You are on page 1of 12

Journal of Colloid and Interface Science 419 (2014) 148–159

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


www.elsevier.com/locate/jcis

Removal of sulfamethoxazole sulfonamide antibiotic from water by high


silica zeolites: A study of the involved host–guest interactions
by a combined structural, spectroscopic, and computational approach
Sonia Blasioli a, Annalisa Martucci b, Geo Paul c,d, Lara Gigli b, Maurizio Cossi c,d, Cliff T. Johnston e,
Leonardo Marchese c,d, Ilaria Braschi a,d,⇑
a
Department of Agricultural Sciences, University of Bologna, Viale G. Fanin 44, 40127 Bologna, Italy
b
Department of Physics and Earth Sciences, University of Ferrara, Via G. Saragat, 1, 44100 Ferrara, Italy
c
Dipartimento di Scienze e Innovazione Tecnologica, Università del Piemonte Orientale A. Avogadro, Viale T. Michel 11, 15121 Alessandria, Italy
d
Centro Interdisciplinare Nano-SiSTeMI, Università del Piemonte Orientale A. Avogadro, Viale T. Michel 11, 15121 Alessandria, Italy
e
Department of Agronomy, Purdue University, 915 West State Street, West Lafayette, IN 47907-2053, USA

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

Article history: Sulfonamide antibiotics are persistent pollutants present in surface and subsurface waters in both agri-
Received 27 August 2013 cultural and urban environments. Sulfonamides are of particular concern in the environment because
Accepted 16 December 2013 they are known to induce high levels of bacterial resistance. Adsorption of sulfamethoxazole sulfonamide
Available online 24 December 2013
antibiotic into three high silica zeolites (Y, mordenite, and ZSM-5) with pore opening sizes comparable to
sulfamethoxazole dimensions is reported. Sulfamethoxazole was almost completely removed from water
Keywords: by zeolite Y and MOR in a few minutes. Adsorption onto ZSM-5 showed an increased kinetics with
Adsorption
increasing temperature. Antibiotic sorption was largely irreversible with little antibiotic desorbed. Sulfa-
Water depollution
Zeolite Y
methoxazole incorporation and localization into the pore of each zeolite system was defined along with
Mordenite medium-weak and cooperative host–guest interactions in which water molecules play a certain role only
ZSM-5 in zeolite Y and mordenite.
Ó 2013 Elsevier Inc. All rights reserved.

1. Introduction employed. Using biological treatment with activate sludge, SMX


removal efficiency ranged between 100% and <25% [36]. Two
Sulfamethoxazole (SMX) is a broad spectrum antibiotic used in reasons for the high variability could have been the reconversion
human and veterinary therapy which belongs to the sulfonamide of the N1-acetyl derivate in the parent compound, which would
family (sulfa drugs). Like other sulfonamides, SMX is a competitor have underestimated SMX removal efficiency, and to its anionic
of p-aminobenzoic acid in the biosynthesis of tetrahydrofolic acid. nature (pKa2 = 5.7 [26]) that limits sorption on activate sludge
In studies on humans it was found that 20% of SMX daily dose was resulting in lower efficiency [38]. As a result, wastewater treat-
excreted via urine/feces in its original form, 50–60% was trans- ment plants represent potential reservoirs of sulfonamides that
formed to acetylated derivative (N1-acetylsulfamethoxazole, see may contribute to the evolution and spread of antibiotic resistance
Fig. 1) and 15–20% to glucuronide conjugate [30,31,13]. [33]. Owing to its anionic nature, SMX is not strongly retained by
The inefficiency of sewage treatment plants to degrade sulfon- soils containing organic carbon and clays [15,16] and moves
amides and the practice to land apply sulfonamide-containing through the soil profile accumulating in the aquatic environment
manure for agronomic purpose contribute to the spreading of anti- [7]. The most important consequence of antibiotic release in natu-
biotics in the environment. As reported by Michael et al. [36], SMX ral environments is the development of resistant bacteria [18].
concentrations in urban wastewater treatment plants depend on Several techniques have been developed to degrade and remove
its consumed amount and the type of wastewater treatment antibiotics from water and wastewater [34]. Among these, adsorp-
tion on activated carbon is one of the most efficient; however, the
⇑ Corresponding author at: Department of Agricultural Sciences, University of high cost of the adsorbent, their susceptibility of pore blocking, and
Bologna, Viale G. Fanin 44, 40127 Bologna, Italy. Fax: +390512096203. the difficulties of regeneration are disadvantages [9]. As an alterna-
E-mail addresses: sonia.blasioli@unibo.it (S. Blasioli), annalisa.martucci@unife.it tive low-cost adsorbent, this study evaluated the mechanisms of
(A. Martucci), geo.paul@mfn.unipmn.it (G. Paul), 90647@studenti.unimore.it SMX removal efficiency from water using several commercial high
(L. Gigli), maurizio.cossi@mfn.unipmn.it (M. Cossi), cliffjohnston@purdue.edu
silica (HS) zeolites.
(C.T. Johnston), leonardo.marchese@mfn.unipmn.it (L. Marchese), ilaria.braschi@
unibo.it (I. Braschi).

0021-9797/$ - see front matter Ó 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.jcis.2013.12.039
S. Blasioli et al. / Journal of Colloid and Interface Science 419 (2014) 148–159 149

Fig. 1. DFT optimized stick and ball (left) and electron density (right) structure of sulfamethoxazole. Geometry optimization with GGA functional BLYP. Plot of total electron
density with isosurface at 8  103 Å3.

Zeolites are microporous crystalline alumino-silicates exten- supernatant was withdrawn and analyzed by HPLC. The amount
sively used because of their high surface area, adsorption capacity of adsorbed antibiotics was calculated by the difference between
and cationic exchange capacity. Moreover, the sorption of the initial and final concentration. Controls run in the absence of
hydrophobic and hydrophilic solutes can be modulated by control zeolites allowed to verify the stability of SMX and any possible
of zeolite SiO2/Al2O3 ratio. Of particular interest in this study is the adsorption of the drug onto the centrifuge tubes. The experiments
relation between the molecular dimensions of the zeolite pores were carried out in triplicate.
relative to the size of SMX.
Studies on the use of high silica (HS) zeolites (high SiO2/Al2O3 2.3. Adsorption kinetics
ratio) adsorbents for pharmaceutical compounds from water are
limited [37,4,12,19,20,6,24,25]. HS zeolite Y has been tested for Kinetics were performed at RT on zeolite Y and MOR and at
the adsorption of sulfa drugs from soil solution containing salts 65 °C on ZSM-5 with a zeolite:antibiotic solution ratio of
and dissolved organic matter with excellent results (ca. 20% zeolite 1 mg:2 mL. SMX was weighted on a Mettler Toledo AT21 Compar-
dry weight – DW [4,5]). In the light of these recent findings, we be- ator balance (weigh limit = 1 lg). The concentration of adsorbed
lieved interesting to investigate the adsorption of SMX from water antibiotic was obtained by HPLC as reported in Section 2.2. The
by a series of HS zeolites and the effect of water on the host–guest experiments were conducted in triplicate. Control experiments
interactions between zeolite and antibiotic. were conducted to verify the SMX stability for the entire duration
of the kinetics.

2. Materials and methods 2.4. Adsorption and desorption isotherms

2.1. Chemicals Owing to the moderate SMX solubility and to the high adsorp-
tion capacity of the selected zeolites, the adsorption isotherms
Sulfamethoxazole (4-amino-N-(5-methylisoxazol-3-yl)ben- were built by mean of subsequent adsorption cycles with zeo-
zenesulfonamide (DFT optimized structure in Fig. 1); MW lite:antibiotic solution ratio of 1 mg:2 mL as follows: suspensions
253.3 g mol1; Kow 0.4: [27]; pKa2 5.7), was purchased as analytical of zeolites Y or MOR in the antibiotic solution at maximal solubility
standard by Dr. Ehrenstorfer GmbH (Germany) with a purity of were shaken for 30 min, whereas ZSM-5 for 24 h. Then the suspen-
99%. sions were centrifuged and the supernatant analyzed by HPLC. The
Because SMX contains ionizable functional groups, its solubility supernatant was then completely removed and substituted by
depends on solution pH [23]. A SMX stock solution at the maximal fresh antibiotic solution. Several subsequent adsorption steps were
solubility was prepared by adding the antibiotic to distilled water performed until the zeolites reached the maximum adsorption
in equilibrium with the atmospheric CO2 (ca. pH 6.0), in amount capacity (from here on called exhausted/loaded zeolites). All
exceeding that required to saturate the solution. The suspension experiments were conducted in triplicate.
was sonicated for 15 min, shaken at 50 °C for 30 min and, after Adsorption data were expressed as Ce (lM) – the antibiotic con-
cooling at room temperature (RT), filtered through 0.45 lm Dura- centration in aqueous phase at equilibrium – and Cs (lmol g1
poreÒ membrane filters to eliminate the undissolved solute. SMX adsorbent) – the amount of antibiotics adsorbed onto zeolite (i.e.
solubility, measured by means of high performance liquid chroma- the sum of adsorbed antibiotic after each adsorption step) – calcu-
tography (HPLC), was 202.3 ± 4.2 lM at RT. lated by the difference between the initial and final (Ce)
Three different HS zeolites have been chosen: Y, mordenite concentrations.
(MOR), and ZSM-5. Powders of zeolites Y and MOR (code Desorption isotherms were performed on loaded zeolites ob-
HSZ-390HUA and HSZ-690 HOA, respectively) were purchased in tained after the last adsorption step by substituting one SMX solu-
their protonic form from Tosoh Corporation (Japan). ZSM-5 (code tion half-volume with distilled water. The new suspension was
TSP-3022) was supplied in its ammonium form (<0.04%) by Tricat shaken for 24 h, centrifuged and the supernatant analyzed by
(USA). HPLC. Then, a supernatant second half-volume was again removed
and substituted with distilled water. The dilution step was re-
2.2. Temperature effect on adsorption peated several times, until no decrease was detected in the SMX
equilibrium concentration.
Adsorption experiments were performed at both RT and 65 °C
by adding each zeolite to a 30 lM antibiotic solution (zeolite: 2.5. Chromatographic analysis
antibiotic solution ratio of 1 mg:2 mL) in polyallomer centrifuge
tubes (Nalgene, NY, USA). Suspensions were shaken for 24 h and SMX concentration was determined by a Jasco HPLC-Diodarray.
then centrifuged at 20,000g for 15 min. Finally, an aliquot of the The system was assembled with the detector set at 267 nm, Jasco
150 S. Blasioli et al. / Journal of Colloid and Interface Science 419 (2014) 148–159

ChromNAV1.14.01 chromatography data software, a Jones Chro- 2.11. Solid state nuclear magnetic resonance spectroscopy
matography model 7971 column heater, and a 4.60 mm  150 mm
Synergi 4 lm Hydro-RP 80A analytical column (Phenomenex, All solid state NMR (SS-NMR) spectra were acquired on a Bruker
USA). The column was kept at 35 °C and eluted with acetoni- Avance III 500 spectrometer and a wide bore 11.7 T magnet with
trile:water eluant (23:77 by volume, pH 2.7 for H3PO4) at 1 mL/ operational frequencies for 1H and 13C of 500.13 and
min. Under these conditions, SMX retention time was 11.5 min. 125.77 MHz, respectively. A 4 mm triple resonance probe with
All solvents were HPLC grade. MAS was employed in all the experiments. The samples were
packed on a Zirconia rotor and spun at a MAS rate between 10
2.6. Thermogravimetric analysis and 15 kHz. For the 13C{1H} CPMAS experiments, the magnetic
fields tHrf of 55 and 28 kHz were used for initial excitation and
Thermogravimetric analyses (TGA) were carried out using a decoupling, respectively. During the CP period the 1H RF field tH rf
TGA/SDTA851e, Mettler Toledo. About 10 mg of sample (SMX, bare was ramped using 100 increments, whereas the 13C RF field tCrf
and loaded zeolites) were heated in a ceramic crucible under air was maintained at a constant level. During the acquisition, the pro-
flux at 25–1000 °C temperature range and 10 °C/min. tons are decoupled from the carbons by using a TPPM decoupling
scheme. A moderate ramped RF field tH rf of 62 kHz was used for
2.7. Specific surface area and pore volume measurements spin locking, while the carbon RF field tCrf was matched to obtain
optimal signal and the CP contact time of 2 ms was used. The relax-
The specific surface area and pore size distribution were mea- ation delay between accumulations was 0.5–75 s and all chemical
sured by nitrogen adsorption–desorption isotherms obtained at li- shifts are reported using d scale and are externally referenced to
quid nitrogen temperature (196 °C) in the 5  106–760 Torr TMS at 0 ppm. Dehydration of the adducts were carried out (sam-
(1 Torr = 133.33 Pa) pressure range using a Micromeritics Tristar ple packed on the NMR rotor) in a special hand-made quartz cell
3000 sorptometer (Norcross, GA, USA). Prior to analysis, samples that was connected to a high vacuum line (residual pressure
were outgassed at 150 °C and 106 Torr for a minimum of 12 h. p < 104 mbar), allowing in situ heat treatment at 120 °C for 2 h.
BET surface areas were calculated from the linear part of the BET Then the rotor was extracted from the cell, rapidly closed and
plot and BJH pore sizes were obtained from adsorption isotherms. was submitted to SS-NMR analysis (both 1H and 13C CPMAS NMR
studies).
2.8. X-ray diffraction and structure determination

3. Results and discussion


X-ray powder diffraction patterns of zeolites after SMX
adsorption (Y-SMX, MOR-SMX, and ZSM-5-SMX, respectively) were
The HS Y zeolite (Y) structure is characterized by eight super-
measured on a Bruker D8 Advance Diffractometer equipped with a
cages per unit cell with a diameter of approximately 12 Å joined
Sol-X detector, using Cu Ka1, a2 radiation. The spectra were col-
via circular windows and delimited by four 12-membered ring
lected in the 3–110° 2h range with a counting time of 12 s step1.
(MR) pore openings (7.4 Å) per cage [2]. The MOR structure is char-
Rietveld profile fittings were performed using the GSAS [17]
acterized by straight 12MR (6.5 Å  7.0 Å) and 8MR (5.7 Å  2.6 Å)
computer program with the EXPGUI interface [29]. Full matrix
channels running along the c axis. The 12MR channels are inter-
least-square refinements of SMX loaded zeolites were carried out
connected along [0 1 0] through 8MR side pockets (3.9 Å). The
in the Fd-3 (for zeolite Y), Cmc21 (for MOR), and P212121 (for
ZSM-5 structure consists of two intersecting sets of tubular chan-
ZSM-5) space groups, respectively. The crystallographic data and
nels, delimited by 10MR openings. The sinusoidal channel system
refinement details are reported as Supporting information (SI) in
is parallel to the [1 0 0] direction and has a cross section of
Table 1S. The final atomic positions and thermal parameters are gi-
5.1 Å  5.5 Å. The straight channel system runs parallel to the
ven in Tables 2S–4S.
[0 1 0] direction and has 5.3 Å  5.6 Å [0 1 0] openings. Structural
and physico-chemical characteristics of the three zeolites are given
2.9. Infrared analysis
in Table 1.
Like other sulfonamides, the molecular structure of SMX is not
Absorption spectra were collected on a Thermo Electron Corpo-
linear, but shows a distorted ‘‘V-shaped’’ geometry. The molecular
ration FT Nicolet 5700 Spectrometer with 4 cm1 resolution. DRIFT
dimensions (Fig. 1). Make possible the diffusion of the antibiotic
and ATR-FTIR spectra were obtained on a Perkin–Elmer GX2000
into the zeolites.
Fourier Transform infrared spectrometer equipped with MCT and
DTGS detectors with 4 and 2 cm1 unapodized resolution, respec-
tively. Further details are available as SI. 3.1. Adsorption

2.10. Computational modeling SMX adsorption onto the selected zeolites has been carried out
at RT and 65 °C for 24 h (see Fig. 1S in the Supplemental informa-
All the calculations were performed with GAUSSIAN03 code tion (SI) section) in order to evaluate the effect of temperature on
[11], at the Density Functional Theory (DFT) level, using the hybrid adsorption. The choice of 65 °C was also done because some prac-
B3LYP functional [3] and the double-zeta polarized 6-31G(d,p) tical applications of these sorbents might have working conditions
[14,10] gaussian basis set. Solvent effects were included through such a high temperature (e.g. the thermophilic phase of manure/
the C-PCM [8] implemented in GAUSSIAN03. sewage composting often reaches this temperature and, as re-
Two SMX conformations have been considered, since two tauto- ported in Section 1, these wastes are usually highly polluted with
meric forms are possible, according to the protonation state of the sulfa drugs). Complete removal of SMX by zeolite Y was observed
sulfonamide and heterocyclic nitrogen atoms: the amide structure at both temperatures, in agreement with prior work [4]. Decreased
(Fig. 1) resulted the most stable, while the imide form relative en- SMX sorption on MOR occurred was observed at elevated temper-
ergy was 8.00 kcal mol1. ature where the sorbed amount was 65% of the initial concentra-
Harmonic vibrational frequencies were computed for amide tion at RT but below 40% at 65 °C. A similar trend was observed
conformation in its optimized geometry, and compared to experi- for sulfachloropyridazine sulfonamide sorption onto zeolite MOR
mental IR spectra. [20]. In contrast, the higher temperature favored SMX adsorption
S. Blasioli et al. / Journal of Colloid and Interface Science 419 (2014) 148–159 151

Table 1
Structural and physico-chemical characteristics of zeolites under investigation.

Zeolite Cation type SiO2/Al2O3 (mol/mol) Unit cell Chemical formula


+
Y H 200 [Si192O384]

MOR H+ 200 [Si48O96]

ZSM-5 NHþ
4
500 [Si96O192]

onto ZSM-5 which increased from 40% at RT to 75% at 65 °C. Given is attributed to the presence of the larger heterocyclic ring (6
that diffusivity of species inside zeolite porosities is favored at in- membered chloropyridazine ring) of sulfachloropyridazine
creased temperature and that adsorption is unfavored (being an compared to the 5 membered methylisoxazole ring of SMX.
exhothermic process), the amount retained by zeolite is a result Taking into consideration the significant differences of SMX
of a counterbalance between the two processes. Among the three sorption on Y and MOR compared to ZSM-5, subsequent adsorption
zeolites studied, ZSM-5 has the smallest channel system, and the experiments on zeolite Y and MOR were performed using 30 min
reduced diffusivity of drug molecules at RT into its pore system contact time compared to 24 h for ZSM-5 (Fig. 3). At low SMX con-
is the limiting factor. Consequently, the SMX adsorption kinetics centrations of <20 lM, SMX showed an affinity for the zeolites in
and isotherms have been collected at RT in zeolite Y and MOR the order: Y  ZSM-5  MOR. At SMX concentrations >180 lM,
and at 65 °C in ZSM-5. the affinity of this zeolite for the drug was undoubtedly the highest
SMX adsorption kinetics are shown in Fig. 2. SMX was quickly being in the order: ZSM-5  Y  MOR. The adsorption isotherms
adsorbed on zeolite Y and MOR with <25% of the initial SMX onto both Y and MOR zeolites were L-type (plateau calculated val-
concentration remaining in solution after <1 min of contact. In ues = 1250 and 270 lmol g1, R2 = 0.991 and 0.985, respectively).
contrast, adsorption equilibrium was not achieved even after The adsorption isotherm of SMX onto ZSM-5 was S-shaped
2 weeks on ZSM-5. In the case of sulfachloropyridazine, 4 h was showing higher affinity of SMX at higher SMX concentrations.
required to reach the adsorption equilibrium onto MOR [20]. This SMX sorption on ZSM-5 did not achieve a sorption plateau that

30 (a) Y MOR ZSM-5


30 (b) ZSM-5
SMX concentration

SMX concentration

25 25
20 20
(µM)

(µM)

15 15
10 10
5 5
0 0
0 20 40 60 0 100 200 300
Time (min) Time (h)

Fig. 2. Adsorption kinetics of sulfamethoxazole onto zeolites Y, MOR (at RT) and ZSM-5 (at 65 °C): (a) 1 h and (b) 14 d contact time.
152 S. Blasioli et al. / Journal of Colloid and Interface Science 419 (2014) 148–159

Y Mor ZSM-5 SSA of zeolite decreased by 79% resulting from SMX sorption; the
des Y des Mor des ZSM-5 corresponding decrease in pore volume was 49%. This sharp reduc-
1200 tion in both SSA and pore volume is attributed to SMX molecules
residing in the pores and channels and is in agreement with the
C s (µmol/g)

800 SMX sorption results (Fig. 3).


In ZSM-5, a decrease in SSA and pore volume of 56% and 48%,
400 respectively, were found. SMX is larger than that the smallest pores
in ZSM-5 resulting in an SMX adsorption capacity that is below sat-
uration. These features are in good agreement with the fact that
0
0 50 100 150 200 the adsorption data point investigated (ca. 200 lM along Ce axis,
Ce (µM) Fig. 3) follows in the isotherm branch at the highest concentration:
the corresponding adsorption isotherm data point is far away from
Fig. 3. Adsorption (colored marks) and desorption (white marks) isotherms of the saturation concentration, given that the slope of the corre-
sulfamethoxazole onto high silica zeolites. Contact time: 30 min at RT for zeolite Y sponding isotherm branch is linear. The loaded ZSM-5 sample
and MOR and 24 h at 65 °C for ZSM-5. Mean of three replicates, standard deviation
hence is seemingly able to adsorb an additional considerable
lower than 4%. (For interpretation of the references to color in this figure legend, the
reader is referred to the web version of this article.) amount of antibiotic.
The observed decrease of both SSA and total pore volume in ex-
hausted MOR was more limited: 32% and 21%, respectively. Though
was evident for Y and MOR zeolites. At SMX concentrations the remaining free surface and volume of the sorbent after antibi-
<180 lM, the initial curve slope was a C type (R2 = 0.998). At con- otic loading were found to be remarkably high, the investigated
centrations >180 lM, an isotherm branch with stiffer slope oc- adsorption data point lays in the plateau region of the adsorption
curred, still recalling a C type (R2 = 0.902). This trend can be isotherm. Likely, the lowest antibiotic loading onto MOR seems
explained assuming that, in 0–180 lM concentration range, the to be due to the limited diffusion of bulky SMX molecules inside
drug interacts with specific adsorption sites which are different its ‘‘mono-dimensional’’ pore system.
from those occupied in the adsorption branch at higher antibiotic
concentrations. Based on the crystal structure of ZSM-5 and the 3.3. Rietveld refinement
molecular dimensions of SMX, two adsorption sites are available
[21,22]. The first and most accessible sorption site is localized at Changes in both the positions and intensities (Figs. 3–5S) of the
the intersection of the linear and sinusoidal channels. Once these diffraction peaks on bare and loaded zeolites revealed structural
sites are occupied, sorption can onto sites within the linear chan- modifications induced by the presence of SMX.
nel. In the case of SMX, this hypothesis is supported by the Rietveld
analysis (vide infra, Fig. 6) that localize SMX molecule at the inter-
3.3.1. Zeolite Y
section of the two channels. The data suggest that once SMX mol-
The observed and difference Fourier maps generated using
ecules are present at the intersection of the linear and sinusoidal
GSAS program, revealed the presence of a number of extraframe-
channels, additional sorption into the smaller, less-accessible lin-
work ions inside the supercages, which can be attributed to encap-
ear channels can occur (>180 lM range). Adsorption can proceed
sulated organic molecules. The largest and intense peak in the
up to the highest concentration investigated where sites are still
difference Fourier map was attributed to the sulfur atom. The re-
available as shown by the almost unchanged slope of the isotherm
fined occupancy factor of sulfur atom allowed us to determine
branch in the 180–200 lM concentration range. The adsorption
the amount of sulfa drugs adsorbed in the zeolite channel systems.
process of SMX onto each zeolite is irreversible with essentially
SMX embedding caused a lowering in real Fd-3m symmetry in the
no SMX desorption was observed by water from any of the three
parent Y zeolite [4] to Fd-3 as well as a significant decrease in unit
zeolites (Fig. 3).
cell volume (Table 1S) in comparison to those of the untreated
According to HPLC analysis, the amount of SMX sorbed was
materials [4].
about 30%, 10%, and 5% (wt/wt) for Y, ZSM-5, and MOR, respec-
On the whole, Rietveld structure refinement revealed the pres-
tively (Table 2). These findings were also confirmed by TGA
ence of about 14 antibiotic molecules per unit cell, which are
(Fig. 2S) and XRPD results. Since SMX has been detected in muni-
hosted in the eight supercages of Y zeolite. As a consequence, about
cipal sewage treatment plant effluents at 0.1–2 lg mL1 concen-
2 molecules per unit cell were embedded inside each zeolite cage.
tration range [39], 1 g of zeolite Y could clean up more than 6 m3
All these molecules are hosted in crystallographic sites partially
water maximally contaminated with the drug.
statistically occupied. The cubic Fd-3 symmetry imposes that some
of these are symmetrically equivalent (for instance, C1 crystallo-
3.2. Porosimetric analysis
graphic site describes C1, C2, C3, C4, C5, and C6 atoms of the aniline
ring at the same time, as reported in Table 2S). The geometry of
Evidence of the entrapment of SMX molecules inside the zeolite
sulfa drug adsorbed in Y zeolite is consistent with that of pure
channels was obtained by comparing the specific surface area
SMX, reported by Takasuka and Nakai [28]. Fig. 4 shows that the
(SSA) and pore volume of the ‘bare’ zeolites (no SMX) with those
aniline ring is almost centered on the supercage with its plane per-
containing the maximum surface loading of SMX (Table 3). The
pendicular to the threefold axes of the unit cell, whereas the

Table 2
Maximal amount of sulfamethoxazole adsorbed onto zeolites. The unit cell composition of zeolite-SMX systems is also reported.

Zeolite SMX adsorbed onto zeolite (mg 100 mg1 DW zeolite) SD in parenthesis Unit cell composition TGA (XRPD)
HPLC TGA XRPD
Y 29.6 (0.6) 23.9 (3.1) 24.0 [Si192O384] 11.5 (14) SMX
MOR 5.2 (0.3) 5.8 (0.3) 5.0 [Si48O96] 0.6 (0.6) SMX
ZSM-5 9.6 (0.1) 8.0 (1.0) 6.5 [Si96O192] 1.9 (1.5) SMX
S. Blasioli et al. / Journal of Colloid and Interface Science 419 (2014) 148–159 153

Table 3
Specific surface area and total pore volume of high silica zeolites before and after sulfamethoxazole adsorption determined by N2 physisorption measurements and applying the
Brunauer–Emmet–Teller (BET) and non-local density functional theory (NLDFT) methods, respectively.

Sample Specific surface area (m2/g) Total pore volume at P/Po = 0.99 (cm3/g, STP)
Bare zeolite Loaded zeolite Bare zeolite Loaded zeolite
Y 648 135 0.57 0.29
MOR 441 301 0.28 0.22
ZSM-5 342 150 0.21 0.11

isoxazole ring is found with six different orientations. The aniline towards the side pocket. The presence of the mirror plane orthog-
ring (C1 site) is situated in the window that joins neighboring onal to [1 0 0] imposes that in isoxazole ring C8 crystallographic
supercages, whereas the positions occupied by isoxazole atoms site can host either C or N atoms whereas in C9 site O or C atoms
appear disordered and consequently are only partially localized can be hosted: owing to the position of SMX molecule, the real
(C7 site in Table 2S). The isoxazole and aniline rings of adsorbed symmetry of MOR-SMX is lower than Cmc21.
SMX form a typical V configuration with the torsion angle The isoxazole and aniline rings form a typical V configuration
SAN2AC7 of about 126°, with respect to that (121.1°) of crystalline with the torsion angle SAN2AC7 of 123.8°, which is slightly larger
SMX [28]. The refined distances of N2 (N2AO1 = 2.68(11) Å and than that reported for crystalline SMX [28] (121.1°). The positions
N2AO4 = 3.29(9) Å, respectively) indicate interactions with of oxygen atoms bound to the sulfur and those of methyl group are
framework oxygens (Fig. 4). The 12MR increased its Crystallo- not localized, due to a strong disorder due to sorption.
graphic Free Area (CFA) when compared with the bare material, The refined distances of SMX atoms with the framework oxy-
whereas the double 6MR (D6R) unit is contracted (Fig. 4). This gens (C9AO9 = 2.80(4) Å, C8/N3AO4 = 2.89(4) Å, respectively)
combined effect of widening/contraction of pore openings justifies indicate that these species strongly interact with the framework
the variations in the unit cell parameters reported in Table 1S thus altering the framework geometry. After SMX adsorption, the
[4,35]. 12MR channel widens and becomes more circular thus increasing
its CFA (Fig. 5). The 12MR widening causes a narrowing of both
3.3.2. MOR zeolite 8MR channel and 8MR of the side pocket (see CFA values in
No significant change of unit cell parameters occurred Fig. 5). This widening/contraction mechanism is responsible of
(Table 1S) for MOR zeolite. However, the TAO11AT angle changed the small unit cell volume variation reported in Table 1S.
from 156.4° (MOR-SMX) to 172.8° (MOR) which clearly revealed a
strong deviation from the topological Cmcm symmetry. A similar 3.3.3. ZSM-5 zeolite
lowering of symmetry in MOR after sulfachloropyridazine [20] As reported in Table 1S, variations of the unit cell parameters in
and toluene [1] adsorption has been reported. comparison to those of the untreated material [20] indicated a
Rietveld structure refinement revealed the presence of about monoclinic (P21/n) orthorhombic (P212121) phase transition
0.6 antibiotic molecules per unit cell, located in the large 12MR depending on the nature and/or amount of encapsulated mole-
channel, with the aniline ring (crystollographic sites C1, C2, and cules. Rietveld structure refinement revealed the incorporation of
C3, corresponding to C1/C6, C2/C5, C3/C4 atom numbering of about 1.5 antibiotic molecules per unit cell: the aniline ring being
Fig. 1, respectively) oriented parallel to the [0 1 0] direction at the intersection of straight and sinusoidal channels, and the
(Fig. 5). The SMX position is similar to that recently found for sul- isoxazole ring protruding towards the sinusoidal channel.
fachloropyridazine loaded into the same high silica MOR [20]. Also The positions occupied by the oxygen atoms bound to the sulfur as
in the case of SMX, the drug molecules show two different orienta- well as those of methyl group in ZSM-5-SMX, appear strongly disor-
tions due to the screw axis parallel to [0 0 1], with the isoxazole dered and consequently are not localized. The torsion angle SAN2AC7
ring (crystallographic sites C7, C8, and C9 corresponding to C7, in SMX loaded into the zeolite was about 118.7°, which is smaller than
C8/N3, C9/O3 atom numbering of Fig. 1, respectively) oriented in pure crystalline antibiotic (121.1° [28]). The distances between N1

Fig. 4. Location of sulfamethoxazole (SMX) and oxygen–oxygen distance (Å) of the apertures in zeolite Y after drug adsorption, assuming an oxygen ionic radius of 1.35 Å.
Crystallographic Free Area (CFA) and ellipticity (e) are also reported.
154 S. Blasioli et al. / Journal of Colloid and Interface Science 419 (2014) 148–159

Fig. 5. Location of sulfamethoxazole (SMX) and oxygen–oxygen distance (Å) of the apertures in mordenite (MOR) after drug adsorption, assuming an oxygen ionic radius of
1.35 Å. Crystallographic Free Areas (CFA) and ellipticity (e) are also reported.

Fig. 6. Location of sulfamethoxazole (SMX) and oxygen–oxygen distance (Å) of the apertures in ZSM-5 after drug adsorption, assuming an oxygen ionic radius of 1.35 Å.
Crystallographic Free Areas (CFA) and ellipticity (e) are also reported.
S. Blasioli et al. / Journal of Colloid and Interface Science 419 (2014) 148–159 155

and N3 with respect to framework oxygen atoms (N1AO2 = 2.90(8) Å, stretching of ANH2 group (masNH2 and msNH2, respectively),
N1AO20 = 2.76(9) Å, N3AO25 = 2.60(11) Å) suggested interactions of whereas the absorption at 3346 cm1 to the strectching of the sul-
SMX with the framework. fonamide ANH group (mNH). The absorption at 1624 cm1 is as-
The incorporation of SMX molecules in the ZSM-5 structure is signed to the bending of ANH2 (dNH2), and the features at lower
confirmed by a change in the dimension of 10MRs, when compared wavenumbers are due to phenyl or heterocycle ring vibrations cou-
to the parent zeolite (Fig. 6). After adsorption both sinusoidal and pled with the bending modes of amino and amide groups, along
straight 10MR channels become more expanded and their CFA in- with deformation modes of the methyl group. The strong OAH
crease (Fig. 6), thus indicating a framework flexibility of ZSM-5. stretching bands in the 3700–3000 cm1 range of water adsorbed
onto both bare and loaded zeolites obfuscated observation of the
3.4. FTIR SMX bands. In order to minimize the interference from H2O, spec-
tral interpretations of host–guest interactions were based upon
FTIR spectra of the three zeolites and SMX bound to Y, MOR and spectra obtained under vacuum.
ZSM-5 zeolites are compared in Fig. 7. Absorbance FTIR spectra re-
corded in vacuum (without water), DRIFT spectra (with a moderate 3.4.1. Zeolite Y-sulfamethoxazole system
water amount present), and ATR-FTIR spectra (with a large water Considering the absorption spectrum of loaded zeolite Y, the
amount present) are reported to evaluate whether the host–guest bands corresponding to the stretching and bending modes of the
zeolite-SMX interactions were affected by water. aniline amino group of the adsorbed antibiotic are slightly shifted
The interpretation of IR spectrum of SMX in CH2Cl2 and ad- with respect to the drug in CH2Cl2 (Table 4) indicating that the NH2
sorbed onto the zeolites was facilitated by comparing the observed group has weak interactions with the zeolite, in agreement with
absorptions with the theoretical frequencies computed for the the localization by Rietveld analysis of this ring in the center of
amidic form of the isolated molecule as shown in Fig. 7. Sulfa drugs the cage and far away from zeolite walls.
can exist as amide (ASO2ANHA) or imide (ASO2AN@) form of A more relevant downshift observed for the sulfonamide
monomeric or dimeric species [5,6,20]. In the case of SMX, the stretching mode (DmNH: 106 cm1) is indicative of either weak
experimental IR in CH2Cl2 matches the amide computed spectrum H-bond involving the sulfonamide group or a geometry modifica-
in good agreement with the stabilization energy of the amide spe- tion of the adsorbed molecule. Although the distance between
cies determined by DFT calculations which was 8.00 kcal/mol low- the sulfonamide nitrogen and the cage wall oxygens is short en-
er than the imide form. The most important SMX vibrational ough to give rise to strong H-bonding, the poor directionality of
frequencies are listed in Table 4. the NH group with framework oxygens likely leads to a weak H-
The experimental bands of the antibiotic in CH2Cl2 at 3503 and bond. The IR features are consistent with a zeolite-induced change
3406 cm1 were assigned to the asymmetric and symmetric in SMX geometry.

Fig. 7. DRIFT and in vacuum absorbance FTIR spectra of zeolite Y, MOR, and ZSM-5 before (dotted curve) and after (solid curve) sulfamethoxazole (SMX) adsorption. The
difference spectrum (gray curve) of the drug adsorbed into MOR obtained by subtracting the bare outgassed sample is reported. ATR spectrum of Y-SMX adduct in water is
also presented. Reference spectra of SMX in CH2Cl2 (black curve) and calculated (red curve) are also reported. Asterisk indicates unreliable band due to solvent subtraction.
(For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)
156 S. Blasioli et al. / Journal of Colloid and Interface Science 419 (2014) 148–159

Table 4
Calculated and experimental IR frequencies of sulfamethoxazole before and after adsorption onto high silica zeolites (n.d.: not detectable; –: absent).

Assignment IR wavenumber (cm1)


Computed In CH2Cl2 Adsorbed on
Y MOR ZSM-5
masNH2 3707 3503 3494, 3478 3516 3520
msNH2 3594 3406 3392 3422 3424
mNH 3565 3346 3240 n.d. 3356
dNH2 1676 1624 1622 1629 n.d.
mNC/CChet ring + dNH 1664 1614(sh) 1635–1610 1616 1635–1610
mPhquad + dNH2 1652 1597 1596 1597 1598
mPhsext 1543 1504 1503 1502 1508
mNC/CChet ring + dNH + mCACH3 1542 – – – –
mCChet ring + dasCH3 1508 1461 1464 1468 1470
mCChet ring + dasCH3 1485 – – – –
mPhsext + masSO2 + dNH 1389 1384 1384 – –
mPhsext + masSO2 + dNH 1371 1375 1373 – –
dCCHin-plane + mCANH2 1341 – – – 1340
mCOhet ring + masSO2 + dNH 1318 1311 1333, 1323 n.d. 1340, 1321
mCN/COhet ring + dNH + dasCH3 1304 – 1308 – –

At lower frequencies, the most significant spectroscopic change band in the 3400–3000 cm1 range with maximum at ca.
is due to the unresolved band of the antibiotic in CH2Cl2 at 3300 cm1). This is a clear-cut evidence of H-bond interactions of
ca.1614 cm1, assigned to the heterocycle ring CAN vibrations medium strength, which occur between silanols and SMX.
coupled with the sulfonamide dNH mode, which is increased in In the silanol-drug adduct, only the msNH2 at 3422 cm1 is
intensity, upshifted and overlapped to the dNH2 band at clearly visible. The weak band of masNH2 can be found in the differ-
1622 cm1 after adsorption into zeolite. A similar feature has been ence spectrum (red dotted curve) at 3516 cm1. The small upshift
observed for other sulfa drugs loaded into zeolite Y [5,6] and is of the NH2 group when the antibiotic is in adsorbed form, indicates
consistent with the occurrence of van der Waals interactions that this group is far apart from the silica walls in agreement with
between SMX heterocycle ring and the zeolite framework. the fact that the aniline ring protrudes toward the MOR channel
In the 1350–1250 cm1 region a complex IR band is found for center (Fig. 5). Unfortunately, the band of the stretching mode of
the adsorbed SMX that is absent in the spectrum in of SMX in sulfonamide NH is not clearly measurable for the adsorbed drug
CH2Cl2 solution. Likely, this difference can be mainly ascribed to because it is overlapped to the intense and broad band of MOR
modifications of the geometry of the sulfonamide moiety as a silanols.
consequence of the heterocycle ring interaction with the zeolite At lower wavenumber, the band at 1614 cm1 of SMX in CH2Cl2
framework. related to the stretching of the heterocycle ring coupled with sul-
Insight of the effect of water on the drug adsorption into zeolite fonamide dNH was strongly perturbed upon adsorption. Moreover,
Y can be obtained observing the ATR-FTIR spectrum of loaded zeo- the two bands at 1384 and 1375 cm1 in CH2Cl2 almost completely
lite Y recorded in water solution (Fig. 7). Interestingly, the band of disappear in loaded MOR, likely because the heterocycle ring is H-
the drug in CH2Cl2 at 1384 cm1 due to vibrations of the phenyl- bonded to zeolite silanols and the collective vibrations of this part
sulfonamide moiety (Table 4) is observed upshifted at 1402 cm1 of the molecule are strongly perturbed. This feature is consistent
in the ATR spectrum maybe due to the solvatation effect of the sul- with the localization of the antibiotic heterocycle ring partly inside
fonamide group. Moreover, the broad band of SMX in CH2Cl2 at the MOR side pocket (Fig. 5) with both the isoxazole ring nitrogen
1311 cm1, which is assigned to deformations of the heterocycle and oxygen atoms at H-bond distance with the oxygens of the MOR
ring coupled with the SO2 asymmetric stretching and the NH bend- side pocket entrance. A similar scenario was already observed for
ing, is found split into two new bands at 1339 and 1327 cm1 in sulfachloropyridazine adsorbed into the same MOR zeolite [20].
the ATR spectrum of the loaded zeolite. Likely, the solvation effect The similarity of the absorbance (outgassed) and DRIFT spectra
of water modifies the electron distribution of the drug in adsorbed rule out a significant water involvement in the stabilization of
form with respect to the molecule in CH2Cl2. the host–guest interactions observed.

3.4.2. MOR-sulfamethoxazole system 3.4.3. ZSM-5-sulfamethoxazole system


The bare MOR shows a broad and intense band in the region be- In the absorbance spectrum of the loaded zeolite, the bands as-
tween 3750 and 3000 cm1 where silanol groups and bound H2O signed to the stretching modes of the drug NH2 group are slightly
absorb (Fig. 7, dotted red curve). This high intensity is attributable upshifted with respect to the SMX in CH2Cl2 (Table 4): similarly to
to the large amount of silanol defects produced by the dealumina- the MOR, these small perturbations reveal that some H-bondings
tion process. A defined and weak band due to isolated silanols in- occur in CH2Cl2 and these are turned off in zeolite. Also the band
side microporosities is found at 3728 cm1 whereas the most related to the stretching mode of the sulfonamide NH is upshifted
intense and broad band in the range 3700–2900 cm1 is assigned after adsorption with respect to the free molecule, ruling out the
to H-bonded donor silanols [32]. involvement of this group in H-bonds with the zeolite network.
After antibiotic adsorption (Fig. 7, solid red curve), the band at At lower frequency, the two bands at 1384 and 1375 cm1,
3728 cm1 almost completely disappears whereas the intense which are active in CH2Cl2, almost completely disappear in the
band of H-bonded silanols is downshifted and enlarged at the same spectrum of the adsorbed drug similarly to the MOR-drug case:
time. These findings indicate that a fraction of silanols are bound to also in the ZSM-5-SMX system the heterocycle ring is the drug
the antibiotic. Clearly the band of silanols is eroded after antibiotic moiety more involved in the adsorption process, being localized
adsorption (negative band at 3750–3400 cm1 in the difference into the zeolite sinusoidal channel. Moreover, two bands of the
spectrum) and a new band is formed at lower frequencies (positive heterocycle ring vibrations at 1340 and 1321 cm1, which are not
S. Blasioli et al. / Journal of Colloid and Interface Science 419 (2014) 148–159 157

resolved in CH2Cl2, become active and well resolved in the spectra appears at 2.3 ppm. Y-SMX exhibits similar spectra before
of the adsorbed antibiotic. These bands are related to vibrations of (Fig. 8(I)a) and after dehydration (Fig. 8(I)b), except the fact that
the heterocycle ring which is localized into the smaller sinusoidal there is a difference in the region where water resonate. Before
channel and therefore they are found strongly perturbed with re- dehydration, a broad peak at 3.9 ppm is clearly visible and is due
spect to the molecule in CH2Cl2. The water amount occurring in contributions from physisorbed water, ANH2 and ANH. In addi-
DRIFT samples does not seem to affect the nature of the host–guest tion, isolated silanols appear as a shoulder of the main peak at
ZSM-5-SMX interactions based on similarity between the absor- 1.7 ppm after dehydration.
bance (water absent) and DRIFT (some water present) spectra. The behavior of MOR-SMX adduct is entirely different from
that of Y-SMX. Before dehydration, the resonance due to protons
3.5. Solid state NMR of water, ANH2 and ANH is dominant in this spectrum (Fig. 8(I)c)
and the amount of adsorbed water is significantly higher than in
Solid state NMR spectroscopy is a powerful method for the Y-SMX. After dehydration (Fig. 8(I)d), resonances due to ANH2
study of host–guest adducts, since it provides information about and isolated silanols are visible at 4.2 and 1.7 ppm, respectively.
the structure and dynamics of both guest and host over a wide- Furthermore, a dramatic increase in intensity for the methyl res-
range of time scales, length scales and temperature scales. In the onance after dehydration is evident and can be due to the reso-
1
H MAS NMR spectrum of pure sulfamethoxazole (SMX, Fig. 8(I)), nance contributions from ANH as well as isolated silanol
two broad signals in the range of 0.5–2.5 and 5.5–9.5 ppm due to resonances being in the same chemical shift range. This demon-
methyl and aromatic protons are found. The broadness is due to strates that both silanol and ANH groups are H-bonded to water
the overlapping of resonances arising from different polymorphs molecules in the hydrated sample. In the aromatic region, after
and/or due to the non-averaging of the 1HA1H dipolar interactions. dehydration, three resonances can be clearly distinguished; how-
On the other hand, the 1H MAS NMR spectra of the 3 adducts ever, the chemical shifts are different from the hydrated sample.
show a good resolution for a solid sample and consist of resonances In addition, a very broad low intense peak in the range 12–
from SMX, physisorbed H2O as well as the host zeolites. Fig. 8(I) 18 ppm is also visible in dehydrated MOR-SMX (Fig. 8(I)d, inset)
show the 1H MAS NMR spectra of adducts both before and after and are associated to H-bonded protons (highlighted by an ar-
dehydration in vacuum at 120 °C for 2 h. For adduct Y-SMX, the row). It is clearly evident that water plays a dramatic role in
peaks from the aromatic and heterocyclic rings are well resolved the arrangement of SMX molecules inside the mordenite cages
compared to that of pure SMX and are associated to the averaging and drug molecules reorganize after dehydration as water leaves
of the 1HA1H dipolar interactions. Resonance due to methyl group the channels/cages.

Fig. 8. 1H solid state MAS-NMR spectra of pure sulfamethoxazole (SMX), Y-SMX, MOR-SMX and ZSM5-SMX at a MAS rate of 15 kHz (I) before (a, c, e) and after dehydration (b,
d, f). Inset show the zoomed 1H MAS spectra of MOR-SMX before (red) and after dehydration (black). 13C solid state CPMAS-NMR spectra of SMX, Y-SMX, MOR-SMX and
ZSM5-SMX (II) before (a, c, e) and after dehydration (b, d, f). A cross polarization contact time of 2 ms and a MAS rate of 10 kHz was used in all the experiments. Inset show the
zoomed methyl region of 13C CPMAS spectra of ZSM5-SMX before (red) and after dehydration (black).  Denotes spinning sidebands. (For interpretation of the references to
color in this figure legend, the reader is referred to the web version of this article.)
158 S. Blasioli et al. / Journal of Colloid and Interface Science 419 (2014) 148–159

ZSM5-SMX adduct shows a similar behavior to that of MOR- positions. However, for the aromatic 13C pairs, namely C3/C5 and
SMX, however, the line widths are narrower. The spectra are iden- C2/C6, only one resonance each is observed in this adduct and
tical before (Fig. 8(I)e) and after dehydration (Fig. 8(I)f) apart from can be attributed to the fast flipping of aromatic ring. While such
the fact that there is a change in the region where physisorbed good resolution with sharp lines can be attributed either to a great-
water, ANH2 and ANH resonate. After dehydration, only resonance er mobility of SMX inside the cages or to a more regular packing
due to ANH2 is visible as a broad peak centered at 4.2 ppm in that with less equivalent orientations as in crystalline samples,
region. Additionally, methyl resonance appears as a doublet, sepa- although long range ordering is highly unlikely in ZSM5 due to
rated by 1 ppm apart, and confirms the presence on two distinct the restriction of space. However, we have observed a splitting in
environments for this group. The existence of multiple resonances the methyl carbon resonance which confirms the distinct environ-
at different chemical shifts, before and after dehydration, (aromatic ments this group experiences (Fig. 8(II)f, inset). It is important to
protons in MOR-SMX) as well as multiple resonances for identical note here that, in ZSM5-SMX, there are no visible changes observed
protons, before and after dehydration (methyl protons in ZSM5- in the spectra before and after dehydration. As was pointed out in
SMX), suggests that SMX molecules exist in a number of different the XRD data recorded at RT, the methyl group position in ZSM5-
chemical environments within the host–guest interfaces. This SMX was highly disordered and could not be localized. Further-
may arise either from the differences in the proximity between more, XRD data show that the aniline ring is at the intersection
heterocyclic ring (and/or aniline ring) and zeolite walls or from dif- of straight and sinusoidal channels, while the heterocyclic ring pro-
ferences in the packing order. Further evidences are available from trudes towards the sinusoidal channel. In fact, the methyl group
the 13C CPMAS NMR data. can experience different environments in such arrangement. Thus,
Standard CPMAS conditions at ambient temperature were used we have performed a low temperature 13C CPMAS experiment
to acquire the 13C[1H] CPMAS NMR spectra (before and after dehy- (data not shown), and the data recorded at 193 K show that the
dration) with short contact time of 2 ms and Fig. 8(II) shows spec- two resonances coalesce into one signal. It can be concluded that,
tra acquired on pure SMX and the three different adducts. The based on the combined 1H MAS and 13C CPMAS NMR as well as
assignments are based on theoretical calculations as well as on li- XRD data, SMX molecules occupy crystallographically unique posi-
quid NMR data (not shown). The sharp and narrow resonances are tion in framework channel intersections which resemble a crystal-
consistent with the highly ordered crystalline nature of the pure line state for the drug in ZSM5 host matrix. Furthermore, water
SMX sample [Fig. 8(II), topmost]. However, the behavior of the plays no role in the arrangement of SMX molecules in ZSM5
three adducts are different to that of pure crystalline SMX. channels.
The 13C CPMAS spectrum of Y-SMX before dehydration In summary, the observation of variable adsorption rates and
(Fig. 8(II)a) shows sharp resonances for each carbon particularly capacities of SMX on different host zeolite framework is of rele-
in the aromatic region, however, peaks broaden after dehydration vance, and suggests that non-covalent guest–host and guest–guest
(Fig. 8(II)b). The peak positions are different from that of the crys- interactions play significant role in the stability of these adduct
talline pure SMX, with the number of resonances as well as chem- systems. Water might play a role in the adsorption process of anti-
ical shifts being different in the aromatic region. This observation biotics in zeolites, however, necessarily not control the structure
shows that there is a fast flipping of rings in the cages of zeolite and dynamics of these molecules in the final adduct as was con-
Y in the hydrated state, probably assisted by water, leading to an firmed by the solid state NMR study. The exact location of the drug
averaging of the resonance peaks from the aromatic carbons, molecules inside the host system depends on different factors
whereas in the crystalline solid, the particular packing of mole- including, dimension of the drug molecules, the size of the cages/
cules implies a crystallographic unequivalence of the various 13C channels, the presence and concentration of other guest molecules
sites. The distinct behaviors in terms of observed line widths in like water, as well the non-covalent interactions between and
Y-SMX adduct, before and after dehydration can be explained by among guests and host systems. Clearly, this multi-technique ap-
steric considerations. It is important to note here that the SMX proach allows the investigation of the adsorption of wide range
loading in zeolite Y is around 24 wt% (TGA results, Table 2). At such of sulfa drug families on different host zeolite systems and to
high loading, it is expected that the SMX molecules are engaged in understand the structural and dynamic details of the adsorbed spe-
the zeolite cages probably through both host–guest and guest– cies. Such unprecedented detail understanding can help to apply
guest interactions. Molecular motion (e.g., flipping of rings and this principle to clean up polluted water bodies.
methyl group rotation) of SMX inside the cages at RT would disrupt
the molecular orientation and the motional rate is fast enough to
average the chemical shift anisotropy (CSA) of the aromatic 4. Conclusions
carbons as it is seen by sharp resonances in the hydrated state.
However, such CSA averaging does not take place when the SMX The efficacy of sulfamethoxazole (SMX) removal by high silica
molecules are physisorbed on the cage surface (as water molecules zeolites was evaluated based on the interaction of SMX with three
leave the cages) and broad peaks are considered to be due either to representative high silica zeolites (Y, MOR, and ZSM-5) each with a
a restriction of mobility or to a distribution of environments. different pore size and architecture. Zeolite Y and MOR adsorbed
The key difference in 13C CPMAS spectrum of MOR-SMX with SMX with a favorable kinetics (a few minutes) up to 24% and 6%
that of Y-SMX is that the resonances are generally broad even in zeolite DW, respectively. No equilibrium adsorption was reached
the hydrated state (Fig. 8(II)c). After dehydration (Fig. 8(II)d), onto ZSM-5 after two weeks, thus revealing a slow adsorption
changes in chemical shifts as well as line widths are observed. As kinetics. However, a surprisingly positive temperature effect was
was observed in the 1H MAS NMR spectra of MOR-SMX adduct, highlighted which can be relevant for applications in manure/sew-
water plays a cooperative role in the guest–host interactions be- age polluted with sulfa drugs, given that the thermophilic phase in
tween SMX and mordenite cages. As the water molecules leave, manure/sewage composting can reach high temperatures. The
SMX reorganizes inside the cages with heterocyclic ring being adsorption of SMX into ZSM-5 after 24 h contact at 65 °C reached
interacting with the cage surface through hydrogen bonding 8% zeolite DW. SMX desorption revealed the irreversibility of the
(Fig. 8(I)d, inset). adsorption process for each zeolite.
As for the remaining ZSM5-SMX adduct, the spectra (Fig. 8(II)e The location of SMX inside the zeolite porosities was defined by
and f) are very different from the other two adducts but similar to XRPD Rietveld analysis. The close vicinity of some drug moieties to
13
C CPMAS spectrum of pure crystalline SMX with matching peak the oxygen atoms of the zeolite voids suggests the occurrence of
S. Blasioli et al. / Journal of Colloid and Interface Science 419 (2014) 148–159 159

multiple host–guest interactions between each sorbent and SMX. [6] I. Braschi, G. Paul, G. Gatti, M. Cossi, L. Marchese, RSC Adv. 3 (2013) 7427–7437.
[7] M. Burkhardt, C. Stamm, J. Environ. Qual. 36 (2007) 588–596.
The type and strength of the interactions, which are responsible
[8] M. Cossi, N. Rega, G. Scalmani, V. Barone, J. Comput. Chem. 24 (2003) 669–681.
for the irreversible extraction of the drug from water, were eluci- [9] R. Crisafully, M.A.L. Milhome, R.M. Cavalcante, E.R. Silveira, D. De Keukeleire,
dated by FTIR and SS-NMR spectroscopy. Medium-weak and coop- R.F. Nascimento, Biores. Technol. 99 (2008) 4515–4519.
erative H-bonds and van der Waals forces occur between zeolite [10] T.H. Dunning, J. Chem. Phys. 90 (1989) 1007–1023.
[11] M.J. Frisch, et al., Gaussian 03, Revision C.02, Gaussian Inc., Wallingford, CT,
porosity walls and the antibiotic: water seems to play a key role 2004.
in the stabilization of the host–guest high silica zeolite-SMX inter- [12] S. Fukahori, T. Fujiwara, R. Ito, N. Funamizu, Desalination 275 (2011) 237–242.
actions only for zeolite Y and mordenite. [13] H.J. Gill, J.L. Maggs, S. Madden, M. Pirmohamed, B.K. Park, Brit. J. Clin. Pharm.
42 (1996) 347–353.
In conclusion, favorable adsorption kinetics along with effective [14] P.C. Hariharan, J.A. Pople, Theor. Chim. Acta 28 (1973) 213–222.
and highly irreversible adsorption for SMX in zeolite Y pores make [15] M. Kahle, C. Stamm, Environ. Sci. Technol. 41 (2007) 132–138.
this environmental friendly material interesting for sulfa drugs re- [16] M. Kahle, C. Stamm, Chemosphere 68 (2007) 1224–1231.
[17] A.C. Larson, R.B. von Dreele, Los Alamos Natl. Lab. Rep. LAUR (2000) 86–748.
moval from water bodies, whereas the positive temperature effect [18] J.L. Martinez, Environ. Pollut. 157 (2009) 2893–2902.
revealed by ZSM-5 towards the antibiotic adsorption makes this [19] A. Martucci, L. Pasti, N. Marchetti, A. Cavazzini, F. Dondi, A. Alberti,
sorbent interesting for manure treatments also at the high temper- Microporous Mesoporous Mater. 148 (2011) 1742–1783.
[20] A. Martucci, M.A. Cremonini, S. Blasioli, L. Gigli, G. Gatti, L. Marchese, I. Braschi,
atures reached during thermophilic phase. Based on the sorption, Microporous Mesoporous Mater. 170 (2013) 274–286.
structural and spectroscopic data presented, zeolite Y showed the [21] B.F. Mentzen, Mater. Res. Bull. 27 (1992) 831–838.
highest affinity for SMX and can be thus considered optimal for [22] K. Nishi, A. Hidaka, Y. Yokomori, Acta Cryst. B 61 (2005) 160–163.
[23] C.F. McDonald, J. Parenteral Sci. Technol. 45 (3) (1991) 147–151.
sulfonamide removal from impacted waters.
[24] L. Pasti, A. Martucci, M. Nassi, A. Cavazzini, A. Alberti, R. Bagatin, Microporous
Mesoporous Mater. 160 (2012) 182–193.
Acknowledgments [25] L. Pasti, E. Sarti, A. Cavazzini, N. Marchetti, F. Dondi, A. Martucci, J. Sep. Sci. 00
(2013) 1–8.
[26] C.P. Price, A.L. Grzesiak, A.J. Matzger, J. Am. Chem. Soc. 127 (2005) 5512–5517.
Research co-funded by the Italian Ministry of Education, Uni- [27] P. Srinivasan, A.K. Sarmah, M. Manley-Harris, A.L. Wilkins, Sorption of
versity, and Research (Project: Zeolites as nano-reactors for the sulfamethoxazole, sulfachloropyridazine and sulfamethazine onto six New
environment: efficiency, selectivity and stability in the adsorption Zealand dairy farm soils, in: 2010, 19th World Congress of Soil Science, Soil
Solutions for a Changing World 1–6 August 2010, Brisbane, Australia.
of drugs from contaminated waters). G.S. Premachandra is [28] M. Takasuka, H. Nakai, Vib. Spectrosc. 25 (2) (2001) 197–204.
acknowledged for laboratory assistance. [29] B.H. Toby, J. Appl. Cryst. 34 (2001) 210–213.
[30] A.J. van der Ven, M.A. Mantel, T.B. Vree, P.P. Koopmans, J.W. van der Meer, Brit.
J. Clin. Pharm. 38 (1994) 147–150.
Appendix A. Supplementary material [31] A.J. van der Ven, T.B. Vree, E.W. van Ewijk-Benken Kolmer, P.P. Koopmans, J.W.
van der Meer, Brit. J. Clin. Pharm. 39 (1995) 621–625.
[32] I. Braschi, G. Gatti, C. Bisio, G. Berlier, V. Sacchetto, M. Cossi, L. Marchese, Phys.
Supplementary data associated with this article can be found, in
Chem. C 116 (2012) 6943–6952.
the online version, at http://dx.doi.org/10.1016/j.jcis.2013.12.039. [33] J. Gao, F.-Q. Yu, L.-P. Luo, J.-Z. He, R.-G. Hou, H.-Q. Zhang, S.-M. Li, J.-L. Su, B.
Han, Veterinary J. 194 (3) (2012) 423–424.
References [34] V. Homem, L. Santos, J. Environ. Manag. 92 (10) (2011) 2304–2347.
[35] M. Martucci, A. Alberti, L. Pasti, M. Nassi, R. Bagatin, J. Solid State Chem. 194
(2012) 135–142.
[1] R. Arletti, A. Martucci, A. Alberti, L. Pasti, M. Nassi, R. Bagatin, J. Solid State [36] I. Michael, L. Rizzo, C.S. McArdell, C.M. Manaia, C. Merlin, T. Schwartz, C. Dagot,
Chem. 194 (2012) 135–142. Water Res. 47 (3) (2013) 957–995.
[2] Ch. Baerlocher, L.B. McCusker, D.H. Olson, Atlas of Zeolite Framework Types, [37] A. Rossner, S.A. Snyder, D.R.U. Knappe, Water Research 43 (15) (2009) 3787–
6th revised ed., Elsevier, Amsterdam, 2007. 3796.
[3] A.D. Becke, J. Chem. Phys. 98 (7) (1993) 5648–5652. [38] T.A. Ternes, Water Res. 32 (1998) 3245–3260.
[4] I. Braschi, S. Blasioli, L. Gigli, C.E. Gessa, A. Alberti, A. Martucci, J. Hazard. Mater. [39] A.G. Trovò, R.F.P. Nogueira, A. Aguera, C. Sirtori, A.R. Fernandez-Alba,
17 (2010) 218–225. Chemosphere 77 (10) (2009) 1292–1298.
[5] I. Braschi, G. Gatti, G. Paul, C.E. Gessa, M. Cossi, L. Marchese, Langmuir 26
(2010) 9524–9532.

You might also like