Elsevier

Polymer

Volume 54, Issue 11, 9 May 2013, Pages 2632-2638
Polymer

Separation of cellulose acetates by degree of substitution

https://doi.org/10.1016/j.polymer.2013.03.041Get rights and content

Abstract

A liquid chromatographic method was developed allowing separating cellulose acetates (CA) with respect to degree of substitution (DS). The normal-phase gradient separation, which is based on an adsorption–desorption mechanism, uses a multi-step gradient of dichloromethane (DCM) and methanol (MeOH) on a bare silica column as stationary phase. Applicability of the developed multi-step gradient allows separating CAs within the DS-range DS = 1.5–2.9, which is the target DS-range for most CA applications. To the best of our knowledge the developed method for the first time allows determination of the DS-distribution of intact CA chains over a wide DS-range.

Introduction

Due to the architectural complexity of cellulose derivatives, analytical methods for comprehensive analysis of the molecular heterogeneities have become increasingly important in cellulose research field. Cellulose derivatives are complex copolymers being heterogeneous at least in molar mass and chemical composition. These heterogeneities need to be characterized as they critically affect many properties of these derivatives such as adhesion strength, solubility, viscoelasticity and drug release from hydrophilic tablets just to name a few [1], [2], [3], [4], [5], [6], [7]. The heterogeneity in chemical composition of cellulose derivatives includes the distribution of the substituents within the individual anhydroglucose units (AGUs), i.e. the partial degree of substitution (DS) of O-2, O-3, and O-6 atoms, as well as the distribution of the substituents among the polymer chains (heterogeneity of 1st order) and along the polymer chains (heterogeneity of 2nd order). DS refers to the average number of substituted hydroxyl groups per AGU. Therefore, DS can take values between 0 and 3 for unbranched cellulose derivatives.

Today cellulose derivatives are mainly characterized in terms of their molar mass and molar mass distribution (MMD), their average DS and of the distribution of the substituents on both the monomer and oligomer levels. NMR techniques are applied to obtain a first insight in the partial DS within the monomer units [8], [9], [10], [11], [12], [13], [14], [15], [16]. Alternatively complete acidic chain degradation and subsequent determination of the resulting differently substituted AGUs can be applied to characterize the composition of cellulose derivatives [17], [18], [19], [20]. In another characterization pathway, the polymer chains are completely or partially degraded using acids or enzymes. The partial degradation of the chains results in a mixture of monomers and/or oligomers present in different molar ratios. These mixtures are separated and characterized subsequently in detail using various analytical techniques, alone or in combination, such as size exclusion chromatography (SEC) [21], [22], [23], [24], anion-exchange chromatography (AEC) [23], [24], [25], [26], [27], [28], gas–liquid chromatography (GLC) [21], [27], [29], [30] and mass spectrometry (MS) [26], [27], [28], [29], [30], [31], [32], [33], [34], [35]. This gives information on the monomer composition as well as the substituent distribution on the oligomeric levels. Such data can be compared with the product distribution expected for a specified monomer distribution along the chain at a certain DP, e.g. a completely random monomer distribution. This procedure therefore yields information on the distribution of the differently substituted AGUs along the polymer chain (2nd order heterogeneity).

In general, any or at least a large amount of information on whether the different monomeric or oligomeric units result from the same or from different chains is lost, if the sample is fully or partially degraded. Therefore, establishing methods to characterize the chemical heterogeneity of cellulose derivatives on the level of intact polymeric chains is still a highly challenging task.

Gradient HPLC has been shown to be a powerful tool for separating (co)polymer molecules according to chemical composition [36], [37]. Such separations allow determining the chemical composition distribution (CCD) of the copolymers, i.e. the distribution of the monomer units among the different polymer chains. Strictly, cellulose derivatives have to be regarded as complex polymers composed of at least eight different monomer units, which are the differently substituted AGUs. Today, it is hardly possible to separate intact cellulose chains of such complexity by the fractions of the differently substituted AGUs. In our somehow more naïve picture we regard cellulose derivatives as being composed of only two kind of structural units, the AGUs and the substituents. A separation with regard to DS is therefore similar to a separation of binary copolymers according to chemical composition. Thus, the DS-distribution is the analogue to the chemical composition distribution of binary copolymers. The efforts on separation and characterization of cellulose derivatives, particularly CAs, with respect to the DS-distribution are very limited and some of the attempts described in literature are briefly discussed here:

SEC with MALLS/RI/UV detections has been used to gain insight on the substitution distribution along the chains of CA [38]. The procedure requires the presence of substituents that can be detected by an UV-detector in relation to the molar mass determined by MALLS/RI. It was found that DS varies with molar mass, with the lower molar mass fractions being less substituted than the higher molar mass fractions. However, the DS gradient over the MMD curve does not represent a true separation in terms of DS, because SEC separation is based on size, not on chemical composition. Furthermore, since CAs do not contain UV-active groups, they had to be modified to the respective amide derivatives. Thus, the modification procedures have to be quantitatively performed in order to achieve reliable results.

Other techniques used to obtain insight into the CCD of CAs are fractionated precipitation and thin layer chromatography (TLC). Fractionated precipitation involves dissolving a solid polymer in a good solvent and precipitating the desired fractions stepwise by decreasing the solubility as a result of the controlled addition of a non-solvent. However, this method is laborious and requires large amounts of solvents and samples. In addition it is not very selective and difficult to automate. In most cases, the isolated fractions of CAs varied in DP but were of the same DS [39], [40], [41], [42], [43]. Thus, the separations were based on molar mass rather than on chemical composition. Kamide et al. evaluated the use of TLC for the separation of CAs according to DS and molar mass [44], [45]. By stepwise changing the eluent composition, they were able to identify experimental conditions where clear dependences of retardation factor (Rf) on either DS or molar mass were observed. The resulting separations according to DS and molar mass were found to be independent of molar mass and DS, respectively. The methods allowed calculating both the DS-distribution and the MMD. However, TLC is difficult to quantify and has a poor reproducibility.

Regarding the application of gradient chromatography, two separation systems for different DS-ranges are reported in literature. Both systems followed reversed-phase liquid chromatography under different conditions. The first system reported by Floyd et al. allowed the separation of cellulose diacetate in the range of DS = 2.3–2.7 [46]. The separation was carried out on a poly(styrene-co-divinyl benzene) based column in a linear gradient from acetone/water/MeOH (4:3:1) to acetone in 15 min at a flow rate of 0.8 mL/min. The samples eluted in the order of increasing average DS. The second system reported by Asai et al. was applied for the separation of cellulose triacetate in the range of DS = 2.7–2.9 [47]. The separation was performed on a Waters Novapak-phenyl column in a gradient from chloroform–MeOH (9/1):MeOH–water (8/1) [2:8] to 100% in 28 min at a flow rate of 0.7 mL/min.

Both systems allow calculating DS-distributions from a linear correlation between DS and retention volume. However, the applicability of both methods is confined to the DS-ranges investigated.

Therefore, the aim of the present paper was to develop a chromatographic method capable of separating CAs according to DS over a wide DS-range.

Section snippets

Materials

CA samples with average DS ranging from DS = 1.5 to DS = 2.9 were used. Five of the samples (sample 4, DS = 1.72; sample 13, DS = 2.42; sample 14, DS = 2.45; sample 15, DS = 2.45; sample 17, DS = 2.92) were kindly provided by Rhodia Acetow GmbH (Freiburg im Breisgau, Germany). The others were prepared in our laboratory by partial saponification of a commercial high DS sample (sample 16, DS = 2.60, Acetati, Italy). Information on the average DS, weight average molar masses (Mw), weight average

Results and discussion

For developing the separation method, a bare silica stationary phase was used.

In a previous study only DMSO and DMAc/LiCl were identified to dissolve all our samples, irrespective of DS [48]. Since the samples were only partially soluble in DMAc without LiCl, DMSO was preferred as solvent due to the incompatibility of non-volatile LiCl with evaporative light scattering detection (ELSD). However, DMSO was not selected as the initial eluent since it creates a high back pressure and detector

Conclusions

CAs of different DS in the DS-range DS = 1.5–2.9 were successfully separated using a normal-phase multi-step gradient from pure DCM to MeOH. The separation is based on adsorption–desorption mechanism rather than on precipitation–redissolution.

The developed method was applied to determine the 1st order chemical heterogeneity (DS-distribution) and the variance of the DS-distribution of CA samples. Significant differences in the chemical heterogeneity were observed even for samples of comparable

Acknowledgement

Hewa Othman Ghareeb would like to acknowledge the Deutscher Akademischer Austauschdienst (DAAD) for providing a PhD scholarship. The allocation of samples by Rhodia Acetow GmbH is gratefully acknowledged.

References (51)

  • M. Scandola et al.

    Polymer

    (1985)
  • C. Clasen et al.

    Progress in Polymer Science

    (2001)
  • A. Viridén et al.

    European Journal of Pharmaceutical Sciences

    (2009)
  • A. Viridén et al.

    International Journal of Pharmaceutics

    (2010)
  • Y. Tezuka et al.

    Carbohydrate Research

    (1995)
  • Y. Tezuka

    Carbohydrate Research

    (1993)
  • J. Reuben et al.

    Carbohydrate Research

    (1983)
  • J. Reuben

    Carbohydrate Research

    (1986)
  • D.L. Verraest et al.

    Carbohydrate Research

    (1997)
  • S. Horner et al.

    Carbohydrate Polymers

    (1999)
  • E.A. Kragten et al.

    Journal of Chromatography

    (1992)
  • A. Cohen et al.

    Journal of Chromatography A

    (2004)
  • J. Cuers et al.

    Carbohydrate Research

    (2012)
  • P.W. Arisz et al.

    Carbohydrate Research

    (1995)
  • T.R. Floyd

    Journal of Chromatography

    (1993)
  • H.O. Ghareeb et al.

    Carbohydrate Polymers

    (2012)
  • C.M. Buchanan et al.

    Macromolecules

    (1991)
  • M. Rinaudo

    Biomacromolecules

    (2004)
  • M. Bonet et al.

    Journal of Adhesion Science and Technology

    (2005)
  • K. Kamide et al.

    Polymer Journal

    (1981)
  • T. Sei et al.

    Polymer Journal

    (1985)
  • K. Kowsaka et al.

    Polymer Journal

    (1986)
  • S.-I. Takahashi et al.

    Journal of Polymer Science Part A: Polymer Chemistry Edition

    (1986)
  • Y. Tezuka et al.

    Macromolecules

    (1987)
  • K. Kowsaka et al.

    Polymer Journal

    (1988)
  • Cited by (10)

    • The retention behavior of diblock copolymers in gradient chromatography; Similarities of diblock copolymers and homopolymers

      2019, Journal of Chromatography A
      Citation Excerpt :

      Liquid chromatography can separate polymers according to molecular size, chemical composition, functionality and other structural features [1–33].

    • NMR characterization of cellulose acetate: Chemical shift assignments, substituent effects, and chemical shift additivity

      2015, Carbohydrate Polymers
      Citation Excerpt :

      As a new technique for obtaining the structural information of CA at the polymeric chain level, liquid chromatography using multi-step gradient separation has been used and allows the determination of the DS distribution of intact CA chains over a wide DS range of 1.5–2.9. Further, this technique reveals the chemical heterogeneity of the CA structure (Ghareeb & Radke, 2013). As mentioned above, NMR spectroscopy has been frequently used to reveal the molecular structure of CA, including substituent distribution, DS, etc., because this method provides structural information at the polymer level without any pretreatments.

    • Polymer separations by liquid interaction chromatography: Principles - prospects - limitations

      2014, Journal of Chromatography A
      Citation Excerpt :

      Therefore the solution to the problem is gradient chromatography, since each high molar mass fraction of a given composition is expected to elute close to its respective critical eluent composition. Linear to weakly curved dependences of the eluent composition at the time of elution on copolymer composition were observed in several studies [62–67], provided the separation is based on adsorption–desorption equilibria. An example of a copolymer separation by chemical composition is given in Fig. 12.

    • Characterization of cellulose acetates according to DS and molar mass using two-dimensional chromatography

      2013, Carbohydrate Polymers
      Citation Excerpt :

      Finally the flow rates in both dimensions need to be optimized to allow for comprehensive characterization in a reasonable analysis time. Taking advantages of the proper sequence of separation methods, it would be beneficial to use the gradient HPLC (i.e. the multi-step gradient from DCM to MeOH) and SEC methods (i.e. DMAc/LiCl) developed recently in our laboratory as the first and second dimension, respectively (Ghareeb & Radke, 2013; Ghareeb et al., 2012). However, the low refractive index increment of CA in DMAc/LiCl requires injection of at least approx. 0.3 mg material into SEC to obtain a sufficiently intense RI-signal.

    View all citing articles on Scopus
    View full text