Sie sind auf Seite 1von 8

Microporous and Mesoporous Materials 179 (2013) 250257

Contents lists available at SciVerse ScienceDirect

Microporous and Mesoporous Materials


journal homepage: www.elsevier.com/locate/micromeso

Nitrate sorption and desorption in biochars from fast pyrolysis


Rajesh Chintala a,, Javier Mollinedo a, Thomas E. Schumacher a, Sharon K. Papiernik b, Douglas D. Malo a,
David E. Clay a, Sandeep Kumar a, Dylan W. Gulbrandson a
a
Department of Plant Science, South Dakota State University, Brookings, SD, USA
b
USDA-ARS, Brookings, SD, USA

a r t i c l e i n f o a b s t r a c t

Article history: Increasing the nitrate (NO 3 ) sorption capacity of Midwestern US soils has the potential to reduce nitrate
Received 13 February 2013 leaching to ground water and reduce the extent of the hypoxia zone in the Gulf of Mexico. The objective
Received in revised form 1 April 2013 of this study was to determine the sorption and desorption capacity of non-activated and chemically acti-
Accepted 27 May 2013
vated biochars from microwave pyrolysis using selected biomass feedstocks of corn stover (Zea mays L.),
Available online 14 June 2013
Ponderosa pine wood chips (Pinus ponderosa Lawson and C. Lawson), and switchgrass (Panicum virgatum
L.). Surface characteristics such as surface area and net surface charge have shown signicant effects on
Keywords:
nitrate sorption and desorption in biochars. Freundlich isotherms performed well to t the nitrate sorp-
Nitrate retention
Nitrate release
tion data (R2 > 0.95) of biochars when compared to Langmuir isotherms. Nitrate sorption and desorption
Biochars was signicantly inuenced by solution pH and presence of highly negative charged potential ions such
Surface charge as phosphate (PO3 2
4 ) and sulfate (SO4 ) in aqueous solution. Chemical activation with concentrated HCl
Point of zero net charge had signicant effect on surface characteristics of biochars and enhanced the nitrate sorption capacity.
The rst order model t the nitrate desorption kinetics of biochars with a high coefcient of determina-
tion (R2 > 0.95) and low standard error (SE).
2013 Elsevier Inc. All rights reserved.

1. Introduction (NO 3 ) in ground water. But this process requires reduced pH and
produces ammonia as by-product [16]. Biological denitrication
In Midwestern US agriculture, ammonium (NH 4 ) based fertiliz- is slow and inefcient to remove nitrate from highly concentrated
ers are routinely applied to meet the nitrogen (N) requirement of systems, and also requires additional organic substrates as electron
agricultural crops. Nitrogen fertilizers generally undergo chemical donors [17]. Reverse osmosis and ion exchange resins are not eco-
transformations such as ammonication, nitrication, de-nitrica- nomical for large scale nitrate removal. Ion exchange resins are not
tion, nitrogen xation, and immobilization in soil [1,2]. Nitrica- selective and also retain other anions (such as SO2 4 and HCO 3)
tion is one of the quickest reaction pathways. It results in the which cause the disequlibrium in the ionic composition of soil
formation of nitrate ion, which is plant available but also highly and water systems and also release chloride (Cl) ions during the
mobile due to its weak afnity to form surface complexes and exchange process [18]. Among these nitrate removal technologies,
therefore prone to leaching [35] or is subject to denitrication. adsorption has been found less expensive and more effective in re-
Increasing nitrate concentrations in surface and groundwater re- moval of nitrate from water systems. Adsorption technologies have
sources across the world have been attributed to agricultural and evolved as highly efcient remediation tools with lesser energy in-
industrial inputs [6,7]. Several nitrate removal technologies have put to produce sorbent materials; adsorption is easy to apply at
been developed such as anion exchange [8], biological denitrica- large scales [15].
tion [9], chemical denitrication [10,11], catalytic denitrication Control of nitrate mobility at its source would eliminate the
[12], reverse osmosis [13], and electrodialyis [14]. These available need for remediation. Various sorbents have been identied which
physico-chemical and biological technologies may be expensive, have potential to reduce of nitrate ion mobility in soil and water
generate additional by-products and toxic wastes, and have limited ecosystems [19,20]. The success of adsorption techniques depends
applicability in large scale remediation scenarios [15]. For instance, on the selection of appropriate sorbents, their removal efcacy,
chemical denitrication (reduction) using reductants such as zero- and economic feasibility. Sorbents (either naturally available or
valent iron (ZVI) have been extensively used to reduce nitrate synthetically produced) to remove nitrate from water and waste
water systems include functionalized mesoporous siliceous and
non-siliceous materials [2123], metal organic frame works
Corresponding author. Address: SNP 247, Box 2140C, South Dakota State
[24], mesoporous carbon materials [25], red mud [26], agricul-
University, Brookings, SD 57006, USA. Tel.: +1 6056884547.
E-mail address: rajesh.chintala@sdstate.edu (R. Chintala).
tural waste [27,28], polypropylene-g-N,N-dimethylaminoethyl

1387-1811/$ - see front matter 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.micromeso.2013.05.023
R. Chintala et al. / Microporous and Mesoporous Materials 179 (2013) 250257 251

methacrylate [29], poly (dimethylaminoehtyl methacrylate/2-hy- time of 18 min. These biochars were activated by heating with con-
droxy ethyl methacrylate) [30], poly (dimethyl diallyl ammonium centrated hydrochloric acid (HCl). Air dried biochar (25 g) was
chloride)/polyacrylamide [31], sepiolite [19], bentonite [32,33], placed in a conical ask covered with a watchglass and treated
slag [19], activated carbon [19,34], chitin and chitosan derivatives with 250 ml of concentrated HCl. The biochar mixed with concen-
[35,36], and MII-Al-Cl layered double hydroxides [37,38]. In recent trated HCl was heated on hot plate at 200 C (placed in laboratory
years, the depletion of petroleum oil reserves and growing demand hood) for 24 h. After heating, the biochars were washed with
for energy across the world has initiated the development of sus- deionized water four times and then washed with 1% sodium
tainable alternative energy technologies based on renewable bio- bicarbonate (NaHCO3) solution (1:500 solid-solution ratio) four
mass. Biochar is the by-product of biomass conversion using times to remove acid residues. Later the biochar materials were
pyrolysis to produce bio-oil and syngas. The scientic community washed with deionized water several times until the pH of the
has found opportunities to develop these biochar materials as cli- supernatant was 7.0. The washed biochars were oven dried at
mate change mitigation tools due to their ability to store carbon 105 5 C for 8 h and considered as acid activated biochars [49].
in soil for long periods. Biochar based materials have the potential These non-activated and activated biochar materials were homog-
to sorb anionic nutrients such as nitrogen (N) and phosphorus (P) enized and ground to pass through a 0.2 mm sieve and used for
from aqueous solutions [39]. Several studies have described that characterization and batch studies.
biochars can also supply nutrients to crop plants and reduce the
leaching of nutrients in soil [40,41]. Unlike other existing nitrate 2.2. Characterization of non-activated and activated biochar materials
removal sorbents, biochar can be produced by a simple, cost effec-
tive process and its application to agricultural lands may improve Specic surface area of biochars, were determined using the io-
soil quality and health. dine absorption method [50]. The biochar acidity, pH, and EC (elec-
The sorption capacity of highly carbon rich biochars have been trical conductivity) were measured at 1:1 water to solid ratio after
found to be excellent due to their unique surface characteristics shaking for 30 min in deionized water [51]. The cation exchange
such as high surface area, pore volume, and surface functional capacity (CEC) of biochar samples were measured using 1 M
groups [42]. The surface area and porosity of biochars can be en- ammonium acetate (CH3COONH4) at pH 7.0 [51]. The quantitative
hanced by either physical or chemical activation methods [43]. analysis of volatile organic compounds (VOCs) in biochar samples
The activation of biochars from agricultural products has been in- were determined [52]. Total carbon and total nitrogen were ana-
creased in recent years due to its use in environmental pollution lyzed by Elementar Vario MAX CNS analyzer [53]. Biochar samples
control [44,45]. Among the activation methods, chemical activa- were wet digested using concentrated nitric acid (HNO3) in a pres-
tion is more economical with higher yield at lower temperature surized (200 psi) microwave oven (MARS 5, CEM Corp., USA)
and less production of burn-off char [46]. Slowing the rate at which [54,55] equipped with teon closed vessels. The wet digested sam-
of nitrate is leached from surface to subsurface soils could improve ples were diluted with distilled water to measure nutrient compo-
water quality and reduce greenhouse gas generation. The sorption sition (total P, Ca, Mg, Na, and K) using Inductively Coupled Plasma
capacity of soils is critical along with the water ow to determine Atomic Emission Spectroscopy (ICP-AES) (Varian, Australia) [56].
the leaching rate of nitrate in soil. Cation transport is slowed by All these determinations were conducted in four separate biochar
sorption on soils with negative charge, while anion transport is slo- samples from different pyrolysis runs.
wed by sorption on positively charged soil particles [47]. Most
Midwestern United States soils have relatively low anion sorption 2.3. Infrared spectroscopic analysis of non-activated and activated
capacity. For example, the anion exchange capacity of Kranzburg biochars
silty clay loam soil (ne-silty, mixed, superactive, frigid, Aeric Cal-
ciaquoll) was ranged from 0.25 to 0.27 cmolc kg1. This relatively Infrared spectrums were recorded for biochar samples using
minor anion sorption capacity must be compared with the soil Fourier Transform Infrared Spectroscopy (Thermo Scientic NICO-
cation exchange capacities that ranged from 27.9 to 33.127 LET 6700 FT-IR spectrometer). Samples were prepared and pressed
cmolc kg1 [47]. into pellets using 10 wt% potassium bromide (KBr). The spectrum
Batch experiments are often used to quantify the ionic retention had wave number (cm1) on x-axis and transmittance (%) on y-axis
of biochar and other sorbents [48]. The determination of sorption with resolution using 40 scans. The peaks of the spectrum were
coefcients (Kd) and sorption maxima (qmax) of these sorbents help studied to understand the surface functionality of non-activated
in developing them as remediation tools to control contaminant and activated biochar samples.
transport in soil and water eco-systems. In this context, it is impor-
tant to investigate the interaction of biochar-based sorbents from 2.4. Surface charge characteristics of non-activated and activated
agricultural products with highly mobile anionic nutrient such as biochars
nitrate is essential. Therefore this study was designed with the
following objectives: (1) determine the surface characteristics of The point of zero net charge (PZNC) was measured by determin-
non-activated and activated biochars produced from microwave ing the cation exchange capacity (CEC) and anion exchange capac-
pyrolysis using corn stover (Zea mays L.), Ponderosa pine (Pinus ity (AEC) simultaneously at eight pH values (pH 210) [57]. The
ponderosa Lawson and C. Lawson) wood residue, and switchgrass PZNC was the pH at which CEC and AEC were equal on sorbent sur-
(Panicum virgatum L.); and (2) determine the nitrate sorption and- face. Biochar samples were initially saturated with 0.01 M KCl and
desorption potential of non-activated and activated biochars. the pH was adjusted with 1 M HCl and 1 M KOH. For the pH adjust-
ment, biochar samples (10 g) were weighed in triplicate. These bio-
char samples were suspended in 100 ml of 0.01 M KCl to prepare
2. Materials and methods subsequent pH levels. These biochar suspensions were adjusted
to different pH by adding predetermined amounts of 1 M HCl
2.1. Production of non-activated and activated biochar materials and 1 M KOH. These predetermined quantities were calculated
by using regression equations tted to the potentiometric titration
Using corn stover, Ponderosa pine wood chips and switchgrass curves of these biochar materials. The biochar samples were sha-
as feedstocks, three biochars were formed as co-products of bio- ken for 12 h and the pH was measured every day until the pH val-
oil production using microwave pyrolysis at 650 C and residence ues were constant. Then suspensions were ltered and washed
252 R. Chintala et al. / Microporous and Mesoporous Materials 179 (2013) 250257

with ethanol and dried to obtain biochar samples of different pH 2.7. Desorption of nitrate from non-activated and activated biochars
levels to use for determining PZNC. Later, 0.5 M NaNO3 was used
as an extractant to exchange K+ and Cl from surface of samples Desorption of nitrate was determined in biochar samples (at the
to determine CEC and AEC. Potassium (K) was analyzed using end of the sorption experiment using an initial nitrate concentra-
Flame Atomic Absorption Spectroscopy (GBC Avanta Corp., USA) tion of 1.29 mmol L1. After equilibration and sorbed nitrate con-
and the Cl ion was determined using Flow Injection Analysis centrations were determined, samples were extracted with 50 mL
(Quick Chem FIA+, 8000 series, Latchat Instruments, USA). The of deionized water (a 1:50 solid to solution ratio) at two pH levels
CEC (K+) and the AEC (Cl) were calculated using Eqs. (1) and (2). 4.0 and 9.0 to study the pH effect on desorption. The tubes with
1
sample and extractant (deionized water) were then shaken for
CEC cmolc kg 0:1C 2 V 2  C 1 V 1 =39W 1 24 h on a reciprocating shaker. Samples from four replications
were then centrifuged, supernatant solutions were ltered using
1
AEC cmolc kg 0:1C 2 V 2  C 1 V 1 =35:5W 2 ashless lter paper (with retention: >2.5 lm) and analyzed for des-
+  orbed nitrate concentration as described above. The desorbed ni-
where C1 = K or Cl concentration in supernatant solution of
trate was calculated as percentage using equation (Eq. (5)).
0.01 M KCl (mg L1), V1 = volume of entrained solution in the bio-
char sediment (mL), C2 = K+ or Cl concentration in the displacing Desorption % desorbed nitrate adsorbed nitrate  100 5
solution of 0.5 M NaNO3 (mg L1), V2 = total volume of displacing
solution of 0.5 M NaNO3 (mmol L1), 39 = atomic weight of K+, Desorption kinetics were studied by successive measurement of
35.5 = atomic weight of Cl, and W = oven dried sample weight (g). nitrate concentration at different agitation times. Following the
sorption step, deionized water (20 ml) with pH levels of 4.0 and
9.0 was added to 0.5 g of nitrate treated biochar samples in centri-
2.5. Nitrate sorption isotherms for non-activated and activated
fuge tube. Samples were equilibrated on reciprocating shaker at
biochars
27 1 C for periods of 15,30,45,60, 75, and 90 min. At the end of
each agitation period the supernatant solutions were collected to
All biochar materials were adjusted to pH 4.0 0.1 after saturat-
measure nitrate concentrations. Then fresh deionized water
ing with K+ or Cl using KCl solution [57]. This process was done to
(20 ml) was added to continue to measure nitrate desorption at
maximize the nitrate sorption and remove other competitive an-
longer agitation times. The rst order kinetic model was tted to
ions from biochar surface. Nitrate sorption isotherms were deter-
the data using equation (Eq. (6)).
mined for biochars produced from the previous process. Biochar
samples (0.5 g) were placed in a 50-mL centrifuge tube. A 20-mL lnP0  Pt a  K d t 6
solution of 0.01 M calcium chloride (CaCl2) containing 0, 0.16,
1
0.32, 0.64, 1.29, and 1.61 mmol L1 of nitrate (NO 3 ) were added
where Pt = desorbed nitrate at time t (mmol kg ), P0 = desorbed ni-
along with two drops of chloroform to inhibit microbial growth. trate at equilibrium (mmol kg1), Kd = kinetic constant (h1),
The samples were equilibrated in an end-over-end shaker at room t = time (min), and a = constant.
temperature for 24 h and then centrifuged (with xed-angle rotor) The goodness of t for this kinetic model was determined by
for 5 min, producing a clear supernatant. The supernatant was l- evaluating its coefcient of determination (R2) and standard error
tered using ashless lter paper with retention: >2.5 lm (Whatman of estimate (SE). A one-way analysis of variance (ANOVA) was per-
42, Whatman Corp., Kent, UK) and then analyzed for nitrate using formed using the SAS Statistical Package, version 9.2 to determine
Flow Injection Analysis (Quick Chem FIA+, 8000 series, Latchat the effect of treatments on characterization, nitrate sorption, and
Instruments, USA). Sorbed nitrate was calculated as the difference desorption. TukeyKramer multiple comparisons analysis was
between equilibrium nitrate concentration (after shaking) and ini- done to determine the signicant differences (p 6 0.05 level) be-
tial nitrate concentration of solution. Adjustable parameters for tween treatments [59].
Freundlich (Eq. (3)) and Langmuir (Eq. (4)) were determined by
non-linear regression using Proc NLIN [58]: 3. Results and discussion
Freundlich equation q K f c1=n 3
3.1. Characterization of non-activated and activated biochars
Langmuir equation q K L bc=1 K L c 4
The physical and chemical properties of non-activated biochars
where q = sorbed nitrate (mmol kg1), c = nitrate concentration in varied signicantly with biomass feedstocks (p < 0.05) (Table 1).
the equilibrium solution (mmol L1), Kf = Freundlich partitioning Corn stover (CSB) and switchgrass biochars (SGB) exhibited signif-
coefcient, 1/n = sorption intensity, b = adsorption maxima icantly higher pH, EC, CEC, PZNC, VOCs, total nitrogen, and base
(mmol kg1 of nitrate), and KL = parameter related to binding en- cation concentration (Ca, Mg, K, and Na) than Ponderosa pine wood
ergy (L kg1). residue biochar (PWRB), whereas PWRB had higher surface area
and total carbon content. The alkaline nature of CSB and SGB
2.6. Effect of solution pH and competitive anions may be due to high base cation concentration which may partici-
pate in proton consuming reactions [60]. The presence of soluble
The pH of the initial nitrate solution of 1.29 m mmol L1 was salts caused higher EC values for these plant based biochars. CEC
adjusted to obtain a range in solution pH from 4.0 to 9.0 [57]. This of CSB and SGB was signicantly higher than PWRB due to their
pH adjusted nitrate solution was added to biochar (adjusted to pH high pH and negative surface charge potential. The source of this
4.0 0.1) to study the effect of solution pH on nitrate sorption. The surface charge potential may be due to the amphoteric nature of
inuence of competitive anions such as phosphate and sulfate on oxygenated functional groups present in VOCs [61]. CSB and SGB
nitrate sorption was also investigated by adding 0.01 M KH2PO4 had signicantly higher VOC concentrations than PWRB.
and K2SO4 solutions (1.29 mmol L1 of phosphate and sulfate) in The chemical activation (with conc. HCl) of these three biochars
separate centrifuge tubes. Biochar (0.5 g) sample (with pH 4.0) signicantly increased their surface area and CEC, but decreased
was added to 20 mL of pH adjusted nitrate solutions with other chemical properties such as pH, EC, PZNC, VOCs, total carbon,
competitive ions (with pH 4.0) into centrifuge tubes. Sorption of and base cation concentration. (Table 1). Surface area of SGB and
nitrate was determined using the procedure described above. CSB was increased by 12 times whereas PWRB increased by 9 times
R. Chintala et al. / Microporous and Mesoporous Materials 179 (2013) 250257 253

Table 1
Physical and chemical properties of biochars produced from fast pyrolysis at 650 C of corn stoverb (CSB), switchgrass (SGB), and Ponderosa pine wood residue (PWRB). Values are
means of four samples with standard errors. Different letters within the row represent signicant differences by Tukeys test at a = 0.05.

Properties CSB PWRB SGB


Non-activated Activated Non-activated Activated Non-activated Activated
Specic surface area (m2 g1) 43.4 2.5f 513 21b 52.1 4.1d 456 15c 48.0 2.6e 582 23a
pH 10.5 0.04a 8.16 0.10c 7.85 0.02f 7.05 0.11e 9.73 0.24b 7.75 0.10d
EC (lS cm1) 3000 61a 140 21e 200 22c 110 15f 890 21b 152 16d
CEC (cmolc kg1) 41.6 5e 134 11b 32.5 3f 115 7c 43.1 2d 149 8a
PZNC 6.67 0.15b 4.63 0.13d 5.75 0.03c 3.04 0.07e 7.08 0.44a 4.92 0.14d
VOC (g kg1) 206 21a 58.5 8e 125 11c 45.1 6f 184 15b 59.1 7d
Total N (g kg1) 13.5 0.3b 12.8 0.1b 4.82 0.1c 4.94 0.03c 15.8 0.7a 15.2 0.5a
Total C (g kg1) 742 10d 726 13e 845 30a 801 14b 763 22c 720 17c
Ca (g kg1) 7.82 0.25a 6.12 0.55c 2.62 0.16d 1.25 0.09e 7.04 0.22b 5.82 0.61c
Mg (g kg1) 5.61 0.26a 4.33 0.15c 0.86 0.31e 0.26 0.03f 5.18 0.06b 4.15 0.05d
K (g kg1) 23.5 2.6a 15.6 1.4b 1.55 0.03e 0.92 0.07f 2.92 0.05c 1.72 0.11d
Na (g kg1) 0.86 0.06a 0.64 0.01c 0.62 0.03c 0.45 0.05d 0.75 0.05b 0.60 0.08c

CSB, Corn stover biochar; PWRB, Ponderosa pine wood residue biochar; SGB, Switchgrass biochar.

Fig. 1. ac FTIR sepctra of non-activated and activated biochars. KBr pellet 10 wt%.

after chemical activation. Biochars may undergo decomposition VOCs, and soluble salts and decreasing pH, EC, PZNC, and total
during the acid activation process and subsequent washing with carbon. For example, the activation may have had a catalytic effect
NaHCO3 solution, resulting in loss of base cation concentration, on the cleavage of bond linkages in hemicellulose and cellulose,
254 R. Chintala et al. / Microporous and Mesoporous Materials 179 (2013) 250257

Fig. 2. ac pH-dependent CEC and AEC of biochars within the pH investigated (2.0 6 pH 6 9.0). Symbols represent data SE. Each data point is average of four replications.

Table 2
Freundlich and Langmuir nitrate sorption parameters with standard errors for biochars. Different letters in the same column represent signicant differences by Tukeys test at
a = 0.05.
Biochar Freundlich Langmuir
Kf 1/n R2 KL b R2
CSB Non-activated 384 21e 0.77 0.07b 0.99 0.95 0.06d 620 26e 0.73
Activated 456 28c 0.60 0.03e 0.99 1.59 0.05b 702 35b 0.90
PWRB Non-activated 158 17f 0.65 0.02c 0.99 0.41 0.04f 184 13f 0.95
Activated 437 26d 0.65 0.08d 0.99 1.09 0.04c 696 24c 0.84
SGB Non-activated 447 44b 0.92 0.12a 0.99 0.92 0.05e 625 35d 0.80
Activated 576 31a 0.57 0.08f 0.99 1.76 0.06a 800 35a 0.92

CSB, corn stover biochar; PWRB, Ponderosa pine wood residue biochar; SGB, switchgrass biochar. Kf and KL are expressed in L kg1 and b is in mmol L1.

increasing the internal pore volume. The acid treatment may have 3.2. Infrared spectroscopic analysis of non-activated and activated
oxidized the carbon and made walls of pores thinner, increasing biochars
both the micro-porosity and surface area of biochars [62]. The in-
creased surface area would also inuence the sorption of ions The baseline corrected FT-IR spectra of non-activated and acti-
which may also result high CEC. Similar changes in pH, EC, and vated biochars were shown the vibrational mode of surface func-
ash fraction were observed for pecan shell-based granular carbon tional groups (Fig. 1ac). The composition of functional groups in
following activation with phosphoric acid (H3PO4) [63]. Changes biochars produced from herbaceous feedstocks (CSB and SGB)
in surface properties due to the activation process are expected was similar. Non-activated and activated biochars of CSB and
to inuence the adsorptive efciency of biochars and their ionic SGB had shown CN stretch (23002360 cm1), C@O stretch
retention mechanism in soil. (17251735 cm1), CAO stretch (11001124 cm1), and aromatic
R. Chintala et al. / Microporous and Mesoporous Materials 179 (2013) 250257 255

CAH vibrational band (700713 cm1). PWRB was recorded


with bands of CN stretch (23502352 cm1), CAO stretch
(10911100 cm1), and aromatic CAH (740743 cm1) in both
non-activated and activated biochars. Chemical activation signi-
cantly decreased the absorbence of FT-IR bands of all functional
groups. The acid-base properties of these functional groups on bio-
char develop surface charge and result in electrostatic attraction
for opposite charged ions in soil solution.

3.3. pH-dependent surface charge characteristics of biochars

Surface charge of three biochars was found to be variable with


pH (Fig. 2ac). All three biochars exhibited a wide range of CEC and
AEC values as pH increased from 2 to 10. The CEC increased signif-
icantly (p < 0.05) with increasing pH in all biochars, but AEC de-
creased with increasing pH. The CEC and AEC of CSB and SGB
were relatively higher than PWRB at all pH. Each unit increase in
pH caused a higher increase of CEC for SGB (3.88 cmol kg1) than
for CSB (3.41 cmol kg1) and PWRB (2.91 cmol kg1). AEC was de-
creased by an average of 1.74, 2.33, and 2.84 cmol kg1 in SGB,
CSB, and PWRB respectively, per unit increase in pH. The PZNC Fig. 3. Quantity of nitrate sorbed per kilogram of biochar (q) across a range of
equilibrium nitrate solution concentrations at pH = 4.0 0.1 for non-activated and
was also signicantly higher for SGB (7.08) than CSB (6.51) and
activated biochars produced from corstover (CSB), Ponderosa pine wood residue
PWRB (5.75). The pH of all the biochars was higher than their PZNC (PWRB), and switchgrass (SGB). Lines represent Freundlich isotherms. Each data
values, implying an overall negative surface charge potential. The point was mean of four replications.
variable surface charge of biochars and their anion exchange
capacity may be caused by the protonation of functional groups
which can be inuenced by the pH and ionic strength of aqueous
solution. The sorption studies were conducted using biochars ad-
justed to pH 4.0 where the net surface charge of biochars was po-
sitive according to their PZNC. The information related to surface
charge and composition would be extremely critical to determine
the utility of these biochars as sorbents to remove anions such as
nitrate from solutions.

3.4. Nitrate sorption in non-activated and activated biochars

Freundlich and Langmuir isotherm models were used to t the


equilibrium and sorbed concentration of nitrate in biochars and
activated biochars for both non-activated and activated biochars
(Table 2). Coefcients of determination (R2 value) were higher for
Freundlich isotherms for both non-activated and activated bioch-
ars when compared to the Langmuir isotherm. The better t of
the Freundlich isotherm model suggests heterogeneity of sorption
sites on the biochar surface which is not restricted to monolayer
formation (Fig. 3). Nitrate sorption at the biochar surface may be
mainly due to electrostatic interaction (outer-sphere complexation Fig. 4. Effect of solution pH on nitrate sorption (q sorbed nitrate concentration) in
mechanism) and to a lesser extent ionic exchange mechanism. non-activated and activated biochars produced from corstover (CSB), Ponderosa
There was a signicant difference in sorption potential of biochars pine wood residue (PWRB), and switchgrass (SGB). The initial nitrate concentration
of solution was 1.29 mmol L1 at pH ranged from 4.0 0.1 to 9.0 .1. Values are the
due to biomass feedstock type and activation. Among non-
mean of four samples. Error bars indicate standard error of the mean.
activated biochars, SGB was found to have the highest sorption
capacity (highest Kf) followed by CSB and PWRB. The higher nitrate
sorption observed for SGB and CSB possibly due to their higher
PZNC, VOCs, and base cation concentration providing a greater po- sorption in activated biochars may be mainly due to increased sur-
sitive charge on surface functional groups. The higher base cation face area.
concentration also may facilitate divalent cation bridging with
negatively charged nitrate ions. 3.5. Effect of solution pH and competitive anions on nitrate sorption of
The acid activation (with conc. HCl) signicantly increased the non-activated and activated biochars
sorption capacity of all biochars. A greater increase in nitrate sorp-
tion was observed in activated PWRB (440%) followed by activated Solution pH had a signicant effect on nitrate sorption of non-
biochars of CSB (240%) and SGB (200%). The acid activation greatly activated and activated biochars (Fig. 4). As the solution pH in-
increased the surface area of all biochars, providing more space for creased from 4.0 to 9.0, the nitrate sorption signicantly decreased
sorption of ions. CEC was also increased in biochars due to in- (p < 0.05). The decrease in nitrate sorption per unit increase in
creased surface area during acid activation. Despite may chemical solution pH was higher in activated biochars than non-activated
changes induced by acid activation (decrease in pH, EC, PZNC, biochars. The oxygenated functional groups of biochar undergo
VOCs, and base cation concentration, Table 1), the increased nitrate protonation and de-protonation reactions as aqueous solution pH
256 R. Chintala et al. / Microporous and Mesoporous Materials 179 (2013) 250257

Table 3
First order kinetic equations for nitrate sorption on biochars at pH 4.0 and initial
nitrate solution concentration of 1.29 mmol L1. Different letters in the same column
represent signicant differences by Tukeys test at a = 0.05.

Biochar type a Kd  102 SE R2


CSB non-activated 4.35c 2.78b 0.05 0.97
CSB activated 5.05b 2.81b 0.08 0.99
PWRB non-activated 3.61d 3.48a 0.07 0.95
PWRB activated 5.09b 3.52a 0.05 0.97
SGB non-activated 4.38c 2.79b 0.03 0.96
SGB activated 5.33a 2.83b 0.09 0.98

CSB, corn stover biochar; PWRB, Ponderosa pine wood residue biochar; SGB,
switchgrass biochar.

sorption sites towards more highly negatively charge potential an-


ions such as phosphate and sulfate. This phenomenon was ob-
served to be more predominant with activated biochars which
may be due to their higher surface area and greater sorption capac-
ity. The decrease in nitrate sorption was higher in CSB (71.3%) and
SGB (74.2%) than PWRB (57.6%) among non-activated biochars in
Fig. 5. Effect of competitive ion concentration in solution on nitrate sorption by
the presence of phosphate ion. When phosphate was present, the
non- activated and activated biochars from corstover (CSB), Ponderosa pine wood % decrease in nitrate sorption (75%) was the same for all activated
residue (PWRB), and switchgrass (SGB). The initial nitrate concentration of solution biochars. Sulfate ion had a lesser effect on nitrate sorption, about
was 1.29 mmol L1 at pH 4.0 0.1. Values are the mean of four samples. Error bars 3854%. In the presence of sulfate ion, the decrease in nitrate sorp-
indicate standard error of mean.
tion was greater for CSB and SGB than for PWRB in activated and
non-activated biochars.

changes. As the solution pH decreases, the positive charge on the


biochar surface increases due to protonation reactions which in- 3.6. Desorption of nitrate from non-activated and activated biochars
creased the electrostatic attraction between the biochar surface
and negatively charged nitrate ions. Anion exchange also took The desorption (%) of sorbed nitrate (at initial nitrate concentra-
place between the anions on the biochar surface and nitrate ions tions of 0.16, 0.64, and 1.29 mmol L1) was signicantly higher in
by breaking up of electrostatic interactions. The functional groups activated biochars than non-activated biochars at both pH 4.0 and
on the biochar surface undergo dissociation (deprotonation) and 9.0 (Fig. 6a and b). Nitrate desorption was signicantly higher at
develop negative charge as the solution pH increases above the pH 9.0 than 4.0 and showed hysteresis. As the initial nitrate concen-
PZNC of biochars. The competition for sorption sites from OH ions tration increased, nitrate desorption also increased. Higher nitrate
then increases which may hinder nitrate sorption. concentration may translate to lower bond energy and more
The presence of competitive anions such as phosphate and sul- exchangeable ions (nitrate more easily desorbed). Nitrate desorp-
fate in solution signicantly decreased nitrate sorption on non- tion was signicantly higher at pH 9.0 than 4.0. The surface of bio-
activated and activated biochars (Fig. 5). The decrease in nitrate char may develop negative charge as the pH increased over PZNC
sorption was signicantly higher in activated biochars than non- which may repel the negatively charged nitrate ions. Moreover,
activated biochars in presence of both phosphate and sulfate ions. the OH concentration in solution may also increase at high pH
The decrease in nitrate sorption may be due to the preference of and displace nitrate ions from sorption sites. Both non-activated

Fig. 6. a and b Desorption of nitrate on biochars and activated biochars from corstover (CSB), Ponderosa pine wood residue (PWRB), and switchgrass (SGB). The initial nitrate
concentration of solution was 1.29 mmol L1 at two pH levels (4.0 0.1 and 9.0 0.1). Values are mean of four samples. Error bars indicate standard error of mean.
R. Chintala et al. / Microporous and Mesoporous Materials 179 (2013) 250257 257

and activated CSB and SGB showed signicantly higher desorption [15] A. Bhatnagar, M. Sillanpaa, Chem. Eng. J. 168 (2011) 493504.
[16] L.F. Cheng, R. Muftikian, Q. Fernando, N. Korte, Chemosphere 35 (1997) 2689
than PWRB at both pH 4.0 and 9.0. Nitrate desorption of non-
2695.
activated and activated biochars followed rst order kinetics with [17] F. Thalasso, A. Vallecillo, P. Garcia-Encina, F. Fdz-Polanco, Water Res. 31 (1997)
higher coefcients of determination (R2) and low standard error 5566.
(SE) (Table 3). Acid activation did not affect rst order rate con- [18] C.T. Matos, A.M. Sequeira, S. Velizarov, J.G. Crespo, M.A.M. Reis, J. Hazard.
Mater. 166 (2009) 428434.
stants (Kd). But nitrate desorption was complete (equal to nitrate [19] N. Ozturk, T.E. Bektas, J. Hazard. Mater. 112 (2004) 155162.
released at equilibrium) in approximately 58 min for non-activated [20] Y. Wang, B.Y. Gao, W.W. Yue, Q.Y. Yue, J. Environ. Sci. 19 (2007) 13051310.
and activated biochars of CSB and SGB and 54 min for PWRB which [21] A. Vinu, K.Z. Hossain, K. Ariga, J. Nanosci. Nanotech. 5 (2005) 347375.
[22] S. Hamoudi, R. Saad, K. Belkacemi, Ind. Eng. Chem. Res. 46 (2007) 88068812.
were not signicantly different. [23] K. Ariga, A. Vinu, Y. Yamauchi, Q. Ji, J.P. Hill, Bull. Chem. Soc. Jpn. 85 (2012) 1
32.
[24] N. Stock, S. Biswas, Chem. Rev. 112 (2012) 933969.
4. Conclusions
[25] A. Vinu, M. Miyahara, V. Sivamurugan, T. Mori, K. Ariga, J. Mater. Chem. 15
(2005) 51225127.
Our results demonstrate strong sorption potential of selected [26] Y. Cengeloglu, A. Tor, M. Ersoz, G. Arslan, Sep. Purif. Technol. 51 (2006) 374
378.
biochars to remove anionic nutrient such as nitrate from aqueous
[27] U.S. Orlando, A.U. Baes, W. Nishijima, M. Okada, Bioresour. Technol. 83 (2002)
solution. The nitrate sorption capacity of biochars was observed 195198.
to depend on surface properties of biochar (surface area and sur- [28] P.C. Mishra, R.K. Patel, J. Environ. Manage. 90 (2009) 519522.
face charge), solution pH, and presence of competitive ions with [29] M.F.A. Taleb, G.A. Mahmoud, S.M. Elsigeny, E.S.A. Hegazy, J. Hazard. Mater. 159
(2008) 372379.
high negative potential. Surface charge of biochars was found to [30] Y. Tian, Chin. Chem. Lett. 19 (2008) 11111114.
be pH dependent due to the amphoteric nature of surface func- [31] Y. Zheng, A. Wang, J. Chem. Eng. Data 55 (2010) 34943500.
tional groups. Study of surface characteristics is essential to under- [32] C.J. Mena-Duran, M.R.S. Kou, T. Lopez, J.A. Azamar-Barrios, D.H. Aguilar, M.I.
Dominguez, J.A. Odriozola, P. Quintana, Appl. Surf. Sci. 253 (2007) 57625766.
stand the potential of biochars as soil amendments at the eld [33] K.G. Bhattacharyya, S.S. Gupta, Adv. Colloid Interface Sci. 140 (2008) 114131.
scale or use biochars as remediation tool. Acid activation of bioch- [34] A. Khani, M. Mirzaei, Comparative study of nitrate removal from aqueous
ars tremendously increased the biochar surface area and the po- solution using powder activated carbon and carbon nanotubes, in: 2nd
International IUPAC Conference on Green Chemistry, Russia, 2008 pp. 1419.
tential to sorb nitrate ions. Acid activation seems to be efcient [35] A. Bhatnagar, M. Sillanpaa, Adv. Colloid Interface Sci. 152 (2009) 2638.
in increasing the sorbing potential of biochars that can immobilize [36] S. Chatterjee, D.S. Lee, M.W. Lee, S.H. Woo, J. Hazard. Mater. 166 (2009) 508
nitrate ions. In higher nitrate concentration systems, the activated 513.
[37] K. Hosni, E. Srasra, Inorg. Mater. 44 (2008) 742749.
biochars have greater utility in reducing the availability of nitrate
[38] M. Islam, R. Patel, Desalination 256 (2010) 120128.
ions for biological de-nitrication and also mitigating the eutrophi- [39] M. Zhang, B. Gao, Y. Yao, Y. Xue, M. Inyang, Chem. Eng. J. 210 (2012) 2632.
cation. Biochars may also increase the residence time of highly mo- [40] J. Lehmann, J. Pereira da Silva, C. Steiner, T. Nehls, W. Zech, B. Glaser, Plant Soil
249 (2003) 343357.
bile nitrate ions and make them more available for plant and
[41] J. Major, M. Rondon, D. Molina, S.J. Riha, J. Lehmann, Plant Soil 333 (2010) 117
nutrient utilization under conditions of limited N availability. Ni- 128.
trate desorption was higher for chemically-activated biochars [42] G.N. Kasozi, A.R. Zimmerman, P. Nkedi-Kizza, B. Gao, Environ. Sci. Technol. 44
and increased solution pH. Feedstock inuenced nitrate sorption (2010) 61896195.
[43] I.A.W. Tan, A.L. Ahmad, B.H. Hameed, J. Hazard. Mater. 154 (2008) 337346.
and desorption by biochar. [44] R.L. Tseung, S.K. Tseng, J. Hazard. Mater. 136 (2006) 671680.
[45] F.C. Wu, R.L. Tseng, R.S. Juang, Sep. Purif. Technol. 47 (2005) 1019.
Acknowledgments [46] C.K. Sing, J.N. Sahu, K.K. Mahalik, C.R. Mohanty, R.B. Mohan, B.C. Meikap, J.
Hazard. Mater. 153 (2008) 221228.
[47] D.E. Clay, Z. Zheng, Z. Lui, S.A. Clay, T.P. Trooien, J. Environ. Qual. 33 (2004)
This project was supported by Agriculture and Food Research 338342.
Initiative Competitive Grant No. 2011-67009-30076 from the [48] K. Jaafari, T. Ruiz, S. Elmaleh, J. Coma, K. Benkhouja, Chem. Eng. J. 99 (2004)
153160.
USDA National Institute of Food and Agriculture. [49] K. Ramakrishnan, C. Namasivayam, J. Environ. Eng. Manage. 19 (2009) 173
178.
Appendix A. Supplementary data [50] ASTM, Standard Test Method for Carbon Black-Iodine Absorption, ASTM
D1510-09A, American Society for Testing and Materials, West Conshohocken,
PA, 2009.
Supplementary data associated with this article can be found, in [51] G.E. Rayment, F.R. Higginson, Australian Laboratory Handbook of Soil and
the online version, at http://dx.doi.org/10.1016/j.micromeso.2013. Water Chemical Methods, Inkata Press, Melbourne, 1992. p. 330.
[52] ASTM, Standard Test Method for the Analysis of Wood Charcoal (D176284),
05.023. 2007.
[53] R. Chintala, D.E. Clay, T.E. Schumacher, D.D. Malo, J. Julson, Anal. Lett. 46
References (2013) 532538.
[54] J.K. Friel, C.S. Skinner, S.E. Jackson, H.P. Longerich, Analyst 115 (1990) 269273.
[55] V.D. Zheljazkov, N.E. Nielson, Plant Soil 178 (1996) 5966.
[1] D.A. Martens, W. Dick, Biol. Fertil. Soils 38 (2003) 144153.
[56] US, Environmental Protection Agency, Microwave Assisted Acid Digestion of
[2] R.A. Ortega, D.G. Westfall, G.A. Peterson, Commun. Soil Sci. Plant Anal. 36
Sediments, Sludges, Soils, and Oils, Method 3051a. Ofce of Solid Waste and
(2005) 28752887.
Emergency Response, US Government Printing Ofce, Washington, DC, 1998.
[3] R.L. Partt, Chemical properties of variable charge soils, in: B.K.G. Theng (Ed.),
[57] L.W. Zelazny, H.E. Liming, A.N. Van Wormhoudt, Charge analysis of soils and
Soils with Variable Charge. New Zealand Society of Soil Science, Offset
anion exchange, in: D.L. Sparks (Ed.), Methods of Soil Analysis. Part 3. SSSA
Publications, Palmerston North, New Zealand, 1980, pp. 167194.
Book Ser. 5, SSSA, Madison, WI, 1996, pp. 12311253.
[4] M. Chabani, A. Amrane, A. Bensmaili, Desalination 197 (2006) 117123.
[58] SAS Institute, SAS Stat., vols. 12. Release 8.2., Cary, NC, 2003.
[5] X. Xu, B.Y. Gao, Q.Y. Yue, Q.Q. Zhang, Bioresour. Technol. 101 (2010) 8558
[59] R.G.W. Steel, J.H. Torrie, Principles and Procedures of Statistics: A Biometrical
8564.
Approach, second ed., McGraw-Hill Book Co., New York, 1980.
[6] Z. Li, Micropor. Mesopor. Mater. 61 (2003) 181188.
[60] J.W. Gaskin, C. Steiner, K. Harris, K.C. Das, B. Bibens, Trans. ASABE 51 (2008)
[7] M. Islam, R. Patel, J. Hazard. Mater. 169 (2009) 524531.
20612069.
[8] A. Pintar, J. Batista, J. Levec, Chem. Eng. Sci. 56 (2001) 15511559.
[61] K.A. Spokas, J.M. Novak, C.E. Stewart, K.B. Cantrell, M. Uchimiya, M.G. Dusaire,
[9] B. Moreno, M.A. Gomez, A. Ramos, J. Gonzalez-Lopez, E. Hontoria, J. Hazard.
K.S. Ro, Chemosphere 85 (2011) 869882.
Mater. 127 (2005) 180186.
[62] C. Moreno-Castilla, M.A. Ferro-Garcia, J.P. Joly, I. Bautista-Toledo, F. Carrasco-
[10] M. Kumar, S. Chakraborty, J. Hazard. Mater. 135 (2006) 112121.
Marin, J. Rivera-Utrilla, Langmuir 11 (1996) 43864392.
[11] Y.H. Huang, T.C. Zhang, J. Environ. Eng. 128 (2002) 604610.
[63] R.R. Bansode, J.N. Losso, W.E. Marshall, R.M. Rao, R.J. Portier, Bioresour.
[12] A. Pintar, J. Batista, Appl. Catal., B 63 (2006) 150159.
Technol. 94 (2004) 129135.
[13] J.J. Schoeman, A. Steyn, Desalination 15 (2003) 1526.
[14] A. Elmidaoui, M.A. Menkouchi Sahli, M. Tahaikt, L. Chay, M. Taky, M. Elmghari,
M. Hafsi, Desalination 153 (2002) 389397.

Das könnte Ihnen auch gefallen