Electrohydrodynamic atomization and spray-drying for the production of pure drug nanocrystals and co-crystals

https://doi.org/10.1016/j.addr.2018.07.012Get rights and content

Abstract

In recent years, nanotechnology has offered attractive opportunities to overcome the (bio)pharmaceutical drawbacks of most drugs such as low aqueous solubility and bioavailability. Among the numerous methodologies that have been applied to improve drug performance, a special emphasis has been made on those that increase the dissolution rate and the saturation solubility by the reduction of the particle size of pure drugs to the nanoscale and the associated increase of the specific surface area. Different top-down and bottom-up methods have been implemented, each one with its own pros and cons. Over the last years, the latter that rely on the dissolution of the drug in a proper solvent and its crystallization or co-crystallization by precipitation in an anti-solvent or, conversely, by solvent evaporation have gained remarkable impulse owing to the ability to adjust features such as size, size distribution, morphology and to control the amorphous/crystalline nature of the product. In this framework, electrohydrodynamic atomization (also called electrospraying) and spray-drying excel due to their simplicity and potential scalability. Moreover, they do not necessarily require suspension stabilizers and dry products are often produced during the formation of the nanoparticles what ensures physicochemical stability for longer times than liquid products. This review overviews the potential of these two technologies for the production of pure drug nanocrystals and co-crystals and discusses the recent technological advances and challenges for their implementation in pharmaceutical research and development.

Introduction

The pharmaceutical industry is found in a persistent and urgent search for new scalable and cost-effective technological strategies to overcome (bio)pharmaceutical drawbacks of drugs such as poor aqueous solubility, low physicochemical stability in the biological milieu, short half-life and reduced bioavailability [[1], [2], [3], [4], [5]]. For instance, >50% of the approved drugs and >70% of new chemical entities under development are classified into Classes II and IV of the Biopharmaceutics Classification System (BCS) [6, 7]. These limitations increase drug attrition rates [8, 9] and lead to a decline in the ability to translate them into new pharmaceutical products [[10], [11], [12]].

Numerous strategies have been applied to overcome these disadvantages. Nanonization of pure drug particles via top-down or bottom-up techniques to produce nanoparticles of amorphous or crystalline nature with sizes ranging from a few nanometers up to 1 μm enhances the dissolution rate and the saturation solubility by increasing the specific surface area-to-volume ratio [1, [13], [14], [15], [16]] (Fig. 1). The Noyes–Whitney Eq. (1) describes the dissolution rate of spherical particles. The process is controlled by diffusion and no chemical reaction takes place [10, [17], [18], [19], [20]].dCxdt=ADhCsCxwhere dCx/dt is the dissolution rate, A is particle surface area, D is the diffusion coefficient, h is the effective thickness of the boundary layer, Cs is particle saturation solubility and Cx is concentration in the surrounding liquid at time x. The Ostwald–Freundlich Eq. (2) establishes the increase in saturation solubility of a given compound based on the increase of the interfacial energy at high curvatures or in other words, for very small particles.S=Sexp2γMrρRTwhere S is the saturation solubility of the nanosized active pharmaceutical ingredient, S is saturation solubility of an infinitely large active pharmaceutical ingredient crystal, γ is the crystal-medium interfacial tension, M is the molecular weight of the compound, r is the particle radius, ρ is the density, R is the gas constant and T is the absolute temperature.

Thus, an increase in the specific surface area-to-volume ratio of the nanonized crystals leads to a faster dissolution rate. In addition, when the particle size is smaller than 100 nm, the saturation solubility increases exponentially. Both phenomena result in an enhancement in the oral bioavailability of the drug [13]. Moreover, the number of contact points with the surrounding tissues (e.g., mucus) increases substantially, favoring the adhesiveness to biological structures and the prolongation of the residence time in specific body sites such as mucosal tissues [13].

Over the last decades, nanosizing techniques have gained increased interest in terms of both new intellectual property and clinical impact [1, 5, 10, 13, 21]. Pure drug nanoparticles can be used dispersed in aqueous media in the so-called nanosuspensions [21, 22] or to produce solid formulations such as tablets [20]. The degree of crystallinity of the drug in the nanoparticle may vary widely and can be controlled by adjusting/optimizing the conditions of the production method [1, 10].

In some cases, the relatively slow dissolution rate of pure nanocrystals of highly hydrophobic drugs combined with their ability to undergo entrapment by the intestinal mucosa resulted in a dramatic increase of the drug half-life with respect to the unprocessed counterpart (Fig. 2), a concept that we coined nanocarrier-less delivery systems [23, 24].

Pure drug nanoparticles have been also used by other minimally-invasive administration routes such as inhalation, transdermal, ophthalmic and buccal [10, 25, 26]. For example, pure drug nanocrystals of hydrophobic drugs (e.g., pranlukat hemihydrate) have been investigated for the treatment of chronic bronchial asthma by pulmonary delivery [[27], [28], [29]]. Drug nanocrystals have been also assessed to improve skin deposition and permeation of the non-steroidal anti-inflammatory drug diclofenac by the transdermal route [30]. More recently, pure drug nanocrystals of hydrophobic drugs (e.g., the antiretroviral rilpivirine) have been investigated to sustain the release upon intramuscular injection in the chronic therapy of the human immunodeficiency virus infection [31, 32]. Unlike other drug nanocarriers that have been extensively researched (i.e., liposomes, nanoemulsions and polymeric nanoparticles) and for which encapsulation efficiency and drug loading have to be defined in the final product, pure drug nanocrystals/co-crystals offer a theoretical drug content of up to 100%. Thus, the encapsulation efficiency is not a constraint which increases the chances of bench-to-bedside translation [13, 33, 34]. On the other hand, small drug particles are thermodynamically instable and they tend to grow in size by agglomeration [25, 35, 36]. This is why they are usually physically stabilized using surfactants or other polymeric stabilizers [1, 25, 33] what brings the typical total drug content to ~50–90% w/w [34]. Another important advantage of drug nanocrystals is the maturity and scalability of their fabrication technologies, which can be demonstrated by multiple commercial products that are currently on the market. Table 1 summarizes oral pharmaceutical formulations based on pure drug nanocrystals currently on the market or under preclinical trials (Table 1) [45]. At the same time, it is important to mention that the administration of nanoparticles with increased oral bioavailability with respect to the unprocessed (non-nanonized) drug might lead to toxicity due to higher Cmax. Thus, bioequivalence and dose adjustment studies need to be conducted to ensure efficacy and safety.

Another approach currently investigated to overcome drug solubility issues is micro/nano-co-crystallization [46, 47]. Pharmaceutical co-crystals are crystalline materials composed of at least two different molecules, typically a drug and a co-crystal former known as a co-former in the same crystal lattice [37, 48, 49]. There are several types of molecular interactions that can generate co-crystals such as π-π stacking, van der Waals forces, H-bonds and ionic bonds [49, 50]. The main advantage of drug co-crystallization is the ability to alter the physicochemical characteristics of the pure drug to improve its solubility, physicochemical stability, dissolution rate and oral bioavailability, while maintaining their therapeutic activity [37, 51]. Conventional co-crystals are formed by a drug and a pharmacologically inert co-former. More recently, drug-drug and multidrug co-crystals were introduced [52, 53]. Their advantage over conventional co-crystals is that the components display a synergistic pharmacological activity as well as enhanced physicochemical properties for at least one of the co-crystal components [52, 54, 55]. Moreover, the combination of multiple therapeutic agents in single unit doses facilitates patient management with complex diseases that require multidrug therapy and increases patient compliance [52, 56].

The techniques applied to nanosize drugs can be classified into three main categories: bottom-up (e.g., conventional nanoprecipitation), top-down (e.g., wet ball milling, high pressure homogenization) and combination techniques [57]. Each one presents pros and cons though in general, all of them have to fulfill similar requirements such as controlled and reproducible size, narrow size distribution, high purity, low content of solvent residues, good physicochemical stability and desired morphology and density [57]. In cases where the nanoparticles are used in the production of solid formulations (e.g., tablets), the mechanical and flow properties of the nanoparticle will govern the tableting process and the final properties of the pharmaceutical product [58, 59]. Moreover, different production methods influence differently the solid state characteristics of the nanoparticles [60]. In general, once the production conditions have been optimized and validated, bottom-up techniques enable a better control of the particle crystallinity/amorphousness and shape, and thus, in recent years, they have gained significant impulse [38]. The current review revisits two bottom-up technologies based on the atomization of a liquid drug solution in a pharmaceutically compatible aqueous or organic solvent into small droplets that undergoes relatively fast drying, namely electrohydrodynamic atomization (EHDA) or electrospraying (these are equivalent terms used to describe the same technique) and spray-drying, for the production of amorphous or crystalline pure drug nanoparticles and nano-co-crystals and critically analyzes their potential to play a fundamental role in the production of innovative pharmaceutical formulations with improved (bio)pharmaceutical properties.

Section snippets

Solvents in pharmaceutical production

Both EHDA and spray-drying rely on the atomization of a drug solution employing different mechanisms and the drying of the formed liquid droplets to produce dry drug particles. In general, both technologies have been developed to enable the safe use of a broad spectrum of aqueous and organic solvents, including flammable (e.g., alcohols, ketones), halogenated (e.g., chloroform, 1,1,1,3,3,3 hexafluoro-2-propanol) and aromatic ones (e.g., toluene). However, this equipment has been developed for

The method

EHDA or electrospraying is a versatile technology based on the use of electrically charged fluids, which derives from the electrospinning technology used to produce micro- and nanofibers, though to obtain particles [63, 64]. It is reproducible due to the ability to control the process parameters and it can be easily operated in a continuous manner. Therefore, it has the potential to replace multiple unit operations in pharmaceutical manufacturing [65, 66]. The main advantages of EHDA over other

The method

Spray-drying is a continuous, cost-effective and scalable process to produce dry powders from a fluid feed by atomization through an atomizer into a hot drying gas medium [[109], [110], [111]]. It is widely used in different industries including food, cosmetics, materials and pharmaceuticals [109, 111]. The first patent concerning this technology can be tracked back to the early 1870s [110]. Thereafter, spray-drying underwent constant evolvement until the more advanced equipment and processes

Electrohydrodynamic atomization and spray-drying to produce drug co-crystals

Production of drug co-crystals/drug-drug co-crystals has emerged as an approach to improve physicochemical properties of drugs such as saturation solubility, dissolution rate and chemical stability in biological media, while preserving pharmacological activity. Better dissolution profiles usually result in higher oral bioavailability. In addition, the mechanical properties of the solid can be modified to comply with a variety of formulation processes (e.g., tableting).

Since traditional

Conclusions and future challenges

EHDA and spray-drying are two bottom-up techniques extensively explored for the fabrication of polymeric nanoparticles. Owing to the great versatility to adjust the process conditions for a broad variety of materials, they recently attracted the attention of pharmaceutical scientists to produce pure drug particles. In both techniques, the evaporation of the solvent is very efficient and, depending on the equipment design, also relatively fast. In fact, most of the solvent is eliminated within

Acknowledgments

This work was funded by the Phyllis and Joseph Gurwin Fund for Scientific Advancement. RSA dedicates this article to the memory of her father, Dov Sverdlov, who recently passed away and whose invaluable help, belief and support has made the writing of this article possible.

References (161)

  • J.C. Imperiale et al.

    Novel protease inhibitor-loaded nanoparticle-in-microparticle delivery system leads to a dramatic improvement of the oral pharmacokinetics in dogs

    Biomaterials

    (2015)
  • L. Wu et al.

    Physical and chemical stability of drug nanoparticles

    Adv. Drug Deliv. Rev.

    (2011)
  • E.S. Ha et al.

    Determination and correlation of solubility of pranlukast hemihydrate in five organic solvents at different temperatures and its dissolution properties

    J. Mol. Liq.

    (2017)
  • T. Mizoe et al.

    Preparation of drug nanoparticle-containing microparticles using a 4-fluid nozzle spray drier for oral, pulmonary, and injection dosage forms

    J. Control. Release

    (2007)
  • R. Pireddu et al.

    Novel nanosized formulations of two diclofenac acid polymorphs to improve topical bioavailability

    Eur. J. Pharm. Sci.

    (2015)
  • L. Baert et al.

    Development of a long-acting injectable formulation with nanoparticles of rilpivirine (TMC278) for HIV treatment

    Eur. J. Pharm. Biopharm.

    (2009)
  • Y. Lu et al.

    Injected nanocrystals for targeted drug delivery

    Acta Pharm. Sin. B

    (2016)
  • J.E. Kipp

    The role of solid nanoparticle technology in the parenteral delivery of poorly water-soluble drugs

    Int. J. Pharm.

    (2004)
  • P.R. Mishra et al.

    Production and characterization of Hesperetin nanosuspensions for dermal delivery

    Int. J. Pharm.

    (2009)
  • G.G. Liversidge et al.

    Particle size reduction for improvement of oral bioavailability of hydrophobic drugs: I. Absolute oral bioavailability of nanocrystalline danazol in beagle dogs

    Int. J. Pharm.

    (1995)
  • G.G. Liversidge et al.

    Drug particle size reduction for decreasing gastric irritancy and enhancing absorption of naproxen in rats

    Int. J. Pharm.

    (1995)
  • G.J. Vergote et al.

    In vivo evaluation of matrix pellets containing nanocrystalline ketoprofen

    Int. J. Pharm.

    (2002)
  • R.H. Müller et al.

    Oral bioavailability of cyclosporine: solid lipid nanoparticles (SLN®) versus drug nanocrystals

    Int. J. Pharm.

    (2006)
  • D. Mou et al.

    Potent dried drug nanosuspensions for oral bioavailability enhancement of poorly soluble drugs with pH-dependent solubility

    Int. J. Pharm.

    (2011)
  • V.B. Junyaprasert et al.

    Nanocrystals for enhancement of oral bioavailability of poorly water-soluble drugs

    Asian J. Pharm. Sci.

    (2015)
  • N. Shan et al.

    The role of cocrystals in pharmaceutical science

    Drug Discov. Today

    (2008)
  • R. Thipparaboina et al.

    Multidrug co-crystals: towards the development of effective therapeutic hybrids

    Drug Discov. Today

    (2016)
  • M.L. Cheney et al.

    Coformer selection in pharmaceutical cocrystal development: a case study of a meloxicam-aspirin cocrystal that exhibits enhanced solubility and pharmacokinetics

    J. Pharm. Sci.

    (2011)
  • J. Lu et al.

    Synthesis and preliminary characterization of sulfamethazine-theophylline co-crystal

    J. Pharm. Sci.

    (2010)
  • X. Wan et al.

    A promising choice in hypertension treatment: fixed-dose combinations

    Asian J. Pharm. Sci.

    (2014)
  • N. Rasenack et al.

    Crystal habit and tableting behavior

    Int. J. Pharm.

    (2002)
  • D.N. Nguyen et al.

    Pharmaceutical applications of Electrospraying

    J. Pharm. Sci.

    (2016)
  • A. Jaworek

    Micro- and nanoparticle production by electrospraying

    Powder Technol.

    (2007)
  • J. Xie et al.

    Electrohydrodynamic atomization: a two-decade effort to produce and process micro−/nanoparticulate materials

    Chem. Eng. Sci.

    (2015)
  • I. Hayati et al.

    Investigations into the mechanism of electrohydrodynamic spraying of liquids

    J. Colloid Interface Sci.

    (1987)
  • A. Jaworek et al.

    Classification of the modes of EHD spraying

    J. Aerosol Sci.

    (1999)
  • R.P.A. Hartman et al.

    Electro hydrodynamic atomization in the cone-jet mode. A physical model of the liquid cone and jet

    J. Aerosol Sci.

    (1999)
  • D.R. Chen et al.

    Electrospraying of conducting liquids for monodisperse aerosol generation in the 4 nm to 1.8 μm diameter range

    J. Aerosol Sci.

    (1995)
  • A. Jaworek et al.

    Electrospraying route to nanotechnology: an overview

    J. Electrost.

    (2008)
  • J. Xie et al.

    Electrohydrodynamic atomization for biodegradable polymeric particle production

    J. Colloid Interface Sci.

    (2006)
  • B. Almería et al.

    Electrospray synthesis of monodisperse polymer particles in a broad (60 nm–2 μm) diameter range: guiding principles and formulation recipes

    J. Colloid Interface Sci.

    (2014)
  • M. Nyström et al.

    Fabrication and characterization of drug particles produced by electrospraying into reduced pressure

    J. Electrost.

    (2010)
  • J. Xie et al.

    Microparticles developed by electrohydrodynamic atomization for the local delivery of anticancer drug to treat C6 glioma in vitro

    Biomaterials

    (2006)
  • A.M. Gañán-Calvo et al.

    Current and droplet size in the electrospraying of liquids. Scaling laws

    J. Aerosol Sci.

    (1997)
  • S.W. Li et al.

    Aspirin particle formation by electric-field-assisted release of droplets

    Chem. Eng. Sci.

    (2006)
  • M. Wang et al.

    Production and characterization of carbamazepine nanocrystals by electrospraying for continuous pharmaceutical manufacturing

    J. Pharm. Sci.

    (2012)
  • J. Yao et al.

    Characterization of electrospraying process for polymeric particle fabrication

    J. Aerosol Sci.

    (2008)
  • S.N. Jayasinghe et al.

    Effect of viscosity on the size of relics produced by electrostatic atomization

    J. Aerosol Sci.

    (2002)
  • A. Gomez et al.

    Production of protein nanoparticles by electrospray drying

    J. Aerosol Sci.

    (1998)
  • K. Tang et al.

    Monodisperse electrosprays of low electric conductivity liquids in the cone-jet mode

    J. Colloid Interface Sci.

    (1996)
  • Cited by (46)

    • Nanocrystals as a versatile platform for theranostic applications

      2023, Advanced Nanoformulations: Theranostic Nanosystems: Volume 3
    View all citing articles on Scopus

    This review is part of the Advanced Drug Delivery Reviews theme issue on “Drug Nanoparticles and Nano-Cocrystals: From Production and Characterization to Clinical Translation”

    View full text