Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter December 12, 2017

Stimulus-responsive liposomes as smart nanoplatforms for drug delivery applications

  • Parham Sahandi Zangabad

    Parham Sahandi Zangabad received his MSc in Nanomaterials and Nanotechnology from Sharif University of Technology (SUT), Tehran, Iran. Since 2012 until January 2015, he was a research assistant at the Research Center for Nanostructured and Advanced Materials (RCNAM) at SUT working on nanocomposites and their mechanical properties. Since late 2014, he joined Professor Karimi’s Advanced Nanobiotechnology and Nanomedicine Research Group (ANNRG) in Iran University of Medical Science, Tehran, Iran, having collaboration with Professor M.R. Hamblin from Harvard Medical School, Boston, MA, USA. To research in nanomedicine, smart drug delivery systems, and biosensors was the aim of this collaboration, resulting in several high-impact review articles and books. Afterward, since February 2016, he became a research assistant in the Research Center for Pharmaceutical Nanotechnology (RCPN), Tabriz University of Medical Sciences, Tabriz, Iran, working on nanocarriers for anticancer drug delivery systems. This led to beneficial experiences in design and applying smart nanocarriers in biomedical applications. He was then a research assistant in the Cellular and Molecular Research Center in Iran University of Medical Sciences, since 2017, working on drug/gene delivery systems.

    , Soroush Mirkiani , Shayan Shahsavari

    Shayan Shahsavari is a member of the Nanoclub Elites Association (NEA) and CEO at Mataab Company, at Tehran, Iran. He joined NEA at the Iran Nanotechnology Initiative Council as a member and as an expert nanobiotechnologist. In 2014, he established Mataab Co to manufacture and commercialize high-tech products at a biotechnology incubator, Pasteur Institute of Iran. His fields of research are theranostics and tissue engineering. Recently, he joined the ANNRG, Iran University of Medical Sciences to collaborate with Professor Mahdi Karimi in the related fields of Nanomedicine including Nanobio-sensors and Drug delivery systems.

    , Behrad Masoudi , Maryam Masroor , Hamid Hamed , Zahra Jafari , Yasamin Davatgaran Taghipour , Hura Hashemi , Mahdi Karimi

    Mahdi Karimi received his BSc degree in Medical Laboratory Science from Iran University of Medical Science (IUMS), in 2005. In 2008, he achieved the MSc degree in Medical Biotechnology from Tabriz University of Medical Science and joined Tarbiat Modares University as a PhD student in the Nanobiotechnology field and completed his research in 2013. During his research, in 2012, he affiliated with the laboratory of Professor Michael Hamblin in the Wellman Center for Photomedicine at Massachusetts General Hospital and Harvard Medical School as a visiting researcher, where he contributed to the design and construction of new smart nanoparticles for drug/gene delivery. On finishing the study, he joined, as assistant professor, the Department of Medical Nanotechnology at IUMS. His current research interests include smart nanoparticles’ design in drug/gene delivery and microfluidic systems. He has established a scientific collaboration between his lab and Professor Michael Hamblin’s lab to design new classes of smart nanovehicles in drug/gene delivery systems. Dr. Karimi was recipient of several grants from Iran Science Elites Federation (ISEF), and Iranian Nanotechnology Initiative Council. He was also ranked 34th among all faculty members in Iran, by Iran Science Elites Federation (ISEF) in Iran and selected as the best young researcher of Iran University of Medical Sciences in 2016.

    EMAIL logo
    and Michael R. Hamblin

    Michael R. Hamblin, PhD, is a principal investigator at the Wellman Center for Photomedicine, Massachusetts General Hospital, an associate professor of Dermatology, Harvard Medical School and the affiliated faculty of Harvard, MIT Division of Health Science and Technology. He directs a laboratory of around 12 scientists who work in photodynamic therapy and low-level light therapy. He has published over 380 peer-reviewed articles and documents (according to scopus), is associate editor for eight journals, and serves on NIH study sections. He has edited 10 proceedings volumes, together with four other major textbooks on PDT and Photomedicine. In 2011, Dr. Hamblin was honored by election as a Fellow of SPIE.

    EMAIL logo
From the journal Nanotechnology Reviews

Abstract

Liposomes are known to be promising nanoparticles (NPs) for drug delivery applications. Among the different types of self-assembled NPs, liposomes stand out for their non-toxic nature and their possession of dual hydrophilic-hydrophobic domains. The advantages of liposomes include the ability to solubilize hydrophobic drugs, the ability to incorporate different hydrophilic and lipophilic drugs at the same time, lessening the exposure of host organs to potentially toxic drugs and allowing modification of the surface by a variety of different chemical groups. This modification of the surface, or of the individual constituents, may be used to achieve two important goals. First, ligands for active targeting can be attached that are recognized by cognate receptors overexpressed on the target cells of tissues. Second, modification can be used to impart a stimulus-responsive or “smart” character to the liposomes, whereby the cargo is released on demand only when certain internal stimuli (pH, reducing agents, specific enzymes) or external stimuli [light, magnetic field, or ultrasound (US)] are present. Here, we review the field of smart liposomes for drug delivery applications.

Introduction

Advances in nanotechnology have made a substantial impact in many different fields and research areas, particularly in biomedicine and pharmaceutics [1], [2], [3], [4], [5], [6]. Studies of nanoparticles (NPs) as vehicles for drug and gene delivery and for controlled release have been carried out for the last 20 years [7], [8], [9], [10]. Among the many different types of NPs, liposomes are one of the most attractive types for use as drug carriers due to their dual hydrophilic-hydrophobic domains [11]. In 1965, Bangham et al. [12] reported the first “swollen” phospholipid system that formed distinct layers, and during the following years, a novel drug delivery system (DDS) based on liposomes was developed. Liposomes are self-assembled, totally enclosed spherical vesicles composed of non-toxic phospholipids, which are capable of encapsulating both hydrophilic and lipophilic cargomolecules [13].

Liposomes have several special advantages for drug delivery: enhancing the solubility of hydrophobic drugs: the ability to incorporate both hydrophilic and lipophilic drugs at the same time; lessening the exposure of host organs to potentially toxic drugs; and allowing modification of the surface by a variety of different chemical groups [14]. The ability to design smart vehicles that can lead to accumulation of drugs or other therapeutic agents specifically at a diseased site has resulted in the development of the concept of active targeting [15]. Surface modification of liposomes can be carried out by attachment of antibodies (or other ligands that are recognized by over- receptors) for site-specific targeting. Also, chemical moieties can be attached to the surface of the liposomes (or to the constituent building blocks) that will respond to various stimuli. These “smart liposomes” can undergo triggered drug release based on various physiology-dependent properties. The stimuli can be either internal (e.g. enzyme activity, pH changes, or the presence of reducing agents) or they can be external stimuli applied from outside (e.g. temperature, light, magnetic field, or ultrasound (US)) [16]. The release of drug from liposomes based on triggering by external stimuli provides a better accuracy concerning the timing and location of release, and a better control over the delivery and dosage of the drug [17].

Over the last few years, many different FDA-approved liposomes have been introduced for clinical applications, and many others are in different phases of clinical trials. (Table 1).

Table 1:

FDA-approved liposomal drug delivery systems used in clinical application.

ProductReleased drugTreatmentReference
DoxilDoxorubicin (DOX)Kaposi’s sarcoma[18], [19]
Breast cancer
MyocetDOXMetastatic breast cancer[20], [21]
LipodoxDOXBreast cancer[22]
Kaposi’s sarcoma[23]
Ovarian cancer
DaunoxomeDaunorubicinKaposi’s sarcoma[24], [25]
MarqiboVincristineAcute lymphoblastic leukemia[26], [27], [28]
AmphotecAmphotericin BInvasive aspergillosis[29], [30]
Serious fungal infections
DepocytCytarabineNeoplastic meningitis and lymphomatous meningitis[31], [32]
EstrasorbEstrogenMenopausal therapy[33]
DepoDurMorphine sulfatePain[34], [35], [36]
VisudyneVerteporfinPDT for age-related molecular degeneration[37], [38]
EpaxalInactivated hepatitis A viral strain RG-SBVaccine for hepatitis A[39], [40]

Temperature-responsive liposomes

Temperature plays a critical role in the metabolic activity of cells. The normal mammalian body temperature of 37°C is an ideal temperature for intracellular chemical reactions. Difficulties arise when tissue temperature exceeds the approximate limit of 43–45°C, hyperthermia, as the cells will not be able to readily maintain their normal activity [41], [42]. The utilization of mild hyperthermia (HT), a few degrees higher than the physiological temperature, has long been applied together with radiotherapy and chemotherapy in order to boost the efficacy of these treatments with minimal side effects [43], [44]. Temperature-sensitive or thermo-sensitive liposomes (TSLs) have been considered as one of the most efficient types of nanocarriers for site-specific drug delivery applications [45], [46], [47].

HT, which can be applied using different modalities, combined with temperature-sensitive liposomes can enhance therapeutic efficiency as follows [48]:

  1. Increasing the accumulation of liposomes in the tumor by increasing tumor vascular permeability

  2. Controlling the release of therapeutic agents from temperature-triggered liposomes into the tumor vascular and interstitial space

  3. Modulating the target cells with enhanced permeability and susceptibility to released drugs

  4. Increasing blood flow at the area exposed to heat

  5. Exerting direct cytotoxicity to cancer cells while sparing normal cells

The combination of TSL therapy with external HT can be carried out with three different liposome formulations, namely, traditional TSLs (TTSLs); lysolipid-containing TSLs (LTSLs); and polymer-modified TSLs (PTSLs). The design, drug release behavior, mechanisms of thermally triggered release, and the clinical potential of each type have been reviewed by Ta and Porter [48]. Here, we briefly summarize each category and its properties, and highlight recent advances in that particular area.

TTSLs contain traditional phospholipids and undergo a phase transition such as a gel-to-liquid crystallite transition, or a structural transition from lamellar to hexagonal morphology [45], [48]. These types of liposomes were introduced for the first time in 1978 by Yatvin et al. [49], [50]. These authors used dipalmitoyl phosphatidylcholine (DPPC) for the primary lipid with a transition temperature of 41°C, and distearoylphosphatidylcholine was added to DPPC in order to adjust the DPPC transition temperature to the desired value and release the cargo, neomycin. Since then, a number of studies have focused on the preparation of DPPC-based liposomes in combination with other lipids and polyethylene glycol-(PEG)-lipid conjugates in order to enhance the permeability of the liposomal membrane and producing long-circulating (stealth) TTSLs [51], [52], [53]. Levachea et al. [52] described a DOX-loaded temperature-sensitive liposome (DOX-TL) composed of a combination of DPPC, 1,2-distearoyl-sn-glycero-3-phosphocholine (DSP), distearoyl-sn-glycero-3-phosphoethanolamine-N-[maleimide-PEG-2000-ammonium salt] (DSPE-PEG-2000), and cholesterol. This liposomal formulation was triggered in response to local hyperthermia, while it remained stable at physiological temperature. By optimization of the liposomal composition, they successfully improved the stability and inhibited premature drug release in antitumor therapy.

LTSLs were first reported by Anyaramabhatla and Needham [54] in 1999 to reduce the phase transition temperature and boost the rapid drug release over a period of tens of seconds. Since then, lysolipid formulations have undergone further development and has shown improved properties in comparison with TTSLs. Studies suggest that the presence of the lysolipid in a relatively low molar percentage in the primary lipid (DPPC) bilayers causes the stabilization of defects in the lipid membrane through the phase transition. The optimized LTSL formulations for rapid drug release and stable liposomal membranes are tailored to have a transition temperature in the range of 39–40°C [48].

Along with the advantages of lysolipids, there is also a drawback consisting of the possibility of the lysolipids leaking from the liposomal shell and degrading the bilayer stability. This led to the idea of synthesizing PTSLs, which has received considerable amount of interest from researchers. The incorporation of polymers can lead to structures that can be either completely or partially degraded, or undergo a phase transition, and disrupt the liposomal membrane in response to heat. These polymers exhibit a lower critical solution temperature (LCST) and an upper critical solution temperature (UCST), below and above which the polymers are soluble, and near these temperatures, the polymers undergo a coil-to-globule transition [45], [48], [55]. Among the different types of temperature-sensitive polymers, poly(N-isopropylacrylamide) (pNIPAAm) has been extensively studied [56], [57]. In a recent study, Pippa et al. [58] synthesized novel temperature-sensitive liposomes by coating DPPC with an end-functionalized C12H25-PNIPAM-COOH copolymer. This polymer is temperature-sensitive leading to membrane disruption, and the polymer chains aggregated at temperatures above 32°C. The composition and the molar weight of the PNIPAM chains played an essential role in the thermo-tropic properties of the mixed liposomal membrane. Also, it was shown that by tailoring the DPPC/PNIPAM ratio in the components, drug release from this nanocarrier could be effectively controlled. Turner et al. [59] engineered a thermo-sensitive liposome formulation containing DPPC, soyPC, and cholesterol with optimized molar ratios. Then, the liposome was coated with PEG and dextran, and they compared the stability and the release behavior of the liposome with these two different coatings. They concluded that dextran and PEG had similar release properties. However, in vitro macrophage uptake was greater with dextran. Another example of polymer-modified thermo-sensitive liposomes was described by Guo and Kim [60]. They produced an electrostatic complex between cinnamic acid (CA) and polyethyleneimine (PEI), and the PEI-CA conjugate was immobilized on the surface of an egg phosphatidylcholine (EPC) liposome formulation. The PEI-CA conjugate could be disassembled above its UCST (hindering drug release), and remained assembled below its UCST, which triggered drug release by increasing stress on the liposomal membrane (Figure 1). This PEI-CA conjugate could change its configuration in response to a temperature below or above its UCST thus controlling drug release.

Figure 1: Schematic of temperature-dependent behavior of liposome-conjugating PEI-CA (A) above UCST and (B) below UCST.
Figure 1:

Schematic of temperature-dependent behavior of liposome-conjugating PEI-CA (A) above UCST and (B) below UCST.

Wang and Kim [26] modified the block co-polymer Pluronic F127 by attaching cinnamoyl groups (CF127) and immobilized it on the surface of EPC liposomes resulting in triggered release of its water-soluble payload in response to a temperature change. The drug release could be triggered by the phase transition of CF127 on the surface of the liposome. When the aqueous solution was heated up, the CF127 underwent micellization and gelation as a result of phase transition, which triggered the drug release from the liposomal membrane. Another approach for producing polymer-modified thermo-sensitive liposomes was recently described as the so-called “polymer-caged nanobins” (PCNs). In this study, a hydrogel network was integrated over the surface of a liposomal carrier. Cholesterol-terminated poly(acrylic acid)(chol-PAAc) chains were immobilized into the bilayer membrane of the drug-loaded liposome, and then, chol-PAAc chains were crosslinked with telechelic diamine linkers. This gel-like three-dimensional network of polymers was thermo-sensitive, which induced drug release from the liposomal core [61].

In addition to the three above-mentioned extensively studied categories (TTSLs, LTSLs, and PTSLs), other novel liposomal formulations have been reported. “Nanohybrid liposomal cerasomes” (NHLC) are one of these formulations [62], [63], [64], [65]. NHLC are liposomal bilayers with an inorganic silicate framework on the surface. This system is biologically stable. However, slow drug release from this carrier can be considered to be one of its drawbacks. Recently, Liang et al. [66] developed a stable NHLC formulation, thermo-sensitive cerasome (HTSC), that displayed rapid thermo-sensitive drug release upon exposure to high-intensity focused ultrasound (HIFU). This system consisted of four types of lipids, each of which played a distinct role in the formulation, including N-[N-(3-triethoxysilyl)propylsuccinamoyl]-dihexadecylamine (CFL), DPPC, mono C17 lipid (MSPC), and DSPE-PEG-2000. Different molar ratios of CFL to DPPC were explored to enhance serum stability and thermal sensitivity. This formulation was prepared using a thin-film hydration method inside and a self-assembly process leading to sol-gel reaction. In this system, the presence of DSPE-PEG-2000 induced rapid drug release. Figure 2A illustrates the formation of the aforementioned nanosystem, its drug release behavior, and a comparison to conventional low temperature-sensitive liposomes (LTSLs).

Figure 2: Ultrasound-sensitive drug-loaded liposomal cerasome.(A) Schematic of the formation of drug-loaded nanohybrid liposomal cerasome and its drug release on exposing the system to HIFU [66]. The release behavior of one of the synthesized HTSC formulations (HTSCs-3) compared to the component lipids of conventional low temperature-sensitive liposomes (LTSLs) is shown in (C). It was concluded that below 41°C, LTSLs released calcein (used as a model drug) more rapidly than HTSCs; however, above 41°C, LTSLs and HTSCs-3 exhibited similar release profiles. Accordingly, HTSCs-3 showed a more narrow temperature-responsive range, which was better for drug release in the hyperthermia range. This formulation overcame one of the disadvantages of temperature-sensitive liposomes, viz variable drug delivery efficiency due to heterogeneous heat distribution in the tumor tissue after mild hyperthermia (that occurs due to variations in tumor vascularity). (B) Release behavior of one of the HTSCs’ formulations compared to LTSLs. Reprinted with permission from Ref. [66], copyright 2015, American Chemical Society.
Figure 2:

Ultrasound-sensitive drug-loaded liposomal cerasome.

(A) Schematic of the formation of drug-loaded nanohybrid liposomal cerasome and its drug release on exposing the system to HIFU [66]. The release behavior of one of the synthesized HTSC formulations (HTSCs-3) compared to the component lipids of conventional low temperature-sensitive liposomes (LTSLs) is shown in (C). It was concluded that below 41°C, LTSLs released calcein (used as a model drug) more rapidly than HTSCs; however, above 41°C, LTSLs and HTSCs-3 exhibited similar release profiles. Accordingly, HTSCs-3 showed a more narrow temperature-responsive range, which was better for drug release in the hyperthermia range. This formulation overcame one of the disadvantages of temperature-sensitive liposomes, viz variable drug delivery efficiency due to heterogeneous heat distribution in the tumor tissue after mild hyperthermia (that occurs due to variations in tumor vascularity). (B) Release behavior of one of the HTSCs’ formulations compared to LTSLs. Reprinted with permission from Ref. [66], copyright 2015, American Chemical Society.

Another approach that can be applied for further enhancing of the therapeutic efficiency and bioavailability of TSLs is active targeting using cell-specific ligands such as antibodies, folate, and peptides, which can be attached to the surface of TSLs. Some degree of targeting can also be achieved by incorporating cationic lipids into the bilayer [44], [67], [68], [69]. Kono et al. [67] synthesized thermo-sensitive liposomes with target specificity. Different types of PEGylated liposomes were functionalized with thermo-sensitive poly (2-2-ethoxy) ethoxyethyl vinyl ether) chains and conjugated to a monoclonal antibody, trastuzumab (Herceptin) that recognizes the HER2 receptor, which is overexpressed on some breast cancers. The liposomes were loaded with indocyanine green for near-infrared fluorescence imaging. Owing to the specific interaction between the antibody and its target cells, the liposomes accumulated efficiently at tumor sites.

Light-responsive liposomes

Among the diverse external stimuli, the use of light as a triggering agent for smart DDSs has been extensively reviewed in the literature [4], [5]. Light is particularly attractive as a wide variety of parameters, such as wavelength, duration, the intensity of the beam, and beam diameter, can be tailored to achieve the desired release profile and tissue penetration. Furthermore, non-invasive light activation can be exerted with high temporal and spatial control [70]. Electromagnetic light wavelengths below 650 nm cannot penetrate deeply through tissue (>1 cm) because of scattering and absorption by endogenous chromophores (like water, lipids, and hemoglobin) [71]; thus, although UV light can trigger a lot of chemical reactions, it can be employed as a triggering agent only for superficial treatments applied to the skin and mucosa [72]. The main strategy to obtain a deeper light penetration is the use of near-infrared (NIR) light within the wavelength range of 650–1000 nm, which also causes less damage to living cells [73].

Liposomes are among the most studied types of light-responsive NPs. The key element in the light-responsive NPs is the necessity for a chromophore (a chemical moiety that can capture light energy) [4]. Many mechanisms have been developed to induce light-triggered release from NPs, such as photo-crosslinking [74], photo-un-crosslinking [75], photochemical hydrophobicity switch [76], and photo-induced cleavage of chemical bonds [77]. Mechanisms used in light-sensitive liposomes can be categorized into three groups, photo-crosslinking, photochemical triggering, and photo-isomerization [78]. Each mechanism works on the basis of a light-triggered change in the chemical structure of a specific molecule. Some important photo-sensitive chemical groups are azobenzene and spiropyran (that undergo a photo-isomerization transition), and coumarin (that undergoes a photo-crosslinking reaction) [79].

Photo-crosslinking (photo-polymerization)

Although photo-crosslinking or photo-polymerization has been used as a means for formation of NPs, this same mechanism can also be employed as a means to trigger drug release from liposomes. Photo-polymerization causes short- or long-lived pores to occur in the liposome membrane by polymerizing double bonds in components in the hydrophobic part of the liposome bilayer, leading to shrinkage of the bilayer, and consequent drug release (Figure 3) [80]. Recent studies have also employed photosensitizers in the liposome structure to increase the sensitivity and to shift the wavelength of light absorption into the visible region [81]. Yavlovich et al. [82] prepared photo-triggerable liposomes from photo-polymerizable diacetylene phospholipid and DPPC, which released calcein (a fluorescent dye entrapped in liposome) upon treatment with UV light (254 nm).

Figure 3: Schematic of light-activated polymerization of liposomes, leading to the release of cargos.
Figure 3:

Schematic of light-activated polymerization of liposomes, leading to the release of cargos.

Photochemical activation

Many mechanisms can lead to the disturbance of structured assembly of the lipidic components of the liposome bilayer and the consequent release of its contents: photosensitization-induced oxidation (photodynamic), triggering with “photo-acid generators”, and photo-deprotection of fusogenic lipids [73], [83]. An example of the use of a photosensitizer in photochemical triggering is the activation of plasmenylcholine by photosensitizers activated by NIR light. Photosensitizers including octabutoxyphthalocyanine, Zn-phthalocyanine, and bacteria-chlorophyll-α, which absorb light between 630 and 820 nm, can lead to the photo-triggering of liposomes and the release of payload [83].

Photo-isomerization

Another approach to develop light-responsive liposomes is to use polymers that have been modified by attaching chromophores that undergo conformational changes upon light irradiation [84]. Lipids can be modified with light-responsive chromophores such as azobenzene, spirooxazine, stilbene, fulgide, or spiropyran groups. All these contain a double bond, which can undergo trans-to-cis isomerization in their structure (Figure 4A). Azobenzene is among the most attractive chromophores due to its favorable properties, including high stability, bright color, low flammability, and reversible and fast isomerization [87]. The mechanism involves switching from nonpolar and more stable trans-isomer, to the more polar cis-isomer upon excitation with visible, UV, or even NIR wavelengths leading to drug release [85], [86], [88] (Figure 4B and C).

Figure 4: Light-sensitive liposomes via cis-trans isomerization of azobenzene. (A) Schematic of a liposomal membrane equipped with azobenzene moieties (left), and the polarity alterations during the reversible cis-trans isomerization of azobenzene (right) [85] (open access). (B) Schematic of photo-isomerization mechanism that occurred in liposomal carriers. (C) Drug release curves under 980-nm irradiation (Release obtained in buffer and under the irradiation duration time of 30 min). Reprinted with permission from Ref. [86], copyright 2016, John Wiley & Sons, Inc.
Figure 4:

Light-sensitive liposomes via cis-trans isomerization of azobenzene. (A) Schematic of a liposomal membrane equipped with azobenzene moieties (left), and the polarity alterations during the reversible cis-trans isomerization of azobenzene (right) [85] (open access). (B) Schematic of photo-isomerization mechanism that occurred in liposomal carriers. (C) Drug release curves under 980-nm irradiation (Release obtained in buffer and under the irradiation duration time of 30 min). Reprinted with permission from Ref. [86], copyright 2016, John Wiley & Sons, Inc.

The most attractive feature of photo-isomerization is its reversibility, i.e. the conversion of cis back to trans is a relaxation back to the more stable state and can be achieved by irradiation with longer wavelengths of light or by thermal relaxation [2]. Many studies have investigated azobenzene-modified liposomes; for example, Ishii et al. introduced azobenzene into the amphiphilic region of the 1,2-dioleoyl-sn-glycero-3-phosphocholineliposomal membrane in order to change the membrane permeability by light irradiation [89]. Yao et al. [87] reported a NIR light-responsive azobenzene-liposome carrier, loaded with DOX. It was found that >90 wt% of DOX was released under 7.8 W/m2 power density (Figure 4C).

Photo-induced thermo-responsive release

In contrast with conventional bulk gold metal salts, gold nanostructures (NPs, nanoshells, nanorods, nanocages, etc.) can absorb light in the NIR region and in the visible region (five times higher absorption coefficients than common photo-absorbing dyes) [90]. The absorption occurs due to the surface plasmons (quasi-particles of plasma oscillation), whereby the conductance electrons possess a high resonance with certain electromagnetic frequencies [91]. The photo-absorption by gold nanostructures increases the temperature by forming a “heated electron gas”, which loses its energy rapidly (within 1 ps) by heat exchange with the nanostructure lattice. Thereby, the surrounding medium is heated locally within a time scale of 100 ps [92]. This thermal property can be employed to trigger release of the drugs entrapped in the liposomes (Figure 5) [93]. Moreover gold NPs can be used on their own in laser-induced hyperthermia-based cancer treatment, in which tumor tissue is targeted by Au-NPs to kill the cancer cells [94], [95].

Figure 5: Schematic of gold NPs being excited by NIR radiation leading to drug release from thermosensitive liposomes. Reprinted with permission from Ref. [93], copyright 2015, Elsevier.
Figure 5:

Schematic of gold NPs being excited by NIR radiation leading to drug release from thermosensitive liposomes. Reprinted with permission from Ref. [93], copyright 2015, Elsevier.

Wu et al. [96] prepared DPPC liposomes loaded with hollow gold NPs and 6-carboxyfluorescein (a fluorescent dye used as a soluble model drug). Here, hollow gold NPs had a maximum absorption at 820 nm, and the maximum drug release occurred at 4.3 W/cm2 power density. Recently, Lajunen et al. [97] described a liposomal DDS loaded with gold NPs prepared in nanostar and nanorod shapes. The release profile of the calcein payload was investigated as a function of temperature (Figure 6).

Figure 6: Curves indicating calcein release profiles (DLD, 1,2-dipalmitoyl-sn-glycero-3-phosphocholine, 1-stearoyl-2-hydroxy-snglycero-3-phosphocholine,1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[methoxy(polyethylene glycol) and DD, 1,2-distearoylsn-glycero-3-phosphocholine, 1,2-dipalmitoyl-sn-glycero-3-phosphocholine). Reprinted with permission from Ref. [97], copyright 2015, Elsevier.
Figure 6:

Curves indicating calcein release profiles (DLD, 1,2-dipalmitoyl-sn-glycero-3-phosphocholine, 1-stearoyl-2-hydroxy-snglycero-3-phosphocholine,1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[methoxy(polyethylene glycol) and DD, 1,2-distearoylsn-glycero-3-phosphocholine, 1,2-dipalmitoyl-sn-glycero-3-phosphocholine). Reprinted with permission from Ref. [97], copyright 2015, Elsevier.

pH-sensitive liposomes

Although significant developments have been achieved in addressing the challenges related to liposomal drug delivery, there are still some major problems [98]. One of the most important challenges is the ability of the delivery system to selectively enhance the bioavailability of the drugs at the target site, as well as to maintain its stability in the blood circulation until it reaches its destination. Currently, one of the most common strategies to overcome the aforementioned problem is “pH-sensitive liposomes”. This strategy is frequently used to target cancer as relatively low pH is one of the most common features of solid tumors. Moreover, the release of liposomal contents inside the acidic compartments of cells (endosomes or lysosomes) can also be achieved. The principle mechanism of these NPs is the destabilization of the structure under acidic conditions, which, in turn, leads to the release of their contents. In other words, the acidification of the surrounding medium induces selective destabilization of the liposomal membrane [98], [99].

However, the aforementioned strategy has two major limitations. First, in the case of tumor therapy, the acidic sections of tumors are located remotely from the microvasculature, as that is where hypoxia is most pronounced leading to accumulation of lactic acid due to the cellular metabolism switching to glycolysis (Warburg effect). This remoteness from the blood supply makes it difficult for IV-injected liposomes to efficiently accumulate. Second, as the tumor pH range rarely reaches lower than pH 6.5, it is technically difficult to fabricate engineered liposomes that are capable of efficiently responding to such a small pH change (~1 pH unit) [98], [100].

The structure of the pH-sensitive liposomes generally consists of a neutral lipid, which is usually a phosphatidylamine derivative such as dioleoylphosphatidyl-ethanolamine (DOPE), dioleoylphosphatidyl choline or N-succinyl-DOPE, and a weakly acidic amphiphile, such as cholesteryl hemisuccinate (CHEMS) [101], [102], [103]. The negatively charged group of the phospholipid undergoes destabilization in an acidic environment, leading to enhanced liposomal fusion with the cell membrane (in tumors) or with the endosomal membrane (inside cancer cells) and, ultimately, the release of its contents [98], [104]. Researchers have been trying to introduce novel lipids as a replacement for DOPE. Several approaches such as the combination of cationic/anionic lipids, a lipid diolein with CHEMS, and a formulation containing eggPC combined with Tween-80 have also been reported [105], [106]. These new formulations have shown much better pH-responsive characteristics compared with DOPE liposomes.

Liposome recognition by opsonins (naturally occurring serum proteins designed to remove foreign bodies from the circulation) and, subsequently, their endocytosis and clearance by the reticulo-endothelial system (RES) are other hindrances to the implementation of the liposomal drug delivery [107]. Moreover, the liposomal carrier and its contents can be totally degraded by various enzymes after its delivery to the lysosomes after endocytosis occurs into target cells (Figure 7) [108]. Therefore, efficient strategies must be developed to design liposomes, which can release their contents into the cytosol (so the drugs can reach the nucleus, which is often the final site of action) before they are degraded in the lysosomes. One useful technique applied for reducing the RES uptake and prolonging the circulation time is the incorporation of PEG-lipid conjugates into the liposomal bilayer [109], [110]. However, the problem with this approach is that PEGylation may interfere with the membrane fusion of the liposome and may even cause membrane destabilization [111]. One suggested solution to this problem was to develop liposomes that were capable of losing their coating polymer before they reached their target sites. This approach resulted in the introduction of liposomes with a cleavable linker between their coating polymer chain and the hydrophobic part of the liposome bilayer. Through surface modification, the stability and cellular uptake efficiency of liposomes can be enhanced; for example, via exploitation of novel materials such as maleimide [112], [113].

Figure 7: Design of MGlu-Dex-modified liposomes as a vaccine for induction of antigen-specific immunity. MGlu-Dex-modified liposomes are taken up by dendritic cells (DC) via endocytosis and trapped in endosome. The weakly acidic DC cytosol induce the formation antigen-specific cytotoxic T lymphocytes (CTL) via presentation by MHC class I, resulting in induction of cellular immunity (Th1 response). Antigen molecules in the endosomes undergo presentation by MHC class II and induce antigen-specific Th2 cells that lead to induction of humoral immunity. Reprinted with permission from Ref. [117], copyright 2014, Elsevier.
Figure 7:

Design of MGlu-Dex-modified liposomes as a vaccine for induction of antigen-specific immunity. MGlu-Dex-modified liposomes are taken up by dendritic cells (DC) via endocytosis and trapped in endosome. The weakly acidic DC cytosol induce the formation antigen-specific cytotoxic T lymphocytes (CTL) via presentation by MHC class I, resulting in induction of cellular immunity (Th1 response). Antigen molecules in the endosomes undergo presentation by MHC class II and induce antigen-specific Th2 cells that lead to induction of humoral immunity. Reprinted with permission from Ref. [117], copyright 2014, Elsevier.

One of the new strategies of designing pH-sensitive liposomes is the modification of the liposome surface with pH-labile polymers [114], [115], [116]. Accordingly, Yuba et al. [117].designed pH-labile polymers using dextran derivatives containing 3-methylglutaraldehyde residues (MGLu-Dex). The surface modification of egg yolk phosphatidylcholine (EYPC) liposomes with MGLu-Dex was reported to show significant destabilization in weakly acidic environments (Figure 7). In another study, Zong et al. [118] reported a controlled release liposome platform using polydopamine (PDA)-coated liposomes as pH-sensitive carriers. They used a chemotherapeutic agent, 5-fluorouracil (5-FU), as the liposomal cargo. It was demonstrated that 5-FU-loaded pH-sensitive liposomes showed improved performance compared to free drugs at tumor pH (pH 6.87).

Another important class of pH-sensitive liposomal delivery systems is attached to target-specific ligands, such as folate and transferrin, or to cell-penetrating peptides, such as TAT [100]. These ligands are capable of recognizing and binding to specific receptors overexpressed on the target cells. This method is known as “active targeting” in contrast with “passive targeting” in which liposomes accumulate at target sites merely according to their size (enhanced permeability and retention effect) [100], [119]. The combination of surface modification to target folate receptors (FR) (overexpressed on many different types of solid tumors), together with a pH-responsive liposome strategy, improved the performance of FR-targeted anticancer drugs as the FR receptor-mediated endocytosis leads to endosomal localization. Other important ligands that are used for surface modification of liposomes include monoclonal antibodies that are against antigens, such as H-2Kk (expressed in different types of tumor cell), E-selectin (on activated vascular endothelial cells), p-glycoproteins (on endothelial cells), CD-19 (on B-lymphoma cells), CD (on T-leukemia cells), and so forth. Immuno-liposomes, which use antibodies as their surface modification on the pH-sensitive liposomes, are also well investigated. Immunoglobulins (IG) of the IgG class and their derivatives are extensively exploited as targeting moieties that can be incorporated to the surface of liposomes with no undesired effects on the liposomal integrity [120], [121], [122], [123], [124], [125].

Although the main focus of most studies carried out on pH-responsive liposomes has been the pH-controlled disturbance of the membrane structure, there are studies on different mechanisms based on internal acidification of liposomes [126], [127]. More recently, there have been several attempts performed to develop pH-labile liposomes for addressing multidrug resistance of cancer cells. Liu et al. prepared novel pH-labile liposomes loaded with DOX using a NH4HCO3 gradient method. The bicarbonate ion (encapsulated within the liposome structure) reacted in acidic environments to produce copious quantities of CO2 gas, which subsequently disturbed the liposome membrane and allowed drug release (Figure 8) [128], [129].

Figure 8: Gas generating-based pH-labile liposome. (A) A novel DOX-encapsulated liposome with NH4HCO3will produce CO2 in an acidic environment leading to drug release. (B) Quantitative DOX release from pH-sensitive liposome (Lipo-c) and normal liposome (Lipo-s). (C) The resistant MCF-7 cell viability after treatment with DOX-encapsulated liposomes. Reprinted with permission from Ref. [128], copyright 2012, Royal Society of Chemistry.
Figure 8:

Gas generating-based pH-labile liposome. (A) A novel DOX-encapsulated liposome with NH4HCO3will produce CO2 in an acidic environment leading to drug release. (B) Quantitative DOX release from pH-sensitive liposome (Lipo-c) and normal liposome (Lipo-s). (C) The resistant MCF-7 cell viability after treatment with DOX-encapsulated liposomes. Reprinted with permission from Ref. [128], copyright 2012, Royal Society of Chemistry.

Generally, pH-sensitive liposomes have demonstrated a variety of applications including (1) delivery systems for chemotherapy, (2) pH-sensitive immunoliposomes, (3) tumor diagnostic tools, and (4) biological macromolecule (vaccine/DNA/gene/oligonucleotide) delivery systems [104], [130], [131], [132], [133], [134], [135], [136], [137], [138], [139], [140]. Table 2 summarizes the applications of anticancer drugs delivered via pH-responsive liposomes.

Table 2:

Several examples of pH-labile liposomes reported in literature for anti-cancer therapy.

Liposome formulationPreparation methodpH-sensitive moietyEncapsulated cargoTarget site/cell lineEncapsulation/release efficiency (%)Size (nm)BioassayReference
DOPE/HSPE/CHEMS/CHOL/mPEG2000–DSPE(4:2:2:2:0.3)DOXB lymphoma cells95/–120In vitro[141]
PEG-HZ-PETAT peptideTAT peptidesH9C2 rat embryonic cardiomyocytes, LLC tumorsIn vitro and in vivo[142]
DOPE/CHEMS/DSPE-PEGUnsaturated phosphatidylethanolamineCisplatinA549 (lung cancer)

A431 (cervix cancer)

GLC4 (lung cancer)

Ehrlich solid tumor
26.3/–

23
110

121, 130
In vitro and in vivo[143], [144], [145]
PEG/PE,TATpCell-penetrating TAT peptideDOXB16-F10,HeLa and MCF-7 cells120In vitro[146]
Fplate-poly(2-ethyl-2-oxazoline)-DSPEFPEOz-DSPEDOXHuman ovarian cancer cell SKOV391/45100–150In vitro[147]
DOPE/DSPG/DSPE-PEG(7:3:0.53)MethotrexateRat femoral artery2.51/37143In vitro and in vivo[148]
PC/DDAB/CHEMS/Tween-80/Folate-PE-PEG (25:25:49:1:0.1)A procedure based on polycarbonate membrane extrusionOAlc-containing liposomes targeted to the FRCytosine-b-d-arabinofuranosideKB human oral cancer cells15/33,83 (dependent on Tween-80 concentration)In vitro[149]
HSPC/CHEM/PEG/diethylenetriaminepentaacetic acid-modified phosphatidylethanolamineCalcein–/>2088–102In vitro[150]
PEtOz-CHEMSLipid film hydration methodDOXA375 cells96/90100–120In vitro and in vivo[151]
PE:Chol:OA (3:2:3)Thin-film dispersion methodDocetaxel–/86.9277In vivo[152]
DOPE:CHEMS:PDP-PEG-DOPE (6:4:0.1)Reverse-phase evaporation vesicle methodEGFR dioleylphosphatidylethanolamine (DOPE)GemcitabineA549 (lung cancer)67/–146In vivo[153]
HHG2C18-L and PEGHG2C18-LFilm dispersion methodZwitterionic oligopeptide lipids hexahydrobenzoic amideTemsirolimus (CCI-779)A498 (renal carcinoma)

Renca (murine renal carcinoma)
93.76/<1(for in vitro)104.5In vitro and in vivo[154]
S100PC:Chol:mPEG-Hz-Chol (90:10:3)HydrazonePaclitaxelMCF7 (breast cancer)92.2/65.8132.6In vitro[155]
  1. DOPE, dioleoylphosphatidylethanolamine; CHEMS, cholesteryl hemisuccinate; Chol, cholesterol; DSPE, 1,2-distearoyl-sn-glycero-3-phosphoethanolamine; HSPC, hydrogenated soy phosphatidylcholine; PC, phosphatidylcholines; PE, pphosphatidylethanolamine; DSPG, dipalmitoylsuccinylglycerol; PEG, poly ethylene glycole; DDAB, dimethyldioctadecylammonium bromide; PEG-HZ-PE, PEG and PE being conjugated with the lowered pH-degradable hydrazone bond; HHG2C18, 1,5-dioctadecyl-lglutamyl 2-histidylhexahydrobenzoic acid; HS, hydrogenated soy; PEtOz, poly(2-ethyl-2-oxazoline).

Ultrasound responsive

US can be simply described as acoustic waves with frequencies higher than 20 kHz, which is higher than the audible range of human hearing [156]. The common medical applications of US include imaging, tissue ablation, kidney stone disruption, physiotherapy, and so forth [157]. However, the original idea of using US as a method for drug delivery dates back to the era when biochemists began using US to rupture biological cell membranes to remove the cell contents before purification [158]. One of the advantages of using US for drug delivery is that it can be focused from outside to any organ almost regardless of depth, which makes it reasonably suitable for the targeted delivery of anticancer drugs/genes to tumors located in the pancreas, brain, as well as any other organ such as the liver [159].

The role of US in drug delivery nanocarriers such as liposomes is to rupture the structure of the phospholipid bilayer. As the pulses of the US waves propagate through tissue, some physical phenomena take place: cavitation, hyperthermia, and acoustic streaming [160]. It is currently understood that cavitation is the main factor that governs the response of liposomes and other lipid-based nanocarriers subjected to US [161]. Gas bubbles (which may be present already) or which can form as the pressure drops below the vapor pressure of the surrounding medium, will undergo rapid oscillations, i.e. expansion and contraction, as the pulses of US arrive. These oscillations in gas bubbles are called cavitation bubbles. Depending on the frequency and amplitude of the US waves as well as the size and properties of the bubbles, there are two types of cavitation: internal and stable [157], [160] (Figure 9). At lower US intensities, stable cavitation occurs, while at higher US intensities, internal cavitation takes place. Internal cavitation can eventually lead to bursting of the bubbles, and hence, drug release occurs [162]. Studies on delivering doxorubicin (DOX) encapsulated in micelles suggested that low-frequency US would cause more efficient drug release than high-frequency US [163]. The mechanism of the drug release is due to the internal cavitation of gas bubbles occurring near these micelles [160] (Figure 10A). Cavitation, in liposomes, which have an additional layer of phospholipids compared to micelles, either occurs near the liposomes (like micelles) or can occur inside their lipid bilayers [160] (Figure 10B). Low-frequency US waves also show better results in rupturing liposomes [164]. It has also been shown that using low-frequency US on liposomes did not alter the chemical properties of the encapsulated drug or destroy its activity [165].

Figure 9: Stable and internal cavitation. (A) Stable cavitation: bubble oscillations due to low power US pulses, which are non-resonant. (B) Internal cavitation: high-power pulses will lead to amplified resonance and finally collapse of the bubble.
Figure 9:

Stable and internal cavitation. (A) Stable cavitation: bubble oscillations due to low power US pulses, which are non-resonant. (B) Internal cavitation: high-power pulses will lead to amplified resonance and finally collapse of the bubble.

Figure 10: Proposed mechanism for (A) cavitation effect on micelle rupture, (B) liposome collapse.
Figure 10:

Proposed mechanism for (A) cavitation effect on micelle rupture, (B) liposome collapse.

Microbubbles (MB), were originally designed as a contrast agent for US imaging. However, they are also able to function in drug delivery applications [2]. MBs consist of a gas-filled core that can contain air, nitrogen, or perfluorocarbons (PFC), which is surrounded by a layer/layers of lipids, proteins, or polymers [2]. PFC compounds loaded into MBs have a pronounced ability to dissolve oxygen. Accordingly, Chang et al. investigated the delivery of oxygen along with a fluorescent dye [1,10-dioctadecyl-3,3,30,30-tetramethylindocarbocyanine iodide] (DiI), a model drug, loaded into the liposomal MBs. Using US and fluorescence imaging in ex vivo tissue models, it was found that multifunctional MBs could provide simultaneous controlled delivery of oxygen and therapeutic agents [166]. Generally speaking, MBs, themselves, have a limited drug loading capacity. This limitation requires using highly active drugs [167]. Although some drugs such as DOX can be loaded onto the coated lipid bilayer of MBs via electrostatic attraction and hydrophobic forces (only for lipophilic drugs), liposome-encapsulated drugs that are conjugated or attached to external MBs can have a higher loading efficiency [168]. Yu et al. performed studies on combining liposomes with MB for delivery of DOX utilizing US. As demonstrated in Figure 11, this method led to increased loading of the drugs, thus, achieving improved results. Their results showed that it was possible for a MB to carry about 1600 liposomes. Furthermore, low-pressure US pulses resulted in more efficient drug release [168].

Figure 11: Potential ability of microbubbles to carry a large number of liposomes and, hence, increasing the drug loading capacity (see text).
Figure 11:

Potential ability of microbubbles to carry a large number of liposomes and, hence, increasing the drug loading capacity (see text).

Recent advancements in the field of US-responsive liposomal DDSs have focused on designing the so-called “sonosensitive” materials to be employed in liposomes to achieve better results. These efforts include a study conducted by Evjen et al. who found promising results using liposomes composed of 1,2 distearoyl-sn-glycero-3-phosphatidylethanolamine (DSPE) and liposomes comprised of dioleoylphosphatidylethanolamine (DOPE) [169], [170]. In vitro study of DSPE showed about 70% release of DOX within 6 min of US exposure, which was a sevenfold improvement compared to standard PEGylated liposomal DOX [170]. DOPE in vitro showed even better results: releasing 95% of DOX from liposomes in 6 min of 40-kHz US exposure; a 35% increase compared with using DSPE, and a ninefold increase compared to standard PEGylated liposomes [169]. Other than using the aforementioned sonosensitive agents to modify the phospholipid bilayer, it is also possible to use bile salts (BS) to make vesicles with higher sensitivity to an US stimulus. It has been reported that BSs destabilize the structure of the phospholipid bilayer [171]. Recently, Mujoo et al. attempted to use BSs (e.g. cholate, chenodeoxycholate, ursodeoxycholate, glycocholate, and taurocholate) to make improved US-sensitive phospholipid bilayers. They also demonstrated significant responses to low-frequency US, whereas non-significant responses were obtained for high-frequency US. They also found that liposomes made of taurocholate and DOPE showed less sensitivity to high-frequency US than liposomes containing only DOPE. It was speculated that this phenomenon was because of a preferred shape arising from the combination of taurocholate and DOPE in the phospholipid bilayer [172].

Sonoporation is another effect that can be exerted by US action on nanocarriers, which is defined as the formation of temporary pores in the cellular membrane. The generation of these pores is an opportunity for the penetration of drugs into the cells especially for delivery of drugs into the brain where the blood brain barrier (BBB) is the most important barrier against drug uptake. The underlying mechanism for such pore formation is mainly due to cavitation and also hyperthermia associated with MB US activation [161], [173]. For example, Lin et al. studied cationic liposomes (CLs) for the delivery of DOX to brain tumors using focused US. The idea of using cationic liposomes originated from the general characteristic of tumor cells, i.e. their more negatively charged surfaces. Their investigation on glioma, a common brain tumor, showed promising results where tumors became smaller, and animal survival was increased [174].

As mentioned previously, MBs generally can be used for the delivery of highly active drugs. However, siRNA, miRNA, and plasmid DNA also have the potential to produce a biological effect at comparatively low doses. This non-viral gene delivery can also be combined with US. Grayburn et al. applied this method with cationic liposomes for diabetes treatment using RIP 3.1, an insulin-specific promoter, to trigger the expression of gene NeuroD. Their histological studies on diabetic rats showed the regeneration of pancreatic islets employing this technique [167].

In spite of many studies on US for drug delivery, it is still difficult to be quantitatively precise concerning US parameters. This is mostly because different US methods use different frequencies and materials. Moreover, the fact that cavitation is also accompanied by hyperthermia makes it difficult to study the effect of cavitation alone on lipid-based nanocarriers. One method to study the cavitation effects more specifically is using different frequencies and investigating the effects of temperature and/or US pulses on drug release [160].

Magneto responsive

In recent years, lipid-based NPs that also contain magnetic substances, the so-called magnetic/magneto-liposomes (MLs) have been developed for the diagnosis and therapy of miscellaneous diseases. As an example, metallic ions and magnetic NPs (MNPs) can be bound to, or encapsulated into, liposomes during the synthetic processes. Metallic ions such as Gd (III) have been used as a contrast agent (CA) for magnetic resonance imaging (MRI). In comparison with metallic ions, MNPs cannot only be used as CAs but also have other applications such as magnetic targeting, therapeutic hyperthermia, and magnetically triggered drug release [175]. Passive and active targeting of magnetic liposomes are potentially applicable in delivering other therapeutic agents at the same time. This ability can lead to the design of multimodal systems for theranostic purposes.

The combination of various methods such as hyperthermia and photodynamic therapy (PDT) is an interesting strategy to enhance therapeutic efficiency, and ultramagnetic liposomes can be employed for this purpose. In one study, the bilayers of liposomes that were highly loaded with MNPs in the hydrophilic core and also contained a photosensitizer (FOSCAN™) in the hydrophobic region were used for cancer therapy. Stimulation with an alternating magnetic field raised the temperature of the MNPs. Besides, magnetic NPs also played a role as MRI CAs. Laser excitation activated the photosensitizer, and the surrounding oxygen was changed into a reactive form (singlet oxygen). Also, the fluorescence emission of the photosensitizer could be used for imaging. The coupling of magnetic hyperthermia and PDT potentiated the tumor destruction caused by the singlet oxygen. The mechanism of this process was triggering apoptotic pathways that led to more complete cancer cell and tumor regression in vivo. This approach could avoid surgical damage and preserve the surrounding normal tissues [176].

Magnetic fluid-loaded liposomes (MFLs) have been reported to address the problem of inefficient chemotherapy of malignant tumors in the brain and central nervous system. In one study, labeled PEGylated liposomes loaded with maghemite crystals could pass the BBB. To further investigate this approach, glioblastoma cells were implanted intracerebrally in nude mice. Then, MFLs were intravenously injected. Because of the vasculature permeability induced by a magnetic field gradient applied to the heads of the mice, the tumor was targeted both by EPR and magnetic targeting. Thus, the concentration of the MFLs was selectively increased in the malignant area. Also, a 7T small animal MRI system was used to monitor the process at the same time [177].

An anti-EGFR antibody conjugated to magnetic liposomes was used as an active targeting system, for co-delivery of drugs and genes. In this platform, DOX was loaded into mesoporous silica magnetic NPs coated by a liposomal bilayer. Mesoporous silica not only facilitated high drug loading (98% efficiency) but also enhanced the interaction of water molecules with the magnetic core, which enhanced the MRI contrast. Here, Plk1 siRNA was first treated with protamine to form a complex; afterward, the complex was incorporated into the liposomal system by the ammonium bicarbonate gradient technique. pH-sensitive drug release was shown at acidic pH values. These multifunctional theranostic particles were tested as image-guided therapy on BxPC3 pancreatic cancer cells [178].

The increased temperature resulting from magnetic-stimulated MLs can damage the healthy surrounding tissue. To prevent this side effect, ~5-nm nitrodopa-palmityl stabilized magnetic NPs were embedded into the liposome bilayers. Magnetic stimulation increased the temperature to the melting point of the liposome bilayers, which caused a change in the permeability. Controlled release and optimized drug loading were the most significant achievement [179]. Another research demonstrated that simultaneous encapsulation of SPIONs and methylene blue dye into the lipid bilayer-coated mesoporous silica NPs (MSNs) could function as a remote smart-triggered DDS. MSNs provided higher encapsulation efficiency and allowed interaction of the SPIONs with lipid bilayers, and the effect of alternating magnetic force on the melting temperature of lipid bilayers made the gates stimulus responsive [180]. Elsewhere, magnetite NPs were deposited on the surface of melamine formaldehyde particles that were coated by polyelectrolyte. The core was dissolved to form a microcapsule structure. The layer-by-layer assembly led to a covering of the temperature-responsive lipid bilayers on the surface of the microcapsules. Under a magnetic stimulus, the microcapsule could trap and release the dye [181]. Recently, Salvatore et al. reported a new type of MLs that contained a fluorescent hydrophilic drug combined with a single-stranded oligonucleotide (ON) conjugated to a cholesteryl unit of the liposome membrane. Two types of MNPs were employed in this platform. A hydrophobic MNP was located in the liposome bilayer, and an ON-decorated hydrophilic gold core-shell MNP was connected to the single-stranded ON on the surface. Controlled and sequential release could be obtained by the frequency changes in an alternating magnetic field (AMF). Five minutes of 3.22-kHz AMF caused the release of a model drug, whereas 15 min of 6.22-kHz AMF released the DNA because the temperature increased to the melting point of the DNA (Figure 12) [182].

Figure 12: Schematic illustration of sequential controlled release of dual MNP-equipped liposomes. Reprinted with permission from Ref. [182], copyright 2016, American Chemical Society.
Figure 12:

Schematic illustration of sequential controlled release of dual MNP-equipped liposomes. Reprinted with permission from Ref. [182], copyright 2016, American Chemical Society.

MRI is a non-invasive diagnostic tool, where the excitation of the nuclear spin leads to an energy transition, and the time for the relaxation of the nuclei can be detected by MRI [183]. In recent years, DDSs have been suggested to simultaneously serve as drug carriers and CAs for enhancing MRI images. This aim could be achieved by increasing longitudinal relaxivity. The study of the abnormal muscle contraction or investigating brain physiology is facilitated by MRI. Fluctuations in cellular calcium (Ca2+) can be used in an intelligent CA for MRI. This probe has features such as specific local tissue targeting, improvement in retention time, and modifying and amplifying the MRI signals. To this end, a formulation of a novel amphiphilic ligand in the liposome structures that contained a Gd(III) complex was reported as a CA. Ca2+ binding to magnetic liposomes changed the coordination of the Gd-chelate. In fact, Ca2 increased hydration and reduced the local motion of the Gd(III) complex, and the increase in relaxivity could be detected even at low-magnetic field values. Therefore, this type of smart CA triggered by molecules or ions involved in neuronal signaling can help neuroscience studies for the better understanding of brain function [184]. Aptamer-conjugated thermosensitive liposomes encapsulating Gd-DTPA were employed as an MRI probe. These biocompatible MLs enhanced cellular uptake, due to active targeting whereby the Gd CA was released at 42°C via mild hyperthermia. Such a CA delivery system could improve early cancer diagnosis [185]. Another study by Zhang et al. showed that the encapsulation of Mn2+ in liposomal NPs not only could be monitored by MRI but also facilitated high encapsulation of As2O3 in the liposomes. Actually, the reaction of Mn2+ and As2O3 produced a complex structure (Mn-As complex). The complex formation and deformation affected the darkness and brightness of the MRI image contrast. Mn2+ alone was a dark T2-weighted CA, whereas the Mn-As complex was a bright T1-weighted CA. This convertible MRI platform could provide drug release inside glioblastoma cells, and furthermore, a high dosage of As overcame the temozolomide resistance of glioblastoma [186].

Carbon nanostructures are promising materials and can be combined with MLs. In one study, buckyball (fullerene C60) was utilized for surface decoration of magnetic NPs and also for the adsorption of hydrophobic drugs or fluorescent dyes by a π-π stacking mechanism. Folate-targeted thermosensitive liposomes encapsulated PEG-functionalized C60-Fe3O4 nanocomposites and docetaxel (DTX) as a multifunctional system. Herein, the fullerene hybrid system was shown to increase the temperature after radio-frequency radiation. Fullerenes could be used for radiofrequency-triggered release of DTX and with increased accumulation of the fullerenes and DTX in tumor tissues. Histological studies showed that the coupling of radiofrequency and magnetic field targeting of this multifunctional system caused cell necrosis, lysis, and fragmentation [187].

Recently, a new type, “nanorobots”, was introduced where the combination of liposomes, carbon nanohorns (CNH), and magnetic iron NPs were investigated as a nanotransporter. This structure allowed monitoring of cancerous cells even at the single-cell level. The liposome structure encapsulated the fluorescein di-β-d-galactopyranoside (FDG) as a quenched non-fluorescent agent that could be activated by β-galactosidase, which was conjugated with a “teddy bear”-shape, made of avidin-biotin-polyethylenimine (PEI)-MAG-functionalized-CNH-ox hybrid. These hybrids were investigated as a novel theranostic platform (Figure 13D). Confocal microscopy showed the magnetic field effect to improve transfection yield in HeLa cells in vitro (magnetotransfection). Cytotoxicity tests confirmed the biocompatibility of the nanohybrids in the presence of the magnetic field. Finally, the nanotransporter was applied for the remote control of enzymatic reactions in vivo (Figure 13). To illustrate this ability, FDG encapsulated nanotransporters injected into transgenic mice that expressed human β-galactosidase in the whole body. Simultaneous magnetic field and laser induction were used to trap the nanotranporters in the vessels of the mouse ear, and release the FDG. Because of the β-Gal enzymatic reaction, the green fluorescence of fluorescein was produced by hydrolysis of FDG and could be directly observed in vivo [188].

Figure 13: In vivo remote control of enzymatic reactions using nanotransporters. (A) Schematic image of the experiment. (1): injection of nanotransporters, (2): Nanotransporters accumulated by a magnet, (3): Manipulation and observation of physicochemical aspects of nanotransporters with NIR radiation. (B) Accumulation of nanotransporters in the vessels by magnetic field. (C) In vivo β-Gal reaction observed in the blood vessel of a living mouse. (D) Nanotransporter structure. Reprinted with permission from Ref. [188], copyright 2016, John Wiley & Sons, Inc.
Figure 13:

In vivo remote control of enzymatic reactions using nanotransporters. (A) Schematic image of the experiment. (1): injection of nanotransporters, (2): Nanotransporters accumulated by a magnet, (3): Manipulation and observation of physicochemical aspects of nanotransporters with NIR radiation. (B) Accumulation of nanotransporters in the vessels by magnetic field. (C) In vivo β-Gal reaction observed in the blood vessel of a living mouse. (D) Nanotransporter structure. Reprinted with permission from Ref. [188], copyright 2016, John Wiley & Sons, Inc.

Other stimulus-responsive smart liposomal carriers

Matrix metalloprotease 2 (MMP2)

Novel multifunctional DDSs for cancer targeting that are responsive to matrix metalloprotease 2 (MMP2), which is often upregulated in a tumor environment, have been developed. In a recent study, a liposomal nanocarrier was designed comprised of (PEG (providing elongated blood circulation and preventing non-specific interactions) on the liposomal surface, which was also conjugated to an antinucleosome monoclonal antibody (mAb2C5). Because of the EPR phenomenon, this liposomal nanocarrier accumulated in the location of the tumor (due to the nanocarrier size) and also exhibited active targeting by the mAb2C5. Therefore, in the microenvironment of the tumor, the MMP2-sensitive linker between PEG and lipid was cleaved because of up-regulated extracellular MMP2; thus, the protective long-chain PEG was removed. This led to the exposure of the hidden surface-functionalized cell-penetrating peptides (TATp), which facilitated intracellular delivery of the drugs. Therefore, targeted and enhanced cell internalization of the nanocarriers was obtained [189].

In another study, copolymers based on the self-assembly of PEG-phosphoethanolamine (PEG-pp-PE) demonstrated MMP2-sensitive drug delivery to tumor cells, and the drug efflux mediated by P-glycoprotein (P-gp) was prevented. In this research, different homologous of PEG-pp-PE were studied. This investigation proved that the total structure of PEG-peptide-lipid was crucial for the inhibition of P-gp, and the equilibrium between its hydrophilic and lipophilic parts was important. Here, PEG2k-pp-PE gave a higher inhibition of P-gp. This copolymer exhibited a promising tumor-targeted drug delivery, which was connected with the enhancement in plasma membrane fluidity and the inhibition of the activity of P-gp ATPase [190].

Redox

A study by Noyhouzer et al. demonstrated the controlled payload release of a ferrocene-modified phospholipid unilamellar liposome as a drug delivery vehicle by a redox-active mechanism. This takes advantage of the fact that the redox potential is higher inside cancer cells due to a higher concentration of glutathione. Here, the exposure of the ferrocene groups on the surface of the vesicle were shown, thus, providing for triggering the redox reaction. Therein, the results of flow cytometry evaluating the drug release in HeLa cells indicated a 200-times stronger signal for the cells treated with redox active vesicles. This implicated an efficient redox-sensitive delivery and provided specificity to the cancer cells [191].

Dual-sensitive liposomes

Stimuli-responsive characteristics are among the efficient approaches to improve the precision of drug delivery; thus, to date, diverse stimuli-responsive systems have been designed [192], [193], [194]. Liposomes can discharge their cargos in response to temperature, pH, light, and redox signals. Normally, the stimulus results in alterations in the structure of the lipid, as well as the destabilization and leakage of the liposome contents. In this section, we will cover liposomes that respond to a combination of these different stimuli [195], [196]. The most common format of dual-responsive liposomes is an “AND” logic gate where both stimuli must be present at the same time to get a drug release.

pH/temperature responsive

Combined temperature-responsive and pH-sensitive liposomes have been investigated as they show controlled release when exposed to increased temperature and an acidic surrounding milieu [68], [134]; thus, the design of the nanocarriers responsive to both temperature and pH changes provides new opportunities for controlled drug delivery [197], [198]. In one study, dual-stimuli responsive liposomes reacting to both pH and temperature were fabricated with biocompatible hyperbranched poly (glycidol) (HPG) containing pH-labile succinyl moieties and temperature-responsive oligo (ethylene glycol) (OEG) moieties. This nanoplatform switched from hydrophilic to hydrophobic in response to pH and temperature changes of the surrounding environment, and the NPs were destabilized [197]. In another study, Chen et al. synthesized a liposomal gel formulation, sensitive to alterations in both temperature and pH. This platform was used for vaginal administration of arctigenin employing a cleavable methoxy polyethylene glycol 2000-hydrazonecholesteryl hemisuccinate (mPEG-Hz-CHEMS). Thus, high stability was found at neutral pH, but in acidic conditions (pH 5), the platform was destabilized via bond cleavage. The results showed that the vaginal gel loaded with arctigenin was an effective formulation for the treatment of vaginal candidiasis with less toxicity [198]. In another study, Kono et al. synthesized hyperbranched poly (glycidol) containing N-isopropylamide (NIPAM), as a thermo-responsive moiety, and succinylate groups as the pH-labile component. In response to a decrease in pH and an increase in temperature, the polymer NP altered from hydrophilic to hydrophobic [199]. Such polymers have been also incorporated onto stable EYPC liposomes to generate dual sensitive liposomes and could deliver payloads into the cytosol of target cells under an acidic environment and local heating [200]. Furthermore, Wu et al. demonstrated that an MSN as the drug-encapsulating core and a copolymer-lipid layer comprised of phospholipids and poly(N-isopropylacrylamide-methacrylic acid-octadecyl acrylate) as the dual-responsive shell could allow the delivery of a drug such as DOX. Here, changes in temperature and pH could each trigger the release of drugs in response to more acidic and higher temperature of cancerous cells, respectively [201].

Temperature/ultrasound responsive

Thermo-sensitive liposomes (TSLs) experience a phase alteration when they are heated, and this phenomenon makes them permeable inducing the release of their cargo [202]. Conventional TSLs are activated between a temperature range of 42°C and 45°C, and release their cargo over 30 min, while LTSLs discharge their drugs at a temperature range of 39°C to 40°C in a few seconds [203].

High-intensity focused HIFU is able to ablate tumors by employing rather long and continuous exposures to produce high temperatures for tumor thermal ablation [204]. Dromi et al. have combined pulsed HIFU with LTSLs to elevate the local delivery of DOX into tumors. The results showed the improvement of the antitumor effects of the drugs [205]. Elsewhere, Chen et al. designed a curcumin-encapsulated liposome MB gel comprised of N-cholesteryl hemisuccinate-O-sulfate chitosan (NCHOSC). This liposomal MB had a high loading efficiency for curcumin. The temperature-responsive CS/GP gel containing liposomal MBs assumed a gel format body temperature where the initial release of curcumin induced by US was found to be 85%. In Figure 14A, a schematic image of the process and the release diagram are shown. The results implied the fact that the double responsive curcumin-loaded liposomal MB gel served as a potential ultrasound and temperature dual-responsive DDS [206]. Oerlemans et al. developed fluorescent liposomes (FCLs) that mediated ablation with the aid of MR-HIFU for the treatment of non-palpable breast lesions. FCLs could liberate their fluorescent contents after 30 s of exposure to MR-HIFU. After exposure and release of the contents, the treated lesions could be imaged by fluorescence excited by UV light. PEG and cholesterol were included in the liposome to provide more stability and elongated circulation time. Through introducing thermo-sensitive phospholipids like 1,2-dipalmitoyl-sn-glycero-3-phosphocholine into the liposomes, MR-HIFU-mediated fluorescein release from the liposome could be accomplished [207].

Figure 14: (A) Schematic illustration of Cur-LM-G and the external morphology of the blank liposome microbubble gel is shown at the middle: (A) under the temperature of gelation, (B) gelation process, (C) the formed Cur-LM-G under the temperature of gelation, and (D) gelation. (B) The percentage of curcumin that is released from Cur-LM-G in the presence and without the presence of US. Reprinted with permission from Ref. [206], copyright 2013, Elsevier.
Figure 14:

(A) Schematic illustration of Cur-LM-G and the external morphology of the blank liposome microbubble gel is shown at the middle: (A) under the temperature of gelation, (B) gelation process, (C) the formed Cur-LM-G under the temperature of gelation, and (D) gelation. (B) The percentage of curcumin that is released from Cur-LM-G in the presence and without the presence of US. Reprinted with permission from Ref. [206], copyright 2013, Elsevier.

pH/redox responsive

The combination of pH and redox is another category of dual-response liposomes. It is known that the pH of the tumor microenvironment is between 6 and 7, which is lower than physiological pH. It also becomes more acidic (e.g. pH between 5 and 6) inside the intracellular endocytic vesicles [208]. Additionally, glutathione (GSH) has a significantly higher concentration inside cells, making it a promising option for smart intracellular drug delivery. Recently, Zhang et al. developed a cleavable polymerized liposome (CPL) for the delivery of pharmaceutical agents. A gallate-derived group with three propargyl moieties was joined to palmitoyl oleoyl phosphoethanolamine (POPE). To create large unilamellar vesicles, the mentioned anionic lipid was formulated with other lipids including palmitoyl oleoyl phosphatidyl choline. Employing the Cu(I)-catalyzed click reaction between the propargyl moieties and the azides in the crosslinker resulted in the polymerization of the unilamellar vesicles. The polymerized liposomes equipped with a disulfide or ketal group in the crosslinker could be depolymerized and discharge their payload through the addition of a reducing thiol or under acidic conditions, respectively [209]. Elsewhere, a PEGylated nano-sized polymeric lipid vesicle (PPLV) loaded with DOX exhibited dual pH and reduction responsiveness. The system enabled efficient antitumor drug delivery via triggered drug release, as well as enhancing tumor cellular internalization. The PPLVs were reported to have good uptake by tumor cells under acidic conditions. On the other hand, an accelerated DOX release at pH 5 as well as with 10 mM of GSH proved its promising potential [210]. In a similar study, another dual-responsive DDS containing folate-PEG-coated polymeric lipid vesicles (FPPLVs) was constructed showing an enhanced DOX release profile under acidic conditions in comparison to neutral pH values due to the loss of PEG in acidic conditions, which, in turn, made it easy for tumor cells to take up the NPs. Fluorescence microscopy of HeLa cells also showed that disrupting the FPPLV structure in response to 2–10 mM of intracellular GSH levels triggered drug release in tumor cells [211].

Xu et al. designed acid/redox dual responsive liposomes, comprised of soy phosphatidylcholine (SPC), 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine (DOPE), and a synthetic functional lipid, 2-[2-(2-carboxylcyclohexylformamido)-3,12-dioxy-1-(1H-imidazolyl-4)-7,8-dithio-4,11-diazapentadecylamide]-glutaric acid ditetradecanoldiester (HH-SS-E2C14). HH-SS-E2C14 contained histidine, an amino acid moiety, and acid-cleavable hexahydrobenzoic (HHB) amide, as a hydrophilic block, and two tetradecyl alkane chains as a hydrophobic block. A disulfide bond was integrated into the structure as a redox-labile linkage between the two blocks. When this formulation was administered intravenously, it had a negative charge, but after the EPR enabled the accumulation in tumors, the surface charge became positive due to the interaction with cell membranes, which were negatively charged. Subsequently, after endocytosis, the imidazole group was protonated and promoted the proton influx into the endocytic vesicle followed by an increase in osmotic pressure and rupture of the endo-lysosomes, allowing endo-lysosomal escape of the cargo [212].

Other dual responsive liposomes

pH/enzyme

Zhang et al. created a dual stimuli-responsive liposome, responsive to pH and esterase for the delivery of camptothecin and siRNA for the treatment of cancer. This dual CPT-PCB (poly carboxy betaine)/siPIK1 lipoplex preparation showed a high capacity for loading of CPT and could boost the serum stability of the liposomes and provide controlled release of CPT and siRNA, due to the protonation of PCB in endosome/lysosomes, and esterase and pH sensitivity of the CPT-PCB prodrug. This system was reported to be a promising delivery platform for cancer therapy and co-delivery of CPT and siPIK1 [213].

Dual redox

Moreover, a co-assembly system comprised of block copolymers containing diselenide and polymeric lipid could show dual redox responsiveness because it could be disrupted when exposed to either allow the concentration of hydrogen peroxide (as an oxidant) or glutathione (as a reductant). In the presence of 0.05 mM of GSH or 0.1% H2O2, most of the co-assembly system was disrupted. Also, the results illustrated that hydrophobic interactions provided the driving force of the co-assembly [214].

Redox/ultrasound

In 2013, Nahire et al. showed that polymer-coated echogenic lipid NPs could release their content via redox triggering and in response to diagnostic frequency US. As a result, a 3-MHz US combined with a reducing agent increased the drug release up to 96%. Such lipid NPs were stable in oxidizing environments and released their content in the reducing environment of the cytosol. The rate of drug release was boosted by simultaneous application of redox trigger and US [215].

Conclusion

Liposomes have long been one of the most studied types of drug delivery nanosystems. In addition to the other advantages described in some detail above, another overriding reason lies in their remarkably low toxicity. Widespread concern about possible hazardous effects of diverse nanostructures such as carbon nanotubes, fullerenes, noble metal nanostructures, various inorganic NPs, quantum dots, and so forth has led to the establishment of a brand new area of toxicology called “nanotoxicology”. In contrast to many of these nanostructures, liposomes are principally composed of naturally occurring lipids that are naturally found as constituents of cell membranes that compose all the cells of our bodies. The body is well used to dealing with the basic components of liposomes, so they can be regarded as the ultimate in biocompatibility and biodegradability. This is probably why there are more liposomal preparations that have gained regulatory approval for medical applications, than any other class of NPs. Now that liposomes have almost progressed to being the “standard of care”, the road is open for researchers to improve upon them by active targeting strategies, and by engineering various stimulus-responsive and smart characteristics.

Smart modification of liposome is an opportunity to impart a stimulus-responsive moiety to the structure. Herein, encapsulated cargos can be released through activating the stimuli-responsive part followed by alterations in the structure of liposome, e.g. conformations, or bond cleavage and the structural destabilization of the carrier, thus, leading to the leakage of the cargos. These alterations can be triggered on-demand and with highly precise control via internal stimuli (pH, reducing agents, specific enzymes) or external stimuli (light, magnetic field, or US). In this regard, smart stimuli-responsive liposome vehicles are opening new horizons for drug/gene delivery applications and innovative therapies for various diseases and disorders, especially in clinical evaluations and exploitations.

About the authors

Parham Sahandi Zangabad

Parham Sahandi Zangabad received his MSc in Nanomaterials and Nanotechnology from Sharif University of Technology (SUT), Tehran, Iran. Since 2012 until January 2015, he was a research assistant at the Research Center for Nanostructured and Advanced Materials (RCNAM) at SUT working on nanocomposites and their mechanical properties. Since late 2014, he joined Professor Karimi’s Advanced Nanobiotechnology and Nanomedicine Research Group (ANNRG) in Iran University of Medical Science, Tehran, Iran, having collaboration with Professor M.R. Hamblin from Harvard Medical School, Boston, MA, USA. To research in nanomedicine, smart drug delivery systems, and biosensors was the aim of this collaboration, resulting in several high-impact review articles and books. Afterward, since February 2016, he became a research assistant in the Research Center for Pharmaceutical Nanotechnology (RCPN), Tabriz University of Medical Sciences, Tabriz, Iran, working on nanocarriers for anticancer drug delivery systems. This led to beneficial experiences in design and applying smart nanocarriers in biomedical applications. He was then a research assistant in the Cellular and Molecular Research Center in Iran University of Medical Sciences, since 2017, working on drug/gene delivery systems.

Shayan Shahsavari

Shayan Shahsavari is a member of the Nanoclub Elites Association (NEA) and CEO at Mataab Company, at Tehran, Iran. He joined NEA at the Iran Nanotechnology Initiative Council as a member and as an expert nanobiotechnologist. In 2014, he established Mataab Co to manufacture and commercialize high-tech products at a biotechnology incubator, Pasteur Institute of Iran. His fields of research are theranostics and tissue engineering. Recently, he joined the ANNRG, Iran University of Medical Sciences to collaborate with Professor Mahdi Karimi in the related fields of Nanomedicine including Nanobio-sensors and Drug delivery systems.

Mahdi Karimi

Mahdi Karimi received his BSc degree in Medical Laboratory Science from Iran University of Medical Science (IUMS), in 2005. In 2008, he achieved the MSc degree in Medical Biotechnology from Tabriz University of Medical Science and joined Tarbiat Modares University as a PhD student in the Nanobiotechnology field and completed his research in 2013. During his research, in 2012, he affiliated with the laboratory of Professor Michael Hamblin in the Wellman Center for Photomedicine at Massachusetts General Hospital and Harvard Medical School as a visiting researcher, where he contributed to the design and construction of new smart nanoparticles for drug/gene delivery. On finishing the study, he joined, as assistant professor, the Department of Medical Nanotechnology at IUMS. His current research interests include smart nanoparticles’ design in drug/gene delivery and microfluidic systems. He has established a scientific collaboration between his lab and Professor Michael Hamblin’s lab to design new classes of smart nanovehicles in drug/gene delivery systems. Dr. Karimi was recipient of several grants from Iran Science Elites Federation (ISEF), and Iranian Nanotechnology Initiative Council. He was also ranked 34th among all faculty members in Iran, by Iran Science Elites Federation (ISEF) in Iran and selected as the best young researcher of Iran University of Medical Sciences in 2016.

Michael R. Hamblin

Michael R. Hamblin, PhD, is a principal investigator at the Wellman Center for Photomedicine, Massachusetts General Hospital, an associate professor of Dermatology, Harvard Medical School and the affiliated faculty of Harvard, MIT Division of Health Science and Technology. He directs a laboratory of around 12 scientists who work in photodynamic therapy and low-level light therapy. He has published over 380 peer-reviewed articles and documents (according to scopus), is associate editor for eight journals, and serves on NIH study sections. He has edited 10 proceedings volumes, together with four other major textbooks on PDT and Photomedicine. In 2011, Dr. Hamblin was honored by election as a Fellow of SPIE.

  1. Funding: National Institute of Allergy and Infectious Diseases, (Grant/Award Number: ‘R01AI050875 and R21AI121700’).

References

[1] Karimi M, Zare H, Bakhshian Nik A, Yazdani N, Hamrang M, Mohamed E, Sahandi Zangabad P, Moosavi Basri SM, Bakhtiari L, Hamblin MR. Nanotechnology in diagnosis and treatment of coronary artery disease. Nanomedicine 2016, 11, 513–530.10.2217/nnm.16.3Search in Google Scholar PubMed PubMed Central

[2] Karimi M, Ghasemi A, Sahandi Zangabad P, Rahighi R, Moosavi Basri SM, Mirshekari H, Amiri M, Pishabad ZS, Aslani A, Bozorgomid M, Ghosh D. Smart micro/nanoparticles in stimulus-responsive drug/gene delivery systems. Chem. Soc. Rev. 2016, 45, 1457–1501.10.1039/C5CS00798DSearch in Google Scholar

[3] Sahandi Zangabad P, Karimi M, Mehdizadeh F, Malekzad H, Ghasemi A, Bahrami S, Zare H, Moghoofei M, Hekmatmanesh A, Hamblin MR. Nanocaged platforms: modification, drug delivery and nanotoxicity. Opening synthetic cages to release the tiger. Nanoscale 2017, 9, 1356–1392.10.1039/C6NR07315HSearch in Google Scholar

[4] Karimi M, Sahandi Zangabad P, Baghaee-Ravari S, Ghazadeh M, Mirshekari H, Hamblin MR. Smart nanostructures for cargo delivery: uncaging and activating by light. J. Am. Chem. Soc. 2017, 139, 4584–4610.10.1021/jacs.6b08313Search in Google Scholar PubMed PubMed Central

[5] Karimi M, Zangabad PS, Ghasemi A, Hamblin MR. Smart External Stimulus-Responsive Nanocarriers for Drug and Gene Delivery, Morgan & Claypool Publishers: San Rafael, CA, USA, 2015, ISBN: 978-1-6817-4202-1.10.1088/978-1-6817-4257-1Search in Google Scholar

[6] Karimi M, Zangabad PS, Ghasemi A, Hamblin MR. Smart Internal Stimulus-Responsive Nanocarriers for Drug and Gene Delivery: IOP Concise Physics, 2015. ISBN: 978-1-6817-4257-1. doi: 10.1088/978-1-6817-4202-1.10.1088/978-1-6817-4202-1Search in Google Scholar

[7] Barar J, Aghanejad A, Fathi M, Omidi Y. Advanced drug delivery and targeting technologies for the ocular diseases. Bioimpacts 2016, 6, 49.10.15171/bi.2016.07Search in Google Scholar PubMed PubMed Central

[8] Massoumi B, Poorgholy N, Jaymand M. Multistimuli responsive polymeric nanosystems for theranostic applications. Int. J. Polym. Mater. Polym. Biomater. 2017, 66, 38–47.10.1080/00914037.2016.1180622Search in Google Scholar

[9] Karimi M, Moosavi Basri SM, Vossoughi M, Pakchin PS, Mirshekari H, Hamblin MR. Redox-sensitive smart nanosystems for drug and gene delivery. Curr. Org. Chem. 2016, 20, 2949–2959.10.2174/1385272820666160510154557Search in Google Scholar

[10] Karimi M, Bahrami S, Ravari SB, Zangabad PS, Mirshekari H, Bozorgomid M, Shahreza S, Sori M, Hamblin MR. Albumin nanostructures as advanced drug delivery systems. Exp. Opin. Drug Deliv. 2016, 13, 1609–1623.10.1080/17425247.2016.1193149Search in Google Scholar PubMed PubMed Central

[11] Akbarzadeh A, Rezaei-Sadabady R, Davaran S, Joo SW, Zarghami N, Hanifehpour Y, Samiei M, Kouhi M, Nejati-Koshki K. Liposome: classification, preparation, and applications. Nanoscale Res. Lett. 2013, 8, 1.10.1186/1556-276X-8-102Search in Google Scholar

[12] Bangham A, Standish MM, Watkins J. Diffusion of univalent ions across the lamellae of swollen phospholipids. J. Mol. Biol. 1965, 13, 238–IN27.10.1016/S0022-2836(65)80093-6Search in Google Scholar

[13] Pattni BS, Chupin VV, Torchilin VP. New developments in liposomal drug delivery. Chem. Rev. 2015, 115, 10938–10966.10.1021/acs.chemrev.5b00046Search in Google Scholar

[14] Deshmukh RR, Gawale SV, Bhagwat MK, Ahire PA, Derle ND. A review on: liposomes. World J. Pharm. Pharm. Sci. 2016, 5, 506–517.Search in Google Scholar

[15] Jone A. Liposomes: a short review. J. Pharm. Sci. Res. 2013, 5, 181–183.Search in Google Scholar

[16] Torchilin VP. Recent advances with liposomes as pharmaceutical carriers. Nat. Rev. Drug Discov. 2005, 4, 145–160.10.1038/nrd1632Search in Google Scholar

[17] Barhoumi A, Liu Q, Kohane DS. Ultraviolet light-mediated drug delivery: principles, applications, and challenges. J. Control. Release 2015, 219, 31–42.10.1016/j.jconrel.2015.07.018Search in Google Scholar

[18] Barenholz YC. Doxil®—the first FDA-approved nano-drug: lessons learned. J. Control. Release 2012, 160, 117–134.10.1016/j.jconrel.2012.03.020Search in Google Scholar

[19] Barenholz YC. Doxil®–the first FDA-approved Nano-drug: from basics via CMC, cell culture and animal studies to clinical use. In Nanomedicines: Design, Delivery and Detection, Braddock, M, Ed., RSC Publishing: London, UK, 2016, 51, pp. 315–345. Available at: http://pubs.rsc.org/en/content/chapter/bk9781849739474-00315/978-1-84973-947-4.10.1039/9781782622536-00315Search in Google Scholar

[20] Leonard R, Williams S, Tulpule A, Levine A, Oliveros S. Improving the therapeutic index of anthracycline chemotherapy: focus on liposomal doxorubicin (Myocet™). Breast 2009, 18, 218–224.10.1016/j.breast.2009.05.004Search in Google Scholar

[21] Swenson C, Perkins W, Roberts P, Janoff A. Liposome technology and the development of Myocet™ (liposomal doxorubicin citrate). Breast 2001, 10, 1–7.10.1016/S0960-9776(01)80001-1Search in Google Scholar

[22] Chou H-H, Wang K-L, Chen C-A, Wei L-H, Lai C-H, Hsieh C-Y, Yang YC, Twu NF, Chang TC, Yen MS. Pegylated liposomal doxorubicin (Lipo-Dox®) for platinum-resistant or refractory epithelial ovarian carcinoma: a Taiwanese gynecologic oncology group study with long-term follow-up. Gynecol. Oncol. 2006, 101, 423–428.10.1016/j.ygyno.2005.10.027Search in Google Scholar PubMed

[23] Smith JA, Mathew L, Burney M, Nyshadham P, Coleman RL. Equivalency challenge: evaluation of Lipodox® as the generic equivalent for Doxil® in a human ovarian cancer orthotropic mouse model. Gynecol. Oncol. 2016, 141(2), 357–363.10.1016/j.ygyno.2016.02.033Search in Google Scholar PubMed

[24] Fassas A, Anagnostopoulos A. The use of liposomal daunorubicin (DaunoXome) in acute myeloid leukemia. Leuk. Lymphoma 2005, 46, 795–802.10.1080/10428190500052438Search in Google Scholar PubMed

[25] Petre CE, Dittmer DP. Liposomal daunorubicin as treatment for Kaposi’s sarcoma. Int. J. Nanomed. 2007, 2, 277.Search in Google Scholar

[26] Silverman JA, Deitcher SR. Marqibo®(vincristine sulfate liposome injection) improves the pharmacokinetics and pharmacodynamics of vincristine. Cancer Chemother. Pharmacol. 2013, 71, 555–564.10.1007/s00280-012-2042-4Search in Google Scholar PubMed PubMed Central

[27] Sarris A, Hagemeister F, Romaguera J, Rodriguez M, McLaughlin P, Tsimberidou A, Medeiros LJ, Samuels B, Pate O, Oholendt M, Kantarjian H. Liposomal vincristine in relapsed non-Hodgkin’s lymphomas: early results of an ongoing phase II trial. Ann. Oncol. 2000, 11, 69–72.10.1023/A:1008348010437Search in Google Scholar

[28] Rodriguez M, Pytlik R, Kozak T, Chhanabhai M, Gascoyne R, Lu B, Deitcher SR, Winter JN. Vincristine sulfate liposomes injection (Marqibo) in heavily pretreated patients with refractory aggressive non-Hodgkin lymphoma. Cancer 2009, 115, 3475–3482.10.1002/cncr.24359Search in Google Scholar PubMed

[29] Bowden R, Chandrasekar P, White MH, Li X, Pietrelli L, Gurwith M, van Burik JA, Laverdiere M, Safrin S, Wingard JR. A double-blind, randomized, controlled trial of amphotericin B colloidal dispersion versus amphotericin B for treatment of invasive aspergillosis in immunocompromised patients. Clin. Infect. Dis. 2002, 35, 359–366.10.1086/341401Search in Google Scholar PubMed

[30] Clemons KV, Stevens DA. Comparative efficacies of four amphotericin B formulations—Fungizone, Amphotec (Amphocil), AmBisome, and Abelcet—against systemic murine aspergillosis. Antimicrob. Agents Chemother. 2004, 48, 1047–1050.10.1128/AAC.48.3.1047-1050.2004Search in Google Scholar PubMed PubMed Central

[31] Glantz MJ, Jaeckle KA, Chamberlain MC, Phuphanich S, Recht L, Swinnen LJ, Maria B, LaFollette S, Schumann GB, Cole BF, Howell SB. A randomized controlled trial comparing intrathecal sustained-release cytarabine (DepoCyt) to intrathecal methotrexate in patients with neoplastic meningitis from solid tumors. Clin. Cancer Res. 1999, 5, 3394–3402.Search in Google Scholar

[32] Chamberlain MC. Neurotoxicity of intra-CSF liposomal cytarabine (DepoCyt) administered for the treatment of leptomeningeal metastases: a retrospective case series. J. Neuro-oncol. 2012, 109, 143–148.10.1007/s11060-012-0880-xSearch in Google Scholar PubMed

[33] Simon JA, Group ES. Estradiol in micellar nanoparticles: the efficacy and safety of a novel transdermal drug-delivery technology in the management of moderate to severe vasomotor symptoms. Menopause 2006, 13, 222–231.10.1097/01.gme.0000174096.56652.4fSearch in Google Scholar PubMed

[34] Gambling D, Hughes T, Martin G, Horton W, Manvelian G, Group S-DES. A comparison of Depodur™, a novel, single-dose extended-release epidural morphine, with standard epidural morphine for pain relief after lower abdominal surgery. Anesth. Analg. 2005, 100, 1065–1074.10.1213/01.ANE.0000145009.03574.78Search in Google Scholar PubMed

[35] Carvalho B, Roland LM, Chu LF, Campitelli III VA, Riley ET. Single-dose, extended-release epidural morphine (DepoDur™) compared to conventional epidural morphine for post-cesarean pain. Anesth. Analg. 2007, 105, 176–813.10.1213/01.ane.0000265533.13477.26Search in Google Scholar

[36] Keck S, Glennon C, Ginsberg B. DepoDur® extended-release epidural morphine: reshaping postoperative care: what perioperative nurses need to know. Orthop. Nurs. 2007, 26, 86–93.10.1097/01.NOR.0000265863.78103.78Search in Google Scholar

[37] Gelisken F, Inhoffen W, Partsch M, Schneider U, Kreissig I. Retinal pigment epithelial tear after photodynamic therapy for choroidal neovascularization. Am. J. Ophthalmol. 2001, 131, 518–520.10.1016/S0002-9394(00)00813-8Search in Google Scholar

[38] Ebrahim S, Peyman GA, Lee PJ. Applications of liposomes in ophthalmology. Surv. Ophthalmol. 2005, 50, 167–182.10.1016/j.survophthal.2004.12.006Search in Google Scholar

[39] Bovier PA. Epaxal®: a virosomal vaccine to prevent hepatitis A infection. Exp. Rev. Vaccines 2008, 7, 1141–1150.10.1586/14760584.7.8.1141Search in Google Scholar

[40] Riedemann S, Reinhardt G, Ibarra H, Frösner G. Immunogenicity and safety of a virosomal hepatitis A vaccine (Epaxal®) in healthy toddlers and children in Chile. Acta Paediatr. 2004, 93, 412–414.10.1111/j.1651-2227.2004.tb02971.xSearch in Google Scholar

[41] Feng Y, Fuentes D. Model-based planning and real-time predictive control for laser-induced thermal therapy. Int. J. Hyperthermia 2011, 27, 751–761.10.3109/02656736.2011.611962Search in Google Scholar

[42] Akimoto J, Nakayama M, Okano T. Temperature-responsive polymeric micelles for optimizing drug targeting to solid tumors. J. Control. Release 2014, 193, 2–8.10.1016/j.jconrel.2014.06.062Search in Google Scholar

[43] Wust P, Hildebrandt B, Sreenivasa G, Rau B, Gellermann J, Riess H, Felix R, Schlag PM. Hyperthermia in combined treatment of cancer. Lancet Oncol. 2002, 3, 487–497.10.1016/S1470-2045(02)00818-5Search in Google Scholar

[44] Dicheva BM, Koning GA. Targeted thermosensitive liposomes: an attractive novel approach for increased drug delivery to solid tumors. Exp. Opin. Drug Deliv. 2014, 11, 83–100.10.1517/17425247.2014.866650Search in Google Scholar PubMed

[45] Kono K. Thermosensitive polymer-modified liposomes. Adv. Drug Deliv. Rev. 2001, 53, 307–319.10.1016/S0169-409X(01)00204-6Search in Google Scholar

[46] Kawakami K, Ebara M, Izawa H, Sanchez-Ballester NM, Hill JP, Ariga K. Supramolecular approaches for drug development. Curr. Med. Chem. 2012, 19, 2388–2398.10.2174/092986712800269254Search in Google Scholar PubMed

[47] Mallick S, Choi JS. Liposomes: versatile and biocompatible nanovesicles for efficient biomolecules delivery. J. Nanosci. Nanotechnol. 2014, 14, 755–765.10.1166/jnn.2014.9080Search in Google Scholar PubMed

[48] Ta T, Porter TM. Thermosensitive liposomes for localized delivery and triggered release of chemotherapy. J. Control. Release 2013, 169, 112–125.10.1016/j.jconrel.2013.03.036Search in Google Scholar PubMed PubMed Central

[49] Yatvin MB, Weinstein JN, Dennis WH, Blumenthal R. Design of liposomes for enhanced local release of drugs by hyperthermia. Science 1978, 202, 1290–1293.10.1126/science.364652Search in Google Scholar PubMed

[50] Weinstein J, Magin R, Yatvin M, Zaharko D. Liposomes and local hyperthermia: selective delivery of methotrexate to heated tumors. Science 1979, 204, 188–191.10.1126/science.432641Search in Google Scholar PubMed

[51] Unezaki S, Maruyama K, Takahashi N, Koyama M, Yuda T, Suginaka A, Iwatsuru M. Enhanced delivery and antitumor activity of doxorubicin using long-circulating thermosensitive liposomes containing amphipathic polyethylene glycol in combination with local hyperthermia. Pharm. Res. 1994, 11, 1180–5.10.1023/A:1018949218380Search in Google Scholar

[52] Levacheva I, Samsonova O, Tazina E, Beck-Broichsitter M, Levachev S, Strehlow B, Baryshnikova M, Oborotova N, Baryshnikov A, Bakowsky U. Optimized thermosensitive liposomes for selective doxorubicin delivery: formulation development, quality analysis and bioactivity proof. Colloids Surf. B Biointerfaces 2014, 121, 248–256.10.1016/j.colsurfb.2014.02.028Search in Google Scholar PubMed

[53] Wardlow R, Bing C, VanOsdol J, Maples D, Ladouceur-Wodzak M, Harbeson M, Nofiele J, Staruch R, Ramachandran A, Malayer J, Chopra R. Targeted antibiotic delivery using low temperature-sensitive liposomes and magnetic resonance-guided high-intensity focused ultrasound hyperthermia. Int. J. Hyperthermia 2016, 32, 254–264.10.3109/02656736.2015.1134818Search in Google Scholar PubMed PubMed Central

[54] Anyarambhatla GR, Needham D. Enhancement of the phase transition permeability of DPPC liposomes by incorporation of MPPC: a new temperature-sensitive liposome for use with mild hyperthermia. J. Liposome Res. 1999, 9, 491–506.10.3109/08982109909035549Search in Google Scholar

[55] Musyanovych A, Landfester K. Polymer micro- and nanocapsules as biological carriers with multifunctional properties. Macromol. Biosci. 2014, 14, 458–477.10.1002/mabi.201300551Search in Google Scholar PubMed

[56] Fathi M, Entezami AA, Pashaei-Asl R. Swelling/deswelling, thermal, and rheological behavior of PVA-g-NIPAAm nanohydrogels prepared by a facile free-radical polymerization method. J. Polym. Res. 2013, 20, 1–11.10.1007/s10965-013-0125-5Search in Google Scholar

[57] Karimi M, Sahandi Zangabad P, Ghasemi A, Amiri M, Bahrami M, Malekzad H, Ghahramanzadeh Asl H, Mahdieh Z, Bozorgomid M, Ghasemi A, Rahmani Taji Boyuk MR. Temperature-responsive smart nanocarriers for delivery of therapeutic agents: applications and recent advances. ACS Appl. Mater. Interfaces 2016, 8, 21107–21133.10.1021/acsami.6b00371Search in Google Scholar PubMed PubMed Central

[58] Pippa N, Meristoudi A, Pispas S, Demetzos C. Temperature-dependent drug release from DPPC: C 12 H 25-PNIPAM-COOH liposomes: control of the drug loading/release by modulation of the nanocarriers’ components. Int. J. Pharm. 2015, 485, 374–382.10.1016/j.ijpharm.2015.03.014Search in Google Scholar PubMed

[59] Turner DC, Moshkelani D, Shemesh CS, Luc D, Zhang H. Near-infrared image-guided delivery and controlled release using optimized thermosensitive liposomes. Pharm. Res. 2012, 29, 2092–2103.10.1007/s11095-012-0738-0Search in Google Scholar PubMed

[60] Guo H, Kim J-C. Upper critical solution temperature behavior of cinnamic acid and polyethyleneimine mixture and its effect on temperature-dependent release of liposome. Int. J. Pharm. 2015, 494, 172–179.10.1016/j.ijpharm.2015.08.034Search in Google Scholar PubMed

[61] Lee S-M, Nguyen ST. Smart nanoscale drug delivery platforms from stimuli-responsive polymers and liposomes. Macromolecules 2013, 46, 9169–9180.10.1021/ma401529wSearch in Google Scholar PubMed PubMed Central

[62] Cao Z, Ma Y, Yue X, Li S, Dai Z, Kikuchi J. Stabilized liposomal nanohybrid cerasomes for drug delivery applications. Chem. Commun. 2010, 46, 5265–5267.10.1039/b926367eSearch in Google Scholar PubMed

[63] Ma Y, Dai Z, Gao Y, Cao Z, Zha Z, Yue X, Kikuchi JI. Liposomal architecture boosts biocompatibility of nanohybrid cerasomes. Nanotoxicology 2011, 5, 622–635.10.3109/17435390.2010.546950Search in Google Scholar PubMed

[64] Liang X, Li X, Jing L, Xue P, Jiang L, Ren Q, Dai Z. Design and synthesis of lipidic organoalkoxysilanes for the self-assembly of liposomal nanohybrid cerasomes with controlled drug release properties. Chemistry 2013, 19, 16113–16121.10.1002/chem.201302518Search in Google Scholar PubMed

[65] Yue X, Dai Z. Recent advances in liposomal nanohybrid cerasomes as promising drug nanocarriers. Adv. Colloid Interface Sci. 2014, 207, 32–42.10.1016/j.cis.2013.11.014Search in Google Scholar PubMed

[66] Liang X, Gao J, Jiang L, Luo J, Jing L, Li X, Jin Y, Dai Z. Nanohybrid liposomal cerasomes with good physiological stability and rapid temperature responsiveness for high intensity focused ultrasound triggered local chemotherapy of cancer. ACS Nano 2015, 9, 1280–1293.10.1021/nn507482wSearch in Google Scholar PubMed

[67] Kono K, Takashima M, Yuba E, Harada A, Hiramatsu Y, Kitagawa H, Otani T, Maruyama K, Aoshima S. Multifunctional liposomes having target specificity, temperature-triggered release, and near-infrared fluorescence imaging for tumor-specific chemotherapy. J. Control. Release 2015, 216, 69–77.10.1016/j.jconrel.2015.08.005Search in Google Scholar PubMed

[68] Shin DH, Koo M-J, Kim JS, Kim J-S. Herceptin-conjugated temperature-sensitive immunoliposomes encapsulating gemcitabine for breast cancer. Arch. Pharm. Res. 2016, 39, 350–358.10.1007/s12272-016-0707-ySearch in Google Scholar PubMed

[69] Patra S, Roy E, Madhuri R, Sharma PK. Creation of ultrasound and temperature-triggered bubble liposomes from economical precursors to enhance the therapeutic efficacy of curcumin in cancer cells. RSC Adv. 2016, 6, 85473–85485.10.1039/C6RA14584ASearch in Google Scholar

[70] Mathiyazhakan M, Yang Y, Liu Y, Zhu C, Liu Q, Ohl C-D, Tam KC, Gao Y, Xu C. Non-invasive controlled release from gold nanoparticle integrated photo-responsive liposomes through pulse laser induced microbubble cavitation. Colloids Surf. B Biointerfaces 2015, 126, 569–574.10.1016/j.colsurfb.2014.11.019Search in Google Scholar PubMed

[71] Weissleder R. A clearer vision for in vivo imaging. Nat. Biotechnol. 2001, 19, 316–317.10.1038/86684Search in Google Scholar PubMed

[72] McCoy CP, Edwards CR, Jones DS, Gorman SP. Light-triggered molecule-scale drug dosing devices. J. Am. Chem. Soc. 2007, 129, 9572–9573.10.1021/ja073053qSearch in Google Scholar PubMed

[73] Alvarez-Lorenzo C, Bromberg L, Concheiro A. Light-sensitive intelligent drug delivery systems. Photochem. Photobiol. 2009, 85, 848–860.10.1111/j.1751-1097.2008.00530.xSearch in Google Scholar PubMed

[74] Jiang J, Qi B, Lepage M, Zhao Y. Polymer micelles stabilization on demand through reversible photo-cross-linking. Macromolecules 2007, 40, 790–792.10.1021/ma062493jSearch in Google Scholar

[75] He J, Tong X, Zhao Y. Photoresponsive nanogels based on photocontrollable cross-links. Macromolecules 2009, 42, 4845–4852.10.1021/ma900665vSearch in Google Scholar

[76] Lin B, Zhou S. Light-responsive nanoparticles with wettability changing from hydrophobicity to hydrophilicity and their application towards highly hydrophilic fluorocarbon coatings. Appl. Surf. Sci. 2015, 359, 380–387.10.1016/j.apsusc.2015.10.111Search in Google Scholar

[77] Bao C, Zhu L, Lin Q, Tian H. Building biomedical materials using photochemical bond cleavage. Adv. Mater. 2015, 27, 1647–1662.10.1002/adma.201403783Search in Google Scholar

[78] Fomina N, Sankaranarayanan J, Almutairi A. Photochemical mechanisms of light-triggered release from nanocarriers. Adv. Drug Deliv. Rev. 2012, 64, 1005–1020.10.1016/j.addr.2012.02.006Search in Google Scholar

[79] Rwei AY, Wang W, Kohane DS. Photoresponsive nanoparticles for drug delivery. Nano Today 2015, 10, 451–467.10.1016/j.nantod.2015.06.004Search in Google Scholar

[80] Baek S, Phan MD, Lee J, Shin K. Packing effects on polymerization of diacetylene lipids in liposomes and monolayers matrices. Polym. J. 2016, 48, 457–463.10.1038/pj.2015.136Search in Google Scholar

[81] Wöll D, Smirnova J, Galetskaya M, Prykota T, Bühler J, Stengele K-P, Pfleiderer W, Steiner UE. Intramolecular sensitization of photocleavage of the photolabile 2-(2-nitrophenyl)propoxycarbonyl (NPPOC) protecting group: photoproducts and photokinetics of the release of nucleosides. Chemistry 2008, 14, 6490–6497.10.1002/chem.200800613Search in Google Scholar

[82] Yavlovich A, Singh A, Blumenthal R, Puri A. A novel class of photo-triggerable liposomes containing DPPC: DC 8, 9 PC as vehicles for delivery of doxorubcin to cells. Biochim. Biophys. Acta (BBA)-Biomembranes 2011, 1808, 117–126.10.1016/j.bbamem.2010.07.030Search in Google Scholar

[83] Shum P, Kim J-M, Thompson DH. Phototriggering of liposomal drug delivery systems. Adv. Drug Deliv. Rev. 2001, 53, 273–284.10.1016/S0169-409X(01)00232-0Search in Google Scholar

[84] Wang Y, Han P, Xu H, Wang Z, Zhang X, Kabanov AV. Photocontrolled self-assembly and disassembly of block ionomer complex vesicles: a facile approach toward supramolecular polymer nanocontainers. Langmuir 2009, 26, 709–715.10.1021/la9023844Search in Google Scholar PubMed

[85] Hester TJ, Dennison SR, Baker MJ, Snape TJ. Functionalising the azobenzene motif delivers a light-responsive membrane-interactive compound with the potential for photodynamic therapy applications. Org. Biomol. Chem. 2015, 13, 8067–8070.10.1039/C5OB00465ASearch in Google Scholar

[86] Yao C, Wang P, Li X, Hu X, Hou J, Wang L, Zhang F. Near-infrared-triggered azobenzene-liposome/upconversion nanoparticle hybrid vesicles for remotely controlled drug delivery to overcome cancer multidrug resistance. Adv. Mater. 2016, 28, 9341–9348.10.1002/adma.201503799Search in Google Scholar PubMed

[87] Wei Y-B, Tang Q, Gong C-b, Lam MH-W. Review of the recent progress in photoresponsive molecularly imprinted polymers containing azobenzene chromophores. Anal. Chim. Acta 2015, 900, 10–20.10.1016/j.aca.2015.10.022Search in Google Scholar PubMed

[88] Sun T, Zhang YS, Pang B, Hyun DC, Yang M, Xia Y. Engineered nanoparticles for drug delivery in cancer therapy. Angew. Chem. Int. Ed. Engl. 2014, 53, 12320–12364.10.1201/9780429027819-2Search in Google Scholar

[89] Ishii Ki, Hamada T, Hatakeyama M, Sugimoto R, Nagasaki T, Takagi M. Reversible control of exo- and endo-budding transitions in a photosensitive lipid membrane. ChemBioChem 2009, 10, 251–256.10.1002/cbic.200800482Search in Google Scholar PubMed

[90] Xia Y, Yang P, Sun Y, Wu Y, Mayers B, Gates B, Yin Y, Kim F, Yan H. One-dimensional nanostructures: synthesis, characterization, and applications. Adv. Mater. 2003, 15, 353–389.10.1002/adma.200390087Search in Google Scholar

[91] Eustis S, El-Sayed MA. Why gold nanoparticles are more precious than pretty gold: noble metal surface plasmon resonance and its enhancement of the radiative and nonradiative properties of nanocrystals of different shapes. Chem. Soc. Rev. 2006, 35, 209–217.10.1039/B514191ESearch in Google Scholar PubMed

[92] Link S, El-Sayed MA. Shape and size dependence of radiative, non-radiative and photothermal properties of gold nanocrystals. Int. Rev. Phys. Chem. 2000, 19, 409–453.10.1080/01442350050034180Search in Google Scholar

[93] Guo H, Kim J-C. Photothermally induced release from liposome suspended in mixture solution of gold nanoparticle and thermo-sensitive polymer. Colloids Surf. A Physicochem. Eng. Asp. 2015, 469, 73–82.10.1016/j.colsurfa.2015.01.010Search in Google Scholar

[94] Kalber TL, Ordidge KL, Southern P, Loebinger MR, Kyrtatos PG, Pankhurst QA, Lythgoe MF, Janes SM. Hyperthermia treatment of tumors by mesenchymal stem cell-delivered superparamagnetic iron oxide nanoparticles. Int. J. Nanomed. 2016, 11, 1973.10.2147/IJN.S94255Search in Google Scholar PubMed PubMed Central

[95] Ma N, Jiang YW, Zhang X, Wu H, Myers JN, Liu P, Jin H, Gu N, He N, Wu FG, Chen Z. Enhanced radiosensitization of gold nanospikes via hyperthermia in combined cancer radiation and photothermal therapy. ACS Appl. Mater. Interfaces 2016, 8, 28480–28494.10.1021/acsami.6b10132Search in Google Scholar

[96] Wu G, Mikhailovsky A, Khant HA, Fu C, Chiu W, Zasadzinski JA. Remotely triggered liposome release by near-infrared light absorption via hollow gold nanoshells. J. Am. Chem. Soc. 2008, 130, 8175–8177.10.1021/ja802656dSearch in Google Scholar

[97] Lajunen T, Viitala L, Kontturi L-S, Laaksonen T, Liang H, Vuorimaa-Laukkanen E, Viitala T, Le Guével X, Yliperttula M, Murtomäki L, Urtti A. Light induced cytosolic drug delivery from liposomes with gold nanoparticles. J. Control. Release 2015, 203, 85–98.10.1016/j.jconrel.2015.02.028Search in Google Scholar

[98] Drummond DC, Zignani M, Leroux J-C. Current status of pH-sensitive liposomes in drug delivery. Prog. Lipid Res. 2000, 39, 409–460.10.1016/S0163-7827(00)00011-4Search in Google Scholar

[99] Karanth H, Murthy R. pH-sensitive liposomes-principle and application in cancer therapy. J. Pharm. Pharmacol. 2007, 59, 469–483.10.1211/jpp.59.4.0001Search in Google Scholar

[100] Ferreira DdS, Lopes SCdA, Franco MS, Oliveira MC. pH-sensitive liposomes for drug delivery in cancer treatment. Ther. Deliv. 2013, 4, 1099–1123.10.4155/tde.13.80Search in Google Scholar

[101] Lai MZ, Duzgunes N, Szoka FC. Effects of replacement of the hydroxyl group of cholesterol and tocopherol on the thermotropic behavior of phospholipid membranes. Biochemistry 1985, 24, 1646–1653.10.1021/bi00328a012Search in Google Scholar

[102] Webb MS, Wheeler JJ, Bally MB, Mayer LD. The cationic lipid stearylamine reduces the permeability of the cationic drugs verapamil and prochlorperazine to lipid bilayers: implications for drug delivery. Biochim. Biophys. Acta (BBA)-Biomembranes 1995, 1238, 147–155.10.1016/0005-2736(95)00121-ISearch in Google Scholar

[103] Nayar R, Schroit AJ. Generation of pH-sensitive liposomes: use of large unilamellar vesicles containing N-succinyldioleoylphosphatidylethanolamine. Biochemistry 1985, 24, 5967–5971.10.1021/bi00342a042Search in Google Scholar PubMed

[104] Paliwal SR, Paliwal R, Vyas SP. A review of mechanistic insight and application of pH-sensitive liposomes in drug delivery. Drug Deliv. 2015, 22, 231–242.10.3109/10717544.2014.882469Search in Google Scholar PubMed

[105] Shi G, Guo W, Stephenson SM, Lee RJ. Efficient intracellular drug and gene delivery using folate receptor-targeted pH-sensitive liposomes composed of cationic/anionic lipid combinations. J. Control. Release 2002, 80, 309–319.10.1016/S0168-3659(02)00017-2Search in Google Scholar

[106] Guo W, Gosselin MA, Lee RJ. Characterization of a novel diolein-based LPDII vector for gene delivery. J. Control. Release 2002, 83, 121–132.10.1016/S0168-3659(02)00167-0Search in Google Scholar

[107] Straubinger RM, Hong K, Friend DS, Papahadjopoulos D. Endocytosis of liposomes and intracellular fate of encapsulated molecules: encounter with a low pH compartment after internalization in coated vesicles. Cell 1983, 32, 1069–1079.10.1016/0092-8674(83)90291-XSearch in Google Scholar

[108] Ohkuma S, Poole B. Fluorescence probe measurement of the intralysosomal pH in living cells and the perturbation of pH by various agents. Proc. Natl. Acad. Sci. USA 1978, 75, 3327–3331.10.1073/pnas.75.7.3327Search in Google Scholar

[109] Woodle MC, Lasic DD. Sterically stabilized liposomes. Biochim. Biophys. Acta (BBA) Rev. Biomembr. 1992, 1113, 171–199.10.1016/0304-4157(92)90038-CSearch in Google Scholar

[110] Lasic DD, Martin FJ. Stealth Liposomes, Vol. 20, CRC Press: Boca Raton, FL, USA, 1995.Search in Google Scholar

[111] Holland JW, Hui C, Cullis PR, Madden TD. Poly (ethylene glycol)-lipid conjugates regulate the calcium-induced fusion of liposomes composed of phosphatidylethanolamine and phosphatidylserine. Biochemistry 1996, 35, 2618–2624.10.1021/bi952000vSearch in Google Scholar

[112] Yu Y, Kang X, Yang X, Yuan L, Feng W, Cui S. Surface charge inversion of self-assembled monolayers by visible light irradiation: cargo loading and release by photoreactions. Chem. Commun. 2013, 49, 3431–3433.10.1039/c3cc39081kSearch in Google Scholar

[113] Paliwal SR, Paliwal R, Agrawal GP, Vyas SP. Hyaluronic acid modified pH-sensitive liposomes for targeted intracellular delivery of doxorubicin. J. Liposome Res. 2015, 26, 1–12.10.3109/08982104.2015.1117489Search in Google Scholar

[114] Seki K, Tirrell DA. pH-dependent complexation of poly (acrylic acid) derivatives with phospholipid vesicle membranes. Macromolecules 1984, 17, 1692–1698.10.1021/ma00139a009Search in Google Scholar

[115] Murthy N, Robichaud JR, Tirrell DA, Stayton PS, Hoffman AS. The design and synthesis of polymers for eukaryotic membrane disruption. J. Control. Release 1999, 61, 137–143.10.1016/S0168-3659(99)00114-5Search in Google Scholar

[116] Yoshizaki Y, Yuba E, Komatsu T, Udaka K, Harada A, Kono K. Improvement of peptide-based tumor immunotherapy using pH-sensitive fusogenic polymer-modified liposomes. Molecules 2016, 21, 1284.10.3390/molecules21101284Search in Google Scholar

[117] Yuba E, Tajima N, Yoshizaki Y, Harada A, Hayashi H, Kono K. Dextran derivative-based pH-sensitive liposomes for cancer immunotherapy. Biomaterials 2014, 35, 3091–3101.10.1016/j.biomaterials.2013.12.024Search in Google Scholar

[118] Zong W, Hu Y, Su Y, Luo N, Zhang X, Li Q, Han X. Polydopamine-coated liposomes as pH-sensitive anticancer drug carriers. J. Microencapsul. 2016, 33, 257–262.10.3109/02652048.2016.1156176Search in Google Scholar

[119] Zhao Y, Ren W, Zhong T, Zhang S, Huang D, Guo Y, Yao X, Wang C, Zhang WQ, Zhang X, Zhang Q. Tumor-specific pH-responsive peptide-modified pH-sensitive liposomes containing doxorubicin for enhancing glioma targeting and anti-tumor activity. J. Control. Release 2016, 222, 56–66.10.1016/j.jconrel.2015.12.006Search in Google Scholar

[120] Bhardwaj A, Grobler A, Rath G, Kumar Goyal A, Kumar Jain A, Mehta A. Pulmonary delivery of anti-tubercular drugs using ligand anchored pH sensitive liposomes for the treatment of pulmonary tuberculosis. Curr. Drug Deliv. 2016, 13, 909–922.10.2174/1567201813666151231093605Search in Google Scholar

[121] Ishida T, Kirchmeier M, Moase E, Zalipsky S, Allen T. Targeted delivery and triggered release of liposomal doxorubicin enhances cytotoxicity against human B lymphoma cells. Biochim. Biophys. Acta (BBA)-Biomembr. 2001, 1515, 144–158.10.1016/S0005-2736(01)00409-6Search in Google Scholar

[122] Mizoue T, Horibe T, Maruyama K, Takizawa T, Iwatsuru M, Kono K, Yanagie H, Moriyasu F. Targetability and intracellular delivery of anti-BCG antibody-modified, pH-sensitive fusogenic immunoliposomes to tumor cells. Int. J. Pharm. 2002, 237, 129–137.10.1016/S0378-5173(02)00044-3Search in Google Scholar

[123] Ng K-Y, Zhao L, Liu Y, Mahapatro M. The effects of polyethyleneglycol (PEG)-derived lipid on the activity of target-sensitive immunoliposome. Int. J. Pharm. 2000, 193, 157–166.10.1016/S0378-5173(99)00330-0Search in Google Scholar

[124] Spragg DD, Alford DR, Greferath R, Larsen CE, Lee K-D, Gurtner GC, Cybulsky MI, Tosi PF, Nicolau C, Gimbrone MA. Immunotargeting of liposomes to activated vascular endothelial cells: a strategy for site-selective delivery in the cardiovascular system. Proc. Natl. Acad. Sci. USA 1997, 94, 8795–8800.10.1073/pnas.94.16.8795Search in Google Scholar PubMed PubMed Central

[125] Wang C-Y, Huang L. pH-sensitive immunoliposomes mediate target-cell-specific delivery and controlled expression of a foreign gene in mouse. Proc. Natl. Acad. Sci. USA 1987, 84, 7851–7855.10.1073/pnas.84.22.7851Search in Google Scholar PubMed PubMed Central

[126] Heinemann L, Pfutzner A, Heise T. Alternative routes of administration as an approach to improve insulin therapy: update on dermal, oral, nasal and pulmonary insulin delivery. Curr. Pharm. Des. 2001, 7, 1327–1351.10.2174/1381612013397384Search in Google Scholar PubMed

[127] Wehunt MP, Winschel CA, Khan AK, Guo TL, Abdrakhmanova GR, Sidorov V. Controlled drug-release system based on pH-sensitive chloride-triggerable liposomes. J. Liposome Res. 2013, 23, 37–46.10.3109/08982104.2012.727423Search in Google Scholar PubMed

[128] Liu J, Ma H, Wei T, Liang X-J. CO2 gas induced drug release from pH-sensitive liposome to circumvent doxorubicin resistant cells. Chem. Commun. 2012, 48, 4869–4871.10.1039/c2cc31697hSearch in Google Scholar PubMed

[129] Liu J, Huang Y, Kumar A, Tan A, Jin S, Mozhi A, Liang XJ. pH-sensitive nano-systems for drug delivery in cancer therapy. Biotechnol. Adv. 2014, 32, 693–710.10.1016/j.biotechadv.2013.11.009Search in Google Scholar PubMed

[130] Watarai S, Iwase T, Tajima T, Yuba E, Kono K. Efficiency of pH-sensitive fusogenic polymer-modified liposomes as a vaccine carrier. Sci. World J. 2013, 2013, 903234.10.1155/2013/903234Search in Google Scholar PubMed PubMed Central

[131] Hatakeyama H, Murata M, Sato Y, Takahashi M, Minakawa N, Matsuda A, Harashima H. The systemic administration of an anti-miRNA oligonucleotide encapsulated pH-sensitive liposome results in reduced level of hepatic microRNA-122 in mice. J. Control. Release 2014, 173, 43–50.10.1016/j.jconrel.2013.10.023Search in Google Scholar

[132] Chen Y, Sun J, Lu Y, Tao C, Huang J, Zhang H, Yu Y, Zou H, Gao J, Zhong Y. Complexes containing cationic and anionic pH-sensitive liposomes: comparative study of factors influencing plasmid DNA gene delivery to tumors. Int. J. Nanomed. 2013, 8, 1573.10.2147/IJN.S42800Search in Google Scholar PubMed PubMed Central

[133] Hatakeyama H, Akita H, Harashima H. The polyethyleneglycol dilemma: advantage and disadvantage of PEGylation of liposomes for systemic genes and nucleic acids delivery to tumors. Biol. Pharm. Bull. 2013, 36, 892–899.10.1248/bpb.b13-00059Search in Google Scholar PubMed

[134] Vila-Caballer M, Codolo G, Munari F, Malfanti A, Fassan M, Rugge M, Balasso A, de Bernard M, Salmaso S. A pH-sensitive stearoyl-PEG-poly (methacryloyl sulfadimethoxine)-decorated liposome system for protein delivery: an application for bladder cancer treatment. J. Control. Release 2016, 238, 31–42.10.1016/j.jconrel.2016.07.024Search in Google Scholar PubMed

[135] Liu X, Huang G. Formation strategies, mechanism of intracellular delivery and potential clinical applications of pH-sensitive liposomes. Asian J. Pharm. Sci. 2013, 8, 319–328.10.1016/j.ajps.2013.11.002Search in Google Scholar

[136] Banerjee S, Sen K, Pal TK, Guha SK. Poly (styrene-co-maleic acid)-based pH-sensitive liposomes mediate cytosolic delivery of drugs for enhanced cancer chemotherapy. Int. J. Pharm. 2012, 436, 786–797.10.1016/j.ijpharm.2012.07.059Search in Google Scholar PubMed

[137] Chang M, Lu S, Zhang F, Zuo T, Guan Y, Wei T, Shao W, Lin G. RGD-modified pH-sensitive liposomes for docetaxel tumor targeting. Colloids Surf. B Biointerfaces 2015, 129, 175–182.10.1016/j.colsurfb.2015.03.046Search in Google Scholar PubMed

[138] de Barros ALB, Mota LdG, Coelho MMA, Corrêa NCR, de Góes AM, Oliveira MC, Cardoso VN. Bombesin encapsulated in long-circulating pH-sensitive liposomes as a radiotracer for breast tumor identification. J. Biomed. Nanotechnol. 2015, 11, 342–350.10.1166/jbn.2015.1910Search in Google Scholar PubMed

[139] Chen Q, Ding H, Zhou J, Zhao X, Zhang J, Yang C, Li K, Qiao M, Hu H, Ding P, Zhao X. Novel glycyrrhetinic acid conjugated pH-sensitive liposomes for the delivery of doxorubicin and its antitumor activities. RSC Adv. 2016, 6, 17782–17791.10.1039/C6RA01580HSearch in Google Scholar

[140] Mota LDG, Barros ALBd, Fuscaldi LL, Oliveira MCd, Cardoso VN. Long-circulating and pH-sensitive liposome preparation trapping a radiotracer for inflammation site detection. J. Nanosci. Nanotechnol. 2015, 15, 4149–4158.10.1166/jnn.2015.9619Search in Google Scholar PubMed

[141] Ishida T, Okada Y, Kobayashi T, Kiwada H. Development of pH-sensitive liposomes that efficiently retain encapsulated doxorubicin (DXR) in blood. Int. J. Pharm. 2006, 309, 94–100.10.1016/j.ijpharm.2005.11.010Search in Google Scholar PubMed

[142] Kale AA, Torchilin VP. “Smart” drug carriers: PEGylated TATp-modified pH-sensitive liposomes. J. Liposome Res. 2007, 17, 197–203.10.1080/08982100701525035Search in Google Scholar PubMed PubMed Central

[143] Júnior ÁD, Mota LG, Nunan EA, Wainstein AJ, Wainstein APD, Leal AS, Cardoso VN, De Oliveira MC. Tissue distribution evaluation of stealth pH-sensitive liposomal cisplatin versus free cisplatin in Ehrlich tumor-bearing mice. Life Sci. 2007, 80, 659–664.10.1016/j.lfs.2006.10.011Search in Google Scholar PubMed

[144] Leite EA, Lana ÂM, Carvalho Junior ÁD, Coelho LG, De Oliveira MC. Acute toxicity study of cisplatin loaded long-circulating and pH-sensitive liposomes administered in mice. J. Biomed. Nanotechnol. 2012, 8, 229–239.10.1166/jbn.2012.1388Search in Google Scholar PubMed

[145] Leite EA, Souza CM, Carvalho-Júnior AD, Coelho L, Lana A, Cassali GD, Oliveira MC. Encapsulation of cisplatin in long-circulating and pH-sensitive liposomes improves its antitumor effect and reduces acute toxicity. Int. J. Nanomed. 2012, 7, 5259–5269.10.2147/IJN.S34652Search in Google Scholar PubMed PubMed Central

[146] Koren E, Apte A, Jani A, Torchilin VP. Multifunctional PEGylated 2C5-immunoliposomes containing pH-sensitive bonds and TAT peptide for enhanced tumor cell internalization and cytotoxicity. J. Control. Release 2012, 160, 264–273.10.1016/j.jconrel.2011.12.002Search in Google Scholar

[147] Xia G, An Z, Wang Y, Zhao C, Li M, Li Z, Ma J. Synthesis of a novel polymeric material folate-poly (2-ethyl-2-oxazoline)-distearoyl phosphatidyl ethanolamine tri-block polymer for dual receptor and pH-sensitive targeting liposome. Chem. Pharm. Bull. 2013, 61, 390–398.10.1248/cpb.c12-00951Search in Google Scholar

[148] Hong MS, Lim SJ, Oh YK, Kim CK. pH-sensitive, serum-stable and long-circulating liposomes as a new drug delivery system. J. Pharm. Pharmacol. 2002, 54, 51–8.10.1211/0022357021771913Search in Google Scholar

[149] Sudimack JJ, Guo W, Tjarks W, Lee RJ. A novel pH-sensitive liposome formulation containing oleyl alcohol. Biochim. Biophys. Acta (BBA)-Biomembr. 2002, 1564, 31–37.10.1016/S0005-2736(02)00399-1Search in Google Scholar

[150] Duan Y, Wei L, Petryk J, Ruddy TD. Formulation, characterization and tissue distribution of a novel pH-sensitive long-circulating liposome-based theranostic suitable for molecular imaging and drug delivery. Int. J. Nanomed. 2016, 11, 5697.10.2147/IJN.S111274Search in Google Scholar PubMed PubMed Central

[151] Xu H, Hu M, Yu X, Li Y, Fu Y, Zhou X, Zhang D, Li J. Design and evaluation of pH-sensitive liposomes constructed by poly (2-ethyl-2-oxazoline)-cholesterol hemisuccinate for doxorubicin delivery. Eur. J. Pharm. Biopharm. 2015, 91, 66–74.10.1016/j.ejpb.2015.01.030Search in Google Scholar PubMed

[152] Zhang H, Li R-Y, Lu X, Mou Z-z, Lin G-m. Docetaxel-loaded liposomes: preparation, pH sensitivity, pharmacokinetics, and tissue distribution. J. Zhejiang Univ. Sci. B 2012, 13, 981–989.10.1631/jzus.B1200098Search in Google Scholar PubMed PubMed Central

[153] Kim I-Y, Kang Y-S, Lee DS, Park H-J, Choi E-K, Oh Y-K, Son HJ, Kim JS. Antitumor activity of EGFR targeted pH-sensitive immunoliposomes encapsulating gemcitabine in A549 xenograft nude mice. J. Control. Release 2009, 140, 55–60.10.1016/j.jconrel.2009.07.005Search in Google Scholar PubMed

[154] Mo R, Sun Q, Li N, Zhang C. Intracellular delivery and antitumor effects of pH-sensitive liposomes based on zwitterionic oligopeptide lipids. Biomaterials 2013, 34, 2773–2786.10.1016/j.biomaterials.2013.01.030Search in Google Scholar PubMed

[155] Chen D, Jiang X, Huang Y, Zhang C, Ping Q. pH-sensitive mPEG-Hz-cholesterol conjugates as a liposome delivery system. J. Bioact. Compat. Polym. 2010, 25, 441–454.Search in Google Scholar

[156] Leighton TG. What is ultrasound? Progr. Biophys. Mol. Biol. 2007, 93, 3–83.10.1016/j.pbiomolbio.2006.07.026Search in Google Scholar PubMed

[157] Schroeder A, Kost J, Barenholz Y. Ultrasound, liposomes, and drug delivery: principles for using ultrasound to control the release of drugs from liposomes. Chem. Phys. Lipids 2009, 162, 1–16.10.1016/j.chemphyslip.2009.08.003Search in Google Scholar

[158] Mason TJ. Chemistry with ultrasound. Published for the Society of Chemical Industry by. Sole distributor in the USA and Canada, Elsevier Science Pub. Co.; 1990.Search in Google Scholar

[159] Rychak JJ, Klibanov AL. Nucleic acid delivery with microbubbles and ultrasound. Adv. Drug Deliv. Rev. 2014, 72, 82–93.10.1016/j.addr.2014.01.009Search in Google Scholar

[160] Sirsi SR, Borden MA. State-of-the-art materials for ultrasound-triggered drug delivery. Adv. Drug Deliv. Rev. 2014, 72, 3–14.10.1016/j.addr.2013.12.010Search in Google Scholar

[161] Husseini GA, Pitt WG. Micelles and nanoparticles for ultrasonic drug and gene delivery. Adv. Drug Deliv. Rev. 2008, 60, 1137–1152.10.1016/j.addr.2008.03.008Search in Google Scholar

[162] Lentacker I, De Cock I, Deckers R, De Smedt SC, Moonen CTW. Understanding ultrasound induced sonoporation: definitions and underlying mechanisms. Adv. Drug Deliv. Rev. 2014, 72, 49–64.10.1016/j.addr.2013.11.008Search in Google Scholar

[163] Husseini GA, Myrup GD, Pitt WG, Christensen DA, Rapoport NY. Factors affecting acoustically triggered release of drugs from polymeric micelles. J. Control. Release 2000, 69, 43–52.10.1016/S0168-3659(00)00278-9Search in Google Scholar

[164] Lin H-Y, Thomas JL. Factors affecting responsivity of unilamellar liposomes to 20 kHz ultrasound. Langmuir 2004, 20, 6100–6106.10.1021/la049866zSearch in Google Scholar PubMed

[165] Schroeder A, Avnir Y, Weisman S, Najajreh Y, Gabizon A, Talmon Y, Kost J, Barenholz Y. Controlling liposomal drug release with low frequency ultrasound: mechanism and feasibility. Langmuir 2007, 23, 4019–4025.10.1021/la0631668Search in Google Scholar PubMed

[166] Chang S, Si T, Zhang S, Merrick MA, Cohn DE, Xu RX. Ultrasound mediated destruction of multifunctional microbubbles for image guided delivery of oxygen and drugs. Ultrason. Sonochem. 2016, 28, 31–38.10.1016/j.ultsonch.2015.06.024Search in Google Scholar PubMed

[167] Unger E, Porter T, Lindner J, Grayburn P. Cardiovascular drug delivery with ultrasound and microbubbles. Adv. Drug Deliv. Rev. 2014, 72, 110–126.10.1016/j.addr.2014.01.012Search in Google Scholar PubMed

[168] Yu FTH, Chen X, Wang J, Qin B, Villanueva FS. Low intensity ultrasound mediated liposomal doxorubicin delivery using polymer microbubbles. Mol. Pharm. 2016, 13, 55–64.10.1021/acs.molpharmaceut.5b00421Search in Google Scholar

[169] Evjen TJ, Nilssen EA, Barnert S, Schubert R, Brandl M, Fossheim SL. Ultrasound-mediated destabilization and drug release from liposomes comprising dioleoylphosphatidylethanolamine. Eur. J. Pharm. Sci. 2011, 42, 380–386.10.1016/j.ejps.2011.01.002Search in Google Scholar

[170] Evjen TJ, Nilssen EA, Rögnvaldsson S, Brandl M, Fossheim SL. Distearoylphosphatidylethanolamine-based liposomes for ultrasound-mediated drug delivery. Eur. J. Pharm. Biopharm. 2010, 75, 327–333.10.1016/j.ejpb.2010.04.012Search in Google Scholar

[171] Walde P, Sunamoto J, O’Connor CJ. The mechanism of liposomal damage by taurocholate. BBA Biomembr. 1987, 905, 30–38.10.1016/0005-2736(87)90005-8Search in Google Scholar

[172] Mujoo H, Reynolds JNJ, Tucker IG. The influence of bile salts on the response of liposomes to ultrasound. J. Liposome Res. 2016, 26, 87–95.10.3109/08982104.2015.1019515Search in Google Scholar PubMed

[173] Aryal M, Arvanitis CD, Alexander PM, McDannold N. Ultrasound-mediated blood–brain barrier disruption for targeted drug delivery in the central nervous system. Adv. Drug Deliv. Rev. 2014, 72, 94–109.10.1016/j.addr.2014.01.008Search in Google Scholar PubMed PubMed Central

[174] Lin Q, Mao KL, Tian FR, Yang JJ, Chen PP, Xu J, Fan ZL, Zhao YP, Li WF, Zheng L, Zhao YZ. Brain tumor-targeted delivery and therapy by focused ultrasound introduced doxorubicin-loaded cationic liposomes. Cancer Chemother. Pharmacol. 2016, 77, 269–280.10.1007/s00280-015-2926-1Search in Google Scholar PubMed

[175] Golovin YI, Gribanovsky SL, Golovin DY, Klyachko NL, Majouga AG, Master AM, Sokolsky M, Kabanov AV. Towards nanomedicines of the future: Remote magneto-mechanical actuation of nanomedicines by alternating magnetic fields. J. Control. Release 2015, 219, 43–60.10.1016/j.jconrel.2015.09.038Search in Google Scholar PubMed PubMed Central

[176] Di Corato R, Béalle G, Kolosnjaj-Tabi J, Espinosa A, Clement O, Silva AK, Menager C, Wilhelm C. Combining magnetic hyperthermia and photodynamic therapy for tumor ablation with photoresponsive magnetic liposomes. ACS Nano 2015, 9, 2904–2916.10.1021/nn506949tSearch in Google Scholar PubMed

[177] Marie H, Lemaire L, Franconi F, Lajnef S, Frapart YM, Nicolas V, Frébourg G, Trichet M, Ménager C, Lesieur S. Superparamagnetic liposomes for MRI monitoring and external magnetic field-induced selective targeting of malignant brain tumors. Adv. Funct. Mater. 2015, 25, 1258–1269.10.1002/adfm.201402289Search in Google Scholar

[178] Kwon YS, Jang SJ, Yoon YI, Cho HS, Lee HJ, Cho YS, Shin HS, Yoon TJ. Magnetic liposomal particles for magnetic imaging, sensing, and the pH-sensitive delivery of therapeutics. Part. Part. Syst. Charact. 2016, 33, 242–247.10.1002/ppsc.201600041Search in Google Scholar

[179] Amstad E, Kohlbrecher J, Müller E, Schweizer T, Textor M, Reimhult E. Triggered release from liposomes through magnetic actuation of iron oxide nanoparticle containing membranes. Nano Lett. 2011, 11, 1664–1670.10.1021/nl2001499Search in Google Scholar PubMed

[180] Bringas E, Köysüren Ö, Quach DV, Mahmoudi M, Aznar E, Roehling JD, Marcos MD, Martínez-Máñez R, Stroeve P. Triggered release in lipid bilayer-capped mesoporous silica nanoparticles containing SPION using an alternating magnetic field. Chem. Commun. 2012, 48, 5647–5649.10.1039/c2cc31563gSearch in Google Scholar PubMed

[181] Katagiri K, Nakamura M, Koumoto K. Magnetoresponsive smart capsules formed with polyelectrolytes, lipid bilayers and magnetic nanoparticles. ACS Appl. Mater. Interfaces 2010, 2, 768–773.10.1021/am900784aSearch in Google Scholar PubMed

[182] Salvatore A, Montis C, Berti D, Baglioni P. Multifunctional magnetoliposomes for sequential controlled release. ACS Nano 2016, 10, 7749–7760.10.1021/acsnano.6b03194Search in Google Scholar PubMed

[183] Edelman RR, Warach S. Magnetic resonance imaging. N. Engl. J. Med. 1993, 328, 708–716.10.1056/NEJM199303113281008Search in Google Scholar PubMed

[184] Garello F, Vibhute S, Gündüz S, Logothetis NK, Terreno E, Angelovski G. Innovative design of Ca-sensitive paramagnetic liposomes results in an unprecedented increase in longitudinal relaxivity. Biomacromolecules 2016, 17, 1303–1311.10.1021/acs.biomac.5b01668Search in Google Scholar PubMed

[185] Zhang K, Liu M, Tong X, Sun N, Zhou L, Cao Y, Wang J, Zhang H, Pei R. Aptamer-modified temperature-sensitive liposomal contrast agent for magnetic resonance imaging. Biomacromolecules 2015, 16, 2618–2623.10.1021/acs.biomac.5b00250Search in Google Scholar PubMed

[186] Zhang L, Zhang Z, Mason RP, Sarkaria JN, Zhao D. Convertible MRI contrast: Sensing the delivery and release of anti-glioma nano-drugs. Sci. Rep. 2015, 5, 9874.10.1038/srep09874Search in Google Scholar PubMed PubMed Central

[187] Du B, Han S, Li H, Zhao F, Su X, Cao X, Zhang Z. Multi-functional liposomes showing radiofrequency-triggered release and magnetic resonance imaging for tumor multi-mechanism therapy. Nanoscale 2015, 7, 5411–5426.10.1039/C4NR04257CSearch in Google Scholar

[188] Chechetka SA, Yuba E, Kono K, Yudasaka M, Bianco A, Miyako E. Magnetically and near-infrared light-powered supramolecular nanotransporters for the remote control of enzymatic reactions. Angew. Chem. 2016, 128, 6586–6591.10.1002/ange.201602453Search in Google Scholar

[189] Zhu L, Kate P, Torchilin VP. Matrix metalloprotease 2-responsive multifunctional liposomal nanocarrier for enhanced tumor targeting. ACS Nano 2012, 6, 3491–3498.10.1021/nn300524fSearch in Google Scholar

[190] Dai Z, Yao Q, Zhu L. MMP2-sensitive PEG–lipid copolymers: a new type of tumor-targeted P-glycoprotein inhibitor. ACS Appl. Mater. Interfaces 2016, 8, 12661–12673.10.1021/acsami.6b03064Search in Google Scholar

[191] Noyhouzer T, L’Homme C, Beaulieu I, Mazurkiewicz S, Kuss S, Kraatz H-B, Canesi S, Mauzeroll J. Ferrocene-modified phospholipid: an innovative precursor for redox-triggered drug delivery vesicles selective to cancer cells. Langmuir 2016, 32, 4169–4178.10.1021/acs.langmuir.6b00511Search in Google Scholar

[192] Lopes JR, Santos G, Barata P, Oliveira R, Lopes CM. Physical and chemical stimuli-responsive drug delivery systems: targeted delivery and main routes of administration. Curr. Pharm. Des. 2013, 19, 7169–7184.10.2174/13816128113199990698Search in Google Scholar

[193] Ganta S, Devalapally H, Shahiwala A, Amiji M. A review of stimuli-responsive nanocarriers for drug and gene delivery. J. Control. Release 2008, 126, 187–204.10.1016/j.jconrel.2007.12.017Search in Google Scholar

[194] Rapoport N. Physical stimuli-responsive polymeric micelles for anti-cancer drug delivery. Progr. Polym. Sci. 2007, 32, 962–990.10.1016/j.progpolymsci.2007.05.009Search in Google Scholar

[195] Zasadzinski JA, Wong B, Forbes N, Braun G, Wu G. Novel methods of enhanced retention in and rapid, targeted release from liposomes. Curr. Opin. Colloid Interface Sci. 2011, 16, 203–214.10.1016/j.cocis.2010.12.004Search in Google Scholar

[196] Chandaroy P, Sen A, Hui SW. Temperature-controlled content release from liposomes encapsulating Pluronic F127. J. Control. Release 2001, 76, 27–37.10.1016/S0168-3659(01)00429-1Search in Google Scholar

[197] Kono K, Kaiden T, Yuba E, Sakanishi Y, Harada A. Synthesis of oligo (ethylene glycol)-modified hyperbranched poly (glycidol) s for dual sensitization of liposomes to pH and temperature. J. Taiwan Inst. Chem. Eng. 2014, 45, 3054–3061.10.1016/j.jtice.2014.06.011Search in Google Scholar

[198] Chen D, Sun K, Mu H, Tang M, Liang R, Wang A, Zhou S, Sun H, Zhao F, Yao J, Liu W. pH and temperature dual-sensitive liposome gel based on novel cleavable mPEG-Hz-CHEMS polymeric vaginal delivery system. Int. J. Nanomed. 2012, 7, 2621–2630.10.2147/IJN.S31757Search in Google Scholar PubMed PubMed Central

[199] Kojima C, Yoshimura K, Harada A, Sakanishi Y, Kono K. Synthesis and characterization of hyperbranched poly (glycidol) modified with pH-and temperature-sensitive groups. Bioconjug. Chem. 2009, 20, 1054–1057.10.1021/bc900016xSearch in Google Scholar PubMed

[200] Kaiden T, Yuba E, Harada A, Sakanishi Y, Kono K. Dual signal-responsive liposomes for temperature-controlled cytoplasmic delivery. Bioconjug. Chem. 2011, 22, 1909–1915.10.1021/bc2000353Search in Google Scholar PubMed

[201] Wu X, Wang Z, Zhu D, Zong S, Yang L, Zhong Y, Cui Y. pH and thermo dual-stimuli-responsive drug carrier based on mesoporous silica nanoparticles encapsulated in a copolymer–lipid bilayer. ACS Appl. Mater. Interfaces 2013, 5, 10895–10903.10.1021/am403092mSearch in Google Scholar PubMed

[202] Kong G, Dewhirst M. Review hyperthermia and liposomes. Int. J. Hyperthermia 1999, 15, 345–370.10.1080/026567399285558Search in Google Scholar PubMed

[203] Kong G, Anyarambhatla G, Petros WP, Braun RD, Colvin OM, Needham D, Dewhirst MW. Efficacy of liposomes and hyperthermia in a human tumor xenograft model: importance of triggered drug release. Cancer Res. 2000, 60, 6950–6957.Search in Google Scholar

[204] Kennedy JE. High-intensity focused ultrasound in the treatment of solid tumours. Nat. Rev. Cancer 2005, 5, 321–327.10.1038/nrc1591Search in Google Scholar PubMed

[205] Dromi S, Frenkel V, Luk A, Traughber B, Angstadt M, Bur M, Poff J, Xie J, Libutti SK, Li KC, Wood BJ. Pulsed-high intensity focused ultrasound and low temperature–sensitive liposomes for enhanced targeted drug delivery and antitumor effect. Clin. Cancer Res. 2007, 13, 2722–2727.10.1158/1078-0432.CCR-06-2443Search in Google Scholar PubMed PubMed Central

[206] Chen D, Yu H, Mu H, Wei J, Song Z, Shi H, Liang R, Sun K, Liu W. Novel chitosan derivative for temperature and ultrasound dual-sensitive liposomal microbubble gel. Carbohydr. Polym. 2013, 94, 17–23.10.1016/j.carbpol.2012.12.069Search in Google Scholar PubMed

[207] Oerlemans C, Nijsen F, van Amersfoort M, van Bloois L, Heijman E, Luijten P, Mali W, Storm G. A novel approach to identify non-palpable breast lesions combining fluorescent liposomes and magnetic resonance-guided high intensity focused ultrasound-triggered release. Eur. J. Pharm. Biopharm. 2011, 77, 458–464.10.1016/j.ejpb.2010.12.028Search in Google Scholar PubMed

[208] Simoes S, Moreira JN, Fonseca C, Düzgüneş N, de Lima MCP. On the formulation of pH-sensitive liposomes with long circulation times. Adv. Drug Deliv. Rev. 2004, 56, 947–965.10.1016/j.addr.2003.10.038Search in Google Scholar PubMed

[209] Zhang S, Zhao Y. Controlled release from cleavable polymerized liposomes upon redox and pH stimulation. Bioconjug. Chem. 2011, 22, 523–528.10.1021/bc1003197Search in Google Scholar PubMed

[210] Wang S, Zhang S, Liu J, Liu Z, Su L, Wang H, Chang J. pH-and reduction-responsive polymeric lipid vesicles for enhanced tumor cellular internalization and triggered drug release. ACS Appl. Mater. Interfaces 2014, 6, 10706–10713.10.1021/am502579eSearch in Google Scholar PubMed

[211] Wang S, Wang H, Liu Z, Wang L, Wang X, Su L, Chang J. Smart pH-and reduction-dual-responsive folate–PEG-coated polymeric lipid vesicles for tumor-triggered targeted drug delivery. Nanoscale 2014, 6, 7635–7642.10.1039/c4nr00843jSearch in Google Scholar PubMed

[212] Xu X, Zhang L, Assanhou AG, Wang L, Zhang Y, Li W, Xue L, Mo R, Zhang C. Acid/redox dual-activated liposomes for tumor-targeted drug delivery and enhanced therapeutic efficacy. RSC Adv. 2015, 5, 67803–67808.10.1039/C5RA06445GSearch in Google Scholar

[213] Li Y, Liu R, Yang J, Ma G, Zhang Z, Zhang X. Dual sensitive and temporally controlled camptothecin prodrug liposomes codelivery of siRNA for high efficiency tumor therapy. Biomaterials 2014, 35, 9731–9745.10.1016/j.biomaterials.2014.08.022Search in Google Scholar PubMed

[214] Wang L, Cao W, Yi Y, Xu H. Dual redox responsive coassemblies of diselenide-containing block copolymers and polymer lipids. Langmuir 2014, 30, 5628–5636.10.1021/la501054zSearch in Google Scholar PubMed

[215] Nahire R, Haldar MK, Paul S, Mergoum A, Ambre AH, Katti KS, Gange KN, Srivastava DK, Sarkar K, Mallik S. Polymer-coated echogenic lipid nanoparticles with dual release triggers. Biomacromolecules 2013, 14, 841–853.10.1021/bm301894zSearch in Google Scholar PubMed PubMed Central

Received: 2017-6-19
Accepted: 2017-9-20
Published Online: 2017-12-12
Published in Print: 2018-2-23

©2017 Walter de Gruyter GmbH, Berlin/Boston

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Downloaded on 19.4.2024 from https://www.degruyter.com/document/doi/10.1515/ntrev-2017-0154/html
Scroll to top button