Next Article in Journal
Analysis of RNA Polyadenylation in Healthy and Osteoarthritic Human Articular Cartilage
Next Article in Special Issue
Synchronization between Attractors: Genomic Mechanism of Cell-Fate Change
Previous Article in Journal
Short Photoperiod-Dependent Enrichment of Akkermansia spec. as the Major Change in the Intestinal Microbiome of Djungarian Hamsters (Phodopus sungorus)
Previous Article in Special Issue
Autophagy: A Potential Therapeutic Target to Tackle Drug Resistance in Multiple Myeloma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Liposomes in Cancer Therapy: How Did We Start and Where Are We Now

by
Melody D. Fulton
1 and
Wided Najahi-Missaoui
2,*
1
Department of Chemistry, College of Arts and Sciences, Washington State University, Pullman, WA 99164, USA
2
Department of Pharmaceutical and Biomedical Sciences, College of Pharmacy, University of Georgia, Athens, GA 30602, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(7), 6615; https://doi.org/10.3390/ijms24076615
Submission received: 28 February 2023 / Revised: 28 March 2023 / Accepted: 29 March 2023 / Published: 1 April 2023
(This article belongs to the Special Issue Latest Review Papers in Molecular Oncology 2023)

Abstract

:
Since their first discovery in the 1960s by Alec Bangham, liposomes have been shown to be effective drug delivery systems for treating various cancers. Several liposome-based formulations received approval by the U.S. Food and Drug Administration (FDA) and European Medicines Agency (EMA), with many others in clinical trials. Liposomes have several advantages, including improved pharmacokinetic properties of the encapsulated drug, reduced systemic toxicity, extended circulation time, and targeted disposition in tumor sites due to the enhanced permeability and retention (EPR) mechanism. However, it is worth noting that despite their efficacy in treating various cancers, liposomes still have some potential toxicity and lack specific targeting and disposition. This explains, in part, why their translation into the clinic has progressed only incrementally, which poses the need for more research to focus on addressing such translational limitations. This review summarizes the main properties of liposomes, their current status in cancer therapy, and their limitations and challenges to achieving maximal therapeutic efficacy.

1. Introduction: Cancer and Chemotherapy

Cancer is the second leading cause of death and constitutes a major public health burden worldwide [1,2]. In the United States, it is estimated that one in three women and one in two men will be diagnosed with cancer in their lifetimes [3]. Prostate and breast cancers account for the most common non-cutaneous cancers among men and women, respectively, in addition to lung cancer [1]. In fact, the World Health Organization listed breast cancer as the most commonly diagnosed cancer worldwide in 2020 [4]. Cancer is classified according to its staging number, which helps clinicians (and patients) assess the extent of the tumor. The most clinically useful staging system is the TNM staging, which refers to tumor, node, and metastasis, respectively. This system was developed by the American Joint Committee on Cancer (AJCC) and has since been referred to as the AJCC TNM system [5]. The TNM staging system assesses the size of the tumor (T), the involvement of regional lymph nodes (N), and any evidence of distant metastasis (M). T is usually associated to a number that can be 0, 1, 2, 3, or 4. Numbers after the T (such as T1, T2, T3, and T4) describe the tumor size and/or cancer spread to nearby structures. A higher T number indicates a larger tumor size and/or increased cancer invasion to local tissues [5,6].
Treatment options for cancer depend on many factors, including the stage, age, and patient comorbidity. Prostatectomy and mastectomy are the first treatment options for more than 50% of men with prostate cancer and women with breast cancer. In addition to surgery, the treatment options usually include a combination of chemotherapy, radiation, and hormonal therapy for advanced stages of cancer [7]. Photodynamic therapy has emerged as a non-invasive and effective approach in the local treatment of cancer. A photosensitizer may be combined with a cancer drug. The activation of the photosensitizer by an appropriate wavelength of light results in the generation of reactive oxygen species that are capable of damaging and killing cancer cells. This process can also cause an immune response inside the tumor that can cause more cancer cells to die [8,9]. Although beneficial, surgery and radiation therapy are usually associated with debilitating side effects, such as incontinence and bowel complications in prostate cancer patients [10]. Female patients under hormonal treatment may suffer from menopausal-like symptoms, such as hot flashes, night sweats, osteoporosis, increased risks of cardiovascular diseases, and obesity [11].
The development of resistance to chemotherapy remains a major problem in the management of many cancer types [12]. The mechanisms of resistance to chemotherapy are still not completely understood, but include the activation of cellular survival pathways to inhibit cell death mechanisms, as well as possible epigenetic mechanisms that are yet to be fully elucidated [13]. Cancer cells may also exhibit multidrug resistance (MDR), which is most commonly associated with drug efflux mechanisms via ATP-binding cassette (ABC) membrane transporters. Examples of the most studied MDR transporters include P-glycoprotein (P-gp), which is overexpressed in various cancers and can bind and efflux a large number of anticancer drugs. Other examples include multidrug resistance-associated protein-1 (MRP1) and breast cancer resistant proteins (ABCG2), which are able to reduce the intracellular delivery of drugs and reduce their efficacy [12]. Various approaches, including nanoparticles, have been studied to overcome resistance to chemotherapeutic drugs [14].

2. Nanoparticles and Liposome-Mediated Drug Delivery

The effectiveness of many chemotherapeutic drugs can be limited by their rapid metabolism, their toxic side effects, and the development of resistance [15]. To overcome these limitations, nanoparticles, such as liposomes, have been used to improve the therapeutic efficacy of various chemotherapeutic drugs. Liposomes provide several advantages, including improved pharmacokinetic properties of the encapsulated drug, long circulation time, and passive targeting and disposition in tumors and inflammatory sites due to the enhanced permeability and retention (EPR) mechanism. They can also reduce systemic toxicity associated with the free drug. In addition, liposomes can improve the solubility of drugs and provide slow and sustained release of encapsulated drugs (Figure 1). However, it is worthy to note that despite their efficacy in treating various cancers, nanoparticles, including liposomes, still have some potential toxicity and lack specific targeting and disposition [16,17,18].
Since their discovery in the 1960s by the late Alec Bangham, liposomes have been extensively studied as drug delivery systems, and they continue to be investigated in various research areas [19]. In the early 1970s, the work lead by Gregory Gregoriadis successfully delivered enzymes into the lysosomes of tissues in the reticuloendothelial system (RES) [20,21]. His following work, along with his colleagues, introduced the potential of exploiting liposomes as a drug delivery system [22,23,24]. Liposomes are considered to be one of the most successful drug-carrier systems, and several liposomal formulations are actively marketed or are in clinical trials [25]. Liposomes result from the self-assembly of phospholipids in an aqueous media, resulting in closed bilayered structures with an aqueous cavity and one or more bilayer phospholipid membranes (Figure 2) [26,27].
Phospholipids are the main components of cell membranes, which make them biocompatible. In addition, their amphiphilic properties enable self-assembly into bilayer membranes in aqueous environments [28,29]. These unique properties make phospholipids suitable for drug delivery systems such as liposomes. Phospholipids are characterized by their phase transition temperature (TC), which is the temperature at which phospholipids transit from gel crystalline to liquid crystalline states [30]. The TC depends on many factors, such as the nature of the polar head group of phospholipids, the length of their aliphatic chains, and the presence of unsaturation in their hydrocarbon chains [31].
Liposomal formulations usually include cholesterol incorporated into the lipid bilayer to decrease membrane fluidity and control the rate of drug release. Cholesterol can reduce the rotational freedom of the phospholipid hydrocarbon chains, which limits liposome interactions with plasma proteins and subsequent loss of the encapsulated material [32,33,34,35]. Cholesterol also plays an essential function in regulating the biophysical states of the phospholipids in the liposomes by controlling the lipid organization and phase behavior. Cholesterol decreases the order of phospholipids in the crystalline gel phase and increases the order in the liquid crystalline phase [31,36]. Studies have shown that adding cholesterol to liposomal formulations shifts the TC of phospholipids to a lower temperature, and a cholesterol composition above 30% abolishes the TC. Moreover, adding cholesterol increases the stability of liposomes and limits their leakage after systemic administration [36,37].
The unique structure of liposomes allows hydrophilic drugs to be retained in the aqueous interior. Hydrophobic drugs are usually inserted into the liposome bilayer; however, caution should be taken with this approach because high drug concentrations can disrupt liposomes (Figure 2). Amphipathic drugs can also be encapsulated in liposomes, provided the drug is partitioned between bilayer and aqueous phases [38].
Additional advantages of liposomes include biocompatibility, biodegradability, and decreased drug side effects [25]. Liposomes allow for controlled drug release and protection from rapid metabolism and clearance. Liposomes are also associated with improved patient compliance because of a decreased frequency of drug administration as compared to unencapsulated drugs [32,39,40]. Liposomes, like other nanoparticles, have some disadvantages, including possible carrier toxicity [41]. Typically, the toxicity of liposomes is lower, as compared to other nanoparticles. Liposomes are primarily composed of phospholipids, and other lipids, that are generally recognized as safe (GRAS), as well as biocompatible, biodegradable, and non-immunogenic [42,43]. Another limitation of liposomes is their preparation on a large industrial scale with reproducible properties [44]. The stability of liposomes constitutes another limitation, and lyophilizing the produced lipid vesicles is one of the proposed solutions to overcome this limitation [45,46].

3. Biophysical Characterization of Liposomes

In general, physical characterization of liposomes and nanoparticles (NPs) is necessary to accurately correlate a particular property of a nanoparticle with biological reactions. Particle size (or diameter) and surface area properties of NPs are key parameters for their characterization.

3.1. Transmission Electron Microscopy (TEM)

Electron microscopy techniques, such as transmission electron microscopy (TEM), can provide accurate size measurements of liposomes. However, this technique is unable to resolve any organic surface ligands conjugated to liposomes because of their low electron density. As such, the size determined by TEM is mainly of the core of liposomes. TEM is also limited in that it requires a high vacuum that may result in liposome aggregation during their preparation for imaging [47,48,49]. TEM suffers from limited sampling, and the data reported are not necessarily representative of all the NPs in one formulation [50]. Therefore, multiple methods should be used to determine the size of liposomes.

3.2. Dynamic Light Scattering (DLS)

Dynamic light scattering is one of the most commonly used methods to determine the size of liposomes. DLS can measure particle size, particle size distribution and surface charge, or zeta potential. DLS requires that the material be dispersed in a solvent and that all liposomes are individually dispersed. DLS is based on the fact that dispersed nanoparticles are in a continuous Brownian motion, and the illumination with a laser results in scattered light fluctuations at a rate dependent upon the size of the particles. The diameter measured by DLS is called the hydrodynamic diameter, which refers to how a particle diffuses within a fluid. While DLS is simple and quick, it has a limitation of low sensitivity towards small particles (i.e., lower limit for DLS measurement is around 10 nm) and interference from light-absorbing species [49].

3.3. Fluorescence Correlation Spectroscopy (FCS)

Fluorescence correlation spectroscopy is used to measure the hydrodynamic diameter of nanoparticles that are either intrinsically fluorescent, or labeled with fluorescent dyes [51]. FCS measures the duration of brief bursts of photons from individual nanoparticles passing through a small focal volume of typically 1 fL (10−15 L), and the size can be determined using autocorrelation analysis [52].

3.4. Encapsulation Efficiency (EE)

EE is defined as the ratio of the amount of encapsulated drug to the initial amount of drug added in the liposome formulation [46,53]. Once prepared, liposome formulations are usually a mixture of encapsulated and un-encapsulated drug fractions. In order to determine the EE of liposomes, a first step is to separate the unencapsulated drug fraction from the encapsulated drug. This can be achieved using size-exclusion chromatography [54]. This method separates the two fractions based on the differences in the size between the drug-encapsulated liposomes and the free drug, where the unencapsulated drug may be retained in the gel while loaded liposomes are not [55,56,57]. Loaded liposomes can also be separated from a free unencapsulated drug using a dialysis membrane with the appropriate cut-off pore size [58]. Once drug-loaded liposomes are separated from a free drug, the lipid bilayer is disrupted, and the released drug is then quantified using various techniques such as spectrophotometry and high-performance liquid chromatography (HPLC) [59].

3.5. Phospholipid Content

Phospholipid content of liposomes can be determined by spectrophotometric techniques using several methods, such as the Bartlett phosphate assay [60]. This method is based on the colorimetric measurement of inorganic phosphate. The phospholipid content of liposomes is measured after acid-based destruction of the phospholipid to inorganic phosphate. This step is followed by the addition of ammonium molybdate and the conversion of inorganic phosphate into phospho-molybdic acid. Ammonium molybdate is then reduced into a blue-colored complex that can be measured spectrophotometrically at 830 nm [60]. In vitro stability of liposome-encapsulated drugs can be assessed by dialyzing samples of liposomal formulations and quantifying the drug content at different time points, as indicated above [61].
Additional biophysical properties of liposomes that affect encapsulation of drugs in liposomes and their efficacy include phospholipid composition and lamellarity, bilayer rigidity, lipid-to-drug ratio, in vitro drug release, and drug-liposome interactions. A more detailed discussion of these properties is beyond the focus of this review. However, these properties were discussed in a couple of studies, and the reader is referred to these reviews that focus on these aspects of liposomes [62,63,64,65,66,67,68].

4. Preparation and Properties of Liposomes

Over the years, research publications reported the use of different methods to make liposomes. Most techniques used in liposome preparation include the dissolution of phospholipids in their appropriate organic solvents, followed by the removal of organic solvents to allow liposomes to form [69]. There are several different methods used to load drugs into liposomes, and these differ depending on the hydrophilicity or hydrophobicity of the drug being encapsulated, as well as whether liposomes are manufactured at a small laboratory scale or an industrial scale [70]. Hydrophobic drugs, such as paclitaxel and docetaxel, have been loaded into liposomes using the lipid film hydration method with sufficient encapsulation [71,72]. Hydrophilic drugs are encapsulated into liposomes using passive or remote (active) loading. Passive loading involves entrapping drugs as the lipid films are hydrated; however, a major limitation of this method is the low encapsulation efficiency, as most hydrophilic drugs remain entrapped in the external aqueous compartment [73,74]. Below are examples of commonly used methods in the preparation of liposomes.

4.1. Thin Lipid Film Hydration

Thin lipid film hydration is the simplest, oldest, and one of the most widely used methods at the research laboratory scale [75,76,77,78]. Phospholipids dissolved in organic solvents are subject to the removal of organic solvent via evaporation, resulting in a thin lipid film. Hydration of the lipid film results in heterogeneous liposomes dispersed in the aqueous solvent. Several techniques can reduce their heterogeneity and narrow their size distribution, including sonication and multiple extrusions through polycarbonate membranes [75,76,79,80]. This lipid film hydration method is easy, simple, and reproducible. However, one major limitation of this method is the low EE of hydrophilic drugs. Applying remote loading techniques can increase the EE, which involves changing the transmembrane pH or ionic gradients across the liposomal membranes. This method was successfully used in the encapsulation of anti-cancer drugs including doxorubicin [81,82].

4.2. Reverse Phase Evaporation

Reverse phase evaporation is a relatively simple method that is used to improve the EE of drugs into liposomes. This method is based on the formation of an emulsion of an aqueous phase (containing the drug) and an organic phase (containing the lipid), followed by the evaporation of the organic solvent, and the formation of an aqueous suspension containing the assembled liposomes [83,84]. Examples of drugs that have been encapsulated using this method include the anti-Alzheimer drug tacrine hydrochloride [85], the anti-cancer drug doxorubicin [86], and carboplatin [87].

4.3. Dehydration-Rehydration

Dehydration-rehydration is another method used to improve drug EE. This approach consists of dispersing preformed and lyophilized liposomes in an aqueous solvent [38]. This method consists of first making liposomal suspensions with small unilamellar vesicles followed by lyophilization. Upon rehydration, the liposomes reform and passively entrap the compound of interest. This method requires the addition of suitable lyoprotectants to preserve the liposomes’ bilayers integrity and prevent aggregation of liposomes during rehydration [88,89]. Radiocontrast agents (e.g., diatrizoate and iotrolan), DNA, and RNA have been encapsulated in liposomes with this method [90,91].
The methods described above have been used successfully in small-scale research laboratories. However, the large-scale industrial manufacturing of liposomes has not been as successful for many reasons. The industrial production of liposomes faces many limitations such as the batch-to-batch variability, which affects the chemical, physical, and performance qualities of the product. In addition, it is a long and laborious process that involves several operations. Additional levels of complexity are usually related to the additional testing performed prior to, during, and after the manufacturing, storage, and clinical applications as well as the lack of well controlled good manufacturing practices (GMPs) [70,92]. Recently, the preparation of drug-loaded liposomes has been revolutionized by the state-of-the-art microfluidic systems [93].

4.4. Microfluidic Techniques

Microfluidic techniques, especially microfluidic hydrodynamic focusing (MHF), have shown great potential to achieve high quality control over the physical properties of liposomes, including their uniformity in size and narrow size distribution. This technique is suitable for both hydrophilic and hydrophobic drugs and is based on solvent displacement methods. This technique uses a microfluidic device that introduces the lipid solution through a central channel, followed by the flow of aqueous solution. The aqueous solvent will then replace the organic solvent and promote the self-assembly of liposomes. This technique has the advantage of producing uniformly sized unilamellar liposomes because of the ability to change the flow rate of the lipid solution in the organic solvent and aqueous phase [92,94]. Other techniques to determine nanoparticle size includes atomic force microscopy (AFM), nanoparticle tracking analysis (NTA), absorption spectroscopy, and analytical ultracentrifugation [95,96,97,98].

5. Active Loading vs. Passive Loading of Drugs in Liposomes

Active or remote loading has been used to increase the EE of hydrophilic drugs. This method was first used by the Deamer research group, who used a pH gradient across the liposomal membranes to load catecholamine remotely into liposomes [99]. Remote loading is based on the fact that uncharged drugs will cross the liposomal membrane and become protonated and entrapped inside the polar cavity. This change in protonation means that these drugs can no longer easily diffuse across the bilayer membrane. Effective remote loading requires drugs that have a pKa ≤ 11 and a distribution coefficient (logD) within −2.5 to 2.0 at pH 7 [38,100]. Successful remote loading was achieved with the anti-cancer anthracycline drug, doxorubicin, known as Doxil®—an FDA-approved liposomal formulation used for the treatment of AIDS-related Kaposi’s sarcoma, recurrent ovarian cancer, and metastatic breast cancer [101,102]. Doxorubicin was stably encapsulated in this liposomal formulation using an ammonium sulfate pH gradient across the liposomal membrane. Doxorubicin is a weak base in the outer compartment and exchanges with the ammonium ions from the inner compartment. Once inside the polar cavity, doxorubicin is precipitated as a sulfate salt, which extends the controlled and sustained release of the drug from the liposomes, and extends its half-life to more than 50 h in plasma [102]. In addition, Doxil® decreased cardiomyopathy associated with free doxorubicin and allowed patients to receive fewer doses of the drug because of the extended half-life of the liposomes [103]. Table 1 lists several drugs that have been encapsulated in liposomes and approved by the U.S. Food and Drug Administration (FDA) and European Medicines Agency (EMA) for the treatment of different diseases or are currently in clinical trials (Table 2).
The surface charge and liposome composition play an important role in the fate of liposomes. Anionic and neutral liposomes have been shown to escape renal clearance; however, cationic liposomes can have specific interactions with various anionic species in the blood that enhance their clearance through the reticuloendothelial system (RES). Neutral liposomes have been shown to accumulate in solid tumors [141]. Increasing the mol % of cationic lipids in liposomes can induce aggregation through electrostatic interactions between the liposomes and the anionic species in the circulation, resulting in reduced liposome disposition at the target site [142]. Studies have also shown that cationic liposomes preferentially accumulate in the angiogenic tumor vessels, and they may be efficient drug carrier systems to target blood vessels of solid tumors. Because of their extravasation, neutral and anionic liposomes have been suggested as potential drug carriers to the extravascular compartments of tumors [143].

6. Strategies for Targeting Liposomes to Tumors

Liposomes have emerged as efficient carrier systems for therapeutic agents, owing, in part, to some of the unique properties discussed above. Various strategies have been developed to target liposomes to tumor sites. Some of these strategies involve using passive targeting and active targeting via surface functionalization as well as using various stimuli to trigger drug release from liposomes.

6.1. Liposomes and the EPR Effect (Passive Targeting)

Nanoparticles, such as long circulating liposomes, take advantage of the leaky nature of the blood vessels in tumor tissues. Because tumor blood vessels have increased fenestrations, liposomes can passively cross the capillary endothelial barrier and reach the interstitial space [144,145]. In normal non-tumor tissues, vascular endothelial cells are tightly connected and have small para-cellular gaps in the 5–10 nm range. In contrast, larger gaps exist between endothelial cells in tumor blood vessels, ranging from 100 to 700 nm, depending on the cancer [146,147]. In addition, because of their disorganized vascular architecture, solid tumors lack a functional lymphatic system (Figure 3). The combination of the leaky tumor vasculature and the limited lymphatic drainage is called the enhanced permeability and retention (EPR) effect, which allows the passive disposition and accumulation of liposomes into the tumor site [145,148,149,150,151,152,153].
First-generation, or conventional, liposomes had limited tumor disposition because of their rapid clearance by the reticular endothelial system (RES), and their opsonization by plasma proteins. In addition, these liposomes suffered from drug leakage during their systemic circulation [154]. Over the years, changes were made to improve these liposomes, including composition and surface modification, to produce the second-generation liposomes. These liposomes had improved stability, disposition, and efficacy compared to first generation liposomes. Cholesterol was added in the lipid bilayers of liposomes to increase their rigidity and reduce drug leakage [25,155]. The incorporation of polyethylene glycol or PEG (PEGylation) provided a steric protection of liposomes from electrostatic and hydrophobic interactions with plasma proteins, which decreased uptake by RES. In addition, PEGylation extended the circulation time of liposomes, allowing for a more effective drug delivery in vivo. These long-circulating liposomes were therefore named “stealth liposomes” [27,39,156,157].
The first stealth liposomal formulation to be approved for cancer therapy in the United States (1995) and European Union (1996) was Doxil®/Caelyx® [102,113]. Doxil® offers reduced cardiotoxicity and myelotoxicity in comparison to free doxorubicin, while achieving higher drug concentrations in tumors by using a liposomal composition of HSPC:CL:MPEG 2000-DSPE (calc. molar ratio 3:2:0.9, w/w 3:1:1) [112,158]. While the Doxil®/Caelyx® liposomal formulation is clinically efficacious, efforts have been made to change the formulation in order to improve the pharmacokinetic properties. Lipo-Dox® was created with a similar lipid molar ratio to Doxil®/Caelyx®, except the formulation consists of DSPC instead of HSPC (HSPC:CL:MPEG-DPSC, 3:2:0.3) [114,115]. DSPC has a higher phase transition temperature than HSPC [159], and DSPC consists of one type of long-chain fatty acid (i.e., stearic acid, C18). On the other hand, HSPC has a varied fatty acid composition of palmitic (C16) and stearic acid (C18) [160,161]. In a phase I study, replacing HSPC with DSPC in the liposomal doxorubicin formulation increased the half-life, as well as reduced the volume of distribution and clearance rate, which overall offered a better plasma area under the curve (AUC) performance than the reported plasma AUC for Doxil®/Caelyx® [162,163]. While stomatitis was noted as a new dose-limiting toxicity at 50 mg/m2 with the Lipo-Dox® formulation, which is not an observed limitation with Doxil®/Caelyx®, combinations of Lipo-Dox® with other cancer treatments have been efficacious for patients with ovarian, AIDS-related Kaposi’s sarcoma, and breast cancer in Taiwan [164,165,166]. Zolsketil®, a recently authorized bioequivalent to Caelyx®, is available in the European Union to treat advanced ovarian cancer, breast cancer, multiple myeloma, and AIDS-related Kaposi’s sarcoma [128]. Comparing the drug leaflets, Zolsketil® uses the same excipients as Caelyx®/Doxil® (i.e., HSPC, MPEG 2000-DSPE, CL, ammonium sulphate, histidine, sucrose, water, HCl, NaOH), but it is unclear at this time the excipient molar ratio or other modifications to the preparation of the liposomes that would distinguish the two stealth liposome formulations.
Although proven to be clinically useful, stealth liposomes depend mostly on their passive accumulation into tumor tissues; they lack the ability to control cellular uptake and drug release and rely only on passive drug efflux, which may result in limited efficacy [25]. For example, SPI-77 is a stealth liposome of cisplatin with a similar formulation to Doxil® (HSPC:CL:MPEG2000-DSPE), but it has not progressed beyond phase II clinical trials. While SPI-77 demonstrated a better toxicity profile over the conventional toxicities observed with cisplatin alone in non-small cell lung cancer patients, most patients in the phase II clinical studies did not respond to SPI-77 [167,168]. The third, or “new generation”, liposomes use ligand-mediated targeting or active targeting to improve biodistribution and liposome-mediated drug delivery at tumor sites [169].

6.2. Active Targeting of Liposomes

Most nanomedicines are using passive mechanisms, such as the EPR effect, to target tumors. Many of these have failed to get FDA approval for clinical use. Passive targeting of liposomes relies only on the pathophysiological properties at the tumor site and has limitations that include decreased efficacy and/or off-target toxicity [170]. One reason for this lack of clinical efficacy is that passively targeted liposomes lack true specificity for the tumor cells themselves. This has led several researchers to focus on more precise forms of targeting liposomes, such as active targeting. Active targeting uses molecular approaches to directly target tumor cells via interactions with tumor-specific markers [171,172]. Actively targeted liposomes are usually prepared by conjugating targeting moieties such as monoclonal antibodies, fragments of antibodies, or peptides to their surface [146]. This approach is a promising strategy for cancer therapy [146,173,174,175]. Active targeting utilizes specific pathological changes in the tumor microenvironment such as the overexpression of several proteins. Therefore, liposomes targeting these markers can be selectively taken up by cells that overexpress these proteins to achieve improved drug delivery [176,177,178,179]. While active targeting has the ability to target cells once liposomes are in the tumor microenvironment, it actually has no tumor targeting ability. Liposomes still rely on their passive movement to reach tumors. Recently, transcytosable nanomedicine has emerged as an alternative approach that has the potential to cross the vascular wall and diffuse more efficiently within the tumor tissue. The design of transcytosable nanomedicines depends on various forms of transcytosis, including receptor-mediated, adsorption-mediated, or fluid-mediated transcytosis [180,181].
Targeted drug delivery to cancer cells has gained significant interest and shown great potential due to the various overexpressed target proteins on cancer cells. Examples of these targets include human epidermal growth factor receptor 2 (HER2), epidermal growth factor receptor (EGFR), transferrin receptors, epithelial cell adhesion molecule (EpCAM), and vascular receptors [179,182]. The conjugation of antibodies to the surface of liposomes creates immunoliposomes that show enhanced cell binding and internalization compared to untargeted liposomes [174]. The majority of immunoliposomes currently studied are used for the delivery of anticancer drugs [183]. Antibodies are usually conjugated to PEG, and not to the liposomal phospholipids, to overcome the steric hindrance possibly caused by PEG interference with antibody-target protein interactions. Thus, the ligand is extended outside the PEG layer and is more accessible for binding to its target [183].
The conjugation of monoclonal antibodies (mAbs), or their Fab fragments against HER2, improved drug delivery to tumor sites and showed therapeutic efficacy in various HER2-overexpressing mouse xenograft models [184,185]. Doxorubicin-loaded liposomes conjugated to the nucleosome mAbs (2C5) showed increased antitumor efficacy in both in vitro and in vivo models of various over-expressing nucleosome tumors compared to non-actively targeted liposomes [186,187,188]. PEGylated liposomes conjugated to the internalizing receptor CD19, which is overexpressed in various B-lymphoid cancers, showed improved efficacy in a human CD19+ B-lymphoma mouse model compared to non-antibody conjugated liposomes [189,190].

6.3. Local Stimuli to Trigger Drug Release from Liposomes

Strategies also exist to increase drug release from liposomes after they accumulate in the tumor. Many of these strategies take advantage of pathological changes in the tumor microenvironment, such as altered pH, increased temperature, and overexpression of proteolytic enzymes such as secretory phospholipases [176,177]. External stimuli can also be applied to enhance or trigger drug release from liposomes [146]. pH-sensitive co-polymers can be added in liposomal formulations that are stable at a physiological pH, but these will be hydrolyzed at an acidic pH of 6 and lower, which is commonly found in the tumor microenvironment. Examples of polymers used in pH-responsive liposomes include poly (acrylic acid) and poly (methacrylic acid) [146,191]. Temperature-triggered drug delivery from liposomes involves local heating of the tumor site to increase tumor permeability by increasing vascular pore size. This increases liposomal extravasation and accumulation in the tumor microenvironment [192,193]. Temperature-sensitive liposomes are usually prepared by incorporating thermosensitive lipids with a specific gel-to-liquid phase transition temperature, such as dipalmitoylphosphatidylcholine (DPPC). These thermosensitive liposomes have been shown to release more than 80% of encapsulated methotrexate in the tumor site after raising the temperature from 37 °C to 41 °C [146,194,195]. ThermoDox® (Celsion, NJ, USA), another thermosensitive liposomal formulation, showed improved doxorubicin delivery and efficacy in mouse models and has advanced to phase III clinical trials for treating hepatocellular carcinoma and breast cancer [194,196].
Taking advantage of specific pathological changes in the tumor microenvironment can increase drug release from liposomes, such as the overexpression of enzymes (e.g., matrix metalloproteinases (MMPs) and phospholipase A2) [78,197]. The activity of these enzymes can mediate the uptake and release of encapsulated drugs from enzyme-sensitive or responsive liposomes [198]. Despite the extensive research and the development of different liposome formulations, the sub-optimal potency is still a major limitation of liposomes. For instance, the most successful nanomedicine, Doxil®, can only achieve modest benefits. Additional work is needed and has to focus on how to improve the therapeutic efficacy of liposomes. While these strategies exist to increase drug delivery, another factor that is limiting drug release includes the PEG layer in stealth liposomes.

7. PEGylation of Liposomes

PEGylation offers stealth properties to liposomes, including evasion of the mononuclear phagocytic system and extended circulation times that are responsive to PEG length and density [199]. Increasing the percent of grafted PEG on liposomes (i.e., 2–5 mol%), and using PEG2000 or PEG5000, markedly reduces protein adsorption, phagocytosis, and cellular adhesion of erythrocytes, lymphocytes, and macrophages [200,201]. However, some of the beneficial properties of PEGylated liposomes can create a few challenges for maximizing drug delivery, cell uptake, and endosomal escape.

7.1. Accelerated Blood Clearance

In animal models, increased blood clearance and increased accumulation in the liver and spleen can occur after a second injection (i.e., <4 weeks from the 1st injection) of PEGylated liposomes, which is known as the accelerated blood clearance phenomenon [202,203,204]. In addition, the second injection (i.e., <1 week) of PEGylated liposomes results in significantly increased IgM production in rats [205], despite using different types of lipids (i.e., EPC, SPC, ESM, HSPC, DPPC) [204]. Interestingly, the accelerated blood clearance phenomenon is not associated with Doxil® in clinical practice, but the significance of this phenomenon in clinical studies is still an area of debate [206,207]. Moreover, various strategies are still being pursued, such as employing alternative PEG-lipid/cholesterol derivatives (e.g., PEG-cholesteryl hemisuccinate) or a cleavable PEG (e.g., pH), to significantly reduce the accelerated blood clearance phenomenon [208,209,210]. One of the additional benefits of using a cleavable PEG coating is that it can help improve cargo delivery to target cells.

7.2. Cell Uptake and Cargo Delivery of PEGylated Nanoparticles

PEGylation of nanoparticles can introduce varying degrees of inhibitory effects on cell uptake, tumor-targeted efficiency, and endosomal escape; however, not all PEGylated formulations lead to these effects, and it can vary depending on the target cells [211,212,213,214]. The inhibitory effects of PEGylation have been reported with solid nanoparticles [215,216,217], lipid-DNA complexes [212,218,219], and chemotherapeutic loaded liposomal formulations [220,221], and recent reviews are available to provide greater depth in this topic [222,223].
PEG length and density play an essential role in cell uptake and endosomal escape. Starting simply on the level of measuring how PEGylation impacts liposome-to-liposome fusion, Holland et al. performed in vitro fusion assays measuring the changes in resonance energy transfer from mixing fluorescently labeled liposomes (i.e., Rh-PE and NBD-PE) with non-fluorescent liposomes, followed by the addition of CaCl2 to promote liposome fusion [224]. They observed that increasing the mol % of the PEG-lipid incorporated in the liposomes reduces the maximal % vesicle fusion, and 2 mol % was enough to completely inhibit vesicle fusion. Additionally, the rate of fusion decreased with increasing acyl chain length and saturation in the PEG-lipid conjugate. Increasing the PEG length (i.e., 2000 to 5000) markedly reduced the maximal % fusion achieved with increasing the mol % of the PEG-lipid conjugate. Interestingly, Brandenberger et al. observed that PEG5000 coated gold nanoparticles had significantly less cell uptake in human lung carcinoma cells (A549) in comparison to plain gold nanoparticles [215]. Furthermore, Song et al. reported that increasing the PEG length (i.e., 220 to 3400) or increasing the hydrophobic anchor of the conjugated PEG lipid (i.e., ceramide C8 to C20) in unilamellar DNA/lipid complexes markedly inhibited gene transfection in liver (HepG2) and cervical cancer (HeLa) cells [212]. Monitoring cell uptake of fluorescently labeled lipid and DNA complexes revealed that while the lipid/DNA complexes were endocytosed, the longer PEG and ceramide chain lengths inhibited cargo release, which resulted in complexes unable to escape the perinuclear region. Hence, the PEG density and length need to be balanced to support optimal cargo delivery.
In the area of liposome-based chemotherapeutics, PEGylation has offered reduced accumulation in the liver, spleen, and heart over free drug, but some studies report that tumor-targeting efficiency (Te = AUCtumor/AUCplasma) is reduced in comparison to non-PEGylated liposomal formulations. For example, Parr et al. reported that DSPC/CL doxorubicbin liposomal formulations, with or without PEG2000-PE, achieved similar drug accumulation levels by day 4 in BDF-1 mice bearing Lewis Lung carcinoma, but the formulation without PEG2000-PE achieved significantly higher drug accumulation levels in tumors earlier (1 h to 48 h) in the animal study [221]. Therapeutic efficacy was not significantly different between formulations with or without PEG2000-PE. The tumor targeting efficiency (Te) of the DSPC/CL doxorubicin formulation without PEG was nearly 2-fold greater than the Te value (0.76) of the DSPC/CL/PEG2000-PE formulation (0.40). Hong et al. also noticed a similar observation of lower tumor targeting efficiency with their own comparison studies of PEGylated and non-PEGylated liposomal doxorubicin in C-26 carcinoma tumor-bearing mice [225]. By 24 h, the doxorubicin concentration in the tumor was higher than plasma with the non-PEGylated formulation, while the drug concentration remained lower than plasma with the PEGylated formulation. Overall, the doxorubicin liposomal formulation without PEG offered a greater accumulation of drug in tumors over 72 h than with PEG. The drug targeting efficiency (Te) was greater than 2-fold for non-PEGylated liposomal doxorubicin (0.87) in comparison to the PEGylated liposomal doxorubicin (0.31). Moreover, while both liposomal formulations were significantly more efficacious at extending survival of C-26 mice in comparison to free drug, there was no significant difference in survival with or without PEG in the liposomal formulation. Of note, PEGylated liposomal doxorubicin had the greatest advantage over the non-PEGylated formulation with reducing drug accumulation in the liver and spleen.

7.3. Cleavable PEG Coatings

In an effort to still use the steric stabilization offered by PEGylation, others have taken the approach to create cleavable PEG-lipid linkages in order to shed the PEG coating. One effective strategy to overcome these limitations is by installing acid-labile acetal, hydrazine, hydrazone, or vinyl ether linkages [226,227] between PEG and the lipid or polymer, which takes advantage of the slightly acidic tumor microenvironment (pH 5.6–7 [228,229]) as well as endosomes (pH 5.5–6.5 [230]). For example, Kanamala et al. linked PEG2000 to 1,2-dipalmitoyl-sn-glycero-3-phosphoethanolamine (DPPE) via an acid-labile hydrazine bond, which was incorporated at 5 mol% into DOPE, DSPC, CHEMS, cholesterol liposomes (4:2:2:2 molar ratio) with calcein loaded in the core, and Nile Red in the bilayer [231]. Visible by live cell imaging and confocal fluorescence microscopy, the non-cleavable and cleavable PEG pH-sensitive liposomes (pSLs) were both internalized by pancreatic cancer cells (MIA PaCa-2) after 30 min, but calcein release was clearly seen in the cytoplasm by 1 h with the cleavable PEG liposomes. On the other hand, the liposomes with covalently linked PEG remained as a punctate distribution. Using the same formulation, gemcitabine was encapsulated in the cleavable PEG pSLs and demonstrated increased cytotoxicity over the covalently linked PEG pSLs in MIA PaCa-2, and even performed better than free gemcitabine in a glioblastoma cell line (U-87) after 24 h of exposure. An in vivo biodistribution study with CD-1 nude mice bearing MIA PaCa-2 tumors demonstrated a statistically significant increase in gemcitabine accumulation in the tumor with cleavable PEG pSLs than the covalently linked PEG pSLs and free drug. While gemcitabine levels were significantly increased in the liver with both types of pSL formulations in comparison to the free drug, both formulations offered significantly less drug accumulation in the spleen and heart at 4 h. In addition to enhancing drug accumulation, endosomal escape, and cell uptake, a cleavable PEG strategy can be useful to shield targeting moieties for enhancing cell uptake upon reaching the tumor site.
Strategies to target liposomes (e.g., peptide-targeted liposomes) or use fusogenic materials to increase tumor cell uptake and drug delivery can be concealed with a cleavable PEG polymeric layer that would be shed upon an external stimulus (i.e., lower pH or reducing conditions) [232,233,234,235] and expose the targeting ligands to guide the liposomes to the tumor cells. Hak et al. noticed that keeping the PEG surface density to <10 mol% in 100 nm of αvβ3-integrin targeted (i.e., using an RGD peptide) nanoemulsions was optimal for targeting tumor sites [236]. Consistent with the idea that reducing the PEG density or shedding PEG supports targeted delivery, Geng et al. recently reported the in vitro and in vivo performance of an αvβ3-integrin targeted (i.e., RGD peptide), PEG cleavable, doxorubicin liposomal formulation that uses near-infrared (NIR) light to shed the PEG thermo-labile linker, 4,4′-azobis (4-cyanovaleric acid). Upon exposure to NIR, the thermo-labile, PEGylated doxorubicin liposomes demonstrate significantly increased cell uptake, enhanced drug accumulation at the tumor site, and efficacy in H22 tumor-bearing mice when compared against doxorubicin or the thermo-stabile PEGylated doxorubicin formulation [237].
Yang et al. designed a charge reversed doxorubicin liposomal formulation (CRDOXIL) using a 4:1:1 weight ratio of HSPC, Chol, and cleavable PEG2000 [238]. PEG2000 is attached to a distearoyl via a diakylmaleamidic amide linkage, which falls apart at a low pH to expose a primary amine. Hence, these liposomes will primarily bear a positively charged surface upon encountering a low pH environment (pH 6.5), which is anticipated to facilitate cell uptake as well as endosomal escape. Indeed, the CRDOXIL formulation had increased cell uptake than the DOXIL (weight ratio of 4:1:1 with HSPC, CHOL, and DSPE-mPEG2000) based on fluorescence microscopy and flow cytometry. Moreover, while both DOXIL and CRDOXIL performed similarly with normal lung (BEAS-2B) and liver (L02) cell lines at pH 7.4, only CRDOXIL could demonstrate nearly the same reduction in cell viability to free doxorubicin when cultured with lung (A549) and liver (HepG2) cancer cells at pH 6.5.
The challenge in employing some of the common acid-labile chemical linkages (i.e., acid-labile acetal, hydrazine, hydrazone, or vinyl ether) is minimizing premature release of drug payloads at physiological pH, while ensuring the release is at pH 4.5–6.5 [223,239]. Unfortunately, many of the acid-labile linkages lack the stability to withstand the extended circulation times of stealth liposomes (e.g., 45 h in humans) at physiological pH [240]. A chemical linkage such as the PhosAm technology, which is stable for >70 h at pH 7.4 but rapidly hydrolyzes (t½ < 1 h) at pH 5.5, would be well matched with the hours of circulation that liposomes achieve [241,242]. Overall, there is promising potential in achieving a balance between the stealth properties of PEGylated liposomes and maximizing drug delivery by shedding the PEG layer.

8. Summary and Conclusions

Liposomes represent an attractive delivery system due to their physicochemical properties that allow overcoming various challenges and limitations with drug delivery. The use of liposomes to improve drug delivery has greatly impacted various biomedical areas. Liposomes have been shown to improve stability and biodistribution of therapeutic agents, overcome limitations to tissue and cellular uptake in target sites in vivo, and reduce systemic toxicity associated with non-encapsulated agents. However, despite the considerable preclinical work on liposomes, their translation into the clinic has progressed only incrementally. Future research will need to focus on addressing such translational limitations. This will require continuous communications and collaborations between experts in all stages of pharmaceutical development, including pre-clinical and clinical applications as well as toxicological evaluations.

Author Contributions

Conceptualization, W.N.-M.; writing—original draft preparation, W.N.-M. and M.D.F.; writing—review and editing, W.N.-M. and M.D.F.; supervision, W.N.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This work received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Torre, L.A.; Bray, F.; Siegel, R.L.; Ferlay, J.; Lortet-Tieulent, J.; Jemal, A. Global cancer statistics, 2012. CA Cancer J. Clin. 2015, 65, 87–108. [Google Scholar] [CrossRef] [Green Version]
  2. Ryerson, A.B.; Eheman, C.R.; Altekruse, S.F.; Ward, J.W.; Jemal, A.; Sherman, R.L.; Henley, S.J.; Holtzman, D.; Lake, A.; Noone, A.M.; et al. Annual Report to the Nation on the Status of Cancer, 1975–2012, featuring the increasing incidence of liver cancer. Cancer 2016, 122, 1312–1337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2018. CA Cancer J. Clin. 2018, 68, 7–30. [Google Scholar] [CrossRef] [PubMed]
  4. Cancer. Available online: https://www.who.int/news-room/fact-sheets/detail/cancer (accessed on 3 March 2023).
  5. Siegel, R.; DeSantis, C.; Virgo, K.; Stein, K.; Mariotto, A.; Smith, T.; Cooper, D.; Gansler, T.; Lerro, C.; Fedewa, S.; et al. Cancer treatment and survivorship statistics, 2012. CA Cancer J. Clin. 2012, 62, 220–241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Cowherd, S.M. Tumor staging and grading: A primer. Methods Mol. Biol. 2012, 823, 1–18. [Google Scholar] [PubMed]
  7. Holohan, C.; Van Schaeybroeck, S.; Longley, D.B.; Johnston, P.G. Cancer drug resistance: An evolving paradigm. Nat. Rev. Cancer 2013, 13, 714–726. [Google Scholar] [CrossRef]
  8. Gamal, H.; Tawfik, W.; Fahmy, H.M.; El-Sayyad, H.H. Breakthroughs of using Photodynamic Therapy and Gold Nanoparticles in Cancer Treatment. In Proceedings of the IEEE International Conference on Nanoelectronics, Nanophotonics, Nanomaterials, Nanobioscience & Nanotechnology (5NANO), Kottayam, Kerala, India, 29–30 April 2021; pp. 1–4. [Google Scholar]
  9. Zhang, Q.; Li, L. Photodynamic combinational therapy in cancer treatment. Off. J. Balk. Union Oncol. 2018, 23, 561–567. [Google Scholar]
  10. Katz, A.; Ferrer, M.; Suarez, J.F. Comparison of quality of life after stereotactic body radiotherapy and surgery for early-stage prostate cancer. Radiat. Oncol. 2012, 7, 194. [Google Scholar] [CrossRef] [Green Version]
  11. Kintzel, P.E.; Chase, S.L.; Schultz, L.M.; O’Rourke, T.J. Increased risk of metabolic syndrome, diabetes mellitus, and cardiovascular disease in men receiving androgen deprivation therapy for prostate cancer. Pharmacotherapy 2008, 28, 1511–1522. [Google Scholar] [CrossRef]
  12. Saraswathy, M.; Gong, S. Different strategies to overcome multidrug resistance in cancer. Biotechnol. Adv. 2013, 31, 1397–1407. [Google Scholar] [CrossRef]
  13. Lackner, M.R.; Wilson, T.R.; Settleman, J. Mechanisms of acquired resistance to targeted cancer therapies. Future Oncol. 2012, 8, 999–1014. [Google Scholar] [CrossRef]
  14. Kirtane, A.R.; Kalscheuer, S.M.; Panyam, J. Exploiting nanotechnology to overcome tumor drug resistance: Challenges and opportunities. Adv. Drug Deliv. Rev. 2013, 65, 1731–1747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Marin, J.J.; Sanchez de Medina, F.; Castano, B.; Bujanda, L.; Romero, M.R.; Martinez-Augustin, O.; Moral-Avila, R.D.; Briz, O. Chemoprevention, chemotherapy, and chemoresistance in colorectal cancer. Drug Metab. Rev. 2012, 44, 148–172. [Google Scholar] [CrossRef] [PubMed]
  16. England, C.G.; Ng, C.F.; van Berkel, V.; Frieboes, H.B. A Review of Pharmacological Treatment Options for Lung Cancer: Emphasis on Novel Nanotherapeutics and Associated Toxicity. Curr. Drug Targets 2015, 16, 1057–1087. [Google Scholar] [CrossRef]
  17. Soliman, G.M. Nanoparticles as safe and effective delivery systems of antifungal agents: Achievements and challenges. Int. J. Pharm. 2017, 523, 15–32. [Google Scholar] [CrossRef] [PubMed]
  18. Straubinger, R.M.; Arnold, R.D.; Zhou, R.; Mazurchuk, R.; Slack, J.E. Antivascular and antitumor activities of liposome-associated drugs. Anticancer Res. 2004, 24, 397–404. [Google Scholar]
  19. Deamer, D.W. From “banghasomes” to liposomes: A memoir of Alec Bangham, 1921–2010. FASEB J. 2010, 24, 1308–1310. [Google Scholar] [CrossRef]
  20. Gregoriadis, G.; Ryman, B.E. Fate of protein-containing liposomes injected into rats. An approach to the treatment of storage diseases. Eur. J. Biochem. 1972, 24, 485–491. [Google Scholar] [CrossRef]
  21. Gregoriadis, G.; Buckland, R.A. Enzyme-containing liposomes alleviate a model for storage disease. Nature 1973, 244, 170–172. [Google Scholar] [CrossRef]
  22. Gregoriadis, G. The carrier potential of liposomes in biology and medicine (first of two parts). N. Engl. J. Med. 1976, 295, 704–710. [Google Scholar] [CrossRef]
  23. Gregoriadis, G.; Neerunjun, E.D. Treatment of tumour bearing mice with liponsome-entrapped actinomycin D prolongs their survival. Res. Commun. Chem. Pathol. Pharmacol. 1975, 10, 351–362. [Google Scholar]
  24. Gregoriadis, G. Liposomes in Drug Delivery: How It All Happened. Pharmaceutics 2016, 8, 19. [Google Scholar] [CrossRef] [Green Version]
  25. Bozzuto, G.; Molinari, A. Liposomes as nanomedical devices. Int. J. Nanomed. 2015, 10, 975–999. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Forssen, E.A.; Tokes, Z.A. Use of anionic liposomes for the reduction of chronic doxorubicin-induced cardiotoxicity. Proc. Natl. Acad. Sci. USA 1981, 78, 1873–1877. [Google Scholar] [CrossRef] [Green Version]
  27. Sercombe, L.; Veerati, T.; Moheimani, F.; Wu, S.Y.; Sood, A.K.; Hua, S. Advances and Challenges of Liposome Assisted Drug Delivery. Front. Pharmacol. 2015, 6, 286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Suetsugu, S.; Kurisu, S.; Takenawa, T. Dynamic shaping of cellular membranes by phospholipids and membrane-deforming proteins. Physiol. Rev. 2014, 94, 1219–1248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Dowhan, W. Understanding phospholipid function: Why are there so many lipids? J. Biol. Chem. 2017, 292, 10755–10766. [Google Scholar] [CrossRef] [Green Version]
  30. Huang, C.H. Mixed-chain phospholipids: Structures and chain-melting behavior. Lipids 2001, 36, 1077–1097. [Google Scholar] [CrossRef]
  31. Feigenson, G.W. Phase behavior of lipid mixtures. Nat. Chem. Biol. 2006, 2, 560–563. [Google Scholar] [CrossRef]
  32. Papahadjopoulos, D.; Nir, S.; Oki, S. Permeability properties of phospholipid membranes: Effect of cholesterol and temperature. Biochim. Biophys. Acta 1972, 266, 561–583. [Google Scholar] [CrossRef]
  33. Deniz, A.; Sade, A.; Severcan, F.; Keskin, D.; Tezcaner, A.; Banerjee, S. Celecoxib-loaded liposomes: Effect of cholesterol on encapsulation and in vitro release characteristics. Biosci. Rep. 2010, 30, 365–373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Bardania, H.; Tarvirdipour, S.; Dorkoosh, F. Liposome-targeted delivery for highly potent drugs. Artif. Cells Nanomed. Biotechnol. 2017, 45, 1478–1489. [Google Scholar] [CrossRef]
  35. Tiwari, G.; Tiwari, R.; Sriwastawa, B.; Bhati, L.; Pandey, S.; Pandey, P.; Bannerjee, S.K. Drug delivery systems: An updated review. Int. J. Pharm. Investig. 2012, 2, 2–11. [Google Scholar] [CrossRef] [Green Version]
  36. Redondo-Morata, L.; Giannotti, M.I.; Sanz, F. Influence of cholesterol on the phase transition of lipid bilayers: A temperature-controlled force spectroscopy study. Langmuir 2012, 28, 12851–12860. [Google Scholar] [CrossRef] [PubMed]
  37. Kaur, R.; Henriksen-Lacey, M.; Wilkhu, J.; Devitt, A.; Christensen, D.; Perrie, Y. Effect of incorporating cholesterol into DDA:TDB liposomal adjuvants on bilayer properties, biodistribution, and immune responses. Mol. Pharm. 2014, 11, 197–207. [Google Scholar] [CrossRef] [Green Version]
  38. Eloy, J.O.; Claro de Souza, M.; Petrilli, R.; Barcellos, J.P.; Lee, R.J.; Marchetti, J.M. Liposomes as carriers of hydrophilic small molecule drugs: Strategies to enhance encapsulation and delivery. Colloids Surf. B Biointerfaces 2014, 123, 345–363. [Google Scholar] [CrossRef]
  39. Drummond, D.C.; Meyer, O.; Hong, K.; Kirpotin, D.B.; Papahadjopoulos, D. Optimizing liposomes for delivery of chemotherapeutic agents to solid tumors. Pharm. Rev. 1999, 51, 691–743. [Google Scholar] [PubMed]
  40. Langer, R. Drug delivery and targeting. Nature 1998, 392 (Suppl. 6679), 5–10. [Google Scholar] [PubMed]
  41. Missaoui, W.N.; Arnold, R.D.; Cummings, B.S. Toxicological status of nanoparticles: What we know and what we don’t know. Chem. Biol. Interact. 2018, 295, 1–12. [Google Scholar] [CrossRef]
  42. Kshirsagar, N.A.; Pandya, S.K.; Kirodian, G.B.; Sanath, S. Liposomal drug delivery system from laboratory to clinic. J. Postgrad. Med. 2005, 51 (Suppl. S1), S5–S15. [Google Scholar]
  43. Najahi-Missaoui, W.; Arnold, R.D.; Cummings, B.S. Safe Nanoparticles: Are We There Yet? Int. J. Mol. Sci. 2020, 22, 385. [Google Scholar] [CrossRef] [PubMed]
  44. Mozafari, M.R. Liposomes: An overview of manufacturing techniques. Cell. Mol. Biol. Lett. 2005, 10, 711–719. [Google Scholar] [PubMed]
  45. Chen, C.; Han, D.; Cai, C.; Tang, X. An overview of liposome lyophilization and its future potential. J. Control. Release 2010, 142, 299–311. [Google Scholar] [CrossRef]
  46. Chang, H.I.; Yeh, M.K. Clinical development of liposome-based drugs: Formulation, characterization, and therapeutic efficacy. Int. J. Nanomed. 2012, 7, 49–60. [Google Scholar]
  47. Shang, L.; Nienhaus, K.; Nienhaus, G.U. Engineered nanoparticles interacting with cells: Size matters. J. Nanobiotechnol. 2014, 12, 5. [Google Scholar] [CrossRef] [Green Version]
  48. Mahl, D.; Diendorf, J.; Meyer-Zaika, W.; Epple, M. Possibilities and limitations of different analytical methods for the size determination of a bimodal dispersion of metallic nanoparticles. Colloids Surfaces Physicochem. Eng. Asp. 2011, 377, 386–392. [Google Scholar] [CrossRef]
  49. Cho, E.J.; Holback, H.; Liu, K.C.; Abouelmagd, S.A.; Park, J.; Yeo, Y. Nanoparticle characterization: State of the art, challenges, and emerging technologies. Mol. Pharm. 2013, 10, 2093–2110. [Google Scholar] [CrossRef] [Green Version]
  50. Pietroiusti, A.; Magrini, A. Engineered nanoparticles at the workplace: Current knowledge about workers’ risk. Occup. Med. 2014, 64, 319–330. [Google Scholar] [CrossRef] [Green Version]
  51. Rocker, C.; Potzl, M.; Zhang, F.; Parak, W.J.; Nienhaus, G.U. A quantitative fluorescence study of protein monolayer formation on colloidal nanoparticles. Nat. Nanotechnol. 2009, 4, 577–580. [Google Scholar] [CrossRef] [PubMed]
  52. Zemanova, L.; Schenk, A.; Valler, M.J.; Nienhaus, G.U.; Heilker, R. Confocal optics microscopy for biochemical and cellular high-throughput screening. Drug Discov. Today 2003, 8, 1085–1093. [Google Scholar] [CrossRef]
  53. Arifin, D.R.; Palmer, A.F. Determination of size distribution and encapsulation efficiency of liposome-encapsulated hemoglobin blood substitutes using asymmetric flow field-flow fractionation coupled with multi-angle static light scattering. Biotechnol. Prog. 2003, 19, 1798–1811. [Google Scholar] [CrossRef] [PubMed]
  54. Grabielle-Madelmont, C.; Lesieur, S.; Ollivon, M. Characterization of loaded liposomes by size exclusion chromatography. J. Biochem. Biophys. Methods 2003, 56, 189–217. [Google Scholar] [CrossRef] [PubMed]
  55. Desormeaux, A.; Bergeron, M.G. Lymphoid tissue targeting of anti-HIV drugs using liposomes. Methods Enzymol. 2005, 391, 330–351. [Google Scholar]
  56. Petkowicz, J.; Byra, A.; Szumilo, T. The hypoglycaemic response of diabetic rats to insulin-liposomes. Acta Physiol. Pol. 1990, 41, 97–103. [Google Scholar]
  57. Lundahl, P.; Yang, Q. Liposome chromatography: Liposomes immobilized in gel beads as a stationary phase for aqueous column chromatography. J. Chromatogr. 1991, 544, 283–304. [Google Scholar] [CrossRef]
  58. Bragagni, M.; Mennini, N.; Ghelardini, C.; Mura, P. Development and characterization of niosomal formulations of doxorubicin aimed at brain targeting. J. Pharm. Pharm. Sci. 2012, 15, 184–196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Rossi, C.; Fardella, G.; Chiappini, I.; Perioli, L.; Vescovi, C.; Ricci, M.; Giovagnoli, S.; Scuota, S. UV spectroscopy and reverse-phase HPLC as novel methods to determine Capreomycin of liposomal fomulations. J. Pharm. Biomed. Anal. 2004, 36, 249–255. [Google Scholar] [CrossRef]
  60. Bartlett, G.R. Phosphorus assay in column chromatography. J. Biol. Chem. 1959, 234, 466–468. [Google Scholar] [CrossRef]
  61. Mayer, L.D.; Tai, L.C.; Ko, D.S.; Masin, D.; Ginsberg, R.S.; Cullis, P.R.; Bally, M.B. Influence of vesicle size, lipid composition, and drug-to-lipid ratio on the biological activity of liposomal doxorubicin in mice. Cancer Res. 1989, 49, 5922–5930. [Google Scholar]
  62. Kulkarni, S.B.; Betageri, G.V.; Singh, M. Factors affecting microencapsulation of drugs in liposomes. J. Microencapsul. 1995, 12, 229–246. [Google Scholar] [CrossRef]
  63. Tardi, P.G.; Boman, N.L.; Cullis, P.R. Liposomal doxorubicin. J. Drug Target. 1996, 4, 129–140. [Google Scholar] [CrossRef] [PubMed]
  64. Fenske, D.B.; Cullis, P.R. Liposomal nanomedicines. Expert Opin. Drug Deliv. 2008, 5, 25–44. [Google Scholar] [CrossRef] [PubMed]
  65. Demetzos, C. Differential Scanning Calorimetry (DSC): A tool to study the thermal behavior of lipid bilayers and liposomal stability. J. Liposome Res. 2008, 18, 159–173. [Google Scholar] [CrossRef] [PubMed]
  66. Franzen, U.; Ostergaard, J. Physico-chemical characterization of liposomes and drug substance-liposome interactions in pharmaceutics using capillary electrophoresis and electrokinetic chromatography. J. Chromatogr. A 2012, 1267, 32–44. [Google Scholar] [CrossRef]
  67. Rohilla, S.; Dureja, H. Recent Patents, Formulation and Characterization of Nanoliposomes. Recent Pat. Drug Deliv. Formul. 2015, 9, 213–224. [Google Scholar] [CrossRef]
  68. Vemuri, S.; Rhodes, C.T. Preparation and characterization of liposomes as therapeutic delivery systems: A review. Pharm. Acta Helvetiae 1995, 70, 95–111. [Google Scholar] [CrossRef]
  69. Ahmed, K.S.; Hussein, S.A.; Ali, A.H.; Korma, S.A.; Qiu, L.; Chen, J. Liposome: Composition, characterisation, preparation, and recent innovation in clinical applications. J. Drug Target. 2019, 27, 742–761. [Google Scholar] [CrossRef]
  70. Shah, S.; Dhawan, V.; Holm, R.; Nagarsenker, M.S.; Perrie, Y. Liposomes: Advancements and innovation in the manufacturing process. Adv. Drug Deliv. Rev. 2020, 154–155, 102–122. [Google Scholar] [CrossRef]
  71. Crosasso, P.; Ceruti, M.; Brusa, P.; Arpicco, S.; Dosio, F.; Cattel, L. Preparation, characterization and properties of sterically stabilized paclitaxel-containing liposomes. J. Control. Release 2000, 63, 19–30. [Google Scholar] [CrossRef]
  72. Hua, H.; Zhang, N.; Liu, D.; Song, L.; Liu, T.; Li, S.; Zhao, Y. Multifunctional gold nanorods and docetaxel-encapsulated liposomes for combined thermo- and chemotherapy. Int. J. Nanomed. 2017, 12, 7869–7884. [Google Scholar] [CrossRef] [Green Version]
  73. Castile, J.D.; Taylor, K.M. Factors affecting the size distribution of liposomes produced by freeze-thaw extrusion. Int. J. Pharm. 1999, 188, 87–95. [Google Scholar] [CrossRef] [PubMed]
  74. Mayer, L.D.; Hope, M.J.; Cullis, P.R.; Janoff, A.S. Solute distributions and trapping efficiencies observed in freeze-thawed multilamellar vesicles. Biochim. Biophys. Acta 1985, 817, 193–196. [Google Scholar] [CrossRef]
  75. Zhang, H. Thin-Film Hydration Followed by Extrusion Method for Liposome Preparation. Methods Mol. Biol. 2017, 1522, 17–22. [Google Scholar]
  76. Olson, F.; Hunt, C.A.; Szoka, F.C.; Vail, W.J.; Papahadjopoulos, D. Preparation of liposomes of defined size distribution by extrusion through polycarbonate membranes. Biochim. Biophys. Acta 1979, 557, 9–23. [Google Scholar] [CrossRef] [PubMed]
  77. Al-Azayzih, A.; Missaoui, W.N.; Cummings, B.S.; Somanath, P.R. Liposome-mediated delivery of the p21 activated kinase-1 (PAK-1) inhibitor IPA-3 limits prostate tumor growth in vivo. Nanomedicine 2016, 12, 1231–1239. [Google Scholar] [CrossRef] [Green Version]
  78. Najahi-Missaoui, W.; Quach, N.D.; Somanath, P.R.; Cummings, B.S. Liposomes Targeting P21 Activated Kinase-1 (PAK-1) and Selective for Secretory Phospholipase A(2) (sPLA(2)) Decrease Cell Viability and Induce Apoptosis in Metastatic Triple-Negative Breast Cancer Cells. Int. J. Mol. Sci. 2020, 21, 9396. [Google Scholar] [CrossRef] [PubMed]
  79. Mendez, R.; Banerjee, S. Sonication-Based Basic Protocol for Liposome Synthesis. Methods Mol. Biol. 2017, 1609, 255–260. [Google Scholar] [PubMed]
  80. Hope, M.J.; Bally, M.B.; Webb, G.; Cullis, P.R. Production of large unilamellar vesicles by a rapid extrusion procedure: Characterization of size distribution, trapped volume and ability to maintain a membrane potential. Biochim. Biophys. Acta 1985, 812, 55–65. [Google Scholar] [CrossRef] [PubMed]
  81. Haran, G.; Cohen, R.; Bar, L.K.; Barenholz, Y. Transmembrane ammonium sulfate gradients in liposomes produce efficient and stable entrapment of amphipathic weak bases. Biochim. Biophys. Acta 1993, 1151, 201–215. [Google Scholar] [CrossRef]
  82. Fritze, A.; Hens, F.; Kimpfler, A.; Schubert, R.; Peschka-Süss, R. Remote loading of doxorubicin into liposomes driven by a transmembrane phosphate gradient. Biochim. Biophys. Acta 2006, 1758, 1633–1640. [Google Scholar] [CrossRef] [Green Version]
  83. Akbarzadeh, A.; Rezaei-Sadabady, R.; Davaran, S.; Joo, S.W.; Zarghami, N.; Hanifehpour, Y.; Samiei, M.; Kouhi, M.; Nejati-Koshki, K. Liposome: Classification, preparation, and applications. Nanoscale Res. Lett. 2013, 8, 102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Szoka, F., Jr.; Papahadjopoulos, D. Procedure for preparation of liposomes with large internal aqueous space and high capture by reverse-phase evaporation. Proc. Natl. Acad. Sci. USA 1978, 75, 4194–4198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Corace, G.; Angeloni, C.; Malaguti, M.; Hrelia, S.; Stein, P.C.; Brandl, M.; Gotti, R.; Luppi, B. Multifunctional liposomes for nasal delivery of the anti-Alzheimer drug tacrine hydrochloride. J. Liposome Res. 2014, 24, 323–335. [Google Scholar] [CrossRef]
  86. Zhu, X.; Xie, Y.; Zhang, Y.; Huang, H.; Huang, S.; Hou, L.; Zhang, H.; Li, Z.; Shi, J.; Zhang, Z. Thermo-sensitive liposomes loaded with doxorubicin and lysine modified single-walled carbon nanotubes as tumor-targeting drug delivery system. J. Biomater. Appl. 2014, 29, 769–779. [Google Scholar] [CrossRef] [PubMed]
  87. Fichtner, I.; Reszka, R.; Goan, S.R.; Naundorf, H. Carboplatin-liposomes (CPL) in immunodeficient mice: Improved antitumor activity for breast carcinomas and stimulation of hematopoiesis. Med. Oncol. 1994, 11, 111–119. [Google Scholar] [CrossRef] [PubMed]
  88. Wang, Y.; Grainger, D.W. Lyophilized liposome-based parenteral drug development: Reviewing complex product design strategies and current regulatory environments. Adv. Drug Deliv. Rev. 2019, 151–152, 56–71. [Google Scholar] [CrossRef]
  89. Franzé, S.; Selmin, F.; Samaritani, E.; Minghetti, P.; Cilurzo, F. Lyophilization of Liposomal Formulations: Still Necessary, Still Challenging. Pharmaceutics 2018, 10, 139. [Google Scholar] [CrossRef] [Green Version]
  90. Seltzer, S.E.; Gregoriadis, G.; Dick, R. Evaluation of the dehydration-rehydration method for production of contrast-carrying liposomes. Investig. Radiol. 1988, 23, 131–138. [Google Scholar] [CrossRef]
  91. Monnard, P.A.; Oberholzer, T.; Luisi, P. Entrapment of nucleic acids in liposomes. Biochim. Biophys. Acta 1997, 1329, 39–50. [Google Scholar] [CrossRef] [Green Version]
  92. Osouli-Bostanabad, K.; Puliga, S.; Serrano, D.R.; Bucchi, A.; Halbert, G.; Lalatsa, A. Microfluidic Manufacture of Lipid-Based Nanomedicines. Pharmaceutics 2022, 14, 1940. [Google Scholar] [CrossRef]
  93. Shah, V.M.; Nguyen, D.X.; Patel, P.; Cote, B.; Al-Fatease, A.; Pham, Y.; Huynh, M.G.; Woo, Y.; Alani, A.W. Liposomes produced by microfluidics and extrusion: A comparison for scale-up purposes. Nanomedicine 2019, 18, 146–156. [Google Scholar] [CrossRef]
  94. Mijajlovic, M.; Wright, D.; Zivkovic, V.; Bi, J.X.; Biggs, M.J. Microfluidic hydrodynamic focusing based synthesis of POPC liposomes for model biological systems. Colloids Surf. B Biointerfaces 2013, 104, 276–281. [Google Scholar] [CrossRef] [Green Version]
  95. Saveyn, H.; De Baets, B.; Thas, O.; Hole, P.; Smith, J.; Van der Meeren, P. Accurate particle size distribution determination by nanoparticle tracking analysis based on 2-D Brownian dynamics simulation. J. Colloid Interface Sci. 2010, 352, 593–600. [Google Scholar] [CrossRef] [PubMed]
  96. Glatzel, T.; Holscher, H.; Schimmel, T.; Baykara, M.Z.; Schwarz, U.D.; Garcia, R. Advanced atomic force microscopy techniques. Beilstein J. Nanotechnol. 2012, 3, 893–894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Haiss, W.; Thanh, N.T.; Aveyard, J.; Fernig, D.G. Determination of size and concentration of gold nanoparticles from UV-vis spectra. Anal. Chem. 2007, 79, 4215–4221. [Google Scholar] [CrossRef] [PubMed]
  98. Planken, K.L.; Colfen, H. Analytical ultracentrifugation of colloids. Nanoscale 2010, 2, 1849–1869. [Google Scholar] [CrossRef] [PubMed]
  99. Deamer, D.W.; Prince, R.C.; Crofts, A.R. The response of fluorescent amines to pH gradients across liposome membranes. Biochim. Biophys. Acta 1972, 274, 323–335. [Google Scholar] [CrossRef] [PubMed]
  100. Zucker, D.; Marcus, D.; Barenholz, Y.; Goldblum, A. Liposome drugs’ loading efficiency: A working model based on loading conditions and drug’s physicochemical properties. J. Control. Release 2009, 139, 73–80. [Google Scholar] [CrossRef]
  101. Li, X.; Hirsh, D.J.; Cabral-Lilly, D.; Zirkel, A.; Gruner, S.M.; Janoff, A.S.; Perkins, W.R. Doxorubicin physical state in solution and inside liposomes loaded via a pH gradient. Biochim. Biophys. Acta 1998, 1415, 23–40. [Google Scholar] [CrossRef] [Green Version]
  102. Barenholz, Y. Doxil(R)—The first FDA-approved nano-drug: Lessons learned. J. Control. Release 2012, 160, 117–134. [Google Scholar] [CrossRef]
  103. Safra, T.; Muggia, F.; Jeffers, S.; Tsao-Wei, D.D.; Groshen, S.; Lyass, O.; Henderson, R.; Berry, G.; Gabizon, A. Pegylated liposomal doxorubicin (doxil): Reduced clinical cardiotoxicity in patients reaching or exceeding cumulative doses of 500 mg/m2. Annal. Oncol. 2000, 11, 1029–1033. [Google Scholar] [CrossRef]
  104. Cancer Medications Enquiry Database (CanMED). Surveillance Research Program SEER Website Tool. National Drug Code, Version 1.14.0; Division of Cancer Control and Population Sciences, National Cancer Institute: Bethesda, MA, USA, 2023.
  105. HRPA. DAUNOXOME 2 mg/mL, Liposomal Dispersion for Injection; HRPA: Toronto, ON, Canada, 2014; pp. 1–14. [Google Scholar]
  106. BioSpace. Clinigen And Galen Enter Exclusive Global Access Agreement for Chemotherapy Drug Daunoxome; BioSpace: Des Moines, IA, USA, 2016. [Google Scholar]
  107. Van Hoogevest, P.; Wendel, A. The use of natural and synthetic phospholipids as pharmaceutical excipients. Eur. J. Lipid Sci. Technol. 2014, 116, 1088–1107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Forssen, E. The design and development of DaunoXome® for solid tumor targeting in vivo. Adv. Drug Deliv. Rev. 1997, 24, 133–150. [Google Scholar] [CrossRef]
  109. European Medicines Agency. DepoCyt; European Medicines Agency: Amsterdam, The Netherlands, 2017.
  110. U.S. Food and Drug Administration. Depocyt (Cytarabine); U.S. Food and Drug Administration: Silver Spring, MD, USA, 2017.
  111. WCG FDANews. Pacira Shutters DepCyt Operations after Years of Manufacturing Problems; WCG FDANews: Falls Church, VA, USA, 2017. [Google Scholar]
  112. U.S. Food & Drug Administration. Doxil (Liposomal) [Doxorubicin Hydrochloride] Label; U.S. Food & Drug Administration: Silver Spring, MD, USA, 2022.
  113. European Medicines Agency. Caelyx Pegylated Liposomal; European Medicines Agency: Amsterdam, The Netherlands, 2023.
  114. Taiwan Liposome Company, Ltd. Form F-1 Registration Statement; Taiwan Liposome Company, Ltd.: Taipei, Taiwan, 2018. [Google Scholar]
  115. Hsieh, Y.J.; Chang, C.H.; Huang, S.P.; Lin, C.W.; Wang, M.N.; Wu, Y.T.; Chen, Y.J.; Tsai, T.H. Effect of cyclosporin A on the brain regional distribution of doxorubicin in rats. Int. J. Pharm. 2008, 350, 265–271. [Google Scholar] [CrossRef]
  116. Hospira Australia Pty Ltd. Marqibo® (vinCRIStine Sulfate LIPOSOME Injection) for Intravenous Infusion [Package Insert]; U.S. Food and Drug Administration: Silver Spring, MD, USA, 2022.
  117. Mifamurtide: CGP 19835, CGP 19835A, L-MTP-PE, liposomal MTP-PE, MLV 19835A, MTP-PE, muramyltripeptide phosphatidylethanolamine. Drugs R D 2008, 9, 131–135. [CrossRef] [PubMed]
  118. Vail, D.M.; MacEwen, E.G.; Kurzman, I.D.; Dubielzig, R.R.; Helfand, S.C.; Kisseberth, W.C.; London, C.A.; Obradovich, J.E.; Madewell, B.R.; Rodriguez, C.O. Liposome-encapsulated muramyl tripeptide phosphatidylethanolamine adjuvant immunotherapy for splenic hemangiosarcoma in the dog: A randomized multi-institutional clinical trial. Clin. Cancer Res. 1995, 1, 1165–1170. [Google Scholar]
  119. European Medicines Agency. Mepact (mifamurtide); European Medicines Agency: Amsterdam, The Netherlands, 2020.
  120. Batist, G.; Ramakrishnan, G.; Rao, C.S.; Chandrasekharan, A.; Gutheil, J.; Guthrie, T.; Shah, P.; Khojasteh, A.; Nair, M.K.; Hoelzer, K.; et al. Reduced cardiotoxicity and preserved antitumor efficacy of liposome-encapsulated doxorubicin and cyclophosphamide compared with conventional doxorubicin and cyclophosphamide in a randomized, multicenter trial of metastatic breast cancer. J. Clin. Oncol. 2001, 19, 1444–1454. [Google Scholar] [CrossRef]
  121. European Medicines Agency. Myocet Liposomal (Previously Myocet); European Medicines Agency: Amsterdam, The Netherlands, 2021.
  122. Swenson, C.E.; Perkins, W.R.; Roberts, P.; Janoff, A.S. Liposome technology and the development of Myocet™ (liposomal doxorubicin citrate). Breast 2001, 10, 1–7. [Google Scholar] [CrossRef]
  123. U.S. Food and Drug Administration. Onivyde (Irinotecan Hydrochloride); U.S. Food and Drug Administration: Silver Spring, MD, USA, 2023.
  124. European Medicines Agency. Onivyde Pegylated Liposomal (Previously Known as Onivyde); European Medicines Agency: Amsterdam, The Netherlands, 2022.
  125. Kalra, A.V.; Kim, J.; Klinz, S.G.; Paz, N.; Cain, J.; Drummond, D.C.; Nielsen, U.B.; Fitzgerald, J.B. Preclinical activity of nanoliposomal irinotecan is governed by tumor deposition and intratumor prodrug conversion. Cancer Res. 2014, 74, 7003–7013. [Google Scholar] [CrossRef] [Green Version]
  126. U.S. Food and Drug Administration. Vyxeos; U.S. Food and Drug Administration: Silver Spring, MD, USA, 2023.
  127. European Medicines Agency. Vyxeos Liposomal (Previously Known as Vyxeos); European Medicines Agency: Amsterdam, The Netherlands, 2022.
  128. European Medicines Agency. Zolsketil Pegylated Liposomal; European Medicines Agency: Amsterdam, The Netherlands, 2022.
  129. Dragovich, T.; Mendelson, D.; Kurtin, S.; Richardson, K.; Von Hoff, D.; Hoos, A. A Phase 2 trial of the liposomal DACH platinum L-NDDP in patients with therapy-refractory advanced colorectal cancer. Cancer Chemother. Pharmacol. 2006, 58, 759–764. [Google Scholar] [CrossRef]
  130. Gutiérrez-Puente, Y.; Tari, A.M.; Stephens, C.; Rosenblum, M.; Guerra, R.T.; Lopez-Berestein, G. Safety, pharmacokinetics, and tissue distribution of liposomal P-ethoxy antisense oligonucleotides targeted to Bcl-2. J. Pharmacol. Exp. Ther. 1999, 291, 865–869. [Google Scholar]
  131. Fasol, U.; Frost, A.; Buchert, M.; Arends, J.; Fiedler, U.; Scharr, D.; Scheuenpflug, J.; Mross, K. Vascular and pharmacokinetic effects of EndoTAG-1 in patients with advanced cancer and liver metastasis. Annal. Oncol. 2012, 23, 1030–1036. [Google Scholar] [CrossRef] [PubMed]
  132. Van, P.; Tiemessen, H.; Metselaar, J.M.; Drescher, S.; Fahr, A. The Use of Phospholipids to Make Pharmaceutical Form Line Extensions. Eur. J. Lipid Sci. Technol. 2021, 123, 2000297. [Google Scholar]
  133. Li, C.; Wang, J.; Wang, C.; Li, Y.; Shen, D.; Guo, W.; Li, J.; Zhang, L. Liposomal Pharmaceutical Preparation and Method for Manufacturing the Same. U.S. 10,028,913B2, 24 July 2018. [Google Scholar]
  134. Gao, Y.; Huang, H.; Wang, X.; Bai, B.; Huang, Y.; Yang, H.; Zhang, Q.; Li, Y.; Li, Y.; Zhou, M.; et al. Safety and Efficacy of Mitoxantrone Hydrochloride Liposome in Patients with Relapsed or Refractory Peripheral T-Cell Lymphoma and Extranodal NK/T-Cell Lymphoma: A Prospective, Single-Arm, Open-Label, Multi-Center, Phase II Clinical Trial. Blood 2020, 136, 36–37. [Google Scholar] [CrossRef]
  135. Yarmolenko, P.S.; Zhao, Y.; Landon, C.; Spasojevic, I.; Yuan, F.; Needham, D.; Viglianti, B.L.; Dewhirst, M.W. Comparative effects of thermosensitive doxorubicin-containing liposomes and hyperthermia in human and murine tumours. Int. J. Hyperth. 2010, 26, 485–498. [Google Scholar] [CrossRef]
  136. Lyon, P.C.; Griffiths, L.F.; Lee, J.; Chung, D.; Carlisle, R.; Wu, F.; Middleton, M.R.; Gleeson, F.V.; Coussios, C.C. Clinical trial protocol for TARDOX: A phase I study to investigate the feasibility of targeted release of lyso-thermosensitive liposomal doxorubicin (ThermoDox(R)) using focused ultrasound in patients with liver tumours. J. Ther. Ultrasound 2017, 5, 28. [Google Scholar] [CrossRef] [Green Version]
  137. Petersen, M.J.; Melander, F.; Vikbjerg, A.F.; Petersen, S.A.; Madsen, M.W. Medical Use of sPLA2 Hydrolysable Liposomes. U.S. Patent 11,207,269B2, 28 December 2021. [Google Scholar]
  138. Jehn, C.F.; Boulikas, T.; Kourvetaris, A.; Possinger, K.; Lüftner, D. Pharmacokinetics of liposomal cisplatin (lipoplatin) in combination with 5-FU in patients with advanced head and neck cancer: First results of a phase III study. AntiCancer Res. 2007, 27, 471–475. [Google Scholar] [PubMed]
  139. Newman, M.S.; Colbern, G.T.; Working, P.K.; Engbers, C.; Amantea, M.A. Comparative pharmacokinetics, tissue distribution, and therapeutic effectiveness of cisplatin encapsulated in long-circulating, pegylated liposomes (SPI-077) in tumor-bearing mice. Cancer Chemother. Pharmacol. 1999, 43, 1–7. [Google Scholar] [CrossRef]
  140. Seetharamu, N.; Kim, E.; Hochster, H.; Martin, F.; Muggia, F. Phase II study of liposomal cisplatin (SPI-77) in platinum-sensitive recurrences of ovarian cancer. AntiCancer Res. 2010, 30, 541–545. [Google Scholar]
  141. Ogihara, I.; Kojima, S.; Jay, M. Tumor uptake of 67Ga-carrying liposomes. Eur. J. Nucl. Med. 1986, 11, 405–411. [Google Scholar] [CrossRef]
  142. Zhao, W.; Zhuang, S.; Qi, X.R. Comparative study of the in vitro and in vivo characteristics of cationic and neutral liposomes. Int. J. Nanomed. 2011, 6, 3087–3098. [Google Scholar]
  143. Krasnici, S.; Werner, A.; Eichhorn, M.E.; Schmitt-Sody, M.; Pahernik, S.A.; Sauer, B.; Schulze, B.; Teifel, M.; Michaelis, U.; Naujoks, K.; et al. Effect of the surface charge of liposomes on their uptake by angiogenic tumor vessels. Int. J. Cancer 2003, 105, 561–567. [Google Scholar] [CrossRef] [PubMed]
  144. Maruyama, K. Intracellular targeting delivery of liposomal drugs to solid tumors based on EPR effects. Adv. Drug Deliv. Rev. 2011, 63, 161–169. [Google Scholar] [CrossRef] [PubMed]
  145. Fang, J.; Nakamura, H.; Maeda, H. The EPR effect: Unique features of tumor blood vessels for drug delivery, factors involved, and limitations and augmentation of the effect. Adv. Drug Deliv. Rev. 2011, 63, 136–151. [Google Scholar] [CrossRef]
  146. Deshpande, P.P.; Biswas, S.; Torchilin, V.P. Current trends in the use of liposomes for tumor targeting. Nanomedicine 2013, 8, 1509–1528. [Google Scholar] [CrossRef] [Green Version]
  147. Haley, B.; Frenkel, E. Nanoparticles for drug delivery in cancer treatment. Urol. Oncol. 2008, 26, 57–64. [Google Scholar] [CrossRef]
  148. Maeda, H. Vascular permeability in cancer and infection as related to macromolecular drug delivery, with emphasis on the EPR effect for tumor-selective drug targeting. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2012, 88, 53–71. [Google Scholar] [CrossRef] [Green Version]
  149. Maeda, H. The enhanced permeability and retention (EPR) effect in tumor vasculature: The key role of tumor-selective macromolecular drug targeting. Adv. Enzym. Regul. 2001, 41, 189–207. [Google Scholar] [CrossRef]
  150. Maeda, H.; Nakamura, H.; Fang, J. The EPR effect for macromolecular drug delivery to solid tumors: Improvement of tumor uptake, lowering of systemic toxicity, and distinct tumor imaging in vivo. Adv. Drug Deliv. Rev. 2013, 65, 71–79. [Google Scholar] [CrossRef]
  151. Torchilin, V. Tumor delivery of macromolecular drugs based on the EPR effect. Adv. Drug Deliv. Rev. 2011, 63, 131–135. [Google Scholar] [CrossRef]
  152. Brannon-Peppas, L.; Blanchette, J.O. Nanoparticle and targeted systems for cancer therapy. Adv. Drug Deliv. Rev. 2004, 56, 1649–1659. [Google Scholar] [CrossRef]
  153. Sawant, R.R.; Torchilin, V.P. Challenges in development of targeted liposomal therapeutics. AAPS J. 2012, 14, 303–315. [Google Scholar] [CrossRef] [Green Version]
  154. Madni, A.; Sarfraz, M.; Rehman, M.; Ahmad, M.; Akhtar, N.; Ahmad, S.; Tahir, N.; Ijaz, S.; Al-Kassas, R.; Lobenberg, R. Liposomal drug delivery: A versatile platform for challenging clinical applications. J. Pharm. Pharm. Sci. 2014, 17, 401–426. [Google Scholar] [CrossRef] [Green Version]
  155. Klibanov, A.L.; Maruyama, K.; Torchilin, V.P.; Huang, L. Amphipathic polyethyleneglycols effectively prolong the circulation time of liposomes. FEBS Lett. 1990, 268, 235–237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Allen, T.M. Long-circulating (sterically stabilized) liposomes for targeted drug delivery. Trends Pharmacol. Sci. 1994, 15, 215–220. [Google Scholar] [CrossRef] [PubMed]
  157. Siegal, T.; Horowitz, A.; Gabizon, A. Doxorubicin encapsulated in sterically stabilized liposomes for the treatment of a brain tumor model: Biodistribution and therapeutic efficacy. J. Neurosurg. 1995, 83, 1029–1037. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Working, P.K.; Newman, M.S.; Huang, K.S.; Mayhew, E.; Vaage, J.; Lasic, D.D. Pharmacokinetics, Biodistribution and Therapeutic Efficacy of Doxorubicin Encapsulated in Stealth® Liposomes (Doxil®). J. Liposome Res. 1994, 4, 667–687. [Google Scholar] [CrossRef]
  159. Phase Transition Temperatures for Glycerophospholipids. Available online: https://avantilipids.com/tech-support/physical-properties/phase-transition-temps (accessed on 26 February 2023).
  160. 18:0 PC (DSPC) 1,2-Distearoyl-sn-glycero-3-phosphocholine. Available online: https://avantilipids.com/product/850365 (accessed on 26 February 2023).
  161. Hydro Soy PC L-α-phosphatidylcholine, Hydrogenated (Soy). Available online: https://avantilipids.com/product/840058 (accessed on 26 February 2023).
  162. Hong, R.L.; Tseng, Y.L. Phase I and pharmacokinetic study of a stable, polyethylene-glycolated liposomal doxorubicin in patients with solid tumors: The relation between pharmacokinetic property and toxicity. Cancer 2001, 91, 1826–1833. [Google Scholar] [CrossRef]
  163. Gabizon, A.; Catane, R.; Uziely, B.; Kaufman, B.; Safra, T.; Cohen, R.; Martin, F.; Huang, A.; Barenholz, Y. Prolonged circulation time and enhanced accumulation in malignant exudates of doxorubicin encapsulated in polyethylene-glycol coated liposomes. Cancer Res. 1994, 54, 987–992. [Google Scholar]
  164. Chou, H.H.; Wang, K.L.; Chen, C.A.; Wei, L.H.; Lai, C.H.; Hsieh, C.Y.; Yang, Y.C.; Twu, N.F.; Chang, T.C.; Yen, M.S.; et al. 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. [Google Scholar] [CrossRef]
  165. Hsiao, S.M.; Chen, C.A.; Lin, H.H.; Hsieh, C.Y.; Wei, L.H. Phase II trial of carboplatin and distearoylphosphatidylcholine pegylated liposomal doxorubicin (Lipo-Dox) in recurrent platinum-sensitive ovarian cancer following front-line therapy with paclitaxel and platinum. Gynecol. Oncol. 2009, 112, 35–39. [Google Scholar] [CrossRef]
  166. Yao, N.; Kao, W.; Chao, T.; Hsieh, R.; Lin, J.; Su, C.; Lo, S. A phase II study of gemcitabine and liposomal doxorubicin (Lipo-Dox) as first line chemotherapy in the treatment of metastatic breast cancer. J. Clin. Oncol. 2006, 24, 10688. [Google Scholar] [CrossRef]
  167. White, S.C.; Lorigan, P.; Margison, G.P.; Margison, J.M.; Martin, F.; Thatcher, N.; Anderson, H.; Ranson, M. Phase II study of SPI-77 (sterically stabilised liposomal cisplatin) in advanced non-small-cell lung cancer. Br. J. Cancer 2006, 95, 822–828. [Google Scholar] [CrossRef] [Green Version]
  168. Kim, E.S.; Lu, C.; Khuri, F.R.; Tonda, M.; Glisson, B.S.; Liu, D.; Jung, M.; Hong, W.K.; Herbst, R.S. A phase II study of STEALTH cisplatin (SPI-77) in patients with advanced non-small cell lung cancer. Lung Cancer 2001, 34, 427–432. [Google Scholar] [CrossRef] [PubMed]
  169. Immordino, M.L.; Dosio, F.; Cattel, L. Stealth liposomes: Review of the basic science, rationale, and clinical applications, existing and potential. Int. J. Nanomed. 2006, 1, 297–315. [Google Scholar]
  170. Kroll, A.; Pillukat, M.H.; Hahn, D.; Schnekenburger, J. Current in vitro methods in nanoparticle risk assessment: Limitations and challenges. Eur. J. Pharm. Biopharm. 2009, 72, 370–377. [Google Scholar] [CrossRef] [PubMed]
  171. Raj, R.; Mongia, P.; Sahu, S.K.; Ram, A. Nanocarriers based anticancer drugs: Current scenario and future perceptions. Curr. Drug Targets 2015, 17, 206–228. [Google Scholar] [CrossRef]
  172. Perez-Herrero, E.; Fernandez-Medarde, A. Advanced targeted therapies in cancer: Drug nanocarriers, the future of chemotherapy. Eur. J. Pharm. Biopharm. 2015, 93, 52–79. [Google Scholar] [CrossRef] [Green Version]
  173. Shi, C.; Cao, H.; He, W.; Gao, F.; Liu, Y.; Yin, L. Novel drug delivery liposomes targeted with a fully human anti-VEGF165 monoclonal antibody show superior antitumor efficacy in vivo. Biomed. Pharmacother. 2015, 73, 48–57. [Google Scholar] [CrossRef]
  174. Gao, J.; Chen, H.; Song, H.; Su, X.; Niu, F.; Li, W.; Li, B.; Dai, J.; Wang, H.; Guo, Y. Antibody-targeted immunoliposomes for cancer treatment. Mini Rev. Med. Chem. 2013, 13, 2026–2035. [Google Scholar] [CrossRef]
  175. Fegan, A.; Kumarapperuma, S.C.; Wagner, C.R. Chemically Self-Assembled Antibody Nanostructures as Potential Drug Carriers. Mol. Pharm. 2012, 9, 3218–3227. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Torchilin, V.P. Targeted pharmaceutical nanocarriers for cancer therapy and imaging. AAPS J. 2007, 9, E128–E147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Danhier, F.; Feron, O.; Preat, V. To exploit the tumor microenvironment: Passive and active tumor targeting of nanocarriers for anti-cancer drug delivery. J. Control. Release 2010, 148, 135–146. [Google Scholar] [CrossRef] [PubMed]
  178. Noble, G.T.; Stefanick, J.F.; Ashley, J.D.; Kiziltepe, T.; Bilgicer, B. Ligand-targeted liposome design: Challenges and fundamental considerations. Trends Biotechnol. 2014, 32, 32–45. [Google Scholar] [CrossRef] [PubMed]
  179. Eloy, J.O.; Petrilli, R.; Trevizan, L.N.F.; Chorilli, M. Immunoliposomes: A review on functionalization strategies and targets for drug delivery. Colloids Surf. B Biointerfaces 2017, 159, 454–467. [Google Scholar] [CrossRef] [Green Version]
  180. Wang, L.; Zhao, C.; Lu, L.; Jiang, H.; Wang, F.; Zhang, X. Transcytosable Peptide-Paclitaxel Prodrug Nanoparticle for Targeted Treatment of Triple-Negative Breast Cancer. Int. J. Mol. Sci. 2023, 24, 4646. [Google Scholar] [CrossRef]
  181. Li, J.; Kataoka, K. Chemo-physical Strategies to Advance the in Vivo Functionality of Targeted Nanomedicine: The Next Generation. J. Am. Chem. Soc. 2021, 143, 538–559. [Google Scholar] [CrossRef]
  182. Hussain, S.; Pluckthun, A.; Allen, T.M.; Zangemeister-Wittke, U. Antitumor activity of an epithelial cell adhesion molecule targeted nanovesicular drug delivery system. Mol. Cancer Ther. 2007, 6, 3019–3027. [Google Scholar] [CrossRef] [Green Version]
  183. Koshkaryev, A.; Sawant, R.; Deshpande, M.; Torchilin, V. Immunoconjugates and long circulating systems: Origins, current state of the art and future directions. Adv. Drug Deliv. Rev. 2013, 65, 24–35. [Google Scholar] [CrossRef] [Green Version]
  184. Park, J.W.; Hong, K.; Kirpotin, D.B.; Meyer, O.; Papahadjopoulos, D.; Benz, C.C. Anti-HER2 immunoliposomes for targeted therapy of human tumors. Cancer Lett. 1997, 118, 153–160. [Google Scholar] [CrossRef]
  185. Park, J.W.; Kirpotin, D.B.; Hong, K.; Shalaby, R.; Shao, Y.; Nielsen, U.B.; Marks, J.D.; Papahadjopoulos, D.; Benz, C.C. Tumor targeting using anti-her2 immunoliposomes. J. Control. Release 2001, 74, 95–113. [Google Scholar] [CrossRef] [PubMed]
  186. Elbayoumi, T.A.; Torchilin, V.P. Enhanced accumulation of long-circulating liposomes modified with the nucleosome-specific monoclonal antibody 2C5 in various tumours in mice: Gamma-imaging studies. Eur. J. Nucl. Med. Mol. Imaging 2006, 33, 1196–1205. [Google Scholar] [CrossRef] [PubMed]
  187. ElBayoumi, T.A.; Torchilin, V.P. Tumor-targeted nanomedicines: Enhanced antitumor efficacy in vivo of doxorubicin-loaded, long-circulating liposomes modified with cancer-specific monoclonal antibody. Clin. Cancer Res. 2009, 15, 1973–1980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Elbayoumi, T.A.; Torchilin, V.P. Enhanced cytotoxicity of monoclonal anticancer antibody 2C5-modified doxorubicin-loaded PEGylated liposomes against various tumor cell lines. Eur. J. Pharm. Sci. 2007, 32, 159–168. [Google Scholar] [CrossRef] [Green Version]
  189. Allen, T.M.; Mumbengegwi, D.R.; Charrois, G.J. Anti-CD19-targeted liposomal doxorubicin improves the therapeutic efficacy in murine B-cell lymphoma and ameliorates the toxicity of liposomes with varying drug release rates. Clin. Cancer Res. 2005, 11, 3567–3573. [Google Scholar] [CrossRef] [Green Version]
  190. Cheng, W.W.; Allen, T.M. Targeted delivery of anti-CD19 liposomal doxorubicin in B-cell lymphoma: A comparison of whole monoclonal antibody, Fab’ fragments and single chain Fv. J. Control. Release 2008, 126, 50–58. [Google Scholar] [CrossRef]
  191. Arias, J.L. Drug targeting strategies in cancer treatment: An overview. Mini Rev. Med. Chem. 2011, 11, 1–17. [Google Scholar] [CrossRef]
  192. Ta, T.; Porter, T.M. Thermosensitive liposomes for localized delivery and triggered release of chemotherapy. J. Control. Release 2013, 169, 112–125. [Google Scholar] [CrossRef] [Green Version]
  193. Kneidl, B.; Peller, M.; Winter, G.; Lindner, L.H.; Hossann, M. Thermosensitive liposomal drug delivery systems: State of the art review. Int. J. Nanomed. 2014, 9, 4387–4398. [Google Scholar]
  194. May, J.P.; Li, S.D. Hyperthermia-induced drug targeting. Expert Opin. Drug Deliv. 2013, 10, 511–527. [Google Scholar] [CrossRef]
  195. Grull, H.; Langereis, S. Hyperthermia-triggered drug delivery from temperature-sensitive liposomes using MRI-guided high intensity focused ultrasound. J. Control. Release 2012, 161, 317–327. [Google Scholar] [CrossRef] [PubMed]
  196. Bulbake, U.; Doppalapudi, S.; Kommineni, N.; Khan, W. Liposomal Formulations in Clinical Use: An Updated Review. Pharmaceutics 2017, 9, 12. [Google Scholar] [CrossRef] [PubMed]
  197. Fouladi, F.; Steffen, K.J.; Mallik, S. Enzyme-Responsive Liposomes for the Delivery of Anticancer Drugs. Bioconjug. Chem. 2017, 28, 857–868. [Google Scholar] [CrossRef] [Green Version]
  198. Mock, J.N.; Costyn, L.J.; Wilding, S.L.; Arnold, R.D.; Cummings, B.S. Evidence for distinct mechanisms of uptake and antitumor activity of secretory phospholipase A2 responsive liposome in prostate cancer. Integr. Biol. 2013, 5, 172–182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Woodle, M.C.; Matthay, K.K.; Newman, M.S.; Hidayat, J.E.; Collins, L.R.; Redemann, C.; Martin, F.J.; Papahadjopoulos, D. Versatility in lipid compositions showing prolonged circulation with sterically stabilized liposomes. Biochim. Biophys. Acta 1992, 1105, 193–200. [Google Scholar] [CrossRef]
  200. Du, H.; Chandaroy, P.; Hui, S.W. Grafted poly-(ethylene glycol) on lipid surfaces inhibits protein adsorption and cell adhesion. Biochim. Biophys. Acta 1997, 1326, 236–248. [Google Scholar] [CrossRef] [Green Version]
  201. Gref, R.; Lück, M.; Quellec, P.; Marchand, M.; Dellacherie, E.; Harnisch, S.; Blunk, T.; Müller, R.H. ‘Stealth’ corona-core nanoparticles surface modified by polyethylene glycol (PEG): Influences of the corona (PEG chain length and surface density) and of the core composition on phagocytic uptake and plasma protein adsorption. Colloids Surf. B Biointerfaces 2000, 18, 301–313. [Google Scholar] [CrossRef] [PubMed]
  202. Dams, E.; Laverman, P.; Oyen, W.; Storm, G.; Scherphof, G.; Van der Meer, J.; Corstens, F.; Boerman, O. Accelerated blood clearance and altered biodistribution of repeated injections of sterically stabilized liposomes. J. Pharm. Exp. Ther. 2000, 292, 1071–1079. [Google Scholar]
  203. Li, C.; Cao, J.; Wang, Y.; Zhao, X.; Deng, C.; Wei, N.; Yang, J.; Cui, J. Accelerated blood clearance of pegylated liposomal topotecan: Influence of polyethylene glycol grafting density and animal species. J. Pharm. Sci. 2012, 101, 3864–3876. [Google Scholar] [CrossRef]
  204. Xu, H.; Ye, F.; Hu, M.; Yin, P.; Zhang, W.; Li, Y.; Yu, X.; Deng, Y. Influence of phospholipid types and animal models on the accelerated blood clearance phenomenon of PEGylated liposomes upon repeated injection. Drug Deliv. 2015, 22, 598–607. [Google Scholar] [CrossRef]
  205. Ishida, T.; Ichihara, M.; Wang, X.; Yamamoto, K.; Kimura, J.; Majima, E.; Kiwada, H. Injection of PEGylated liposomes in rats elicits PEG-specific IgM, which is responsible for rapid elimination of a second dose of PEGylated liposomes. J. Control. Release 2006, 112, 15–25. [Google Scholar] [CrossRef]
  206. Verhoef, J.J.F.; Carpenter, J.F.; Anchordoquy, T.J.; Schellekens, H. Potential induction of anti-PEG antibodies and complement activation toward PEGylated therapeutics. Drug Discov. Today 2014, 19, 1945–1952. [Google Scholar] [CrossRef] [PubMed]
  207. Li, Z.; Gao, X.; Yan, X.; Deng, Y.; Ma, H. PEGylated nanoemulsions containing 1,2-distearoyl-sn-glycero-3-phosphoglycerol induced weakened accelerated blood clearance phenomenon. Drug Deliv. Transl. Res. 2022, 12, 2569–2579. [Google Scholar] [CrossRef] [PubMed]
  208. Chen, D.; Liu, W.; Shen, Y.; Mu, H.; Zhang, Y.; Liang, R.; Wang, A.; Sun, K.; Fu, F. Effects of a novel pH-sensitive liposome with cleavable esterase-catalyzed and pH-responsive double smart mPEG lipid derivative on ABC phenomenon. Int. J. Nanomed. 2011, 6, 2053–2061. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Xu, H.; Wang, K.; Deng, Y.; Chen, D. Effects of cleavable PEG-cholesterol derivatives on the accelerated blood clearance of PEGylated liposomes. Biomaterials 2010, 31, 4757–4763. [Google Scholar] [CrossRef] [PubMed]
  210. Han, X.; Zhang, T.; Liu, M.Y.; Song, Y.Z.; Liu, X.R.; Deng, Y.H. Polysialic Acid Modified Liposomes for Improving Pharmacokinetics and Overcoming Accelerated Blood Clearance Phenomenon. Coatings 2020, 10, 834. [Google Scholar] [CrossRef]
  211. Sadzuka, Y.; Kishi, K.; Hirota, S.; Sonobe, T. Effect of polyethyleneglycol (PEG) chain on cell uptake of PEG-modified liposomes. J. Liposome Res. 2003, 13, 157–172. [Google Scholar] [CrossRef]
  212. Song, L.Y.; Ahkong, Q.F.; Rong, Q.; Wang, Z.; Ansell, S.; Hope, M.J.; Mui, B. Characterization of the inhibitory effect of PEG-lipid conjugates on the intracellular delivery of plasmid and antisense DNA mediated by cationic lipid liposomes. Biochim. Biophys. Acta 2002, 1558, 1–13. [Google Scholar] [CrossRef] [Green Version]
  213. Ghaferi, M.; Raza, A.; Koohi, M.; Zahra, W.; Akbarzadeh, A.; Ebrahimi Shahmabadi, H.; Alavi, S.E. Impact of PEGylated Liposomal Doxorubicin and Carboplatin Combination on Glioblastoma. Pharmaceutics 2022, 14, 2183. [Google Scholar] [CrossRef]
  214. Miller, C.R.; Bondurant, B.; McLean, S.D.; McGovern, K.A.; O’Brien, D.F. Liposome-cell interactions in vitro: Effect of liposome surface charge on the binding and endocytosis of conventional and sterically stabilized liposomes. Biochemistry 1998, 37, 12875–12883. [Google Scholar] [CrossRef]
  215. Brandenberger, C.; Mühlfeld, C.; Ali, Z.; Lenz, A.G.; Schmid, O.; Parak, W.J.; Gehr, P.; Rothen-Rutishauser, B. Quantitative evaluation of cellular uptake and trafficking of plain and polyethylene glycol-coated gold nanoparticles. Small 2010, 6, 1669–1678. [Google Scholar] [CrossRef]
  216. Zhang, Y.; Kohler, N.; Zhang, M. Surface modification of superparamagnetic magnetite nanoparticles and their intracellular uptake. Biomaterials 2002, 23, 1553–1561. [Google Scholar] [CrossRef]
  217. Pelaz, B.; del Pino, P.; Maffre, P.; Hartmann, R.; Gallego, M.; Rivera-Fernández, S.; de la Fuente, J.M.; Nienhaus, G.U.; Parak, W.J. Surface Functionalization of Nanoparticles with Polyethylene Glycol: Effects on Protein Adsorption and Cellular Uptake. ACS Nano 2015, 9, 6996–7008. [Google Scholar] [CrossRef] [PubMed]
  218. Mishra, S.; Webster, P.; Davis, M.E. PEGylation significantly affects cellular uptake and intracellular trafficking of non-viral gene delivery particles. Eur. J. Cell Biol. 2004, 83, 97–111. [Google Scholar] [CrossRef] [PubMed]
  219. Chan, C.L.; Majzoub, R.N.; Shirazi, R.S.; Ewert, K.K.; Chen, Y.J.; Liang, K.S.; Safinya, C.R. Endosomal escape and transfection efficiency of PEGylated cationic liposome-DNA complexes prepared with an acid-labile PEG-lipid. Biomaterials 2012, 33, 4928–4935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. 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, 527–542. [Google Scholar] [CrossRef]
  221. Parr, M.J.; Masin, D.; Cullis, P.R.; Bally, M.B. Accumulation of liposomal lipid and encapsulated doxorubicin in murine Lewis lung carcinoma: The lack of beneficial effects by coating liposomes with poly(ethylene glycol). J. Pharmacol. Exp. Ther. 1997, 280, 1319–1327. [Google Scholar]
  222. Zalba, S.; ten Hagen, T.L.M.; Burgui, C.; Garrido, M.J. Stealth nanoparticles in oncology: Facing the PEG dilemma. J. Controll. Release 2022, 351, 22–36. [Google Scholar] [CrossRef]
  223. Fang, Y.; Xue, J.; Gao, S.; Lu, A.; Yang, D.; Jiang, H.; He, Y.; Shi, K. Cleavable PEGylation: A strategy for overcoming the “PEG dilemma” in efficient drug delivery. Drug Deliv. 2017, 24 (Suppl. S1), 22–32. [Google Scholar] [CrossRef] [Green Version]
  224. Holland, J.W.; Hui, C.; Cullis, P.R.; Madden, T.D. Poly(ethylene glycol)—Lipid conjugates regulate the calcium-induced fusion of liposomes composed of phosphatidylethanolamine and phosphatidylserine. Biochemistry 1996, 35, 2618–2624. [Google Scholar] [CrossRef]
  225. Hong, R.L.; Huang, C.J.; Tseng, Y.L.; Pang, V.F.; Chen, S.T.; Liu, J.J.; Chang, F.H. Direct comparison of liposomal doxorubicin with or without polyethylene glycol coating in C-26 tumor-bearing mice: Is surface coating with polyethylene glycol beneficial? Clin. Cancer Res. 1999, 5, 3645–3652. [Google Scholar]
  226. Zhao, G.; Long, L.; Zhang, L.; Peng, M.; Cui, T.; Wen, X.; Zhou, X.; Sun, L.; Che, L. Smart pH-sensitive nanoassemblies with cleavable PEGylation for tumor targeted drug delivery. Sci. Rep. 2017, 7, 3383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Shin, J.; Shum, P.; Thompson, D.H. Acid-triggered release via dePEGylation of DOPE liposomes containing acid-labile vinyl ether PEG-lipids. J. Control. Release 2003, 91, 187–200. [Google Scholar] [CrossRef] [PubMed]
  228. Griffiths, J.R. Are cancer cells acidic? Br. J. Cancer 1991, 64, 425–427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Vaupel, P.; Kallinowski, F.; Okunieff, P. Blood flow, oxygen and nutrient supply, and metabolic microenvironment of human tumors: A review. Cancer Res. 1989, 49, 6449–6465. [Google Scholar]
  230. Hu, Y.B.; Dammer, E.B.; Ren, R.J.; Wang, G. The endosomal-lysosomal system: From acidification and cargo sorting to neurodegeneration. Transl. Neurodegener. 2015, 4, 18. [Google Scholar] [CrossRef] [Green Version]
  231. Kanamala, M.; Palmer, B.D.; Jamieson, S.M.; Wilson, W.R.; Wu, Z. Dual pH-sensitive liposomes with low pH-triggered sheddable PEG for enhanced tumor-targeted drug delivery. Nanomedicine 2019, 14, 1971–1989. [Google Scholar] [CrossRef]
  232. McNeeley, K.M.; Karathanasis, E.; Annapragada, A.V.; Bellamkonda, R.V. Masking and triggered unmasking of targeting ligands on nanocarriers to improve drug delivery to brain tumors. Biomaterials 2009, 30, 3986–3995. [Google Scholar] [CrossRef]
  233. Kale, A.A.; Torchilin, V.P. “Smart” drug carriers: PEGylated TATp-modified pH-sensitive liposomes. J. Liposome Res. 2007, 17, 197–203. [Google Scholar] [CrossRef] [Green Version]
  234. Kale, A.A.; Torchilin, V.P. Enhanced transfection of tumor cells in vivo using “Smart” pH-sensitive TAT-modified pegylated liposomes. J. Drug Target. 2007, 15, 538–545. [Google Scholar] [CrossRef] [Green Version]
  235. Kuai, R.; Yuan, W.; Li, W.; Qin, Y.; Tang, J.; Yuan, M.; Fu, L.; Ran, R.; Zhang, Z.; He, Q. Targeted delivery of cargoes into a murine solid tumor by a cell-penetrating peptide and cleavable poly(ethylene glycol) comodified liposomal delivery system via systemic administration. Mol. Pharm. 2011, 8, 2151–2161. [Google Scholar] [CrossRef] [PubMed]
  236. Hak, S.; Helgesen, E.; Hektoen, H.H.; Huuse, E.M.; Jarzyna, P.A.; Mulder, W.J.; Haraldseth, O.; Davies, C.e.L. The effect of nanoparticle polyethylene glycol surface density on ligand-directed tumor targeting studied in vivo by dual modality imaging. ACS Nano 2012, 6, 5648–5658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Geng, S.; Guo, M.; Zhan, G.; Shi, D.; Shi, L.; Gan, L.; Zhao, Y.; Yang, X. NIR-triggered ligand-presenting nanocarriers for enhancing synergistic photothermal-chemotherapy. J. Control. Release 2023, 353, 229–240. [Google Scholar] [CrossRef] [PubMed]
  238. Yang, J.; Yin, Z.; Chang, Y.; Wang, H.; Xu, J.-F.; Zhang, X. Tumor acidity-induced charge-reversal liposomal doxorubicin with enhanced cancer cell uptake and anticancer activity. Giant 2021, 6, 100052. [Google Scholar] [CrossRef]
  239. Savoy, E.A.O.; Yoon, F.P.; Mesbahi, H.; Knight, N.J.R.; Berkman, C.E. Chapter 6 Acid-labile Linkers. In Chemical Linkers in Antibody-Drug Conjugates (ADCs); The Royal Society of Chemistry: London, UK, 2022; pp. 213–231. [Google Scholar]
  240. Moghimi, S.M.; Szebeni, J. Stealth liposomes and long circulating nanoparticles: Critical issues in pharmacokinetics, opsonization and protein-binding properties. Prog. Lipid Res. 2003, 42, 463–478. [Google Scholar] [CrossRef]
  241. Olatunji, F.P.; Herman, J.W.; Kesic, B.N.; Olabode, D.; Berkman, C.E. A click-ready pH-triggered phosphoramidate-based linker for controlled release of monomethyl auristatin E. Tetrahedron Lett. 2020, 61, 152398. [Google Scholar] [CrossRef]
  242. Choy, C.J.; Ley, C.R.; Davis, A.L.; Backer, B.S.; Geruntho, J.J.; Clowers, B.H.; Berkman, C.E. Second-Generation Tunable pH-Sensitive Phosphoramidate-Based Linkers for Controlled Release. Bioconjug. Chem. 2016, 27, 2206–2213. [Google Scholar] [CrossRef]
Figure 1. Advantages of liposomes as drug delivery systems.
Figure 1. Advantages of liposomes as drug delivery systems.
Ijms 24 06615 g001
Figure 2. Schematic of liposomes and their different components.
Figure 2. Schematic of liposomes and their different components.
Ijms 24 06615 g002
Figure 3. Tumor vasculature leakage. Tumor vasculature exhibits structural abnormalities, fenestrated blood vessels, and absence of a lymphatic system (bottom) compared to normal/healthy vasculature (top). Image was created with BioRender.com (accessed on 1 February 2023).
Figure 3. Tumor vasculature leakage. Tumor vasculature exhibits structural abnormalities, fenestrated blood vessels, and absence of a lymphatic system (bottom) compared to normal/healthy vasculature (top). Image was created with BioRender.com (accessed on 1 February 2023).
Ijms 24 06615 g003
Table 1. Approval and Marketing History of Liposomal Drugs.
Table 1. Approval and Marketing History of Liposomal Drugs.
Product NameDrugLipid Composition
(Molar Ratio)
IndicationApproval DateMarketing StatusRef.
DaunoXome® (US)DaunorubicinDSPC:CL (2:1)AIDS-related Kaposi’s sarcoma1995–1997 (EMA)
1996 (FDA)
US discontinued (2016),
global on-demand access (2016, EU, UK, AU, NZ, HK)
[104,105,106,107,108]
DepoCyt®
(US, EU)
CytarabineCL:TR:DOPC:DPPG (w/w/4.4:1.2:5.7:1)Lymphomatous meningitis1999 (FDA)
2001 (EMA)
Withdrawn production by company (2017)[109,110,111]
Doxil® (US)/
Caelyx® (EU) a
DoxorubicinHSPC:CL:MPEG 2000-DSPE (calc. 3:2:0.9, w/w 3:1:1)AIDS-related Kaposi’s sarcoma, recurrent ovarian cancer, multiple myeloma, metastatic breast cancer (EU only)1995 (FDA)
1996 (EMA)
Active (US, EU)[102,112,113]
Lipo-Dox®
(TW)
DoxorubicinDSPC:CL:MPEG 2000-DSPE (3:2:0.3)AIDS-related Kaposi’s sarcoma,
ovarian cancer,
breast cancer,
multiple myeloma
1998 (TW)Active (TW)[114,115]
Marqibo®
(US)
VincristineSM:CL (60:40)Acute lymphoblastic leukemia2012 (FDA)US discontinued (2020)[104,116]
Mepact®
(EU)
MifamurtideDOPS:POPC (3:7)Osteosarcoma2009 (EMA)Active (EU)[117,118,119]
Myocet®
(EU)
DoxorubicinPC:CL (55:45)Metastatic breast cancer2000 (EMA)Active (EU)[120,121,122]
Onivyde®/Nal-IRI
(EU, US)
IrinotecanDSPC:CL:MPEG 2000-DSPE (3:2:0.015)Pancreatic cancer1996 (FDA)
2016 (EMA)
Active (US, EU)[123,124,125]
Vyxeos®/CPX-351
(EU, US)
Cytarabine: daunorubicin
(5:1 mol. ratio)
DSPG:DSPC:CL (7:2:1)Newly diagnosed therapy–related acute myeloid leukemia,
acute myeloid leukemia with myelodysplasia-related changes
2017 (FDA)
2018 (EMA)
Active (US, EU)[126,127]
Zolsketil®
(EU) a
DoxorubicinHSPC:CL:MPEG 2000-DSPEMetastatic breast cancer, advanced ovarian cancer, multiple myeloma, AIDS-related Kaposi’s sarcoma2022 (EMA)Active (EU)[128]
Notes: Molar ratios are based on the literature unless indicated as calculated (Calc.) from a drug label. AU: Australia; Calc.: calculated; CL: cholesterol; DOPC: dioleoylphosphatidylcholine; DOPS: dioleoyl-phosphatidylserine; DPPG: dipalmitoylphosphatidylglycerol; DSPC: distearoylphosphatidylcholine; DSPG: distearoylphosphatidylglycerol; EU: Europe; EMA: European Medicines Agency; HK: Hong Kong; HSPC: fully hydrogenated soy phosphatidylcholine; NZ: New Zealand; MPEG2000-DSPE: N-(carbonylmethoxypolyethylene glycol-2000)-1,2-distearoyl-sn-glycero-3-phosphoethanolamine sodium salt; TR: triolein sodium salt; SM: sphingomyelin; sPLA2: secretory phospholipase A2; POPC: 1-palmitoyl-2-oleoyl-phosphatidylcholine; PC: phosphatidylcholine; w/w: weight for weight. a: bioequivalent. However, it is unclear if Lipo-Dox® is currently used for treating myelomas.
Table 2. Liposomal Drugs in Clinical Trials *.
Table 2. Liposomal Drugs in Clinical Trials *.
Product NameDrugLipid Composition
(Molar Ratio)
ConditionsDelivery
Mechanism
StatusRef.
L-NDDP/
Aroplatin™
cis-bis-neodecanoato-trans-R,R-1,2-
diaminocyclohexane platinum (II)
DMPC:DMPGB-cell lymphoma, malignant mesothelioma, pancreatic cancer, colorectal cancer, solid tumorsEPRPhase I/II[129]
BP1002
(US)
Antisense oligonucleotide against BCl-2DOPC:ASO (20:1)Acute myeloid leukemia, advanced lymphoid malignanciesEPRPhase I[130]
EndoTAG®
(EU, US, TW, UA)
PaclitaxelDOTAP:DOPC (53:47)Breast cancer, pancreatic cancer, liver cancerElectrostaticPhase II/III[131,132]
PLM60
(US, CN)
MitoxantroneHSPC:CL:MPEG 2000-DSPE (w/w 3:1:1)Advanced hepatocellular carcinoma, small-cell lung cancer, non-Hodgkin’s lymphoma, recurrent/refractory lymphomasEPRPhase I/II[133,134]
ThermoDox®
(US)
DoxorubicinDPPC:MSPC:MPEG2000-DSPE (90:10:4)Hepatocellular carcinoma, colorectal cancer, pediatric cancer, liver neoplasms, pancreatic cancer, breast cancerTemperaturePhase I/II/III[135,136]
LiPlaCis
(DK)
CisplatinDSPC:DSPG:MPEG 2000-DSPE (mol.% 70:25:5)Adv./refractory solid tumors, metastatic breast cancer, prostate cancer, skin cancersPLA2 targetedPhase I/II[137]
Lipoplatin™CisplatinSPC-3: DPPG: CL: MPEG2000-DSPEMalignant pleural effusionsFusogenicPhase I[138]
SPI-077CisplatinHSPC:CL:MPEG2000-DSPE (51:44:5)Ovarian cancerEPRPhase II[139,140]
* Searched trials are listed as either “active”, “unknown”, “not yet recruiting”, “terminated”, or “completed” in the clinicaltrials.gov database as of 26 March 2023. The listed liposomal drugs under clinical evaluation are not exhaustive. ASO: antisense oligonucleotide; CL: cholesterol; DMPC: 1,2-dimyristoylphosphatidylcholine; DMPG: 1,2-dimyristoylphosphatidylglycerol; DOTAP: dioleoyl-3-trimethylammonium propane; DPPC: dipalmitoylphosphatidylcholine; DPPG: dipalmitoylphosphatidylglycerol; HSPC, hydrogenated soy phosphatidylcholine; MPEG2000-DSPE: α-(2-(1,2-distearoyl-sn-glycero(3)phosphooxy)ethylcarbamoyl)-ω-methoxypoly(oxyethylen)-40; MSPC: monostearoylphosphatidylcholine; SPC-3, soy phosphatidyl choline.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fulton, M.D.; Najahi-Missaoui, W. Liposomes in Cancer Therapy: How Did We Start and Where Are We Now. Int. J. Mol. Sci. 2023, 24, 6615. https://doi.org/10.3390/ijms24076615

AMA Style

Fulton MD, Najahi-Missaoui W. Liposomes in Cancer Therapy: How Did We Start and Where Are We Now. International Journal of Molecular Sciences. 2023; 24(7):6615. https://doi.org/10.3390/ijms24076615

Chicago/Turabian Style

Fulton, Melody D., and Wided Najahi-Missaoui. 2023. "Liposomes in Cancer Therapy: How Did We Start and Where Are We Now" International Journal of Molecular Sciences 24, no. 7: 6615. https://doi.org/10.3390/ijms24076615

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop