07.04.2013 Views

Drug nanocrystals for the formulation of poorly soluble drugs and its ...

Drug nanocrystals for the formulation of poorly soluble drugs and its ...

Drug nanocrystals for the formulation of poorly soluble drugs and its ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

J Nanopart Res (2008) 10:845–862<br />

DOI 10.1007/s11051-008-9357-4<br />

TECHNOLOGY AND APPLICATIONS<br />

<strong>Drug</strong> <strong>nanocrystals</strong> <strong>for</strong> <strong>the</strong> <strong>for</strong>mulation <strong>of</strong> <strong>poorly</strong> <strong>soluble</strong><br />

<strong>drugs</strong> <strong>and</strong> <strong>its</strong> application as a potential drug delivery system<br />

Lei Gao Æ Dianrui Zhang Æ Minghui Chen<br />

Received: 13 April 2007 / Accepted: 24 December 2007 / Published online: 22 March 2008<br />

Ó Springer Science+Business Media B.V. 2008<br />

Abstract Formulation <strong>of</strong> <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong> is a<br />

general intractable problem in pharmaceutical field,<br />

especially those compounds <strong>poorly</strong> <strong>soluble</strong> in both<br />

aqueous <strong>and</strong> organic media. It is difficult to resolve<br />

this problem using conventional <strong>for</strong>mulation<br />

approaches, so many <strong>drugs</strong> are ab<strong>and</strong>oned early in<br />

discovery. Nanocrystals, a new carrier-free colloidal<br />

drug delivery system with a particle size ranging<br />

from 100 to 1000 nm, is thought as a viable drug<br />

delivery strategy to develop <strong>the</strong> <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong>,<br />

because <strong>of</strong> <strong>the</strong>ir simplicity in preparation <strong>and</strong> general<br />

applicability. In this article, <strong>the</strong> product techniques <strong>of</strong><br />

<strong>the</strong> <strong>nanocrystals</strong> were reviewed <strong>and</strong> compared, <strong>the</strong><br />

special features <strong>of</strong> drug <strong>nanocrystals</strong> were discussed.<br />

The researches on <strong>the</strong> application <strong>of</strong> <strong>the</strong> drug<br />

<strong>nanocrystals</strong> to various administration routes were<br />

described in detail. In addition, as introduced later,<br />

<strong>the</strong> <strong>nanocrystals</strong> could be easily scaled up, which was<br />

<strong>the</strong> prerequisite to <strong>the</strong> development <strong>of</strong> a delivery<br />

system as a market product.<br />

L. Gao D. Zhang (&)<br />

Department <strong>of</strong> Pharmaceutics, College <strong>of</strong> Pharmacy,<br />

Sh<strong>and</strong>ong University, 44 Wenhua Xilu, Jinan 250012,<br />

China<br />

e-mail: zhangdianrui2006@163.com<br />

M. Chen<br />

Department <strong>of</strong> Pharmacology, College <strong>of</strong> Pharmacy,<br />

Sh<strong>and</strong>ong University, 44 Wenhua Xilu, Jinan 250012,<br />

China<br />

Keywords Poorly <strong>soluble</strong> <strong>drugs</strong><br />

Nanocrystals Bioavailability Precipitation<br />

Pearl milling High-pressure homogenization<br />

Colloids Nanomedicine<br />

Introduction<br />

It is estimated that approximately 40% or more <strong>of</strong> <strong>the</strong><br />

new chemical entities (NCE) generated through drug<br />

discovery programs are <strong>poorly</strong> <strong>soluble</strong> in water<br />

(Lipinski 2002). Formulation <strong>of</strong> this class <strong>of</strong> compounds<br />

is a challenging problem faced by <strong>the</strong><br />

pharmaceutical researcher, because typical problems<br />

associated with <strong>the</strong>se <strong>drugs</strong> are a too low oral<br />

bioavailability <strong>and</strong> erratic absorption due to <strong>the</strong>ir too<br />

low saturation solubility <strong>and</strong> dissolution velocity. For<br />

oral administration, <strong>the</strong> low concentration gradient<br />

between <strong>the</strong> gut <strong>and</strong> blood vessel due to <strong>the</strong> poor<br />

solubility <strong>of</strong> <strong>the</strong> drug leads to a limited transport,<br />

consequently influences <strong>the</strong> oral absorption. Parenteral<br />

administration as microsuspensions (e.g. i.m. or<br />

i.p.) frequently cannot lead to sufficient drug levels<br />

due to <strong>the</strong> limited solute volume at <strong>the</strong> injection site.<br />

To obtain a sufficiently high bioavailability, intravenous<br />

injection is necessarily <strong>for</strong>med by using a<br />

solvent mixture (e.g. water/ethanol), solubilizing<br />

agents (e.g. surfactants like Cremophor EL). The<br />

problems associated with <strong>the</strong>se solubilizing agents<br />

are <strong>the</strong>ir limited ability to increase <strong>the</strong> drug solubility<br />

123


846 J Nanopart Res (2008) 10:845–862<br />

(<strong>of</strong>ten show a large injection volume) or/<strong>and</strong> some<br />

side effects or toxic reaction to <strong>the</strong> body. For<br />

example, to increase <strong>the</strong> bioavailability <strong>of</strong> paclitaxel,<br />

an injectable solution Taxol Ò was produced using<br />

Cremophor EL as solubilizing agent, however, some<br />

adverse effects such as allergic shock caused by <strong>the</strong><br />

Cremophor EL were reported. (Muller <strong>and</strong> Peters<br />

1998).<br />

There<strong>for</strong>e, <strong>the</strong> urgent need <strong>for</strong> <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong><br />

is to efficiently <strong>and</strong> safely increase <strong>the</strong>ir saturation<br />

solubility in <strong>the</strong> body fluid. Generally, it is less cost<br />

effective to chemically modify <strong>the</strong> molecules than<br />

resolve it with <strong>the</strong> <strong>for</strong>mulation methods, because <strong>the</strong><br />

pharmacological activity may be changed when <strong>the</strong><br />

chemical structure is altered, <strong>and</strong> it is costly to<br />

redemonstrate <strong>the</strong> efficacy <strong>and</strong> safety <strong>of</strong> <strong>the</strong> newly<br />

generated chemical species. So in many cases<br />

screening an ideal <strong>for</strong>mulation is a preferred way to<br />

develop <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong>.<br />

Some conventional <strong>for</strong>mulation approaches are<br />

used to tackle <strong>the</strong> problems, such as generating a salt,<br />

using solvent mixture or solubilizing agents, incorporating<br />

into complexing agents, using some drug<br />

carrier (e.g. o/w emulsions delivery). But <strong>the</strong> successful<br />

application <strong>of</strong> all <strong>the</strong>se approaches is limited<br />

because <strong>the</strong>y required <strong>the</strong> <strong>drugs</strong> to have processspecific<br />

properties, <strong>for</strong> instance possessing sufficient<br />

ionizable ability, possessing sufficient solubility in<br />

oils or o<strong>the</strong>r hydrophobic mediums, having a suitable<br />

molecular size <strong>and</strong> shape to incorporate in <strong>the</strong><br />

cyclodextrine ring. In <strong>the</strong> 1990s, liposomes achieved<br />

reasonable success in <strong>for</strong>mulating <strong>poorly</strong> <strong>soluble</strong><br />

<strong>drugs</strong> (Mohammed et al. 2004), but <strong>the</strong> poor physical<br />

<strong>and</strong> chemical stability <strong>and</strong> expensive costs were<br />

major obstacles <strong>for</strong> launching liposomes to <strong>the</strong><br />

market. So <strong>the</strong>se approaches are not suitable <strong>for</strong> all<br />

<strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong>, especially those <strong>drugs</strong> that are<br />

not <strong>soluble</strong> in both aqueous <strong>and</strong> organic media.<br />

There<strong>for</strong>e, <strong>the</strong>re is an imminent need <strong>for</strong> an economic<br />

<strong>and</strong> a universal approach applicable to all <strong>drugs</strong>.<br />

Micronization is a classic method applied to <strong>the</strong><br />

<strong>for</strong>mulation <strong>of</strong> <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong>; <strong>for</strong> many years<br />

mechanical milling (e.g. jet milling or colloid milling)<br />

is a commonly used technique to enhance dissolution<br />

velocity <strong>of</strong> <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong> (Rasenack <strong>and</strong><br />

Muller 2003). When <strong>the</strong> particles are comminuted<br />

into micrometer range, <strong>the</strong> surface-to-volume ratio<br />

increases significantly which leads to an increase <strong>of</strong><br />

123<br />

dissolution velocity according to <strong>the</strong> Noyes–Whitney<br />

equation, <strong>and</strong> subsequently an improvement <strong>of</strong><br />

absorption in <strong>the</strong> gastrointestinal tract (GIT). If,<br />

however, <strong>the</strong> compound has a very poor solubility,<br />

e.g. below <strong>the</strong> mg/mL level, <strong>the</strong> definite increase <strong>of</strong><br />

dissolution velocity is not sufficient to obtain a high<br />

bioavailability.<br />

Consequently, <strong>the</strong> next step is nanosizing, i.e. <strong>the</strong><br />

trans<strong>for</strong>mation <strong>of</strong> <strong>the</strong> micronized drug particles to<br />

nanoparticles (Liversidge et al. 1991; Muller et al.<br />

1998). In 1990s a new drug solid nanoparticle<br />

technology came about, i.e. drug <strong>nanocrystals</strong>. Nanocrystals,<br />

consisting <strong>of</strong> pure <strong>drugs</strong> <strong>and</strong> a minimum <strong>of</strong><br />

surface active agents required <strong>for</strong> stabilization, are a<br />

carrier-free submicron colloidal drug delivery with a<br />

mean particle size in <strong>the</strong> nanometer range, typically<br />

between 10 <strong>and</strong> 1000 nm. The drug <strong>nanocrystals</strong> not<br />

only fur<strong>the</strong>r increase <strong>the</strong> dissolution velocity but also<br />

<strong>the</strong> saturation solubility (<strong>for</strong> details see section<br />

‘Characterization <strong>and</strong> evaluation <strong>of</strong> nanocrystal suspensions’),<br />

so <strong>the</strong>y could more efficiently improve<br />

<strong>the</strong> oral absorption <strong>of</strong> <strong>the</strong> <strong>drugs</strong> <strong>and</strong> achieve a higher<br />

bioavailability compared to <strong>the</strong> microparticles. Following<br />

some simple post process, more convenient<br />

dosage <strong>for</strong>ms could be obtained, such as tablets,<br />

pellets, capsules. Apart from <strong>the</strong>se, due to <strong>the</strong><br />

sufficiently small size <strong>and</strong> safe composition, <strong>nanocrystals</strong><br />

can be injected parenterally, even injected<br />

intravenously; <strong>the</strong>n a 100% bioavailability can be<br />

obtained. At present, several production techniques<br />

are applied to produce drug <strong>nanocrystals</strong>, such as<br />

precipitation (Gassmann et al. 1994), pearl milling<br />

(Liversidge et al. 1991), <strong>and</strong> high-pressure homogenization<br />

(Krause <strong>and</strong> Muller 2001). Basically, <strong>the</strong>se<br />

techniques produce a suspension (so-called nanosuspensions)<br />

<strong>of</strong> drug <strong>nanocrystals</strong> dispersed in liquid<br />

medium <strong>and</strong> stabilized by surface active agents. The<br />

use <strong>of</strong> drug <strong>nanocrystals</strong> is a universal approach<br />

generally applicable to all <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong><br />

because all <strong>the</strong>se <strong>drugs</strong> could be directly disintegrated<br />

into nanometer-sized drug particles.<br />

This paper introduces different techniques preparing<br />

drug <strong>nanocrystals</strong>, especially <strong>the</strong> high-pressure<br />

homogenization methods. The special properties <strong>and</strong><br />

general applications <strong>of</strong> <strong>the</strong> <strong>nanocrystals</strong> are discussed<br />

at length. The possibilities <strong>of</strong> large-scale production,<br />

<strong>the</strong> prerequisite to bring a <strong>for</strong>mulation to market, are<br />

also discussed.


J Nanopart Res (2008) 10:845–862 847<br />

Overview <strong>of</strong> production techniques <strong>of</strong> drug<br />

<strong>nanocrystals</strong><br />

Generally, <strong>the</strong> drug <strong>nanocrystals</strong> techniques are<br />

classified as ‘‘bottom up’’ methods <strong>and</strong> ‘‘top down’’<br />

methods (Xing 2004). The ‘‘bottom up’’ methods,<br />

such as <strong>the</strong> precipitation technique, mean that <strong>nanocrystals</strong><br />

can be constructed from <strong>the</strong> molecules;<br />

however, <strong>the</strong> ‘‘top down’’ methods, such as <strong>the</strong> pearl<br />

milling <strong>and</strong> homogenization techniques, mean that<br />

<strong>the</strong> <strong>nanocrystals</strong> can be disintegrated step by step<br />

from <strong>the</strong> coarse powder.<br />

Precipitation techniques<br />

Precipitation has been applied to prepare small<br />

particles <strong>for</strong> many years,but until <strong>the</strong> 1980s it was<br />

used in <strong>the</strong> preparation <strong>of</strong> nanoparticles <strong>for</strong> drug<br />

delivery (Siostrom et al. 1993). Sucker <strong>and</strong> coworkers<br />

prepared drug nanoparticles using a precipitation<br />

technique (Gassmann et al. 1994).<br />

Basically, <strong>the</strong> <strong>poorly</strong> water-<strong>soluble</strong> drug is dissolved<br />

in a solvent (in general organic solvent), <strong>and</strong> <strong>the</strong><br />

solution is added into a miscible non-solvent (aqueous<br />

solvent) under agitation. This leads to a sudden high<br />

supersaturation, resulting in rapid nucleation <strong>and</strong><br />

precipitation. This technique simply takes advantage<br />

<strong>of</strong> <strong>the</strong> variation in <strong>the</strong> solubility <strong>of</strong> <strong>the</strong> same <strong>drugs</strong> in<br />

different but miscible liquids. However, <strong>the</strong> tendency<br />

<strong>of</strong> <strong>the</strong> pharmaceutical particles to grow <strong>and</strong> being hard<br />

to inhibit <strong>the</strong> growth brings problems to <strong>the</strong> production.<br />

Stabilizers should be added to avoid <strong>for</strong>mulation<br />

<strong>of</strong> <strong>the</strong> microparticles due to <strong>the</strong> spontaneous aggregation<br />

<strong>of</strong> <strong>the</strong> particles. To obtain uni<strong>for</strong>m nanoparticles,<br />

several parameters should be optimized:<br />

1. Stirring rate<br />

2. The volume ratio <strong>of</strong> antisolvent to solvent<br />

3. <strong>Drug</strong> content<br />

4. Temperature<br />

The increase in stirring rate favors decrease <strong>of</strong> <strong>the</strong><br />

particle size. With <strong>the</strong> increasing stirring rate <strong>the</strong><br />

micromixing (i.e. mixing on <strong>the</strong> molecular level)<br />

between <strong>the</strong> two phases is intensified. High micromixing<br />

efficiency enhanced <strong>the</strong> rate <strong>of</strong> diffusion <strong>of</strong><br />

<strong>drugs</strong> between <strong>the</strong> two phases, which induces a rapid<br />

<strong>and</strong> high homogenous supersaturation <strong>and</strong> increases<br />

<strong>the</strong> nuclei number to produce smaller drug particles<br />

(Douroumis <strong>and</strong> Fahr 2006). A larger volume ratio <strong>of</strong><br />

antisolvent to solvent also contributes to a higher<br />

supersaturation in <strong>the</strong> interface <strong>of</strong> two phases leading<br />

to a rapid nucleation. A moderate drug content is<br />

better <strong>for</strong> <strong>the</strong> precipitation progress. The increased<br />

viscocity due to a higher drug concentration will<br />

hinder <strong>the</strong> diffusion between <strong>the</strong> two phases consequently,<br />

<strong>the</strong>reby leading to a non-uni<strong>for</strong>m<br />

supersaturation. On <strong>the</strong> o<strong>the</strong>r h<strong>and</strong>, higher drug<br />

content will increase <strong>the</strong> probability <strong>of</strong> aggregation<br />

<strong>of</strong> particles (Zhang et al. 2006). A lower temperature<br />

also favors <strong>the</strong> reduction <strong>of</strong> particle size. A lower<br />

saturation solubility at a lower temperature makes <strong>the</strong><br />

supersaturation easier to reach. In addition <strong>the</strong><br />

nucleating process was a process <strong>of</strong> free-energy<br />

decrease <strong>and</strong> heat release, so a lower temperature will<br />

make <strong>for</strong> <strong>the</strong> nucleation.<br />

The precipitation technique is simple <strong>and</strong> cost<br />

effective; no expensive equipments are required,<br />

compared with milling <strong>and</strong> high-pressure homogenization<br />

technique. Fur<strong>the</strong>rmore, this method avoids <strong>the</strong><br />

use <strong>of</strong> high energy like in disintegration technique,<br />

which prevents denaturation <strong>of</strong> drug due to high<br />

energy input (Zhong et al. 2005). Amorphous drug<br />

state <strong>of</strong>ten <strong>for</strong>ms in this progress (Kipp 2004),<br />

enabling an increase in <strong>the</strong> solubility <strong>and</strong> dissolution<br />

velocity <strong>of</strong> <strong>the</strong> drug. However, <strong>drugs</strong> in this highenergy<br />

state tend to convert to more stable crystalline;<br />

so this metastable state needs to be maintained<br />

during both <strong>the</strong> production progress <strong>and</strong> shelf time to<br />

maintain good stability. For precipitation technique<br />

<strong>the</strong> following prerequisites should be satisfied:<br />

1. <strong>the</strong> drug should be <strong>soluble</strong> at least in one solvent.<br />

2. <strong>the</strong> solvent is miscible with a non-solvent.<br />

3. <strong>the</strong> solvents used in this techniques should be<br />

eliminated to an acceptable level in <strong>the</strong> end<br />

products.<br />

The above mentioned all limit <strong>the</strong> application <strong>of</strong> <strong>the</strong><br />

precipitation technique; it cannot be generally used to<br />

all <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong>.<br />

Pearl milling techniques<br />

The first-generation disintegration technique is pearl<br />

milling developed by Liversidge, leading to <strong>the</strong><br />

product Nanocrystals Ò in 1990 (Liversidge et al.<br />

1991). In this method a nanocrystal suspension is<br />

produced using a high shear pearl mill consisting <strong>of</strong><br />

milling chamber, milling pearls, milling motor, <strong>and</strong> a<br />

123


848 J Nanopart Res (2008) 10:845–862<br />

recirculation chamber. Basically, <strong>the</strong> drug, surfactant<br />

<strong>and</strong> water are filled into <strong>the</strong> milling chamber charged<br />

with milling pearls (typically made from glass, zircon<br />

oxide, or polystyrene resin); <strong>the</strong>n <strong>the</strong> pearls are<br />

rotated at a high rate driven by <strong>the</strong> motor <strong>and</strong> <strong>the</strong> drug<br />

was comminuted into nanosized crystals by <strong>the</strong> high<br />

shear <strong>for</strong>ces. A fine nanocrystal nanosuspension can<br />

be obtained through a pearl milling progress <strong>for</strong> hours<br />

to several days, depending on <strong>the</strong> drug hardness,<br />

quantities, <strong>and</strong> requested particle size <strong>for</strong> different<br />

administration routes. The progress can be per<strong>for</strong>med<br />

in ei<strong>the</strong>r batch or recirculation mode. A narrow<br />

particle size distribution with a very little batch-tobatch<br />

variation can be obtained once <strong>the</strong> optimal<br />

<strong>for</strong>mulation <strong>and</strong> progress is achieved. In general <strong>the</strong><br />

following process parameters should be investigated<br />

to obtain an optimal <strong>for</strong>mulation:<br />

1. <strong>Drug</strong> amount<br />

2. Number <strong>of</strong> milling pearls<br />

3. Milling speed<br />

4. Milling time<br />

5. Temperature<br />

One should bear in mind that particle size would<br />

simultaneously exhibit two opposite effects in <strong>the</strong><br />

milling chamber: decrease occurring through fragmentation<br />

to smaller particles <strong>and</strong> particle growth<br />

occurring through collisions between particles<br />

(Annapragada <strong>and</strong> Adjei 1996). To comminute particles<br />

into nanopowders mean that <strong>the</strong> probability <strong>of</strong><br />

collisions between particles in milling chamber<br />

should be reduced to a minimum. <strong>Drug</strong> amount <strong>and</strong><br />

milling time are two main parameters involved in <strong>the</strong><br />

two opposite effects: <strong>the</strong> dispersion <strong>and</strong> <strong>the</strong> <strong>for</strong>mation<br />

<strong>of</strong> secondary aggregates. Larger drug amount <strong>and</strong><br />

longer milling time mean a larger probability <strong>of</strong><br />

collisions between particles, leading to aggregation.<br />

So it should be necessary to use a moderate time<br />

associated with a sufficient drug amount to reduce<br />

particle size (Geze et al. 1999). Milling speed also<br />

should be compromised to reduce <strong>the</strong> occurrence <strong>of</strong><br />

secondary aggregation, while excessive speed will<br />

favor <strong>the</strong> particle agglomeration.<br />

The number <strong>of</strong> milling pearls is ano<strong>the</strong>r important<br />

parameter; more milling pearls lead to more contact<br />

points between pearls <strong>and</strong> <strong>drugs</strong>, but too many<br />

milling pearls in <strong>the</strong> chamber can increase <strong>the</strong> weight<br />

loading <strong>of</strong> machine <strong>and</strong> collisions between pearls<br />

which will consume <strong>the</strong> mechanical energy. Also<br />

123<br />

moderate number <strong>of</strong> milling pearls will be beneficial<br />

<strong>for</strong> production.<br />

It was reported that low milling temperature<br />

slowed down <strong>the</strong> process <strong>of</strong> aggregation <strong>and</strong> provided<br />

high efficiency (Gubskaya et al. 1995). An increase<br />

in friability <strong>of</strong> particle <strong>and</strong> a low heat energy <strong>of</strong><br />

substance at a low temperature are both favorable <strong>for</strong><br />

grinding. In many case, cooling was carried out by<br />

<strong>the</strong> addition <strong>of</strong> liquid nitrogen in milling chamber<br />

be<strong>for</strong>e powder deposition. It was observed that this<br />

short cooling step had an important effect on particle<br />

size reduction <strong>and</strong> narrowed <strong>the</strong> particle size<br />

distribution.<br />

In fact, all <strong>the</strong>se parameters are known to affect<br />

particle size, but <strong>the</strong> interrelation among <strong>the</strong>m is still<br />

not clear; so <strong>the</strong>ir adjustment is <strong>of</strong>ten empirical at<br />

present.<br />

This technique is simple <strong>and</strong> applicable to <strong>drugs</strong><br />

simultaneously in<strong>soluble</strong> in aqueous <strong>and</strong> non-aqueous<br />

media. Product quantity can be flexibly adapted,<br />

ranging from milliliters to liters; so this technique<br />

cannot only provide a new avenue to improve<br />

screening ef<strong>for</strong>ts in <strong>the</strong> lab scale but also provide a<br />

large-scale procedure suitable <strong>for</strong> commercialization<br />

<strong>of</strong> various dosage <strong>for</strong>ms. The first product based on<br />

this technique is Rapamune Ò , an immunosuppressant<br />

agent, marketed by Wyeth Research Laboratories in<br />

2002. Compared with precipitation technique, highenergy<br />

input is needed <strong>and</strong> <strong>the</strong> equipment expenditure<br />

is higher. A major concern associated with pearl<br />

milling is potential erosion <strong>of</strong> material from <strong>the</strong><br />

pearls, leading to product contamination (Keck <strong>and</strong><br />

Muller 2006). The product contamination by erosion<br />

may be problematic, especially <strong>the</strong> final product is<br />

intended to be administrated <strong>for</strong> a chronic <strong>the</strong>rapy.<br />

Ano<strong>the</strong>r problem is that pearl milling is <strong>of</strong>ten a timeconsuming<br />

progress, ranging from hours to several<br />

days. It not only means a low product efficiency but<br />

also bears a risk <strong>of</strong> microbiological pollution, especially<br />

in <strong>the</strong> case <strong>of</strong> <strong>the</strong> progress being per<strong>for</strong>med at<br />

room temperature <strong>and</strong> a dispersion media providing<br />

nutrition to bacteria being used in <strong>the</strong> procedure.<br />

High-pressure homogenization techniques<br />

In <strong>the</strong> mid-1990s, Müller et al. invented <strong>the</strong> secondgeneration<br />

disintegration technique <strong>for</strong> production <strong>of</strong><br />

drug <strong>nanocrystals</strong>, that is nanosuspensions (Disso-<br />

Cubes Ò ) produced by high-pressure homogenization


J Nanopart Res (2008) 10:845–862 849<br />

with a piston-gap homogenizer (Muller et al. 1998).<br />

Basically, macrosuspensions consisting <strong>of</strong> drug, surfactant,<br />

<strong>and</strong> water passes a very thin gap with a high<br />

velocity; <strong>the</strong> energy caused by cavitation <strong>for</strong>ces <strong>and</strong><br />

crystal collision is high enough to comminute <strong>the</strong><br />

drug particles into nanosized crystals. Also, this<br />

technique is applicable to all <strong>drugs</strong> <strong>poorly</strong> <strong>soluble</strong> in<br />

both aqueous <strong>and</strong> non-aqueous media. The first<br />

technology DissoCubes Ò was based on homogenization<br />

<strong>of</strong> particles in water, because <strong>of</strong> universal<br />

application <strong>and</strong> appealing advantages; nanocrystal<br />

technology based on <strong>the</strong> high-pressure homogenization<br />

was developed rapidly within <strong>the</strong> last decade. In<br />

<strong>the</strong> first few years <strong>of</strong> 2000 a new homogenization<br />

technique was developed, which means homogenization<br />

<strong>of</strong> drug particles in non-aqueous media (e.g.<br />

propylene glycol) or mixtures <strong>of</strong> water with watermiscible<br />

liquids (e.g. PEG, glycerol). The registered<br />

trade name is Nanopure Ò . To resolve <strong>the</strong> problem in<br />

<strong>the</strong> precipitation technique, a homogenization progress<br />

is per<strong>for</strong>med following <strong>the</strong> precipitation to<br />

avoid <strong>the</strong> growth <strong>of</strong> drug <strong>nanocrystals</strong>, i.e. NANO-<br />

EDGE Ò (Rabinow 2004), developed recently by <strong>the</strong><br />

company Baxter. This paper briefly introduces <strong>the</strong><br />

principle <strong>of</strong> <strong>the</strong> homogenization progress <strong>and</strong> reviews<br />

<strong>the</strong> piston-gap homogenization technique. The possibility<br />

<strong>of</strong> large-scale production <strong>and</strong> <strong>the</strong> application <strong>of</strong><br />

<strong>the</strong> products based on this technique are also covered.<br />

At present, <strong>the</strong> high-pressure homogenizer lines<br />

have been widely applied in food <strong>and</strong> cosmetic<br />

industry, <strong>and</strong> have already been approved <strong>for</strong> <strong>the</strong><br />

parenteral production (Muller <strong>and</strong> Peters 1998). The<br />

homogenization principle is not a nascent concept;<br />

<strong>for</strong> many years cavitation has been considered as <strong>the</strong><br />

most important effect to diminish particles in a<br />

piston-gap homogenizer. Figure 1 (Muller <strong>and</strong> Keck<br />

2004) shows <strong>the</strong> progress <strong>and</strong> principle <strong>of</strong> producing<br />

drug <strong>nanocrystals</strong> by <strong>the</strong> piston-gap homogenizer. In<br />

this progress <strong>the</strong> raw material suspensions (in microsize)<br />

is suddenly pumped from a pipe into a thin gap<br />

(generally 25 lm at 1500 bar). The high energy<br />

provided by <strong>the</strong> pump converts to <strong>the</strong> increased<br />

kinetic energy as <strong>the</strong> suspension passes through <strong>the</strong><br />

narrow gap. According to Bernoulli’s equation <strong>the</strong><br />

static pressure <strong>of</strong> <strong>the</strong> fluid simultaneously decreases<br />

to maintain constant energy. The static pressure falls<br />

below <strong>the</strong> vapor pressure <strong>of</strong> water causing <strong>the</strong> boiling<br />

<strong>of</strong> <strong>the</strong> fluid. As <strong>the</strong> suspension leaves <strong>the</strong> gap, <strong>the</strong><br />

pressure suddenly rises to normal pressure <strong>and</strong> <strong>the</strong><br />

vapor bubbles implode vigorously, <strong>and</strong> <strong>the</strong> cleavage<br />

along <strong>the</strong> crystal dislocation is caused by cavitation,<br />

fluid shear, <strong>and</strong> particle collision against each o<strong>the</strong>r.<br />

The energy generated in this instantaneous progress<br />

(generally within milliseconds) is high enough to<br />

break down <strong>the</strong> drug microparticles into nano-sizing<br />

particles (Liedtke et al. 2000).<br />

Several tapes <strong>of</strong> piston-gap homogenizers have<br />

been reported to be used on a lab scale production <strong>of</strong><br />

nanosuspension, <strong>for</strong> example, <strong>the</strong> APV Micron LAB<br />

40 (APV Deutschl<strong>and</strong> GmbH, Lubeck, Germany)<br />

(Jacobs et al. 2000), Avestin EmulsiFlex-C5 (Avestin<br />

Inc., Ottawa, Canada) (Hecq et al. 2005), Niro Soavi<br />

NS1001L (Niro Soavi, S.P.A., Italy) (Zhang et al.<br />

2007). Figure 2 shows <strong>the</strong> transmission electron<br />

microscope photos <strong>of</strong> <strong>the</strong> <strong>for</strong>mulation progress <strong>of</strong><br />

<strong>the</strong> azithromycin nanosuspensions produced by a<br />

NS1001L piston-gap homogenizer (Zhang et al.<br />

2007).<br />

To obtain an optimized <strong>for</strong>mulation, <strong>the</strong> effects <strong>of</strong><br />

<strong>the</strong> following process parameters should be<br />

investigated<br />

1. Applied pressure<br />

2. Number <strong>of</strong> homogenization cycles<br />

3. Temperature<br />

Usually, <strong>the</strong> homogenizer can h<strong>and</strong>le varying pressures,<br />

ranging from 100 to 1500 bar <strong>for</strong> most labscale<br />

ones, so <strong>the</strong> effect <strong>of</strong> <strong>the</strong> homogenization<br />

pressure on <strong>the</strong> particle size should be investigated to<br />

optimize <strong>the</strong> final <strong>for</strong>mulation. The pressure provided<br />

by <strong>the</strong> pump converts <strong>the</strong> kinetic energy <strong>of</strong> <strong>the</strong> fluid<br />

in <strong>the</strong> gap. The higher <strong>the</strong> homogenization pressure<br />

is, <strong>the</strong> higher velocity <strong>of</strong> <strong>the</strong> fluid in <strong>the</strong> gap is, <strong>and</strong><br />

<strong>the</strong> static pressure will drop to a larger extent leading<br />

to generation <strong>of</strong> more bubbles <strong>and</strong> <strong>the</strong>n higher energy<br />

to comminute <strong>the</strong> particles. This is consistent with <strong>the</strong><br />

law <strong>of</strong> conservation <strong>of</strong> energy. So it is anticipated that<br />

<strong>the</strong> higher <strong>the</strong> homogenization pressure, <strong>the</strong> smaller<br />

<strong>the</strong> particle size obtained. Usually <strong>for</strong> <strong>the</strong> production<br />

<strong>of</strong> <strong>the</strong> drug <strong>nanocrystals</strong>, a maximum pressure (<strong>for</strong><br />

most lab homogenizers this value is 1500 bar) is<br />

needed.<br />

The progress that <strong>the</strong> fluid passes <strong>the</strong> gap is<br />

per<strong>for</strong>med instantaneously, in general within several<br />

milliseconds. The energy generated in such short time<br />

is not sufficient to comminute all particles into<br />

uni<strong>for</strong>m drug <strong>nanocrystals</strong> even at <strong>the</strong> highest applied<br />

pressure 1500 bar, so more homogenization cycles<br />

123


850 J Nanopart Res (2008) 10:845–862<br />

Fig. 1 Schematic<br />

representation <strong>of</strong><br />

diminution mechanism in<br />

homogenization gap <strong>of</strong> a<br />

piston-gap homogenizer<br />

(Muller <strong>and</strong> Keck 2004)<br />

Fig. 2 Formulation progress <strong>of</strong> azithromycin nanosuspensions<br />

by high-pressure homogenization with <strong>the</strong> increase <strong>of</strong> applied<br />

pressure <strong>and</strong> cycles. (1% azithromycin, 3% Pluronic F68)<br />

(Zhang et al. 2007). (a) TEM Micrographs <strong>of</strong> <strong>the</strong> azithromycin<br />

nanosuspensions after 2 Cycles at 100 bar (10,0009). (b) TEM<br />

123<br />

Micrographs <strong>of</strong> <strong>the</strong> azithromycin nanosuspensions after 5<br />

Cycles at 500 bar (10,0009). (c) TEM Micrographs <strong>of</strong> <strong>the</strong><br />

azithromycin nanosuspensions after 15 Cycles at 1500 bar<br />

(20,0009)


J Nanopart Res (2008) 10:845–862 851<br />

are needed to be per<strong>for</strong>med on <strong>the</strong> suspensions. The<br />

increased cycle number provided more energy to<br />

break down <strong>the</strong> crystalline. There<strong>for</strong>e, homogenization<br />

<strong>of</strong>ten is per<strong>for</strong>med in five, ten, or more cycles,<br />

that depends on <strong>the</strong> hardness <strong>of</strong> <strong>the</strong> drug <strong>and</strong> <strong>the</strong><br />

desired particle size. The studies carried out on<br />

RMKP 22 revealed that an inverse relationship exists<br />

between <strong>the</strong> number <strong>of</strong> homogenization cycles <strong>and</strong><br />

<strong>the</strong> particle size (Muller <strong>and</strong> Peters 1998).<br />

Apart from reducing <strong>the</strong> particle size, more cycles<br />

lead to a more homogenous nanocrystal suspension,<br />

i.e. a narrower size distribution. Because <strong>the</strong> flow rate<br />

<strong>of</strong> fluid in <strong>the</strong> gap is not identical among different<br />

zones, i.e. fluid on central zone <strong>of</strong> <strong>the</strong> pipe has a<br />

higher velocity than <strong>the</strong> fluid near <strong>the</strong> wall, <strong>the</strong><br />

energy dispersed among <strong>the</strong> fluid is not uni<strong>for</strong>m,<br />

leading to an inhomogenous particle size distribution.<br />

With increasing number <strong>of</strong> cycles, <strong>the</strong> probability that<br />

larger particles pass <strong>the</strong> zone <strong>of</strong> high-power density in<br />

<strong>the</strong> middle <strong>of</strong> <strong>the</strong> gap increases; thus <strong>the</strong>se particles<br />

are also diminished.<br />

So <strong>the</strong> particle size is a function <strong>of</strong> <strong>the</strong> pressure<br />

<strong>and</strong> number <strong>of</strong> cycles, <strong>and</strong> a desired particle size can<br />

be achieved by adjusting <strong>the</strong> procedure parameters,<br />

pressure <strong>and</strong> cycle number. The optimal particle size<br />

<strong>of</strong> drug <strong>nanocrystals</strong> depends on <strong>the</strong> <strong>the</strong>rapeutic<br />

purpose <strong>of</strong> <strong>the</strong> drug. For example, a relatively small<br />

particle size, about 100–200 nm, is required <strong>for</strong> a fast<br />

dissolution; <strong>for</strong> a prolonged dissolution, <strong>the</strong> particle<br />

size should be adjusted to a larger level, e.g. 800–<br />

1000 nm.<br />

Temperature is an important parameter which<br />

should be strictly controlled when <strong>the</strong> drug is<br />

temperature sensitive. There is an increase <strong>of</strong> temperature<br />

in <strong>the</strong> homogenization progress, which is not<br />

favorable to <strong>the</strong> temperature-sensitive <strong>drugs</strong>. The<br />

temperature can be promptly reduced by placing a<br />

heat exchanger ahead <strong>of</strong> <strong>the</strong> homogenizer valve, so<br />

<strong>the</strong> sample temperature can be maintained at about<br />

10 °C or even below. For <strong>the</strong> DissoCubes Ò <strong>and</strong><br />

Nanopure Ò technology production at temperatures <strong>of</strong><br />

0 °C <strong>and</strong> even below (e.g. -20 °C, so-called deep<br />

freeze homogenization technique) was described to<br />

process chemically labile compounds. For example,<br />

to avoid <strong>the</strong> degradation <strong>of</strong> omeprazole during<br />

production <strong>the</strong> nanosuspensions were prepared at<br />

0 °C, <strong>and</strong> <strong>the</strong> samples were cooled down to 0 °C<br />

between each cycle (Schwitzera et al. 2004). Hecq<br />

et al. (2005) produced nifedipine nanosuspensions<br />

using an Avestin EmulsiFlex-C5 homogenizer<br />

equipped with a heat exchanger to maintain a lower<br />

temperature. The fact that high-pressure homogenization<br />

is applicable to <strong>the</strong> temperature-sensitive <strong>drugs</strong><br />

also shows that this is a universal technique in some<br />

respects.<br />

To sum up, high-pressure homogenization is a<br />

simple technique. When an optimized procedure was<br />

achieved following <strong>the</strong> adjustment <strong>of</strong> <strong>the</strong> production<br />

parameters, high-quality nanosuspensions with little<br />

patch-to-patch variation can be obtained, i.e. <strong>the</strong><br />

production <strong>of</strong> nanosuspensions by high-pressure<br />

homogenization possesses a high reproducibility<br />

(Grau et al. 2000). An important advantage <strong>of</strong> this<br />

technique is that a considerably high productivity can<br />

be obtained with a very low microparticle content in<br />

<strong>the</strong> product, which is favorable <strong>for</strong> implementation <strong>of</strong><br />

industrial products. In addition, compared with pearl<br />

milling technique, <strong>the</strong> contamination due to <strong>the</strong><br />

erosion from <strong>the</strong> wall <strong>of</strong> <strong>the</strong> homogenizer is at a<br />

lower level. Muller et al. investigated <strong>the</strong> metal<br />

contamination <strong>of</strong> <strong>the</strong> nanosuspensions under a harsh<br />

production condition, i.e. 20 cycles at a maximum<br />

pressure <strong>of</strong> 1500 bar. The most dominant ion iron in<br />

steel is analyzed in nanosuspensions <strong>and</strong> was found to<br />

be below 1 ppm; it is an uncritical level <strong>and</strong> safe even<br />

<strong>for</strong> a chronic <strong>the</strong>rapy (Borner et al. 1999; Krause<br />

et al. 2000).<br />

Characterization <strong>and</strong> evaluation <strong>of</strong> nanocrystal<br />

suspensions<br />

The essential characterization parameters <strong>for</strong> nanocrystal<br />

suspensions are as follows:<br />

Size <strong>and</strong> size distribution<br />

Size <strong>and</strong> size distribution are important characterizations<br />

<strong>of</strong> <strong>the</strong> nanosuspensions because <strong>the</strong>y govern <strong>the</strong><br />

o<strong>the</strong>r characterizations, such as saturation solubility<br />

<strong>and</strong> dissolution velocity, physical stability, or even<br />

biological per<strong>for</strong>mances. The mean particle size <strong>of</strong><br />

nanosuspensions is typically analyzed by photon<br />

correlation spectroscopy (PCS) (Muller <strong>and</strong> Muller<br />

1984). Apart from <strong>the</strong> mean particle diameter, PCS<br />

can also yield <strong>the</strong> width <strong>of</strong> <strong>the</strong> particle size distribution,<br />

i.e. polydispersity index (PI). The PI value<br />

ranges from 0 (monodisperse particles) to 0.500<br />

123


852 J Nanopart Res (2008) 10:845–862<br />

(broad distribution), <strong>and</strong> is an important index that<br />

governs <strong>the</strong> physical stability. For a long-term<br />

stability <strong>the</strong> PI should be as low as possible.<br />

However, due to a narrow measuring range <strong>of</strong> PCS,<br />

approximately from 3 nm to 3 lm, it shows an<br />

incapability in <strong>the</strong> detection <strong>of</strong> <strong>the</strong> larger microparticles.<br />

Then <strong>the</strong> laser diffractometry (LD) is needed to<br />

investigate <strong>the</strong> content <strong>of</strong> particles in <strong>the</strong> micrometer<br />

range or aggregates <strong>of</strong> drug nanoparticles, especially<br />

<strong>for</strong> <strong>the</strong> nanosuspensions that are meant <strong>for</strong> parenteral<br />

<strong>and</strong> pulmonary delivery. The LD yields a volume<br />

distribution <strong>and</strong> possesses a measuring range <strong>of</strong><br />

approximately 0.05–80 lm up to a maximum <strong>of</strong><br />

2000 lm, depending on <strong>the</strong> type <strong>of</strong> equipment<br />

employed. Typical characterization parameters <strong>of</strong><br />

LD are diameters 50, 90, 99%, represented by D50,<br />

D90, <strong>and</strong> D99, respectively (i.e. <strong>the</strong> D50 means that<br />

50% <strong>of</strong> <strong>the</strong> volume <strong>of</strong> <strong>the</strong> particles is below <strong>the</strong> given<br />

size). It should be noted that <strong>the</strong> particle size data <strong>of</strong> a<br />

nanosuspension obtained by LD <strong>and</strong> PCS are not<br />

identical, because <strong>the</strong> LD data are volume based <strong>and</strong><br />

<strong>the</strong> PCS mean particle size is a light-intensityweighted<br />

size (Fig. 3).<br />

A Coulter counter analysis is essential <strong>for</strong> nanosuspensions<br />

to be administrated intravenously.<br />

Compared with a volume distribution <strong>of</strong> <strong>the</strong> LD<br />

analysis, <strong>the</strong> Coulter counter data give an absolute<br />

value, that is <strong>the</strong> absolute number <strong>of</strong> particles per<br />

volume unit <strong>for</strong> <strong>the</strong> different size classes. The size <strong>of</strong><br />

<strong>the</strong> smallest blood capillary is about 5 lm, so even a<br />

Fig. 3 The D90 (-m-) <strong>and</strong> D99 (-j-) <strong>of</strong> RMKP 22 as a<br />

function <strong>of</strong> cycle number (9% drug, 0.6% Phospholipon 90,<br />

2.2% glycerol 85%, pressure 1500 bar, n = 3) (Muller <strong>and</strong><br />

Peters 1998)<br />

123<br />

small content <strong>of</strong> particles greater than 5 lm may<br />

cause capillary blockade or emboli <strong>for</strong>mation. So <strong>the</strong><br />

content <strong>of</strong> microparticles in nanosuspensions should<br />

be controlled strictly by Coulter counter analysis.<br />

Shape <strong>and</strong> morphous<br />

Typically, <strong>the</strong> shape or <strong>the</strong> morphous <strong>of</strong> <strong>the</strong> <strong>nanocrystals</strong><br />

can be determined using a transmission<br />

electron microscope (TEM) <strong>and</strong>/or a scanning electron<br />

microscope (SEM). A wet sample <strong>of</strong> suitable<br />

concentration is needed <strong>for</strong> <strong>the</strong> TEM analysis. When<br />

<strong>the</strong> original nanosuspensions are required to be<br />

processed into dried powder (e.g. by spray drying<br />

or lyophilization), a SEM analysis is essential to<br />

monitor changes <strong>of</strong> <strong>the</strong> particle size be<strong>for</strong>e <strong>and</strong><br />

following <strong>the</strong> progress <strong>of</strong> <strong>the</strong> water removal. In<br />

general, agglomeration phenomenon may occur following<br />

water removal, leading to an increase <strong>of</strong> <strong>the</strong><br />

particles’ size which can be viewed through SEM. To<br />

minimize <strong>the</strong> extent <strong>of</strong> <strong>the</strong> increase <strong>of</strong> particle size,<br />

some excipients should be added as a protectant. For<br />

example, mannitol, generally used as a cryoprotectant<br />

in lyophilization, can recrystallize around <strong>nanocrystals</strong><br />

during <strong>the</strong> water-removal operation, thus<br />

preventing particle interaction <strong>and</strong> agglomeration.<br />

An agglomeration within a certain extent is permitted<br />

when <strong>the</strong> particle size still in an accepted range,<br />

particularly <strong>the</strong> dried powder can be well redispersed<br />

into stable nanosuspensions. The shape <strong>of</strong> <strong>the</strong> drug<br />

crystals depends on <strong>the</strong>ir crystalline structure; different<br />

crystal shapes <strong>of</strong> different <strong>drugs</strong> were viewed<br />

under SEM (For details see ‘Scale up <strong>and</strong><br />

commercialization’).<br />

Zeta potential<br />

The measurement <strong>of</strong> <strong>the</strong> zeta potential allows <strong>the</strong><br />

prediction about <strong>the</strong> storage stability <strong>of</strong> submicron<br />

colloidal dispersion (Li et al. 2006). In general,<br />

particle aggregation is less likely to occur if particles<br />

possess enough zeta potential providing sufficient<br />

electric repulsion, or enough steric barrier providing<br />

sufficient steric repulsion between each o<strong>the</strong>r.<br />

According to <strong>the</strong> literature, a zeta potential <strong>of</strong> at<br />

least -30 mV <strong>for</strong> electrostatic <strong>and</strong> -20 mV <strong>for</strong><br />

sterically stabilized systems is desired to obtain a<br />

physically stable nanocrystal suspensions (Jacobs<br />

et al. 2000).


J Nanopart Res (2008) 10:845–862 853<br />

Crystalline state<br />

The evaluation <strong>of</strong> crystalline state is necessary to<br />

underst<strong>and</strong> <strong>the</strong> polymorphic changes that a drug might<br />

undergo when subjected to nanosizing. Differential<br />

scanning calorimetry (DSC) <strong>and</strong> X-ray diffraction can<br />

be used to evaluate <strong>the</strong> crystalline structure <strong>of</strong> <strong>the</strong> drug<br />

<strong>nanocrystals</strong>. It was reported that some <strong>drugs</strong> retained<br />

<strong>the</strong>ir crystalline state during homogenization, such as<br />

danazol nifedipine, <strong>and</strong> ucb-35440-3 (Hecq et al.<br />

2006). However, <strong>for</strong> some <strong>drugs</strong>, <strong>for</strong> example azithromycin,<br />

<strong>the</strong> results <strong>of</strong> DSC <strong>and</strong> X-ray showed that<br />

amorphous state was generated during homogenization<br />

(Zhang et al. 2007). Though it increases <strong>the</strong><br />

saturation solubility, <strong>the</strong> amorphous state is a metastable<br />

state with a higher energy <strong>and</strong> leads to an<br />

instability during shelf time. Trans<strong>for</strong>ming <strong>the</strong> nanosuspension<br />

into stable dried solid powder by water<br />

removal can resolve this problem.<br />

Saturation solubility <strong>and</strong> dissolution velocity<br />

Determination <strong>of</strong> <strong>the</strong> saturation solubility <strong>and</strong> dissolution<br />

velocity is very important as it not only can<br />

help to assess <strong>the</strong> benef<strong>its</strong> compared to <strong>the</strong> microparticle<br />

<strong>for</strong>mulation but also help to anticipate <strong>the</strong><br />

in vivo per<strong>for</strong>mance (e.g. blood pr<strong>of</strong>iles, plasma<br />

peaks, <strong>and</strong> bioavailability). The nanosuspensions<br />

should be transferred into a dried powder be<strong>for</strong>e <strong>the</strong><br />

investigation <strong>of</strong> <strong>the</strong> dissolution behavior. In <strong>the</strong><br />

shakering experiments a different temperature can<br />

be used to determined <strong>the</strong> saturation solubility <strong>of</strong> a<br />

dried powder in a different artificial medium (i.e.<br />

physiologic saline, artificial gastric juice, or artificial<br />

intestinal juice). To determine <strong>the</strong> dissolution velocity,<br />

<strong>the</strong> methods described in <strong>the</strong> Pharmacopoeia can be<br />

used; also experiments in various physiological<br />

buffers should be per<strong>for</strong>med.<br />

Surface properties<br />

The research on <strong>the</strong> surface parameters <strong>of</strong> nanosuspensions<br />

is very important, especially <strong>for</strong> <strong>the</strong><br />

nanosuspensions to be administrated intravenously.<br />

The fate <strong>of</strong> <strong>the</strong> <strong>nanocrystals</strong> in vivo following injection,<br />

such as organ distribution, depends on <strong>its</strong> surface<br />

properties, such as surface hydrophobicity <strong>and</strong> interactions<br />

with plasma proteins (Schmidt <strong>and</strong> Muller<br />

2003; Goppert <strong>and</strong> Muller 2005). There<strong>for</strong>e, some<br />

special techniques have to be used to evaluate <strong>the</strong><br />

surface properties to give an idea <strong>of</strong> in vivo behavior.<br />

Hydrophobic interaction chromatography has been<br />

used to determine <strong>the</strong> surface hydrophobicity (Wallis<br />

<strong>and</strong> Muller 1993), <strong>and</strong> 2-D PAGE can be per<strong>for</strong>med to<br />

measure <strong>the</strong> protein absorption after intravenous<br />

injection <strong>of</strong> nanosuspensions (Blunk et al. 1996).<br />

Physical <strong>and</strong> chemical properties <strong>of</strong> <strong>nanocrystals</strong><br />

Increase in saturation solubility <strong>and</strong> dissolution<br />

velocity<br />

An important feature <strong>of</strong> drug <strong>nanocrystals</strong> is <strong>the</strong><br />

increase <strong>of</strong> <strong>the</strong> saturation solubility <strong>and</strong> dissolution<br />

velocity. This advantage makes drug <strong>nanocrystals</strong><br />

amenable to many applications. In textbooks, <strong>the</strong><br />

saturation solubility in a given solvent is defined as a<br />

compound-specific constant depending only on <strong>the</strong><br />

temperature. However, <strong>the</strong> saturation solubility is<br />

also a function <strong>of</strong> <strong>the</strong> crystalline structure (i.e. lattice<br />

energy) <strong>and</strong> particle size. In general, solubility is best<br />

<strong>for</strong> <strong>the</strong> polymorphic modification that is characterized<br />

by highest energy <strong>and</strong> lowest melting point. Amorphous<br />

fraction generated in homogenization progress<br />

is also conducive to an increased solubility due to<br />

high inner energy <strong>of</strong> substance in this state (Hancock<br />

et al. 2002). The reason why saturation solubility is<br />

also a function <strong>of</strong> particle size can be explained by<br />

<strong>the</strong> Kelvin <strong>and</strong> <strong>the</strong> Ostwald–Freundlich equations<br />

(Böhm <strong>and</strong> Müller 1999).<br />

The Kelvin equation (Eq. 1) is originally used to<br />

describe <strong>the</strong> vapor pressure over a curved surface <strong>of</strong> a<br />

liquid droplet in gas; it is also applicable to explain<br />

<strong>the</strong> relation between <strong>the</strong> dissolution pressure <strong>and</strong> <strong>the</strong><br />

curvature <strong>of</strong> <strong>the</strong> solid particles in liquid:<br />

Ln Pr<br />

¼<br />

P1<br />

2cMr<br />

ð1Þ<br />

rRTq<br />

where Pr is <strong>the</strong> dissolution pressure <strong>of</strong> a particle with<br />

<strong>the</strong> radius r, P ? is <strong>the</strong> dissolution pressure <strong>of</strong> an<br />

infinitely large particle, c is <strong>the</strong> surface tension, R is<br />

<strong>the</strong> gas constant, T is <strong>the</strong> absolute temperature, r is<br />

<strong>the</strong> radius <strong>of</strong> <strong>the</strong> particle, Mr is <strong>the</strong> molecular weight,<br />

<strong>and</strong> q is <strong>the</strong> density <strong>of</strong> <strong>the</strong> particle.<br />

According to <strong>the</strong> Kevin equation, <strong>the</strong> dissolution<br />

pressure increases with increasing curvature, which<br />

means decreasing particle size. The curvature is<br />

123


854 J Nanopart Res (2008) 10:845–862<br />

enormous when <strong>the</strong> particle size is in <strong>the</strong> nanometer<br />

range; <strong>the</strong>n a large dissolution pressure can be<br />

achieved leading to a shift <strong>of</strong> <strong>the</strong> equilibrium<br />

toward dissolution. The Ostwald–Freundlich (Eq. 2)<br />

directly describes <strong>the</strong> relation between <strong>the</strong> saturation<br />

solubility <strong>of</strong> <strong>the</strong> drug <strong>and</strong> <strong>the</strong> particle size:<br />

log CS<br />

¼<br />

Ca<br />

2rV<br />

2:303RTqr<br />

ð2Þ<br />

where C S is <strong>the</strong> saturation solubility, C a is <strong>the</strong><br />

solubility <strong>of</strong> <strong>the</strong> solid consisting <strong>of</strong> large particles, r<br />

is <strong>the</strong> interfacial tension <strong>of</strong> substance, V is <strong>the</strong> molar<br />

volume <strong>of</strong> <strong>the</strong> particle material, R is <strong>the</strong> gas constant,<br />

T is <strong>the</strong> absolute temperature, q is <strong>the</strong> density <strong>of</strong> <strong>the</strong><br />

solid, <strong>and</strong> r is <strong>the</strong> radius.<br />

It is obvious that <strong>the</strong> saturation solubility (CS) <strong>of</strong><br />

<strong>the</strong> drug increases with a decrease <strong>of</strong> particle size (r).<br />

However, this effect is pronounced <strong>for</strong> materials that<br />

have mean particle size <strong>of</strong> less than 2 lm.<br />

The increase <strong>of</strong> <strong>nanocrystals</strong> in <strong>the</strong> dissolution<br />

velocity can be explained by <strong>the</strong> Noyes–Whitney<br />

equation (Eq. 3) (Dressman et al. 1998). For drug<br />

<strong>nanocrystals</strong>, <strong>the</strong> increased saturation solubility (CS)<br />

<strong>and</strong> surface area (A) lead to an increase in <strong>the</strong><br />

dissolution velocity (dX/dt).<br />

dX<br />

dt<br />

¼ DA<br />

hD<br />

ðCS CtÞ ð3Þ<br />

where dX/dt is <strong>the</strong> dissolution velocity, D is <strong>the</strong><br />

diffusion coefficient, A is <strong>the</strong> surface area, h D is <strong>the</strong><br />

diffusional distance, CS is <strong>the</strong> saturation solubility,<br />

<strong>and</strong> Ct is <strong>the</strong> concentration around <strong>the</strong> particles.<br />

Ano<strong>the</strong>r important factor is <strong>the</strong> diffusional distance<br />

hD, which, as a part <strong>of</strong> <strong>the</strong> hydrodynamic boundary<br />

layer hH, is also strongly dependent on <strong>the</strong> particle<br />

size, as <strong>the</strong> Pr<strong>and</strong>tl equation shows (Mosharraf <strong>and</strong><br />

Nystrom 1995):<br />

hH ¼ k (L 1=2 =V 1=2 ) ð4Þ<br />

where hH is <strong>the</strong> hydrodynamic boundary layer, k<br />

denotes a constant, L is <strong>the</strong> length <strong>of</strong> <strong>the</strong> particle<br />

surface, <strong>and</strong> V is <strong>the</strong> relative velocity <strong>of</strong> <strong>the</strong> flowing<br />

liquid surrounding <strong>the</strong> particle.<br />

According to <strong>the</strong> Pr<strong>and</strong>tl equation, <strong>the</strong> reduced<br />

particle size leads to a decreased diffusional distance<br />

hD <strong>and</strong> consequently an increased dissolution velocity,<br />

as described by <strong>the</strong> Noyes–Whitney equation. There<strong>for</strong>e,<br />

to sum up, a reduction in <strong>the</strong> drug particle size in<br />

<strong>the</strong> nanometer range leads to an increase in solubility<br />

123<br />

as well as <strong>the</strong> dissolution velocity. Both are very<br />

important factors with regard to <strong>the</strong> aim <strong>of</strong> improving<br />

<strong>the</strong> bioavailability <strong>of</strong> <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong>.<br />

Increased adhesiveness<br />

Ano<strong>the</strong>r outst<strong>and</strong>ing feature <strong>of</strong> nanosuspensions is<br />

<strong>the</strong> distinct increased adhesiveness compared to<br />

microparticles. The high adhesiveness, a general<br />

characteristic <strong>of</strong> nanoparticles (Duchene <strong>and</strong> Ponchel<br />

1997), can be considered as ano<strong>the</strong>r factor improving<br />

<strong>the</strong> oral absorption <strong>of</strong> <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong> apart from<br />

<strong>the</strong> increased saturation solubility <strong>and</strong> dissolution<br />

velocity. In section ‘Oral administration route’<br />

positive effects due to this feature are discussed at<br />

length.<br />

A good long-term stability<br />

A long-term stability is ano<strong>the</strong>r special feature <strong>of</strong> <strong>the</strong><br />

nanosuspensions, which is also one <strong>of</strong> prerequisites<br />

<strong>for</strong> a qualified <strong>for</strong>mulation product. The good physical<br />

stability <strong>of</strong> nanosuspensions is mainly reflected<br />

through <strong>the</strong> absence <strong>of</strong> aggregation <strong>and</strong> Ostwald<br />

ripening phenomenon.<br />

The absence <strong>of</strong> <strong>the</strong> aggregation phenomenon is<br />

realized by <strong>the</strong> addition <strong>of</strong> suitable stabilizer. It is<br />

well known that nanoparticles dispersed in a medium<br />

have a nature <strong>of</strong> aggregation (Drews <strong>and</strong> Tsapatsis<br />

2007). The physical instability <strong>of</strong> <strong>the</strong> system varies<br />

from <strong>the</strong> natural tendency <strong>of</strong> <strong>the</strong> nanoparticles to<br />

reduce <strong>the</strong> high surface energy created by <strong>the</strong> large<br />

interface between <strong>the</strong> solid <strong>and</strong> <strong>the</strong> medium. Physical<br />

stability may be achieved by using <strong>the</strong> surface-active<br />

agents as a stabilizer, such as <strong>the</strong> ionic surfactants,<br />

non-ionic surfactants, <strong>and</strong> amphiphilic copolymers<br />

(Kocbek et al. 2006). During <strong>the</strong> homogenization<br />

progress, <strong>the</strong> surfactant diffuse rapidly <strong>and</strong> cover <strong>the</strong><br />

surface <strong>of</strong> <strong>the</strong> crystals, stabilizing <strong>the</strong> system by<br />

providing an electrostatic <strong>and</strong> static repulsion<br />

between <strong>the</strong> crystals (Muller <strong>and</strong> Jacobs 2002). To<br />

sufficiently stabilize special drug <strong>nanocrystals</strong>, <strong>the</strong><br />

used surfactants should meet <strong>the</strong> following requests:<br />

firstly, <strong>the</strong> stabilizer should have a sufficient affinity<br />

<strong>for</strong> <strong>the</strong> particle surface <strong>of</strong> <strong>the</strong> special drug to stabilize<br />

<strong>the</strong> <strong>nanocrystals</strong> (Lee et al. 2005). For example, <strong>the</strong><br />

ionic surfactants such as sodium dodecyl sulfate<br />

(SDS) <strong>and</strong> lecithin could migrate to <strong>the</strong> solid–liquid<br />

intersurface <strong>and</strong> provide an electrostatic barrier


J Nanopart Res (2008) 10:845–862 855<br />

against aggregation <strong>of</strong> <strong>the</strong> particles. The non-ionic<br />

surfactants, such as Pluronic F68 <strong>and</strong> Birj 78, have<br />

multiple attachments <strong>of</strong> hydrophobic domains at<br />

surface which could adequately interact with <strong>the</strong><br />

hydrophobic functional groups <strong>of</strong> <strong>the</strong> compounds <strong>and</strong><br />

<strong>the</strong>n stabilize <strong>the</strong> system by providing <strong>the</strong> steric<br />

barrier between particles (Kipp 2004). Secondly, <strong>the</strong><br />

surfactant should possess an adequately high diffusion<br />

velocity to cover <strong>the</strong> generated surface rapidly,<br />

because <strong>the</strong> homogenization is a very fast progress.<br />

Lastly, <strong>the</strong> amount <strong>of</strong> <strong>the</strong> stabilizer should be<br />

sufficient <strong>for</strong> full coverage on <strong>the</strong> particle surface to<br />

provide enough electronic or steric repulsion between<br />

<strong>the</strong> particles. However, it not true that <strong>the</strong> more <strong>the</strong><br />

stabilizer, <strong>the</strong> better <strong>the</strong> excess surfactants tend to<br />

<strong>for</strong>m micelles containing a small number <strong>of</strong> dissolved<br />

drug molecules (Choi et al. 2005). So screening <strong>for</strong><br />

an optimal surfactant <strong>and</strong> <strong>its</strong> amount is very important<br />

<strong>for</strong> <strong>the</strong> product quality. It was reported that a<br />

binary or ternary mixture <strong>of</strong> electrostatic <strong>and</strong> static<br />

stabilizer usually had a better effectiveness. At<br />

present <strong>the</strong> choice <strong>of</strong> <strong>the</strong> type <strong>and</strong> amount <strong>of</strong> <strong>the</strong><br />

surfactant are mainly based on experiences because<br />

<strong>of</strong> <strong>the</strong> lack <strong>of</strong> systematic underst<strong>and</strong>ing <strong>of</strong> how <strong>the</strong><br />

stabilizer <strong>and</strong> <strong>drugs</strong> interact with each o<strong>the</strong>r.<br />

The capital principle that must be observed to<br />

choose a stabilizer is that it must be acceptable <strong>and</strong><br />

safe to <strong>the</strong> body. For <strong>the</strong> nanosuspensions to be<br />

administrated intravenously, <strong>the</strong> stabilizers should be<br />

more strictly chosen. In addition, <strong>the</strong> choice <strong>of</strong> <strong>the</strong><br />

stabilizers is limited by <strong>the</strong> procedure, e.g. <strong>the</strong><br />

surfactant which is in <strong>the</strong> liquid state (e.g. Tween<br />

80) cannot be used if <strong>the</strong> nanosuspensions need to be<br />

trans<strong>for</strong>med into a dried powder.<br />

The absence <strong>of</strong> Ostwald ripening phenomenon is<br />

mainly attributed to <strong>the</strong> narrow size distribution <strong>of</strong><br />

nanosuspensions (Mantzaris 2005). It was reported<br />

that <strong>the</strong> particles in <strong>the</strong> highly dispersed systems tend<br />

to grow, which is caused by <strong>the</strong> differences in<br />

saturation solubility in <strong>the</strong> vicinity <strong>of</strong> different sized<br />

particles. The phenomenon was called Ostwald<br />

ripening (Jacobs et al. 2000). The solute concentration<br />

is higher in <strong>the</strong> vicinity <strong>of</strong> <strong>the</strong> smaller particles<br />

than that <strong>of</strong> <strong>the</strong> larger ones due to <strong>the</strong> higher<br />

saturation solubility <strong>of</strong> <strong>the</strong> small particles. So <strong>the</strong><br />

molecules will diffuse from <strong>the</strong> surrounding <strong>of</strong> <strong>the</strong><br />

small particles to <strong>the</strong> surrounding <strong>of</strong> <strong>the</strong> large<br />

particles driven by <strong>the</strong> concentration gradient <strong>and</strong><br />

recrystallize on <strong>the</strong> surface <strong>of</strong> <strong>the</strong> larger particles. The<br />

continual dissolution <strong>of</strong> <strong>the</strong> small particles <strong>and</strong><br />

recrystallization <strong>of</strong> <strong>the</strong> solute on <strong>the</strong> surface <strong>of</strong> <strong>the</strong><br />

large particles lead to <strong>the</strong> <strong>for</strong>mulation <strong>of</strong> <strong>the</strong> microparticles.<br />

A narrow size distribution <strong>of</strong> <strong>the</strong><br />

nanosuspensions created by adequate homogenization<br />

progress avoids <strong>the</strong> different saturation solubility<br />

between particles in different sizes <strong>and</strong> conduces to<br />

<strong>the</strong> absence <strong>of</strong> <strong>the</strong> Ostwald ripening <strong>and</strong> a long-term<br />

stability.<br />

For <strong>the</strong> chemically labile <strong>drugs</strong>, <strong>the</strong> nanosuspensions<br />

can also ensure a good chemical stability both<br />

in vitro <strong>and</strong> in vivo. These <strong>drugs</strong>, containing a certain<br />

functional group sensitive to water, hydrogen ions,<br />

oxygen, or various enzymes in <strong>the</strong> body, are easy to<br />

degrade as a <strong>for</strong>m <strong>of</strong> molecule both in vitro <strong>and</strong><br />

in vivo which leads to a short half-time. The surface<br />

layer <strong>of</strong> <strong>the</strong> drug <strong>nanocrystals</strong> can protect <strong>the</strong> drug<br />

below <strong>the</strong> surface against degradation, just like <strong>the</strong><br />

<strong>for</strong>mation <strong>of</strong> an oxide layer on aluminum. A good<br />

example is paclitaxel, a water-sensitive drug. Figure 4<br />

(Muller <strong>and</strong> Keck 2004) shows <strong>the</strong> HPLC diagram <strong>of</strong><br />

an aqueous paclitaxel solution (methanol 10 mL,<br />

water (demineralized) 5 mL, paclitaxel 20.8 mg)<br />

degraded <strong>for</strong> 48 h at room temperature; <strong>the</strong> main<br />

peak degradation products are clearly visible. However,<br />

only <strong>the</strong> paclitaxel peak occurs in <strong>the</strong> HPLC<br />

diagram <strong>of</strong> <strong>the</strong> nanosuspension (stabilized with poloxamer<br />

188) after 4 years’ storage at 4–8 °C, with no<br />

optically visible peaks <strong>of</strong> degradation products.<br />

Improved biological per<strong>for</strong>mance<br />

All <strong>the</strong> advantages <strong>of</strong> nanosuspensions mentioned<br />

above lead to an improvement in <strong>the</strong> in vivo per<strong>for</strong>mance<br />

<strong>of</strong> <strong>drugs</strong> irrespective <strong>of</strong> <strong>the</strong> administration<br />

route used. The improved biological per<strong>for</strong>mance in<br />

various delivery <strong>for</strong>m is discussed later in detail.<br />

Administrations <strong>of</strong> nanosuspensions in different<br />

routes<br />

Oral administration route<br />

The oral route is <strong>the</strong> first choice <strong>for</strong> drug delivery<br />

because <strong>of</strong> <strong>its</strong> numerous advantages, such as safety,<br />

convenience, etc. At present, most nanosuspension<br />

products on <strong>the</strong> market are <strong>for</strong> oral delivery.<br />

123


856 J Nanopart Res (2008) 10:845–862<br />

Fig. 4 HPLC diagram <strong>of</strong> an aqueous paclitaxel solution<br />

degraded <strong>for</strong> 48 h at room temperature; <strong>the</strong> main peak<br />

degradation products are clearly visible (a) However, only<br />

<strong>the</strong> paclitaxel peak occurs in <strong>the</strong> HPLC diagram <strong>of</strong> <strong>the</strong><br />

nanosuspension (stabilized with poloxamer 188) after 4 years’<br />

storage at 4–8 °C, with no optically visible peaks <strong>of</strong><br />

degradation products (b) (Muller <strong>and</strong> Keck 2004)<br />

Following oral administration, <strong>the</strong> drug is absorbed<br />

from <strong>the</strong> gut <strong>and</strong> enters into blood circulation, <strong>the</strong>n<br />

distributed to various tissues. The per<strong>for</strong>mance <strong>of</strong> <strong>the</strong><br />

drug depends on <strong>its</strong> solubility in digestive juice <strong>and</strong><br />

transportation through <strong>the</strong> gastrointestinal tract. The<br />

poor solubility <strong>and</strong>/or dissolution rate <strong>of</strong> <strong>poorly</strong><br />

<strong>soluble</strong> <strong>drugs</strong> limit in vivo absorption <strong>and</strong> cannot<br />

reach an effective <strong>the</strong>rapeutic concentration in <strong>the</strong><br />

blood. In addition, it is well known that <strong>poorly</strong><br />

<strong>soluble</strong> <strong>drugs</strong> <strong>of</strong>ten exhibit increased or accelerated<br />

absorption when <strong>the</strong>y are administered with food.<br />

This positive food effect would be attributed to <strong>the</strong><br />

enhancement <strong>of</strong> <strong>the</strong> dissolution rate in <strong>the</strong> gastrointestinal<br />

tract (GIT) caused by many factors such as<br />

delayed gastric emptying, increased bile secretion,<br />

larger volume <strong>of</strong> <strong>the</strong> gastric fluid, increased gastric<br />

pH (<strong>for</strong> acidic <strong>drugs</strong>), <strong>and</strong>/or increased splanchnic<br />

blood flow (Jinno et al. 2006). Thus <strong>the</strong>re is <strong>of</strong>ten an<br />

absorption variability resulting from <strong>the</strong> presence or<br />

123<br />

absence <strong>of</strong> food. To sum up, <strong>the</strong> problems associated<br />

with <strong>the</strong>se <strong>drugs</strong> are as follows:<br />

1. a low/variable bioavailability<br />

2. a large oral dose<br />

3. a variation in bioavailability resulting from fed/<br />

fasted state<br />

4. a retarded onset <strong>of</strong> action<br />

Many examples verify that <strong>the</strong> nanosuspensions have<br />

a lot <strong>of</strong> positive effects on <strong>the</strong> oral drug delivery <strong>and</strong><br />

can be applied to overcome <strong>the</strong> above issues. When<br />

<strong>the</strong> drug are administered as nanosuspensions, a high<br />

drug concentration gradient between <strong>the</strong> gastrointestinal<br />

tract <strong>and</strong> blood vessel, resulting from <strong>the</strong><br />

increased saturation solubility <strong>and</strong> dissolution velocity<br />

in digestive juice, improves absorption to a large<br />

extent <strong>and</strong> leads to a high bioavailability. A classic<br />

example is danazol, a <strong>poorly</strong> <strong>soluble</strong> gonadotropin<br />

inhibitor (Liversidge <strong>and</strong> Cundy 1995). The absolute<br />

bioavailability <strong>of</strong> marketed danazol microsuspension<br />

Danocrine was only 5.2%. When administered as a<br />

nanosuspension, an absolute bioavailability <strong>of</strong> 82.3%<br />

could be achieved. Amphotericin B, a highly effective<br />

polyene antibiotic, shows a low oral<br />

bioavailability due to <strong>its</strong> low solubility. However,<br />

when administered as nanosuspensions, <strong>the</strong> absorption<br />

<strong>of</strong> Amphotericin B was significantly improved<br />

compared to orally administered conventional commercial<br />

<strong>for</strong>mulations such as Fungizone Ò ,<br />

AmBisome Ò , <strong>and</strong> micronized amphotericin B. The<br />

oral Amphotericin B nanosuspensions significantly<br />

reduced parasite numbers in <strong>the</strong> liver <strong>of</strong> infected<br />

female Balb/c mice by 28.6%, whereas all liposomal<br />

Amphotericin B (Ambisome Ò ), micronized Amphotericin<br />

B <strong>and</strong> Fungizone Ò did not show any curative<br />

effect at all upon oral administration (Fig. 5) (Kayser<br />

et al. 2003).<br />

The adhesiveness to <strong>the</strong> gut wall is also conducive<br />

to a high bioavailability by prolonging residence <strong>and</strong><br />

contact time in <strong>the</strong> GIT. Buparvaquone, a naphthoquinone<br />

antibiotic, is used <strong>for</strong> <strong>the</strong> treatment <strong>of</strong><br />

Cryptosporidium parvum (C. parvum). Because <strong>of</strong><br />

<strong>its</strong> low solubility, <strong>the</strong> bioavailability is very low when<br />

given orally. What makes <strong>the</strong> matter worse is that <strong>the</strong><br />

infection with C. parvum usually causes watery<br />

diarrhea leading to a fast clearance from <strong>the</strong> GIT.<br />

The buparvaquone nanosuspensions cannot only<br />

increase drug solubility but also show a mucoadhesiveness<br />

to <strong>the</strong> gut wall to overcome drug loss


J Nanopart Res (2008) 10:845–862 857<br />

Fig. 5 Comparison <strong>of</strong> percentage reduction <strong>of</strong> Leishmania<br />

donovani parasite load in livers <strong>of</strong> infected Balb/c mice<br />

following oral administration <strong>of</strong> untreated Amphotericin B (as<br />

control), AmBisome Ò , Fungizone Ò , micronized Amphotericin<br />

B <strong>and</strong> Amphotericin B nanosuspensions (Kayser et al. 2003)<br />

resulting from <strong>the</strong> diarrhea (Jacobs et al. 2001).<br />

Fur<strong>the</strong>rmore, to enhance <strong>the</strong> mucoadhesive, <strong>the</strong><br />

buparvaquone nanosuspensions are incorporated into<br />

a mucoadhesive polymer to get a mucoadhesive<br />

nanosuspension. The data illustrate that <strong>the</strong> mucoadhesive<br />

nanosuspensions can more effectively clear<br />

C. parvum from <strong>the</strong> gastrointestinal tract than<br />

unmodified nanosuspensions due to <strong>the</strong> prolonged<br />

gastrointestinal residence time resulting from mucoadhesiveness<br />

(Kayser 2001).<br />

When <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong> are <strong>for</strong>mulated as a<br />

uni<strong>for</strong>m nanosuspension, <strong>the</strong> variation in bioavailability<br />

resulting from <strong>the</strong> fed/fasted state can be<br />

minimized (Hecq et al. 2006; Jinno et al. 2006;<br />

Liversidge <strong>and</strong> Cundy 1995). The reason is that <strong>the</strong><br />

dissolution rate <strong>of</strong> <strong>nanocrystals</strong>, enhanced significantly<br />

due to <strong>the</strong> increase in solubility <strong>and</strong> enormous<br />

particle surface, is fast enough even under <strong>the</strong> fasted<br />

condition. Then, <strong>the</strong> absorptions both in fasted <strong>and</strong><br />

fed state might be a permeability-limited progress,<br />

<strong>and</strong> <strong>the</strong> difference in <strong>the</strong> absorption due to different<br />

dissolution between <strong>the</strong> two conditions will be<br />

eliminated.<br />

Nanosuspensions are also advantageous in achieving<br />

a quick onset <strong>of</strong> action <strong>for</strong> <strong>drugs</strong> with a slow dissolution<br />

rate. It can be illustrated by a study involving a<br />

comparison <strong>of</strong> <strong>the</strong> pharmacokinetic pr<strong>of</strong>iles <strong>of</strong><br />

naproxen after oral administration in <strong>the</strong> <strong>for</strong>m <strong>of</strong><br />

nanosuspensions (mean size is 270 nm) <strong>and</strong> unwilling<br />

suspensions (Liversidge <strong>and</strong> Conzentino 1995).<br />

The data showed that Cmax was achieved at 33.5 min<br />

<strong>for</strong> <strong>the</strong> unwilling naproxen. However, <strong>the</strong> time <strong>for</strong><br />

naproxen nanosuspensions to reach an equivalent<br />

mean plasma concentration to this Cmax was only<br />

about 8 min. It suggested that <strong>the</strong> nanosuspensions<br />

appeared to be absorbed approximately 4-fold faster<br />

than <strong>the</strong> unwilling suspension. The increase in<br />

absorption rate <strong>for</strong> <strong>the</strong> nanosuspensions was attributed<br />

to <strong>the</strong> increased solubility <strong>and</strong> dissolution rate<br />

compared to <strong>the</strong> unwilling microparticles.<br />

Because <strong>the</strong> increase in bioavailability <strong>for</strong> nanosuspensions<br />

is attributed to small particle size <strong>and</strong><br />

large particle surface, an unchanged disparity <strong>of</strong><br />

nanosuspensions in GIT following oral administration<br />

is <strong>the</strong> prerequisite <strong>for</strong> a high bioavailability. If<br />

aggregation <strong>of</strong> particles occurs in <strong>the</strong> GIT, <strong>the</strong><br />

positive effects <strong>of</strong> <strong>the</strong> nanosuspensions on <strong>the</strong> oral<br />

bioavailability may be reduced. For instance, <strong>the</strong><br />

cyclosporine A nanosuspensions showed a disappointing<br />

oral absorption <strong>and</strong> bioavailability due to<br />

<strong>the</strong> aggregation <strong>of</strong> <strong>the</strong> drug <strong>nanocrystals</strong> in <strong>the</strong> GIT<br />

(Muller et al. 2006). So it is important that <strong>the</strong><br />

stabilizer used is able to prevent aggregation <strong>of</strong> <strong>the</strong><br />

nanoparticles in both in vitro <strong>and</strong> in vivo circumstances.<br />

Thus, many physiological factors in <strong>the</strong> GIT<br />

should be considered an ideal stabilizer <strong>for</strong> oral<br />

nanosuspensions. For example, <strong>the</strong> variation in pH<br />

during transit through <strong>the</strong> GIT may preclude <strong>the</strong> use<br />

<strong>of</strong> a charge stabilizing suspending agent, such as<br />

sodium lauryl sulfate (Liversidge <strong>and</strong> Cundy 1995).<br />

Parenteral administration route<br />

For many cases, intravenous injection is requested to<br />

meet some treatment purpose, such as immediate<br />

effect, targeting effect, overcoming <strong>the</strong> first pass<br />

effect, <strong>and</strong> so on. Through intravenous administration,<br />

<strong>drugs</strong> can reach a 100% bioavailability in <strong>the</strong><br />

body. However, as known to us, <strong>the</strong> parenteral<br />

administration is a critical route, <strong>for</strong> which many<br />

requirements should be achieved. For example, <strong>the</strong><br />

products should be sterile, <strong>and</strong> <strong>the</strong> ingredients in <strong>the</strong>m<br />

should not cause any biological problems such as<br />

toxic reactions <strong>and</strong> allergic reactions. As <strong>for</strong> <strong>the</strong><br />

123


858 J Nanopart Res (2008) 10:845–862<br />

submicron delivery system, <strong>the</strong> particle size is also a<br />

crucial factor in determining whe<strong>the</strong>r or not it can be<br />

used in parenteral route. For intravenous injection <strong>the</strong><br />

content <strong>of</strong> particles larger than 5 lm should be<br />

controlled strictly, because <strong>the</strong> smallest size <strong>of</strong> blood<br />

capillaries is about 5 lm, <strong>the</strong> existence <strong>of</strong> a high<br />

content <strong>of</strong> particles larger than 5 lm can lead to<br />

capillary blockade <strong>and</strong> embolism.<br />

Nanosuspensions can be made into an absolutely<br />

safe <strong>and</strong> injectable product <strong>for</strong> parenteral administration.<br />

As mentioned above, nanosuspensions contain<br />

only pure drug <strong>and</strong> slight stabilizers, <strong>and</strong> many<br />

surface active agents that have ever been approved to<br />

be used <strong>for</strong> intravenous injection can be used as<br />

stabilizers. There<strong>for</strong>e, besides reducing <strong>the</strong> injection<br />

volume <strong>and</strong> increasing <strong>the</strong> tolerated dose, nanosuspensions<br />

can also avoid biological problems caused<br />

by some excipients. For example, compared to<br />

Taxol Ò , paclitaxel nanosuspension showed a higher<br />

LD50 (100 mg/kg), <strong>and</strong> no allergic reaction occurred<br />

(Böhm <strong>and</strong> Müller 1999).<br />

Section ‘Increase in saturation solubility <strong>and</strong><br />

dissolution velocity’ has mentioned that nanosuspensions,<br />

that had undergone strict control over <strong>the</strong> larger<br />

particles by Coulter counter analysis, can be safely<br />

administrated intravenously. Injections <strong>of</strong> 0.3 mL<br />

RMKP 23 <strong>nanocrystals</strong> (2.5% solid content) in<br />

C57B1/6 mice were well tolerated without any signs<br />

<strong>of</strong> acute toxicity (Muller <strong>and</strong> Peters 1998). When<br />

judging <strong>the</strong> tolerability one should bear in mind that<br />

<strong>the</strong> blood volume <strong>of</strong> <strong>the</strong> mice is about 2.0 mL, i.e.<br />

15% <strong>of</strong> <strong>the</strong> blood volume was injected as nanosuspensions<br />

(Muller <strong>and</strong> Peters 1998).<br />

At present, several sterilization approaches have<br />

successfully been applied to <strong>the</strong> nanosuspensions,<br />

such as gamma irradiation, filtration sterilization, <strong>and</strong><br />

<strong>the</strong>rmal sterilization (only in case <strong>the</strong> drug <strong>its</strong>elf is not<br />

sensitive to temperature <strong>and</strong> <strong>the</strong> stabilizers are suitable)<br />

(Na et al. 1999). An aseptic production line <strong>for</strong><br />

nanosuspension products has been realized recently<br />

by <strong>the</strong> Baxter company. Such a production line can<br />

even been successfully applied to <strong>the</strong> production <strong>of</strong><br />

cytotoxics, such as paclitaxel. In addition, <strong>the</strong> most<br />

noteworthy point is that <strong>the</strong> high-pressure homogenization<br />

process <strong>its</strong>elf has a germ-reducing effect. In <strong>the</strong><br />

homogenization progress, <strong>the</strong> bacteria as well as <strong>the</strong><br />

crystals are simultaneously disintegrated.<br />

According to <strong>the</strong> required treatment purpose, <strong>the</strong><br />

fate <strong>of</strong> <strong>nanocrystals</strong> in vivo following parenteral<br />

123<br />

administration varies in different cases. This can be<br />

realized by adjusting several parameters, such as<br />

particle size <strong>and</strong> surface properties. In some case a<br />

relatively small particle size is needed <strong>for</strong> <strong>nanocrystals</strong><br />

to achieve a rapid dissolution. For example, <strong>for</strong><br />

cyclosporine, it has been reported that <strong>the</strong> same<br />

pharmacokinetic parameters were obtained after<br />

intravenous injection <strong>of</strong> a solution <strong>and</strong> nanosuspensions<br />

(Muller <strong>and</strong> Keck 2004). In o<strong>the</strong>r cases,<br />

however, it is not necessary to make <strong>the</strong> <strong>nanocrystals</strong><br />

dissolve rapidly. That is <strong>the</strong> drug <strong>nanocrystals</strong> can<br />

circulate in <strong>the</strong> blood circulation as submicron<br />

particles <strong>for</strong> a certain time period. Then <strong>the</strong> <strong>nanocrystals</strong><br />

may be recognized as <strong>for</strong>eign matter <strong>and</strong><br />

rapidly cleared by phagocytic cells <strong>of</strong> mononuclear<br />

phagocyte system (MPS) which abounds in special<br />

tissues <strong>and</strong> organs, such as liver, lung <strong>and</strong> spleen.<br />

This is what is termed as passive targeting progress. It<br />

is important <strong>for</strong> some cases, <strong>for</strong> example, in <strong>the</strong><br />

treatment <strong>of</strong> infections caused by parasites residing<br />

<strong>the</strong> macrophages <strong>of</strong> <strong>the</strong> MPS <strong>and</strong> cancers in special<br />

parts, such as liver <strong>and</strong> lung. It was reported that<br />

following i.v. administration <strong>of</strong> cl<strong>of</strong>azimine nanosuspensions<br />

<strong>and</strong> a control liposomal at a dose <strong>of</strong> 20 mg<br />

drug/kg to mice, drug concentration in livers, spleens,<br />

<strong>and</strong> lungs reached comparably high concentrations,<br />

well in excess <strong>of</strong> <strong>the</strong> MIC <strong>for</strong> most Mycobacterium<br />

avium strains, <strong>and</strong> nanosuspensions showed a better<br />

targeting effect to liver, compared to <strong>the</strong> liposomal<br />

(Peters et al. 2000).<br />

However, <strong>the</strong> clearance <strong>of</strong> <strong>the</strong> MPS is not desirable<br />

when a comparatively enduring <strong>and</strong> high drug<br />

concentration in <strong>the</strong> circulation is requested. At<br />

present studies are in progress to identify how best<br />

to manipulate <strong>the</strong> surface properties, size, <strong>and</strong> shape<br />

<strong>of</strong> <strong>the</strong> <strong>nanocrystals</strong> to eliminate, when desirable,<br />

uptake by <strong>the</strong> phagocytic cells <strong>of</strong> MPS-enriched<br />

organs (Liversidge et al. 2003).<br />

Ocular administration route<br />

For ocular delivery <strong>of</strong> <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong>, preparations<br />

such as suspensions <strong>and</strong> ointments have been<br />

developed to meet <strong>the</strong> treatment request. Although<br />

both two approaches show some advantageous such<br />

as a relative higher drug dose <strong>and</strong> a prolonged<br />

residence time in site <strong>of</strong> action, <strong>the</strong>ir actual per<strong>for</strong>mance<br />

is greatly constrained by <strong>the</strong> limited intrinsic<br />

solubility <strong>of</strong> drug in lachrymal fluids. The low drug


J Nanopart Res (2008) 10:845–862 859<br />

intrinsic solubility in <strong>the</strong> lachrymal fluids leads to a<br />

low level <strong>of</strong> drug concentration in <strong>the</strong> site <strong>of</strong> action,<br />

so an effective per<strong>for</strong>mance can hardly be obtained.<br />

However, when <strong>the</strong> drug is disintegrated into nanosized<br />

crystals <strong>the</strong> remarkably increased saturation<br />

solubility <strong>of</strong> <strong>the</strong> drug proves to be a benefit <strong>for</strong> ocular<br />

delivery <strong>of</strong> <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong>. Moreover, <strong>the</strong><br />

inherent adhesive features <strong>of</strong> <strong>nanocrystals</strong> not only<br />

reduce <strong>the</strong> loss <strong>of</strong> drug caused by <strong>the</strong> outflow <strong>of</strong><br />

lachrymal fluids but also are conducive to a sustained<br />

release <strong>of</strong> <strong>the</strong> drug. In addition, <strong>nanocrystals</strong> can be<br />

incorporated into a suitable mucoadhesive base or<br />

ocular inserts to achieve a sustained release <strong>of</strong> <strong>the</strong><br />

drug <strong>for</strong> a stipulated time period. These all ensure a<br />

high bioavailability <strong>and</strong> effective per<strong>for</strong>mance. At<br />

present, <strong>the</strong> approach <strong>of</strong> a <strong>for</strong>mulation <strong>of</strong> polymeric<br />

nanosuspensions loaded with drug is being investigated<br />

to achieve <strong>the</strong> desired duration <strong>of</strong> action. Some<br />

bioerodible polymers possessing ocular tolerability<br />

can be used to produce sustain-release nanosuspensions.<br />

A good example is <strong>the</strong> polymeric<br />

nanosuspension <strong>of</strong> ibupr<strong>of</strong>en <strong>for</strong> ocular delivery<br />

using acrylate polymers such as Eudragit RS 100 Ò .<br />

The nanosuspensions have been characterized <strong>for</strong><br />

particle size, drug loading, zero potential, in vitro<br />

drug release <strong>and</strong> ocular tolerability, <strong>and</strong> have proven<br />

to be safely used <strong>for</strong> ocular delivery (Rosario et al.<br />

2002). The rabbit in vivo experiment indicated that<br />

<strong>the</strong> polymeric nanosuspensions revealed superiority<br />

in in vivo per<strong>for</strong>mance over <strong>the</strong> existing eye drop<br />

<strong>for</strong>mulations in <strong>the</strong> market <strong>and</strong> could sustain drug<br />

release <strong>for</strong> 24 h (Ponchel et al. 1997).<br />

Pulmonary administration route<br />

Nanosuspensions also show a great potential <strong>for</strong> <strong>the</strong><br />

pulmonary delivery <strong>of</strong> <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong>. Some <strong>of</strong><br />

<strong>the</strong>se <strong>drugs</strong> have been successfully made into aerosols<br />

in <strong>the</strong> <strong>for</strong>m <strong>of</strong> microparticles that can be<br />

nebulized <strong>for</strong> pulmonary administration. But such<br />

microparticle aerosol products are <strong>of</strong>ten accompanied<br />

with some disadvantages. An unwanted deposition <strong>of</strong><br />

<strong>the</strong> microparticles in <strong>the</strong> mouth <strong>and</strong> pharynx, <strong>and</strong><br />

clearance <strong>of</strong> <strong>the</strong> drug by cilia movement in <strong>the</strong> lungs<br />

are <strong>the</strong> main causes leading to <strong>the</strong> loss <strong>of</strong> <strong>the</strong> drug.<br />

Moreover, a low level <strong>of</strong> drug concentration in <strong>the</strong><br />

action site due to a poor drug saturation solubility to a<br />

great extent lim<strong>its</strong> <strong>the</strong> absorption <strong>of</strong> <strong>the</strong> drug in lungs.<br />

All <strong>the</strong>se lead to a lower bioavailability greatly<br />

affecting <strong>the</strong> per<strong>for</strong>mance. However, <strong>the</strong>se problems<br />

can be overcome by application <strong>of</strong> nanosuspensions<br />

<strong>for</strong> pulmonary delivery. The increased dissolution<br />

velocity <strong>and</strong> saturation solubility can rapidly realize a<br />

larger concentration <strong>of</strong> drug in <strong>the</strong> lung, leading to<br />

higher local drug levels at <strong>the</strong> absorption site. An<br />

additional advantage is <strong>the</strong> natural tendency <strong>of</strong> <strong>the</strong><br />

nanoparticles to stick to mucosal surfaces at <strong>the</strong><br />

absorption site over an extended period <strong>of</strong> time. The<br />

longer residence time on <strong>the</strong> mucosal surface not only<br />

reduces <strong>the</strong> loss <strong>of</strong> <strong>the</strong> drug due to <strong>the</strong> clearance by<br />

cilia movement (Ponchel et al. 1997) but also contributes<br />

to a larger extent <strong>of</strong> absorption because <strong>of</strong> an<br />

increased dissolution time compared with microparticles.<br />

In addition, nanosuspensions generally possess<br />

a very low fraction <strong>of</strong> microparticles (Gassmann<br />

et al. 1994), which reduces <strong>the</strong> deposition <strong>of</strong> particles<br />

in respiratory passage. Fur<strong>the</strong>rmore, compared to <strong>the</strong><br />

microparticles, nanoparticles increase <strong>the</strong> number <strong>of</strong><br />

particles in each aerosol droplet (see Fig. 6); <strong>the</strong>n a<br />

more homogenous distribution <strong>of</strong> <strong>the</strong> drug in <strong>the</strong><br />

aerosol can be obtained. This leads to a more efficient<br />

delivery to <strong>the</strong> lungs <strong>and</strong> in this way decreases local<br />

<strong>and</strong> systemic side effects <strong>of</strong> drug microparticles. A<br />

long-term stable budesonide nanosuspension which<br />

could be nebulized <strong>for</strong> pulmonary administration was<br />

prepared by Claudia Jacobs by <strong>the</strong> high-pressure<br />

homogenization technique (Jacobs <strong>and</strong> Muller 2002).<br />

Mean PCS particle size <strong>of</strong> this nanosuspension was<br />

about 500–600 nm, <strong>and</strong> analysis by laser diffraction<br />

showed that <strong>the</strong> diameters <strong>of</strong> 95% <strong>and</strong> 99% were<br />

below 3 lm. The PCS diameter be<strong>for</strong>e <strong>and</strong> after<br />

aerosolization did not change <strong>and</strong> <strong>the</strong> LD diameters<br />

increased negligibly, showing <strong>the</strong> suitability <strong>for</strong><br />

pulmonary delivery. The scale-up from 40 mL up to<br />

Fig. 6 Comparison <strong>of</strong> particle distribution in <strong>the</strong> droplets, left<br />

as a 3-lm particle, right as 500-nm particles (Jacobs <strong>and</strong><br />

Muller 2002)<br />

123


860 J Nanopart Res (2008) 10:845–862<br />

300 mL could be per<strong>for</strong>med successfully through this<br />

technology (Jacobs <strong>and</strong> Muller 2002).<br />

Surface modification <strong>and</strong> targeted delivery<br />

Nanosuspensions can be used <strong>for</strong> targeted delivery as<br />

<strong>the</strong>ir surface properties <strong>and</strong> in vivo behavior can be<br />

easily be altered by changing <strong>the</strong> used stabilizer. That<br />

is what Muller (Muller <strong>and</strong> Keck 2004) raised as <strong>the</strong><br />

concept <strong>of</strong> ‘‘differential protein absorption.’’ After<br />

intravenous injection, <strong>the</strong> particles absorb proteins<br />

from <strong>the</strong> blood; <strong>the</strong>se absorbed proteins determine <strong>the</strong><br />

in vivo fate <strong>of</strong> <strong>the</strong> particles. The qualitative <strong>and</strong><br />

quantitative composition <strong>of</strong> <strong>the</strong> proteins adsorption<br />

pattern depends on <strong>the</strong> physico-chemical surface<br />

properties <strong>of</strong> <strong>the</strong> particle, which can be adjusted by<br />

choosing a different stabilizer (Goppert <strong>and</strong> Muller<br />

2005). In <strong>the</strong> PathFinder Ò technology proposed by<br />

Muller (Muller <strong>and</strong> Keck 2004), atvaquone <strong>nanocrystals</strong><br />

were produced by high-pressure<br />

homogenization, <strong>the</strong> surface was modified with <strong>the</strong><br />

surfactant Tween 80 leading to in vivo preferential<br />

absorption <strong>of</strong> apolipoprotein E, <strong>the</strong> particles accumulated<br />

to a sufficient amount in <strong>the</strong> blood brain<br />

barrier (BBB), where <strong>the</strong> receptors <strong>of</strong> apolipoprotein<br />

E abound <strong>and</strong> <strong>the</strong>n a <strong>the</strong>rapeutic level was achieved.<br />

Natural targeting <strong>of</strong> MPS by nanosuspensions has<br />

already been described. However, <strong>the</strong> natural targeting<br />

progress could pose an obstacle when<br />

macrophages are not <strong>the</strong> desired targets. Then, to<br />

bypass <strong>the</strong> phagocytic uptake <strong>of</strong> <strong>the</strong> drug, <strong>its</strong> surface<br />

properties need to be altered, as in <strong>the</strong> case <strong>of</strong> stealth<br />

liposomes (Allen 1997). To sum up, <strong>for</strong> <strong>the</strong> targeted<br />

delivery <strong>of</strong> nanosuspensions, <strong>the</strong> development <strong>of</strong><br />

stealth <strong>nanocrystals</strong> <strong>and</strong> active targeting <strong>nanocrystals</strong><br />

modified with functionalized surface coatings will be<br />

<strong>the</strong> next focal point.<br />

Scale-up <strong>and</strong> commercialization<br />

The vitality <strong>of</strong> a pharmaceutical technique to a great<br />

extent relies on <strong>the</strong> possibility <strong>of</strong> <strong>its</strong> commercialization,<br />

or in o<strong>the</strong>r words, <strong>the</strong> large-scale production.<br />

The time period between invention <strong>of</strong> <strong>the</strong> conception<br />

<strong>and</strong> <strong>the</strong> first products on <strong>the</strong> market is much shorter<br />

<strong>for</strong> <strong>the</strong> drug <strong>nanocrystals</strong>, compared to <strong>the</strong> o<strong>the</strong>r<br />

submicron drug delivery system. For example <strong>the</strong><br />

liposomes were raised by Bingham in 1968, but <strong>the</strong><br />

123<br />

first pharmaceutical products appeared on <strong>the</strong> market<br />

at <strong>the</strong> beginning <strong>of</strong> 1900s. However, only about<br />

10 years after <strong>the</strong> appearance <strong>of</strong> <strong>the</strong> concept <strong>of</strong> <strong>the</strong><br />

drug <strong>nanocrystals</strong>, <strong>the</strong> first product Rapamune Ò was<br />

placed in <strong>the</strong> market by <strong>the</strong> company Wyeth. This<br />

rapid commercialization verifies <strong>the</strong> simple realization<br />

<strong>of</strong> <strong>the</strong> large scale production. First, <strong>the</strong> highpressure<br />

homogenization was a well-developed<br />

procedure in <strong>the</strong> pharmaceutical field be<strong>for</strong>e it was<br />

applied to <strong>the</strong> production <strong>of</strong> <strong>the</strong> drug nanosuspensions.<br />

At present, <strong>the</strong> homogenizers with a large yield<br />

are successfully brought to operation <strong>for</strong> scale-up<br />

purpose, <strong>for</strong> example, an APV Rannie 118 with a<br />

capacity <strong>of</strong> 1 ton/h <strong>and</strong> an Avestin 1000 with a<br />

capacity <strong>of</strong> 1000 L/h. In addition, <strong>for</strong> economical<br />

purposes, a continuous model has successfully been<br />

carried out in <strong>the</strong> production <strong>of</strong> <strong>the</strong> nanosuspensions,<br />

i.e. placing several homogenizations in series; this<br />

increases <strong>the</strong> production rate by many folds. For<br />

example, assuming a homogenizer with a capacity <strong>of</strong><br />

1000 kg/h, <strong>the</strong> production <strong>of</strong> one-ton nanosuspension<br />

would require 20 h <strong>and</strong> 20 cycles. In case <strong>of</strong> four<br />

homogenizers in series, this would reduce to 5 h <strong>for</strong><br />

<strong>the</strong> production <strong>of</strong> 1000-kg nanosuspensions.<br />

Conclusion <strong>and</strong> perspective<br />

<strong>Drug</strong> <strong>nanocrystals</strong>, regardless <strong>of</strong> <strong>the</strong>ir production<br />

method, can be applied to all <strong>the</strong> <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong><br />

to overcome <strong>the</strong> solubility <strong>and</strong> bioavailability problems,<br />

because all <strong>the</strong> <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong> can be<br />

disintegrated into <strong>nanocrystals</strong>. In some cases,<br />

besides improving drug dissolution, drug <strong>nanocrystals</strong><br />

also show o<strong>the</strong>r biological activities, such as<br />

realization <strong>of</strong> a sustained release <strong>and</strong> targeting to <strong>the</strong><br />

special tissues or organs. An important advantage <strong>of</strong><br />

<strong>the</strong> drug <strong>nanocrystals</strong> is that <strong>the</strong>y can be applied to<br />

various administration routes, such as oral, parenteral,<br />

ocular, <strong>and</strong> pulmonary delivery, <strong>and</strong> have<br />

shown great superiority over <strong>the</strong> counterparts <strong>of</strong> <strong>the</strong><br />

traditional <strong>for</strong>mulation products in every administration<br />

route. In addition, combined with <strong>the</strong> traditional<br />

<strong>for</strong>mulation approaches, drug <strong>nanocrystals</strong> also can<br />

be trans<strong>for</strong>med into various dosage <strong>for</strong>ms, such as<br />

tablets, capsules, injections, aerosols, <strong>and</strong> so on.<br />

Production techniques such as high-pressure homogenization<br />

have been successfully employed <strong>for</strong><br />

large-scale production <strong>of</strong> drug <strong>nanocrystals</strong>. So far,


J Nanopart Res (2008) 10:845–862 861<br />

many <strong>nanocrystals</strong> products have been placed in <strong>the</strong><br />

market (Keck <strong>and</strong> Muller 2006). Due to <strong>the</strong> unique<br />

advantage <strong>and</strong> pharmaeconomical value, drug <strong>nanocrystals</strong><br />

are paid increasingly more attention as a<br />

promising approach. In <strong>the</strong> future, <strong>the</strong> development <strong>of</strong><br />

stealth <strong>nanocrystals</strong> <strong>and</strong> active targeting <strong>nanocrystals</strong><br />

modified with functionalized surface coatings will be<br />

<strong>the</strong> next focal point.<br />

References<br />

Allen TM (1997) Liposomes: opportunities in drug delivery.<br />

<strong>Drug</strong>s 54:8–14<br />

Annapragada A, Adjei A (1996) Numerical simulation <strong>of</strong><br />

milling processes as an aid to process design. Int J Pharm<br />

136:1–11<br />

Blunk T, Hochstrasser DF, Luck MA et al (1996) Kinetics <strong>of</strong><br />

plasma protein absorption on model particles <strong>for</strong> controlled<br />

drug delivery <strong>and</strong> drug targeting. Eur J Pharm<br />

Biopharm 42:262–268<br />

Böhm BHL, Müller RH (1999) Lab-scale production unit<br />

design <strong>for</strong> nanosuspensions <strong>of</strong> sparingly <strong>soluble</strong> cytotoxic<br />

<strong>drugs</strong>. Pharm Sci Tech Today 2:336–339<br />

Borner K, Hartwig H, Leitzke S et al (1999) HPLC determination<br />

<strong>of</strong> cl<strong>of</strong>azimine in tissues <strong>and</strong> serum <strong>of</strong> mice after<br />

intravenous administration <strong>of</strong> nanocrystalline or liposomal<br />

<strong>for</strong>mulations. Int J Pharm 11:75–79<br />

Choi JY, Yoo JY, Kwak HS et al (2005) Role <strong>of</strong> polymeric<br />

stabilizers <strong>for</strong> drug nanocrystal dispersions. Curr Appl<br />

Phys 5:472–474<br />

Douroumis D, Fahr A (2006) Nano- <strong>and</strong> micro-particulate<br />

<strong>for</strong>mulations <strong>of</strong> <strong>poorly</strong> water-<strong>soluble</strong> <strong>drugs</strong> by using a<br />

novel optimized technique. Eur J Pharm Biopharm<br />

63:173–175<br />

Dressman JB, Amidon GL, Reppas C et al (1998) Dissolution<br />

testing as a prognostic tool <strong>for</strong> oral drug absorption:<br />

immediate release dosage <strong>for</strong>ms. Pharm Res 15:11–22<br />

Drews TO, Tsapatsis M (2007) Model <strong>of</strong> <strong>the</strong> evolution <strong>of</strong><br />

nanoparticles to crystals via an aggregative growth<br />

mechanism. Micropor Mesopor Mat 101:97–107<br />

Duchene D, Ponchel G (1997) Bioadhesion <strong>of</strong> solid oral dosage<br />

<strong>for</strong>ms, why <strong>and</strong> how? Eur J Pharm Biopharm 44:15–23<br />

Gassmann P, List M, Schweitzer A et al (1994) Hydrosolsalternatives<br />

<strong>for</strong> <strong>the</strong> parenteral application <strong>of</strong> <strong>poorly</strong> water<br />

<strong>soluble</strong> <strong>drugs</strong>. Eur J Pharm Biopharm 40:64–72<br />

Geze A, Julienne M, Mathieu D et al (1999) Development <strong>of</strong><br />

5-iodo-2%-deoxyuridine milling process to reduce initial<br />

burst release from PLGA microparticles. Int J Pharm<br />

178:257–268<br />

Goppert TM, Muller RH (2005) Adsorption kinetics <strong>of</strong> plasma<br />

proteins on solid lipid nanoparticles <strong>for</strong> drug targeting. Int<br />

J Pharm 302:172–186<br />

Grau MJ, Kayser O, Muller RH (2000) Nanosuspensions <strong>of</strong><br />

<strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong> – reproducibility <strong>of</strong> small scale<br />

production. Int J Pharm 196:155–157<br />

Gubskaya A, Lisnyak Y, Blagoy Y (1995) Effect <strong>of</strong> cryogrinding<br />

on physico-chemical properties <strong>of</strong> <strong>drugs</strong>. I.<br />

Theophylline: evaluation <strong>of</strong> particle sizes <strong>and</strong> <strong>the</strong> degree<br />

<strong>of</strong> crystallinity, relation to dissolution parameters. <strong>Drug</strong><br />

Dev Ind Pharm 21:1953–1964<br />

Hancock BC, Carlson GT, Ladipo DD et al (2002) Comparison<br />

<strong>of</strong> <strong>the</strong> mechanical properties <strong>of</strong> <strong>the</strong> crystalline<br />

<strong>and</strong> amorphous <strong>for</strong>ms <strong>of</strong> a drug substance. Inter J Pharm<br />

241:73–85<br />

Hecq J, Deleers M, Fanara D et al (2005) Preparation <strong>and</strong><br />

characterization <strong>of</strong> <strong>nanocrystals</strong> <strong>for</strong> solubility <strong>and</strong> dissolution<br />

rate enhancement <strong>of</strong> nifedipine. Int J Pharm<br />

299:167–177<br />

Hecq J, Deleers M, Fanara D et al (2006) Preparation <strong>and</strong><br />

in vitro/in vivo evaluation <strong>of</strong> nano-sized crystals <strong>for</strong> dissolution<br />

rate enhancement <strong>of</strong> ucb-35440–3, a highly dosed<br />

<strong>poorly</strong> water-<strong>soluble</strong> weak base. Eur J Pharm Biopharm<br />

64:360–368<br />

Jacobs C, Kayser O, Muller RH (2000) Nanosuspensions as a<br />

new approach <strong>for</strong> <strong>the</strong> <strong>for</strong>mulation <strong>for</strong> <strong>the</strong> <strong>poorly</strong> <strong>soluble</strong><br />

drug tarazepide. Int J Pharm 196:161–164<br />

Jacobs C, Kayser O, Muller RH (2001) Production <strong>and</strong> characterisation<br />

<strong>of</strong> mucoadhesive nanosuspensions <strong>for</strong> <strong>the</strong><br />

<strong>for</strong>mulation <strong>of</strong> buparvaquone. Int J Pharm 214:3–7<br />

Jacobs C, Muller RH (2002) Production <strong>and</strong> characterization <strong>of</strong><br />

a budesonide nanosuspension <strong>for</strong> pulmonary administration.<br />

Pharm Res 19:189–194<br />

Jinno J, Kamada N, Miyake M et al (2006) Effect <strong>of</strong> particle<br />

size reduction on dissolution <strong>and</strong> oral absorption <strong>of</strong> a<br />

<strong>poorly</strong> water-<strong>soluble</strong> drug, cilostazol, in beagle dogs.<br />

J Control Release 111:56–64<br />

Kayser O (2001) A new approach <strong>for</strong> targeting to Cryptosporidium<br />

parvum using mucoadhesive nanosuspensions:<br />

research <strong>and</strong> applications. Int J Pharm 214:83–85<br />

Kayser O, Olbrich C, Yardley V et al (2003) Formulation <strong>of</strong><br />

amphotericin B as nanosuspension <strong>for</strong> oral administration.<br />

Int J Pharm 254:73–75<br />

Keck C, Muller RH (2006) <strong>Drug</strong> <strong>nanocrystals</strong> <strong>of</strong> <strong>poorly</strong> <strong>soluble</strong><br />

<strong>drugs</strong> produced by high pressure homogenisation. Eur J<br />

Pharm Biopharm 62:3–16<br />

Kipp JE (2004) The role <strong>of</strong> solid nanoparticle technology in <strong>the</strong><br />

parenteral delivery <strong>of</strong> <strong>poorly</strong> water-<strong>soluble</strong> <strong>drugs</strong>. Int J<br />

Pharm 284:109–122<br />

Kocbek P, Baumgartner S, Kristl J. (2006) Preparation <strong>and</strong><br />

evaluation <strong>of</strong> nanosuspensions <strong>for</strong> enhancing <strong>the</strong> dissolution<br />

<strong>of</strong> <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong>. Int J Pharm 312:179–186<br />

Krause KP, Kayser O, Mader K et al (2000) Heavy metal<br />

contamination <strong>of</strong> nanosuspensions produced by highpressure<br />

homogenisation. Int J Pharm 196:169–172<br />

Krause KP, Muller RH (2001) Production <strong>and</strong> characterisation<br />

<strong>of</strong> highly concentrated nanosuspensions by high pressure<br />

homogenisation. Int J Pharm 214:21–24<br />

Lee J, Lee SJ, Choi JY et al (2005) Amphiphilic amino acid<br />

copolymers as stabilizers <strong>for</strong> <strong>the</strong> preparation <strong>of</strong> nanocrystal<br />

dispersion. Eur J Pharm Sci 24:441–449<br />

Li YCh, Dong L, Jiaa A et al (2006) Preparation <strong>and</strong> characterization<br />

<strong>of</strong> solid lipid nanoparticles loaded traditional<br />

chinese medicine. Int J Biol Macromol 38:296–299<br />

Liedtke S, Wissing S, Muller RH et al (2000) Influence <strong>of</strong> high<br />

pressure homogenization equipment on nanodispersions<br />

characteristics. Int J Pharm 196:183–185<br />

Lipinski C (2002) Poor aqueous solubility – an industry wide<br />

problem in drug discovery. Am Pharm Rev 5:82–85<br />

123


862 J Nanopart Res (2008) 10:845–862<br />

Liversidge EM, Liversidge GG, Cooper ER (2003) Nanosizing:<br />

a <strong>for</strong>mulation approach <strong>for</strong> <strong>poorly</strong>-water-<strong>soluble</strong> compounds.<br />

Eur J Pharm Sci 18:113–120<br />

Liversidge GG, Conzentino P (1995) <strong>Drug</strong> particle size<br />

reduction <strong>for</strong> decreasing gastric irritancy <strong>and</strong> enhancing<br />

absorption <strong>of</strong> naproxen in rats. Int J Pharm 125:309–313<br />

Liversidge GG, Cundy KC (1995) Particle size reduction <strong>for</strong><br />

improvement <strong>of</strong> oral bioavailability <strong>of</strong> hydrophobic <strong>drugs</strong>:<br />

I. Absolute oral bioavailability <strong>of</strong> nanocrystalline danazol<br />

in beagle dogs. Int J Pharm 125:91–97<br />

Liversidge GG, Cundy KC, Bishop J, et al (1991) Surface<br />

modified drug nanoparticles. US Patent 5145684, 8 Sept<br />

1992<br />

Mantzaris NV (2005) Liquid-phase syn<strong>the</strong>sis <strong>of</strong> nanoparticles:<br />

particle size distribution dynamics <strong>and</strong> control. Chem Eng<br />

Sci 60:4749–4770<br />

Mohammed AR, Weston N, Coombes AGA et al (2004)<br />

Liposome <strong>for</strong>mulation <strong>of</strong> <strong>poorly</strong> water <strong>soluble</strong> <strong>drugs</strong>:<br />

optimisation <strong>of</strong> drug loading <strong>and</strong> ESEM analysis <strong>of</strong> stability.<br />

Int J Pharm 285:23–34<br />

Mosharraf M, Nystrom C (1995) The effect <strong>of</strong> particle size <strong>and</strong><br />

shape on <strong>the</strong> surface specific dissolution rate <strong>of</strong> micronized<br />

practically in<strong>soluble</strong> <strong>drugs</strong>. Int J Pharm 122:35–47<br />

Muller BW, Muller RH (1984) Particle size analysis <strong>of</strong> latex<br />

suspensions <strong>and</strong> microemulsions by photon correlation<br />

spectroscopy. J Pharm Sci 73:915–918<br />

Muller RH, Becker R, Kruss B, et al (1998) Pharmaceutical<br />

nanosuspensions <strong>for</strong> medicament administration as system<br />

<strong>of</strong> increased saturation solubility <strong>and</strong> rate <strong>of</strong> solution. US<br />

Patent 5858410, 12 Jan 1999<br />

Muller RH, Jacobs C (2002) Buparvaquone mucoadhesive<br />

nanosuspension: preparation, optimisation <strong>and</strong> long-term<br />

stability. Int J Pharm 237:151–161<br />

Muller RH, Keck CM (2004) Challenges <strong>and</strong> solutions <strong>for</strong> <strong>the</strong><br />

delivery <strong>of</strong> biotech <strong>drugs</strong> – a review <strong>of</strong> drug nanocrystal<br />

technology <strong>and</strong> lipid nanoparticles. J Biotechnol 113:<br />

151–170<br />

Muller RH, Peters K (1998) Nanosuspensions <strong>for</strong> <strong>the</strong> <strong>for</strong>mulation<br />

<strong>of</strong> <strong>poorly</strong> <strong>soluble</strong> <strong>drugs</strong> I. Preparation by a sizereduction<br />

technique. Int J Pharm 160:229–237<br />

Muller RH, Rungea S, Ravelli V et al (2006) Thunemannc <strong>and</strong><br />

E. B. Souto. Oral bioavailability <strong>of</strong> cyclosporine: solid<br />

lipid nanoparticles (SLN Ò ) versus drug <strong>nanocrystals</strong>. Int J<br />

Pharm 317:82–89<br />

Na GC, Stevens HJ, Yuan B et al (1999) Physical stability <strong>of</strong><br />

ethyl diatrizoate nanocrystalline suspension in steam<br />

sterilization. Pharm Res 16:569–574<br />

Peters K, Leitzke S, Diederichs JE et al (2000) Preparation <strong>of</strong> a<br />

cl<strong>of</strong>azimine nanosuspension <strong>for</strong> intravenous use <strong>and</strong><br />

123<br />

evaluation <strong>of</strong> <strong>its</strong> <strong>the</strong>rapeutic efficacy in murine Mycobacterium<br />

avium infection. J Antimicrob Chemoth 45:<br />

77–83<br />

Ponchel G, Montisci MJ, Dembri A et al (1997) Mucoadhesion<br />

<strong>of</strong> colloidal particulate systems in <strong>the</strong> gasrointesinal tract.<br />

Eur J Pharm Biopharm 4:25–31<br />

Rabinow BE (2004) Nanosuspensions in drug delivery. Nat<br />

Rev <strong>Drug</strong> Discov 3:785–796<br />

Rasenack N, Muller BW (2003) Micro-size drug particles:<br />

common <strong>and</strong> novel micronization techniques. Pharm Dev<br />

Technol 9:1–13<br />

Rosario P, Claudio B, Piera F et al (2002) Eudragit RS100<br />

nanosuspensions <strong>for</strong> <strong>the</strong> ophthalmic controlled delivery <strong>of</strong><br />

ibupr<strong>of</strong>en. Eur J Pharm Sci 16:53–61<br />

Schmidt S, Müller RH (2003) Plasma protein adsorption patterns<br />

on surfaces <strong>of</strong> Amphotericin B-containing fat<br />

emulsions. Int J Pharm 254:3–5<br />

Schwitzera JM, Achleitner G, Pomperb H et al (2004) Development<br />

<strong>of</strong> an intravenously injectable chemically stable<br />

aqueous omeprazole <strong>for</strong>mulation using nanosuspension<br />

technology. Eur Pharm Biopharma 58:615–619<br />

Siostrom B, Kronberg B, Carl<strong>for</strong>s J (1993) A method <strong>for</strong> <strong>the</strong><br />

preparation <strong>of</strong> submicron particles <strong>of</strong> sparingly water<strong>soluble</strong><br />

<strong>drugs</strong> by precipitation in oil-in-water emulsions. 1.<br />

influence <strong>of</strong> emulsification <strong>and</strong> surfactant concentration.<br />

J Pharm Sic 82:579–583<br />

Wallis KH, Muller RH (1993) Determination <strong>of</strong> <strong>the</strong> surface<br />

hydrophobicity <strong>of</strong> colloidal dispersions by mini-hydrophobic<br />

interaction chromatography. Pharm Ind 55:1124–<br />

1128<br />

Xing CS (2004) Preparation <strong>and</strong> physico-chemical characterization<br />

<strong>of</strong> nanoparticles. In: Xu BH (ed) Nano-medicine.<br />

Tsinghua University Publishers, pp 9–10<br />

Zhang DR, Tan TW, Gao L et al (2007) Preparation <strong>of</strong><br />

Azithromycin Nanosuspensions by High Pressure<br />

Homogenization <strong>and</strong> <strong>its</strong> Physicochemical Characteristics<br />

Studies. <strong>Drug</strong> Dev Ind Pharm 33:569–575<br />

Zhong J, Shen Zh, Yang Y et al (2005) Preparation <strong>and</strong> characterization<br />

<strong>of</strong> uni<strong>for</strong>m nanosized cephradine by<br />

combination <strong>of</strong> reactive precipitation <strong>and</strong> liquid anti-solvent<br />

precipitation under high gravity environment. Int J<br />

Pharm 301:286–293<br />

Zhang J, Shen Zh, Zhong J et al (2006) Preparation <strong>of</strong> amorphous<br />

cefuroxime axetil nanoparticles by controlled<br />

nanoprecipitation method without surfactants. Int J Pharm<br />

323:153–160

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!