You are on page 1of 12

Environmental Nanotechnology, Monitoring & Management 20 (2023) 100788

Contents lists available at ScienceDirect

Environmental Nanotechnology, Monitoring & Management


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

U(VI) removal from water by novel P-modified activated carbon derived


from polymeric microspheres
R. Olchowski a, b, B. Podkościelna c, B. Zawisza d, K. Morlo a, R. Dobrowolski a, *
a
Department of Analytical Chemistry, Institute of Chemical Sciences, Faculty of Chemistry, Maria Curie-Sklodowska University, M. C. Sklodowska Sq. 3, 20-031 Lublin,
Poland
b
Department of Pharmacology, Toxicology and Environmental Protection, Faculty of Veterinary Medicine, University of Life Sciences, Akademicka Sq. 12, 20-950
Lublin, Poland
c
Department of Polymer Chemistry, Institute of Chemical Sciences, Faculty of Chemistry, Maria Curie-Sklodowska University, M. C. Sklodowska Sq. 2, 20-031 Lublin,
Poland
d
Institute of Chemistry, University of Silesia, Szkolna 9, 40-006 Katowice, Poland

A R T I C L E I N F O A B S T R A C T

Keywords: Phosphorous-modified mesoporous carbon was obtained by thermochemical modification-carbonization of


Uranium adsorption originally synthesized polymeric microspheres with orthophosphoric acid. The obtained material was charac­
Seawater terized by physicochemical methods. The adsorption properties of the obtained carbon material towards U(VI)
River water
ions were studied and compared to the pristine polymeric microspheres. The highest static adsorption capacity of
TXRF
U(VI) onto studied carbonaceous material (150 mg ‧ g− 1) was observed at pH = 1.8 after 2 min contact time. The
Mesoporous carbon
pseudo-first-order and Langmuir theoretical models were best fitted to the experimental kinetics and equilibrium
data. Studied ions present at concentration levels found in natural waters had a negligible effect on U(VI)
adsorption onto synthesized carbonaceous material. The adsorption mechanism of U(VI) onto the studied carbon
material consisted of surface complexation and interactions with π-π electrons and ion exchange. The synthesized
material can be easily regenerated by 1 mol ‧ L-1 NaHCO3 (pH = 12.6). The phosphorous-modified carbon was
successfully used both for efficient removal of U(VI) ions from model water samples and for the enrichment of
trace levels of U from water samples prior to its determination by the TXRF technique. The LOD for the proposed
procedure of U determination by TXRF technique after the enrichment step was 0.011 µg ‧ L-1.

1. Introduction 2014). The mine tailings near rivers pose a significant threat to the
environment. The uranium content in this waste can be high (ca.
Uranium is an alpha-decay radioactive heavy metal found in many 1.5–2.0 mg kg− 1) compared to naturally occurring uranium (0.02–43.7
naturally occurring minerals, the most common of which is uraninite µg ‧ L-1). A possible flood could breach the waste and consequently, huge
(also known as pitchblende). The U(IV) and U(VI) are the main oxida­ amounts of uranium (mainly as U(VI)) would get into the river water
tion states of this metal in various compounds, but only U(VI) com­ (Garboś et al., 2014; Borzęcki et al., 2018). Furthermore, seawater used
pounds are readily soluble in water and thus highly mobile in the as a cooling agent in the nuclear industry can contain at least 1.3 mg ‧ L-1
environment. Moreover, U(VI) can bond to inorganic (e.g., CO2– 3 , OH ,

of U (Dauer et al., 2011; Koma and Murakami, 2019). For comparison
3-
PO4 ) and organic ligands (e.g., carboxyl, citrate, oxalate). The chemical the uranium is present in natural seawater at 2.9–6.1 µg ‧ L-1. The nuclear
form of U(VI) in an aqueous solution depends on the pH. The uranium power plant disaster (e.g. Fukushima accident in Japan) can be
presence in the environment is not only controlled by natural sources responsible for the release of huge amounts of radioactive uranium into
but also by human activities. It is related to nuclear power engineering, the environment (Garboś et al., 2014; Dauer et al., 2011; Koma and
nuclear weapons tests, the mining industry, the fluorescent paints in­ Murakami, 2019). Such accidents could be responsible for the increased
dustry, photographic and ceramic industries, aircraft, the military in­ U(VI) concentration in the naturally occurring water, possibly higher
dustry, and using phosphate-fertilizers in agriculture (Garboś et al., than the WHO maximum tolerable concentration of uranium in drinking

* Corresponding author.
E-mail addresses: rafal.olchowski@poczta.umcs.lublin.pl (R. Olchowski), beatapod@umcs.pl (B. Podkościelna), beata.zawisza@us.edu.pl (B. Zawisza), kinga.
morlo1@gmail.com (K. Morlo), rdobrow@poczta.umcs.lublin.pl (R. Dobrowolski).

https://doi.org/10.1016/j.enmm.2023.100788
Received 25 August 2022; Received in revised form 22 January 2023; Accepted 26 January 2023
Available online 28 January 2023
2215-1532/© 2023 Elsevier B.V. All rights reserved.
R. Olchowski et al. Environmental Nanotechnology, Monitoring & Management 20 (2023) 100788

water (30 µg/L) (Who, 2003). The U(VI) compounds are highly toxic for Aldrich, MO, USA), potassium chloride (>99 %) (Sigma-Aldrich, MO,
living organisms, especially if absorbed. Its toxicity is not only related to USA), sodium sulfate (>90 %) (Sigma-Aldrich, MO, USA), iron(III) ni­
the emission of alpha radiation but also the binding of this element to trate nonahydrate (>90 %) (Sigma-Aldrich, MO, USA), calcium nitrate
the organic ligands of proteins and nucleotides, blocking their biological monohydrate (>90 %) (Sigma-Aldrich, MO, USA), magnesium nitrate
activity. U(VI) accumulates mainly in bones and kidneys. In the former (>90 %) (Sigma-Aldrich, MO, USA), sodium hydroxide (>90 %) (Sigma-
Ca2+ ions are replaced by UO2+ 2 ions with their half-time removal from Aldrich, MO, USA), poly(vinyl alcohol) (PVA Mw = 72000) (Merck,
bones being 180–833 days. Chronic exposure to U(VI) can lead to bone Darmstadt, Germany), benzyl alcohol (Merck, Darmstadt, Germany),
and kidney damage and it can also result in cancer (Garboś et al., 2014; styrene (St) (Merck, Darmstadt, Germany), diethylene glycol dimethy­
Craft et al., 2004). Thus, the efficient removal of U(VI) from water is a lacrylate (DEGDMA) (Merck, Darmstadt, Germany), Ga standard solu­
crucial task. tion (1000 mg L-1) (Merck, Darmstadt, Germany), U(VI) standard
Among various metal ion removal technologies, adsorption is the solution (1000 mg L-1) (Merck, Darmstadt, Germany), α,α’-azoiso-bis-
most popular method due to its versatility, simplicity and efficiency, butyronitrile (AIBN) (Fluka, Buchs, Switzerland), acetone (Ace) (Avan­
especially using carbonaceous materials. The large specific surface area tor Performance Materials Poland S.A., Gliwice, Poland), uranyl acetate
of the carbonaceous materials and the easy chemical surface modifica­ dihydrate (>99 %) (Fisher Scientific, Lund, Sweden).
tion are responsible for their high maximum adsorption capacity, and Certified reference materials of seawater (CASS-6, National Research
selectivity towards selected metal ions. Moreover, carbonaceous mate­ Council, Canada) and river water (SLRS-6, National Research Council,
rials are characterized by good mechanical strength, chemical, thermal Canada) with known U concentration were obtained from LGC Stan­
and radiation resistance, low economic cost, and high re-usability dards (Teddington, UK). Throughout all experiments, the double
(Soylak and Doan, 1996; Caccin et al., 2013; Ma et al., 2020). In the distilled Milli-Q water from Millipore (Merck, Darmstadt, Germany) was
literature, various pristine and chemically modified carbonaceous ma­ used.
terials, such as activated carbons (Caccin et al., 2013; Ma et al., 2020; Li
et al., 2014; Han et al., 2018; Mellah et al., 2006), graphene oxides 2.2. Synthesis of DEGDMA based polymeric microspheres
(Verma and Dutta, 2015; Dutta et al., 2019; Sun et al., 2013), carbon
nanotubes (Zare et al., 2015; Mishra et al., 2016; Fasfous and Dawoud, The kraft lignin (5 g) was dissolved in benzyl alcohol (14 mL). The
2012; Shah et al., 2013), carbon nanofibers (Sun et al., 2016), biochars PVA (2.2 g) and purified water (180 mL) were placed in a 250 mL round-
(Qiu et al., 2022; Liang et al., 2021)or ordered mesoporous carbons bottomed flask equipped with a mechanical stirrer and a thermometer.
(Wang et al., 2012; Tian et al., 2011)were specified as adsorbents for The mixture was stirred to dissolve the suspension stabilizer for 0.5 h at
uranyl ion removal from aqueous solutions. Additionally, it is well- 80 ◦ C. Next, the monomers: DEGDMA and St (1:1 mol.%) were added to
known that U(VI) ions possess a strong affinity toward phosphorous the dissolved lignin. Finally, the organic mixture was added to the
groups, which is related to the creation of complex compounds between aqueous phase. The copolymerization of the reactive mixture was per­
uranyl ions and phosphorous functional groups. Incorporating phos­ formed in the aqueous medium in the presence of the initiator (2 wt%).
phorous groups on the surface of the carbonaceous material during The reaction mixture was stirred at 300 rpm for 8 h at 85 ◦ C. When the
carbonization can improve the maximum static adsorption capacity and reaction was completed the resulting microspheres were filtered and
selectivity of the material towards U(VI) ions (Puziy et al., 2020). washed with hot water (2 L). The yield of reaction: 96 % (Goliszek et al.,
In this work, novel activated carbon modified by orthophosphoric 2018; Podkościelna et al., 2017; Puziy et al., 2020; Wnuczek et al.,
acid during carbonization of the polymeric microspheres has been ob­ 2020). The obtained polymeric microspheres were marked as PS.
tained and applied for the first time for the efficient U(VI) removal from
model polluted water samples. First, the obtained carbonaceous mate­ 2.3. Chemical modification and carbonization of polymeric microspheres
rial was subjected to physicochemical studies. In the second step, studies
concerning U(VI) ions removal from an aqueous solution by the syn­ The polymeric microspheres (1 g) were mixed with 60 wt% H3PO4
thesized activated carbon were performed. Especially the effect of so­ (15 mL) and the obtained mixture was set aside for 24 h. After that, the
lution pH, contact time, U(VI) concentration, and presence of mixture was dried in a laboratory oven at 120 ◦ C for another 24 h and
competitive ions (HCO–3, Cl-, NO–3, SO2- 2+ 2+ 3+
4 , Mg , Ca , Fe ) on the U(VI) heated in a quartz tubular furnace under nitrogen atmosphere (1 L ·
adsorption onto the synthesized carbonaceous material were studied. min− 1) according to the following time–temperature program:
The U(VI) adsorption properties of the synthesized activated carbon 25–360 ◦ C (ramp 10 ◦ C · min− 1), 360–450 ◦ C (ramp 2 ◦ C · min− 1),
were compared to the initial polymeric microspheres. Moreover, the 450–700 ◦ C (ramp 10 ◦ C · min− 1) and at 700 ◦ C for 3 h (pyrolysis step).
theoretical adsorption kinetic (pseudo-first order, pseudo-second order, Next, the black solid was washed thoroughly with distilled water until
Elovich) and isotherm (Langmuir, Freundlich) models were fitted to the the effluent pH was close to 7. Finally, the black solid was dried at 120 ◦ C
experimental data for the carbonaceous material. Furthermore, the for 24 h (Puziy et al., 2020; Huang et al., 2018). The obtained carbo­
mechanism of U(VI) adsorption onto the synthesized carbonaceous naceous material was denoted as PS_C.
material was proposed. The regeneration studies of U-loaded activated
carbon were also performed. Finally, the phosphorous-modified acti­ 2.4. Physicochemical characteristics
vated carbon was successfully utilized for U(VI) removal from the model
polluted water samples. The nitrogen adsorption/desorption experiments were conducted at
− 196 ◦ C with an ASAP 2420 analyzer (Micromeritics Inc., Norcross, GA,
2. Materials and methods USA). The samples were degassed at 120 ◦ C in a vacuum for 12 h before
measurements. The BET surface area (SBET), total pore volume (VT), and
2.1. Chemicals and eluents BJH pore size distribution (PSD) were estimated using the desorption
branch of the nitrogen adsorption/desorption isotherm.
The following materials and chemicals were used during the Morphological analysis of the PS and PS_C particles was carried out
research: kraft lignin (Sigma-Aldrich, MO, USA), orthophosphoric acid by the automated morphology particle size and shape analyzer (Mor­
(85 %) (Sigma-Aldrich, MO, USA), arsenazo III sodium salt (>90 %) phologi G3, Malvern) microscope and the scanning electron microscope
(Sigma-Aldrich, MO, USA), nitric acid (65 %) (Sigma-Aldrich, MO, (SEM) Carl Zeiss Ultra Plus (Carl Zeiss, Jena, Germany) equipped with
USA), hydrochloric acid (36 %) (Sigma-Aldrich, MO, USA), sodium bi­ an energy dispersive X-ray (EDX) detector BrukerAXS (Bruker, Karls­
carbonate (>90 %) (Sigma-Aldrich, MO, USA), potassium nitrate (>90 ruhe, Germany). The SEM microscope was also equipped with secondary
%) (Sigma-Aldrich, MO, USA), sodium chloride (>99 %) (Sigma- electron (SE) and backscattered electron (BSE) detectors. All

2
R. Olchowski et al. Environmental Nanotechnology, Monitoring & Management 20 (2023) 100788

experiments were carried out under the required conditions (20-kV ac­ The effect of competitive ions (HCO–3, Cl-, NO–3, SO2- 2+
4 , Mg , Ca ,
2+

celeration voltage and 5-nA probe current). Fe3+) on the U(VI) adsorption onto the synthesized carbonaceous ma­
The elemental composition of the synthesized materials was studied terial was studied. The 20 mg of PS_C was mixed with 5 mL of an
using SEM-EDX and X-ray photoelectron spectroscopy (XPS) techniques. aqueous solution containing 5.5 mg ‧ L-1 U(VI) and competitive ions in
XPS measurements were conducted using a Multi-Chamber Analytical the concentration range 0–1000 mol ‧ L-1 for anions and 0–10 mmol ‧ L-1
System (Prevac, Rogów, Poland) equipped with monochromatic Kα-Al for cations. The pH of all aqueous solutions was adjusted to 1.8. After
radiation (1486.6 eV) (Gammadata Scienta, Uppsala, Sweden) and an X- shaking the carbonaceous slurries for 24 h, the initial and equilibrium U
ray power of 450 W. Binding energies were referenced to the carbon C1s (VI) concentrations were determined, and the adsorption of U(VI) ions
peak at 285 eV. onto PS_C was calculated, as described above. Finally, the relative
The presence of surface functional groups in the studied materials adsorption was calculated according to the following equation:
was examined by the XPS and attenuated total reflectance (ATR) with a
the Fourier transform infrared (FT-IR) spectroscopy techniques. The Adsorption = 100% (2)
amax.
ATR FT-IR spectra of all samples were recorded using the Bruker
TENSOR 27 FTIR spectrophotometer. There were 32 scans accumulated where Adsorption is the relative U(VI) adsorption onto the PS_C [%] and
in the range of 600 – 4000 cm− 1. The spectrophotometer was equipped amax. is the maximum adsorption of U(VI) in the studied adsorption
with a diamond crystal. system [mg ‧ g− 1].
Thermal analysis was carried out on a Netzsch, STA 449 Jupiter F1
(Germany). The samples were heated from 30 to 800 ◦ C at a rate of 2.6. Regeneration studies
10 ◦ C⋅ min− 1 in the dynamic atmosphere of helium (25 mL ⋅ min− 1). The
sensor thermocouple type S TG–DSC was used with an empty Al2O3 The various liquid media (HCl, HNO3, NaHCO3) with concentrations
crucible as the reference. The loss mass temperatures (T5,10,50%), peak
in the range 0.01–1.00 mol ‧ L-1 were used for regeneration studies of U-
maximum decomposition temperatures (Tmax), and residual mass at loaded PS-C. Also, the pH effect of 1 mol ‧ L-1 NaHCO3 on the regener­
800 ◦ C (RM) were determined.
ation degree of U-loaded PS_C was studied in the pH range of 8.9–12.6.
The TXRF analysis was carried out using a S4 T-STAR (Bruker AXS First, 100 mg of PS_C was mixed with 25 mL of an aqueous solution
Microanalysis GmbH, Berlin, Germany) equipped with a molybdenum
containing 317 mg ‧ L-1 U(VI) with adjusted pH = 1.8. The obtained
target X-ray tube of 50 W power, a multilayer monochromator, and a slurry was shaken for 24 h and after that, the U(VI) adsorption on the
silicon drift detector. The X-ray tube was operating at 50 kV and 1000 PS_C was determined (aU(VI)_PS_C [mg ‧ g− 1]), according to the Eq. (1).
μA. Uranium was determined in an air atmosphere, with a counting time Next, 5 mg of U-loaded PS_C was mixed with the 2 mL of a liquid me­
of 600 s and Ga as an internal standard. dium with the proper concentration and pH. The obtained slurry was
The pH measurements were carried out by the pH meter CP-401 shaken for 24 h, and then the solid phase was separated from the liquid
(Elmetron, Zabrze, Poland) equipped with a glass electrode. The acid- by centrifugation. The U(VI) concentration in the liquid medium was
base properties of the studied materials were estimated by the pH determined by the UV–Vis technique (arsenazo III method (Wang et al.,
measurement of an aqueous KCl solution (10-3 mol · L-1) being in contact 2012). The desorption degree of uranium was calculated according to
with the studied solid phase for 24 h (msolid phase/VKCl solution = 20 mg/5 the following relationship:
mL) (pHequil. measurements).
CU V liquid medium
Desorption = 100% (3)
mU PS C aU(VI) PS C
2.5. U(VI) adsorption studies
where Desorption is the desorption degree of uranium [%], CU is the
20 mg of the PS or PS_C were mixed with 5 mL of an aqueous solution uranium concentration in the liquid medium after desorption [mg ‧ L-1],
of U(VI) ions with specified pH value and U(VI) concentration. The mU_PS_C is the mass of the U-loaded PS_C [mg] and Vliquid_medium is the
obtained slurries were shaken by an Elpin Plus orbital shaker (Elpin Plus, volume of the liquid medium used for uranium desorption [mL].
Lubawa, Poland) at 150 rpm for 24 h (except for kinetics studies). After
equilibration, the liquid phase was separated from the solid phase by 2.7. U(VI) removal from model water samples
centrifugation. The U(VI) concentration in the aqueous solution before
and after the sorption process was determined spectrophotometrically Studies concerning the removal of U(VI) ions by PS_C material were
by a Cary 50 Bio (Varian, Australia) at 652 nm according to the arsenazo conducted using the model water samples: river water (SLRS-6, NRC,
III method (Wang et al., 2012). The adsorption quantity of U(VI) on the Canada) and seawater (CASS-6, NRC, Canada). Each sample was spiked
studied adsorbents (a) [mg · g− 1] was calculated with the following with UO2+2 so as to reach the final uranium concentration in the range of
equation: 0.05 mg ‧ L-1 – 1.0 mg ‧ L-1. The pH value of model water samples was
( ) adjusted to 7.0.
Cin − Ceq V
a= (1) Briefly, 5 mg of PS_C was mixed with 50 mL of model natural water.
m
Then the suspension was shaken for 24 h and dried in the oven at 120 ◦ C
where Cin and Ceq were the initial and final U(VI) concentration in for the next 24 h. The concentration of U in the liquid phase was
aqueous solution [mg · L-1], respectively; V was the volume of the determined by UV–vis and TXRF techniques. The TXRF technique was
aqueous U(VI) solution [mL] and m was the mass of the adsorbent [mg]. proposed because of the low UV–vis limit of detection (LOD) towards
The U(VI) adsorption on the studied sorbents was optimized by the uranium.
effect of pH, contact time, and U(VI) concentration. The effect of pH was UV–vis determination: 5 mg of U-loaded PS_C was mixed with the 1
studied in the pH range of 1–7 and for U(VI) concentration of 65 mg · L-1. mL of 1 mol ‧ L-1 NaHCO3 (pH = 12.6) for 24 h. The suspension was
The pH of an aqueous solution of U(VI) ions was adjusted by the addition centrifuged and 1 mL of liquid phase was added to the mixture, con­
of the proper amount of 1 mol · L-1 HNO3 or 1 mol · L-1 NaOH solution. taining 0.1 mL of 0.07 wt% arsenazo III, 0.1 mL of 1 mol ‧ L-1 HNO3 and
The effect of time on the U(VI) adsorption was studied in the time range 0.8 mL of distilled water. Measurements were conducted at 652 nm. The
of 1–240 min and for U(VI) concentration of 65 mg · L-1. The U(VI) calibration was performed with the U standard solutions with concen­
adsorption isotherms were determined in the U(VI) concentration range trations between 0.5 mg ‧ L-1 and 6.2 mg ‧ L-1.
of 7–536 mg · L-1 and at 25 ◦ C. All U(VI) adsorption studies were per­ TXRF determination: 1 mg of U-loaded PS_C material was placed in a
formed in static conditions. 2-mL Eppendorf vial. Then, 0.5 mL of 1 mg L− 1 Ga (as an internal

3
R. Olchowski et al. Environmental Nanotechnology, Monitoring & Management 20 (2023) 100788

standard) in 2 mol L− 1 HNO3 was added. In order to obtain a homo­ specific surface area (1 m2 ‧ g− 1) and a total pore volume close to 0 cm3 ‧
geneous suspension, the sample was placed in the ultrasonic bath for 10 g− 1, which confirms the nonporous character of this solid material. In
min and vortexed immediately before pipetting onto a siliconized quartz comparison, the carbonaceous material possesses a high specific surface
reflector. Finally, 10 μL of the suspension was transferred onto a area (437 m2 ‧ g− 1) and total pore volume equal to 0.29 cm3 ‧ g− 1.
reflector, dried, and directly analysed. Additionally, the BJH average pore diameter>2.0 nm confirms the
presence of mesopores in the PS_C material. The high specific surface
3. Results and discussion area of PS_C is probably related to the presence of micropores in its
structure. The increase of the specific surface area and total pore volume
3.1. Physicochemical studies after thermochemical treatment of PS material by H3PO4 confirms that
the orthophosphoric acid is the efficient pores forming agent (Puziy
Fig. 1 shows the chemical structures of the monomers and lignin used et al., 2020). The H3PO4 acts as the initiator of the bond breaking in the
for the copolymerization reaction. DEGDMA belongs to the popular carbon precursor (PS material), leading to the dehydration and elimi­
group of monomers based on a very reactive methacrylic group. Addi­ nation reactions that release volatile products (e.g., CO, CO2, H2O)
tionally, it possesses 3 ether and two carboxylic oxygens, which signif­ (Peng et al., 2017; El-Hendawy, 2006). These volatile compounds are
icantly affects lignin’s affinity for this monomer. The addition of lignin probably responsible for the creation of porous structure in PS_C
increases the functionality of microspheres and further increases the material.
attractiveness of these materials as potential precursors to the carbon­ Table 2 presents the elemental composition of PS and PS_C materials
ization process. This is due to the fact that lignin has numerous aromatic obtained from SEM-EDX and XPS studies. Differences in SEM-EDX and
(phenolic) groups in its structure, which grow its thermal resistance XPS data for a given material are related to the various sampling depth
compared to other biopolymers, such as cellulose or starch. (1 µm or 10 nm for SEM-EDX and XPS technique, respectively). Thus, the
In Fig. 2A, photos of PS microspheres are presented. The micro­ XPS data include only a few layers of atoms in contrast to the SEM-EDX
spheres possess a regular spherical shape and diameter in the range of data. According to these data, the PS microspheres consist of C
30–50 µm. In Fig. 2B, the material PS_C is presented. During the ther­ (73.1–86.3 wt%), O (13.5–24.7 wt%), and S (0.2–1.0 wt%). More O and
mochemical treatment of PS material by H3PO4 polymeric microspheres S atoms are present in deeper layers of PS microspheres than on the
were destroyed, and the powder particles of PS_C were created. It can be surface. All these atoms come from lignin and monomers used during the
seen that some of PS_C particles have diameters much smaller than 30 synthesis of PS material. After thermochemical treatment of PS material
µm. Microspheres show a slight tendency to swell, which is presented in by H3PO4, the overall C content decreases (by ca. 12 wt%), O content
Fig. S1. The swelling did not exceed 20 %. slightly decreases (by ca. 0.7 wt%), and S atoms are removed from the
The SEM microphotographs of the studied materials are presented in material. At the same time, the P atoms appeared (10.6–13.5 wt%).
Fig. 3. The spherical shape of PS particles with diameters between 30 µm Some surface O atoms are probably replaced by P atoms simultaneously,
and 50 µm is confirmed by SEM. In the case of PS_C material, it can be and the content of the surface O decreased by ca. 10 wt% and the content
seen that its particles have a cauliflower-like morphology (Rezaeipayam of surface P increased similarly. The thermochemical H3PO4 treatment
et al., 2017). It means that the high-temperature carbonization and of PS microspheres resulted in the removal of S from its structure and the
activation of PS by H3PO4 have a noticeable impact on the morphology incorporation of P on the newly created carbonaceous surface of PS_C.
of the obtained carbonaceous product (PS_C). The ATR-FTIR results are shown in Fig. 6. In the spectrum of PS, the
Fig. 4 and Fig. 5 present the nitrogen adsorption/desorption iso­ band around 3400 cm− 1 corresponding to νOH vibrations is visible.
therms and BJH pore size distribution patterns for PS and PS_C mate­ Moreover, νCH2, CH3, νC=O, νAr, νC-O and γAr, ArH bands are observed at
rials, respectively. The PS material is characterized by the isotherm of III 2940–2930 cm− 1, 1723 cm− 1, 1475 cm− 1, 1120 cm− 1, and 900–650
type according to IUPAC. This type of isotherm indicates a nonporous or cm− 1, respectively (Zielinski and Rajca, 2000). All observed bands come
macroporous solid. In contrast, the PS_C isotherm is of the IVa type, from the monomers and lignin used during the synthesis of PS micro­
which is typical for micro-mesoporous material (Thommes et al., 2015). spheres. After modification/carbonization of PS material new bands νPO-
− 1 − 1
These considerations are confirmed by the BJH pore size distributions. H (2700–2560 cm ), νO=C–O (2360 cm ), νP-OH, νP–O–C, νP=O (1200

− 1 − 1
PS material is nonporous, and PS_C material has mesopores in its cm ) and νP O P (980–900 cm ) appeared, corresponding to phos­
– –
structure, mainly with a width of 3.8 nm. There is also a large population phates, polyphosphates, and carbon dioxide (Zielinski and Rajca, 2000).
of mesopores with a width equal to 2.0 nm in the carbonaceous struc­ Furthermore, the thermochemical treatment of PS material by H3PO4
ture. In Table 1, the porosity data for both studied materials are sum­ resulted in the decrease of intensity of bands related to the aromatic
marized. The polymeric material is characterized by the negligible carbon rings, carbonyl, methyl, methylene, and C–O groups. These

H3C O
CH 2OH
H2C O
HO
lignin

+ O
+ HS

H3C O H3C O H (or lignin) ST-DEGDM-L


OH microspheres
H2C O
ST DEGDM Lignin (L)
Fig. 1. Chemical structures of monomers and kraft lignin used for polymerization.

4
R. Olchowski et al. Environmental Nanotechnology, Monitoring & Management 20 (2023) 100788

Fig. 2. Obtained PS microspheres before (A) and after (B) carbonization.

Fig. 3. SEM microphotographs of the PS (magnification 1 000x) and PS_C (magnification 25 000x) materials.

Fig. 5. BJH pore size distribution of studied materials calculated from the
Fig. 4. Nitrogen adsorption/desorption isotherms of PS and PS_C materials. desorption branch of nitrogen isotherms.

changes can be attributed to the graphene domains formation and re­ microspheres proceeds through three stages (three peaks in the DTG
actions between the surface carbon–oxygen groups with the H3PO4. The plot). The main decomposition takes place at 365 ◦ C. Additionally, two
latter probably was responsible for the porous structure development in small decomposition maxima’s are located at 190 ◦ C and 610 ◦ C. In the
the PS_C material (Peng et al., 2017; El-Hendawy, 2006). case of PS material, the main decomposition peak is connected with the
Fig. 7 illustrates the thermogravimetric (TG/DTG) curves for the thermal degradation of the polymeric microspheres. Two other peaks
obtained PS and PS_C materials. The thermal degradation process of PS are probably connected with the degradation of the C–O–C aliphatic

5
R. Olchowski et al. Environmental Nanotechnology, Monitoring & Management 20 (2023) 100788

Table 1 533.9 eV, 12–11 %). The P2p band for PS_C material consists of the three
Porous characteristics of PS and PS_C materials (SBET, VT and dBJH,des.). peaks located at 132.9 eV (P–C, 47.3 %), 133.8 eV (PO2- 3 , 48.3 %), and
Symbol SBET VT dBJH,des. 135.8 eV (H3PO4, 4.4 %) (Oh et al., 2014; Stobinski et al., 2014; Barinov
of material [m2 ‧ g¡1] [cm3 ‧ g¡1] [nm] et al., 2009; Wagstaffe et al., 2016). The deconvolution of the S2p band
PS 1* 0.01* – for PS material revealed the presence of the three main groups: C-SH
PS_C 437 ± 13# 0.29 ± 0.01# 3.8 ± 0.1# (163.4 eV, 53.9 %), C–S (165.5 eV, 33.2 %), and C-S– – O (167.5 eV,
*
- SD close to 0; # - SD from 3 replicates.

chains (190 ◦ C) and the partial collapse of the aromatic part of lignin
(610 ◦ C). In contrast, the H3PO4 decomposes at a temperature higher
than 300 ◦ C (Xu et al., 2021). Thus, during the modification-pyrolysis
process of PS material with H3PO4 the slow heating step in the tem­
perature range of 360–450 ◦ C was applied. Moreover, the applied py­
rolysis temperature of PS with the orthophosphoric acid (700 ◦ C) was
crucial for the efficient incorporation of the phosphorous atoms into the
emerging carbonaceous surface (Puziy et al., 2020); since the modifi­
cation temperature was closer to 800 ◦ C the higher P content in the
obtained carbon material. After the carbonization process, no significant
changes to the mass sample were observed up to 600 ◦ C; only about 10 %
mass loss of adsorbed water or solvents was noticeable. When the tem­
perature exceeded 600 ◦ C, significant mass changes were observed. It
means that the PS_C material is thermally stable up to 600 ◦ C, and at
higher temperatures, the decomposition of some functional groups can
take place (Puziy et al., 2020).
The pHequil. values of PS and PS_C materials were 5.47 and 3.11,
respectively. The weak acidic character of the PS surface can be related
to the phenolic and thiol groups present on the PS surface. The presence
of the carboxyl and orthophosphoric groups, which can be easily
dissociated in an aqueous medium, is probably responsible for the strong
acidic character of the PS_C surface.
The deconvoluted high-resolution XPS spectra of C1s, O1s, S2p, and
P2p core energy levels for PS, and PS_C materials are presented in
Fig. S2-S5 (Supplementary Data). Moreover, the contributions of each
Fig. 6. ATR-FTIR spectra of PS (bottom) and PS_C (top) materials.
deconvoluted component in the overall C1s, O1s, or P2p XPS band for
both studied materials are depicted in Fig. S6-S8. In the case of the PS
material, there are three main components of C1s band: C– – C sp2 (284.6
eV, 26.7 %), C–C sp3 (285.3 eV, 29.3 %), and C–O (286.5 eV, 26.3 %).
There are also components related to the C– – O (287.9 eV, 0.4 %), O-
C– O (288.8 eV, 3.9 %), and carbon vacancies (284.1 eV, 13.4 %)

(Rabchinskii et al., 2020). These data are in agreement with the struc­
ture of chemical compounds used for PS microsphere synthesis. After
carbonization-modification, the increase of C– – C sp2 intensity (up to
52.5 %) and simultaneously the decrease of C–O and O-C– – O intensities
(up to 7.1 % and 1.0 %, respectively) were observed. The modification
had a negligible impact on the intensity of C–C sp3 (26.3 %), C– – O (1.0
%), and carbon vacancies (11.4 %). One additional peak located at
290.1 eV (CO2– 3 , 0.7 %) has appeared. It means that the modification of
PS microspheres by H3PO4 at high temperature resulted in the formation
of the aromatic benzene rings, carbonate groups, and decomposition of
C–O groups (phenolic, alcoholic, ether). These considerations are
confirmed by O1s and P2p deconvoluted bands. During modification
process the content of O–C groups (533.2 eV) decreased (from 53.5 %
to 33.6 %), the content of O– – P groups (530.9 eV) increased (from 0 % to
18.6 %), the water (535.5 eV) content increased (from 0 % to 1.9 %) and
the content of the other groups did not change significantly (O– – C:
Fig. 7. Thermogravimetry analysis of studied materials.
530.0 eV, 0.3–0.7 %; O– – C-O: 532.0 eV, 34.2–34.3 %; O– – C-O, O–P:

Table 2
Elemental composition of PS and PS_C obtained from SEM-EDX and XPS studies.
Symbol SEM-EDX XPS$
of material C O P S C O P S
[wt. %] [wt. %] [wt. %] [wt. %] [wt. %] [wt. %] [wt. %] [wt. %]

PS 86.3 ± 0.5# 13.5 ± 0.5# 0.0* 0.20 ± 0.02# 73.1 ± 2.0# 24.7 ± 1.4# 0.0* 1.0 ± 0.6#
PS_C 73.7 ± 0.3# 12.8 ± 0.5# 13.5 ± 0.4# 0.0* 74.2 ± 2.0# 13.8 ± 1.5# 10.6 ± 0.8# 0.0*
*
- SD close to 0; # - SD from 3 replicates, $
- for both materials Si and N content < 1.4 wt%.

6
R. Olchowski et al. Environmental Nanotechnology, Monitoring & Management 20 (2023) 100788

12.9 %) (https://srdata.nist.gov/xps/EngElmSrchQuery.aspx­ groups on the surface of PS_C than on PS. The fast equilibration in the
EType=PECSOpt=Retri_ex_dat Elm=S, 2022). During the modification case of PS_C is beneficial for the efficient U(VI) removal from the model
of PS microspheres with H3PO4 and probably its thermal carbonization, water samples.
the graphene domains are formed, sulfur is volatilized and removed The experimental U(VI) adsorption kinetics data for PS_C were
from the material. The O–C groups are decomposed in favor of the matched to the three theoretical kinetics models: pseudo-first-order (Eq.
attachment of phosphorous groups to the surface of the forming carbon (4)), pseudo-second-order (Eq. (5)), and Elovich (Eq. (6)) (Olchowski
by linking with the oxygen atoms. Some water may be formed as a result et al., 2021; Godlewska et al., 2020). These models are described by the
of the reaction between the H3PO4 and oxygen functional groups of following linear equations:
forming carbonaceous material. ( )
ln qeq − qt = lnqeq − k1 t (4)

1 1 1
3.2. U(VI) adsorption tests = + (5)
qt qeq q2eq k2 t
The effect of pH on the U(VI) adsorption from aqueous solution onto
1 1
the PS and PS_C materials was studied. In Fig. 8A, the U(VI) adsorption qt = ln(αβ) + lnt (6)
as a function of the final pH for both studied materials was presented. In β β
the case of the PS material U(VI) adsorption values are negligible and
where qt is the U(VI) adsorption at time t [mg ‧ g− 1], qeq is the maximum
there is no dependency between the U(VI) adsorption and the final pH of
adsorption of U(VI) in the studied system [mg ‧ g− 1], t is the time of
an aqueous solution. Conversely, for PS_C it can be seen that the
adsorption process [min], k1 is the kinetics rate constant for pseudo-
maximum U(VI) adsorption was achieved for the pHfin. = 1.8. Taking
first-order-model [min− 1], k2 is the kinetics rate constant for pseudo-
into account the presence of the U(VI) species in the studied pH range
second-order model [g ‧ mg− 1 ‧ min− 1], α is the initial adsorption rate
(Gładysz-Płaska et al., 2018)it may be expected, that UO2+ 2 ions can be
[mg ‧ g− 1 ‧ min− 1], β is the constant related to the chemisorption acti­
complexed mainly by the surface orthophosphate groups on the PS_C
vation energy and the extent of surface coverage [g ‧ mg− 1]. In Table 3
carbon material. The course of the presented curve can be related to the
there are collected data obtained from matching the experimental ki­
simultaneous decrease of UO2+ 2 fractionation and increase of the com­
netics data with the theoretical one for the studied adsorption system.
plexed U(VI) form contribution. Fig. 8B shows the relationship between
The U(VI) adsorption kinetics onto PS_C is well described by both the
pHfin. and pHin. for both tested materials during U(VI) adsorption. For
pseudo-first-order (R2 = 0.9957) and pseudo-second-order (R2 =
the PS material, there was no change in pH associated with either
0.9949) kinetics model. Moreover, for both models the qeq, theoret. (15.9
dissociation of surface functional groups or the adsorption of U(VI) ions.
mg ‧ g− 1) was close to the qeq, exp. (16.0 mg ‧ g− 1). It means that the
For PS_C, there was a decrease in pHfin. to a value of about 2.5 for pHin.
physisorption and chemisorption processes have a similar impact on the
> 2. It may be the result of the dissociation of carboxyl and ortho­
U(VI) adsorption kinetics in the studied adsorption system. Further­
phosphate groups present on its surface and following cationic exchange
more, the large α constant obtained from the Elovich model can be
between U(VI) cations (Gładysz-Płaska et al., 2018)and H3O+ ions. Thus
related to the very fast dynamic of the adsorption kinetic process, which
the second probable mechanism of U(VI) adsorption onto the PS_C
confirms the fast equilibrium in the studied system.
material may be related to the cation exchange.
Also, the U(VI) adsorption isotherms for PS and PS_C materials were
In Fig. 9, the adsorption kinetics curves of U(VI) ions onto the PS
obtained (Fig. 10). In the case of the PS microspheres, the maximum
(Fig. 9A) and PS_C (Fig. 9 B) materials are presented. The U(VI)
static adsorption capacity of U(VI) ions was estimated as 0.095 mg ‧ g− 1.
adsorption equilibrium state was established after 5 min. or 2 min. for PS
In contrast, the value obtained for PS_C is ca. 150 mg ‧ g− 1. Moreover, for
and PS_C materials, respectively. In the case of PS material, the
carbonaceous material, the first part of the adsorption isotherm was
adsorption process consisted of two steps: a sharp increase of U(VI)
characterized by the paraxial mileage up to high U(VI) adsorption values
adsorption in ca. 1 min. and a slower one. The first step can be caused by
(ca. 100 mg ‧ g− 1), which means that in the system U(VI) – PS_C the
the fast diffusion rate of U(VI) ions to the surface of PS microspheres.
adsorption equilibrium is strongly shifted towards the carbonaceous
However, the second step was probably related to the reaction of U(VI)
surface. Thus the PS_C material can be successfully used for the efficient
ions mainly with the surface oxygen-rich groups. The adsorption process
removal of U(VI) ions from the natural water.
of U(VI) ions onto PS_C was one-step, probably due to the presence of the
The U(VI) adsorption isotherm data for PS_C were compared to the
mesopores in the structure of PS_C material. The faster establishment of
Langmuir (Eq. (7)) and Freundlich (Eq. (8)) theoretical adsorption
the adsorption equilibrium state for PS_C than for PS microspheres can
isotherm models (Olchowski et al., 2021). These two models are
be explained by the higher content of the negatively charged functional

Fig. 8. The U(VI) adsorption as a function of pHfin. (A) and pHfin. vs pHin. (B) for studied materials (adsorption system: mmaterial = 20 mg, Vsolution = 5 mL, C0U(VI) =
61 mg L-1, teq = 24 h).

7
R. Olchowski et al. Environmental Nanotechnology, Monitoring & Management 20 (2023) 100788

Fig. 9. The U(VI) adsorption kinetics for PS (A) and PS_C (B) (adsorption system: mmaterial = 20 mg, Vsolution = 5 mL, C0U(VI) = 54 mg L-1, pHfin. = 1.8).

Table 3
Fitting results of kinetics and isotherm experimental data for PS_C material with theoretical kinetics (pseudo-first-order, pseudo-second-order, Elovich) and isotherm
(Langmuir, Freundlich) models in studied adsorption systems.
Adsorption kinetics model Pseudo first order Pseudo second order Elovich
2 2
Parameters qeq,theoret. k1 R qeq,theoret. k2 R α β R2
[mg g− 1] [min− 1 105] [mg g− 1] [g mg− 1 min− 1] [mg g− 1 min− 1
1044] [g mg− 1]
15.9 ± 0.1# 6.43* 0.9957 15.9 ± 0.2# 10.0 ± 0.4# 0.9949 2.50 ± 0.12# 6.73 ± 0.13# 0.9916
* #
- SD close to 0; - SD from 3 replicates.

0 (0.18), which means that the surface of PS_C material was energeti­
cally heterogeneous. Additionally, the RL (separation factors constant)
values (Eq. (9)) were calculated to estimate the degree of suitability of
PS_C towards U(VI) ions, according to the following equation (Dobr­
zyńska et al., 2021):
1
RL = (9)
1 + kL cin
The RL values in the studied system U(VI) – PS_C were in the range
between 0 and 1 (Table 3), which means that the PS_C is suitable for the
removal of U(VI) ions from the naturally occurring water.

3.3. Competitive ions effect

Natural water contains many various ions. Some of them are present
at the mg ‧ L-1 concentration level, such as HCO–3, Cl-, NO–3, SO2-
4 , Mg
2+
or
Fig. 10. The U(VI) adsorption isotherms for studied materials (adsorption Ca2+. In turn, the Fe3+ ions are present in river water and seawater at
system: mmaterial = 20 mg, Vsolution = 5 mL, pHfin. = 1.8, teq = 24 h, T = 20.0 concentrations below 80 µg ‧ L-1 and 0.8 µg ‧ L-1, respectively, because of
± 0.4 ◦ C). the pH close to 7 or even higher. In these conditions, iron is mainly
present as precipitated iron(III) hydroxides (Fox, 1988; Kim et al., 2013;
described by the following formulas: Su et al., 2019). All these ions can be competitive for the adsorption of U
1 1 1 (VI) ions on the surface active centers of PS_C material. Thus, the study
= + (7) of their impact on the U(VI) adsorption onto PS_C material is a crucial
qeq qm Ceq qm kL
task because it affects the effectiveness of U(VI) removal from waste­
( ) (
ln qeq = ln(kF ) + nF ln Ceq
)
(8) water. In Fig. 11, the effect of competitive ions on the U(VI) adsorption
onto PS_C was presented. It can be seen that some ions (HCO–3, NO–3, Cl,
where qm is the U(VI) adsorption capacity also known as monolayer Mg2+, and Ca2+) have a negligible impact on the U(VI) adsorption onto
capacity [mg ‧ g− 1], Ceq is the U(VI) equilibrium concentration [mg ‧ L-1], the studied material, even at their highest studied concentrations. A
kL is the Langmuir equilibrium constant [L ‧ mg− 1], kF is the Freundlich different situation was observed for SO2- 4 and Fe
3+
ions. In the first case,
equilibrium constant [mg1-nF ‧ LnF ‧ g− 1] and nF is the Freundlich con­ the U(VI) adsorption onto PS_C decreases by about 20 % for a sulfate
stant. The results of matching the experimental adsorption isotherm concentration equal to 100 mmol ‧ L-1 and may further decrease by as
data with the theoretical one are presented in Table 3. The best matching much as 70 % for a higher sulfate concentration. The studied system is
was shown for the Langmuir model (R2 = 0.9154), for which the very sensitive to the presence of Fe3+ ions. The Fe3+ concentration of
maximum static U(VI) adsorption capacity in the studied system was 0.2 mmol ‧ L-1 (5 mg ‧ L-1) can dramatically deteriorate the adsorption
149 mg ‧ g− 1. The good matching of U(VI) adsorption isotherm data onto properties of PS_C towards U(VI) ions (up to 80 %). It was simulta­
PS_C with the Langmuir model can be related to the chemisorption as the neously observed that the Fe3+ ions have a negligible impact on the PS_C
main U(VI) adsorption mechanism on the surface of PS_C material. adsorption capacity toward U(VI) ions up to 0.1 mmol ‧ L-1. The strong
Moreover, the nF value obtained for the Freundlich model was close to competitiveness of sulfate anions and iron cations to uranyl cations can
be the result of their favorable binding to the uranyl cations or their

8
R. Olchowski et al. Environmental Nanotechnology, Monitoring & Management 20 (2023) 100788

Fig. 11. The effect of competitive ions on the U(VI) adsorption onto PS_C (adsorption system: mmaterial = 20 mg, Vsolution = 5 mL, pHfin. = 1.8, teq = 24 h, C0U(VI) =
5.5 mg L-1).

higher affinity than uranyl cations towards orthophosphoric groups conducted. In Fig. 13, the deconvoluted U4f7/2 XPS spectrum for U-
present on the PS_C surface, respectively. Natural water may contain loaded PS_C was presented. Several signals are located at 381.3 eV (U
sulfate at concentrations below 100 mmol ‧ L-1 and trace concentrations (IV)), 382.4 eV (U(VI)A), 383.4 eV (U(VI)B), 384.5 eV (U(VI)C), 385.4
of iron. Thus, the PS_C can effectively remove the U(VI) ions from nat­ eV (U(VI)Asat.), and 386.6 eV (U(VI)Bsat.) (Ilton et al., 2007; Kowal-
ural water. Fouchard et al., 2004). The most intense signal (62.6 %) is related to
the A phase of U(VI), the second one (19.2 %) is connected with the B
3.4. Regeneration studies phase of U(VI), and the third one (10.2 %) is the C phase of U(VI). There
is also the signal from U(IV) (8.0 %), which can be related to the
The regeneration of the adsorbent allows for its use several times, reduction of U(VI) by the X-ray beam during the XPS measurement (Ilton
which provides the cost reduction related to the adsorbent usage and it et al., 2007). Additionally, there are satellite signals for U(VI)A and U
has a positive impact on the environment. The PS_C material loaded with (VI)B (Ilton et al., 2007). The overall content of U(VI) on the surface of
a known amount of uranium was tested for its regeneration by various PS_C after U(VI) adsorption is ca. 92 %. The comparison of the signal
liquid media: HCl, HNO3, and NaHCO3 with concentrations between intensities for C1s, O1s, and P2p deconvoluted spectra between pristine
0.01 and 1.00 mol ‧ L-1. In Fig. 12A, the obtained regeneration results are PS_C and U-loaded PS_C can be helpful to further investigation of the U
presented. The highest uranium desorption degree from U-loaded PS_C (VI) adsorption mechanism. During U(VI) adsorption onto PS_C, the
intensity of the C– – C sp2 signal decreases (52.5–49.4 %) without sig­
material was obtained for 1 mol ‧ L-1 NaHCO3 (ca. 30 %). This value was
not satisfying; thus, the pH effect on the uranium desorption by 1 mol ‧ L- nificant intensity changes for other C1s signals (Fig. S6). It can be related
1
NaHCO3 was also tested (Fig. 12B). It can be seen that the higher pH of to the formation of the complex between uranyl cation and π-π electrons
1 mol ‧ L-1 NaHCO3 the more effective uranium desorption (up to ca. 85 from graphene plains. Moreover, the O–C signal intensity decreases
% at pH = 12.6). This effect has already been described in the literature (33.6–28.3 %), and simultaneously an O2– signal appears together with
(Gorman-Lewis et al., 2008). The HCO–3 anions in alkaline conditions can the increase of the O– – P signal intensity (52.8–56.7 %) (Fig. S7). These

be strongly competitive with orthophosphate functional groups present changes in the O1s signal intensities can be the result of the uranyl
on the PS_C, and the stable bicarbonate complexes with uranium can be cation binding to the surface C-OH and C–O–C groups of PS_C. In turn,
formed. the P–C signal intensity increases (47.3–57.5 %), and PO2- 3 signal in­
tensity decreases (48.3–37.1 %), which can be related to the U(VI)
complexation by the surface PO2- 3 groups. Also, the binding energies for
3.5. U(VI) adsorption mechanism the above mentioned XPS signals changed after U(VI) adsorption onto
PS_C carbon: from 284.6 eV to 284.5 eV (C = Csp2), from 533.1 eV to
In order to investigate the U(VI) adsorption mechanism onto the 533.0 eV (O–C), from 532.1 eV to 532.0 eV (O– – P) and from 133.8 eV
surface of PS_C, the XPS studies for U-loaded PS_C material were

Fig. 12. The desorption results of U from U-loaded PS_C for three different liquid media (A) and pH effect of 1 mol ‧ L-1 NaHCO3 on the U desorption degree from U-
loaded PS_C (desorption system: mU-loaded PS_C = 5 mg, Vliquid medium = 2 mL, teq = 24 h) (B).

9
R. Olchowski et al. Environmental Nanotechnology, Monitoring & Management 20 (2023) 100788

deposited on the reflector is 20 µg, which undoubtedly allows meeting


the conditions described above. The method of sample preparation for
TXRF analysis was extremely simple, as it consisted only of dispersing
and depositing U-loaded PS_C and an internal standard on the TXRF
reflector.
Gallium was used as an internal standard to correct the non-
reproducibility of the sample position on the TXRF reflector and the
inhomogeneous excitation caused by the standing-wave field over the
reflector surface. The IU/IGa ratio does not depend on the position of the
residue on the reflector and the measurement conditions. The precision
of the method was studied by analyzing five U-loaded PS_C samples as
suspensions pipetted at 20 µg onto the reflectors. The relative standard
deviation (RSD) characterizing the precision of pipetting the suspension
onto quartz reflectors is only 2.0 %. This value should be considered
very good for ultra-trace analysis. An internal standard can also be used
in TXRF spectrometry to determine the content of an analyte in the
analyzed material. In this case, performing a calibration with a series of
reference samples is not necessary. The internal standard method was
used to determine uranium. The calculation of uranium content in U-
loaded PS_C was based on the net intensity of the analyte and the net
intensity of the internal standard, as well as the instrumental sensitiv­
ities for these elements (Kregsamer et al., 2002). Previous studies have
shown that using a low-power TXRF instrument to determine uranium
resulted in an extremely low LOD at the ppt level (Kocot et al., 2022).
Thus, TXRF allows the determination of uranium at concentrations
much lower (ca. 2000 times) than the maximum levels of uranium
contamination of environmental samples allowed by EPA regulations
(30 µg/L) (Uranium - Water Quality Association; https://wqa.org/learn-
about-water/common contaminants/uranium, 2022).
The results of this quantification method are given in Table 4.

3.7. U(VI) removal from the natural water


Fig. 13. The deconvoluted high-resolution U4f7/2 XPS spectrum for U-
loaded PS_C. The results concerning the U(VI) removal from the model seawater
and river water are presented in Table 4. Due to the limited LOD of the
to 133.6 eV (PO2- UV–vis technique for U (2.16 µg ‧ L-1), concentrations of this analyte in
3 ). These binding energy changes during U(VI)
adsorption are indicating the formation of the coordination bonds be­ the non-spiked water samples were determined only by the TXRF
tween surface functional groups of PS_C carbon and uranyl cations technique (LOD = 0.011 µg ‧ L-1). The obtained values were in good
(Wang et al., 2022; Hu et al., 2021; Liu et al., 2022). Considering the XPS agreement with certified values. Moreover, the precision estimated for
investigations and our previous studies, it can be concluded that the the TXRF technique was greater than for the UV–vis technique. Ac­
main mechanism of U(VI) adsorption onto PS_C is the surface cording to these data it can be stated that the U(VI) ions were success­
complexation by PO2- fully removed from the studied model water samples. Thus, the PS_C
3 , C-OH, C–O–C groups, and π-π electrons from
the graphene domains. Additionally, according to the pH changes material can be successfully used both for U(VI) removal from natural
observed during U(VI) adsorption onto PS_C carbon, the ion exchange water and for uranium enrichment prior to its determination by the
between uranyl cations and protons from surface groups of PS_C also can
take place. Table 4
Results of U(VI) removal from natural water by PS_C (m = 5 mg (UV–vis) or 1 mg
(TXRF), V = 50 mL, pHwater = 7.0, teq = 2.5 min, T = (20.0 ± 0.4)oC).
3.6. U-loaded PS_C in ultrasensitive determination of U(VI) using TXRF Sample CU(VI) CU(VI)_determ. Certified U
_added [µg ‧ L-1] concentration
-
The composition and form of PS_C also make it attractive from the [µg ‧ L UV–vis TXRF [µg ‧ L-1]
1
]
point of TXRF analysis. Carbon and oxygen - the main components of
this material emit X-rays that are not detected by TXRF instruments. In River water 0 <LODUV–VIS 0.070 ± 0.070 ± 0.003*
contrast, the X-ray characteristic radiation of phosphorus present in the (SLRS-6) 0.005#
50 49 ± 4# 51 ± 1#
material is very low-energy, which in practice does not disturb the direct 100 103 ± 5# 99 ± 2#
determination of U(VI) ions adsorbed on PS_C. Another advantage of U- 500 507 ± 13# 498 ± 7#
loaded PS_C is the possibility of depositing smooth and thin layer sam­ 1000 997 ± 21# 1002 ±
ples on quartz reflectors, which is especially important for measure­ 10#
Seawater 0 3.05 ± 2.92 ± 0.42*
ments carried out under total reflection conditions. The mass of U- <LOQUV–VIS
(CASS-6) 0.53#
loaded PS_C should be as small as possible to achieve minimal scattering 50 52 ± 5# 53 ± 2#
of the primary X-ray radiation and low background in TXRF measure­ 100 98 ± 4# 104 ± 1#
ments. The ability to deposit small masses of adsorbent onto the reflector 500 504 ± 17# 501 ± 9#
for uranium determination is due primarily to its excellent adsorption 1000 993 ± 15# 1004 ± 4#

properties. Thus, the adsorbent amount is essential for U(VI) ion #


SD from 5 replicates, LODUV–VIS = 2.16 µg ‧ L-1, LODTXRF = 0.011 µg ‧ L-1, * -
adsorption and TXRF measurement. The mass of U-loaded PS_C finally expanded measurement uncertainty for k = 2.

10
R. Olchowski et al. Environmental Nanotechnology, Monitoring & Management 20 (2023) 100788

TXRF technique. The separation of the analyte from the rich/complex Data curation, Investigation, Resources, Visualization. B. Zawisza:
matrix (e.g., high salt concentrations) is important for the TXRF mea­ Conceptualization, Data curation, Investigation, Methodology, Re­
surement. Direct analysis of water with a complex matrix by the TXRF sources, Supervision, Validation, Writing – review & editing. K. Morlo:
can be possible, but only after multiple dilutions prior to deposition on a Conceptualization, Visualization, Writing – review & editing. R.
quartz reflector to form a thin film on it. Nevertheless, even then, the Dobrowolski: Conceptualization, Supervision, Validation, Writing –
intensity of the analyte can be significantly underestimated compared to review & editing.
the radiation intensity of the analyte after matrix separation caused by
the significant absorption of the analyte characteristic radiation by the Declaration of Competing Interest
dried residue (e.g., sodium and magnesium chlorides) (Sitko et al.,
2021). Thus, the use of PS_C is also beneficial in avoiding this problem. The authors declare that they have no known competing financial
Moreover, due to the low detection limits for TXRF, in the case of higher interests or personal relationships that could have appeared to influence
analyte addition, the volume of water may be significantly smaller, e.g., the work reported in this paper.
5 mL.
Data availability
4. Conclusion
The raw/processed data required to reproduce these findings cannot
We significantly improved U(VI) removal performance by devel­ be shared at this time due to technical or time limitations.
oping PS_C material using the high-temperature H3PO4 activation-
modification. The application of H3PO4 not only resulted in the intro­ Acknowledgements
duction of the micropores into the PS_C structure, but it also had an
influence on the particle morphology (from spherical to cauliflower- The authors are grateful to Professor Dr. Joseph H. Rule from Old
like) and the surface chemistry (introduction of the phosphorous-rich Dominion University (Norfolk. VA) USA and his wife Dr. Anna M. Rule
groups) of the obtained material. Also, its thermal stability was for their help with the English preparation of the manuscript.
improved during modification. Generally, the advantages of the devel­
oped PS_C material compared to the initial polymer microspheres are as Appendix A. Supplementary data
follows: (1) The carbonaceous meso-microporous material possesses a
much higher specific surface area (437 m2 ‧ g− 1) than the nonporous Supplementary data to this article can be found online at https://doi.
polymeric material with a negligible specific surface area (1 m2 ‧ g− 1); org/10.1016/j.enmm.2023.100788.
(2) Adsorption equilibrium in the case of PS_C resulting from the higher
content of the negatively charged functional groups on the PS_C surface References
is 2.5 times faster than in the case of PS material; (3) The maximum
adsorption capacity of PS_C towards U(VI) is ca. 1500 times higher than Barinov, A., Malcioǧlu, O.B., Fabris, S., Sun, T., Gregoratti, L., Dalmiglio, M.,
Kiskinova, M., 2009. Initial stages of oxidation on graphitic surfaces: photoemission
of PS; (4) The adsorption equilibrium is strongly shifted towards the study and density functional theory calculations. J. Phys. Chem. C 113, 9009–9013.
carbonaceous surface in the system U(VI) – PS_C. https://doi.org/10.1021/jp902051d.
The PS_C material shows a strong affinity for U(VI) ions. The primary Borzęcki, R., Wójcik, D., Kalisz, M., 2018. Processing of uranium ores in Kowary area.
Hereditas Minariorum 5, 181–213. https://doi.org/10.5277/hm180508.
mechanism of U(VI) adsorption onto PS_C is the complexation by PO2- 3, Caccin, M., Giacobbo, F., Da Ros, M., Besozzi, L., Mariani, M., 2013. Adsorption of
C-OH, C–O–C groups, and π-π electrons from the graphene domains; uranium, cesium and strontium onto coconut shell activated carbon. J. Radioanal.
the ion exchange between uranyl ions and protons from surface groups Nucl. Chem. 297, 9–18. https://doi.org/10.1007/s10967-012-2305-x.
Craft, E.S., Abu-Qare, A.W., Flaherty, M.M., Garofolo, M.C., Rincavage, H.L., Abou-
of PS_C is also possible. XPS analysis, adsorption isotherms, and kinetic Donia, M.B., 2004. Depleted and natural uranium: chemistry and toxicological
data indicate U(VI) chemisorption. The studied system is robust to the effects. J. Toxic. Environ. Health, Part B 7, 297–317. https://doi.org/10.1080/
presence of HCO–3, Cl-, NO3, SO2- 2+ 2+
4 , Mg , Ca , and Fe
3+
at the concen­ 10937400490452714.
Dauer, L.T., Zanzonico, P., Tuttle, R.M., Quinn, D.M., Strauss, H.W., 2011. The Japanese
tration level found in natural waters. Thus, the PS_C material is bene­
tsunami and resulting nuclear emergency at the Fukushima Daiichi power facility:
ficial for efficient U(VI) removal from wastewater. technical, radiologic and response perspectives. J. Nucl. Med. 52, 1423–1432.
The unique adsorption properties of the PS_C material allow the use https://doi.org/10.2967/jnumed.111.091413.
of microgram masses of this sorbent for ultra-sensitive determination of Dobrzyńska, J., Dąbrowska, M., Olchowski, R., Zięba, E., Dobrowolski, R., 2021.
Development of a method for removal of platinum from hospital wastewater by
adsorbed uranyl ions by TXRF using a convenient and simplified novel ion-imprinted mesoporous organosilica. J. Environ. Chem. Eng. 9 (4), 105302.
approach based solely on an internal standard. The sample preparation Dutta, R.K., Shaida, M.A., Singla, K., Das, D., 2019. Highly efficient adsorptive removal
method for TXRF analysis does not require the isolation of uranium ions of uranyl ions by a novel graphene oxide reduced by adenosine 5’-monophosphate
(RGO-AMP). J. Mater. Chem. A 7, 664–678. https://doi.org/10.1039/c8ta09746a.
from the sorbent, which is a huge advantage and complies with green El-Hendawy, A.-N.-A., 2006. Variation in the FTIR spectra of a biomass under
analytical principles. impregnation, carbonization and oxidation conditions. J. Anal. Appl. Pyrolysis 75,
159–166. https://doi.org/10.1016/j.jaap.2005.05.004.
Fasfous, I.I., Dawoud, J.N., 2012. Uranium (VI) sorption by multiwalled carbon
5. Data availability nanotubes from aqueous solution. Appl. Surf. Sci. 259, 433–440. https://doi.org/
10.1016/j.apsusc.2012.07.062.
The raw/processed data required to reproduce these findings cannot Fox, L.E., 1988. The solubility of colloidal ferric hydroxide and its relevance to iron
concentrations in river water. Geochim. Cosmochim. Acta 52, 771–777. https://doi.
be shared at this time due to technical or time limitations. org/10.1016/0016-7037(88)90337-7.
Garboś S., Święcicka D., Bratkowski J., The assessment of exposure to uranium taken
Funding with water intended for human consumption in Poland, National Institute of The
Public Health – State Hygiene Facility, Warsaw, 2014, ISBN 97 883 893 791 91.
Gładysz-Płaska, A., Grabias, E., Majdan, M., 2018. Simultaneous adsorption of uranium
This research did not receive any specific grant from funding (VI) and phosphate on red clay. Prog. Nucl. Energy 104, 150–159. https://doi.org/
agencies in the public, commercial, or not-for-profit sectors. 10.1016/j.pnucene.2017.09.010.
Godlewska, P., Bogusz, A., Dobrzyńska, J., Dobrowolski, R., Oleszczuk, P., 2020.
Engineered biochar modified with iron as a new adsorbent for treatment of water
CRediT authorship contribution statement contaminated by selenium. J. Saudi Chem. Soc. 24, 824–834. https://doi.org/
10.1016/j.jscs.2020.07.006.
R. Olchowski: Conceptualization, Data curation, Formal analysis, Goliszek, M., Podkościelna, B., Fila, K., Riazanova, A.V., Aminzadeh, S.,
Sevastyanova, O., Gun’ko, V.M., 2018. Gun’ko V., Synthesis and structure
Investigation, Methodology, Resources, Visualization, Writing – original characterization of polymeric nanoporous microspheres with lignin. Cellulose 25
draft, Writing – review & editing. B. Podkościelna: Conceptualization, (10), 5843–5862.

11
R. Olchowski et al. Environmental Nanotechnology, Monitoring & Management 20 (2023) 100788

Gorman-Lewis, D., Burns, P.C., Fein, J.B., 2008. Review of uranyl mineral solubility Rabchinskii, M.K., Ryzhkov, S.A., Kirilenko, D.A., Ulin, N.V., Baidakova, M.V.,
measurements. J. Chem. Thermodynamics 40, 335–352. https://doi.org/10.1016/j. Shnitov, V.V., Pavlov, S.I., Chumakov, R.G., Stolyarova, D.Y., Besedina, N.A.,
jct.2007.12.004. Shvidchenko, A.V., Potorochin, D.V., Roth, F., Smirnov, D.A., Gudkov, M.V.,
Han, B., Zhang, E., Cheng, G., Zhang, L., Wang, D., Wang, X., 2018. Hydrothermal carbon Brzhezinskaya, M., Lebedev, O.I., Melnikov, V.P., Brunkov, P.N., 2020. From
superstructures enriched with carboxyl groups for highly efficient uranium removal. graphene oxide towards aminated graphene: facile synthesis, its structure and
Chem. Eng. J. 338, 734–744. https://doi.org/10.1016/j.cej.2018.01.089. electronic properties. Sci. Rep. 10, 6902. https://doi.org/10.1038/s41598-020-
https://srdata.nist.gov/xps/EngElmSrchQuery.aspx?EType=PE&CSOpt=Retri_ex_dat& 63935-3.
Elm=S, (Accessed 01 August 2022). Rezaeipayam, M., Javanbakht, M., Ghafarian-Zahmatkesh, H., Golestani, E., Ghaemi, M.,
Hu, B., Wang, H., Liu, R., Qiu, M., 2021. Highly efficient U(VI) capture by amidoxime/ 2017. In-situ synthesis of LiFePO4/carbon cauliflower-like by hydrothermal reaction
carbon nitride composites: Evidence of EXAFS and modeling. Chemosphere 274, for using in lithium ion batteries. Anal. Bioanal. Electrochem. 9, 630–639.
129743. https://doi.org/10.1016/j.chemosphere.2021.129743. Shah F., Soylak M., Kazi T. G., Afridi H. I., Development of an extractive
Huang, C., Puziy, A.M., Poddubnaya, O.I., Hulicova-Jurcakova, D., Sobiesiak, M., spectrophotometric method for uranium using MWCNTs as solid phase and arsenazo
Gawdzik, B., 2018. Phosphorus, nitrogen and oxygen co-doped polymer-based core- (III) as chromophore, J. Radioanal. Nucl. Chem., 296 (2013), 1239–1245, doi:
shell carbon sphere for high-performance hybrid supercapacitors. Electrochim. Acta 10.1007/s10967-012-2376-8.
270, 339–351. https://doi.org/10.1016/j.electacta.2018.02.115. Sitko, R., Musielak, M., Serda, M., Talik, E., Gagor, A., Zawisza, B., Malecka, M., 2021.
Ilton, E.S., Boily, J.-F., Bagus, P.S., 2007. Beam induced reduction of U(VI) during X-ray Graphene oxide decorated with fullerenol nanoparticles for highly efficient removal
photoelectron spectroscopy: The utility of the U4f satellite structure for identifying of Pb(II) ions and ultrasensitive detection by total-reflection X-ray fluorescence
uranium oxidation states in mixed valence uranium oxides. Surf. Sci. 601, 908–916. spectrometry. Sep. Purif. Technol. 277, 119450–119460. https://doi.org/10.1016/j.
https://doi.org/10.1016/j.susc.2006.11.067. seppur.2021.119450.
Kim, J., Tsouris, C., Mayes, R.T., Oyola, Y., Saito, T., Janke, C.J., Dai, S., Schneider, E., Soylak, M., Doan, M., 1996. Column preconcentration of trace amounts of copper on
Sachde, D., 2013. Recovery of uranium from seawater: a review of current status and activated carbon from natural water samples. Anal. Letters 29, 635–643. https://doi.
future research needs. Sep. Sci. Technol. 48, 367–387. https://doi.org/10.1080/ org/10.1080/00032719608000426.
01496395.2012.712599. Stobinski, L., Lesiak, B., Malolepszy, A., Mazurkiewicz, M., Mierzwa, B., Zemek, J.,
Kocot, K., Pytlakowska, K., Talik, E., Krafft, C., Sitko, R., 2022. Sensitive determination Jiricek, P., Bieloshapka, I., 2014. Graphene oxide and reduced graphene oxide
of uranium using β-cyclodextrin modified graphene oxide and X-ray fluorescence studied by the XRD, TEM and electron spectroscopy methods. J. Electr., Spectr. Rel.
techniques: EDXRF and TXRF. Talanta 246, 123501–123509. https://doi.org/ Phenom. 195, 145–154. https://doi.org/10.1016/j.elspec.2014.07.003.
10.1016/j.talanta.2022.123501. Su, C.-K., Chen, Y.-T., Sun, Y.-C., 2019. Speciation of trace iron in environmental water
Koma, Y., Murakami, E., 2019. Contamination of Fukushima Daiichi nuclear power using 3D-printed minicolumns coupled with inductively coupled plasma mass
station with actinide elements. Radiochim. Acta 107, 965–977. https://doi.org/ spectrometry. Microchem. J. 146, 835–841. https://doi.org/10.1016/j.
10.1515/ract-2019-3126. microc.2019.02.015.
Kowal-Fouchard, A., Drot, R., Simoni, E., Ehrhardt, J.J., 2004. Use of spectroscopic Sun, Y., Shao, D., Chen, C., Yang, S., Wang, X., 2013. Highly efficient enrichment of
techniques for uranium(VI)/montmorillonite interaction modeling. Environ. Sci. radionuclides on graphene oxide-supported polyaniline. Environ. Sci. Technol. 47,
Technol. 38, 1399–1407. https://doi.org/10.1021/es0348344. 9904–9910. https://doi.org/10.1021/es401174n.
Kregsamer, P., Streli, P., Wobrauschek, C., 2002. Total-Reflection X-ray Fluorescence, in Sun, Y., Wu, Z.-Y., Wang, X., Ding, C., Cheng, W., Yu, S.-H., Wang, X., 2016. Macroscopic
Handbook of X-ray Spectrometry, 2nd ed. Marcel Dekker, New York. and microscopic investigation of U(VI) and Eu(III) adsorption on carbonaceous
Li, B., Ma, L., Tian, Y., Yang, X., Li, J., Bai, C., Yang, X., Zhang, S., Li, S., Jin, Y., 2014. nanofibers. Environ. Sci. Technol. 50, 4459–4467. https://doi.org/10.1021/acs.
A catechol-like phenolic ligand-functionalized hydrothermal carbon: One-pot est.6b00058.
synthesis, characterization and sorption behavior toward uranium. J. Hazard. Mater. Thommes, M., Kaneko, K., Neimark, A.V., Olivier, J.P., Rodriguez-Reinoso, F.,
271, 41–49. https://doi.org/10.1016/j.jhazmat.2014.01.060. Rouquerol, J., Sing, K.S.W., 2015. Physisorption of gases, with special reference to
Liang, L., Xi, F., Tan, W., Meng, X., Hu, B., Wang, X., 2021. Review of organic and the evaluation of surface area and pore size distribution (IUPAC Technical Report).
inorganic pollutants removal by biochar and biochar-based composites. Biochar 3, Pure Appl. Chem. 87, 1051–1069. https://doi.org/10.1515/pac-2014-1117.
255–281. https://doi.org/10.1007/s42773-021-00101-6. Tian, G., Geng, J., Jin, Y., Wang, C., Li, S., Chen, Z., Wang, H., Zhao, Y., Li, S., 2011.
Liu, F., Hua, S., Wang, C., Hu, B., 2022. Insight into the performance and mechanism of Sorption of uranium(VI) using oxime-grafted ordered mesoporous carbon CMK-5.
persimmon tannin functionalized waste paper for U(VI) and Cr(VI) removal. J. Hazard. Mater. 190, 442–450. https://doi.org/10.1016/j.jhazmat.2011.03.066.
Chemosphere 287, 132199. https://doi.org/10.1016/j.chemosphere.2021.132199. Uranium - Water Quality Association; https://wqa.org/learn-about-water/common
Ma, D., Hua, S., Lia, Y., Xu, Z., 2020. Adsorption of uranium on phosphoric acid- contaminants/uranium (Accessed 8 August 2022).
activated peanut shells. Sep. Sci. Technol. 55, 1623–1635. https://doi.org/10.1080/ Verma, S., Dutta, R.K., 2015. A facile method of synthesizing ammonia modified
01496395.2019.1606016. graphene oxide for efficient removal of uranyl ions from aqueous medium. RSC Adv.
Mellah, A., Chegrouche, S., Barkat, M., 2006. The removal of uranium(VI) from aqueous 5, 77192–77203. https://doi.org/10.1039/c5ra10555b.
solutions onto activated carbon: Kinetic and thermodynamic investigations. J. Coll. Wagstaffe, M., Thomas, A.G., Jackman, M.J., Torres-Molina, M., Syres, K.L., Handrup, K.,
Inter. Sci. 296, 434–1411. https://doi.org/10.1016/j.jcis.2005.09.045. 2016. An experimental investigation of the adsorption of a phosphonic acid on the
Mishra, S., Dwivedi, J., Kumar, A., Sankararamakrishnan, N., 2016. The synthesis and anatase TiO2(101) surface. J. Phys. Chem. C 120, 1693–1700. https://doi.org/
characterization of tributyl phosphate grafted carbon nanotubes by the floating 10.1021/acs.jpcc.5b11258.
catalytic chemical vapor deposition method and their sorption behavior towards Wang, S., Shi, L., Yu, S., Pang, H., Qiu, M., Song, G., Fu, D., Hu, B., Wang, X., 2022. Effect
uranium. New J. Chem. 40, 1213–1221. https://doi.org/10.1039/c5nj02639c. of Shewanella oneidensis MR-1 on U(VI) sequestration by montmorillonite.
Oh, Y.J., Yoo, J.J., Kim, Y.I., Yoon, J.K., Yoon, H.N., Kim, J.-H., Park, S.B., 2014. Oxygen J. Environ. Radioact. 242, 106798 https://doi.org/10.1016/j.jenvrad.2021.106798.
functional groups and electrochemical capacitive behavior of incompletely reduced Wang, Y.-Q., Zhang, Z.-B., Liu, Y.-H., Cao, X.-H., Liu, Y.-T., Li, Q., 2012. Adsorption of U
graphene oxides as a thin-film electrode of supercapacitor. Electrochim. Acta 116, (VI) from aqueous solution by the carboxyl-mesoporous carbon. Chem. Eng. J.
118–128. https://doi.org/10.1016/j.electacta.2013.11.040. 198–199, 246–253. https://doi.org/10.1016/j.cej.2012.05.112.
Olchowski, R., Giannakoudakis, D.A., Anastopoulos, I., Barczak, M., Zięba, E., Who, 2003. Guidelines for Drinking Water Quality, 3rd edition,.
Dobrowolski, R., Dobrzyńska, J., 2021. Arsenazo III removal from diagnostic Wnuczek, K., Podkościelna, B., Sobiesiak, M., Szajnecki, L., Goliszek, M., 2020. Synthesis
laboratories wastewater by effective adsorption onto thermochemically modified and modification by carbonization of styrene-ethylene glycol dimethacrylate-lignin
ordered mesoporous carbon. Environ. Nanotech. Monitor. Managem. 16, 100607 sorbents and their sorption of acetylsalicylic acid. Materials 13, 1761. https://doi.
https://doi.org/10.1016/j.enmm.2021.100607. org/10.3390/ma13071761.
Peng, H., Gao, P., Chu, G., Pan, B., Peng, J., Xing, B., 2017. Enhanced adsorption of Cu(II) Xu, W., Liu, J., Sun, K., Liu, Y., Chen, C., Wang, A., Sun, H., 2021. Effect of activation
and Cd(II) by phosphoric acid-modified biochars. Environ. Pollut. 229, 846–853. temperature on properties of H3PO4-activated carbon. Bio Resources 16, 4007–4020.
https://doi.org/10.1016/j.envpol.2017.07.004. https://doi.org/10.15376/biores.16.2.4007-4020.
Podkościelna, B., Goliszek, M., Sevastyanova, O., 2017. New approach in the application Zare, F., Ghaedi, M., Daneshfar, A., Agarwal, S., Tyagi, I., Saleh, T.A., Gupta, V.K., 2015.
of lignin for the synthesis of hybrid materials. Pure Appl. Chem. 89, 161–171. Efficient removal of radioactive uranium from solvent phase using AgOH–MWCNTs
https://doi.org/10.1515/pac-2016-1009. nanoparticles: Kinetic and thermodynamic study. Chem. Eng. J. 273 (2), 96–306.
Puziy, A.M., Poddubnaya, O.I., Gawdzik, B., Tascón, J.M.D., 2020. Phosphorus- https://doi.org/10.1016/j.cej.2015.03.002.
containing carbons: preparation, properties and utilization. Carbon 157, 796–846. Zielinski, W., Rajca, A., 2000. Spectroscopic methods and their application for
https://doi.org/10.1016/j.carbon.2019.10.018. identification of organic compounds, 2nd Ed. PWN, Warsaw.
Qiu, M., Liu, L., Ling, Q., Cai, Y., Yu, S., Wang, S., Fu, D., Hu, B., Wang, X., 2022. Biochar
for the removal of contaminants from soil and water: a review. Biochar 4, 1–25.
https://doi.org/10.1007/s42773-022-00146-1.

12

You might also like