Sie sind auf Seite 1von 27

Sorptive removal of salicylic acid and ibuprofen from

aqueous solutions using pine wood fast pyrolysis biochar

By
Matthew Essandoh , Bidhya Kunwar , Charles U.
Pittman Jr. , Dinesh Mohan , Todd Mlsna
Presented by
Muhammad Zeeshan Ul-Haq

Abstract
Pine wood biochar, prepared at 698 K with a residence time of 2030 s in an auger-fed reactor, was
used as a 3-dimensional adsorbent to remove salicylic acid and ibuprofen from aqueous solutions.
This biochar was characterized by FT-IR spectroscopy, scanning electron microscopy, transmission
electron microscopy, surface area determination, and zero point charge measurements. Batch
sorption studies were carried out at pH values from 2 to 10, adsorbate concentrations from 25 to
100 mg/L and temperatures from 298 to 318 K. The adsorption of both adsorbates was highest at
low pH values, dropped as pH increased and then exhibited a second increase related to the pKa of
these carboxylic acid adsorbates. This was followed by a further drop at high pH. Conjugate
acid/base equilibria of the adsorbates and the phenolic hydroxyl and carboxylic acid biochar sites
versus pH dominated the mechanism. Sorption followed pseudo-second order kinetics. Sorption
was evaluated from 298 to 318 K using the Freundlich, Langmuir, RedlichPeterson, Toth, Sips, and
RadkePrausnitz adsorption isotherm models. Langmuir adsorption capacities for both salicylic acid
and ibuprofen were 22.70 and 10.74 mg/g, respectively. This low surface area pinewood biochar
(1.35 m2/g) can adsorb far more adsorbate compared to commercial activated carbons per unit of
measured surface area. Methanol stripping achieved 93% and 88% desorption of salicylic acid and
ibuprofen, respectively, from the spent biochar, and 76% and 72% of the initial salicylic acid and
ibuprofen adsorption capacity, respectively, remained after four full capacity equilibrium recycles.

Introduction
Wastewater pollution is an urgent environmental concern. The negative
effects of pollution on both humans and the environment have become a
subject of intense discussion. Pharmaceutical pollutants are increasingly
important and require immediate attention . Pharmaceutical products enter
the aquatic environment through municipal waste water treatment, plant
effluent, animal excreta, direct discharge of pharmaceuticals into
waterbodies, and septic tank leakage. Negative effects on humans and
animals include the suppression of embryonic cell growth in humans, the
incapacitation of liver and gills in fish, and the inhibition of Gram-positive
bacteria. Despite public concerns about the negative effects that
pharmaceuticals have on aquatic ecosystems, there are currently no federal
laws defining the amount of these chemicals permitted in waste water
streams or in drinking water.

Introduction (contd..)
Ibuprofen 2 is a non-steroidal anti-inflammatory drug with a wide global
consumption. It is produced on an industrial scale as a race-mate which is
generally excreted in the form of conjugates such as 2-[4-(2-hydroxy-2methylpropyl)phenyl]propanoic acid, 2-[4-(2carboxypropyl)phenyl]propanoic acid, and 2-phenylpropanoic acid.
Ibuprofen is known to have endocrine disruptive activity . Aqueous solutions
of these pharmaceutical pollutants have been treated by physical,
biochemical and chemical processes . However, these techniques can be
expensive. Thus, low cost methods to remove pharmaceutical contaminants
from aqueous solution will be well embraced. Low cost adsorbents can be
employed to remove recalcitrant compounds from aqueous solution
inexpensively and can be effective for removing both organic and inorganic
contaminants .

Method
The biochar used for adsorption experiments was produced as a byproduct of
bio-oil produced by fast pyrolysis of pine wood chips in an auger-fed reactor at a
temperature of 698 K using a feed rate of 12.5 kg/h . The residence time of the
pine wood chips in the hotzone was 2030 s and the approximate gas residence
time in the reactor was 2 s. The chips were preheated to between 383 and 393
K, and a proprietary heat transfer method was used to speed the temperature
rise to 698 K in the hotzone. The powdered biochar was collected from the
system by capturing it in an enclosed pot after moving past the hotzone. After
subsequent removal from this container at room temperature, it was washed
several times with distilled water to remove salt impurities and water-soluble
organic residuals. Biochar was sieved to a particle size distribution of 100600
lm (0.10.6 mm), and moisture was removed by heating to 383 K for 12 h. The
adsorbent was then stored in a closed vessel at 338 K for several days until
needed.

Biochar Properties
High carbon content (60 95% C)
Resistant to biodegradation
Significant adsorptive qualities (similar to activated
carbon)
Nutrients (& contaminants) essentially lock on to the
structure
Increases moisture holding capacity
Enhances microbial biomass

Biochar Properties:
Process Conditions
Enhanced biochar yields from:
Lower temperatures (<400C)
Higher pressures
Longer vapour residence time
Slower heating rate
Larger particle size

Biochar Properties:
Process Conditions
As temperature
increases:
Biochar yield
decreases
Fixed carbon
increases
Surface area
increases
Ash content
increases

Tem
pe

Stru
Dev ctural
e lo
pm
ent

ratu
re

FT-IR
The infrared spectrum of powdered biochar was
obtained using a Thermo Nicolet 6700 FT-IR
spectrometer. A total of 64 scans were taken
from 4000 to 500 cm_1 with a resolution of 4
cm_1.

Scanning electron microscopy


(SEM)
A JEOL JSM-6500F FE-SEM operated at 5 kV was
used to examine the biochars surface
morphology.

Transmission electron microscopy


(TEM)
The biochar was examined using a JEOL model
2100 TEM operated at 80 kV. A small amount of
adsorbent was mixed with 100% ethanol,
sonicated for about 4 min, and allowed to stand
for 24 h. A drop of this suspension was applied
to a carbon film on 300 mesh copper grid and
allowed to dry in air before TEM analysis

Surface area measurements and


elemental analysis
The biochars BET surface area (1.35 m2/g) was
determined using a Gemini 2375 V 1.00 surface
area analyzer

Zero point charge determination


Zero point charge was determined using 0.01 M NaCl
aqueous solutions with pH values ranging from 2 to
10 at pH intervals of 2. About 10 mL was added to
0.1 g of the biochar and the mixtures were swirled
for 48 h. pH was adjusted using either 0.1 N HCl or
0.1 N NaOH solutions. The supernatant was decanted
and its pH measured using an ORION model 210 pH
meter. The point of zero charge was obtained by
plotting supernatant pH verses the initial solution pH.

Adsorption studies
The adsorbate concentrations were varied from 25 to 100 mg/L for
kinetic studies. Kinetic studies for ibuprofen and salicylic acid
adsorption were carried out at pH values of 3 and 2.5, respectively,
at 298, 308, and 318 K. A known amount of biochar was added to
25 mL solutions containing different adsorbate concentrations in 40
mL amber glass vials. Samples were then swirled at 200 rpm for 16
h. After equilibration, the samples were filtered through Whatman
No. 1 filter paper. The amount of adsorbate remaining in the filtrate
was determined at their kmax wavelengths of 220 nm for ibuprofen
and 298 nm for salicylic acid. Samples were analyzed in duplicate
and their average absorbances used

Objectives

The objective of this work is to establish if pine wood biochar can be used to
successfully remove ibuprofen and salicylic acid from water. Hence, it is available
as a byproduct of bio-oil, an emerging renewable liquid fuel. Biochars have fuel
value (as a coal substitute) or can be used for carbon sequestration or soil
conditioners . A value-added application as a low cost widely applicable adsorbent
is desirable and could increase the value of this biofuel byproduct. This research
investigates the adsorption behavior of salicylic acid and ibuprofen at different
concentrations, pH values, and temperatures.

pH ADSORPTION
pcz

The point of zero charge (pHpzc) is the pH at


which the net charge on the surface is zero. This
was found to be 2 for the pine-wood biochar.
When pHpzc < pH of the solution, the biochars
surface will be negatively charged. When the
pHpzc > pH of the solution, the surface of the
biochar will be positively charged

SEM Images
SEM images of the biochar.
The biochar has a macroporous surface texture.

pH effect on adsorption
At higher pH, (pH greater than the pKa of the adsorbates), both adsorbates
are increasingly present as their carboxylate anions (RACOO _).
Simultaneously, the biochar becomes increasingly negatively charged as
pH rises (pH > pzc), causing increased electrostatic repulsion of adsorbate
carboxylate anions and decreasing the adsorption capacity. This trend is
seen from pH 2 or 3 through 10. However, both adsorbates show a region
of intermediate pH values where adsorption increases, maximizes and then
continues to drop as pH increases. These adsorptions peak at about pH 8
for ibuprofen and 6 for salicylic acid, where both adsorbates are 99.9% in
their carboxylate anion forms (each approximately 3 pH units above their
pKa). For example, at pH values above 8, almost all salicylic acid (pKa =
2.98) exists as RCOO_ and the surface of the bio-char will have a substantial
negative charge density. Thus, little salicylic acid can adsorb.

Effect of solution pH on adsorbate


adsorption

Solution pH plays a crucial role in ibuprofen and salicylic


acid adsorption from aqueous solution. The percent
removal of both ibuprofen and salicylic acid is higher in
acidic solutions than in the basic region

pH effect on adsorption

Temperature effect on adsorption

Adsorption at 318 K at different ibuprofen [pH = 3, amount of biochar was 4 g/L] and salicylic acid [pH = 2.5, amount of biochar
was 4 g/L] concentrations.

Temperature effect on adsorption

Adsorption at 318 K at different ibuprofen [pH = 3, amount of biochar was 4 g/L] and salicylic acid [pH = 2.5, amount of biochar
was 4 g/L] concentrations.

Repeatability

Biochar recycling using an adsorbent dose of 4 g/L and an adsorbate concentration of 100 mg/L, at a pH of 2.5 for
salicylic acid and pH 3 for ibuprofen at 318 K. Methanol (20 mL _ 2) was used as a stripping agent.

Conclusions

Fast pyrolysis pinewood biochar was characterized and this biochar


successfully adsorbed ibuprofen and salicylic acid from aqueous solutions.
Adsorption of both ibuprofen and salicylic acid followed pseudo-second order
kinetics.
Adsorption was favored at low pH values because the chars pHpzc is 2 relative
to the pKas of ibuprofen (4.91) and salicylic acid (2.98).
Adsorption capacities were 22.70 and 10.74 mg/g at 318 and 308 K, for
salicylic acid and ibuprofen, respectively.
Despite low surface area, the biochar (19% O) appears to adsorb by imbibing
water and adsorbate within its solid structure.
Methanol was able to achieve 93% and 88% desorption of salicylic acid and
ibuprofen, respectively, from the spent biochar.
Additionally, 76% and 72% of the biochars initial adsorption capacity for
salicylic acid and ibuprofen, respectively, remained after four full capacity
equilibrium recycles.

References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.

16.
17.

Y. Zhang, S.U. Geiben, C. Gal, Carbamazepine and diclofenac: removal in wastewater treatment plants and occurrence in water bodies, Chemosphere 73
(2008) 11511161.
F.A. Caliman, M. Gavrilescu, Pharmaceuticals, personal care products and endocrine disrupting agents in the environment a review, Clean Soil Air
Water 37 (2009) 277303.
F. Pomati et al., Effects of a complex mixture of therapeutic drugs at environmental levels on human embryonic cells, Environ. Sci. Technol. 40
(2006) 24422447.
K. Fent, A.A. Weston, D. Caminada, Ecotoxicology of human pharmaceuticals, Aquat. Toxicol. 76 (2006) 122159.
K.T. Elvers, S.J.L. Wright, Antibacterial activity of the anti-inflammatory compound ibuprofen, Lett. Appl. Microbiol. 20 (1995) 8284.
T. Heberer, Occurrence, fate, and removal of pharmaceutical residues in the aquatic environment: a review of recent research data, Toxicol. Lett. 131
(2002) 517.
T.A. Ternes, Occurrence of drugs in German sewage treatment plants and rivers, Water Res. 32 (1998) 32453260.
H.-R. Buser, T. Poiger, M.D. Muller, Occurrence and environmental behavior of the chiral pharmaceutical drug ibuprofen in surface waters and in wastewater,
Environ. Sci. Technol. 33 (1999) 25292535.
A. Ziylan, N.H. Ince, The occurrence and fate of anti-inflammatory and analgesic pharmaceuticals in sewage and fresh water: treatability by
conventional and non-conventional processes, J. Hazard. Mater. 187 (2011)
2436.
D. Mohan, A. Sarswat, Y.S. Ok, C.U. Pittman Jr., Organic and inorganic contaminants removal from water with biochar, a renewable, low cost and
sustainable adsorbent a critical review, Bioresour. Technol. 160 (2014) 191 202.
C.U. Pittman Jr. et al., Characterization of bio-oils produced from fast pyrolysis of corn stalks in an auger reactor, Energy Fuels 26 (2012) 38163825.
B. Kunwar, T. Mlsna, Upgrading bio-oil with a combination of synthesis gas and alcohol, Energy Fuels 58 (2013) 9092.
M. Rondon, J. Lehmann, J. Ramirez, M. Hurtado, Biological nitrogen fixation by common beans (Phaseolus vulgaris L.) increases with bio-char additions, Biol.
Fertil. Soils 43 (2007) 699708.
D.A. Laird, The charcoal vision: a winwinwin scenario for simultaneously producing bioenergy, permanently sequestering carbon, while improving soil
and water quality, Agron. J. 100 (2008) 178181.
D. Mohan, S. Rajput, V.K. Singh, P.H. Steele, C.U. Pittman Jr., Modeling and evaluation of chromium remediation from water using low cost bio-char, a
green adsorbent, J. Hazard. Mater. 188 (2011) 319333.
C.E. Brewer, K. Schmidt-Rohr, J.A. Satrio, R.C. Brown, Characterization of biochar from fast pyrolysis and gasification systems, Environ. Prog. Sustainable
28 (2009) 386396.
F.-M. Pellera et al., Adsorption of Cu(II) ions from aqueous solutions on biochars prepared from agricultural by-products, J. Environ. Manage. 96 (2012) 354
2.

References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.

A.S. Mestre, J. Pires, J.M.F. Nogueira, A.P. Carvalho, Activated carbons for the adsorption of ibuprofen, Carbon 45 (2007) 19791988.
S. Lagergren, Zur theorie der sogenannten adsorption geloster stoffe, K. Sven. Vetenskapsakad. Handl. 24 (1898) 139.
Y.-S. Ho, Review of second-order models for adsorption systems, J. Hazard. Mater. 136 (2006) 681689.
Z. Aksu, D. Akpinar, Modelling of simultaneous biosorption of phenol and nickel(II) onto dried aerobic activated sludge, Sep. Purif. Technol. 21 (2000) 879
9.
R.S. Summers, Strategies and Technologies for Meeting SDWA Requirements, Technomic Publishing Company Inc., Lancaster, 1993.
J.M. Montgomery, Water Treatment Principles and Design, Wiley-Interscience, New York, 1985.
K.Y. Foo, B.H. Hameed, Insights into the modeling of adsorption isotherm systems, Chem. Eng. J. 156 (2010) 210.
D. Mohan, A. Sarswat, V.K. Singh, M. Alexandre-Franco, C.U. Pittman Jr., Development of magnetic activated carbon from almond shells for
trinitrophenol removal from water, Chem. Eng. J. 172 (2011) 11111125.
M. Otero, C.A. Grande, A.E. Rodrigues, Adsorption of salicylic acid onto polymeric adsorbents and activated charcoal, React. Funct. Polym. 60 (2004)
203213.
R.G. Combarros, I. Rosas, A.G. Lavin, M. Rendueles, M. Diaz, Influence of biofilm on activated carbon on the adsorption and biodegradation of salicylic acid i
n wastewater, Water Air Soil Pollut. 225 (2014) 1858.
J. Huang, G. Wang, K. Huang, Enhanced adsorption of salicylic acid onto a b- naphthol-modified hyper-cross-linked poly(styrene-co-divinylbenzene) resin
from aqueous solution, Chem. Eng. J. 168 (2011) 715721.
F. Liu, M. Xia, Z. Fei, J. Chen, A. Li, Adsorption selectivity of salicylic acid and 5- sulfosalicylic acid onto hypercrosslinked polymeric adsorbents, Front. Environ
. Sci. Eng. 1 (2007) 7378.
A.S. Mestre et al., Waste-derived activated carbons for removal of ibuprofen from solution: role of surface chemistry and pore structure, Bioresour.
Technol. 100 (2009) 17201726.
R. Baccar, M. Sarra, J. Bouzid, M. Feki, P. Blanquez, Removal of pharmaceutical
compounds by activated carbon prepared from agricultural by-product, Chem. Eng. J. 211212 (2012) 310317.
S.P. Dubey, A.D. Dwivedi, M. Sillanpaa, K. Gopal, Artemisia vulgaris-derived mesoporous honeycomb-shaped activated carbon for ibuprofen adsorption,
Chem. Eng. J. 165 (2010) 537544.
I. Langmuir, The adsorption of gases on plane surfaces of glass, mica and platinum, J. Am. Chem. Soc. 40 (1918) 13611403.
H.M.F. Freundlich, Over the adsorption in solution, J. Phys. Chem. 57 (1906) 385471.
R. Sips, On the structure of a catalyst surface, J. Chem. Phys. 16 (1948) 490 495.
O. Redlich, D.L. Peterson, A useful adsorption isotherm, J. Phys. Chem. 63 (1959). 10241024.
C.J. Radke, J.M. Prausnitz, Adsorption of organic solutes from dilute aqueous solution on activated carbon, Ind. Eng. Chem. Fundam. 11 (1972) 445451.

Das könnte Ihnen auch gefallen