You are on page 1of 6

Computational and Theoretical Chemistry 1217 (2022) 113891

Contents lists available at ScienceDirect

Computational and Theoretical Chemistry


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

Quantum chemical calculations of IR spectra of heparin


disaccharide subunits
Yulia B. Monakhova a, b, c, *, Polina M. Soboleva a, Elena S. Fedotova a, Kristina T. Musina a,
Natalia A. Burmistrova a
a
Institute of Chemistry, Saratov State University, Astrakhanskaya Street 83, 410012 Saratov, Russia
b
Spectral Service AG, Emil-Hoffmann-Straße 33, 50996 Cologne, Germany
c
University of Applied Sciences Aachen, Faculty of Chemistry and Biotechnology, Heinrich-Mußmann-Straße 1, 52428 Jülich, Deutschland

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

Keywords: Heparin is a natural polysaccharide, which plays essential role in many biological processes. Alterations in
IR spectroscopy building blocks can modify biological roles of commercial heparin products, due to significant changes in the
Heparin conformation of the polymer chain. The variability structure of heparin leads to difficulty in quality control using
Quality control
different analytical methods, including infrared (IR) spectroscopy. In this paper molecular modelling of heparin
Molecular modelling
disaccharide subunits was performed using quantum chemistry. The structural and spectral parameters of these
Quantum chemistry
Chemometrics disaccharides have been calculated using RHF/6-311G. In addition, over-sulphated chondroitin sulphate disac­
charide was studied as one of the most widespread contaminants of heparin. Calculated IR spectra were analyzed
with respect to specific structure parameters. IR spectroscopic fingerprint was found to be sensitive to substi­
tution pattern of disaccharide subunits. Vibrational assignments of calculated spectra were correlated with
experimental IR spectral bands of native heparin. Chemometrics was used to perform multivariate analysis of
simulated spectral data.

1. Introduction a role in inflammation, cell growth, angiogenesis, differentiation and


invasion [8]. There is increasing evidence of their involvement in a wide
Due to its vitally important properties, such as anticoagulant activity, range of pathologies as disaccharides. Changes in GAG disaccharide
heparin (HEP) has been widely used in clinical practice for many years composition in pancreatic carcinoma or in human chondrodysplasia are
[1]. HEP is a natural polysaccharide that belongs to glycosaminoglycans some of such examples [9–12]. Alterations in sulphation pattern may
(GAG) and has a structural order with up to four levels of complexity well reflect modified biological roles of these molecules, because
[2,3]. The structure of HEP is variable and depends both on the animal different sulphation of the basic disaccharide subunits may significantly
origin and on the tissue it is extracted from. Considering the natural affect the conformation of the polymer chain. On the other hand,
sources of HEP, the numbers of potential contaminants are relatively changes of important physicochemical properties (e.g. concentration,
large compared to synthetic substances [4]. Moreover, HEP can include pH, temperature, and the presence of cations) do not have any signifi­
structurally related contaminants that affect the quality of finished cant effect on the secondary structure [13]. In this regard, reliable
product. It is known, that structural analogues of HEP do not exhibit any methods of structural analysis of HEP in terms of sulphation pattern of
anticoagulant activities typical for HEP caused by a pentasaccharide disaccharide subunits are critically needed.
sequence containing a glucosamine residue sulphated at position 3 [5]. Generally, characterization strategies for heparin can be divided into
In this regard, it has been found in 2007–8 that the presence of over- analysis without any chemical or enzymatic pretreatment (e.g. size
sulphated chondroitin sulphate (OSCS) in pharmaceutical HEP results exclusion chromatography, liquid chromatography-mass spectrometry
in severe adverse health effects [6,7]. (LC-MS), MS-MS or NMR) and analysis based on controlled depoly­
Besides the well-known anticoagulant role of heparin, GAG also play merization of polysaccharide chains (e.g. high-performance liquid

* Corresponding author at: University of Applied Sciences Aachen, Faculty of Chemistry and Biotechnology, Heinrich-Mußmann-Straße 1, 52428 Jülich,
Deutschland.
E-mail address: monakhova@fh-aachen.de (Y.B. Monakhova).

https://doi.org/10.1016/j.comptc.2022.113891
Received 2 June 2022; Received in revised form 13 September 2022; Accepted 21 September 2022
Available online 25 September 2022
2210-271X/© 2022 Elsevier B.V. All rights reserved.
Y.B. Monakhova et al. Computational and Theoretical Chemistry 1217 (2022) 113891

chromatography (HPLC), capillary electrophoresis (CE) and LC-MS 2. Methods and computational details
disaccharide compositional analysis or oligosaccharide sequence anal­
ysis) [14]. Obviously, single analytical methods such as NMR, HPLC-MS Quantum chemical calculations were performed using the Firefly QC
etc. are high-quality analytical techniques, however, researchers package [26], which is partially based on the GAMESS (US) [27] source
encounter some difficulties during screening stage because data coming code using the Saratov State University cluster. The geometry of HEP
from these instruments is very complex and difficult to interpret. and OSCS disaccharide subunits were optimized using the RHF/6-311G.
Simpler alternative as IR spectroscopy also suitable for quality con­ The basis set was chosen according to the required calculation accuracy.
trol of heparin due to chemometrical calculations that improve the ef­ Visualization of obtained results was carried out using the ChemCraft v.
ficiency of instrumental methods [3,15]. Although IR spectrum of HEP is 1.8 graphical software for visualization of quantum chemistry compu­
well-known, the majority of studies describing quality control of HEP tations (https://www.chemcraftprog.com).
did not completely assign all IR bands to corresponding structural fea­ Chemometrics Add-in for Microsoft Excel MS was used for chemo­
tures [2,16,17]. Often vibrational assignment is performed based on metric modelling by PCA as described in [28]. PCA is multivariate
features arising from the higher order structure. Therefore, the differ­ decomposition method, whose idea is to reduce the dimensionality of a
ences in spectral characteristics and reasons for classifying a sample as data set consisting of a large number of interrelated variables, while
pure or contaminated are not always clear. To find structural features retaining as much as possible of the variation present in the data set. This
related to variability of IR spectra it is necessary to compare the spectra is achieved by transforming to a new set of variables (principle com­
to structure of the HEP building blocks. Disaccharide compositional ponents), which are uncorrelated and ordered so that the first few
analysis by IR spectroscopy remains a challenge due to the need the components retain most of the variation in all the original variables
detection of relatively subtle structural features (e.g. sulpho group po­ [29]. The rows of a data matrix X correspond to samples while columns
sition or differences in linkage position), so theoretical study of struc­ correspond to variables. PCA decomposes the data matrix X as the sum
tural and spectral characteristics is of great importance to improve the of the outer product of vectors score t (i.e. contain information regarding
quality of interpretation of spectral data. the interaction of the samples to each other) and loading p (i.e. eigen­
Currently, there are some papers dedicated to theoretical de­ vectors of the covariance matrix) plus a residual matrix E.
scriptions of the major disaccharide sequences of HEP and its structural The PCA model was performed on calculated IR spectral data of HEP
analogs [4]. Quantum chemical calculations have been successfully used disaccharides in 2000–900 cm− 1 range. IR spectral data with Gaussian
for modelling of GAGs fragments and were effective for evaluation of broadening were exported and organized in the form of a matrix for the
their properties. The approaches for simulation of HEP and other chemometric data processing. The order of resulting matrix was 12 ×
sulphated GAGs were based mostly on Hartree–Fock (HF) ab initio and 291.
B3LYP Density Functional Theory (DFT) methods (using Pople basis sets We chose porcine HEP powder as an example of native HEP for
6-31G*, 6–31 + G* and 6–311 + G**) [18]. For example, M. Hricovíni experimental IR spectrum. The sample was supplied by Spectral Ser­
[19] used DFT to make a detailed study of the molecular structure and vice’s quality control laboratory (Cologne, Germany). IR measurements
spin–spin coupling constants of HEP tetrasaccharides representing the were performed on IRAffinity-1S FTIR spectrometer (Shimadzu, Kyōto,
predominant HEP repeating-sequence, while M. Remko et al. [13] used Japan) in the spectral region of 4000–400 cm− 1 using KBr pellet. The
HF/6–31G(d), B3LYP/6–31G(d), and B3LYP/6–311 + G(d, p) methods spectrum was obtained by averaging 20 scans at a resolution of 2 cm− 1.
to study the acidic dimeric structural units and the pentamer of HEP, as For sample preparation, 3 mg of a HEP sample was mixed with 297 mg of
well as their anionic form. The molecular structure of trimeric units and KBr, and sample was evacuated for 15 min and pressed under vacuum on
the pentamer of HEP (including their anionic forms and sodium salts) a laboratory press PGR-400 (Monitoring LLC, Russia).
was also studied by B3LYP/6-31G(d) method [20], while authors of
[21–23] used DFT modelling to explain the sulphate patterns in GAGs 3. Results and discussion
(mono-, di-, and oligosaccharides) and found strong correlations be­
tween spectral and structural characteristics found via cryogenic 3.1. Geometric and energetic parameters of HEP and OSCS disaccharide
infrared spectroscopy and InfraRed Multiple Photon Dissociation subunits
(IRMPD). It should be mentioned that previous works were mainly
aimed directly at investigating the structure of various GAG. At the same The structure of HEP has increasing levels of complexity starting
time, the molecular structure is responsible for the unique vibrational from monosaccharide composition to higher orders of structure [3]. In
spectral characteristics of HEP. Moreover, all levels of a HEP structure this work, disaccharides were chosen as our primary object of structural
can and do affect spectral and activity data [3]. According to our liter­ theoretical analysis. HEP is usually considered a mixture of variously
ature search, spectral characteristics of HEP were previously determined sulphated linear polysaccharides consisting of 1–4 linked uronic acid
only in several studies [19,23,24] that focused on laborious instru­ and D-glucosamine [3,30]; while OSCS consists of 1,3-linked glucuronic
mental techniques. acid and N-acetylgalactosamine residues containing four sulphate
In the present study we were focused on theoretical calculation and groups per building block [3]. The disaccharides composition of the
studying of IR spectral characteristics of twelve HEP and OSCS disac­ underlying structure of HEP is known to be different for various animal
charide subunits with different sulphation. Investigated structures were species. Typically, about 70 % of constituent disaccharides of native HEP
optimized by the restricted Hartree-Fock (RHF) method with the 6-311G are trisulphated [3,4]. The flexibility of the pyranose rings and their
basis set [25]. HOMO and LUMO energies, the ionization energy, the ability to rotate around glycosidic linkage regions result in existence of
absolute hardness and dipole moment were calculated and discussed. In several HEP conformations that depend both on the structure of residue
addition, we thoroughly investigated the normal vibration frequencies and on the sulphation degree [3,31]. In this regard, we optimized the
and intensities of bands in the IR spectra of studied compounds. Vibra­ geometry of twelve disaccharide subunits of HEP and OSCS (Fig. 1,
tional assignments were performed to characterize and clearly describe Fig. S1, Table 1). Using the RHF/6-311G method, the obtained struc­
IR bands. In addition, we performed multivariate analysis of calculated tures did not have any negative values in the Hessian matrix, thus,
IR spectra using principal component analysis (PCA) method that confirming the attribution of optimized structures to corresponding
allowed us to evaluate the inherent similarities and differences. It ap­ conformers. We have observed that the β-glycosidic linkage between
pears that substitution pattern of the HEP disaccharides causes well- monosaccharides has been formed. It was found that the bond angle
defined changes of IR spectral profile. between the sulphated residues of iduronic acid and glucosamine in HEP
is almost independent of the substituents position, while the main
changes in geometry are associated with the values of torsion angles

2
Y.B. Monakhova et al. Computational and Theoretical Chemistry 1217 (2022) 113891

Fig. 1. Repeating disaccharide subunits of HEP (A) and OSCS (B) (R1 = H, SO−3 ; R2 = H, SO−3 ; R3 = H, acetyl (Ac), SO−3 ); optimized structure of HEP disaccharide
6 (C).

Table 1
Substituents of disaccharide subunits of HEP.
1 2 3 4 5 6 7 8 9 10 11 12

R1 H H H H SO3H SO3H SO3H SO3H H H SO3H SO3H


R2 H H SO3H SO3H SO3H SO3H H H H SO3H SO3H H
R3 H SO3H H SO3H H SO3H H SO3H Ac Ac Ac Ac

(Fig. 2, Table S1). 3.2. IR spectral analysis of HEP and OSCS disaccharides subunits
Comparison of HEP disaccharide subunits geometry has shown that
they can be divided into three groups based on the values of the torsion Calculated IR spectra of HEP and OSCS disaccharide subunits are
angles: disaccharides 1, 2, and 9 (group I), disaccharides 3, 4, 5, 6, 10 presented in Fig. 4 and Table S3. The primary IR spectral bands of HEP
and 11 (group II) and disaccharides 7, 8 and 12 (group III). Group I were attributed to absorbance of N-sulfonate groups, N-acetyl groups
contain disaccharides with H atoms as R1 and R2, group II has SO3H and iduronate residues rings [33,34]. IR spectroscopic profiles and as­
group as R2, and group III has SO3H group as R1 and H atoms as R2. R2 signments of absorption bands of native HEP samples have already been
has the strongest influence on the values of torsion angles; the influence described previously [15]. The main difficulty in analysis of IR spectrum
of R3 is insignificant. In addition, when the R2 is changed from -SO3H of native HEP lies in variability of analytical signal due to the structure
group to H atom, the influence of R1 increases. The values of torsion complexity. Besides, IR spectroscopic bands are broad and usually
angles for OSCS and HEP disaccharide subunits is significantly different overlap, so differences in spectral profiles of samples is difficult to
(Fig. 2, Table S1); visual comparison between trisulphated HEP and associate with vibrations of certain groups. Here, we compared IR
OSCS disaccharides is shown in Fig. 3. spectra of HEP and OSCS disaccharide subunits computed via RHF/6-
Calculated parameters, such as HOMO and LUMO energies, the 311G method with the experimental IR spectrum of native HEP
ionization energy, the absolute hardness and dipole moment of the (Fig. 4, Table S3). Before that, both the computed and the corresponding
studied disaccharides are summarized in Table S2. Comparison of experimental spectra were normalized to the intensity of the strongest
electronic characteristics shows that disaccharides 3, 11 and 13 has the band.
least absolute hardness and the largest reactivity; the opposite was Calculated frequencies of HEP disaccharide subunits were found to
observed for disaccharides 2, 4, 6 and 8. HEP disaccharide containing H be quite similar to the frequencies in the experimental spectra of native
atoms as R1 and R2 acts as the best electron donor, while HEP disac­ HEP. The only exception were asymmetrical stretching vibrations of
charide with three -SO3H groups is the worst electron donor. Di­ –COOH group where frequencies differed by more than 200 cm− 1. Di­
saccharides 3, 5, 7 and 13 have the largest values of dipole moment, saccharides forming the backbone of HEP have similar structure, how­
while disaccharides 2 and 12 have the least values of dipole moment. ever, various substitution patterns affect their IR spectral profiles in
Representative UV–vis spectrum of native HEP are shown in Fig. S2 different ways (Fig. 4, Table S3). The band corresponding to –COOH
[32]. asymmetrical stretching at around 1900 cm− 1 is very similar for all
calculated disaccharides. This resemblance in terms of the peak position
and intensity without any significant effect from various substituents is
observed for all studied structures. N-acetylated HEP has changes in the

3
Y.B. Monakhova et al. Computational and Theoretical Chemistry 1217 (2022) 113891

Fig. 2. Torsion angles deviation for calculated HEP and OSCS disaccharide subunits.

two -SO3H groups leads to a decrease in resolution for this spectral


range. The IR spectrum of OSCS disaccharide is more resolved.
Calculated and experimental IR spectra were analyzed via chemo­
metric (PCA) modelling to find any inherent similarities and differences.
All HEP disaccharides can be projected onto the first four principal
components (PCs), which described 92.67 % of the total variance (with
39.30 % described by the PC1, 26.31 % by the PC2, 14.38 % by the PC3,
and 12.68 % by the PC4). The PCA loadings were in accordance with the
main differences in IR spectra (Fig. S3). The PCA model showed that
samples formed four isolated clusters in the PC1-PC2-PC4 space
(Fig. S3). HEP disaccharides from Group I and Group III had the negative
score values along PC2, that can be attributed to their lower absorbance
at 1390 cm− 1, 1270 cm− 1 and 1100–1000 cm− 1. According to the
loadings, the highest contribution to PC4 came from the signals pro­
vided by bands located at 1400–1300 cm− 1 range, 1820 cm− 1, 1500
cm− 1 and 965 cm− 1 (positive) and those at 1237 cm− 1, 1100 cm− 1 and
1420 cm− 1 (negative). HEP disaccharides from Group I and Group II had
Fig. 3. Arrangement of atoms of trisulphated HEP (blue) and OSCS (violet)
higher positive scores along PC4 compared to HEP disaccharides from
disaccharides in a three-dimensional space. (For interpretation of the references
to colour in this figure legend, the reader is referred to the web version of
Group III and Group IV. Moreover, HEP disaccharides from Group I and
this article.) Group II were characterized by both higher intensity at 1400–1300 cm− 1
and lower intensity at around 1237 cm− 1. The opposite was observed for
HEP disaccharides from Group III and Group IV.
IR spectral profile associated with amide group vibrations at around
In summary, the differentiation of HEP disaccharides was observed
1850, 1650, 1570 and 1490 cm− 1. Moreover, changes in substitution
according to their substitution patterns. Each cluster contains di­
pattern lead to changes in spectral profiles associated with non-specific
saccharides with the same R1 and R2 substituents, thus, confirming that
vibrations in the ring and vibrations of the –COH side groups and the
the influence of R3 is insignificant to both torsion angles and IR spectra.
glycosidic bonds. Sulphated disaccharides show more complex IR profile
Native HEP and OSCS disaccharides have been used as test structures to
in the range between 1300 and 1000 cm− 1. In addition, the presence of

4
Y.B. Monakhova et al. Computational and Theoretical Chemistry 1217 (2022) 113891

Fig. 4. IR spectra of native HEP, HEP and OSCS disaccharide subunits plotted with a Gaussian broadening. Group differentiation of HEP disaccharides made in
accordance with PCA. The legend reflects the disaccharide numbers.

evaluate similarity with HEP disaccharides. It was found that native HEP affect the conformation of the β-glycosidic linkage, while O-sulphation
and HEP disaccharides from Group II had similar scores that indicates appears to have a measurable effect on the conformation of the
the similarity of their spectral characteristics. Moreover, Group II con­ β-glycosidic linkage.
tains trisulphated HEP disaccharides that confirms the actual predomi­ In addition, multivariate approach for analysis of calculated IR
nance of these disaccharides in native HEP. OSCS disaccharide has closer spectra allowed to evaluate the inherent similarities and differences in
characteristics with HEP disaccharides from Group IV. Their spectral the set of calculated and experimental spectra. PCA modelling of
similarities seem to make it difficult to detect OSCS in native HEP. calculated IR spectral data differentiated the disaccharides according to
their structures. PCA model also confirmed the plausibility of the
4. Conclusions calculated IR spectra of disaccharides. In this regard, an important
practical result is that theoretical IR spectral profiles combined with
The vibrational infrared spectra of HEP and OSCS disaccharide appropriate multivariate modelling can be used to assess the disaccha­
subunits were investigated using quantum mechanics methods. RHF/6- ride composition of native HEP that may facilitate HEP authentication.
311G level of theory provided good performance for the calculations of Therefore, we are proceeding to further studies of the larger fragments
relative intensities and frequencies for vibrational bands in the range of of HEP that ultimately can provide more detailed information about
1800–1000 cm− 1. Comparing the results of quantum–mechanical HEP.
approach with experimental IR spectrum of native HEP revealed that the
treatment at the RHF/6-311G level of theory offers reliable modelling of CRediT authorship contribution statement
the IR spectra. RHF modelling demonstrated that geometric parameters
as well as the features of the IR spectrum of HEP disaccharides depended Yulia B. Monakhova: Writing – review & editing, Supervision,
on the substitution pattern. We have shown that N-sulphation doesn’t Project administration. Polina M. Soboleva: Formal analysis,

5
Y.B. Monakhova et al. Computational and Theoretical Chemistry 1217 (2022) 113891

Investigation, Writing – original draft, Visualization. Elena S. Fedo­ [14] M. Qiao, L. Lin, K.e. Xia, J. Li, X. Zhang, R.J. Linhardt, Recent advances in
biotechnology for heparin and heparan sulfate analysis, Talanta. 219 (2020)
tova: Formal analysis, Investigation. Kristina T. Musina: Formal
121270, https://doi.org/10.1016/j.talanta.2020.121270.
analysis, Investigation. Natalia A. Burmistrova: Conceptualization, [15] N.A. Burmistrova, P.M. Soboleva, Y.B. Monakhova, Is infrared spectroscopy
Investigation, Writing – original draft, Writing – review & editing. combined with multivariate analysis a promising tool for heparin authentication?
J. Pharm. Biomed. Anal. 194 (2021) 113811, https://doi.org/10.1016/j.
jpba.2020.113811.
Declaration of Competing Interest [16] S. Alban, S. Lühn, S. Schiemann, T. Beyer, J. Norwig, C. Schilling, O. Rädler,
B. Wolf, M. Matz, K. Baumann, U. Holzgrabe, Comparison of established and novel
purity tests for the quality control of heparin by means of a set of 177 heparin
The authors declare that they have no known competing financial samples, Anal. Bioanal. Chem. 399 (2) (2011) 605–620, https://doi.org/10.1007/
interests or personal relationships that could have appeared to influence s00216-010-4169-7.
[17] J. Norwig, T. Beyer, D. Brinz, U. Holzgrabe, M. Diller, D. Manns, Prediction of the
the work reported in this paper. oversulphated chondroitin sulphate contamination of unfractionated heparin by
ATR-IR spectrophotometry, Pharmeur. Sci. Notes. 1 (2009) 17–24. PMID:
Data availability 19275869.
[18] A. Almond, Multiscale modeling of glycosaminoglycan structure and dynamics:
current methods and challenges, Curr. Opin. Struct. Biol. 50 (2018) 58–64, https://
Data will be made available on request. doi.org/10.1016/J.SBI.2017.11.008.
[19] M. Hricovíni, M. Hricovíni, Solution Conformation of Heparin Tetrasaccharide.
DFT Analysis of Structure and Spin-Spin Coupling Constants, Mol. 23 (2018) 3042,
Acknowledgements https://doi.org/10.3390/MOLECULES23113042.
[20] M. Remko, C.W. Von Der Lieth, Conformational structure of some trimeric and
This work was supported by the Russian Science Foundation, Russia pentameric structural units of heparin, J. Phys. Chem. A. 111 (2007)
13484–13491, https://doi.org/10.1021/JP075330L/SUPPL_FILE/JP075330L-
(project 18-73-10009). FILE003.PDF.
[21] M. Lettow, K. Greis, M. Grabarics, J. Horlebein, R.L. Miller, G. Meijer, G. Von
References Helden, K. Pagel, Chondroitin Sulfate Disaccharides in the Gas Phase:
Differentiation and Conformational Constraints, J. Phys. Chem. A. 125 (2021)
4373–4379, https://doi.org/10.1021/ACS.JPCA.1C02463/SUPPL_FILE/
[1] C. Hao, H. Xu, L. Yu, L. Zhang, Heparin: An essential drug for modern medicine, JP1C02463_SI_001.PDF.
Prog. Mol. Biol. Transl. Sci. 163 (2019) 1–19, https://doi.org/10.1016/bs. [22] M. Lettow, M. Grabarics, K. Greis, E. Mucha, D.A. Thomas, P. Chopra, G.-J. Boons,
pmbts.2019.02.002. R. Karlsson, J.E. Turnbull, G. Meijer, R.L. Miller, G. von Helden, K. Pagel,
[2] A. Devlin, C. Mycroft-West, M. Guerrini, E. Yates, M. Skidmore, Analysis of solid- Cryogenic Infrared Spectroscopy Reveals Structural Modularity in the Vibrational
state heparin samples by ATR-FTIR spectroscopy, BioRxiv. (2019), 538074, Fingerprints of Heparan Sulfate Diastereomers, Anal. Chem. 92 (15) (2020)
https://doi.org/10.1101/538074. 10228–10232, https://doi.org/10.1021/ACS.ANALCHEM.0C02048.
[3] A. Devlin, C. Mycroft-West, P. Procter, L. Cooper, S. Guimond, M. Lima, E. Yates, [23] B. Schindler, L. Barnes, C.J. Gray, S. Chambert, S.L. Flitsch, J. Oomens, R. Daniel,
M. Skidmore, Tools for the quality control of pharmaceutical heparin, Med. 55 A.R. Allouche, I. Compagnon, IRMPD Spectroscopy Sheds New (Infrared) Light on
(2019) 636, https://doi.org/10.3390/medicina55100636. the Sulfate Pattern of Carbohydrates, J. Phys. Chem. A. 121 (2017) 2114–2120,
[4] S. Beni, J.F.K. Limtiaco, C.K. Larive, Analysis and characterization of heparin https://doi.org/10.1021/ACS.JPCA.6B11642/SUPPL_FILE/JP6B11642_SI_001.
impurities, Anal. Bioanal. Chem. 399 (2) (2011) 527–539, https://doi.org/ PDF.
10.1007/s00216-010-4121-x. [24] M. Hricovíni, P.-A. Driguez, O.L. Malkina, NMR and DFT Analysis of Trisaccharide
[5] U. Lindahl, G. Bäckström, L. Thunberg, I.G. Leder, Evidence for a 3-O-sulfated D- from Heparin Repeating Sequence, J. Phys. Chem. B. 118 (41) (2014)
glucosamine residue in the antithrombin-binding sequence of heparin, Proc. Natl. 11931–11942, https://doi.org/10.1021/jp508045n.
Acad. Sci. 77 (11) (1980) 6551–6555, https://doi.org/10.1073/PNAS.77.11.6551. [25] R. Krishnan, J.S. Binkley, R. Seeger, J.A. Pople, Self-consistent molecular orbital
[6] H. Liu, Z. Zhang, R.J. Linhardt, Lessons learned from the contamination of heparin, methods. XX. A basis set for correlated wave functions, J. Chem. Phys. 72 (1)
Nat. Prod. Rep. 26 (2009) 313–321, https://doi.org/10.1039/b819896a. (1980) 650–654, https://doi.org/10.1063/1.438955.
[7] M. Guerrini, D. Beccati, Z. Shriver, A. Naggi, K. Viswanathan, A. Bisio, I. Capila, J. [26] A.A. Granovsky, Firefly (Version 8.1.0), http://classic.chem.msu.su/gran/fi
C. Lansing, S. Guglieri, B. Fraser, A. Al-Hakim, N.S. Gunay, Z. Zhang, L. Robinson, refly/index.html.
L. Buhse, M. Nasr, J. Woodcock, R. Langer, G. Venkataraman, R.J. Linhardt, [27] M.W. Schmidt, K.K. Baldridge, J.A. Boatz, S.T. Elbert, M.S. Gordon, J.H. Jensen,
B. Casu, G. Torri, R. Sasisekharan, Oversulfated chondroitin sulfate is a S. Koseki, N. Matsunaga, K.A. Nguyen, S. Su, T.L. Windus, M. Dupuis, J.
contaminant in heparin associated with adverse clinical events, Nat. Biotechnol. 26 A. Montgomery, General atomic and molecular electronic structure system,
(6) (2008) 669–675, https://doi.org/10.1038/nbt1407. J. Comput. Chem. 14 (11) (1993) 1347–1363.
[8] R. Wieboldt, H. Läubli, Glycosaminoglycans in cancer therapy, AJP-Cell, [28] A.L. Pomerantsev, Chemometrics in Excel, John Wiley & Sons Inc, Hoboken, New
Physiology. 322 (6) (2022) C1187–C1200. Jersey, 2014.
[9] I.U. Antia, K. Mathew, D.R. Yagnik, F.A. Hills, A.J. Shah, Analysis of procainamide- [29] N. Kumar, A. Bansal, G.S. Sarma, R.K. Rawal, Chemometrics tools used in
derivatised heparan sulphate disaccharides in biological samples using hydrophilic analytical chemistry: An overview, Talanta. 123 (2014) 186–199, https://doi.org/
interaction liquid chromatography mass spectrometry, Anal. Bioanal. Chem. 410 10.1016/J.TALANTA.2014.02.003.
(1) (2018) 131–143. [30] N. Mainreck, S. Brézillon, G.D. Sockalingum, F.-X. Maquart, M. Manfait,
[10] A.D. Theocharis, M.E. Tsara, N. Papageorgacopoulou, D.D. Karavias, D. Y. Wegrowski, Rapid characterization of glycosaminoglycans using a combined
A. Theocharis, Pancreatic carcinoma is characterized by elevated content of approach by infrared and Raman microspectroscopies, J. Pharm. Sci. 100 (2)
hyaluronan and chondroitin sulfate with altered disaccharide composition, (2011) 441–450, https://doi.org/10.1002/jps.22288.
Biochim. Biophys. Acta - Mol. Basis Dis. 1502 (2) (2000) 201–206, https://doi.org/ [31] I. Capila, R.J. Linhardt, Heparin-protein interactions, Angew. Chem. Int. Ed. 41
10.1016/S0925-4439(00)00051-X. (2002) 390–412, https://doi.org/10.1002/1521-3773(20020201)41:3<390::aid-
[11] A. Pudełko, G. Wisowski, K. Olczyk, E.M. Koźma, The dual role of the anie390>3.0.co;2-b.
glycosaminoglycan chondroitin-6-sulfate in the development, progression and [32] N.A. Burmistrova, B.W.K. Diehl, P.M. Soboleva, E. Rubtsova, E.A. Legin, A.
metastasis of cancer, FEBS J. 286 (10) (2019) 1815–1837, https://doi.org/ V. Legin, D.O. Kirsanov, Y.B. Monakhova, Quality Control of Heparin Injections:
10.1111/FEBS.14748. Comparison of Four Established Methods, Anal. Sci. 36 (12) (2020) 1467–1472,
[12] H. Thiele, M. Sakano, H. Kitagawa, K. Sugahara, A. Rajab, W. Höhne, H. Ritter, https://doi.org/10.2116/analsci.20p214.
G. Leschik, P. Nürnberg, S. Mundlos, Loss of chondroitin 6-O-sulfotransferase-1 [33] D. Grant, W.F. Long, C.F. Moffat, F.B. Williamson, Infrared spectroscopy of
function results in severe human chondrodysplasia with progressive spinal chemically modified heparins, Biochem. J. 261 (1989) 1035–1038, https://doi.
involvement, Proc. Natl. Acad. Sci. U. S. A. 101 (27) (2004) 10155–10160, https:// org/10.1042/bj2611035.
doi.org/10.1073/PNAS.0400334101. [34] N. Mainreck, S. Brézillon, G.D. Sockalingum, F.X. Maquart, M. Manfait,
[13] V. Profant, C. Johannessen, E.W. Blanch, P. Bouř, V. Baumruk, Effects of sulfation Y. Wegrowski, Characterization of glycosaminoglycans by tandem vibrational
and the environment on the structure of chondroitin sulfate studied via Raman microspectroscopy and multivariate data analysis, Proteoglycans. 836 (2012)
optical activity, Phys. Chem. Chem. Phys. 21 (14) (2019) 7367–7377, https://doi. 117–130, https://doi.org/10.1007/978-1-61779-498-8_8.
org/10.1039/C9CP00472F.

You might also like