You are on page 1of 7

Chemosphere 72 (2008) 1448–1454

Contents lists available at ScienceDirect

Chemosphere
journal homepage: www.elsevier.com/locate/chemosphere

Re-evaluation of the conformational structure of sulfadiazine species using


NMR and ab initio DFT studies and its implication on sorption and degradation
Gerd Huschek a,c,*, Dirk Hollmann b,*, Nadine Kurowski a, Martin Kaupenjohann a, Harry Vereecken c
a
Institute of Ecology, Technische Universität Berlin, Salzufer 12, D-10587 Berlin, Germany
b
Leibniz-Institut für Katalyse e.V. an der Universität Rostock, Albert-Einstein-Strasse 29a, D-18059 Rostock, Mecklenburg-Vorpommern, Germany
c
Agrosphere Institute, ICG-4, Forschungszentrum Jülich GmbH, D-52425 Jülich, Germany

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

Article history: In the environment, the sorption and the degradation of organic pollutants are of increasing interest.
Received 3 January 2008 The investigation of the chemical structures provides a basis for the development of a suitable binding
Received in revised form 7 May 2008 model approach and for the mechanistic understanding of the chemical fate processes. The aim of this
Accepted 12 May 2008
study was the identification of different species of the antibiotic compound sulfadiazine (SDZ) using
Available online 3 July 2008 1
H and 13C NMR experiments and ab initio density functional theory (DFT) calculations. In the neutral,
aprotic solvent dimethylsulfoxide-d6 (DMSO-d6), a new sulfadiazine structure containing an O–H–N
Keywords:
hydrogen bond was identified. In the protic solvent water-d2 and in dependence on pH and the position
Sulfadiazine
NMR
of the amidogen hydrogen atom nine possible SDZ conformations were analyzed and five structures were
Structure elucidation identified. Good conformity between theory and calculation of 1H NMR was observed. Unfortunately, 13C
Antibiotic NMR is not sensitive enough for comparison and differentiation. In order to verify the identified struc-
Sulfonamide tures, additional NBO/NLMO (natural localized molecular orbital) analyses were conducted (calculation
of net atomic charges, bond polarity, atomic valence, and electron delocalization). Finally, conformation
optimizations were performed in order to investigate the stability of the SDZ species.
We showed that SDZ contains no S@O double bond, but that it has two S–O single bonds. Surprisingly,
negative charges were observed at the pyrimidine nitrogen atom. With these results, the known structure
of SDZ was revised. Studies of the geometrical structure and the torsion angles showed that SDZ is very
flexible and can be easily fitted to the sorbent.
These observations would explain the strong sorbance and hence the rapid formation of non-extract-
able residues in the environment because SDZ acts as a strong ligand. These results show that that the
sulfonamide hydrogen is important for the biological activity but the pyrimidine nitrogen and the sulfon-
amide oxygen is responsible for the sorbance in environment.
Ó 2008 Elsevier Ltd. All rights reserved.

1. Introduction benzenesulfonamide), is widely used for specific therapeutic and


prophylactic purposes (Mitema et al., 2001). Unaltered SDZ and
Pharmaceutical antibiotics are very important as veterinary its conjugates are excreted and enter the environmental compart-
drugs. Over the last decade, the extensive use of these chemical ments (Lindberg et al., 2005; Sukul and Spiteller, 2006). In the
substances in animal husbandry has caused considerable environ- environment, the rapid formation of non-extractable residues of
mental contamination (Thiele-Bruhn, 2003; Boxall et al., 2004). SDZ has been identified as the predominant process in manured
Most antibiotics are quickly excreted from the body, either unal- soil (Heise et al., 2006; Thiele-Bruhn and Peters, 2007). Up to
tered or as metabolites (Halling-Sørensen et al., 1998; Kumar 95% of 14C-labeled SDZ was found to be transferred into the non-
et al., 2005). In the environment, these substances can cause extractable fractions (Kreuzig and Höltge, 2005). Hence, intensive
microbial resistance of soil microorganisms (Thiele-Bruhn, 2005). investigations on the mechanism why and how SDZ or other anti-
For instance Sulfadiazine (SDZ, 4-amino-N-(pyrimidin-2yl)- biotics are bonded and sorbed to the environment are of high
interest. Initial investigations on the sorption of sulfonamides in
soil have already be published by Thiele-Bruhn et al. (2004), Gao
* Corresponding authors. Address: IGV GmbH, Arthur-Scheunert-Allee 40/41, and Pedersen, (2005) and Kahle and Stamm (2007).
14558 Nuthetal, OT Bergholz-Rehbrücke, Germany. Tel.: +49 3320089 131; fax: +49 Sulfonamides act as competitive inhibitors of the enzyme dihy-
3320089 251 (G. Huschek), tel.: +49 3811281 215; fax: +49 3811281 51215
(D. Hollmann).
dropteroate synthase (DHPS) in bacteria, and catalyzes the conver-
E-mail addresses: g_huschek@igv-gmbh.de (G. Huschek), dirk.hollmann@ sion of 4-aminobenzoic acid (PABA) into an essential nutrient for
catalysis.de (D. Hollmann). some bacteria (Roland et al., 1979). The sulfonamide group

0045-6535/$ - see front matter Ó 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.chemosphere.2008.05.038
G. Huschek et al. / Chemosphere 72 (2008) 1448–1454 1449

(-SO2NHR1) in SDZ is very important for biological activity. Li et al. provide first insights into the still poorly understood chemical
(2005) pointed out that the antimicrobial activity disappears, if the bonding of SDZ complexes to soil.
hydrogen atom on the dissociate amidogen is replaced by other In brief, the objective of this study is to investigate the chemical
atoms or groups. Hence, sulfadiazine is an excellent representative structures of SDZ species using 1H and 13C NMR experiments in
for a biomedical active substance where the position of the hydro- dependence on pH and comparison with the calculated chemical
gen atom is essential for the binding mechanism and the antimi- shifts of optimized structures in solvents. To verify these struc-
crobial effects. Relationships between antimicrobial activities of tures, further NBO/NLMO (Natural Localized Molecular Orbital)
sulfonamides and the physico-chemical property pKa highlight analyses were carried out for example calculation of net atomic
the dominant role played of the degree of ionization in the activity charges, bond polarity, atomic valence, and electron delocalization.
(Mengelers et al., 1997). In principle, sulfonamide (SDZ) can be Finally, to investigate the stability of the SDZ species, conformation
present as cationic, zwitterionic, uncharged or anionic species. optimizations were performed.
Based on the two pKa values of 1.8 (protonated form) and 6.4
(deprotonated form), uncharged SDZ exists from approximately
pH 4 until pH 8. At pH 7, the neutral and the anionic form are pres- 2. Methods and materials
ent in almost equal proportions in the environment. Consequently,
these species of SDZ are most important for bonding in the 2.1. Sample preparation
environment.
For SDZ, depending on the pH and position of the dissociate Due to the zwitterion structure, SDZ can act as a weak organic
amidogen hydrogen atom, nine possible SDZ conformations can acid (Nesbitt and Sandmann, 1977) and as a weak Lewis base
be described for the neutral (N), deprotonated (D), and protonated (LB). Hence, two different pKa values exist. For the characterisation
(P) forms. These relevant structures are sketched in Fig. 1: P-1: and identification of the proposed resulting structures of SDZ
+
NHþ 3 cation; P-2: –OH S cation; P-3: O–H–Npyrimidine –NHþ 3 (Fig. 1), sample solutions were prepared. Sulfadiazine (20 mg)
cation; P-4: O–H–N–pyrimidine S+ cation; N-1: uncharged; N-2: and 10 ll acetone were added to 2 ml deuterated water. To adjust
tautomeric imido O–H; N-3: zwitterionic; N-4: O–H–N-amidogen; the samples to five different pH values, 20 ll of 2 N hydrogen chlo-
N-5: tautomeric imido O–H–N-pyrimidine; D: anionic. In past, ride solution (pH 1.35), 20 ll of 2 N hydrogen chloride solution and
attention was mainly devoted to the N-1 structure. However, from 30 mg ammonium chloride (pH 2.90), 50 mg ammonium chloride
a chemical point of view, all other structures are possible. (pH 4.40), no additives (pH 5.30), 20 ll of a 15 N sodium hydroxide
For the development of binding approaches for SDZ in the envi- solution (pH 12.00) were added. Due to the low solubility of SDZ in
ronment, the identification of the active SDZ structures present is water (20 mg l1), the excess SDZ was separated by filtration.
important. Neither experimental nor theoretical data is available After 10 min ultrasonication and pH controlling, the suspension
for the identification of water-solved SDZ species. Theoretically, was filtered into an NMR tube.
NMR spectroscopy could be useful for identifying chemical struc-
tures. Benoit and Preston (2000) used this method to identify atra- 2.2. NMR
zine structures in soil. Theoretical ab initio studies could
1
supplement these measurements. Additionally, calculations of H and 13C NMR spectra were recorded on a Bruker AV 300 MHz
energetics, atomic charges, minimum energy structures, geometry, spectrometer. 13C NMR measurements were performed using stan-
and natural bond orbital (NBO) could indicate the electronic den- dard procedure (30° pulse, no signal saturation). 1H and 13C NMR
sity distribution of each atom. Finally, by taking NBO results show- chemical shifts were reported relative to added acetone
ing the presence of S–O single bonds in consideration, realistic (2.22 ppm (1H) and 30.9 ppm (13C) in water-d2, 2.09 ppm (1H),
Lewis structures can be determined (Stefan and Janoschek, 2000). and 30.56 ppm (13C) in acetone-d6, and DMSO-d6 (1H) (Gottlieb
These systematic data, regarding the variation of molecular prop- et al., 1997). Deutero solvents, NaSDZ, and SDZ were purchased
erties, are important for the chemical structure and could therefore from Sigma–Aldrich and used without further purification.

N N N N
O N O N O N O N
H2N S NH H2N S N H3N+ S N- H2N S N
O OH O O H

N-1 N-2 N-3 N-4

H N N N H N
O N O N O N O N
H2N S N H3N+ S NH H2N S+ NH H3N+ S N
O O OH O

N-5 P-1 P-2 P-3

H N N N
O N O N O- N
H2N S+ NH H2N S N H2N S2+ N
O O O-

P-4 D L-1

Fig. 1. Protonated (P), neutral (N), deprotonated (D) and Lewis (L) structures of sulfadiazine studied in this paper.
1450 G. Huschek et al. / Chemosphere 72 (2008) 1448–1454

Due to the solvents and concentration dependence, the NH-shifts were assigned with the off-resonance and selective proton decou-
were not representative. The aniline protons form an AA’BB’-spin pling technique. Within the margin of error, the calculated 13C
system. The assignment of the 1H and 13C signals were assigned shifts of N-5 conformed well to the experimental 13C shifts. Only
HMQC NMR. All NMR spectra are found in the Supporting informa- the calculated C7 shift (172.0 ppm) showed a slight variation of
tion. The pH was measured with a pH-meter WT pH 330 and a pH- 14 ppm compare to the experimental C7 shift (157.1 ppm), which
electrode SenTix 41 pH 0–14 at 22 °C. close to the maximum error of the GIAO method. A combination
of the 1H and 13C NMR showed that N-5 exists in DMSO.
2.3. Theoretical section
3.2. 1H and 13
C shifts in water at different pH values
In order to facilitate the interpretation of the NMR spectra,
density functional type quantum-chemical calculations were Based on the observations in DMSO we focused our attention on
performed using the Gaussian 03 software (Frisch et al., 2004). the identification of chemical shifts for the proposed SDZ struc-
For the calculations, the B3LYP/6-311G+(d) basis set was tures (Fig. 1) in water in dependence on the pH. A buffer was re-
employed, and the default convergence criteria of the G03 were quired to obtain the corresponding species of SDZ. The buffer and
adopted. This set-up provided good results for hydrogen bonding the low solubility of SDZ in water at pH < 6 proved very problem-
modeling (Musin and Mariam, 2006). X-ray crystallography data atic for 13C NMR experiments. Hence, we could only observe 13C
was used as the start configuration for SDZ (Sin et al., 1974; Heren signals for pH 1.35 and pH 12.00 (high solubility as anion and cat-
et al., 2006). The 1H and 13C NMR chemical shifts of the different ion). In the other cases (pH 2.90, 4.40, 5.30), no 13C signals could be
SDZ species were calculated using the GIAO method with water determined even after scanning for several days.
and DMSO as solvents in order to allow this data be compared with In comparison to DMSO, the high intermolecular rate of
the experimental NMR data. The elucidation of the SDZ species was exchange in water resulted in the disappearance of OH and NH
accompanied by calculation of net atomic charges, bond polarity, protons. Only the CH protons of the aniline and the pyrimidine ring
atomic valence, and electron delocalization using NBO/NLMO were determined (Table 2). Moreover, the solvent effect of DMSO
(natural localized molecular orbital) analysis (Glendening et al., and water ruled out a direct correlation between DMSO and water.
1988). A compendium of the NMR spectra at different pH values has been
The electronic density distribution in the SDZ molecule as a included in the Supplementary information Fig. 1.
consequence of intra- and interhydrogen bonding is essential for pH 1.35: In contrast to the other pH values, a very high shift of
binding model approaches to SDZ in the environment. Solvents the aniline protons was measured at pH 1.35. The protons H18/21
could influence the order of preference of different structures and H19/20 were observed at 8.07 ppm and 7.51 ppm, respectively.
(see Fig. 1), which is why conformations of the SDZ species were Compared to the 1H shifts at pH 2.90, the difference was +0.30 ppm
optimized in the gas phase, in water and in dimethylsulfoxide. and +0.64 ppm. This is in accordance with decreased magnetic
The calculations were performed using the integral equation for- shielding of the aniline protons by an ammonium group (R-NHþ 3 ).
malism polarized continuum model (IEFPCM) (Cancès et al., 1997 This ammonium group militates in favour of the postulated struc-
and Cancès and Mennucci, 2001). The optimized gas phase struc- tures P-1 and P-3. To support these structures P-1 and P-3, we per-
tures were used as starting point for these calculations. formed ab initio calculations of the 1H shifts (Supplementary
information Table 2). For both structures P-1 and P-3, the hydro-
3. Results and discussion gen shifts for H18/21 and H22/24 agreed with the calculated shifts.
A differentiation of P-1 and P-3 was determined for the H19/20
3.1. Structure identification using 1H and 13
C NMR measurements and H23 protons. For P-1 the calculated shifts were 7.8 ppm (H19/
20) and 6.9 ppm (H23). For P-3 the calculated shifts were 8.0/
Until now, no NMR studies have been performed at different pH 7.7 ppm (H19/20) and 6.6 ppm (H23). The best agreement with the
levels in deutero water for the identification of the structure of SDZ experimental data (7.51 ppm for H19/20 and 7.05 ppm for H23)
species. We began our investigations with the analysis of the neu- was confirmed for P-1. In addition, due to the P@N double bond,
tral form of SDZ. Due to the slow intermolecular rate of exchange, a high rotation barrier in P-3 was assumed (Jordan, 1982). This
the polar and aprotic solvent dimethylsulfoxide (d6-DMSO) was was identified in the calculations of the H19 and H20 shifts, which
chosen (Table 1 and Fig. 2, for the atom-labeling scheme of SDZ resulted in two different shifts. This suggests that one clear confor-
see Fig. 3). mation exists. However, in the experimental spectra only one
The aniline hydrogen atoms (H25/H26) were observed at signal for H19/20 appeared. This observation favours the P-1 form.
6.01 ppm. Surprisingly, a broad signal appeared at 11.28 ppm. This Additionally, we compared the experimental 13C shifts (pH
high shift did not correspond with the sulfonamide structure N-1 1.35) with the calculated 13C shifts of P-1 and P-3 (see Supplemen-
nor the imido tautomeric forms N-2. The NH proton of N-1 would tary information Table 1). Unfortunately, if the margin of error is
be observed around 7 ppm and the acidic OH proton of N-2 at taken into account, it became clear that no clear differentiation is
10 ppm (Pretsch et al., 2004). Shift over 11 ppm are well known possible for P-1 and P-3 using 13C measurements. The results of
from hydrogen bonds which are formed in 1,3-diketons by further calculations (inclusion of important additional water mol-
keto-enol tautomerism. When the hydrogen bonds in SDZ were ecules) were no better than those of the 13C NMR calculations. In
also taken into account, we assumed a new structure (N-5) was summary we propose that the protonated form P-1 exists at pH
formed which bears a hydrogen between the oxygen of the sulfon- 1.35 which agrees with the known knowledge.
amide and the nitrogen of the pyrimidine ring. pH 2.90–5.30: The comparison of the proton spectra at pH 2.90,
In order to determine the present form, ab initio calculations 4.40, 5.30, and 12.00 revealed that a gradual shielding of all pro-
were performed (Table 1). Calculations of the proton H27 gave tons occurred. The position of the protons H18/21 and H19/20 chan-
the following results: 6.7 ppm for N-2 and 11.5 ppm for N-5. More- ged only slightly (7.7 ppm and 6.8 ppm) for all pH. This
over, other calculated shifts of N-5 showed good agreement with means that no changes occurred at the aniline ring. In contrast,
the measured proton shifts. proton H22/24 was found to be stabile up to pH 4.40 (8.46 ppm),
In addition, we examined the 13C-spectra of SDZ in DMSO-d6 but at higher pH values the shift decreased to 8.28 ppm. A similar
(Table 1). The 13C signals correspond exactly with the 13C signals observation was also made for H23. Unfortunately, because the buf-
reported by Chang et al. (1975). In this group the chemical shifts fer signal of ammonium chloride overlapped with the proton signal
Table 1
Comparison of experimental and calculated 1H and 13
C NMR shifts of neutral and deprotonated species in DMSO-d6a
Species Experimental Calculated
SDZ Na-SDZ N-2 N-5 D-1 (Na)
H number H shift ppm H shift ppm H shift ppm H shift ppm H shift ppm
300 MHz DMSO 300 MHz DMSO (ref. TMS = 0) GIAO, (ref. TMS = 0) GIAO, (ref. TMS = 0) GIAO, DMSO
DMSO DMSO
H18/21 7.61 7.45 7.5/7.7 7.6 7.5/7.9
H19/20 6.56 6.44 6.6/6.8 6.7/6.7 6.6/6.6
H22/24 8.48 8.08 8.2/8.4 8.3/8.6 8.2/8.2
H23 7.00 6.34 6.6 6.7 6.1
H25/26 6.01 5.33 4.4/4.4 4.1 3.4/3.4
H27 11.27 – 6.7 11.5 –

Species Experimental Calculated


SDZ Na-SDZ N-5 D-1 (Na)
C number C shift ppm C shift ppm C shift ppm C shift ppm
75.5 MHz DMSO 75.5 MHz DMSO (ref. TMS = 0) GIAO, DMSO (ref. TMS = 0) GIAO, DMSO
C1 124.7 133.6 131.1 146.4/139.5

G. Huschek et al. / Chemosphere 72 (2008) 1448–1454


C2/C6 129.7 128.2 133.8/134.1 131.1/136.4
C3/C5 112.0 111.8 115.8/117.7 116.2/115.0
C4 152.9 149.7 160.2 157.6
C7 157.1 164.5 172.0 171.3
C8/C10 158.1 157.1 165.9/160.6 160.9/162.9
C9 115.4 108.8 116.5 111.2
a
300.1 MHz for 1H and 75.5 MHz for 13C NMR, d in ppm, experimental shifts are determined with acetone as internal standard. Calculated 1H NMR shifts using ab initio calculations (B3LYP/6-311G+(d) basis set) in DMSO against
TMS as reference. For 13C calculations margin of error ±6 ppm.

Table 2
Comparison of experimental and calculated 1H NMR shifts of Sulfadiazine in D2O at different pHa

Species N-1 N-2 N-3 N-5 P-1 D-1


H H shift ppm H shift ppm H shift ppm H shift ppm H shift ppm H shift ppm (ref. H shift ppm (ref. H shift ppm (ref. H shift ppm (ref. H shift ppm (ref. H shift ppm (ref.
number pH 1.35 D20 pH 2.90 D20 pH 4.40 D20 pH 5.30 D20 pH 12.00 D20 TMS = 0) GIAO, TMS = 0) GIAO, TMS = 0) GIAO, TMS = 0) GIAO, TMS = 0) GIAO, TMS = 0) GIAO,
water water water water water water
H18/21 8.07 7.77 7.75 7.70 7.67 8.0/8.2 7.9/8.1 7.9/8.1 7.7/7.7 8.3/8.3 7.9/8.1
H19/20 7.51 6.87 6.83 6.82 6.84 7.1/7.1 7.0/7.2 7.6/7.6 6.8/6.8 7.8/7.8 7.0/7.0
H22/24 8.47 8.46 8.46 8.36 8.28 8.7/8.8 8.5/8.8 7.9/8.2 8.4/8.7 8.4/8.6 8.4/8.5
H23 7.05 7.07 n.d 6.92 6.79 7.3 7.0 6.2 6.8 6.9 6.7
H25/26 n.d. n.d. n.d. n.d. n.d. 4.8 4.9/4.9 6.1/6.3/ 6.4 4.6/4.6 6.3/6.4 6.5 4.4/4.4
H27 n.d. n.d. n.d. n.d. n.d. 8.4 7.4 (O12–H) – 12.2 (O12–H) 8.0 –
H28 n.d. n.d. n.d. n.d. n.d.
a
300.13 MHz, d in ppm, experimental shifts are determined with acetone as internal standard. Calculated 1H NMR shifts in ppm using ab initio calculations (B3LYP/6-311G+(d) basis set) in water against TMS as reference.

1451
1452 G. Huschek et al. / Chemosphere 72 (2008) 1448–1454

11.5 11.0 10.5 10.0 9.5 9.0 8.5 8.0 7.5 7.0 6.5 6.0 ppm
1.0

2.0

2.0

1.0

2.0

1.9
Fig. 2. 1H NMR spectra of sulfadiazine in DMSO-d6.

Table 3
Comparison of DESE [kJ mol1] for SDZ species of optimized structures using DFT/
B3LYP/ 311G(d) of G03, N-1 (for N-2–N-5) and P-4 (for P-1–P-3) normalized to zeroa

kcal/mol N-1 N-2 N-3 N-5 P-1 P-2 P-3 P-4


Gas phase 0 +117 +284 +79 +26 +23 +29 0
(0) (+1.2) (+2.9) (+0.8) (+1.1) (+1.0) (+1,2) (0)
Water 0 +32 +83 +92 +5 +38 +9 0
(0) (+0.3) (+0.8) (+0.9) (+0.2) (+0.9) (+0.3) (0)
a
Gas phase: 1154,883 kJ mol1; water: 1154,921 kJ mol1; values in paren-
theses are in eV.

By comparing the experimental 13C shifts (pH 12.00) with the


calculated 13C shifts of D a good conformity was observed (see Sup-
plementary information). Only small variations were determined
for the C-1 and C-5 (+8 ppm). To prove the D form, additional 1H
Fig. 3. Atom-labelling scheme of SDZ.
and 13C measurements of sodium sulfadiazine (NaSDZ) were done
in DMSO-d6 (Table 1). In consideration of the margin of error of the
GIAO method, a good conformity of the measured and calculated
of H23, a shift for H23 could not be detected at pH 4.40. However, chemical shifts of D is given. Therefore at pH 12.00, the form D
because of the unchanged chemical shift of H22/24 at pH 2.90 and exists which agrees with the known knowledge, too.
pH 4.40, we assume that the chemical shift of H23 also remained
unchanged. 3.3. Molecular properties and relative energies
A comparison of the chemical shifts at pH 1.35 and pH 2.90 re-
vealed that only the chemical shifts of the aniline changed. This All structures shown in Fig. 1 could be optimized except for
means that the chemical environment of the protons of the pyrim- structure N-4. In addition to G03 calculations, Vienna ab initio
idine ring remained unchanged. This would therefore eliminate simulation package (VASP) calculations were used to confirm the
structure N-3. At pH 1.35 we succeeded in confirming the presence optimum of these conformations (Kresse and Hafner, 1993 and
of form P-1. At pH 2.90 the analogue neutral form N-1 was ob- Kresse and Furthmüller, 1996). The energetic, geometrical optimi-
served. Unfortunately, ab initio calculation of N-1, but also the con- zations showed that the energetic global minimum of the N-1
templable forms N-2 and N-5 showed good conformity with the structure, the pharmaceutically active substance, is the preferred
observed chemical shifts. By comparing all of the calculated neu- energetic form. Hence, in order to compare the energy of the differ-
tral structures, we recognized that the calculated values for N-1, ent structures for neutral species in gas phase or water, respec-
N-2, and N-5 were very close together. This could also be seen in tively, the computed energy of N-1 was normalized to zero, as
the calculated stability of these complexes (see Chapter 3.3 and shown in Table 3. In general, the optimized structures of the SDZ
Table 3). No exact structure could be assigned to the experimental species exhibited planar aniline and pyrimidine rings, and the
data at pH 2.90 and pH 4.40. Under these conditions (water and geometry around the sulfur atom was distorted from the ideal tet-
room temperature) no rotation barrier existed and these forms rahedral symmetry. The gauche conformation about the S–N bond,
were quickly transformed into each other. Therefore we could indicated by the torsional angle C1–S–N–C7, is important for the
not no determination of the exact species could be conducted. energy minimum of the SDZ forms. The most relevant bond lengths
We suppose that between pH 2.90 and pH 5.50 a state of equilib- and bond angles of the optimized structures are listed in Table 5 in
rium exists between N-1 and the imido tautomeric forms N-2 the Supplementary information, together with the experimental
and N-5 which expand the knowledge about SDZ structures. data on the crystal structures of SDZ. Two different dihedral angles
pH 12.00: At pH 12.00 we assume that only one species, D is of 76° (Sin et al., 1974) and 59.6° (Heren et al., 2006) were reported
present. The calculated 1H chemical shifts of D showed excellent between the planes through the benzene ring and the pyrimidine
conformity with the measured chemical shifts. ring for SDZ structures based on X-ray crystallographic data.
G. Huschek et al. / Chemosphere 72 (2008) 1448–1454 1453

Differences of DESE among structures within the gas phase and sulfur bears two positive charges. This agrees with the results for
between the gas phase and water as a solvent arose because of the sulfurous and sulfuric acid molecules (Stefan and Janoschek,
rotation barriers about the S–N bond in the molecule. For water- 2000). Hence, the calculated resonance structures for all forms
solved structures, the rotation about the S–N bond was restricted were strongly delocalized and indicated a strong electron defi-
as a result of the surrounding water molecules. By comparison of ciency on sulfur. Formulas without multiple S@O bonds were
bond lengths, the best agreement was achieved for N-1 and D with therefore favoured according to the valence Lewis net charge. Thus,
the reported experimental crystal data. No experimental data was under conservation of the octet rule on sulfur and the Lewis struc-
available for the cation. tures of SDZ should be described by ionic S–O bonds, structure of L-
Depending on the position of the hydrogen atom of the amido- 1 (Fig. 1), instead of covalent p-bonds. As reported by Stefan and
gen group two tautomeric imido forms N-2 and N-5 (with an addi- Janoschek (2000), the traditionally assumed violation of the octet
tional hydrogen bond of 1.787 Å) between O13H27 and N17 of rule on the sulfur atom of SDZ conflicts with the charge distribu-
pyrimidine) were obtained from our calculations using G03. The tion of sulfur and oxygen and the low total occupancy of 0.16 for
O13H27 bond was confirmed by different start positions of H27. Tau- the natural 3d orbital of sulfur.
tomerization to N-2 and N-5 was achieved by the formal migration Surprisingly, in addition to the negative charge on oxygen, neg-
of the hydrogen atom to the oxygen atom. Imido tautomeric forms ative charges were also calculated for the nitrogen atoms N16 and
of sulfanilamide derivates were described by 13C- NMR measure- N17 of the pyrimidine ring (0.5) and 0.92 for N14 of D. The neg-
ments (Bult, 1983) but no structures were reported for these forms. ative charges are important for the environmental fate of SDZ. On
In solutions where tautomerization was possible, a chemical the basis of the NBO analysis with a great number of lone pair orbi-
equilibrium of N-1 to N-2 could be reached, because the tautomer- tals (LP) and donated electronic charges in antibonding non-Lewis
ization barrier only amounted to 32 kJ mol1 and 92 kJ mol1 for orbitals, it can be hypothesized that neutral and deprotonated
N-5, respectively. It is possible that the mechanism, whereby a forms act as a Lewis base in solution. Lewis bases can form new
high tautomerization barrier maintains the tautomeric transforma- covalent coordinative bonds by donating a pair of electrons.
tion, is far from a state of equilibrium. The exact ratio of the tauto- If this happens, SDZ acts as a strong ligand. The best examples
mer concentration depends on several factors, including have been shown in sulfadiazine complexes with silver (de Wit
temperature, solvent, and pH. N-1, N-2, and N-5 forms were iden- et al., 1983), copper (Hossain et al., 2006), zinc (Yuan et al.,
tified by NMR measurements. 2001), cobalt (Ajibade et al., 2006), mercury (Garcia-Raso et al.,
The protonated form P-4 was the most energetically stable form 1997), platinum (Pasini et al., 1983), and rhodium (Katsaros and
in the gas phase. The DESE of this structure compared to the next Anagnostopoulou, 2002). Furthermore we investigated the sterical
cation form P-2 was about 23 kJ mol1. The deprotonated form D properties of the SDZ molecule, especially the dihedral angle of the
was computed using DFT calculations. The species D of SDZ is pyrimidine ring to the S–N bond, in dependence on the properties
not stable in solution and reacted spontaneously with H3O+ to of the sorbent. We observed a very flexible S–N bond of 53.6 to
N-1 and with other cations such as Ag+, Cu2+, and Ca2+. In these 152.4°. SDZ can easily adopt to heavy metal cations in solution or
drugs, the structure D acts as ligand. to bounded organic matter.
The protonated form P-4 was the most energetically stable form All of these observations can be used to explain the strong sorp-
in the gas phase. This form made it possible to explain, the LC-MS/ tion on soil. We suggest that ion–ion salt bridges and metal com-
MS observations regarding SDZ mass fragments made by Pfeifer plexes are responsible for the non-extractable fractions of SDZ in
et al. (2005) and Lamshöft et al. (2007). Studies on the SDZ struc- the environment.
tures using positive (protonated) LC-ESI/MS/MS measurements
revealed two key mass fragments (m/z = 159 and 92) in the MS/
MS spectrum. Both positive fragment structures had no additional 4. Conclusions
hydrogen atom on the amino group. Pfeifer et al. (2005) and Lam-
1
shöft et al. (2007) indicated a possible modification of the pyrimi- H und 13C NMR measurements combined with ab initio calcu-
dine moiety but they could not explain how these fragments are lations has been shown to be efficient for the analysis of different
formed [starting from the P-1 form]. If we assume the P-4 struc- SDZ species. In dimethylsulfoxide (aprotic polar solvent), a new
ture, the mass fragment m/z = 159 is caused by the [M+H]+ ion from (N-5) of SDZ was identified, which contains a O–H–N hydro-
(m/z = 251) losing the [NH–C4H3N2–H] fragment (m/z = 95), by gen bond. Compared to DMSO, in water (protic solvent), at pH 2–4
means of the cleavage of the S–N bond and the O13–H28N17 bond. an equilibrium of the tautomeric forms N-1, N-2 and N-5 is
The dihedral angle about rotational S–N bond with 145.3° is assumed. This is reflected in the similar stability of these confor-
responsible for the P-4 structure. This angle is stabilized by the H mations. Under acid conditions (pH 1), these forms were converted
bond between O13 and N17–H28 of the pyrimidine ring and will into one stable form, namely P-1. Under basic conditions, the form
be lost in Q2 of MS. The S–N (S–C1) bond is broken by the rotation D was identified. In all solvents, the measured and calculated 1H
energy of the pyrimidine ring. The hydrogen bond bridge O13–H28– chemical shifts conformed very well. Assignment based on calcu-
N17 was confirmed using different start positions of H28. lated 13C signals was restricted. Due to low sensitivity, these shifts
In order to understand a binding approach to sorbat and sorbent provided only limited information on existing forms.
in the environment, it is important to understand the electronic The analysis of the electronic properties of SDZ point out nega-
structure of a molecule. Thus, the calculation of Lewis structures tive charges for oxygen (1), and surprisingly about 0.5 for the
and the characterization of Lewis orbitals and their occupancies nitrogen atoms of the pyrimidine ring. A positive charge of +2
were performed by NBO/NLMO (natural bond/natural localized was calculated for sulfur means that two S–O single bonds exist.
molecular orbital) analysis. Nine SDZ species were investigated in Due to these properties we predict that SDZ can act as a strong li-
terms of net atomic charge, bond polarity, and electron delocaliza- gand and Lewis base. In addition we observed a very flexible rota-
tion (see Supplementary information Table 4). tion of the S–N bond, which allowed SDZ to easily adopt to heavy
The collected results of the natural population analysis (NPA) metal cations or to bonded organic matter. This would explain the
(see Supplementary information Table 3) indicates a remarkably rapid formation of non-extractable residues and the strong sor-
high positive natural net atomic charge of about 2.2 on sulfur bance of SDZ in the environment.
and of about 0.8 to 1.0 on both oxygen atoms for all SDZ forms. We want to emphasize that the sulfonamide nitrogen is impor-
This means that both oxygen atoms bear one negative charge and tant for the biological activity but the pyrimidine hydrogen and the
1454 G. Huschek et al. / Chemosphere 72 (2008) 1448–1454

sulfonamide oxygen is responsible for the sorbance in environ- Kahle, M., Stamm, C., 2007. Sorption of the veterinary antimicrobial sulfathiazole to
organic materials of different origin. Environ. Sci. Technol. 41, 132–138.
ment. Future research will focus on the revision of analogous sul-
Katsaros, N., Anagnostopoulou, A., 2002. Rhodium and its compounds as potential
fonamide structures. agents in cancer treatment. Crit. Rev. Oncol. Hemat. 42, 297–308.
Kresse, G., Furthmüller, J., 1996. Efficient iterative schemes for ab initio total-energy
Acknowledgements calculations using a planewave basis set. Comput. Mater. Sci. 6, 15–50.
Kresse, G., Hafner, J., 1993. Ab initio molecular dynamics for open-shell transition
metals. Phys. Rev. B 48, 13115–13118.
The authors gratefully acknowledge the financial support pro- Kreuzig, R., Höltge, S., 2005. Investigations on the fate of sulfadiazine in manured
vided by the German Research Foundation (DFG – Deutsche Fors- soil: laboratory experiments and test plot studies. Environ. Toxicol. Chem. 24,
771–776.
chungsgemeinschaft) for Research Unit 566 Veterinary Medicines Kumar, K., Gupta, S.C., Chander, Y., Singh, A.K., 2005. Antibiotic use in agriculture
in Soil. We would also like to thank Dr. A. Koch for his excellent and its impact on the terrestrial environment. Adv. Agron. 87, 1–54.
technical and analytical support. Lamshöft, M., Sukul, P., Zühlke, S., Spiteller, M., 2007. Metabolism of 14C-labelled
and non labelled Sulfadiazine after administration to pigs. Anal. Bioanal. Chem.
388, 1733–1745.
Appendix A. Supplementary material Li, X.H., Cheng, X.L., Zhang, R.Z., Yang, X.D., 2005. Quantum chemical study on
structure-activity relationship of several kinds of drugs. Chinese J. Struct. Chem.
5, 513–520.
Supplementary data associated with this article can be found, in Lindberg, R.H., Wennberg, P., Johansson, M.I., Tysklind, M., Andersson, B.A.V., 2005.
the online version, at doi:10.1016/j.chemosphere.2008.05.038. Behavior of fluoroquinolones and trimethoprim during mechanical, chemical
and active sludge treatment of sewage water and digestion of sludge. Environ.
Sci. Technol. 39, 3421–3429.
References Mengelers, M.J.B., Hougee, P.E., Janssen, L.H.M., Van Miert, A.S.J.P.A.M., 1997.
Structure-activity relationships between antibacterial activities and
Ajibade, P.A., Kolawole, G.A., O’Brien, P., Helliwell, M., Raftery, J., 2006. Cobalt(II) physicochemical properties of sulfonamides. J. Vet. Pharmacol. Therap. 20,
complexes of the antibiotic sulfadiazine, the X-ray single crystal structure of 276–283.
[Co(C10H9N4O2S)2 (CH3OH)2]. Inorg. Chim. Acta 359, 3111–3116. Mitema, E.S., Kikuvi, G.M., Wegener, H.C., Stohr, K., 2001. An assessment of
Benoit, P., Preston, C.M., 2000. Transformation and binding of 13C and 14C-labeled antimicrobial consumption in food producing animals in Kenya. J. Vet.
atrazine in relation to straw decomposition in soil. Eur. J. Soil Sci. 51, 43–54. Pharmacol. Therap. 24, 385–390.
Boxall, A.B., Fogg, L.A., Blackwell, P.A., Kay, P., Pemberton, E.J., Croxford, A., 2004. Musin, R.N., Mariam, Y.H., 2006. An integrated approach to the study of
Veterinary medicines in the environment. Rev. Environ. Contam. Toxicol. 180, intramolecular hydrogen bonds in malonaldehyde enol derivatives and
1–91. naphthazarin: trend in energetic versus geometrical consequences. J. Phys.
Bult, A., 1983. Distinction between the tautomeric forms of sulphanilamide Org. Chem. 19, 425–444.
derivates by 13C NMR. Pharm. Weekblad 5, 77–78. Nesbitt, R.U., Sandmann, B.J., 1977. Solubility studies of silver sulfadiazine. J. Pharm.
Cancès, E., Mennucci, B., 2001. The escaped charge problem in solvation continuum Sci. 66, 519–522.
models. J. Chem. Phys. 115, 6130–6135. Pasini, A., Bersanetti, E., Zunino, F., Savi, G., 1983. Antitumor complexes of platinum
Cancès, E., Mennucci, B., Tomasi, J., 1997. A new integral equation formalism for the with carrier molecules. I. Sulfadiazine derivatives of platinum(II). Inorg. Chim.
polarizable continuum model: theoretical background and applications to Acta 80, 99–102.
isotropic and anisotropic dielectrics. J. Phys. Chem. 107, 3032–3041. Pfeifer, T., Tuerk, J., Fuchs, R., 2005. Structural characterization of sulfadiazine
Chang, C., Floss, H.G., Peck, G.E., 1975. Carbon-13 magnetic resonance spectroscopy metabolites using H/D exchange combined with various MS/MS experiments. J.
of drugs Sulfonamides. J. Med. Chem. 18, 505–509. Am. Soc. Mass Spectrum. 16, 1687–1694.
De Wit, P.P., Van Doorne, H., Bult, A., 1983. Synthesis, physical properties and Pretsch, E., Bühlmann, P., Affolder, C., Badertscher, M., 2004. Structure
microbiological activities of more water soluble silver sulfadiazine derivates. Determination of Organic Compounds: Tables of Spectral Data, 4th ed.
Pharm. Weekblad 5, 298–301. Springer-Verlag, Berlin-Heidelberg.
Frisch, M.J. et al. 2004. Gaussian 03, Revision C02, Gaussian, Inc., Wallingford CT. Roland, S., Ferone, R., Harvey, R.J., Styles, V.L., Morrison, R.W., 1979. The
Full reference please see Supporting information. characteristics and significance of sulfonamides as substrates for Escherichia
Gao, J., Pedersen, J.A., 2005. Adsorption of sulfonamide antimicrobial agents to clay coli dihydropteroate synthase. J. Biol. Chem. 254, 10337–10345.
minerals. Environ. Sci. Technol. 39, 9509–9516. Sin, H.S., Ihn, G.S., Kim, H.S., Koo, C.H., 1974. The crystal and molecular structure of
Garcia-Raso, A., Fiol, J.J., Martorell, G., Lopez-Zafra, A., Quiros, M., 1997. Metallation sulfadiazine. J. Korea Chem. Soc. 5, 327–340.
of 2-sulfanilamidopyrimidine(sulfadiazine). X-ray diffraction structure and Stefan, T., Janoschek, R., 2000. How relevant are S@O and P@O double bonds for the
solution behaviour of bis(sulfadiazinato)mercury(II) bis(dimethylsulfoxide). description of the acid molecules H2SO3, H2SO4, and H3PO4, respectively. J. Mol.
Polyhedron 16, 613–621. Model. 6, 282–288.
Glendening, E.D., Reed, A.E., Carpenter, J.E., Weinhold, F., 1988. NBO Version 3.1. Sukul, P., Spiteller, M., 2006. Fluoroquinolone antibiotics in the environment. Rev.
Wisconsin. Environ. Contam. Toxicol. 187, 67–101.
Gottlieb, H.E., Kotlyar, V., Nudelman, A., 1997. NMR chemical shifts of common Thiele-Bruhn, S., 2003. Pharmaceutical antibiotic compounds in soils – a review. J.
laboratory solvents as trace impurities. J. Org. Chem. 62, 7512–7515. Plant Nutr. Soil Sci. 166, 145–167.
Halling-Sørensen, B., Nors Nilsen, S., Lanzky, P.F., Ingerslev, F., Holten-Lützhoft, H.C., Thiele-Bruhn, S., 2005. Microbial inhibition by pharmaceutical antibiotics in
Jørgensen, S.E., 1998. Occurrence, fate, and effects of pharmaceutical substances different soil – dose–response relations determined with iron(III) reduction
in the environment – a review. Chemosphere 36, 357–393. test. Environ. Toxicol. Chem. 24, 869–876.
Heise, J., Höltge, S., Schrader, S., Kreuzig, R., 2006. Chemical and biological Thiele-Bruhn, S., Peters, D., 2007. Photodegradation of pharmaceutical antibiotics
characterization of non-extractable residues in soil. Chemosphere 65, 2352– on soil surfaces. FAL Agric. Res. 57, 13–24.
2357. Thiele-Bruhn, S., Seibicke, T., Schulten, H.-R., Leinweber, P., 2004. Sorption of
Heren, Z., Pasglu, H., Kastas, G., 2006. Butylammonium sulfadiazinate monohydrate. sulfonamide pharmaceutical antibiotics at whole soils and particle-size
Acta Crystallogr. E 62, o3437–o3439. fractions. J. Environ. Qual. 33, 1331–1342.
Hossain, G.M., Banu, A., Amoroso, A.J., 2006. Trans-bis(ethylenediamine) Yuan, R.X., Xiong, R.G., Chen, Z.F., Zhang, P., Ju, H.X., Dai, Z., Guo, Z.J., Fun, H.K., You,
bis(sulfadiazinato)copper(II). Acta Crystallogr. E 62, m2727–m2729. X.Z., 2001. Crystal structure of zinc(II) 2-sulfanilamidopyrimidine: a widely
Jordan, F., 1982. 1H NMR evidence for high barriers to amino group rotation in 4- used topical burn drug. J. Chem. Soc., Dalton Trans., 774–776.
aminopyrimidines, including thiamin, at low pH in water. J. Org. Chem. 47,
2748–2753.

You might also like