochars may provide more sites for the loading of acidic or cationic species which increase the number of Brønsted and Lewis acid sites respectively.
Studies in the literature directly comparing ammonia sorption of hydrochar and biochar are limited and show mixed results. An emissions and growth study showed that hydrochar lowered NH3 emissions compared to biochar which was attributed to the greater acidity of hydrochar [21]. Contrarily, it was found that hydrochar and biochar increased the emissions of NH3 when applied as a soil amender with the addition of pig slurry. Moreover, hydrochar had a reduced ability to absorb NH4+, which was attributed to its increased hydrophobicity and pH buffering capabilities [22]. Both studies used the same hydrochar and biochar production approach and both used Miscanthus as a feedstock although it is speculated that the production of hydrochar was not performed in subcritical water but was rather steam treated, akin to the autoclaving process.
This study is therefore aimed at contributing to the growing research on factors influencing char NH3 and NH4+ sorption. In particular, understanding the difference in behaviour between hydrochar and biochar and investigating the potential for enhancing sorption using post chemical modification. Owing to their low levels of contaminants, hydrochar and biochars derived from bark-free holm oak wood were evaluated before and after treatment with various chemical agents. It is anticipated that this study will contribute to growing research on the factors influencing nitrogen recovery by adsorbents derived from a wider range of agricultural residues.

2. Materials and Methods

2.1. Materials

Oak hydrochar was produced using a 600 mL stainless steel Parr 4836 bench-top reactor (Parr, Moline, IL, USA) by heating 10 wt% bark-free holm oak wood (particle size < 1 mm) in distilled water at 250 °C for 1 h at autogenous pressure (approximately 4 MPa). The reactor and its contents were allowed to cool to about 50 °C before recovering the resulting hydrochar from the process water by filtration followed by air-drying. The hydrochar was designated as Oak 250. Biochars were supplied and produced from Oak wood (particle size < 2 mm) were obtained from a commercial pyrolysis plant operated by Proininso S.A. (Málaga, Spain) at 450 °C and 650 °C (henceforth referred to as Oak 450 and Oak 650, respectively). Particles were ground and sieved through 1 mm and 2 mm apertures until homogenous.

2.2. Methods

2.2.1. Acid Treatment

For char chemical modification, as-received analytical-grade reagents were used. Chars were prepared to particle size <2 mm then weighed to 4 g and mixed with 2 g of 99% phosphoric acid in 20 mL distilled water. The mixtures were shaken briefly and left to stand for 24 h at room temperature after which the mixtures were heated for about 15 h at 80 °C in an oven and rinsed with distilled water. Hydrochars and biochars were oven-dried at 80 °C and 100 °C, respectively.
For sulphuric acid treatment, 5 g of char (≤2 mm) was mixed with 100 mL of 0.1 M H2SO4 and heated for 3 h at 80 °C after which treated chars were washed with distilled water until a stable pH was attained and oven-dried at 80–100 °C.
H2O2 treatment involved soaking 2 g of char (≤2 mm) in 20 mL of 10% H2O2 for 48 h at room temperature, followed by heating at 80 °C for 24 h and rinsing with distilled water until the pH was between 6–7 for biochars or close to the hydrochars original pH before oven-drying at 80–100 °C.

2.2.2. KOH Treatment

KOH treatment involved mixing 4 g of char (≤2 mm) in a solution of 2 g KOH and 20 mL of distilled water. The mixture was stirred for 2 h at 75 °C with a magnetic stirrer. Treated chars were subsequently rinsed with HCl followed by distilled water until the leachate pH values ranged between 6–7 for biochars or close to the hydrochars original pH, after which treated hydrochars and biochars were oven-dried for 2 h at 80 °C and 100 °C, respectively.

2.2.3. Agronomical Analysis

Ultimate analyses of biochar and hydrochar samples were determined using a CHN Elemental Analyser (Flash 2000, Thermo Scientific, Waltham, MA, USA). Proximate analysis was performed using a muffle furnace and analysis of char pH, CEC, N2 BET, surface area, macro-, and micro-nutrient contents have been described elsewhere [23].

2.2.4. Ammonium Sorption Tests

Details of the batch ammonium adsorption tests have been provided in [24]. Briefly, 0.1 g of char (≤850 µm) is mixed with 100 mL of either 43 mg or 450 mg NH4+ L−1 solution (pH 7) prepared from ammonium chloride salt in deionised water. NH4+ sorption capacities for the various treated chars were evaluated at 450 mg NH4+ L−1 to investigate their potential for NH4+ recovery from a representative concentration range encountered in real-case conditions. The mixture was shaken at 160 rpm for 24 h at room temperature. To determine whether losses in NH4+ occurred during sorption tests, the same procedure was performed without char addition. As only small amounts of NH4+ were lost after sorption tests with such controls, reductions in NH4+ after sorption tests were assumed to be entirely due to char sorption. Char NH4+ sorption capacity was determined as:
q e =   ( C 0 C e ) V M
where C0 and Ce = initial and equilibrium ammonium adsorbate concentrations respectively (mg L−1); V = volume of solution (L); M = mass of char (g).
Batch desorption followed a similar procedure as outlined above, where adsorbed NH4+ was extracted from chars using 0.01 M KCl and analysed by ion chromatography. This procedure was also performed on chars to determine their native NH4+ contents to correct for any NH4+ release from the initial char. Desorbability was determined as the ratio of desorbed NH4+ to total adsorbed NH4+.

2.2.5. Ammonia Sorption Test

Ammonia adsorption was evaluated in a simple batch set-up, slightly modified from the methodology outlined in Taghizadeh-Toosi et al. [11]. Briefly, 60 mL of NaOH was transferred into a 250 mL Schott Duran bottle after which a weighing boat containing about 0.2 g char was carefully placed into the bottle so it floats on top of the solution. The bottle was fitted with a septa and closed to obtain a gas tight seal. Carefully, 25 mL of 0.05 M (NH4)2SO4 was injected into the bottle through the septa and gaseous ammonia was produced via the following chemical reaction:
2 NaOH ( aq ) + ( NH 4 ) 2 SO 4 ( aq ) Na 2 SO 4 ( aq ) + 2 H 2 O ( l ) + 2 NH 3 ( g )
NaOH was present in excess such that a known amount of gaseous NH3 was produced; in the current case this was 43 mg of NH3 gas and average headspace volumes were equivalent to about 242.6 mL. After seven days the samples were removed from the bottles and changes to nitrogen content were analysed via elemental analysis. The nitrogen content was adjusted for increased moisture content with respect to the char samples prior to NH3 sorption. Changes to NH4+ content were also determined by ion chromatography after shaking a mixture of the char in a solution of 0.01 M CaCl2 in a 1:20 ratio for 2 h at 160 rpm. The NH3 sorption capacity of the raw holm oak biomass (from which the biochar and hydrochar were produced) was also evaluated following the same procedure outlined above for comparative purposes, since unprocessed feedstocks are often used as litter material for NH3 abatement in livestock systems. The NH3 sorption capacities of Oak 250 were further investigated at various concentrations of NH3 following the same procedure as outlined above, generating about 450, 1000, and 1500 mg NH3 by adjusting molar concentrations of (NH4)2SO4 and NaOH according to Equation (2).

2.2.6. Assessment of Error and Statistical Analysis

Hydrothermal carbonization experiments were performed until repeatability of hydrochar yields was typically ±3 wt%. Whereas pyrolysis was performed by a commercial plant (Proininso S.A. Málaga, Spain), therefore, yields of biochar were unavailable. Biochar and hydrochar samples were both homogenized prior to use. All absorption tests were performed in duplicate, with selected data being performed in triplicate and products combined for a representative sample for analysis. Metal analysis and ultimate and proximate analysis were performed in duplicate until a maximum standard deviation of ±2% was achieved. Mean values are reported together with standard error ranges and bars in tables and figures. Selected data for sorption of NH3 on modified chars was also analysed using a one-way ANOVA with a significance level (α = 0.05), followed by a Tukey or Dunnett’s test for examining each variable as separate. The value of p < 0.05 was considered significant. Data for statistical analysis can be found in the Supplementary Materials.

3. Results

3.1. Physicochemical Properties of Untreated Hydrochars and Biochars

The unprocessed (as-received) bark-free oak wood possesses a low ash content of 2.6% and appreciable concentrations of Ca, K, Mg, and P. Its carbon content of 50.8% and molar O/C ratio of 0.6 is within the range frequently reported for agricultural feedstocks. Following thermochemical treatment, a reduction in volatile matter content is evident while ash content increases (Figure 1).
Other hydrochar and biochar physicochemical properties have been provided elsewhere [24], in which all carbon contents are >50% in accordance with EBC and IBI product specification guidelines. Previous studies have observed that with increasing pyrolysis temperature, relative carbon contents increase and this is also observed in this study, with Oak 650 exhibiting the highest carbon content and surface area, the latter attributed to loss of volatiles with increasing temperature. Differences between the hydrochar and biochar properties are observed: O/C ratios range from 0.2–0.4 and are lowest for Oak 650. Differences in ash content and surface area are also observed with Oak 250 hydrochar having the lowest ash content (6.2%) and surface area (6 m2 g−1) compared to biochar ash contents and surface areas of 11.7–14.3% and 180–280 m2 g−1, respectively. Oak 250 hydrochar also possesses the highest CEC as expected of a low temperature char [25] although this trend is not followed in the biochars since the CEC of Oak 650 is greater than Oak 450. A similar trend in biochar CEC was observed in Yuan et al. [26] and, to some extent, in Jassal et al. [27].

3.2. NH3/NH4+ Sorption by Untreated Hydrochar and Biochars

The NH3 and NH4+ sorption profiles of untreated oak biochars and hydrochars are presented in Figure 2 for concentrations equating to (a) 43 mg and (b) 450 mg NH3/NH4+. The different conditions are used to assess the influence of the concentration of N species on sorption capacity and to simulate low to high levels of N species. The increase in total N content after exposure to 43 mg NH3 is 11.0 ± 0.4 mg g−1 for oak biomass and between 0.58–15.4 mg g−1 for the resulting chars. These levels are comparable to those reported in the literature; an average increment of 6.7 ± 0.6 mg g−1 as reported by Taghizadeh-Toosi et al. [11].
Figure 2a shows that the % sorption efficiency of NH3 is higher for the hydrochar than for both biochars (45% compared to <10%, respectively), equivalent to 18.8 mg g−1 for the hydrochar and <3 mg g−1 for the biochars (Table 1). It is also worth noting that the performance of the raw biomass is comparable to that of the hydrochar at the lower levels of NH3 (Figure 2a) and is significantly higher than the biochar performance. This is less evident at higher NH3 levels (Figure 2b), however, the sorption efficiency of the raw biomass is still comparable with that of the biochars. NH3 sorption capacities of the raw feedstock and hydrochar are, in fact, comparable in spite of the lower mineral matter concentrations present in the former (Figure 1). While the % sorption is lower with the higher initial NH3/NH4+ concentration, the same general trend is observed for NH3 sorption with Oak 250 hydrochar retaining about 20.4 mg g−1 more NH3 than Oak 650 (Table 1). The sorption capacity is approximately a factor of 10 lower as expected.
Figure 3 illustrates the NH3 sorption capacities for Oak 250 over a range of NH3 concentrations. The analysis was found to be statistically significant F(4,4) = 60.57, p = 0.00086. A Tukey test was further implemented to demonstrate all the treatments were significantly different. The hydrochar shows a change in NH3 sorption with increasing NH3 concentration indicating that while a greater capacity is attained at higher initial NH3 concentration, the % sorption capacity reduces substantially and suggests that while sites are available for NH3 sorption even at 1500 mg NH3, its removal efficiency is greatly reduced. NH3 sorption appears to start levelling off at higher concentration suggesting a maximum sorption in the order of about 45 mg g−1 NH3 for the hydrochar.
Subsequent washing of the chars in CaCl2 does not release the entire adsorbed N (Table 1). Typically, 30–40% of the N is released upon washing with CaCl2 in the form of NH4+. As observed by Taghizadeh-Toosi et al. [11,14], no changes in NO3 contents are observed in oak hydrochar or biochars following ammonia sorption tests. CaCl2-extractable ammonium contents are highest for the hydrochar and generally decrease in proportion to ammonia gas concentration exposure. Following exposure to 1000 and 1500 mg gaseous NH3, only 6.7 ± 0.3 and 8.6 ± 2.0 mg of NH4+ are respectively recoverable per gram of Oak 250 hydrochar. Taghizadeh-Toosi et al. [11] similarly observed that 2 M KCl-extractable NH4+ represented only a fraction of the increase in char total N following exposure to NH3. As CaCl2-extractable ammonium is considered plant-available [28,29], these results suggest that hydrochar may be the most beneficial in terms of release of N in soils or soil amendment products. Further investigations are required to confirm whether more ammonium is recoverable with serial extractions using 2 M KCl and water [30], although N species other than inorganic N may be present in post-NH3 sorption chars [14] which are not easily recoverable with such extractants. For instance, Petit et al. [31] showed that ammonia retention in the form of amines is also possible. Studies by Thorn and Mikita [32] on ammonia fixation by organic matter, suggested that organic matter rich in phenolic and hydroxyl groups are capable of retaining ammonia in the form of heterocyclic nitrogen (indole, pyrrole, pyridine).
The NH4+ sorption profiles for the hydrochar and biochars are also presented in Figure 2a,b at initial concentrations (C0) of 43 and 450 mg NH4+ L−1, respectively. Char NH4+ removal efficiencies range from 15.1–23.3% at 43 mg L−1 and decrease to 6.6–10% at 450 mg L−1. The trends are less clear for the sorption of NH4+. There is less distinction between the different samples for NH4+ sorption and there is little statistical difference between the raw biomass and the different chars. At lower levels of NH4+ (Figure 2a), Oak 250 hydrochar demonstrates a lower capacity for NH4+ sorption compared to NH3 whereas for the biochars, the opposite is true. For higher levels of NH4+ (Figure 2b), the hydrochar sorption is comparable to that of the biochars. In both cases, the capacity of the chars for sorption of NH4+ is higher than for NH3. This suggests that surface functionality is an important consideration in retention of NH3/NH4+. While the biochars have a higher surface area, this does not equate to a higher NH3 adsorption compared to the hydrochar, indicating that functionality is more important than surface area for NH3. The levels of NH4+ sorption by the biochars increases over three-fold compared to NH3. This is thought to be related to the presence of mineral matter in the chars which is enhanced in the higher temperature biochars (Figure 1). The levels of inorganics in the chars are often attributed to higher levels of sorption. Zhang and Wang [33] recovered >60% NH4+ partly due to the biochar composites being enhanced with Mg and PIn this study, the higher temperature biochars contain higher levels of ash than the raw biomass and hydrochars, however, the raw biomass shows a significant level of sorption capacity compared to the hydrochars and biochars. Detailed analysis of the inorganic content of the chars and raw feedstock indicates there is no obvious correlation between % NH4+ sorption and char ash, P, Mg, or Ca content. NH4+ release from chars are also low following desorption tests.
The levels of NH4+ sorption in this study are comparable to reports in the literature for biochar. Wang et al. reported 12.1–16.1% removal efficiencies of untreated and La-treated 300 °C biochars (C0 = 25.7 mg NH4+ L−1) [25]. Other studies have reported higher NH4-N removal efficiencies however: 37.3% sorption by activated phytoremediation plant char (C0 = 39 mg NH4+ L−1) in Zeng et al. [34] and 62–83% sorption by Brewer’s grains-sewage sludge biochar (C0 = 100 mg NH4+ L−1) in Zhang and Wang [33]. 0.01 M KCl was used to determine the readily recoverable NH4+ post-sorption from the chars and a selected number were also extracted with 0.01 M CaCl2 to compare performance of both extractants. NH4+ desorption was found to be <10 mg g−1 in both cases thus the desorbability ratio ranged between a factor of 0.2–0.4.
Previous studies have reported a positive relationship between NH4+ sorption and char CEC, surface area, oxygen content and metal content [9,12,33,34,35]. In this study, NH4+ sorption followed a similar trend to CEC, yet the amounts of NH4+ retained by the chars at both 43 and 450 mg NH4+ L−1 are substantially higher than predicted by their corresponding CEC values, recalculated as NH4+eq to estimate maximum theoretical NH4+ sorption capacities (Table 2). Some studies have also observed that CEC underestimates NH4+ sorption capacity [14,27,34,36]; this discrepancy has been attributed to physical sorption of N into the char pores as opposed to ion exchange [27], or to conversion of NH4+ into organic nitrogen [36]. Conversely, Nguyen and Tanner [37] observed maximum NH4+ sorption by a zeolite to be substantially lower than the CEC-predicted values.

3.3. Physicochemical Properties of Modified Hydrochars and Biochars

Table 2 lists the elemental content and CEC of the modified chars and indicates a decrease in carbon and nitrogen content for all the treatments of Oak 250 hydrochar and Oak 650, and a corresponding increase in oxygen content. The carbon content for the Oak 450 chars increase possibly due to the relative decrease of inorganic content following treatment [34,35] while oxygen contents decrease. These findings are in general agreement with Zheng et al. who observed increases in carbon, hydrogen, and nitrogen for acid-washed biochars with a corresponding decrease in oxygen content [38]. This anomaly could be related to the relative ease of leaching of inorganics and the difference in reactivity of the surface towards decarboxylation and hydrolysis. Similar observations were reported by Latham et al. after acid and base treatment during hydrothermal treatment of sucrose [39]. The nitrogen contents of the untreated chars are presented in Table S1 and range from 4.3–5.5 mg N g−1 char. The N contents of the treated chars generally decrease with the exception of KOH treated biochars, which result in slight increases in N (for example an increase in N d.3–6.3 mg N g−1 for Oak 450 treated in KOH and 5.5–7.0 mg N g−1 for Oak 650 treated in KOH). The relative increase in N content following KOH treatment is speculated to be primarily due to loss of carbon, while the reduction in N is possibly due to modification disrupting N-containing compounds in the biomass thus increasing N release.
Acid treatment has varying effects on char CEC: the CEC of the treated hydrochars are generally higher than the biochars, due to increased surface functionality. There is little change in CEC for the hydrochar treated with H2SO4 and H3PO4, whereas all other treatments increase char CEC. For instance, following H2O2 treatment, a two-fold increase in the CEC of Oak 450 is observed although the increase for Oak 250 hydrochar is less. Previous studies [40,41] have attributed this effect to an increase in oxygen functional groups resulting from the oxidation of aromatic carbon and an increase in carbonyl groups. What is also apparent is that the increase in CEC cannot be attributed to changes in surface area since H2O2 treatment results in a drastic reduction in surface area (>50%) for both oak biochars [24].
Changes in functionality following char modification has been investigated using FTIR. The FTIR spectra can identify changes in the relative abundance of carbonyl, hydroxyl, and other reactive functional groups as shown in the Supplementary Material (Figure S1, Table S2). However, no major differences in functional group intensities are observed, suggesting that chemical treatment does not alter char functionality substantially. A slight relative increase in the intensity of carbonyl functional groups (1700 cm−1) is observed following H2O2 treatment.

3.4. NH3/NH4+ Sorption by Treated Hydrochar and Biochars

The NH3 sorption capacities of the treated chars are shown in Figure 4 using a fixed concentration of 43 mg of NH3. The NH3 sorption for Oak 250 hydrochar and modified hydrochar are shown in Figure 4a, for Oak 450 biochar in Figure 4b and for Oak 650 biochar in Figure 4c. The results show that NH3 sorption is higher for the hydrochar than for the two biochar samples, in agreement with initial testing. There was a significant difference between the means of the different treatments F(4,5) = 21.43, p = 0.0024. Treatment with H2O2 and KOH appear to have the largest impact on increasing NH3 sorption in Oak 250 hydrochar whereas, treatment with H2SO4 and H3PO4 have little impact on sorption capacity (Figure 4a). The individual evaluation with Dunnett’s test demonstrated that the treatments with H2SO4 and H3PO4, do not exhibit a significant effect, whereas the treatments with H2O2 and KOH do exhibit a significant effect. Secondly, in the case of Oak 450 there was a significant difference between the means of the different treatments F(4,5) = 12.10, p = 0.0088. Dunnett’s test demonstrates that all the treatments (H2SO4 and H3PO4, H2O2, and KOH) exhibit a significant effect. Modification of Oak 450 with acid and peroxide appear to enhance sorption capacity resulting in a three-fold increase in NH3 sorption from about 6.8% to 20.5% NH3 removal efficiency. Modification of Oak 650 produces negligible enhancement of sorption capacity with the exception of H3PO4 treatment which shows a four-fold enhancement. Thirdly, in the case of Oak 650 there was no significant difference between the means of the different treatments F(4,5) = 1.88, p = 0.2522. Dunnett’s test here demonstrates that none of the treatments exhibit a significant effect.
Increases in sorption capacity following acid treatment are expected and supported in the literature. The increase in NH3 sorption of H2SO4-treated biochars is in agreement with the findings of Petit et al. who showed that sulfur-containing groups may be as influential as oxygen functional groups for NH3 sorption [31]. Similar increases to NH3 sorption have been reported by Chou et al. and Ritz et al. [42,43]. Such increases likely result from the formation of recoverable ammonium sulphate salts on adsorbent sites possibly following similar reactions to those outlined in Chou et al. [42]. It is also likely that increasing H2SO4 concentration up to a certain point could enhance NH3 sorption capacity, based on findings of Chou et al. [42]. The increase in NH3 sorption of H3PO4 treated biochars is also in agreement with the findings of Oya and Iu. Acid-catalyzed hydrolysis of polymeric units in lignocellulosic materials and labile carbon in biochars result in ketone formation after H3PO4 treatment [44,45]. Additionally, the formation of ammonium complexes with H3PO4 may have occurred as observed by Oya and Iu whose extensive studies confirmed the presence of NH4H2PO4 and/or (NH4)2HPO4 depending on H3PO4 loading ratio employed [46]. H2O2 treatment has a marked effect on NH3 sorption for the lower temperature chars with H2O2-treated Oak 250 and Oak 450 showing an increase compared to the control. Similarly, H2O2 modification performed by Gómez-Serrano et al. resulted in an increase in surface OH groups and with increasing treatment time, an increase in carbonyl groups [47]. Increases in NH3 sorption capacity following H2O2 treatment may therefore occur due to interactions between NH3 and additional carbonyl groups. Similarly, treatment of biochar in KOH can also increase the level of surface OH groups (Figure S1) which may account for the increases in NH3 sorption by KOH treated Oak 250 and 450 chars.
Figure 5 shows the NH4+ sorption following modification. In general, modification of the chars results in either no difference in sorption capacity or a reduction in NH4+ sorption capacity for all the chars investigated. This reduction in sorption capacity is particularly apparent for treatment with H2O2 and KOH treatment. There is some evidence of enhancement of sorption capacity using acid treatment, however when investigated further using ANOVA tests, there is no significant difference compared to the control. For all modified chars, the release of NH4+ is also measured following extraction in 0.01 M KCl. As observed for untreated chars, NH4+ release from modified chars are also low following desorption tests. For instance, KOH treated Oak 450 and 650 released 8.5 and 9.7 mg g−1, respectively; H2O2 treated Oak 450 and Oak 650 released 9.4 and 6.7 mg g−1, respectively. As blank sorption tests performed alongside the sample char sorption tests show negligible NH4+ losses, it is speculated that NH4+ release from chars is low either due to strong interactions between NH4+ and the chars.
While treatment in H2O2 and KOH generally show the highest increase in CEC (Table 2), this is not reflected in sorption experiments. All oak chars experience a reduction in NH4+ sorption following KOH and H2O2 treatment despite the substantial increases in char CEC (Figure 5). Huff et al. similarly noted that while H2O2 treatment increased CEC, methylene blue adsorption capacity decreased, which emphasises that CEC may not be a suitable predictor of char sorption capacity [40].

4. Conclusions

Results from this study show that oak hydrochars possess higher sorption capacities for gaseous NH3 and aqueous NH4+ relative to their corresponding biochars. Post chemical modification of hydrochars and biochars has only limited effect on enhancing sorption capacities. Enhancement is more evident for NH3 sorption capacity than for NH4+ sorption capacity which remains similar and in some cases is reduced compared to the control. Overall, the findings show that char surface functionality is more influential than surface area, thus, modification processes aimed at modifying the former are more beneficial in terms of char NH3 and NH4+ sorption. Of the chemical treatments performed in this study, phosphoric acid appears to show the greatest potential for increasing both NH4+ and NH3 sorption capacities of biochars. Treatment of hydrochars in H2O2 and KOH show potential for enhancement of NH3 sorption but result in a significant drop in NH4+ sorption.
The ammonia batch set-up used in this study may be regarded as a useful tool for screening potential NH3 and NH4+ adsorbents although benefits could also arise from the comparison of adsorbent performance in dynamic test conditions or in batch set-ups where humidity is better controlled.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2077-0472/9/1/16/s1, Figure S1. FTIR spectra of treated and untreated (a) OAK 250 (b) OAK 450, and (c) OAK 650 chars including unprocessed oak biomass (as received) (4000–500 cm−1; arbitrary y-axis values); Table S1. Char nitrogen content before and after exposure to NH3; Table S2. ATR-FTIR functional group assignment of prominent bands, Tables S3–S6: ANOVA analysis for Figure 3 and Figure 4a–c, respectively.

Author Contributions

The authors contributed to this article in the following ways: Conceptualization: C.A.T., L.A.F., and A.B.R.; formal analysis: C.A.T., L.A.F., and A.B.R.; investigation: C.A.T.; supervision: L.A.F. and A.B.R.; visualization: C.A.T. and K.R.P.; writing—original draft: C.A.T.; writing—review and editing: K.R.P.

Funding

The authors would like to thank the EPSRC Centre for Doctoral Training in Bioenergy at the University of Leeds (EP/L014912/1) for financial support of Kiran Parmar. The Petroleum Technology Development Fund (PTDF) Nigeria for the financial support of Chibi Takaya and the Fertiplus Consortium (grant agreement no. 289853), co-funded by the European Commission, Directorate General for Research and Innovation, within the 7th Framework Programme of RTD, Theme 2—Biotechnologies, Agriculture and Food for the supply of biochar. The views and opinions expressed in this study are purely those of the authors and may not in any circumstances be regarded as stating an official position of the European Commission.

Acknowledgments

The authors wish to thank Surjit Singh and Patrick Biller for their analytical assistance and Simon Lloyd, Adrian Cunliffe, David Elliot, and Sheena Bennet for their technical assistance, including Jessica Quintana-Najera for the statistical analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Arogo, J.; Westerman, P.W.; Heber, A.J. A review of ammonia emissions from confined swine feeding operations. Trans. Am. Soc. Agric. Eng. 2003, 46, 805. [Google Scholar] [CrossRef]
  2. Kastner, J.R.; Miller, J.; Das, K.C. Pyrolysis conditions and ozone oxidation effects on ammonia adsorption in biomass generated chars. J. Hazard. Mater. 2009, 164, 1420–1427. [Google Scholar] [CrossRef] [PubMed]
  3. Pollard, S.J.T.; Fowler, G.D.; Sollars, C.J.; Perry, R. Low-cost adsorbents for waste and wastewater treatment: A review. Sci. Total Environ. 1992, 116, 31–52. [Google Scholar] [CrossRef]
  4. Steiner, C.; Das, K.C.; Melear, N.; Lakly, D. Reducing Nitrogen Loss during Poultry Litter Composting Using Biochar. J. Environ. Qual. 2010, 39, 1236–1242. [Google Scholar] [CrossRef] [PubMed]
  5. Corre, Y.; Seredych, M.; Bandosz, T.J. Analysis of the chemical and physical factors affecting reactive adsorption of ammonia on graphene/nanoporous carbon composites. Carbon N. Y. 2013, 55, 176–184. [Google Scholar] [CrossRef]
  6. Tran, H.N.; You, S.J.; Hosseini-Bandegharaei, A.; Chao, H.P. Mistakes and inconsistencies regarding adsorption of contaminants from aqueous solutions: A critical review. Water Res. 2017, 120, 88–116. [Google Scholar] [CrossRef] [PubMed]
  7. Liang, P.; Yu, H.; Huang, J.; Zhang, Y.; Cao, H. The review on adsorption and removing ammonia nitrogen with biochar on its mechanism. MATEC Web Conf. 2016, 67, 07006. [Google Scholar] [CrossRef]
  8. Zhang, S.; Yang, X.; Liu, L.; Ju, M.; Zheng, K. Adsorption behavior of selective recognition functionalized biochar to Cd(II) in wastewater. Materials (Basel) 2018, 11, 299. [Google Scholar] [CrossRef] [PubMed]
  9. Le Leuch, L.M.; Bandosz, T.J. The role of water and surface acidity on the reactive adsorption of ammonia on modified activated carbons. Carbon N. Y. 2007, 45, 568–578. [Google Scholar] [CrossRef]
  10. Petit, C.; Bandosz, T.J. Role of surface heterogeneity in the removal of ammonia from air on micro/mesoporous activated carbons modified with molybdenum and tungtsen oxides. Microporous Mesoporous Mater. 2009, 118, 61–67. [Google Scholar] [CrossRef]
  11. Taghizadeh-Toosi, A.; Clough, T.J.; Sherlock, R.R.; Condron, L.M. Biochar adsorbed ammonia is bioavailable. Plant Soil 2012, 350, 57–69. [Google Scholar] [CrossRef]
  12. Canals-Batlle, C.; Ros, A.; Lillo-Ródenas, M.A.; Fuente, E.; Montes-Morán, M.A.; Martin, M.J.; Linares-Solano, A. Carbonaceous adsorbents for NH3 removal at room temperature. Carbon N. Y. 2008, 46, 176–178. [Google Scholar] [CrossRef]
  13. Yin, X.; Han, H.; Gunji, I.; Endou, A.; Cheettu Ammal, S.S.; Kubo, M.; Miyamoto, A. NH3 Adsorption on the Brönsted and Lewis Acid Sites of V2O5 (010): A Periodic Density Functional Study. J. Phys. Chem. B 1999, 103, 4701–4706. [Google Scholar] [CrossRef]
  14. Taghizadeh-Toosi, A.; Clough, T.J.; Sherlock, R.R.; Condron, L.M. A wood based low-temperature biochar captures NH3-N generated from ruminant urine-N, retaining its bioavailability. Plant Soil 2012, 353, 73–84. [Google Scholar] [CrossRef]
  15. Ro, K.; Lima, I.; Reddy, G.; Jackson, M.; Gao, B. Removing Gaseous NH3 Using Biochar as an Adsorbent. Agriculture 2015, 5, 991–1002. [Google Scholar] [CrossRef]
  16. Wang, Z.; Guo, H.; Shen, F.; Yang, G.; Zhang, Y.; Zeng, Y.; Wang, L.; Xiao, H.; Deng, S. Biochar produced from oak sawdust by Lanthanum (La)-involved pyrolysis for adsorption of ammonium (NH4+), nitrate (NO3-), and phosphate (PO43−). Chemosphere 2015, 119, 646–653. [Google Scholar] [CrossRef]
  17. Zheng, W.; Sharma, B.K.; Rajagopalan, N. Using Biochar as a Soil Amendment for Sustainable Agriculture. 2010. Available online: http://hdl.handle.net/2142/25503 (accessed on 10 September 2018).
  18. Gao, F.; Xue, Y.; Deng, P.; Cheng, X.; Yang, K. Removal of aqueous ammonium by biochars derived from agricultural residuals at different pyrolysis temperatures. Chem. Speciat. Bioavailab. 2015, 27, 92–97. [Google Scholar] [CrossRef] [Green Version]
  19. Zheng, X.; Yang, Z.; Xu, X.; Dai, M.; Guo, R. Characterization and ammonia adsorption of biochar prepared from distillers’ grains anaerobic digestion residue with different pyrolysis temperatures. J. Chem. Technol. Biotechnol. 2018, 93, 198–206. [Google Scholar] [CrossRef]
  20. Ismadji, S.; Tong, D.S.; Soetaredjo, F.E.; Ayucitra, A.; Yu, W.H.; Zhou, C.H. Bentonite hydrochar composite for removal of ammonium from Koi fish tank. Appl. Clay Sci. 2016, 119, 146–154. [Google Scholar] [CrossRef] [Green Version]
  21. Schimmelpfennig, S.; Müller, C.; Grünhage, L.; Koch, C.; Kammann, C. Biochar, hydrochar and uncarbonized feedstock application to permanent grassland-Effects on greenhouse gas emissions and plant growth. Agric. Ecosyst. Environ. 2014, 191, 39–52. [Google Scholar] [CrossRef]
  22. Subedi, R.; Kammann, C.; Pelissetti, S.; Taupe, N.; Bertora, C.; Monaco, S.; Grignani, C. Does soil amended with biochar and hydrochar reduce ammonia emissions following the application of pig slurry? Eur. J. Soil Sci. 2015, 66, 1044–1053. [Google Scholar] [CrossRef]
  23. Takaya, C.A.; Fletcher, L.A.; Singh, S.; Okwuosa, U.C.; Ross, A.B. Recovery of phosphate with chemically modified biochars. J. Environ. Chem. Eng. 2016, 4, 1156–1165. [Google Scholar] [CrossRef] [Green Version]
  24. Takaya, C.A.; Fletcher, L.A.; Singh, S.; Anyikude, K.U.; Ross, A.B. Phosphate and ammonium sorption capacity of biochar and hydrochar from different wastes. Chemosphere 2016, 145, 518–527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Tan, X.; Liu, Y.; Zeng, G.; Wang, X.; Hu, X.; Gu, Y.; Yang, Z. Application of biochar for the removal of pollutants from aqueous solutions. Chemosphere 2015, 125, 70–85. [Google Scholar] [CrossRef] [PubMed]
  26. Yuan, J.H.; Xu, R.K.; Zhang, H. The forms of alkalis in the biochar produced from crop residues at different temperatures. Bioresour. Technol. 2011, 102, 3488–3497. [Google Scholar] [CrossRef] [PubMed]
  27. Jassal, R.S.; Johnson, M.S.; Molodovskaya, M.; Black, T.A.; Jollymore, A.; Sveinson, K. Nitrogen enrichment potential of biochar in relation to pyrolysis temperature and feedstock quality. J. Environ. Manag. 2015, 152, 140–144. [Google Scholar] [CrossRef]
  28. Houba, V.J.G.; Novozamsky, I.; Huybregts, A.W.M.; van der Lee, J.J. Comparison of soil extractions by 0.01 M CaCl2, by EUF and by some conventional extraction procedures. Plant Soil 1986, 96, 433–437. [Google Scholar] [CrossRef]
  29. Lazányi, J.; Loch, J. Evaluation of 0.01 M CaCl2 Extractable Nitrogen Forms in a Long-term Experiment. Agrokémia és Talajt. 2006, 55, 135–144. [Google Scholar] [CrossRef]
  30. Wang, B.; Lehmann, J.; Hanley, K.; Hestrin, R.; Enders, A. Adsorption and desorption of ammonium by maple wood biochar as a function of oxidation and pH. Chemosphere 2015, 138, 120–126. [Google Scholar] [CrossRef] [Green Version]
  31. Petit, C.; Kante, K.; Bandosz, T.J. The role of sulfur-containing groups in ammonia retention on activated carbons. Carbon N. Y. 2010, 48, 654–667. [Google Scholar] [CrossRef]
  32. Thorn, K.A.; Mikita, M.A. Ammonia fixation by humic substances: A nitrogen-15 and carbon-13 NMR study. Sci. Total Environ. 1992, 113, 67–87. [Google Scholar] [CrossRef]
  33. Zhang, J.; Wang, Q. Sustainable mechanisms of biochar derived from brewers’ spent grain and sewage sludge for ammonia-nitrogen capture. J. Clean. Prod. 2016, 112, 3927–3934. [Google Scholar] [CrossRef]
  34. Zeng, Z.; Zhang, S.; Li, T.; Zhao, F.; He, Z.; Zhao, H.; Yang, X.; Wang, H.; Zhao, J.; Rafiq, M.T. Sorption of ammonium and phosphate from aqueous solution by biochar derived from phytoremediation plants. J. Zhejiang Univ. Sci. B 2013, 14, 1152–1161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Gai, X.; Wang, H.; Liu, J.; Zhai, L.; Liu, S.; Ren, T.; Liu, H. Effects of feedstock and pyrolysis temperature on biochar adsorption of ammonium and nitrate. PLoS ONE 2014, 9, e113888. [Google Scholar] [CrossRef]
  36. Bolan, N.S.; Saggar, S.; Luo, J.; Bhandral, R.; Singh, J. Gaseous Emissions of Nitrogen from Grazed Pastures: Processes, Measurements and Modelling, Environmental Implications, and Mitigation. Adv. Agron. 2004, 84, 120. [Google Scholar]
  37. Nguyen, M.L.; Tanner, C.C. Ammonium removal from wastewaters using natural New Zealand zeolites. N. Z. J. Agric. Res. 1998, 41, 427–446. [Google Scholar] [CrossRef] [Green Version]
  38. Zheng, H.; Wang, Z.; Zhao, J.; Herbert, S.; Xing, B. Sorption of antibiotic sulfamethoxazole varies with biochars produced at different temperatures. Environ. Pollut. 2013, 181, 60–67. [Google Scholar] [CrossRef]
  39. Latham, K.G.; Jambu, G.; Joseph, S.D.; Donne, S.W. Nitrogen doping of hydrochars produced hydrothermal treatment of sucrose in H2O, H2SO4, and NaOH. ACS Sustain. Chem. Eng. 2014, 2, 755–764. [Google Scholar] [CrossRef]
  40. Huff, M.D.; Lee, J.W. Biochar-surface oxygenation with hydrogen peroxide. J. Environ. Manag. 2016, 165, 17–21. [Google Scholar] [CrossRef]
  41. Liang, B.; Lehmann, J.; Solomon, D.; Kinyangi, J.; Grossman, J.; O’Neill, B.; Skjemstad, J.O.; Thies, J.; Luizão, F.J.; Petersen, J.; et al. Black Carbon Increases Cation Exchange Capacity in Soils. Soil Sci. Soc. Am. J. 2006, 70, 1719–1730. [Google Scholar] [CrossRef]
  42. Chou, L.H.; Tsai, R.I.; Chang, J.R.; Lee, M.T. Regenerable adsorbent for removing ammonia evolved from anaerobic reaction of animal urine. J. Environ. Sci. (China) 2006, 18, 1176–1181. [Google Scholar] [CrossRef]
  43. Ritz, C.W.; Tasistro, A.S.; Kissel, D.E.; Fairchild, B.D. Evaluation of surface-applied char on the reduction of ammonia volatilization from broiler litter. J. Appl. Poult. Res. 2011, 20, 240–245. [Google Scholar] [CrossRef] [Green Version]
  44. Lin, Y.; Munroe, P.; Joseph, S.; Henderson, R.; Ziolkowski, A. Water extractable organic carbon in untreated and chemical treated biochars. Chemosphere 2012, 87, 151–157. [Google Scholar] [CrossRef] [PubMed]
  45. Marsh, H.; Rodríguez-Reinoso, F. Chapter 5—Activation Processes (Thermal or Physical). Activated Carbon; Elsevier Science Ltd.: Oxford, UK, 2006; pp. 243–321. [Google Scholar]
  46. Oya, A.; Iu, W.G. Deodorization performance of charcoal particles loaded with orthophosphoric acid against ammonia and trimethylamine. Carbon N. Y. 2002, 40, 1391–1399. [Google Scholar] [CrossRef]
  47. Gómez-Serrano, V.; Acedo-Ramos, M.; López-Peinado, A.J.; Valenzuela-Calahorro, C. Oxidation of activated carbon by hydrogen peroxide. Study of surface functional groups by FT-ir. Fuel 1994, 73, 387–395. [Google Scholar] [CrossRef]
Figure 1. Distribution of the main micro- and macro-mineral contents within oak biomass, hydrochar, and biochars (mg/kg, dry basis). Inset: Ash and volatile content of oak biomass and chars (%wt., as received basis). Based on duplicate data with repeatability typically within ±2%. Error bars represent deviation around the mean.
Figure 1. Distribution of the main micro- and macro-mineral contents within oak biomass, hydrochar, and biochars (mg/kg, dry basis). Inset: Ash and volatile content of oak biomass and chars (%wt., as received basis). Based on duplicate data with repeatability typically within ±2%. Error bars represent deviation around the mean.
Agriculture 09 00016 g001
Figure 2. NH3 and NH4+ removal efficiencies of unprocessed oak biomass (as received) and oak chars at approximately (a) 43 mg NH3 and 43 mg NH4+ L−1 (b) 450 mg NH3 and NH4+ L−1. Based on duplicate data and error bars represent deviation around the mean.
Figure 2. NH3 and NH4+ removal efficiencies of unprocessed oak biomass (as received) and oak chars at approximately (a) 43 mg NH3 and 43 mg NH4+ L−1 (b) 450 mg NH3 and NH4+ L−1. Based on duplicate data and error bars represent deviation around the mean.
Agriculture 09 00016 g002
Figure 3. NH3 sorption capacities for Oak 250 hydrochar at 43, 450, 1000, and 1500 mg NH3 gas generated according to Equation (2). Error bars represent deviation around the mean. For statistical analysis please refer to Supplementary Material Table S3.
Figure 3. NH3 sorption capacities for Oak 250 hydrochar at 43, 450, 1000, and 1500 mg NH3 gas generated according to Equation (2). Error bars represent deviation around the mean. For statistical analysis please refer to Supplementary Material Table S3.
Agriculture 09 00016 g003
Figure 4. NH3 sorption by (a) treated oak 250 °C hydrochar; (b) treated oak 450 °C biochar; and (c) treated oak 650 °C biochar d3 mg NH3 gas generated according to Equation (2). Error bars represent deviation around the mean. For statistical analysis please refer to Supplementary Material Tables S4–S6.
Figure 4. NH3 sorption by (a) treated oak 250 °C hydrochar; (b) treated oak 450 °C biochar; and (c) treated oak 650 °C biochar d3 mg NH3 gas generated according to Equation (2). Error bars represent deviation around the mean. For statistical analysis please refer to Supplementary Material Tables S4–S6.
Agriculture 09 00016 g004
Figure 5. NH4+ sorption by (a) treated Oak 250 hydrochar; (b) treated Oak 450 biochar; (c) treated Oak 650 biochar from 450 mg NH4+ L−1 solution. Based on duplicate data, error bars represent deviation around the mean.
Figure 5. NH4+ sorption by (a) treated Oak 250 hydrochar; (b) treated Oak 450 biochar; (c) treated Oak 650 biochar from 450 mg NH4+ L−1 solution. Based on duplicate data, error bars represent deviation around the mean.
Agriculture 09 00016 g005
Table 1. Char sorption and release profiles for gaseous NH3 and aqueous NH4+.
Table 1. Char sorption and release profiles for gaseous NH3 and aqueous NH4+.
CharNH3 Sorption (mg g−1)CaCl2-Extractable NH4+ (mg g−1)NH4+ Sorption (mg g−1)NH4+ Desorbed (mg g−1)
a 43 mg NH343 mg NH4+ L−1
OAK 25018.8 ± 1.67.2 ± 0.36.0 ± 0.50.0 ± 0.0
OAK 4502.9 ± 1.41 b9.2 ± 0.21.1 ± 1.6
OAK 6500.7 ± 0.91 b8.9 ± 0.80.0 ± 0.0
a 450 mg NH3450 mg NH4+ L−1
OAK 25028.5 ± 0.36.1 ± 0.845.1 ± 1.79 a
OAK 4504.3 ± 1.22.68 ± 0.0429.9 ± 1.812 a
OAK 6508.1 ± 2.22.1 ± 0.132.7 ± 6.711 a
a 43 and 450 mg NH3 generated based on Equation (2); b single analyses reported while other values are reported as mean uptake ± standard deviation; undetected NH4+ concentrations reported as zero.
Table 2. Elemental content and CEC values for a selection of treated chars.
Table 2. Elemental content and CEC values for a selection of treated chars.
CharCECb NH4+eqCHNSa O
(cmolc kg−1)(mg g−1)(%)(%)(%)(%)(%)
OAK 25088.3 ± 9.715.967.96.51.40.124.1
OAK250-H2SO481.1 ± 4.714.661.34.90.40.0433.4
OAK 250-H3PO486.8 ± 4.715.660.24.80.3034.6
OAK 250-H2O2102.4 ± 7.318.456.14.50.3039.1
OAK 250-KOH102.0 ± 8.218.4624.90.4032.7
OAK 45059.4 ± 8.110.765.72.70.6031
OAK 450-H2SO466.3 ± 15.611.973.64.80.5021.1
OAK 450-H3PO4103.5 ± 33.018.6704.10.4025.5
OAK 450-H2O2147.2 ± 6.926.571.33.90.5024.3
OAK 450-KOH141.0 ± 13.225.477.72.80.6019
OAK 65076.6 ± 0.713.876.51.40.8021.3
OAK 650-H2SO4106.8 ± 18.719.262.92.40.6034.2
OAK 650-H3PO4126.3 ± 63.922.750.53.60.5045.4
OAK 650-H2O271.2 ± 7.412.863.72.30.5033.5
OAK 650-KOH132.3 ± 11.223.868.31.90.7029.2
a Oxygen content determined as difference between % C, H, N and S from 100 (dry basis); b NH4+eq refers to the maximum equivalent char NH4+ sorption capacity since 1 cmolc kg−1 = 0.18 mg NH4+ kg−1 char; char. CEC expressed as average of duplicates ± standard deviation.

Share and Cite

MDPI and ACS Style

Takaya, C.A.; Parmar, K.R.; Fletcher, L.A.; Ross, A.B. Biomass-Derived Carbonaceous Adsorbents for Trapping Ammonia. Agriculture 2019, 9, 16. https://0-doi-org.brum.beds.ac.uk/10.3390/agriculture9010016

AMA Style

Takaya CA, Parmar KR, Fletcher LA, Ross AB. Biomass-Derived Carbonaceous Adsorbents for Trapping Ammonia. Agriculture. 2019; 9(1):16. https://0-doi-org.brum.beds.ac.uk/10.3390/agriculture9010016

Chicago/Turabian Style

Takaya, Chibi A., Kiran R. Parmar, Louise A. Fletcher, and Andrew B. Ross. 2019. "Biomass-Derived Carbonaceous Adsorbents for Trapping Ammonia" Agriculture 9, no. 1: 16. https://0-doi-org.brum.beds.ac.uk/10.3390/agriculture9010016

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