COREL-07149; No of Pages 6 Journal of Controlled Release xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Journal of Controlled Release journal homepage: www.elsevier.com/locate/jconrel

3Q1

Davide Brambilla, Paola Luciani, Jean-Christophe Leroux ⁎

4

Institute of Pharmaceutical Sciences, Department of Chemistry and Applied Biosciences, ETH Zurich, Vladimir-Prelog-Weg 1-5/10, 8093 Zurich, Switzerland

5

a r t i c l e

6 7 8 9 Q2 10

Article history: Received 8 February 2014 Accepted 21 March 2014 Available online xxxx

a b s t r a c t

P

What if we could open the 2044 special issue of the Journal of Controlled Release? Which drug delivery technologies will have led the field? Which ones will have successfully reached the marketplace? In attempting to answer these questions, we selected a few recent technologies and concepts that could, in our opinion, play a crucial role in coming years. In each case, emblematic papers are cited to introduce and discuss the selected topic. © 2014 Published by Elsevier B.V.

. . . . . . . . . . . .

. . . . . . . . . . . .

R

R

34

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

1. Introduction

36 37

In this special anniversary issue, the Journal of Controlled Release takes readers on a spatio-temporal journey through the world of controlled release, directly escorted by main actors in the field. The objective of this manuscript is to provide a personal viewpoint on drug delivery topics that could play an important role in the near future. We are aware that this is a delicate and subjective exercise, but are confident that some of the presented technologies and concepts in their current or modified form will still be present 30 years from now. An analysis of the literature published in the last 5 years might shed light on rising trends in controlled drug release. By entering the words “drug delivery” in Scopus, the top cited original papers in the last 5-year period (2008–2013) deal with functional nanocarriers [1–5]. More generally,

40 41 42 43 44 45 46 47

U

35

38 39

⁎ Corresponding author. Tel.: +41 446337310. E-mail address: [email protected] (J.-C. Leroux).

. . . . . . . . . . . .

T

. . . . . . . . . . . .

C

Introduction . . . . . . . . . . . . . . . . . . . EPR enhancement . . . . . . . . . . . . . . . . Beyond EPR: extracorporeally-guided drug delivery . Biotechnological approaches to drug delivery . . . . 4.1. Cell-derived membrane vesicles (CMVs) . . . 4.2. Cells as in vivo drug bioreactors . . . . . . . 4.3. Cells as carriers of oncolytic viruses . . . . . 5. “Computer-assisted” drug delivery . . . . . . . . . 6. Transdermal delivery, the next generation of patches. 7. Conclusions . . . . . . . . . . . . . . . . . . . Acknowledgments . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . .

E

1. 2. 3. 4.

E

Contents

N C O

22 23 24 25 26 27 28 29 30 31 32 33

11 12 13 14 15

D

19 21

O

R O

i n f o

F

2

Breakthrough discoveries in drug delivery technologies: The next 30 years

1

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . .

0 0 0 0 0 0 0 0 0 0 0 0

nanoparticle-mediated drug delivery thoroughly dominated the research landscape in recent decades [6–10]. Unfortunately, despite its exciting translational potential, this great scientific effort resulted in a mitigated clinical success. With relatively few products on the market, there is a discrepancy between the amount of research devoted to the topic and its translation to patients, leading some observers to call for trend reversal [11,12]. However, on the timescale of pharmaceutical research, the field of nanotechnology is still young, and it is not surprising that more time is needed to surmount all obstacles to commercialization. Delays in reaching the clinical stage can partly be attributed to the fact that functional nanocarriers are often quite sophisticated in their design and difficult to characterize from a pharmaceutical and toxicological viewpoint. Sensibly, most of the work has been devoted to cancer chemotherapy, and the major achievement of nanocarriers so far has been to reduce systemic drug toxicity (e.g. doxorubicin cardiotoxicity), while their success in improving therapeutic efficacy is more debatable [13]. The tumoral

http://dx.doi.org/10.1016/j.jconrel.2014.03.056 0168-3659/© 2014 Published by Elsevier B.V.

Please cite this article as: D. Brambilla, et al., Breakthrough discoveries in drug delivery technologies: The next 30 years, J. Control. Release (2014), http://dx.doi.org/10.1016/j.jconrel.2014.03.056

48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64

90 91 92 93

98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128

C

88 89

E

86 87

R

84 85

R

82 83

O

80 81

C

78 79

137

The limitations of nanoparticle-mediated drug delivery become clear when organs other than the liver, spleen or the vascular system are targeted. Even when the EPR effect occurs, and when active targeting moieties (antibodies, peptides, aptamers) are employed, total amounts of drugs that reach desired tissues remain relatively low [30]. Achieving spatial control of carrier movements in the body could be instrumental in favoring carrier accumulation at sites of interest [31]. Extracorporeal devices to guide nanocarriers in real time – and trigger the release of their cargo in specific locations – are particularly attractive. Magnetic resonance navigation (MRN) has recently been investigated to steer particles for tumor chemo-embolization [32]. Magnetic microparticles loaded with a cytotoxic anticancer agent were steered transversally in the blood flow of rabbits by a modified magnetic resonance imaging system to embolize specific liver sections (Fig. 1). This strategy differs from conventional magnetic targeting [33,34], as it follows and precisely guides magnetically-responsive carriers to deep tissues. MRN is only in the proof-of-concept stage, and will have to be optimized for smaller particles and applications other than embolization. Interestingly, recent reports have described the magnetic steering of intact cells, with different therapeutic functions, to target tissues [35,36]. The combination of cell therapeutics (vide infra) with advanced guidance strategies could represent major advance in drug delivery and tissue engineering.

138

4. Biotechnological approaches to drug delivery

161

The biotechnology revolution remarkably altered the pharmaceutical landscape in the last 30 years. It stimulated the development of potent and specific drugs, paving the way for personalized medicine. As a result, biopharmaceuticals now represent about 30% of new medicines brought to market [37]. The area of controlled release has certainly benefited from progress in biotechnology, and different bio-derived/inspired targeting systems are being proposed to achieve favorable drug biodistribution. We have selected 3 particularly interesting strategies to exemplify the potential of biotechnology in drug delivery (Fig. 2).

162 163

N

76 77

3. Beyond EPR: extracorporeally-guided drug delivery

U

74 75

F

The EPR effect, first postulated by Matsumura and Maeda in the 1980s [15], represents the main tumor targeting principle of nanocarriers and macromolecules. It has been well-described in rodent models [16], and is the very basis for the clinical evaluation of most nanosized anticancer drug delivery systems, including those that specifically recognize solid tumor cells with targeting ligands. Recently, the relevance of EPR in a number of human tumors has been questioned partly because of its high clinical failure rate, emphasizing the need for improvement [17,18]. The chaotic organization of excessive and leaky tumor vessels is one of the factors that hinders nanocarrier penetration into tumoral matrices, leading to heterogeneous blood perfusion and elevated interstitial fluid pressure [19]. Indeed, it has been determined that tumor perfusion is important for efficient nanocarrier accumulation [20]. Understanding EPR in humans will allow the implementation of strategies that could enhance this effect via modulation of the tumor environment. Anti-angiogenic therapies, originally conceived to correct excessive neo-angiogenesis signaling and reduce tumor blood supply, have the potential to repair abnormalities in vascular functions, facilitating convective penetration of molecules within the tumor parenchyma [19]. EPR modulation by pharmacological means was first demonstrated with angiotensin-II [21] and then confirmed with vasoactive molecules (e.g. TβR-I inhibitors, nitroglycerin, heme oxigenase-1, telmisartan) [22–25]. In 2013, the group of Jain reported that normalization of the tumor microenvironment by vascular endothelial growth factor (VEGF) receptor inhibition enhanced the efflux of smaller particles (12 nm), whereas no effect was observed with larger particles (up to 250 nm). The authors also conceived a mathematical model to elucidate and predict the effect. These findings were reproduced in mice with the commercial nanomedicines Doxil® (100 nm) and Abraxane® (10 nm). VEGF inhibition enhanced Abraxane's effectiveness but did not impact Doxil's antitumor activity [26]. This work demonstrated the worth of tumor/formulation-specific vascular normalization in cancer targeting.

72 73

O

97

71

R O

2. EPR enhancement

69 70

129 130

P

96

67 68

Concomitantly with pharmacological approaches, consistent research has lately been devoted to physical methods, such as hyperthermia or ultrasound-induced cavitation, to improve tumor mass penetration by particles, macromolecules and oncolytic viruses [27–29]. Interestingly, these methods might also facilitate the spatio-temporal control of drug release. Overall, such investigation will hopefully provide a crucial boost to nanomedicine's translation to the clinic and foster studies of related strategies that could enhance nanodrug efficacy.

T

94 95

microenvironment is complex, with strong inter/intra-tumor heterogeneity influencing the effectiveness of anticancer drug delivery systems. Systematic quantitative analyses of extravasation and the intratumoral distribution of differentially-formulated drugs will certainly be performed extensively in the next decade. On the heels of the bioinformatics revolution in cancer typing, a visionary goal would be the implementation of accessible and reliable databases listing the main features (size, shape, surface chemistry, etc.) required for nanocarriers to efficiently reach (and disseminate within) precise tumor classes. Confident that nanoparticles (with their “unlimited” variations and improvements) and other well-established technologies will continue to play a fundamental role in the future and will be comprehensively discussed in this special issue, we decided to focus our attention more on recent trend-setting drug delivery principles than on nanoparticle technologies per se. Enhanced permeation and retention (EPR) remains the governing principle with most nanocarrier systems. However, the EPR effect, which is often inadequately characterized in humans, may be insufficient for high drug deposition at tumoral sites in a number of neoplasias. Therefore, we believe that approaches aimed at better exploiting and increasing this effect are of paramount importance and will therefore be discussed here. We will also present extracorporeal tools to precisely guide drug carriers in the body in real time, introduce biotechnological systems that take advantage of Nature's optimized machinery for targeting purposes, and discuss the possible impact of bioengineering and computer sciences on controlled drug delivery. A final topic that will be addressed is the rapid development of transdermal microneedle patches. Although this simple, ground-breaking technology was reported as early as the 1970s [14], it only recently unveiled its full potential. The approach is expected to gain further popularity, especially when considering the value of worldwide accessibility to vaccination programs.

D

65 66

D. Brambilla et al. / Journal of Controlled Release xxx (2014) xxx–xxx

E

2

Fig. 1. Schematic representation of magnetic-based extracorporeal navigation of drug delivery to the hepatic left lobe. Adapted from [32] with permission.

Please cite this article as: D. Brambilla, et al., Breakthrough discoveries in drug delivery technologies: The next 30 years, J. Control. Release (2014), http://dx.doi.org/10.1016/j.jconrel.2014.03.056

131 132 133 134 135 136

139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160

164 165 166 167 168 169 170 171

D. Brambilla et al. / Journal of Controlled Release xxx (2014) xxx–xxx

3

214 215

4.3. Cells as carriers of oncolytic viruses

223

Oncolytic viruses preferentially infect and kill tumor cells without harming normal tissues. This specificity could be a natural feature of viruses or a result of genetic engineering [49]. However, the administration of free oncolytic viruses has two main limitations: (i) immune responses to the virus capsid, and (ii) poor preferential migration to tumors. Generally, it is believed that around 0.001% of injected viruses reach target tissues, a proportion that can be improved several folds by deploying specific cells as carriers [50]. By mimicking the natural ability of certain viruses to exploit circulating cells to migrate and invade target cell populations, it is possible to enhance the targeting efficiency of oncolytic viruses via their incorporation into cell carriers. Qiao et al. [51] investigated autologous T cells that naturally reside in lymph nodes and lymphoid organs, to carry therapeutic oncolytic viruses and purge metastatic cells from tissues. They demonstrated that adoptive transfer of T cells loaded with vesicular stomatitis virus reduced/ eradicated lymph node-resistant metastatic cells from draining lymph nodes and spleen in mice. Importantly, similar results were obtained in virus-immune animals, although the viral load had to be carefully selected. Such an approach could be applicable to various cancers in which malignant metastases spread to distal areas via lymphoid organs. The technology has been adopted in other studies to target different tissues, and readers are referred to a recent review by Roy and Bell [52] for more information on this fascinating topic. Among the challenges associated with such therapies are the selection of the best virus/cell combination and its clinical translation. In this regard, a study describing the use of human adipose tissue cells from patients to derive tumor-homing MSCs targeting ovarian cancer was published in 2013. The authors proposed a protocol to obtain, expand, and load clinically-relevant amounts of cells with the measles virus within 14 days and showed their capacity to migrate to ovarian cancer cells in mice [53].

224

5. “Computer-assisted” drug delivery

255

Combination of bioengineering, informatics and material sciences recently helped shape new, promising drug delivery concepts. Indeed, aside from data storage and management, computer science and bioelectronics are expected to drastically influence diagnostic and therapeutic protocols. Large investments have lately been made to develop miniaturized diagnostic biosensors for real time disease tracking and prediction [54]. Implantable microdevices for controlled drug release are emblematic examples of the effort devoted to implement bioelectronics in drug delivery [55,56]. These “clever” devices can enable the control of dose amount, time, location, and drug-release rate. A recent paper by Langer's group and microCHIPS, Inc. described the first human test of a breakthrough technology initially conceived about a decade ago (Fig. 3) [57]. In this work, teriparatide, a drug usually administered daily via subcutaneous injection to treat osteoporosis, was delivered in 8 post-menopausal women by implanted microchip consisting of 20 individually-sealed reservoirs. To control drug release, the reservoirs were capped with a thin metal layer which melted when a small electrical current was applied. Importantly, the device was equipped with a miniaturized, wireless controller, allowing bidirectional communication with a computer to schedule drug liberation.

256

O

F

conditions could represent safe rescue approaches. In this regard, Altaner and coworkers [47,48] proposed an interesting variant of this approach, known as “prodrug enzyme gene therapy”. MSCs derived from human adipose tissue were engineered to express cytosine deaminase–uracil phosphotransferase, which converts the inactive prodrug 5-fluorocytosine to highly-cytotoxic 5-fluorouracil. At tumor sites, the enzyme locally produced 5-fluorouracil from the prodrug, resulting in complete tumor regression in up to 50% of treated animals. This approach was tested successfully on different tumor models.

196

4.2. Cells as in vivo drug bioreactors

197 198

Cells represent potentially powerful drug transporters in view of their multiple functions and ability to respond to the environment and exogenous triggers. Three main classes of cells (transformed, immune and progenitor) have been proposed so far for the transport of various drug classes. Particular attention has been paid to mesenchymal stem cells (MSCs) in the treatment of cancer. These cells exhibit pathotropic (tumor microenvironment, metastasis and ischemia) properties, although the mechanisms of migration are not fully understood [43,44]. MSCs have also been engineered to operate as in vivo bioreactors for the sustained production and release of therapeutic molecules at action sites. This concept has been applied to produce different kinds of anticancer biological drugs: interleukins, pro-apoptotic proteins, growth factor antagonists, etc. [45,46]. However, engineering of cell vectors to construct bioactive compounds could be hazardous to hosts if they remain active after cancer eradication. The integration of human-inducible suicide genes or regulatable promoters to activate gene production only under desired

187 188 189 190 191 192 193 194

199 200 201 202 203 204 205 206 207 208 209 210 211 212 213

C

185 186

E

183 184

R

181 182

R

179 180

N C O

177 178

U

175 176

P

195

CMVs are important tools for cell-to-cell material/information transfer, making them very attractive drug carriers. Different CMV subsets exist, and are usually categorized according to their intracellular origin. Exosomes, which derive from multivesicular bodies, are certainly the most investigated CMVs for drug delivery purposes [38]. One of the main advantages of autologous carriers is the expected absence of immunogenicity. In a breakthrough work, Alvarez-Erviti et al. [39] demonstrated that exosomes could overcome the blood–brain-barrier. Mouse-harvested immature dendritic cells served as sources of “selfexosomes” for autologous administration. The targeting ligand was obtained by fusing a brain-specific peptide coding sequence (encoding a peptide able to bind acetylcholine receptors) to the gene encoding an abundant protein in exosome membranes. Exosomes were loaded with glyceraldehyde-3-phosphate dehydrogenase (GAPDH)-specific small interfering RNA (siRNA). These siRNA-loaded exosomes significantly/specifically reduced mRNA and GAPDH expression within the brain after systemic injection. Although peptide targeting was shown to be necessary to generate RNAi responses within the brain, the blood–brain-barrier crossing mechanism remains unknown at the moment. Other papers have reported the relevance of CMVs in drug delivery, and databases on lipids, proteins and nucleic acids composing the CMVs (Exocarta, Vesiclepedia) epitomize the rising attention being paid to endogenous vesicles [40–42].

D

173 174

E

4.1. Cell-derived membrane vesicles (CMVs)

T

172

R O

Fig. 2. Schematic representation of the 3 cell-based biotechnological approaches to drug delivery. A) Cells as in vivo drug bioreactor; B) Cells as carriers of oncolytic viruses; C) Cell-derived membrane vesicles.

Please cite this article as: D. Brambilla, et al., Breakthrough discoveries in drug delivery technologies: The next 30 years, J. Control. Release (2014), http://dx.doi.org/10.1016/j.jconrel.2014.03.056

216 217 218 219 220 221 222

225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254

257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275

D. Brambilla et al. / Journal of Controlled Release xxx (2014) xxx–xxx

320 321

Inadequate vaccine coverage across developing countries is responsible every year for millions of casualties caused mainly by the relatively high costs of vaccination programs (healthcare personnel, vaccine stability, etc.) [61,62]. The quest for an inexpensive way to reach more patients worldwide has in recent years boosted the field of non-invasive transdermal delivery. The presence of dendritic cells in the dermis triggers strong immune responses, making transdermal administration an appealing route for vaccination [63]. A significant step forward lies in the optimization of transdermal microneedle patches [64]. Microneedles as safe and pain-free alternatives to conventional injection modes were introduced decades ago [65], but the most recent examples are perhaps the most relevant. In the milestone paper published by Prausnitz and coworkers [66], a new concept was proposed for the transdermal delivery of influenza virus vaccine. The system, which consisted of poly(N-vinylpyrrolidone) microneedles, dissolved within minutes under the skin after administration, eliminating the production of biohazardous waste while retaining the delivery properties of vaccines in lyophilized form. Humoral and cellular responses equaling or even outperforming those obtained by intramuscular injection were achieved in mice. Such a result, together with the practical advantage of not needing vaccine reconstitution prior to administration, makes this contribution highly promising. Interesting developments arose from this original concept. Microneedles were equipped with a poly(vinyl alcohol)-based detachable arrowhead that allowed the metallic section to be discarded a few seconds after application to the skin [67]. Detachable chitosan matrices embedded with a model antigen, ovalbumin, were designed recently and tested in vivo in rats as warheadlike parts of a microneedle platform for deposit under the skin via poly(N-vinylpyrrolidone)-coated, poly(lactic acid)-supporting arrays [68]. Poly(lactic acid) microneedles coated with adenoviral vectors dispersed in a sucrose–sugar-glass matrix showed translational potential in recent successful trials in non-human primates [69]. Transdermal delivery via microneedle patches is not limited to vaccines, and holds much hope for the administration of potent drugs with low bioavailability and biomacromolecules that are poorly permeable through the skin.

322

7. Conclusions

323

This review reflects our opinion on plausible, leading directions in drug delivery in the future. Of course, many other emerging concepts

301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319

324

C

E

299 300

R

297 298

R

295 296

O

293 294

C

291 292

N

289 290

U

287 288

Acknowledgments

369

F

285 286

O

6. Transdermal delivery, the next generation of patches

280 281

R O

284

278 279

T

282 283

The results demonstrated a pharmacokinetic profile comparable to subcutaneous injection and greater control over amounts released. Generally, the implant was as well tolerated as already-used electronic implants (e.g. pacemakers) [58]. Further improvement of this technology would be the design of closed-loop, implantable systems able to autonomously “detect” a pathological state and accordingly release appropriate drug doses [59,60]. We refer to Santini et al. in this issue for more details on this subject.

P

276 277

325 326

D

Fig. 3. Schematic representation of computer assisted micro-implants drug delivery.

(e.g. controlled release systems based on 3D-printing technologies, advanced gene delivery systems) have not been discussed here and are also rapidly evolving [70,71]. In cancer chemotherapy, it is clear that better understanding of tumor physiopathology will be instrumental in the optimal use of nanoparticles and theranostic agents. Moreover, the continuing commitment to increasingly personalized medicine is giving birth to a great diversity of experimental cell-based therapeutics. These biotechnologies will foster versatile and multifunctional nonimmunogenic human cells as well as cell derivatives with targeting and controlled drug-producing properties [38,50]. Bacteria will also be harnessed to deliver in body areas like the gastrointestinal tract a wide range of therapeutic compounds such as anti-inflammatory interleukins, toxin-neutralizing antibodies, antigens and allergens (to induce tolerogenic responses) [72]. Similarly to conventional drug carriers, thorough characterization of these novel systems will be required to avoid unforeseen side-effects. However, the use of “living” materials heavily complicates the quality control process due to their intrinsic complexity, potential out-ofhand evolution and batch-to-batch variability. Concomitant with scientific challenges, the evolution of these new-generation medicines will require the design of adapted drug approval protocols. Fundamental support will also derive from informatics and bioelectronics in general. Multitasking biosensors able to constantly monitor body parameters and accordingly trigger the release of precise amounts of medicine will profoundly change the daily life of chronic patients. Generally, fundamental research using breakthrough technologies such as the optogenetics (to regulate the activity of individual neurons in vivo) and CLARITY (for the three dimensional staining of intact organs) will notably improve our understanding of human physiology and diseases. Major fundamental discoveries will in the future be probably accompanied by innovative delivery concepts that are today difficult to predict with the current knowledge [73,74]. For instance, the recent finding that sleep induces a striking increase in convective exchange of cerebrospinal–interstitial fluids within the brain opens new interesting avenue for brain drug delivery [75]. The launch of ambitious super-computational biology projects (Human Brain Project, BRAIN Initiative) might lay the foundations to body-in-a-box machines allowing disease and drug (delivery) simulations, drastically accelerating the characterization of new delivery systems while decreasing their development costs. In conclusion, the controlled release field is entering a fascinating era, and we hope that the Journal of Controlled Release will continue to be a mainstream venue for diffusing innovative data at the forefront of drug delivery research.

E

4

327 328 329 330 331 332 333 334 335 336 Q3 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368

D. Brambilla gratefully acknowledges support from the ETH Zurich 370 Fellowship (2012-01). 371 References [1] Z. Liu, J.T. Robinson, X.M. Sun, H.J. Dai, PEGylated nanographene oxide for delivery of water-insoluble cancer drugs, J. Am. Chem. Soc. 130 (2008) 10876–10877. [2] X.M. Qian, X.H. Peng, D.O. Ansari, Q. Yin-Goen, G.Z. Chen, D.M. Shin, L. Yang, A.N. Young, M.D. Wang, S.M. Nie, In vivo tumor targeting and spectroscopic detection with surface-enhanced Raman nanoparticle tags, Nat. Biotechnol. 26 (2008) 83–90. [3] X.M. Sun, Z. Liu, K. Welsher, J.T. Robinson, A. Goodwin, S. Zaric, H.J. Dai, Nanographene oxide for cellular imaging and drug delivery, Nano Res. 1 (2008) 203–212. [4] P. Horcajada, T. Chalati, C. Serre, B. Gillet, C. Sebrie, T. Baati, J.F. Eubank, D. Heurtaux, P. Clayette, C. Kreuz, J.S. Chang, Y.K. Hwang, V. Marsaud, P.N. Bories, L. Cynober, S. Gil, G. Ferey, P. Couvreur, R. Gref, Porous metal-organic-framework nanoscale carriers as a potential platform for drug delivery and imaging, Nat. Mater. 9 (2010) 172–178. [5] M. Liong, J. Lu, M. Kovochich, T. Xia, S.G. Ruehm, A.E. Nel, F. Tamanoi, J.I. Zink, Multifunctional inorganic nanoparticles for imaging, targeting, and drug delivery, ACS Nano 2 (2008) 889–896. [6] J. Nicolas, S. Mura, D. Brambilla, N. Mackiewicz, P. Couvreur, Design, functionalization strategies and biomedical applications of targeted biodegradable/biocompatible polymer-based nanocarriers for drug delivery, Chem. Soc. Rev. 42 (2013) 1147–1235.

Please cite this article as: D. Brambilla, et al., Breakthrough discoveries in drug delivery technologies: The next 30 years, J. Control. Release (2014), http://dx.doi.org/10.1016/j.jconrel.2014.03.056

372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390

D. Brambilla et al. / Journal of Controlled Release xxx (2014) xxx–xxx

N C O

R

R

E

C

D

P

R O

O

F

[38] S.M. van Dommelen, P. Vader, S. Lakhal, S.A.A. Kooijmans, W.W. van Solinge, M.J.A. Wood, R.M. Schiffelers, Microvesicles and exosomes: opportunities for cell-derived membrane vesicles in drug delivery, J. Control. Release 161 (2012) 635–644. [39] L. Alvarez-Erviti, Y.Q. Seow, H.F. Yin, C. Betts, S. Lakhal, M.J.A. Wood, Delivery of siRNA to the mouse brain by systemic injection of targeted exosomes, Nat. Biotechnol. 29 (2011) 341-U179. [40] D. Sun, X. Zhuang, X. Xiang, Y. Liu, S. Zhang, C. Liu, S. Barnes, W. Grizzle, D. Miller, H.-G. Zhang, A novel nanoparticle drug delivery system: the anti-inflammatory activity of curcumin is enhanced when encapsulated in exosomes, Mol. Ther. 18 (2010) 1606–1614. [41] X. Zhuang, X. Xiang, W. Grizzle, D. Sun, S. Zhang, R.C. Axtell, S. Ju, J. Mu, L. Zhang, L. Steinman, D. Miller, H.-G. Zhang, Treatment of brain inflammatory diseases by delivering exosome encapsulated anti-inflammatory drugs from the nasal region to the brain, Mol. Ther. 19 (2011) 1769–1779. [42] J.L. Munoz, S.A. Bliss, S.J. Greco, S.H. Ramkissoon, K.L. Ligon, P. Rameshwar, Delivery of functional anti-miR-9 by mesenchymal stem cell-derived exosomes to glioblastoma multiforme cells conferred chemosensitivity, Mol. Ther. Nucleic Acids 2 (2013) e126. [43] E.N. Momin, G. Vela, H.A. Zaidi, A. Quinones-Hinojosa, The oncogenic potential of mesenchymal stem cells in the treatment of cancer: directions for future research, Curr. Immunol. Rev. 6 (2010) 137–148. [44] J.M. Karp, G.S.L. Teol, Mesenchymal stem cell homing: the devil is in the details, Cell Stem Cell 4 (2009) 206–216. [45] K. Shah, Mesenchymal stem cells engineered for cancer therapy, Adv. Drug Deliv. Rev. 64 (2012) 739–748. [46] L.G. Menon, K. Kelly, H.W. Yang, S.-K. Kim, P.M. Black, R.S. Carroll, Human bone marrow-derived mesenchymal stromal cells expressing S-TRAIL as a cellular delivery vehicle for human glioma therapy, Stem Cells 27 (2009) 2320–2330. [47] I.T. Cavarretta, V. Altanerova, M. Matuskova, L. Kucerova, Z. Culig, C. Altaner, Adipose tissue-derived mesenchymal stem cells expressing prodrug-converting enzyme inhibit human prostate tumor growth, Mol. Ther. 18 (2010) 223–231. [48] L. Kucerova, V. Altanerova, M. Matuskova, S. Tyciakova, C. Altaner, Adipose tissuederived human mesenchymal stem cells mediated prodrug cancer gene therapy, Cancer Res. 67 (2007) 6304–6313. [49] R. Cattaneo, T. Miest, E.V. Shashkova, M.A. Barry, Reprogrammed viruses as cancer therapeutics: targeted, armed and shielded, Nat. Rev. Microbiol. 6 (2008) 529–540. [50] C. Willmon, K. Harrington, T. Kottke, R. Prestwich, A. Melcher, R. Vile, Cell carriers for oncolytic viruses: fed ex for cancer therapy, Mol. Ther. 17 (2009) 1667–1676. [51] J. Qiao, T. Kottke, C. Willmon, F. Galivo, P. Wongthida, R.M. Diaz, J. Thompson, P. Ryno, G.N. Barber, J. Chester, P. Selby, K. Harrington, A. Melcher, R.G. Vile, Purging metastases in lymphoid organs using a combination of antigen-nonspecific adoptive T cell therapy, oncolytic virotherapy and immunotherapy, Nat. Med. 14 (2008) 37–44. [52] D.G. Roy, J.C. Bell, Cell carriers for oncolytic viruses: current challenges and future directions, Oncolytic Virother. 2 (2013) 47–56. [53] E.K. Mader, G. Butler, S.C. Dowdy, A. Mariani, K.L. Knutson, M.J. Federspiel, S.J. Russell, E. Galanis, A.B. Dietz, K.W. Peng, Optimizing patient derived mesenchymal stem cells as virus carriers for a phase I clinical trial in ovarian cancer, J. Transl. Med. 11 (2013) 14. [54] P. Van Arnum, The future of dosage forms, Pharm. Technol. 38 (2014) 28–32. [55] J.M. Maloney, S.A. Uhland, B.F. Polito, N.F. Sheppard Jr., C.M. Pelta, J.T. Santini Jr., Electrothermally activated microchips for implantable drug delivery and biosensing, J. Control. Release 109 (2005) 244–255. [56] Y. Li, R.S. Shawgo, B. Tyler, P.T. Henderson, J.S. Vogel, A. Rosenberg, P.B. Storm, R. Langer, H. Brem, M.J. Cima, In vivo release from a drug delivery MEMS device, J. Control. Release 100 (2004) 211–219. [57] J.T. Santini, M.J. Cima, R. Langer, A controlled-release microchip, Nature 397 (1999) 335–338. [58] R. Farra, N.F. Sheppard Jr., L. McCabe, R.M. Neer, J.M. Anderson, J.T. Santini Jr., M.J. Cima, R. Langer, First-in-human testing of a wirelessly controlled drug delivery microchip, Sci. Transl. Med. 4 (2012). [59] M.T. Salam, M. Mirzaei, M.S. Ly, N. Dang Khoa, M. Sawan, An implantable closedloop asynchronous drug delivery system for the treatment of refractory epilepsy, IEEE Trans. Neural Syst. Rehabil. Eng. 20 (2012) 432–442. [60] V. Ravaine, C. Ancla, B. Catargi, Chemically controlled closed-loop insulin delivery, J. Control. Release 132 (2008) 2–11. [61] S.K. Peasah, E. Azziz-Baumgartner, J. Breese, M.I. Meltzer, M.-A. Widdowson, Influenza cost and cost-effectiveness studies globally — a review, Vaccine 31 (2013) 5339–5348. [62] T.P. Monath, Vaccines against diseases transmitted from animals to humans: a one health paradigm, Vaccine 31 (2013) 5321–5338. [63] M.R. Prausnitz, R. Langer, Transdermal drug delivery, Nat. Biotechnol. 26 (2008) 1261–1268. [64] J.W. Lee, J.H. Park, M.R. Prausnitz, Dissolving microneedles for transdermal drug delivery, Biomaterials 29 (2008) 2113–2124. [65] S. Henry, D.V. McAllister, M.G. Allen, M.R. Prausnitz, Microfabricated microneedles: a novel approach to transdermal drug delivery, J. Pharm. Sci. 87 (1998) 922–925. [66] S.P. Sullivan, D.G. Koutsonanos, M.D. Martin, J.W. Lee, V. Zarnitsyn, S.O. Choi, N. Murthy, R.W. Compans, I. Skountzou, M.R. Prausnitz, Dissolving polymer microneedle patches for influenza vaccination, Nat. Med. 16 (2010) 915-U116. [67] L.Y. Chu, M.R. Prausnitz, Separable arrowhead microneedles, J. Control. Release 149 (2011) 242–249. [68] M.C. Chen, S.F. Huang, K.Y. Lai, M.H. Ling, Fully embeddable chitosan microneedles as a sustained release depot for intradermal vaccination, Biomaterials 34 (2013) 3077–3086. [69] P.C. DeMuth, A.V. Li, P. Abbink, J. Liu, H. Li, K.A. Stanley, K.M. Smith, C.L. Lavine, M.S. Seaman, J.A. Kramer, A.D. Miller, W. Abraham, H. Suh, J. Elkhader, P.T. Hammond, D.H. Barouch, D.J. Irvine, Nat. Biotechnol. 31 (2013) 1082–1085.

E

T

[7] N.R. Patel, B.S. Pattni, A.H. Abouzeid, V.P. Torchilin, Nanopreparations to overcome multidrug resistance in cancer, Adv. Drug Deliv. Rev. 65 (2013) 1748–1762. [8] R. van der Meel, L.J.C. Vehmeijer, R.J. Kok, G. Storm, E.V.B. van Gaal, Ligand-targeted particulate nanomedicines undergoing clinical evaluation: current status, Adv. Drug Deliv. Rev. 64 (2013) 1284–1298. [9] Y. Barenholz, Doxil® — the first FDA-approved nano-drug: lessons learned, J. Control. Release 160 (2012) 117–134. [10] R. Duncan, R. Gaspar, Nanomedicine(s) under the microscope, Mol. Pharm. 8 (2011) 2101–2141. [11] A.T. Florence, Pharmaceutical Nanotechnology: more than size ten topics for research, Int. J. Pharm. 339 (2007) 1–2. [12] V.J. Venditto, F.C. Szoka Jr., Cancer nanomedicines: so many papers and so few drugs! Adv. Drug Deliv. Rev. 65 (2013) 80–88. [13] D.L. Stirland, J.W. Nichols, S. Miura, Y.H. Bae, Mind the gap: a survey of how cancer drug carriers are susceptible to the gap between research and practice, J. Control. Release 172 (2013) 1045–1064. [14] M.S. Gerstel, V.A. Place, Drug delivery device, US Pat. 3964482 (A), 1976. [15] Y. Matsumura, H. Maeda, A new concept for macromolecular therapeutics in cancerchemotherapy — mechanism of tumoritropic accumulation of proteins and the antitumor agent smancs, Cancer Res. 46 (1986) 6387–6392. [16] T. Lammers, P. Peschke, R. Kuehnlein, V. Subr, K. Ulbriich, J. Debus, P. Huber, W. Hennink, G. Storm, Effect of radiotherapy and hyperthermia on the tumor accumulation of HPMA copolymer-based drug delivery systems, J. Control. Release 117 (2007) 333–341. [17] P. Ruenraroengsak, J.M. Cook, A.T. Florence, Nanosystem drug targeting: facing up to complex realities, J. Control. Release 141 (2010) 265–276. [18] K. Park, Questions on the role of the EPR effect in tumor targeting, J. Control. Release 172 (2013) 391. [19] R.K. Jain, Normalizing tumor microenvironment to treat cancer: bench to bedside to biomarkers, J. Clin. Oncol. 31 (2013) 2205-U2210. [20] S. Stapleton, C. Allen, M. Pintilie, D.A. Jaffray, Tumor perfusion imaging predicts the intra-tumoral accumulation of liposomes, J. Control. Release 172 (2013) 351–357. [21] C.J. Li, Y. Miyamoto, Y. Kojima, H. Maeda, Augmentation of tumor delivery of macromolecular drugs with reduced bone-marrow delivery by elevating blood-pressure, Br. J. Cancer 67 (1993) 975–980. [22] M.R. Kano, Y. Bae, C. Iwata, Y. Morishita, M. Yashiro, M. Oka, T. Fujii, A. Komuro, K. Kiyono, M. Kaminishi, K. Hirakawa, Y. Ouchi, N. Nishiyama, K. Kataoka, K. Miyazono, Improvement of cancer-targeting therapy, using nanocarriers for intractable solid tumors by inhibition of TGF-beta signaling, Proc. Natl. Acad. Sci. U. S. A. 104 (2007) 3460–3465. [23] T. Seki, J. Fang, H. Maeda, Enhanced delivery of macromolecular antitumor drugs to tumors by nitroglycerin application, Cancer Sci. 100 (2009) 2426–2430. [24] J. Fang, H. Qin, H. Nakamura, K. Tsukigawa, T. Shin, H. Maeda, Carbon monoxide, generated by heme oxygenase-1, mediates the enhanced permeability and retention effect in solid tumors, Cancer Sci. 103 (2012) 535–541. [25] C. Godugu, A.R. Patel, R. Doddapaneni, S. Marepally, T. Jackson, M. Singh, Inhalation delivery of Telmisartan enhances intratumoral distribution of nanoparticles in lung cancer models, J. Control. Release 172 (2013) 86–95. [26] V.P. Chauhan, T. Stylianopoulos, J.D. Martin, Z. Popovic, O. Chen, W.S. Kamoun, M.G. Bawendi, D. Fukumura, R.K. Jain, Normalization of tumour blood vessels improves the delivery of nanomedicines in a size-dependent manner, Nat. Nanotechnol. 7 (2012) 383–388. [27] A.J. Gormley, N. Larson, A. Banisadr, R. Robinson, N. Frazier, A. Ray, H. Ghandehari, Plasmonic photothermal therapy increases the tumor mass penetration of HPMA copolymers, J. Control. Release 166 (2013) 130–138. [28] L. Li, T.L.M. ten Hagen, M. Bolkestein, A. Gasselhuber, J. Yatvin, G.C. van Rhoon, A.M. M. Eggermont, D. Haemmerich, G.A. Koning, Improved intratumoral nanoparticle extravasation and penetration by mild hyperthermia, J. Control. Release 167 (2013) 130–137. [29] M. Bazan-Peregrino, B. Rifai, R.C. Carlisle, J. Choi, C.D. Arvanitis, L.W. Seymour, C.C. Coussios, Cavitation-enhanced delivery of a replicating oncolytic adenovirus to tumors using focused ultrasound, J. Control. Release 169 (2013) 40–47. [30] U. Prabhakar, H. Maeda, R.K. Jain, E.M. Sevick-Muraca, W. Zamboni, O.C. Farokhzad, S.T. Barry, A. Gabizon, P. Grodzinski, D.C. Blakey, Challenges and key considerations of the enhanced permeability and retention effect for nanomedicine drug delivery in oncology, Cancer Res. 73 (2013) 2412–2417. [31] E. Mastrobattista, Advanced drug delivery in motion, Int. J. Pharm. 454 (2013) 517–520. [32] P. Pouponneau, J.-C. Leroux, G. Soulez, L. Gaboury, S. Martel, Co-encapsulation of magnetic nanoparticles and doxorubicin into biodegradable microcarriers for deep tissue targeting by vascular MRI navigation, Biomaterials 32 (2011) 3481–3486. [33] C. Sun, J.S.H. Lee, M. Zhang, Magnetic nanoparticles in MR imaging and drug delivery, Adv. Drug Deliv. Rev. 60 (2008) 1252–1265. [34] J.L. Arias, L.H. Reddy, M. Othman, B. Gillet, D. Desmaële, F. Zouhiri, F. Dosio, R. Gref, P. Couvreur, Squalene based nanocomposites: a new platform for the design of multifunctional pharmaceutical theragnostics, ACS Nano 5 (2011) 1513–1521. [35] A. Yanai, U.O. Hafeli, A.L. Metcalfe, P. Soema, L. Addo, C.Y. Gregory-Evans, K. Po, X.H. Shan, O.L. Moritz, K. Gregory-Evans, Focused magnetic stem cell targeting to the retina using superparamagnetic iron oxide nanoparticles, Cell Transplant. 21 (2012) 1137–1148. [36] E.-S. Jang, J.-H. Shin, G. Ren, M.-J. Park, K. Cheng, X. Chen, J.C. Wu, J.B. Sunwoo, Z. Cheng, The manipulation of natural killer cells to target tumor sites using magnetic nanoparticles, Biomaterials 33 (2012) 5584–5592. [37] R.A. Rader, FDA biopharmaceutical product approvals and trends in 2012, BioProcess Int. 11 (2013) 18–27.

U

391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 Q4 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476

5

Please cite this article as: D. Brambilla, et al., Breakthrough discoveries in drug delivery technologies: The next 30 years, J. Control. Release (2014), http://dx.doi.org/10.1016/j.jconrel.2014.03.056

477 478 479 480 Q5 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 Q6 553 554 555 556 557 558 559 560 561 562

6

563 564 565 566 567 568 569 570

D. Brambilla et al. / Journal of Controlled Release xxx (2014) xxx–xxx

[70] W.E. Katstra, R.D. Palazzolo, C.W. Rowe, B. Giritlioglu, P. Teung, M.J. Cima, Oral dosage forms fabricated by three dimensional printing™, J. Control. Release 66 (2000) 1–9. [71] N. Scoutaris, M.R. Alexander, P.R. Gellert, C.J. Roberts, Inkjet printing as a novel medicine formulation technique, J. Control. Release 156 (2011) 179–185. [72] L.G. Bermúdez-Humarán, C. Aubry, J.P. Motta, C. Deraison, L. Steidler, N. Vergnolle, J.M. Chatel, P. Langella, Engineering lactococci and lactobacilli for human health, Curr. Opin. Microbiol. 16 (2013) 278–283.

[73] K. Chung, J. Wallace, S.Y. Kim, S. Kalyanasundaram, A.S. Andalman, T.J. Davidson, J.J. Mirzabekov, K.A. Zalocusky, J. Mattis, A.K. Denisin, S. Pak, H. Bernstein, C. Ramakrishnan, L. Grosenick, V. Gradinaru, K. Deisseroth, Structural and molecular interrogation of intact biological systems, Nature 497 (2013) 332–337. [74] G. Miesenböck, The optogenetic catechism, Science 326 (2009) 395–399. [75] L. Xie, H. Kang, Q. Xu, M.J. Chen, Y. Liao, M. Thiyagarajan, J. O'Donnell, D.J. Christensen, C. Nicholson, J.J. Iliff, T. Takano, R. Deane, M. Nedergaard, Sleep drives metabolite clearance from the adult brain, Science (2013) 373–377

U

N

C

O

R

R

E

C

T

E

D

P

R O

O

F

579

Please cite this article as: D. Brambilla, et al., Breakthrough discoveries in drug delivery technologies: The next 30 years, J. Control. Release (2014), http://dx.doi.org/10.1016/j.jconrel.2014.03.056

571 572 573 574 575 576 577 578

Breakthrough discoveries in drug delivery technologies: the next 30 years.

What if we could open the 2044 special issue of the Journal of Controlled Release? Which drug delivery technologies will have led the field? Which one...
624KB Sizes 2 Downloads 3 Views