Mathematical models describing polymer dissolution: consequences for drug delivery

https://doi.org/10.1016/S0169-409X(01)00117-XGet rights and content

Abstract

Polymer dissolution is an important phenomenon in polymer science and engineering that has found applications in areas like microlithography, controlled drug delivery, and plastics recycling. This review focuses on the modeling efforts to understand the physics of the drug release process from dissolving polymers. A brief review of the experimentally observed dissolution behavior is presented, thus motivating the modeling of the mechanism of dissolution. The main modeling contributions have been classified into two broad approaches — phenomenological models and Fickian equations, and anomalous transport models and scaling law-based approaches. The underlying principles and the important features of each approach are discussed. Details of the important models and their corresponding predictions are provided. Experimental results seem to be qualitatively consistent with the present picture.

Introduction

Dissolution of polymers in solvents is an important phenomenon in polymer science and engineering. It has found significant applications in a variety of areas. For example, in microlithography, a process used in fabricating integrated circuits, a photosensitive polymer, called a photoresist, is coated onto a substrate to form a thin film [1], [2], [3], [4], [5]. Irradiation through a glass plate or ‘mask’ bearing an array of circuit patterns allows selected areas of the photoresist to be exposed. The modified or exposed regions of the polymer exhibit an altered rate of dissolution in certain solvents, resulting in the formation of a polymeric image of the mask pattern.

In controlled release applications of polymers, a solute is molecularly dispersed in a polymer phase, which is usually in the glassy state. In the presence of a thermodynamically compatible solvent, swelling occurs and the polymer begins to release its contents to the surrounding fluid. This release process can be controlled either by solute diffusion or by polymer dissolution. In this case, the presence of permanent entanglements in the polymer becomes significant. These systems can give constant release rates under special conditions [6], [7], [8], [9], [10], [11], [12].

Polymer dissolution finds applications in membrane science. In phase inversion, a technique to form asymmetric membranes, a thin film of polymer solution is cast onto a suitable substrate followed by immersion in a coagulation bath (quench step) [13], [14], [15]. During the quench period, solvent/nonsolvent exchange and eventual polymer precipitation occur. The extent of dissolution of the polymer is instrumental in determining the ultimate structure of the membrane. Many crystallizable polymer films such as polycarbonates can be made porous by exposing a uniform film to a beam of alpha particles. The crystalline structure is disrupted and the film then chemically treated with an etchant such as caustic potash. The pores produced are nearly cylindrical and of uniform radius. Such membranes are used for microfiltration. Another technique is to cast films from pairs of compatible, non-complexing polymers. Treatment of such films with a solvent that dissolves only one of the polymers leaves behind a polymer with interconnected microvoids. This acts as a microfiltration membrane. Knowledge of the dissolution behavior of the polymers enables us to select suitable solvents. The membranes produced by the above methods are of uniform texture and are not skinned or asymmetric [16].

Polymer dissolution has also been applied in the treatment of unsorted plastics for recycling [17], [18], [19]. A single solvent such as xylene is used to dissolve five polymers — poly(vinylchloride), polystyrene, low density polyethylene, high density polyethylene and polypropylene — at five different temperatures, ranging from room temperature to 138°C. The remaining poly(ethyleneterephthalate) can be dissolved using a separate solvent. In this process, xylene is placed in a chip-filled vat first at room temperature. Polystyrene dissolves while the other five do not. The xylene solution is then drained to a separate part of the system where it is heated under pressure to about 250°C. The solution is then sent to a vacuum chamber to undergo flash devolatilization, causing xylene to vaporize instantaneously, leaving behind pure polystyrene. The same xylene is then sent back to dissolve another polymer at a different temperature and the process continues with the other polymers.

The dissolution of novolak resins in solvents is an important process in many semiconductor applications [20], [21], [22]. Hence, it becomes necessary to understand the dissolution characteristics of novolaks. Novolak resins, as the world’s oldest synthetic materials, play an important role in today’s semiconductor industry. Their advantages are in their non-swelling nature, aqueous –base developability and etching resistance [23], [24].

Some of the potential applications of the dissolution of polymers are in scaffolding for tissue regeneration [25]. Polymers are shaped into scaffolds that resemble the structure of tissues or organs. They are treated with compounds that help cells adhere onto their surface and multiply. They are then ‘seeded’ with the cells that grow and multiply as the polymer gradually dissolves. The new permanent tissue or the organ is implanted in the patient. So far, only biodegradable polymers have been used for such purposes, thereby considerably narrowing down the number of polymers that can be used. Moreover, most degradable polymers are too weak to be used in load bearing implants [26]. By using semicrystalline polymers, which combine strength with ease of control of the degree of crystallinity, it is possible to obtain desired dissolution rates that correspond to the growth rates of different tissues.

In addition to the above, polymer dissolution rate data have been used to determine glass transition temperature and other thermodynamic parameters associated with polymorphic changes [27]. Dissolution has also found a variety of uses in the pharmaceutical sciences. In the development of microcapsules for sustained release dosage forms [28], the mechanism of drug transport is governed by the dissolution of the polymer. Cooney [29] studied the dissolution of pharmaceutical tablets in the design of sustained release forms. Ozturk and co-workers [30], [31] showed that the dissolution of the polyacid, which is used in enteric-coated tablets, was the controlling step in the release kinetics mechanism.

The dissolution of a polymer into a solvent involves two transport processes, namely solvent diffusion and chain disentanglement. When an uncrosslinked, amorphous, glassy polymer is in contact with a thermodynamically compatible solvent, the latter diffuses into the polymer. A gel-like layer is formed adjacent to the solvent–polymer interface due to plasticization of the polymer by the solvent. After an induction time, the polymer is dissolved. However, there are also cases where a polymer crazes when placed in a solvent. When semicrystalline polymers are exposed to thermodynamically compatible solvents, the unfolding of the crystalline regions is an additional step accompanying solvent diffusion and chain disentanglement. Hence, semicrystalline polymer dissolution becomes equivalent to amorphous polymer dissolution after the crystals in the polymer unfold. A schematic of polymer dissolution is shown in Fig. 1.

In general, polymer dissolution differs from dissolution of a non-polymeric material in two aspects. Polymers require an induction time before starting to dissolve, while non-polymeric materials dissolve instantaneously. Also, polymer dissolution can be controlled either by the disentanglement of the polymer chains or by the diffusion of the chains through a boundary layer adjacent to the solvent–polymer interface. However, the dissolution of non-polymeric materials is generally controlled by the external mass transfer resistance through a liquid layer adjacent to the solid–liquid interface.

This review focuses on understanding the mathematical modeling of the mechanism of polymer dissolution in drug delivery applications. An appropriate understanding of the various controlling steps in the dissolution process greatly enhances the tailoring of the polymer to achieve not only desired rates of drug release, but also the desired profiles.

Section snippets

Mechanisms and models

This section details the current understanding in the field with respect to mechanisms and models for dissolution of amorphous as well as semicrystalline polymers. The various experimental techniques used to characterize dissolution behavior are summarized in Table 1.

Current models for dissolution-controlled drug delivery systems

Controlled delivery of drugs, proteins and other bioactive agents can be achieved by incorporating them either in dissolved or in dispersed form in polymers [61]. In the development of controlled release systems, mathematical modeling of the release process plays a significant role as it establishes the mechanism(s) of drug (solute) release and provides more general guidelines for the development of other systems. It is accepted that numerous successful controlled delivery systems have been

Phenomenological models

Harland et al. [70] formulated for the first mathematical model for drug release in a dissolving polymer–solvent system. The transport was assumed to be Fickian and mass balances were written for the drug and the solvent at the glassy–rubbery interface and at the rubbery–solvent interface. The important parameters identified in the phenomenon were the polymer volume fraction c* at the glassy–rubbery transition, the polymer volume fraction cd for disentanglement of the chains and the

Molecular models

The above model by Harland et al. [70] was modified by Narasimhan and Peppas [10] by accounting for macromolecular chain disentanglement (Fig. 6). This enabled a molecular understanding of the dissolution mechanism of the polymer. This information is important to design tailor-made drug delivery systems for specific applications. A one-dimensional water and drug diffusion is followed by chain disentanglement in amorphous, uncrosslinked, and linear polymers. This model describes transport in a

Conclusions and future outlook

The phenomenological models identified a polymer ‘dissolution rate’ as a key parameter. However, they failed to quantitatively predict this rate from the molecular properties of the polymer and the solvent as treated this rate as a model parameter. Additionally, there are parameters used in these approaches that cannot be obtained from experiments. This is a key area where experimentalists and modelers need to work together since there are numerous models containing parameters that cannot be

References (82)

  • T.T. Moore et al.

    Experimental investigation and mathematical modeling of pluronic F127 gel dissolution: drug release in stirred systems

    J. Contr. Rel.

    (2000)
  • H.R. Lee et al.

    Mathematical models and experiments for swelling phenomena before dissolution of polymer films

    Chem. Eng. Sci.

    (1991)
  • J. Siepmann et al.

    A new model describing the swelling and drug release kinetics from hydroxypropyl methyl cellulose tablets

    J. Pharm. Sci.

    (1999)
  • J. Siepmann et al.

    Modeling plasticizer uptake in aqueous polymer dispersions

    Int. J. Pharm.

    (1998)
  • S.K. Mallapragada et al.

    Crystal dissolution-controlled release systems: I. Physical characteristics and modeling analysis

    J. Contr. Rel.

    (1997)
  • R.T.C. Ju et al.

    Drug release from hydrophilic matrices. 1. New scaling laws for predicting polymer and drug release based on the polymer disentanglement concentration and the diffusion layer

    J. Pharm. Sci.

    (1995)
  • R.T.C. Ju et al.

    Drug release from hydrophilic matrices. 2. A mathematical model based on the polymer disentanglement concentration and the diffusion layer

    J. Pharm. Sci.

    (1995)
  • R.T.C. Ju et al.

    Diffusion coefficients of polymer chains in the diffusion layer adjacent to a swollen hydrophilic matrix

    J. Pharm. Sci.

    (1997)
  • M.J. O’Brien et al.

    Resists in microlithography

  • T. Iwayanagi et al.

    Materials and processes for deep-UV lithography

  • W.J. Cooper et al.

    Dissolution Rates of Poly(methyl methacrylate) Films in Mixed Solvents

    J. Appl. Polym. Sci.

    (1986)
  • E.E. Parsonage et al.

    Properties of positive resists. II. Dissolution characteristics of irradiated poly(methyl methacrylate) and poly(methyl methacrylate–co-maleic anhydride)

    J. Vac. Sci. Technol.

    (1987)
  • E.E. Parsonage et al.

    Properties of positive resists. I. Irradiation-induced degradation and sensitivity of certain methyl methacrylate copolymers

    Br. Polym. J.

    (1987)
  • P. Colombo et al.

    In vitro programmable zero-order release drug delivery system

    Acta Pharm. Technol.

    (1987)
  • P.I. Lee

    Initial concentration distribution as a mechanism for regulating drug release from diffusion-controlled and surface erosion-controlled matrix systems

    J. Contr. Rel.

    (1986)
  • P.I. Lee

    Diffusional release of a solute from a polymeric matrix — approximate analytical solutions

    J. Membr. Sci.

    (1980)
  • C.S. Tsay et al.

    Mass transfer modeling of asymmetric membrane formation by phase inversion

    J. Polym. Sci., Polym. Phys.

    (1990)
  • L. Yilmaz et al.

    Modeling of asymmetric membrane formation. II. The effects of surface boundary conditions

    J. Appl. Polym. Sci.

    (1988)
  • R.E. Kesting
  • E.B. Nauman, J.C. Lynch, Polymer Recycling by Selective Dissolution. US Patent 5,198,471,...
  • E.B. Nauman, J.C. Lynch, Polymer Recycling by Selective Dissolution. US Patent 5,278,282,...
  • A.P. Srivastava et al.

    Kinetics of solvent induced melting of semicrystalline polymers

    Polym. Mater. Sci. Eng. Proceed.

    (1993)
  • R. Dammel

    Recent developments in the dissolution of Novolak-based resins

    Polym. Mater. Sci. Eng. Proceed.

    (1993)
  • T.F. Yeh et al.

    A percolation view of Novolak dissolution. 2. The statistics of a three-dimensional cubic percolation field and a generalized scaling law

    Macromolecules

    (1993)
  • H.-Y. Shih et al.

    Percolation view of Novolak dissolution. 5. The dissolution of exposed resist films

    Macromolecules

    (1996)
  • J.P. Huang et al.

    On the dissolution of Novolak in aqueous alkali

    Macromolecules

    (1989)
  • J.P. Huang et al.

    The dissolution of novolak resins

    Polym. Mater. Sci. Eng. Proceed.

    (1989)
  • R. Langer et al.

    Artificial organs

    Sci. Am.

    (1995)
  • N.A. Peppas et al.

    New challenges in biomaterials

    Science

    (1994)
  • Y. Koida et al.

    Studies on dissolution mechanism of drugs from ethylcellulose microcapsules

    Chem. Pharm. Bull.

    (1987)
  • D.O. Cooney

    Effect of geometry on the dissolution of pharmaceutical tablets and other solids: surface detachment kinetics controlling

    AIChE J.

    (1972)
  • Cited by (165)

    • Polymer dissolution

      2024, Polymers for Pharmaceutical and Biomedical Applications: Fundamentals, Selection, and Preparation
    View all citing articles on Scopus
    View full text