Back to Journals » International Journal of Nanomedicine » Volume 20
Advancements in Liposomal Nanomedicines: Innovative Formulations, Therapeutic Applications, and Future Directions in Precision Medicine
Authors Izadiyan Z, Misran M, Kalantari K, Webster TJ , Kia P , Basrowi NA, Rasouli E, Shameli K
Received 27 July 2024
Accepted for publication 1 January 2025
Published 31 January 2025 Volume 2025:20 Pages 1213—1262
DOI https://doi.org/10.2147/IJN.S488961
Checked for plagiarism Yes
Review by Single anonymous peer review
Peer reviewer comments 3
Editor who approved publication: Prof. Dr. RDK Misra
Zahra Izadiyan,1 Misni Misran,1 Katayoon Kalantari,2 Thomas J Webster,3,4 Pooneh Kia,5 Noor Ashyfiyah Basrowi,1 Elisa Rasouli,6 Kamyar Shameli7
1Department of Chemistry, Universiti Malaya, Kuala Lumpur, Malaysia; 2Department of Chemical Engineering, Northeastern University, Boston, MA, USA; 3Biomedical Engineering, Hebei University of Technology, Tianjin, People’s Republic of China; 4School of Engineering, Saveetha University, Chennai, India; 5Institute of Bioscience, Universiti Putra Malaysia, Serdang, Selangor, Malaysia; 6Department of Electrical and Electronics Engineering, Nanyang Technological University, Nanyang, Singapore; 7School of Medicine, Institute of Virology, Technical University of Munich, Munich, Germany
Correspondence: Zahra Izadiyan; Kamyar Shameli, Email [email protected]; [email protected]
Abstract: Liposomal nanomedicines have emerged as a pivotal approach for the treatment of various diseases, notably cancer and infectious diseases. This manuscript provides an in-depth review of recent advancements in liposomal formulations, highlighting their composition, targeted delivery strategies, and mechanisms of action. We explore the evolution of liposomal products currently in clinical trials, emphasizing their potential in addressing diverse medical challenges. The integration of immunotherapeutic agents within liposomes marks a paradigm shift, enabling the design of ‘immuno-modulatory hubs’ capable of orchestrating precise immune responses while facilitating theranostic applications. The recent COVID-19 pandemic has accelerated research in liposomal-based vaccines and antiviral therapies, underscoring the need for improved delivery mechanisms to overcome challenges like rapid clearance and organ toxicity. Furthermore, we discuss the potential of “smart” liposomes, which can respond to specific disease microenvironments, enhancing treatment efficacy and precision. The integration of artificial intelligence and machine learning in optimizing liposomal designs promises to revolutionize personalized medicine, paving the way for innovative strategies in disease detection and therapeutic interventions. This comprehensive review underscores the significance of ongoing research in liposomal technologies, with implications for future clinical applications and enhanced patient outcomes.
Keywords: liposomes, active targeting, targeted drug delivery, nano-carriers, cancer therapy
Introduction
Cancer refers to the uncontrolled growth of abnormal cells in the body and is the second leading cause of death from non-communicable diseases (NCDs) worldwide. Cancer is a significant societal, public health, and economic challenge in the 21st century, accounting for nearly one in six deaths (16.8%) and over one in five deaths (22.8%) globally.1 It is responsible for 30.3% of premature deaths among individuals aged 30–69 and ranks among the top three causes of deaths in this age group in 177 out of 183 countries.1 Cancer not only poses a significant obstacle to extending life expectancy but it also incurs considerable societal and macroeconomic costs, which differ depending on the type of cancer, geographic region, and gender.2 A recent study highlighted the significant impact of disproportionate cancer mortality among women: in 2020, approximately one million children lost their mothers to cancer, with nearly half of these maternal deaths caused by breast or cervical cancer.3 As shown in Figure 1, the top 10 cancer types in both men and women contribute to more than 60% of all new cancer cases and deaths.4 Specifically, lung cancer is the most frequently diagnosed worldwide, representing 12.4% of cases, followed by breast cancer in women (11.6%), colorectal cancer (9.6%), prostate cancer (7.3%), and stomach cancer (4.9%). Lung cancer is also the leading cause of cancer-related deaths, accounting for 18.7%, with colorectal (9.3%), liver (7.8%), breast (6.9%), and stomach (6.8%) cancers following. In women, breast cancer is the most diagnosed and the leading cause of death, followed by lung and colorectal cancers in both diagnoses and mortality. For men, lung cancer is the most common in terms of both cases and deaths, followed by prostate and colorectal cancers for new diagnoses, and liver and colorectal cancers for deaths.4
![]() |
Figure 1 New cases and mortality rates for the top 10 leading cancers in 2022. Used from Bray F, Laversanne M, Sung H, et al. Global cancer statistics 2022: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. Ca a Cancer J Clinicians. 2024;74(3):229–263. © 2024 The Authors. CA: A Cancer Journal for Clinicians published by Wiley Periodicals LLC on behalf of American Cancer Society.4 |
Cancer treatment involves various approaches, including surgery, chemotherapy, and radiation therapy.5 However, these treatments can lead to side effects such as healthy cell damage, hair loss, infections, pain, nausea, mucositis, and vomiting. To mitigate these adverse effects and enhance treatment effectiveness, nanocarrier-based drug delivery systems have been introduced as alternatives to traditional cancer therapies. Among these systems, liposomes are one of the most widely used.6,7 Liposomes offer significant advantages for cancer treatment due to their unique biological properties, including biocompatibility, biodegradability, cell-like membrane characteristics, low immunogenicity, and non-toxicity. They also help protect drugs from degradation and extend their biological half-life, while allowing for easy modification of size, charge, and surface properties.8,9 Extensive research in recent years has advanced the development of liposomal drug delivery systems, leading to new formulations for cancer therapy.10
In a study published in 1964, Bangham et al provided the first description of liposomes.11 The vesicles were once known as “banghasomes” or “multilamellar smectic mesophases”, but Gerald Weissmann eventually dubbed these systems “liposomes” rather than “banghosomes”, and he was granted the Nobel Prize for this discovery.12 According to these investigations, liposomes are tiny vesicles that can be formed from cell membrane proteins, phospholipids, cholesterol, and nontoxic surfactants.11 Bilayer vesicles are formed when phospholipids envelop the hydrophilic core of liposomes.13 The hydrophobic tails of the phospholipids point inward (against the membrane), whilst the hydrophilic heads point outward (toward the aqueous phase). Because liposomes are amphipathic, they can be loaded with hydrophilic or hydrophobic medicines.14 Liposomes can be classed according to their structure, composition, and manufacturing method. Liposomes can be classified as multilayer, monolayer, or multivesicular based on their structural makeup. Liposomes can be categorized as conventional, fusogenic, long-lived, pH-sensitive, ionic, magnetic, heat-sensitive, and immuno-liposomes based on their composition.15 The physicochemical characteristics of the membrane components, the charge, and the dispersion medium are taken into consideration while selecting liposome manufacturing techniques.6 The preparation of liposomes can be achieved through three methods: solvent dispersion (ethanol/ether injection, double emulsion, and reverse-phase evaporation), mechanical dispersion (hydration of lipid films, sonication, micro emulsification, French pressure cells, membrane extrusion, and freeze-thawing), and detergent solubilization (dialysis, column chromatography, and dilution).15
Liposomal formulations offer a number of advantages over drug solutions including reduced toxicity of the encapsulated drug,16 prolonged systemic circulation when surface-modified (eg PEGylated liposome), improved pharmacokinetics,17 controlled drug release kinetics,18 and tumor targeting.19 Drugs with various physicochemical characteristics can be delivered by liposomes due to their special capacity to encapsulate both lipophilic and hydrophilic substances. These characteristics include overcoming multidrug resistance (MDR), improving the therapeutic index, enhancing drug solubility, sustaining drug release, decreasing drug adverse effects, increasing the concentration of the medication at the target site, and biocompatibility and biodegradability. They also include non-immunogenicity.19 However, these systems’ short half-lives, instability, heightened susceptibility to sterilizing procedures, and high production costs due to costly raw materials and the manufacturing equipment needed are major drawbacks.
The response of the immune system is another important issue. Liposomes are recognized as foreign materials by optogenins, which causes reticuloendothelial system (RES) macrophages to absorb them. Sterically stabilized liposomes (covered with PEG or other hydrophilic polymers) have been created as a solution to the RES absorption issue. The most popular method providing liposomes with a longer half-life in circulation is PEGylation. Additionally, silic acid, polyvinyl alcohol, and poly-N-vinylpyrrolidone—other PEG substitutes—have been employed for the same objective.17
One of the most popular nanocarriers for delivering anticancer drugs to tumor locations include liposomes. Targeting tactics alone or in combination can accomplish this. Because of the increased permeability and retention effect (EPR) brought on by the decreased lymphatic drainage and increased vascular permeability of the tumor microenvironment, liposomes are preferentially absorbed by solid tumors.20
Currently, liposomes are used as active and smart carriers helping bioactive agent accumulate in a specific part of the body.21 Liposomes can be modified by the attachment of antibodies or ligands on their surface which can then be recognized by cellular receptors.22 Because they lack selectivity, the majority of anticancer medications have harmful effects on both malignant and healthy cells. Although there have been attempts to choose treatments that eradicate tumor cells without endangering healthy tissue, the outcomes of chemotherapy typically fall short of these goals. Therefore, there is high scientific attention for the development of new anticancer therapies and new drug delivery strategies that can selectively deliver anticancer drugs to malignant tissues, thus increasing drug efficacy and possibly reducing their toxicity and adverse effects on normal cells.23,24 Liposomal systems with a potential to enhance medication delivery for cancer therapy have made substantial progress in recent years. As a result, scientists have concentrated their efforts on making liposomal delivery systems for active cancer medication targeting to the tumor site, followed by organelle-specific targeting and triggered release of loaded pharmaceuticals that take advantage of the tumor’s microenvironment.25
This review article aims to provide a concise overview of liposomes in modern drug delivery, particularly for cancer therapy. It covers liposomal structure, synthesis methods, drug encapsulation strategies, and their potential in targeted cancer therapies. The focus is on their biocompatibility, ability to overcome biological barriers, and applications in personalized cancer treatment. The goal is to enhance the understanding of liposomal drug delivery systems and their role in improving efficacy and reducing toxicity in cancer treatment.
Liposome Structure
Liposomes have a spherical bilayer structure which consistently includes one or more layers of a phospholipid that can be produced from cholesterol and natural/synthetic phospholipids. Lipophilic and hydrophilic materials are embedded in a lipid bilayer and interior aqueous region, respectively.26 Liposomes with a phospholipid-based structure are known as amphipathic nanocarriers.27 They have advantages such as extending the release of active pharmaceutical agents, biocompatibility, and biodegradability.28 Liposomes are categorized based on the vesicle size and number of lipid bilayers (lamellae) which is presented in Figure 2. 29 The main types of liposomes can be considered as multilamellar (ML, 0.5 to 5 μm) and multivesicular (MV, > 1μm).30,31 Unilamellar also can be classified into three different types of vesicles including small unilamellar (SU), large unilamellar (LU), and giant unilamellar (GU) of different sizes 20–200 nm, > 200 nm and ≥ 1μm, respectively. Unilamellar vesicles are described by the presence of a single bilayer, with the extra capacity for the enclosure of a hydrophilic material. Multilamellar vesicles are desirable choices for the enclosure of a lipophilic material and also, they represent two or more concentric lipid bilayers structured through an onion-like structure. Multi vesicular vesicles are perfectly suitable for the enclosure of a great number of hydrophilic substances. Besides that, they contain a few small non-concentric vesicles trapped inside a lipid bilayer.31,32 Furthermore, the amount of encapsulation of the compounds in a liposome depend on the number of lamellae and the vesicle size.15
Liposome Synthesis Methods
There are different methods to synthesize liposomes.33 The most common methods such as thin film hydration, solvent injection, reverse phase evaporation, dehydration-rehydration, hydration in a packed bed of colloidal particles, pH jumping, freeze-thaw, and detergent removal are discussed in the following sections.34,35 Table 1 provides a detailed overview of the advantages and disadvantages associated with different liposome synthesis methods, offering insights into their efficacy, scalability, and application-specific suitability which are further described below.
Thin Film Hydration
Liposome synthesis began with the Bangham method, also known as thin film hydration.11 This technique involves dissolving lipids in organic solvents like chloroform, ether, or methanol. The solution is then evaporated in a round-bottom flask to yield a thin lipid film, which, upon hydration with an aqueous solvent, forms liposomes. The conditions during hydration play a pivotal role in shaping the liposome structures. Intense agitation leads to the creation of multilamellar vesicles with various sizes, while a gentler hydration process results in the formation of giant unilamellar vesicles.36,37 However, this method comes with limitations: it tends to produce larger and more diverse liposomes, has restricted containment capacity, poses challenges in removing organic solvents, and can present scalability issues.35
Solvent Injection Techniques
The solvent injection technique, first detailed in 1973 by Batzri and Korn, offers an alternate method for liposome production.38 Here, lipids are dissolved in an organic solvent like ethanol or ether and are injected into an aqueous phase, leading to the formation of liposomes.39 It is widely used due to its scalability, reproducibility, simplicity, rapid implementation, and absence of oxidative changes or lipid degradation.40 This method harnesses ethanol, acknowledged by the European Pharmacopeia for its suitability in diverse applications, including in vivo drug delivery.41 However, this approach is not without drawbacks: ethanol might present challenges with lipid solubility, efficient removal from liposomes, agitation-related liposome heterogeneity, and low encapsulation efficiency (EE) for hydrophilic compounds.32 Modifying parameters such as drug-to-lipid ratio, injection rate, the nature of the lipid, orifice diameter during injection, and ethanol lipid concentration allow for control over particle size and EE achieved in the ethanol injection process.42
Reverse Phase Evaporation
Reverse phase evaporation is another approach to prepare liposomes and the reverse-phase evaporation process was first described by Szoka and Papahadjopoulos.43 The initial steps are similar to the thin film hydration. Firstly, phospholipids are dissolved in an organic solvent to produce the film, then the solvent is removed by evaporation. The new film is resolved in the organic solvent again which is normally diethyl ether and/or isopropyl ether, followed by an additional aqueous phase. The final output is oil in a water emulsion formulation.37 Then, the new formulations are sonicated to generate inverted micelles, resulting in the formation of a homogeneous emulsion. In the final step, the organic solvent is evaporated under reduced pressure. The resulting liposomes are in the form of a viscous gel.15,32 One of the main advantages of this method is high EE.42 Also, this technique is defined as time-consuming.35 On the other hand, the weaknesses of this method is based on the encapsulation of the mixture because of sonication and even the organic solvents used.44
Detergent Removal
Another method used to synthesize liposomes is the detergent removal technique. In this process, at the critical micelle concentration, phospholipids are solubilized with detergents.36 After removal of the detergent by column chromatography or dialysis bags and with a suitable aqueous medium, the phospholipid molecules self-assemble into liposomes.15,33 Some of the parameters can affect the homogeneity and the size of the liposomes, counting rate of detergent elimination and initial ratio of phospholipids to detergents.32,42 The presence of contaminants in the final liposomal formulation, the possibility of interaction between the encapsulated drug and the detergent, and the fact that this procedure is time intensive are all disadvantages of the detergent removal process.35,45
Hydration in a Packed Bed
This method represents a one-step hydration-based approach to produce liposomes with a remarkably low polydispersity index (PDI 0.2) and eliminates the need for post-processing. Prior to hydration, lipid molecules, absorbed in a solvent, undergo drying within a densely packed bed of highly asymmetrical colloidal particles with rough surfaces, enabling the creation of liposomes within a specific size range. The resulting size distribution remains consistent regardless of the flow rate of the hydrating medium, indicating that extrusion does not influence the narrow size distribution of these liposomes. By subjecting a milky white dispersion of large and polydispersed liposomes (in the micrometer range) to drying in a packed bed and subsequent rehydration with an aqueous buffer, a monodispersed liposome dispersion below 100 nm can be achieved. Notably, the final size distribution remains unaffected by the size of the colloidal particles or the percentage of bed packing, emphasizing that the highly asymmetrical particles and porous packing structure dictate liposome size. Sundar et al highlight the robustness of this one-step hydration method, its lack of post-processing requirements, and the precise control it offers over liposome size, positioning it as suitable for point-of-care therapies utilizing liposomal drug delivery systems.46
Dehydration-Rehydration
Large unilamellar liposomes can be synthesized by dehydration-rehydration without using detergents or organic solvents. In this technique, lipid or amphiphilic molecules are dispersed into the aqueous phase at a low concentration along with sonication to produce the liposome.47 The drug for closure could be mixed with the formulated vesicles in the aqueous phase. Liposomes combine to form a multilayer film that traps drug molecules when water evaporates under the flow of nitrogen gas. After adding water, large vesicles are formed, encasing the active ingredient.
Detergent Removal
Another method to synthesize liposomes is the detergent removal technique. In this process, at the critical micelle concentration, phospholipids are solubilized with detergents.36 After removal of the detergent by column chromatography or dialysis bags and with a suitable aqueous medium, the phospholipid molecules self-assemble into liposomes.15,33 Some of the parameters can affect the homogeneity and the size of the liposomes, counting rate of detergent elimination and initial ratio of phospholipids to detergents.32,42 The presence of contaminants in the final liposomal formulation, the possibility of interaction between the encapsulated drug and the detergent, and the fact that this procedure is time intensive are all disadvantages of the detergent removal process.35,45
pH Jumping
Another quick method for making liposomes is pH jumping, which does not require organic solvents. Small unilamellar structures form when a phosphatidic acid solution in water is exposed to a 3.5-fold increase in pH (from 3 to 10.5–11) for a short time period (less than 2 minutes).48 When the same procedure is performed on a mixture of phosphatidic acid and phosphatidylcholine, similar results can be obtained, with a specific ratio of phosphatidic acid to phosphatidylcholine, a controlled percentage of small unilamellar versus large unilamellar structures can be obtained.48
Freeze-Thaw
Freeze-thaw cycles are commonly used in liposome production to improve lipid formation and unilamellar vesicle packing.49 The freeze-thaw process might be included in any liposome manufacturing method. For example, following thin filming, the mixture is sonicated at ambient temperature and then frozen in a liquid nitrogen atmosphere at −196 °C. After that, the sample is kept at room temperature to melt. The above-mentioned cycles might be repeated up to ten times to get the desired outcome. The final result is a huge number of unilamellar vesicles. As a final point, if smaller vesicles are needed, the resultant solution can be resonated at room temperature. When lipid concentrations are high, freeze-thaw is not an appropriate approach.15 By using this strategy, the drug enclosure efficiency was reported to be between 20% and 30%.15,50 Freeze-thaw cycling is a common liposome synthesis strategy for increasing the EE.51 The liposomes are usually frozen in liquid nitrogen (−196°C) and then thawed at a temperature above the lipid phase transition temperature.52,53 Freeze-thaw cycling is used to reduce lamellarity,54 reduce polydispersion, and/or tear the liposomal bilayer,51 allowing drug molecules to enter the liposome and facilitate encapsulation.55,56 The needed number of freeze-thaw cycles to encapsulate psychoactive compounds varies widely in the literature, with some papers claiming as many as ten.57 The goal of performing a large number of freeze-thaw cycles is to achieve drug concentration equilibrium.
Extrusion Techniques
Most of the previous methods need additional steps to reduce liposome size such as extrusion, homogenization, and sonication.58 Bath and probe sonication techniques are used to control the size of the liposomes.15 The disadvantages of the sonication technique include that it may be difficult to supply similar ultrasonic energy in a large volume of liposomal suspension (scale-up), and the probe tip may be a source of metal contamination. Furthermore, there is a risk of phospholipid breakdown and subsequent compound enclosure, as well as reduced EE.38,59 Liposomes can be driven through a high-pressure aperture to reduce their size in a homogenization process, resulting in a high-speed collision idea. Size reduction procedures include shear force-induced homogenization processes, micron fluidization, and homogenization.42 Another way for reducing the size of liposomes is the extrusion procedure. Following liposome formation, extrusion cycles are passed a few times through a membrane with set pore size, which is typically a polycarbonate filter to ensure consistent size distribution.35,60 Comparing the extrusion process with using homogenizers requires a lower volume of liposomes and a much lower pressure.58
Characterization of Liposomes
Once the liposomes are formulated and before they are used, they need to be evaluated for physical and chemical properties and they should be extensively characterized to guarantee their in vitro and in vivo performance.61 The main characterization of liposomes are size, polydispersity index (PDI), and zeta potential which are related to the stability, shape, phase behavior, lamellarity, in vitro drug release, and EE.33,36 In Table 2, a thorough list of lipid classification and respective characterization techniques are presented, detailing the merits and limitations of each method.
Size and Polydispersity Index (PDI)
Size and PDI are the most important characteristics of liposomes. Size is known to be a critical factor for inhalation and parenteral administration,27 and finding the liposome’s circulation half-life.62 The small size of liposomes allows them to circulate in the organism for a long time period while larger liposomes are not suitable and are speedily eliminated from the blood circulation system.63 The acceptable size range for liposomes in drug delivery is usually between 50 to 200 nm.39 The PDI value indicates the size of the sample heterogeneity, which can be monodispersed or polydispersed. The PDI can be in a range from 0 to 1 and dimensionless while the desirable range of PDI in drug delivery should be below 0.3 or equal to this value.64 The high PDI can be caused by a very wide range of size distribution or heterogeneity and also several populations of liposomes in the sample.65 Based on the particle size, the PDI can be calculated with the solvent refractive index, the distribution variance, and the angle of measurement.66 The measurement is mostly carried out by using dynamic light scattering (DLS) moreover being identified with photon correlation spectroscopy (PCS). DLS analysis is based on the continuous motion of dispersed particles in the solution, resulting in scattering of the incident light. Light scattering is proportional to the diffusion level of liposomes in suspension, implying that tiny particles diffuse more quickly than big ones. The quantity of light dispersed is used to compute the mean size of the liposome. DLS is considered a quick, straightforward, simple, and dependable method for determining the size of liposomes in their natural habitat. Furthermore, DLS can measure a wide variety of sizes from nanometers to micrometers.33,36 Nevertheless, this technique has certain limitations, including difficulty in distinguishing individual particles from aggregates and it has a high sensitivity in detecting a small amount of impurities (contaminants) which can confound results.67
Recently, a new tool for size characterization called Nanoparticle Tracking Analysis (NTA) was presented to measure the diffusion coefficient of particles in a sample by determining size.68 DLS calculates the diffusion coefficients of the particles based on the intensity change of the scattered light measurements. The dispersion coefficient of individual particle motions in successive optical video pictures can be determined by NTA. Because they measure the same physical attribute, NTA can be a useful approach for confirming the size as determined by DLS. As a result, the NTA size measurements should be the same as those obtained using the DLS approach.69,70 The ability of NTA to simultaneously measure the size and intensity of particle dispersion allows, in addition to differentiating between particles with different refractive indices within the same sample solution, a direct estimation of the particle concentration.71
Zeta Potential
A “colloidal system” is formed when one of the states of matter is finely scattered in another. In aqueous media, the majority of colloidal dispersions carry an electric charge. The surface charges can come from a variety of sources, depending on the particles and their surroundings. The two most important mechanisms are ionization of the surface groups (acidic groups dissociate on a particle’s surface, resulting in a negatively charged surface; conversely, a basic surface takes on a positive charge) and adsorption of a charged species (ie, surfactant ions may be precisely adsorbed on the particle surface, resulting in a positively charged surface in the case of cationic surfactants and a negatively charged surface in the case of anionic surfactants).27 The total charge that a particle accumulates in a given medium is defined by its zeta potential. It is a physical property shared by all particles in suspension. The zeta potential has long been acknowledged as a reliable indicator of colloidal particle interaction. Zeta potential measurements are commonly used to estimate colloidal system stability. If all of the particles in the suspension have a large negative or positive zeta potential, they will not aggregate. If the particles’ zeta potential values are low, however, there will be no force to keep them from flocculating. To determine the zeta potential, a laser is used to create a light source that illuminates particles within the samples. At an angle of about 13 degrees, the incoming laser beam passes through the sample cell’s center and is detected as scattered light.
Any particles moving across the volume measured cause the measured light to fluctuate at a frequency proportional to the particle speed when an electric field is applied to the cell. This information is sent to a digital signal processor, which then sends it to a computer, which calculates the potential zeta. Particle suspensions with a zeta potential greater than or equal to +30 mV or less than −30 mV are considered stable.72 Each particle has a charge and overall this charge usually is stated as zeta potential or surface charge.58 For nanoparticle surface characterization, zeta potential has become a common analytical measurement. Nanoparticle stability, circulation times, protein interactions, particle cell permeability, and biocompatibility can all be determined using the potential at the hydrodynamic shear boundary (also known as the sliding plane).73 However, to draw meaningful conclusions from this data, it is necessary to grasp the technique’s limitations and to define the measurement conditions properly. Zeta potential is influenced by temperature, pH, conductivity (ionic strength), and solvent (viscosity).74 Small adjustments in one of these factors can have a big impact on the zeta potential readings. Zeta potential is considered a necessary physical property of liposomes to electrostatic interactions between particles in a sample solution.75 The surface charge of liposomes are related to several parameters which include a head group of lipids, lipid composition, and ligands, also the surface charge could be present as different charges such as negative, neutral, or positive. Besides, ionic strength has an effect on zeta potential as well and one can consider this as an external environmental effect.76
Liposomal surface charge is a crucial property for tumor dissemination that must be carefully examined. Cationic liposomes accumulate in the tumor vasculature due to electrostatic interactions with angiogenic endothelial cells found in tumor blood vessels. However, due to the extracellular matrix and electrostatic attachment to cancerous cells, highly charged cationic liposomes do not diffuse efficaciously into the tumor site, whereas less cationic or neutral liposomes have shown more efficient penetration into tumor spheroids in vitro and extravasation of blood vessels in vivo.77–79 Lipid functionalization with polyethylene glycol (PEG), according to a study, protects cationic groups from potentially damaging electrostatic interactions with tumor cells and the extracellular matrix. PEGylated cationic liposomes have clumped together in the tumor vasculature in vivo, resulting in a uniform tumor distribution. This method of PEGylation to maintain cationic liposomes (in a positive charge) may be a good way to make liposomes that can penetrate solid tumors and target sites effectively.80,81
Shape
Morphological analysis parameters are critical for effective liposome characterization.71 Liposomal images can be captured using electron microscopy (TEM) and cryo-TEM methods.36 Because the original environment of the liposomes must be removed, the TEM technique has several constraints in terms of sample preparation. Because this is a time-consuming procedure, it is not suitable for routine measurements. Furthermore, this approach has the potential to cause changes in the structure of the liposomes, such as vesicle shrinking, swelling, or the production of artifacts in the produced image.33 There is another alternative technology, cryo-TEM, which can circumvent the sample preparation constraint. By adopting a liquid nitrogen flash-freezing process followed by direct visualization of the liposomes in a controlled environment, this technique retains liposomes near to their original condition and reduces shape distortion or shrinkage. Nonetheless, cryo–TEM typically produces better results with smaller particle sizes than with larger particle sizes because larger particle sizes may be removed from the samples during the preparation process. The AFM approach has been employed for direct investigation of liposomes in their natural environment without any alteration. This approach is thought to be non-invasive, powerful, and quick.27 The main benefit of AFM over electron microscopy has been the greater resolution of three-dimensional micrographs offered by AFM, which can be down to the nanometer and Angstrom scales.82
Lamellarity
Lamellarity is also a property that due to its influence on EE and drug release profile, can influence subsequent liposomal applications. One of the most useful methods which provides valuable information, such as inter-bilayer and their bilayer thickness regarding liposome lamellarity, is 31P-NMR.32 Other methods for accessing lamellarity information are based on variations in the visible signal or fluorescence of the lipid marker upon the addition of certain reagents.36 The 31P-NMR technique has also been used to estimate the liposomal lamellarity value, in particular, the ratio between the number of phospholipids in the outer layers and the inner layers. Paramagnetic ions (Mn2+, Co2+ and Pr3+) are often used to prepare an NMR sample to deactivate the 31P- NMR signal of the phospholipids. The interaction of ions within the bilayer can change the NMR spectrum. Thus, by comparing the two spectra before and after the paramagnetic ion incorporation, it is possible to estimate lamellarity.83 Other techniques, such as SAXS and trapped volume, are also used to estimate the lamellarity of liposomes.84
Phase Behavior
In drug delivery applications, phase behavior is taken into account, since the permeability of the lipid bilayer for the hydrophilic active ingredients increases with the fluidity of the lipid membrane.85 There are a few other properties that are dependent on the phase behavior of the liposomal membrane such as stability fusion, protein binding, and aggregation.32 DSC is generally the most common method for studying and determining transition temperature (Tc). The thermal analysis method depends on estimating the heat flow differential between a reference sample and a real sample. Both samples are exposed to planned heating, cooling, or isothermal treatment while the environment, which is generally saturated with nitrogen gas, is carefully controlled.36 The transition temperature of phospholipids (Tc) can be determined by other approaches like fluorescence probe polarization, TGA, NMR, electron paramagnetic resonance, FTIR and XRD.86,87 Molecular dynamic simulations can be explored by calculating the phase behavior of phospholipids in the lipid bilayers.88
Encapsulation Efficiency (EE)
Optimal liposomal property studies can lead to the creation of liposomal formulations with optimal EE and medication release control. The composition of liposomes, the generation of liposomes, and the stiffness of the bilayer membrane can all have a significant impact on the EE of a particular medication.32 The quantity of medication-loaded is an important point of therapeutic effectiveness in the medical industry.89 The EE is the proportion of drug quantity incorporated into liposomes (encapsulated drugs) including the total amount of drug utilized in manufacturing liposomes (encapsulated and non-encapsulated drugs). The resulting liposomal formulation comprises a combination of encapsulated and non-encapsulated medication. As a result, separating the free (non-encapsulated) drug is the first step in assessing the amount of drug in liposomes and therefore evaluating EE. Many methods have been utilized for this purpose, including size exclusion chromatography based on size differences (liposome vs free drug), gravity or centrifugation, ultracentrifugation, and a dialysis membrane with an adequate cut-off.27 The medication quantity is then measured to determine the encapsulated amount inside the liposomes in the following phase.
There are two recognized methods for the determination of EE: direct and indirect methods. The indirect technique focuses on determining the concentration of the non-encapsulated drug in the elution and subtracting it from the overall concentration of the medication employed in the liposome production. The direct approach, on the other hand, determines EE by physically breaking the liposomes with an organic solvent and then measuring the material liberated.90 Traditional methods for estimating drug concentration in liposomes include UV-Vis and fluorescence spectroscopy, and enzyme or protein tests.27 Furthermore, more advanced technologies such as UPLC, HPLC, gas chromatography-mass spectrometry (LC-MS or GC-MS), and liquid chromatography can be used to determine the quantity of medication.91 Other techniques like 1H- NMR and ESR have also been utilized to quantify the drug amount.92,93 Liposomes are water compartment-encapsulating vesicles made up of one or more lipid layers. Because of their high biocompatibility, liposomes have been used to deliver a wide range of chemicals. They improve the therapeutic indices of the pharmacological molecules enclosed in them dramatically. AmBisome (amphotericin B), DaunoXome (daunorubicin citrate), and Doxil were among the first commercial liposomal dose formulations (doxorubicin). Many more are now being tested in clinical studies.94 Liposomes can be used as carriers in both lipophilic and hydrophilic medications due to their biphasic composition. Depending on their solubility and partitioning qualities, drug molecules are positioned differently in the liposomal environment and exhibit different entrapment and release behaviors (Figure 3). Based on these characteristics, the medications are divided into four categories: very hydrophilic, drugs with biphasic insolubility, amphiphilic pharmaceuticals with good biphasic solubility, and highly lipophilic. Only the watery compartments of liposomes contain very hydrophilic drugs with a log P < - 0.3, such as cytosine arabinoside and CDP choline. The bilayer’s composition affects the transport of such compounds through the liposomal membrane.92 Highly lipophilic medicines, such as cyclosporin, are virtually entirely entrapped in the lipid bilayer of the liposomes (log P oct > 5). Because they are very poorly soluble in water, difficulties like as entrapment drug loss during storage is minor with this category of drugs. Drugs having intermediate partition coefficients, such as 1.7 < log Poct < 4, offer a significant difficulty since they rapidly partition between the lipid and aqueous phases and are quickly lost from liposomes. Mitomycin C,95 actinomycin D, and vinblastine are a few examples.96 Only when these molecules form compounds with the membrane lipids can they generate stable liposomal systems. The most difficult choices for liposomal entrapment, however, are pharmacological compounds with low biphasic solubility. Because they are insoluble in either the aqueous or lipid phases, they are only a minor component of liposome absorption. 6-mercaptopurine, azathioprine, and allopurinol are typical instances.96 Because the liposome was utilized as a biomembrane model, the EE of the drug was shown to be proportional to the partition coefficient between 1-octanol and water.97 The log P octanol/water values of 5-fluorouracil (5-FU) (0.78), ibuprofen (3.72), flurbiprofen (4.11), and ketoprofen (2.81) were obtained from SciFinder Scholar and computed using Advanced Chemistry Development (ACD) Software Solaris V 4.67. The literature cited the log P octanol/water of diclofenac sodium as 0.70. The log P octanol/water ratio was substantially linked with the EE of the five medicines (rs = 0.97).98
![]() |
Figure 3 Different drugs and their location in the liposomal vesicle. |
Drug Loading and in vitro Release
Due to the ability of drug loading with different techniques, liposomes are a desirable system for drug delivery.27 The choice of a suitable approach for drug enclosure into liposomes depends on numerous factors which are cost efficiency of EE, liposome stability, drug/lipid ratio, sterility, facility of production and scale-up, and drug leakage and retention.32,99 Moreover, the volume of encapsulated drugs is related to the method used to produce the liposomes and also to the liposome’s composition and type of drug.33 There are two different methods to load the drug into liposomes which are active and passive.15 In the passive method, the drug is encapsulated during the preparation of the liposome. Hydrophilic drugs were spread in the aqueous phase while hydrophobic drugs were placed in the bilayer of the liposome.32 In this process, as liposomes form, they can capture the aqueous volume that contains the previously dissolved hydrophilic drug. Therefore, the drug concentration in the aqueous core is similar to the volume of water trapped by the liposomes. In the passive loaded drug, the EE can change due to a few characteristics such as production method, drug solubility, lipid concentration, liposome size, and zeta potential.100 The drug and charged ions are not able to penetrate the liposomal membrane. Otherwise, uncharged drugs can diffuse through the lipid membrane, which can lead to drug leakage. Typically, this approach results in low EE, involving a large number of unencapsulated drugs and a large loss of drugs for drugs that are permeable to the liposomal bilayer.101
There is another method of drug loading which is called active or remote loading, which contains the making of an ion gradient or transmembrane pH, which efficiently drives the drug over the lipid bilayer and in some drugs lead up to a 100% loading. After liposome preparation, this approach is applied. The gradient is created between the interior of intact (already formed) liposomes and the exterior of the liposome, the drug is solubilized in the aqueous medium. According to previous work, drugs can cross the lipid membrane, they convert them to a protonated form, prevent them from spreading outside the liposomes and also improving the EE and retention inside the liposome.101 When the drugs are weakly acidic (pKa>3) or amphipathically a weak base (pKa ≤ 11), the desirable loading efficiency was achieved.89 There are different methods to actively load drugs which includes following a calcium acetate gradient for weakly acidic drugs, ammonium sulfate transmembrane gradient for amphipathically weak bases, a phosphate gradient method, an EDTA gradient method and ionophore loading method.101
The in vitro drug release could be evaluated by using dialysis conditions. The choice of a dialysis bag should match the specifications of the drug. It must be freely penetrable to the drug and there must be no adsorption of the drug.27 The liposome sample with a specific molecular weight is placed in the dialysis bags hermetically tied and cut off. Usually, the tubing membrane is put in the buffer at pH 7.4, and the system is kept at a 37 ◦C simulated environment, and under continuous stirring. At defined times, an aliquot of the sample is taken and analyzed according to standard drug quantification methods and the sample volumes need to be kept at a constant level. Hence, an equal volume is added to the system from fresh medium.36 The cumulative release percentage is plotted against the selected time points to create the release profile. The results from drug release research are evaluated in the development of liposomes for the controlled release of pharmaceuticals as an extrapolation to in vivo liposome performance.102
Classification of Liposomes
Liposomes are small, artificial, enclosed spherical vesicles with a phospholipid bilayer separating one aqueous medium from another, capable of encapsulating hydrophilic molecules in the internal aqueous core or sequestering hydrophobic drugs in the lipid bilayer, and providing a controlled release system (Figure 4).103–105 Liposomes are commonly employed as drug delivery nanocarriers because they may carry drugs to target areas while minimizing systemic exposure.63,103,106,107 They are characterized as conventional, theranostic, PEGylated, and ligand-targeted,63 as well as by size, lamellarity, and surface charge (Figure 4).63,103,106,107 Liposome formulations are biocompatible and biodegradable,103 because they are made of mammalian cell membrane-like constituents and may permeate biofilms,108 and intracellular regions such as macrophages.109,110 Liposomes are appealing for drug delivery for a variety of reasons, including their pharmacological inactivity, ability to self-assemble, possession of a large aqueous center to carry large drug “payloads”, controlled drug release, and potential for enhanced pharmacokinetics and reduced toxicity;111 they are particularly useful for drugs that require cell membrane penetration.109,110 Liposomal behavior may be modified in vivo and liposomes can be directed to a specific area in the body. The stealth and targeted liposomes along with recent advancements in liposome design have resulted in a variety of liposomes, including immunological liposomes and stimulus-sensitive liposomes.106,112
![]() |
Figure 4 Structure of liposomes for drug delivery. (a) Conventional liposome, (b) Theranostic liposome, (c) PEGylated liposome, and (d) Ligand-targeted liposome. Reproduced Sercombe L, Veerati T, Moheimani F, Wu SY, Sood AK, Hua S. Advances and Challenges of Liposome Assisted Drug Delivery. Front Pharmacol. 2015 Dec 1;6:286. Copyright © 2015 Sercombe, Veerati, Moheimani, Wu, Sood and Hua. Creative Commons Attribution License (CC BY).63Abbreviation: PEG, polyethylene glycol. |
Conventional liposomes were the first liposomes generated used for therapeutic applications.113–115 To promote the stabilization of the liposomal bilayer, these liposomes can be made up of cationic, neutral, or anionic phospholipids in combination with CH.63,116 Nevertheless, there are still several challenges with this type of liposome, such as plasma instability, which results in a short blood circulation half-life. RES captures liposomes quickly and removes them from the bloodstream,115 The binding of opsonins to liposomes from serum proteins is the first signal for the elimination of liposomes. Conventional liposomes are recognized by opsonins as foreign particles and are therefore destroyed by mononuclear phagocytic system (MPS) phagocytes.22
To circumvent the limitations of regular liposomes, a second generation of liposomes was produced, resulting in the development of so-called stealthy, long-circulating, or PEGylated liposomes.117 The capacity to coat the surface of the liposome membrane with biocompatible hydrophilic polymer conjugates such as chitosan, PEG, and others, therefore increases the repulsive forces between serum components and liposomes, and is fundamental to the stealth method.118 As a result, macrophage immunogenicity and absorption are reduced, resulting in a longer circulatory half-life and lower toxicity of the encapsulated compound.119 Physical absorption of the polymer onto the liposomes surface, integration of PEG-lipid conjugates during liposome production, and covalent coupling of reactive groups to the surface are all methods for attaching PEG to the liposome membrane.115 Nonetheless, the high body bio-distribution of stealth liposomes is a serious drawback. As a result, drug encapsulation cannot be administered to a specific target location.9
Based on this constraint, ligand-targeted liposomes have been designed to transport drugs to specific tissues, allowing for more advanced and selective therapeutic activity.115 Target liposomes are additionally functionalized with glycoproteins, ligands, or polysaccharides for specific receptors such as antibodies, peptides, or small molecules, in addition to PEG surface modification.9,22 The ligand can target and bind to particular receptors overexpressed on diseased cell surfaces, with little off-target effects to healthy cells.113,120,121 Antibody-functionalized liposomes (immune liposomes) and stimuli sensitive liposomes have been proposed based on the previous technique concepts.106 Immune liposomes are made by chemically attaching antibodies or fragments of antibodies to the liposome surface, resulting in a more specific target antigen.122 When specific physicochemical or biological stimuli, such as pH, temperature, redox potential, enzyme and electrolyte concentrations, ultrasonic, electric, or magnetic fields, alter in a stimulus-sensitive liposomal system, the medication is released.123,124 The most prevalent stimuli responsive liposomes are temperature-sensitive and pH-sensitive liposomes.125,126 In addition to medication delivery, liposomes may be used for various purposes, such as making minor changes to their composition and charge.106 Cationic liposomes are a good example of a transfection vector utilized in gene therapy to carry genes. Gene encapsulation in liposomes allows nucleic acids to be protected from destruction during storage and circulation.115 Due to the significant potential of multifunctional liposomes, surface modification approaches have recently been researched for performing a combination of diverse functionalities, resulting in liposomes with a wide variety of functions.22 Another form of liposomes include theranostic liposomes, which can include imaging and therapeutic agents (diagnostic and therapy functions) within the liposome.63,127 Dual targeting liposomes is another example of liposome that involves having two different ligands.22
Targeting Strategies of Liposomes
The targeting strategy is an intense area for researchers to develop liposome formulations. One of the primordial functional properties of liposomes is specific targeting in drug delivery.107 As a result, high attention on specific sites puts emphasis on both the discovery of novel diagnostic tools and the improvement of therapeutic agent efficacy.128 Liposomes can currently be divided into two categories: active and passive targeting. Active tissue targeting is achieved with receptor-specific ligands on the surface of liposomes targeted for cellular uptake, while passive tissue targeting is accomplished primarily through the characteristics of the cancer or tumor vascular system, as illustrated in Figure 5.129 The passive targeting method has been used mostly in the field of oncology due to the pathophysiological characteristics of cancer.130 Liposomes passively target cancerous tissue or cancer cells by transporting and distributing them through the leaky tumor vasculature in the tumor interstitium via a molecular drive in the fluid.131 Therefore, passive targeted liposomes with a size of 10 to 500 nm can accumulate preferentially in the tumor and inflamed tissues through the permeability and retention effect RPE of the vascular system due to increased vasculature, vascular leaks, blood abnormalities, and dysfunction of the lymphatic vessels.132,133
Non-targeted liposomes prevent quick removal through the body’s defense systems, such as phagocyte absorption or clearance by mononuclear phagocyte system (MPS) cells.62 As a result, the production of stealth liposomes by PEG surface modification of liposomes can be an excellent example of applications in passive targeting techniques, and their circulation time can be increased.107 This technique also makes use of liposome-specific features, such as charge, which can promote selective targeting of cancer cells. Cationic liposomes are another example. By electrostatic interactions, this class of liposomes has been shown to bind the head of the phospholipid that has a negative charge, especially expressed on tumor endothelial cells.134 A targeting method based simply on the EPR effect is insufficient to prevent cytotoxic medication side effects. The variability of EPR effects in tumors, as well as their restriction to certain solid tumors, can have an impact on the efficacy of medications supplied by passive targeting.62,135 As a result, researchers studied the creation for new targeting methods with expanded functionality, such as active targeting.130
Paul Ehrlich, who coined the term “magic bullet” to describe the need for precise drug delivery inside the body, introduced the first concept of active targeting in 1906.129,136 Since then, researchers around the world have indeed been looking for a “miracle cure” which targets specific cells to make disease diagnosis and treatment easier.137 For improved liposome delivery systems, active targeting includes applying a targeting ligand to the liposome surface.22 Many target ligands, such as antibodies, peptides, nucleic acids, and whole proteins, as well as small molecules such as vitamins, are used for active targeting. For the identified target ligands, factors like the relative degree of overexpression or specific expression on the target, cellular uptake of the ligand-targeting formula, and degree of coverage of the target molecule have been considered.138,139 These ligands must also be chosen in such a way that they can bind to target cells while avoiding healthy cells.129 There are three methods to choose from when it comes to liposome functionalization. The first step is to attach the ligand to a lipid before mixing it with some other lipids in the liposomes. In the second method, the liposomes are activated immediately after preparation by targeting the required ligand (Table 3).140 The PEG spacer-modified lipid group, which is activated with the amine at the end or using thiol, carboxylic acid, or malamide groups, demonstrates the options available for this method.141 In another approach, it was suggested that functional lipids can be introduced into prefabricated liposomes. This approach is based on the spontaneous combination of activated lipids from the micelle phase to be prefabricated and even liposome-containing drugs. Derivation of the target molecule takes place in a separate step, as a way to prevent the activated lipids from interfering with other liposomal components, such as those compounds, which are present in the buffer.142
Active Targeting of Liposomal Anticancer Drugs
By using active targeting as a new approach, it is possible to overcome targeting barriers by adding a targeting moiety to the drug carrier’s surface. It is expected that the inventory used for the target materials will detect tumor-associated receptors or antigens. As a result, drugs are targeted to the site of action because drug uptake in the target cells increases while drug uptake in non-specific, healthy cells is reduced. In addition, through the receptor-mediated endocytosis process, some ligands can cause drug release from liposomes into target cells.143
As a result, the accumulation of drugs in the cells increases, and the overall effect of the therapies improve. Liposomes that target receptor internalization may also overcome drug resistance, at least in part.144 Many different methods are used for liposome active targeting. Target ligands attached to the surfaces of drug delivery devices are used in active targeting. These target ligands can bind to the receptors that are expressed in the target locations. The ligand must bind to a receptor that is overexpressed by tumor vasculature or cancer cells but not by healthy cells. Similarly, the particular receptors generated by tumor cells must be distributed uniformly. The antibody fragments, monoclonal antibodies, and non-antibody ligands are the proper ligands used for targeting purposes. The degree of ligand binding plays a crucial role in tumor penetration. To specifically target cells that are easily accessible, typically due to the tumor vasculature, high affinity binding appears to be preferred due to the dynamic flow environment of the bloodstream.145,146 Targeted liposomes for anticancer drugs will be discussed in the following section for different reported approaches (Table 4).
Receptor-Based Liposomal Anticancer Drug Targeting
Active targeting by cell surface receptors have been widely explored because many cancer cell types have tumor-specific receptors. Using receptor particular ligands or antibodies are some of the most common strategies to attack overexpressed cell surface receptors on cancer cells.147 The next sections summarize the most frequently activated receptors for liposomal medication delivery against cancer that are over expressed by tumor cells.
Folate Receptor-Based Liposomal Anticancer Drug Targeting
Folic acid is required in one carbon metabolic process and also plays a vital role in nucleotide base synthesis. Overexpression of the folate receptor-α isoform is seen in around 40% of human cancer cells. However, activated macrophages and hematological malignant cells have been observed to overexpress the folate receptor-β.148 Anticancer drugs are frequently targeted using folate-modified liposomes. Due to their site not on the apical surface of the epithelium but rather on the luminal side, targeting folate receptors (FRs) have been shown to be quite more efficient than targeting other receptors in reducing chemotherapeutic toxicities. The FR is a well-known tumor marker that binds with strong affinity to vitamin folic acid and folate-grafted drug carriers or folate drug conjugates, delivering them into cells via receptor-mediated endocytosis. When administered to multiple malignant cells by FR, doxorubicin (DOX) and daunorubicin (DUNO) have been demonstrated to have higher cytotoxicity.149,150 In another research, all-trans retinoic acid activation of FR was combined with folate-modified liposomes loaded with DOX for the treatment of acute myelogenous leukemia.151 Similarly, a diacid metabolite of norcantharidin, which is therapeutically effective against hepatocellular carcinoma and has been loaded into FR-modified polyethyleneglycolated liposomes, has been demonstrated to exhibit more cytotoxicity than plain PEGylated liposomes against the H22 hepatoma cell line. The researchers also observed that FR-modified liposomes were more effective in targeting tumors.152 Another study looked at the use of FR-modified liposomes to deliver Paclitaxel (PTX) to specific areas of the body.153
Liposomal Anticancer Medication Targeting Based on Transferrin Receptors
Transferrin, a glycoprotein, carries iron through the bloodstream and into cells by attaching to the transferrin receptor (TR) and lastly, receptor-mediated endocytosis internalizes the iron. TRs are important proteins that regulate iron homeostasis and cell development. TRs are overexpressed on the surfaces of many tumor cells as a result of their high iron needs and their rapid growth. TR targeted tailored anticancer medication delivery has been a significant technique. These receptors can have a dual impact when used for medication targeting. They carry medications into cells while suppressing their normal function, depriving the cells of iron, when they are targeted for drug delivery. It is also suggested that they have a role in the transport of iron to the brain. This also provides a novel medication targeting method for passing medicines past the blood-brain barrier.154,155 Transferrin coupled DOX-loaded liposomes bind and kill C6 glioma cells more effectively.156 A TR-targeted DOX-loaded liposomal technology, on the other hand, enhanced DOX intracellular absorption, pharmacokinetic profile, and biodistribution, leading to improved efficacy of treatment against liver cancer.157 Sharma et al developed a liposomal system that included transferring and poly-L-arginine. The strategy worked: the transferrin-modified liposomes targeted tumors, and poly-L-arginine increased cell penetration, allowing drugs to pass the blood-brain barrier endothelium.158 Dual functioning liposomes have also been described for penetrating the blood-brain barrier and targeting tumors. When tested using bEnd3 blood-brain barrier models, DOX liposomes supplemented with transferrin and folate were found to be useful for bioavailability in cells, P-glycoprotein (P-gp) expression, and drug transfer throughout the blood-brain barrier. The dual targeting DOX liposomes were found to be able to cross the blood brain barrier and distribute primarily in brain gliomas during in vivo tests. This effectiveness of the dual targeting method has been established in terms of tumor size reduction and extended survival time.159
Epidermal Growth Factor Receptor-Based Liposomal Anticancer Drug Targeting
The epidermal growth factor receptor (EGFR) is a protein that controls cell growth, differentiation, and repair in non-cancerous cells. However, in cancer cells, EGFR is often overactive, leading to uncontrolled cell growth and division.160 EGFR are overexpressed in a variety of solid tumors, including colon cancer, non-small cell lung cancer, ovarian, kidney, head, pancreas, neck, and prostate cancer, as well as breast cancer161,162 This makes it a promising target for therapeutic drug delivery. Proliferation, angiogenesis, and metastasis are only a few of the mechanisms that EGFR controls in cancer cells. EGFR-targeted immune system liposomes have been shown to increase intracellular DOX delivery to tumor cells, as well as enhance cytotoxicity against targeted tumors in xenograft animal models.163,164 The use of EGFR-targeted monoclonal antibodies in combination with liposomal systems has been studied extensively for signs of enhanced active targeted therapy. These antibodies, which act as targeting ligands on the surface of liposomes, have emerged as among the most effective drug candidate delivery techniques due to their high selectivity. At a DOX dose of 10 mM, cetuximab (an EGFR antibody)-biotin liposomes exhibited greater cytotoxicity for SKOV-3 cells than non-targeted biotin liposomes. On SKOV-3 cells, targeted liposomes demonstrated 22- to 38-fold greater binding than non-targeted liposomes.25 These data point to the efficacy of this method in the treatment of ovarian cancer.
Other Receptor-Based Liposomal Anticancer Drug Targeting
In addition to the receptor-based liposomal anticancer drug delivery described previously, other receptors have been found and are being used for specifically targeting anticancer drug delivery. Vasoactive intestinal peptide (VIP) receptors have been studied as therapeutic targets because they are abundant on the surfaces of tumor cells. In mice, VIP-coated PEG liposomes with radionuclides were found to be more effective at suppressing breast cancer than uncoated PEG liposomes with radionuclides.165 Immuno liposomes based on EGFR have also been reported for delivery to malignant cells overexpressing EGFR.163 Many tumor cells also overexpress hyaluronan receptors (HRs), which can be used to target liposomal anticancer treatments. When mitomycin C was encapsulated in long-circulating hyaluronan-targeted liposomes, it was more effective against tumors with HR overexpressed on their surfaces.166 Liposome-loaded cisplatin has been effectively used in vivo to prevent tumor formation and metastasis by selectively binding to chondroitin sulfate, which is overexpressed in many tumor cells.167 Furthermore, galactosylated liposomes have been shown to concentrate preferentially in parenchymal cells. They have been successfully used to transfer genes to these cells.168 Another method for liposome targeting is surface functionalization using peptide amphiphiles.169 Endothelial cells generate a variety of cell adhesion molecules (CAMs), which are necessary for the recruitment of leukocytes from the circulating blood to the endothelium following an inflammatory stimulus. CAMs are a natural target for anticancer therapy since they have a role in inflammatory disorders, including cancer.25 VCAM-1 is one of the CAMs that is overexpressed in tumor vasculature and is a promising target for anticancer medication delivery.170 Integrins, which are overexpressed in many cancers, aid in invasion and metastasis by allowing tumor cells to attach to the endothelium lining of blood arteries in various organs and tissues. Arginine-glycine-aspartate (RGD), a tripeptide, has a high integrin binding efficiency and has been shown to inhibit tumor cell adhesion and angiogenesis.171 Targeted medication delivery has made use of tumor tissue-specific expression of integrin receptors. Chen et al created an integrin-targeted liposomal method for DOX administration. The liposomes were covalently linked with cyclic RGD. In the U87MG cell line, the RGD coupled liposomal system showed a 2.5-fold greater cellular absorption of DOX than the unmodified liposomes. The liposomes were internalized via an integrin receptor-mediated endocytic routes, according to a competitive binding assay.172
Stimulus-Responsive Liposomal Anticancer Drug Targeting
Anticancer medication accumulation in malignant cells is not enough for successful treatment. Furthermore, liposome surface modification with PEG can prolong their circulation duration in the bloodstream while decreasing cellular internalization owing to steric hindrance. This problem can be handled by utilizing both external and internal cues. After liposomal accumulation at the target locations, these stimuli can disrupt the PEG protective layer.173 The notion of stimuli sensitivity is based on tumor microenvironmental features such as lower pH, greater temperature, and overexpression of various proteolytic enzymes.174 The stimuli-sensitive liposomes maintain their shape and physical characteristics throughout circulation. When exposed to a specific tumor microenvironment, they are engineered to undergo rapid changes (aggregation, disruption, and permeability) that result in drug release.25 The section that follows describes stimulus-responsive liposomal targeted delivery of anticancer medicines (Table 5).
Targeted pH-Responsive Liposomal Anticancer Delivery
PEGylation of liposomes has been shown to be a promising method for extending their systemic circulation time. It does, however, prevent emulation-induced drug release and intracellular drug delivery. Because of the glycolytic conversion of glucose to lactate in tumor cells, the tumor microenvironment has been demonstrated to be somewhat acidic (pH 6.0–7.0), lowering the pH value from that of normal tissues (pH 7.4) and has been exploited to develop pH-sensitive drug liposomes.175 The pH- sensitive breakdown of a liposomal carrier liberates encapsulated payloads in low-pH tissues like tumors, cell cytoplasm, or endosomes. Because of the low endosomal pH, liposomes containing pH-sensitive materials fuse with the endovascular outer layer during endocytosis and release their substances into the cytoplasm.176 Wang et al employed modified liposomes to construct a pH-sensitive drug delivery device.177 In a healthy environment (pH 7.4), the system releases slowly and gradually, but releases fast in a sub-acidic environment mimicking tumor tissue (pH 6.0). Tumor cells treated with pH-sensitive liposomes survived only 35% of the time after 48 hours, but normal cells survived 100% of the time, according to in vitro tumor cytotoxicity assays. Junior et al compared the tissue distribution of stealth pH-sensitive liposomes containing cisplatin to that of free cisplatin in solid Ehrlich tumor-bearing mice.178 The longer the stealth pH-sensitive liposomes circulated, the more medication that was released in the blood and accumulated in the tumor.
Temperature-Responsive Liposomal Anticancer Targeted Delivery
Tumor tissues typically display hyperthermia due to their fast metabolism, and as a result, increased temperatures are frequently observed in tumor locations, similar to the inflammatory response. Temperature-sensitive liposomes were motivated by this phenomenon. During pathological hyperthermia or external warming, temperature-sensitive liposomes release anticancer medications at tumor areas. A controlled device can also be used to heat solid tumors using an external energy source, like infrared irradiation. This is because temperature-sensitive liposomes are made up of lipids, which may change phase from gel to liquid at a certain temperature. Following that, when the temperature rises, the phospholipid double molecular chain becomes more disordered and active, resulting in drug release from the liposome vesicles.179 Temperature-triggered liposomal technologies are gaining popularity for targeted anticancer medication delivery. Also lipid temperature-sensitive liposomes have shown increased effectiveness in cancer-targeted medication delivery.180 This formulation is now being studied in Phase III clinical trials for hepatocellular carcinoma, as well as Phase II trials for breast cancer and colorectal liver metastases. Temperature-sensitive liposomes containing cis-platinum were developed by Kakinuma et al for the treatment of animals with brain glioma. Furthermore, the researchers discovered a much greater quantity of cis-platinum in brain tumor locations.181 Yatvin et al described temperature-sensitive liposomes that might release a hydrophilic medicine when heated to just a few degrees above physiological temperatures.182
Targeted Anticancer Enzyme-Responsive Liposomal Delivery
Over-expressed enzyme processes in the tumor environment, such as matrix metalloproteinases, have recently been used to improve the release of anticancer drugs from liposomes. Mura et al created enzyme-sensitive liposomes by using an MMP2- cleavable linker to connect a monoclonal antibody 2C5 to a PEG chain.183 Furthermore, tumors have been reported to over secrete phospholipase A2 (sPLA2), which can be exploited to induce medication release from enzyme-sensitive liposomes. Human sPLA2 activity was found to be particularly sensitive to liposome phospholipid acyl-chain length and negative surface charge density, resulting in drug release of enzyme-sensitive liposomes, according to Hansen et al.183
Targeting Physically Adsorbed Liposomal Anticancer Drugs
By adsorbing onto the membrane of cancer cells, physical adsorption-mediated liposomes, which employ cationic materials to modify the surface of liposomes into positively charged liposomes, can have a targeted impact. Cancer cells have electronegative cell membranes and electropositive liposomes can attach to them. Cationic liposomes can also accumulate in living cell mitochondria in response to mitochondrial membrane potential,179 after being taken up by cancer cells. Wang et al created mitochondrial targeting resveratrol liposomes by combining a dequalinium (DQA) molecule with polyethylene glycol stearoyl phosphatidylethanolamine (PEG2000- DSPE). The findings showed notable antitumor effects in both cancerous cells and drug-resistant cancerous cells.184 Ma et al also created mitochondrial targeting berberine liposomes by customizing DQA-PEG2000-DSPE-200.185 Berberine liposomes that target mitochondria might penetrate across cancer stem cell membranes and accumulate preferentially in cancer cell mitochondria. When coupled with PTX liposomes, mitochondrial targeting berberine liposomes greatly boosted anticancer efficiency in human breast cancer stem cell xenografts in nude mice.
Liposomal Anticancer Targeted Delivery Using a Magnetic Response
Magnetic liposomes are nanoparticles of maghemite (γ-Fe2O3) or magnetite (Fe3O4) that have been placed into liposomes (MLs). They are used to target medicines to specific sites using an external magnetic field as a stimulus.186 Such MLs have a wide range of applications in cancer. They are utilized in diagnostic applications such as MRI contrast agents. They are effective in the treatment of cancer using hyperthermia-based therapy. When an external magnetic field is supplied to MLs, they are used as heat mediators.187 Furthermore, they are employed in combination therapy with medicines for triggered release to provide a more safe and more efficient customized treatment.188 The toxicity of nano-carriers has long been a source of concern, limiting their application in medication delivery. When magnetic nanoparticles are encased in liposomes, the toxicity of the magnetic nanoparticles intended for targeted medication administration and diagnostic purposes is decreased or diminished.189
Magnetic nanoparticles are used in cancer treatments to increase medication accumulation at the tumor sites while minimizing the negative influence of chemotherapeutic medications on other normal tissues.190 Furthermore, such systems are more effective at imaging the entire site and may efficiently transport medicines throughout the cell membrane, thereby maintaining the required concentration levels of pharmaceuticals or diagnostic agents for the diagnosis of brain leukemia.191 Using magnetic gradients, 5-FU loaded MLs have been shown to improve biocompatibility and drug ability control. Surprisingly, the formulation was capable of exhibiting hyperthermia-triggered release of the drug, as well as an enhanced overall combination anticancer efficacy.192 Another research group looked at the co-delivery of glutamic acid-chelated γFe2O3 and methotrexate in the aqueous core of liposomes. The investigation yielded intriguing results. When exposed to an external magnetic field, the formulation produced an increase in the concentration of the medication deposited in the targeted tumor tissues compared to the findings demonstrated by the same formulation without the application of a magnetic field.193
Liposomal Anticancer Targeted Delivery Using an Ultrasound Response
Because of its non-invasiveness, deep penetration into the body, and permeability of blood tissue barriers, ultrasound-based targeted delivery has received a great deal of attention.194 Air is included in ultrasound-triggered drug release devices and is able to respond to ultrasonic stimulation to release a loaded material. The medication in ultrasonic-responsive liposomes can be released in line with ultrasound parameters. As a result, medication released from such stimuli sensitive liposomes may be tailored to meet the needs of the patient. If a medication burst release is desired, a high-intensity single ultrasonic pulse must be used. To induce sustained medication release, multiple low ultrasonic pulses are delivered over a long period.187 It has been demonstrated that ultrasound-responsive liposomal formulations improve cellular transfection by increasing membrane permeability during drug administration into the artery wall.195 A controlled release of DOX using an ultrasound-responsive liposomal formulation has been described. A perfluoropentane nanodroplet emulsion was used in the system, which was loaded with DPPC-based liposomes. The method was utilized for low-intensity DOX administration to the tumor. When the formulation was subjected to low-intensity ultrasound, it was able to release 80% of the loaded drug content, which was significantly higher than the basic emulsion of the same medication. When compared to free drugs, plain emulsions, and liposomal emulsions without ultrasound, the formulation also demonstrated greater anticancer efficacy of the medication against HeLa cells.196
Anticancer Delivery Using Light-Sensitive Liposomes
Because of their great spatial imaging resolution and the potential for targeted therapies, optical techniques for diagnosing and treating illnesses (such as skin wounds, inflammation, and cancer) are gaining scientific interest.197 Light in the near-infrared range has been discovered to penetrate deeply into tissues, making it useful in the treatment of cancer. Photodynamic therapy is now widely used to treat superficial cancers. Photosensitizing compounds including chlorins, porphycenes, porphyrin derivatives, and phthalocyanines, can produce radical oxygen species when exposed to light. As a result, they are utilized to sensitize and eradicate cancerous cells.198 Temoporfin, an amphiphilic molecule, is one of the most widely used photosensitizers in clinical practice. Foscan is an FDA approved treatment for advanced squamous cell carcinoma of the neck and head. Temoporfin, a photosensitizer, is included in the formulation, along with ethanol and propylene glycol. Fospeg and Foslip are two further liposomal formulations based on PEGylated liposomes and DPPC, respectively.199 Another liposomal formulation that is light and temperature sensitive was recently disclosed. It was made up of hollow gold nanospheres with DOX medication. When exposed to light, the formulation demonstrated light-triggered DOX release. When compared to control groups, this formulation based therapy demonstrated improved antitumor effectiveness.200
Therapeutic Applications of Liposomes
Liposomes have shown beneficial results as a drug delivery mechanism for several medications. Thus, rigorous studies of liposome use in medicine have led to the creation of diverse liposomal formulations for the control and management of a wide variety of illnesses, as well as a wide range of therapeutic applications such as fungal infections, analgesics, viral vaccines, photodynamic treatment, and cancer therapy. Because of alterations in pharmacokinetics and pharmacodynamics, encapsulating medications into liposomes increases their therapeutic effectiveness.201 Modification of in vivo drug behavior and reduction of drug toxicity in organisms are essential elements in developing an effective liposomal formulation. In clinical applications, liposomes are used to treat and diagnose cancer. The pH-sensitive liposomal nanocarriers have shown great promise for the delivery of chemotherapeutic drugs to tumor locations, greatly increasing their efficacy in cancer suppression. Liposomes are often PEGylated to increase their blood circulation time.202 Despite the fact that PEGylation reduces off-target toxicity in liposomes, PEGylated liposomes have low extravascular transport, limiting their survival advantage.203 Furthermore, PEGylated liposomes cannot escape endosomes after endocytosis.204
To circumvent the aforementioned constraints, pH-sensitive PEGylated liposomes have been developed. Liposomes were created by thin-film hydration in an experiment, and then SN25860, a small chemical having anticancer activity, was loaded to increase its accumulation at the tumor location. The liposomes demonstrated a high drug loading efficiency (7.0 0.2% w/w) and could boost SN25860 cytotoxicity by up to 21- to 24-fold. pH-sensitive PEGylated liposomes increased anticancer drug internalization and cellular uptake, and they entered cancer cells via clathrin-mediated endocytosis. According to Figure 6, liposomes can be constructed to release their therapeutic cargo prior to cellular uptake due to the intratumoral acidic pH, during cellular uptake by merging with a cell lipid membrane, or after endocytosis upon nanoparticle entry into the tumor location. In the latter situation, the pH-sensitive portion of the liposome degrades, resulting in medication release and targeted tumor suppression. The therapeutic potential of liposomes, however, is not limited to cancer therapy. Liposomes are regarded to be a very adaptable platform that may be used in a variety of research domains.205 The next section will focus on liposomes and their application to the most prevalent cancers (Table 6 and Figure 6).
![]() |
Figure 6 pH-sensitive liposomes exploit the lower pH in tumors versus normal cellular environments for drug release. They can release drugs before or after cell uptake, responding to the acidic tumor environment to target and treat tumors effectively. Adapted from Zhu L, Torchilin VP. Stimulus-responsive nanopreparations for tumor targeting. Integr Biol. 2013;5(1):96–107. © The Royal Society of Chemistry 2013.198 |
Breast Cancer
Breast cancer is classified clinically based on the presence of the estrogen receptor (ER), progesterone receptor (PR), and human epidermal growth factor receptor type 2 (HER-2). Breast cancer types that express receptors are treated with receptor-specific treatment. When the cells do not express any hormone receptor (triple negative), they become resistant, and chemotherapy is used to treat them.206 Monotherapy is currently ineffective because of the risk of chemo-resistance and tumor recurrence. Thus, the combination of anticancer medicines has enabled the treatment to have a synergistic impact.29 Nanoparticle-based drug delivery carriers have great promise as effective chemotherapeutic delivery methods for breast cancer.207 Targeting drug delivery systems (TDDS) are becoming more important for improving the therapeutic effect against cancer diseases,208 and several TDDS have been used in the research of therapies for breast cancer, including micelles, albumin, gold nanoparticles, and liposomes.209
Liposomes are highly researched and have been authorized for clinical use because of their greater biocompatibility, safety, and half-life in circulation compared to other formulations.210,211 Surface ligand functionalization to increase targeting is a key liposomal development theme. Gkionis et al reported the physicochemical features and cytotoxic effects of a new co-loaded liposomal formulation made utilizing two different preparative methods: the classic thin-film hydration approach and the alternative and speedier microfluidic technology.212 The size, zeta potential, stability, and drug loading capacity of liposomal formulations made using the microfluidic method were found to be equal to those created using the thin-film method. When generated utilizing microfluidic technology, lipid formulations were more homogeneous in size and shape, as well as more cytotoxic to the tested breast cancer cell types. Furthermore, the toxicity of all liposomal formulations was evaluated using a panel of human breast cancer cells (MCF-7, MDA-MB 231, and BT-474 cells) to determine the most powerful formulation per liposomal manufacturing technique and loaded chemical(s). Because of the delayed release of DOX from liposomes, the toxicity of DOX: umbelliprenin co-loaded liposomes were less than that of free DOX.212 Table 7 shows the liposomal medicine delivery strategy for breast cancer.
![]() |
Table 1 Comparative Analysis of Various Liposome Synthesis Methods Highlighting Their Advantages and Disadvantages |
![]() |
Table 2 Comprehensive Overview of Lipid Classification and Their Characterization Techniques, Highlighting the Advantages and Disadvantages of Each Method |
![]() |
Table 4 The Target Liposomes for Anticancer Drugs are Divided According to the Advantages and Disadvantages of Each of Them in the Listed Cases |
![]() |
Table 5 Overview of Targeted Anticancer Drug Delivery Methods via Liposomal Stimulation, Including pH, Temperature, Enzyme, Physical Adsorption, Magnetic, Ultrasound, and Light-Responsive Techniques |
![]() |
Table 6 Liposomes for Targeted Delivery of Chemotherapeutic Agents and Synergistic Tumor Targeting in Various Cancers |
![]() |
Table 7 Summary of Liposome Experimental Results for Breast Cancer |
As previously documented, the ICAM-1 antibody has been identified as a highly effective ligand for targeting TNBC in vivo. Through the conjugation of ICAM-1 antibodies to liposomes, we achieved specific delivery of encapsulated siRNA to TNBC tumors and cells. The engineered ICAM-Lcn2- liposomes were designed to hinder angiogenic activities in TNBC, as depicted in Figure 7A. The pH-responsive liposomal delivery system was composed of a blend of DOPC, DODAP, and DSPE-PEG-COOH. DODAP, incorporated into the liposome, responds to the acidic endosomal environment, enhancing its cationic character, facilitating fusion with the endosomal membrane, and delivering encapsulated siRNA to the cytoplasm (30,31). The 2 kDa PEG chain in DSPE-PEG-COOH demonstrated an ability to enhance liposome biocompatibility and circulation duration (32,33). The carboxyl group of DSPE-PEG-COOH serves as a site for conjugation with either the ICAM-1 antibody or the nonspecific immunoglobulin G (IgG). EDC/NHS chemistry was employed to covalently bond the carboxylic acid on DSPE-PEG-COOH to a primary amine group presented on the ICAM-1 antibody or the IgG. We examined the knockdown efficacy of ICAM-Lcn2-LPs by qRT-PCR. Lcn2 expression was measured after MDA-MB-231 cells were treated with PBS (control), free Lcn2 siRNA, ICAM-SCR-LPs, Lcn2-LIPO, IgG-Lcn2-LPs, and ICAM-Lcn2-LPs. As shown in Figure 7B, MDA-MB-231 cells treated with PBS (control), free Lcn2 siRNA and ICAM-SCR-LP demonstrated no change in their Lcn2 expression levels. Lcn2-LIPO and IgG-Lcn2-LP showed a reduction in Lcn2 of 41–56%. ICAM-Lcn2-LP was significantly more efficient than all other formulations, with a reduction in Lcn2 expression of 78.3 ± 1.7% (1.9-fold higher than IgG-Lcn2-LP). This was confirmed by immunoblot assays and densitometric analyses.
![]() |
Figure 7 Example of ICAM-1 liposomal targeting for breast cancer. (A) Schematic of the Lnc2-encapsulating ICAM-1-functionalized liposomes (ICAM-1-Lnc2-LP). (B) Relative Lnc2 protein levels in MDA-MB-231 cells after Lnc2 gene knockdown by the ICAM-1-Lnc2-LPs, accompanied by VEGF concentration in the conditioned media (CM) collected from the knockdown MDA-MB-231 cells. ***: Very significant, P value < 0.001, *: Significant, P value 0.01 to 0.05, NS: Not significant, P value ≥ 0.05. Notes: Reproducd from Nel J, Elkhoury K, Velot É, et al. Functionalized liposomes for targeted breast cancer drug delivery, Bioactive Materials, 24, 2023, 401-437. © 2022 The Authors. Publishing services by Elsevier B.V. on behalf of KeAi Communications Co. Ltd.262 |
Combined qRT-PCR and immunoblot results indicated that engineered TNBC-targeted, siRNA-encapsulating immunoliposomes, significantly inhibited the expression of a specific molecular target in TNBC cells at both mRNA and protein levels. During angiogenesis, tumor cells release VEGF to promote new vessel growth, crucial for developing a blood supply supporting tumor growth and metastasis. Previous findings revealed that Lcn2 stimulates neovascularization in breast cancer, and silencing Lcn2 reduces VEGF production. In this study using specific ELISA for VEGF, ICAM-Lcn2-LP treatment significantly reduced VEGF by 58% in MDA-MB-231 cell conditioned media, compared to reductions of 27% and 19% with Lcn2-LIPO and IgG-Lcn2-LPs, respectively. No change in VEGF concentration occurred with free Lcn2 siRNA or ICAM-SCR-LPs treatment. This demonstrates that simultaneous targeting of overexpressed ICAM-1 and silencing Lcn2 through ICAM-Lcn2-LP effectively suppresses VEGF secretion from MDA-MB-231 cells (Figure 7 and Table 7).
Lung Cancer
Lung cancer has become a major threat to human health, with an estimated 1.38 million cancer-related deaths in males and females in recent decades.228 Non-small cell lung cancer (NSCLC) accounts for approximately 85% of all lung cancer diagnoses, with small cell lung cancer (SCLC) accounting for the remaining 15%.229 Aside from initial lung carcinoma, lung metastases are likely to arise at a high incidence of 20–50% from other malignancies, such as colorectal cancer or breast cancer.230 Lung metastases always appear as numerous lesions, limiting the efficacy of surgery or radiation treatment. Furthermore, because of widespread drug dispersion, conventional systemic chemotherapy has a poor clinical prognosis. As a result, a unique effective treatment for this lethal illness is urgently required.231 Poor outcomes of lung cancer patients treated with traditional methods such as surgical resection, radiation, and chemotherapy have been reported.232 The bulk of chemotherapy is administered intravenously, resulting in significant adverse effects owing to systemic drug distribution. Furthermore, first-pass metabolism typically reduces the bioavailability of orally administered anticancer drugs.233 The cytotoxic effects of chemotherapeutic drugs against normal cells have been documented using dose response effects, resulting in patient frailty and mortality.233 As a result, scientific studies have focused on the targeted delivery of anticancer medicines.
The treatment of non-small cell lung cancer can be improved by using a targeted administration of chemotherapeutic agents to inhibit the key signaling pathways implicated in lung cancer. Then, by delivering anticancer medicines directly into the lungs, they can accumulate in tumor cells while reducing unwanted side effects.234 Price et al reported the combination of cationic liposomal hydroxycamptothecin (CLH) and 5-aminolevulinic acid (5- ALA) administration via intratracheal (i.t.) administration for the chemo-sonodynamic therapy of metastatic lung cancer. Hydroxycamptothecin and a lipid combination of soybean lecithin/cholesterol/octadecylamine were used to make cationic liposomal hydroxycamptothecin with a film technique. For sonosensitizer accumulation, ie, protoporphyrin IX, the metabolite of 5-ALA, an optimal pre-incubation period of 5-ALA with tumor cells before ultrasonic exposure was found at 4 h. In vitro investigations revealed that chemo-sonodynamic therapy had greater cytotoxicity than other therapies such as intratracheal cationic liposomal hydroxycamptothecin, intravenous cationic liposomal hydroxycamptothecin, and sonodynamic therapy alone.
The combination of pulmonary administration with chemo-sonodynamic therapy had the greatest anticancer impact on metastatic lung tumor-bearing mice, as determined by tumor appearance and pathological sections.231 Chemo-sonodynamic treatment via primary anticancer mechanisms improved apoptosis of cancer cells and increased the generation of ROS, as well as the combination of chemotherapy and sonodynamic therapy.231 A potential method for treating lung cancer is the pulmonary administration of chemotherapeutics and sonosensitizers.231
A drug delivery system (DDS) overcomes the limitations of liposomes and polymeric nanoparticles by introducing hybrid NPs. These self-assembled nanoscaled vehicles (<1000 nm) offer multiple benefits in cancer treatment, including enhanced sustained release, targeted delivery, biocompatibility, prolonged circulation time, and efficient surface modifications with ligands. Hybrid NPs consist of three primary components: (i) a hydrophobic polymeric core incorporating lipophilic drugs, (ii) a lipid layer serving as a biocompatible shell and enhancing drug retention within the polymeric core, and (iii) a hydrophilic PEG stealth layer surrounding the lipid shell. The lipid-PEG shell is crucial for enhancing stability, and PEG offers functional groups for additional modification with targeting ligands.25,75 Figure 8 visually depicts the targeting of tumor cells using such hybrid NPs in the context of nanoparticulate formulations designed for lung cancer therapy.229
![]() |
Figure 8 A schematic illustrating the targeting of tumor cells with hybrid NPs for lung cancer therapy. |
Polymeric materials (such as PLGA, dextran, albumin, and PCL) are commonly utilized as the core of hybrid NPs due to their non-toxic and biodegradable nature. The lipid shell of these NPs is typically composed of cationic, anionic, or neutral phospholipids.76 While small interference RNA (siRNA) is a crucial advancement in cancer diagnosis and treatment, its lack of specific targeting due to instability and insufficient bio distribution is a challenge.8,77 To address this, hybrid NPs have been investigated for delivering siRNA, employing PEGylated polyethyleneimine (PEI) with an Arg-Gly-Asp peptide ligand to inhibit vascular endothelial growth factor receptor-2. This approach enables tissue-specific and gene pathway-specific targeting of siRNA.78 In a study by Lakshmikuttyamma et al (2014), hybrid NPs were employed to deliver Kirsten rat sarcoma (KRAS) siRNA to A549 lung adenocarcinoma cells. Human IgG antibodies were also attached to the NPs to prevent immune activation associated with most NPs. The siRNA was efficiently delivered to the mutated KRAS cell line without being captured by the RES and without immune response activation.79 Furthermore, the overexpression of neurotensin receptor 1 (NTSR1) in most non-small cell lung cancers (NSCLC) was targeted by modifying hybrid NPs with an anti-NTSR1 monoclonal antibody, facilitating the efficient delivery of anti-mutant KRAS siRNA to NTSR1-overexpressing tumor cells.80 Table 8 summarizes the key experimental outcomes of liposomes.
![]() |
Table 8 Summary of Liposome Experiments Results for Lung Cancer |
Prostate Cancer
Prostate cancer (PC) has been the most frequently diagnosed cancer and the second leading cause of cancer death globally.254,255 Anti-androgenic medications are beneficial in hormone-dependent prostate cancer, but tumors develop a hormone-refractory phenotype that is resistant to chemotherapy. This stresses the significance of establishing innovative therapeutic techniques for the treatment of PC.256 Cancer stem cells (CSCs) initiate tumors, undergo epithelial-mesenchymal transition, and develop chemo resistance, culminating in metastatic dissemination.257 As a result, for improving the treatment, a combination pharmacotherapy is recommended to target cancer stem cells with traditional cytotoxic agents capable of effectively eradicating cancer stem cells and bulk tumor cells at the same time,258 and could potentially be applied to the field of cancer stem cells. Aside from delaying or suppressing cancer adaptability, mutation, and development, a combination therapy decreases individual medicine dosage, resulting in fewer adverse effects.259,260 Liposomes are an excellent candidate for combinational chemotherapy because of their capacity to carry a wide range of medicines, high surface-to-volume ratios, and adjustable surfaces for targeting.227 VyxeosTM (daunorubicin and cytarabine) is a liposomal injectable authorized by the FDA for the combinatorial treatment for acute myeloid leukemia.261 This opens the door to future effective and safe cancer treatments provided by nanomedicine-based synergistic medication combinations.263
Kroon et al showed that the increased expression of COX-2 and Glut-1 proteins is key in the initiation and progression of prostate cancer via altering related signaling pathways.264 The combined action of these medicines causes prostate cancer cells to be more selectively induced to apoptosis than normal fibroblast cells. According to a mechanistic study, the major mechanisms behind the inhibition of prostate cancer cells include increased reactive oxygen species (ROS) production and a reduction in cellular glutathione concentration, as well as inhibition of COX-2 synthesis and Glut-1 receptors. Although the combination of celecoxib and genistein reduced prostate cancer cell growth by up to 90%, there was no significant damage to normal fibroblast cells, implying that perhaps the dosage of genistein and celecoxib required to eliminate prostate cancer cells seems to be non-toxic to healthy cells. A nanoliposomal formulation of celecoxib and genistein was shown to produce ROS, significantly decrease cellular glutathione concentration, and inhibit glucose uptake. When these events occur together, they successfully limit the growth of prostate cancer cells. Although the created nano-liposomes demonstrated promising in vitro results and thus have the possibility to be further improved for cancer therapeutic applications, more extensive research is needed to realize the full potential of this composition or the treatment of prostate and other types of cancer. Table 9 presents various liposomal medication delivery methods for prostate cancer.
![]() |
Table 9 Summary of Liposome Experimental Results for Prostate Cancer |
Colorectal Cancer
In normal settings, human cells multiply and divide to generate new cells as the body needs. When cells get old or injured, they die, and new cells replace them.282 Cancer is caused by the failure of this mechanism. Cancer is a condition in which some of the body’s cells proliferate and spread into the tissues around them.283 There are trillions of cells in the human body. As a result, cancer can begin everywhere in the body. When colorectal cancer is in its early stages, there may be no symptoms. In certain situations, diarrhea, constipation, blood in the stool, rectum bleeding, severe gas, stomach cramps and abdominal discomfort may be symptoms of colon cancer (CC). In the last 25 years, much progress has been achieved in understanding the molecular and biological aspects as well as stages connected with colon carcinogenesis. It has resulted in more reasonable and successful therapeutic approaches to colorectal cancer (CRC) therapy.284 Traditional adenomas along with the traditional adenoma-to-carcinoma sequence and serrated adenomas via two different routes are both precursor polyps for CC.285 The progression to CC is a multistep process; typical adenomas are caused by mutations in the APC gene. Serrated adenomas have an unusual main genetic abnormality.286 Liposomes have become the most commonly employed nanocarriers for targeted medication delivery for CC.287 Liposomes have several advantages, including biodegradability, biocompatibility, low toxicity, and the ability to entrap both lipophilic and hydrophilic medicines.288,289 However, because hydrophilic medicines are more soluble in water and dissolve in the aqueous layer of liposome synthesis, formulation design and processing of hydrophilic pharmaceutical enclosures into liposomes is a significant challenge.290
Lip-F1 (non-PEGylated liposomes) and Lip-F2 (PEGylated liposomes) substances were created for in vivo as well as in vitro investigations, with interferon-gamma (IFN-) included measuring the impact on antitumor and macrophage activities. The liposomal substances LIP-F1 and LIP-F2 were 120 and 135 nm in size. LIP-F1 and LIP-F2 efficiency was 52.79 and 49.2%, respectively. These findings demonstrated that treatment reactions act as a moderator. LIP-F1 and LIP-F2 efficiency was 52.79 and 49.2%. These results indicated that the treatment reactions influenced by IFN-liposomes in the CRC animal studies were related directly to the lethal effects of IFN-liposomes on a C26 malignant cell line, which corresponded with the polarization of TAMs to exhibit antitumoral activity. IFN-produced PEGylated liposomes showed significant anticancer efficacy leading to enhanced drug delivery to the immune system and antitumor immune responses.291 The goal of the study was to create and design 5-FU, including using tailored liposomes, to increase the drug’s effectiveness and safety and to employ folic acid as a target ligand. CT26, HT-29, HeLa, Caco-2, and MCF 7 cell lines were tested in vitro for cytotoxicity from the formulations using the MTT assay; results showed that the targeted liposomes caused cell death via ROS. After giving the medication and the targeted 5-FU liposome, inhibition tests were performed, revealing that the optimized formulation’s EE was 39.71%. The liposomes had a particle size of 174 nm, in spherical form, and a Differential Scanning Calorimetry (DSC) study demonstrated that the drug was present in the amorphous state in liposomes. MTT findings showed that the targeted liposomes were more cytotoxic than 5-FU and liposomal 5-FU. In vivo, folate liposomal 5-FU suppressed tumors more effectively than the free medicine and control groups (p < 0.05). Please note that the control groups were the groups that received either free 5-FU and no treatment at all. Furthermore, as compared to the control group, the folate-liposomal 5-FU therapy group showed lower cell density in tumor tissue. As a result, folic acid-targeted liposomes might be the next drug carrier for selective drug delivery in CC cells.292 The benefits of using nano- drug delivery systems include increased bioavailability by decreasing the dose and targeting the target cells with anticancer medications to decrease adverse effects. 5-FU in NPs provided an appropriate and safe treatment for CC with decreased side effects and dosage.293 In vitro, folate- targeted liposomal 5-FU enhanced the absorption in B16F10 cells 11 times more than non-targeted liposomes; folate-liposomal 5-FU had a greater tumor inhibitory impact than free 5-FU.294 A mouse model was used to detect ROS by accumulating liposomes in wounded colon regions, administering FL-labeled liposomes, and examining luminescence. Table 10 summarizes the applications of liposome medication delivery for colorectal cancer.
![]() |
Table 10 Summary of Liposome Experimental Results for Colorectal Cancer |
This passage discusses the use of liposomes as nanocarriers for the co-delivery of hydrophilic (OHP) and lipophilic (CUR) drugs, with a focus on enhancing their effectiveness through active targeting. Liposomes are lipid-based structures that can release drugs synchronously, reduce drug accumulation in tumors, and minimize toxicity to non-cancerous cells.10,11 The hydrophilic core and lipid bilayer of liposomes make them suitable for encapsulating both types of drugs. Active targeting is employed to enhance the effectiveness of the liposomal nanocarriers by exploiting interactions between receptors on cancer cell surfaces and targeting groups on liposomes. Hyaluronic acid (HA) has been introduced as a targeting ligand due to its interaction with overexpressed HA receptors (CD44 and RHAMM) on cancer cells, especially in colorectal carcinoma. To address the potential degradation of liposomes in the gastrointestinal tract (GIT), the study proposed entrapping liposomes in alginate beads. These beads, coated with pH-sensitive polymer eudragit S-100 (ES-100), remained intact in the upper GIT and reached the colon. Upon dissolution of the coating in the ileocecal region, uncoated alginate beads entered the colon and underwent biodegradation, releasing surface-modified liposomes. These HA-conjugated liposomes have a higher affinity for HA receptors on colon cancer cells, achieving cell-specific targeting. The study’s objective was to develop HA-anchored liposomes co-loaded with OHP and CUR, entrapped in ES-100 coated alginate beads, for specific delivery to colon cancer cells. The therapeutic efficacy and biocompatibility of these co-loaded liposomes were assessed using in vitro cytotoxic activity on OHP-resistant HT-29 cancer cell lines. Figure 9 shows a schematic of colon-specific targeting of eudragit coated bead encapsulating liposomes.290,308
![]() |
Figure 9 Schematic representation of colon-specific targeting of eudragit coated bead encapsulating liposomes. Notes: Reproduced from Tiwari A, Gajbhiye V, Jain A, et al. Hyaluronic acid functionalized liposomes embedded in biodegradable beads for duo drugs delivery to oxaliplatin-resistant colon cancer. J Drug Delivery Sci Technol. 2022;77:103891. © 2022 Elsevier B.V. All rights reserved.309 |
Liposomal Formulations in the Clinic
Due to their appropriate size, biocompatibility, biodegradability, low toxicity, and immunogenicity, liposomes have proven to be one of the most mature nanomedicine platforms currently used in clinical settings.309 Table 11 presents comprehensive information about different liposomal products, including details about their liposome composition, intended indications, and the preferred route of administration for clinical trials. Several of these liposomes have even been approved by the FDA for the treatment of cancer. Furthermore, the protective function of the liposomal encapsulation can lessen negative reactions, improve absorption, and ultimately improve the therapeutic impact of medications.
![]() |
Table 11 Liposomal Formulations in Clinical Trials. |
Future Research and Development
The future of liposomal nanomedicine is set to undergo transformative advancements, particularly through the incorporation of immunotherapeutic agents. Researchers are exploring the potential of liposomes not just as passive drug carriers but as active “immuno-modulatory hubs” capable of delivering antigens, adjuvants, and gene-editing tools. This paradigm shift in therapeutic approaches aims to evoke and regulate precise immune responses against cancer and infectious diseases. In addition to their therapeutic functions, these innovative liposomes are envisioned to possess theranostic capabilities, enabling simultaneous treatment and monitoring of disease progression. By engineering liposomes to track their accumulation in specific tissues and provide real-time imaging of immune cell dynamics, personalized medicine can be significantly enhanced. Furthermore, the development of “smart” liposomes that respond dynamically to the disease microenvironment holds immense promise. These liposomes could trigger drug release or immune activation based on specific environmental cues, thereby improving treatment precision and efficacy. The COVID-19 pandemic has underscored the need for effective liposomal formulations in vaccines and antiviral agents, revealing challenges such as rapid clearance and suboptimal biodistribution that must be addressed through nanotechnology-based solutions. Incorporating artificial intelligence (AI) and machine learning (ML) into the design of liposomal systems offers a pathway to optimize formulations, predict immune responses, and tailor treatment regimens to individual patients. This integration could lead to a revolutionary era in nanomedicine, enabling researchers to design liposomes that evade immune detection while delivering therapies directly to targeted tissues. Ultimately, the future of liposomal nanomedicines lies in their ability to evolve into sophisticated platforms that not only deliver drugs effectively but also actively participate in modulating immune responses. This advancement could lead to safer and more effective therapeutic options for a variety of diseases, paving the way for a new frontier in personalized medicine and immunology.
Conclusions
This manuscript explores the extensive potential of liposomal drug delivery systems, particularly in the realm of anticancer therapies. It highlights the advancements made in liposome design, characterization, and targeting strategies, emphasizing how these innovations have enhanced the precision and efficacy of drug delivery to cancer cells. By addressing challenges such as biodistribution, drug loading, and targeted delivery, liposomal formulations have shown great promise in improving therapeutic outcomes for various cancers, including breast, lung, prostate, and colorectal cancers.
The discussion extends to recent innovations, such as stimulus-responsive liposomes that react to environmental triggers like pH, temperature, or enzymes, thereby allowing for more controlled and effective drug release. Moreover, liposomes have moved beyond cancer treatment to applications in infectious diseases, as evidenced by their role in COVID-19 vaccines. These nanocarriers are increasingly being designed to modulate immune responses, with the potential to evolve into immuno-modulatory platforms capable of orchestrating targeted immune reactions.
The manuscript also emphasizes the future direction of liposomal nanomedicines, where integrating artificial intelligence and machine learning could optimize design and therapeutic outcomes. By coupling drug delivery with real-time diagnostic capabilities, liposomes have the potential to revolutionize personalized medicine. Ultimately, the conclusion underscores the vital role liposomal drug delivery systems will continue to play in both cancer therapy and broader medical applications, pushing the boundaries of precision medicine and nanotechnology.
Acknowledgments
The authors are thankful to the Department of Chemistry of the Universiti Malaya for the facilities provided throughout this research.
Disclosure
The authors declare no conflicts of interest in this work.
References
1. Bray F, Laversanne M, Weiderpass E, Soerjomataram I. The ever‐increasing importance of cancer as a leading cause of premature death worldwide. Cancer. 2021;127(16):3029–3030. doi:10.1002/cncr.33587
2. Chen S, Cao Z, Prettner K, et al. Estimates and projections of the global economic cost of 29 cancers in 204 countries and territories from 2020 to 2050. JAMA Oncol. 2023;9(4):465–472. doi:10.1001/jamaoncol.2022.7826
3. Guida F, Kidman R, Ferlay J, et al. Global and regional estimates of orphans attributed to maternal cancer mortality in 2020. Nat Med. 2022;28(12):2563–2572. doi:10.1038/s41591-022-02109-2
4. Bray F, Laversanne M, Sung H, et al. Global cancer statistics 2022: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. Ca a Cancer J Clinicians. 2024;74(3):229–263. doi:10.3322/caac.21834
5. Bardania H, Tarvirdipour S, Dorkoosh F. Liposome-targeted delivery for highly potent drugs. Artif Cells Nanomed Biotechnol. 2017;45(8):1478–1489. doi:10.1080/21691401.2017.1290647
6. Fidan Y, Muçaj S, Timur SS, Gürsoy RN. Recent advances in liposome-based targeted cancer therapy. J Liposome Res. 2024;34(2):316–334. doi:10.1080/08982104.2023.2268710
7. Morales-Cruz M, Delgado Y, Castillo B, et al. Smart targeting to improve cancer therapeutics. Drug Des Devel Ther. 2019;Volume 13:3753–3772. doi:10.2147/DDDT.S219489
8. Crommelin DJ, Storm G. Liposomes: from the bench to the bed. J Liposome Res. 2003;13(1):33–36. doi:10.1081/LPR-120017488
9. Torchilin VP. Recent advances with liposomes as pharmaceutical carriers. Nat Rev Drug Discov. 2005;4(2):145–160. doi:10.1038/nrd1632
10. AlSawaftah N, Pitt WG, Husseini GA. Dual-targeting and stimuli-triggered liposomal drug delivery in cancer treatment. ACS Pharmacol Transl Sci. 2021;4(3):1028–1049. doi:10.1021/acsptsci.1c00066
11. Bangham AD, Standish MM, Watkins JC. Diffusion of univalent ions across the lamellae of swollen phospholipids. J Mol Biol. 1965;13(1):238–IN27. doi:10.1016/S0022-2836(65)80093-6
12. Deamer DW. From “banghasomes” to liposomes: a memoir of Alec Bangham, 1921-2010. FASEB J. 2010;24(5):1308. doi:10.1096/fj.10-0503
13. Jain A, Kumari R, Tiwari A, et al. Nanocarrier based advances in drug delivery to tumor: an overview. Curr Drug Target. 2018;19(13):1498–1518. doi:10.2174/1389450119666180131105822
14. Alavi M, Karimi N, Safaei M. Application of various types of liposomes in drug delivery systems. Adv Pharm Bull. 2017;7(1):3. doi:10.15171/apb.2017.002
15. Akbarzadeh A, Rezaei-Sadabady R, Davaran S, et al. Liposome: classification, preparation, and applications. Nanoscale Res Lett. 2013;8(1):1–9. doi:10.1186/1556-276X-8-102
16. O’Brien ME, Wigler N, Inbar M, et al. Reduced cardiotoxicity and comparable efficacy in a phase IIItrial of pegylated liposomal doxorubicin HCl (CAELYX™/Doxil®) versus conventional doxorubicin forfirst-line treatment of metastatic breast cancer. Ann Oncol. 2004;15(3):440–449. doi:10.1093/annonc/mdh097
17. Gabizon A, Papahadjopoulos D. Liposome formulations with prolonged circulation time in blood and enhanced uptake by tumors. Proc Natl Acad Sci. 1988;85(18):6949–6953. doi:10.1073/pnas.85.18.6949
18. Lee JH, Yeo Y. Controlled drug release from pharmaceutical nanocarriers. Chem Eng Sci. 2015;125:75–84. doi:10.1016/j.ces.2014.08.046
19. Huwyler J, Drewe J, Krähenbühl S. Tumor targeting using liposomal antineoplastic drugs. Int j Nanomed. 2008;3(1):21–29. doi:10.2147/IJN.S1253
20. Maeda H, Wu J, Sawa T, Matsumura Y, Hori K. Tumor vascular permeability and the EPR effect in macromolecular therapeutics: a review. J Controlled Release. 2000;65(1–2):271–284. doi:10.1016/S0168-3659(99)00248-5
21. Yan W, Leung SS, To KK. Updates on the use of liposomes for active tumor targeting in cancer therapy. Nanomedicine. 2020;15(3):303–318. doi:10.2217/nnm-2019-0308
22. Riaz MK, Riaz MA, Zhang X, et al. Surface functionalization and targeting strategies of liposomes in solid tumor therapy: a review. Int J Mol Sci. 2018;19(1):195. doi:10.3390/ijms19010195
23. Basile L, Pignatello R, Passirani C. Active targeting strategies for anticancer drug nanocarriers. Curr Drug Delivery. 2012;9(3):255–268. doi:10.2174/156720112800389089
24. Rumjanek VM, Trindade GS, Wagner-Souza K, et al. Multidrug resistance in tumour cells: characterisation of the multidrug resistant cell line K562-Lucena 1. Anais da Academia Brasileira de Ciências. 2001;73(1):57–69. doi:10.1590/S0001-37652001000100007
25. Deshpande PP, Biswas S, Torchilin VP. Current trends in the use of liposomes for tumor targeting. Nanomedicine. 2013;8(9):1509–1528. doi:10.2217/nnm.13.118
26. Senapati S, Mahanta AK, Kumar S, Maiti P. Controlled drug delivery vehicles for cancer treatment and their performance. Signal Transduc Target Ther. 2018;3(1):1–19. doi:10.1038/s41392-017-0004-3
27. Laouini A, Jaafar-Maalej C, Limayem-Blouza I, Sfar S, Charcosset C, Fessi H. Preparation, characterization and applications of liposomes: state of the art. J Colloid Sci Biotechnol. 2012;1(2):147–168. doi:10.1166/jcsb.2012.1020
28. Shazly G, Nawroth T, Langguth P. Comparison of dialysis and dispersion methods for in vitro release determination of drugs from multilamellar liposomes. Dissolution Technol. 2008;15(2):7. doi:10.14227/DT150208P7
29. Novais M, Gomes E, Miranda M, et al. Liposomes co-encapsulating doxorubicin and glucoevatromonoside derivative induce synergic cytotoxic response against breast cancer cell lines. Biomed Pharmacother. 2021;136:111123. doi:10.1016/j.biopha.2020.111123
30. Cabaleiro D, Pastoriza-Gallego MJ, Gracia-Fernández C, Piñeiro MM, Lugo L. Rheological and volumetric properties of TiO2-ethylene glycol nanofluids. Nanoscale Res Lett. 2013;8(1):1–13. doi:10.1186/1556-276X-8-286
31. Emami S, Azadmard-Damirchi S, Peighambardoust SH, Valizadeh H, Hesari J. Liposomes as carrier vehicles for functional compounds in food sector. J Exp Nanosci. 2016;11(9):737–759. doi:10.1080/17458080.2016.1148273
32. Maherani B, Arab-Tehrany ER, Mozafari M, Gaiani C, Linder M. Liposomes: a review of manufacturing techniques and targeting strategies. Curr Nanosci. 2011;7(3):436–452. doi:10.2174/157341311795542453
33. Pattni BS, Chupin VV, Torchilin VP. New developments in liposomal drug delivery. Chem Rev. 2015;115(19):10938–10966. doi:10.1021/acs.chemrev.5b00046
34. Karn PR, Cho W, Hwang S-J. Liposomal drug products and recent advances in the synthesis of supercritical fluid-mediated liposomes. Nanomedicine. 2013;8(9):1529–1548. doi:10.2217/nnm.13.131
35. Meure LA, Foster NR, Dehghani F. Conventional and dense gas techniques for the production of liposomes: a review. AAPS Pharm Sci Tech. 2008;9(3):798–809. doi:10.1208/s12249-008-9097-x
36. Nkanga CI, Bapolisi AM, Okafor NI, Krause RWM. General perception of liposomes: formation, manufacturing and applications. Liposomes-Advances and Perspectives. 2019.
37. Guimarães D, Cavaco-Paulo A, Nogueira E. Design of liposomes as drug delivery system for therapeutic applications. Int J Pharm. 2021;601:120571. doi:10.1016/j.ijpharm.2021.120571
38. Batzri S, Korn ED. Single bilayer liposomes prepared without sonication. Biochimica et Biophysica Acta (BBA)-Biomembranes. 1973;298(4):1015–1019. doi:10.1016/0005-2736(73)90408-2
39. William B, Noemie P, Brigitte E, Geraldine P. Supercritical fluid methods: an alternative to conventional methods to prepare liposomes. Chem Eng J. 2020;383:123106. doi:10.1016/j.cej.2019.123106
40. Justo OR, Moraes AM. Economical feasibility evaluation of an ethanol injection liposome production plant. Chem Eng Technol. 2010;33(1):15–20. doi:10.1002/ceat.200800502
41. Marasini N, Ghaffar KA, Skwarczynski M, Toth I. Liposomes as a vaccine delivery system. Micro and Nanotechnology in Vaccine Development. 2017;2017:221–239.
42. Wagner A, Vorauer-Uhl K. Liposome technology for industrial purposes. J Drug Delivery. 2011;2011:1–9. doi:10.1155/2011/591325
43. Szoka Jr F, Papahadjopoulos D. Procedure for preparation of liposomes with large internal aqueous space and high capture by reverse-phase evaporation. Proc Natl Acad Sci. 1978;75(9):4194–4198. doi:10.1073/pnas.75.9.4194
44. Gad SC. Pharmaceutical Manufacturing Handbook: Production and Processes. John Wiley & Sons; 2008.
45. Schubert R. Liposome preparation by detergent removal. Methods Enzymol. 2003;367:46–70.
46. Sundar S, Tirumkudulu MS. Synthesis of sub-100-nm liposomes via hydration in a packed bed of colloidal particles. Ind Eng Chem Res. 2014;53(1):198–205. doi:10.1021/ie402567p
47. Shew R, Deamer D. A novel method for encapsulation of macromolecules in liposomes. Biochimica et Biophysica Acta (BBA)-Biomembranes. 1985;816(1):1–8. doi:10.1016/0005-2736(85)90386-4
48. Large DE, Abdelmessih RG, Fink EA, Auguste DT. Liposome composition in drug delivery design, synthesis, characterization, and clinical application. Adv Drug Delivery Rev. 2021;176:113851. doi:10.1016/j.addr.2021.113851
49. Cruz LJ, Tacken PJ, Rueda F, Domingo JC, Albericio F, Figdor CG. Targeting nanoparticles to dendritic cells for immunotherapy. Methods Enzymol. 2012;509:143–163.
50. Pick U. Liposomes with a large trapping capacity prepared by freezing and thawing of sonicated phospholipid mixtures. Arch Biochem Biophys. 1981;212(1):186–194. doi:10.1016/0003-9861(81)90358-1
51. MacDonald RC, Jones FD, Qui R. Fragmentation into small vesicles of dioleoylphosphatidylcholine bilayers during freezing and thawing. Biochimica et Biophysica Acta (BBA)-Biomembranes. 1994;1191(2):362–370. doi:10.1016/0005-2736(94)90187-2
52. Xu X, Khan MA, Burgess DJ. Predicting hydrophilic drug encapsulation inside unilamellar liposomes. Int J Pharm. 2012;423(2):410–418. doi:10.1016/j.ijpharm.2011.12.019
53. Aliño S, García M, Lejarreta M, Bobadilla M, Pérez-Yarza G, Unda F. Trapping Drug Efficiency in Liposomes Produced by Extrusion of Freeze-Thaw Multilamellar Vesicles. Portland Press Ltd.; 1989.
54. Traïkia M, Warschawski DE, Recouvreur M, Cartaud J, Devaux PF. Formation of unilamellar vesicles by repetitive freeze-thaw cycles: characterization by electron microscopy and 31P-nuclear magnetic resonance. Eur Biophys J. 2000;29(3):184–195. doi:10.1007/s002490000077
55. Hope M, Bally M, Mayer L, Janoff A, Cullis P. Generation of multilamellar and unilamellar phospholipid vesicles. Chem Phys Lipids. 1986;40(2–4):89–107. doi:10.1016/0009-3084(86)90065-4
56. Xu X, Costa A, Burgess DJ. Protein encapsulation in unilamellar liposomes: high encapsulation efficiency and a novel technique to assess lipid-protein interaction. Pharm Res. 2012;29(7):1919–1931. doi:10.1007/s11095-012-0720-x
57. Hwang SY, Kim HK, Choo J, Seong GH, Hien TBD, Lee E. Effects of operating parameters on the efficiency of liposomal encapsulation of enzymes. Colloids Surf B. 2012;94:296–303. doi:10.1016/j.colsurfb.2012.02.008
58. Kraft JC, Freeling JP, Wang Z, Ho RJ. Emerging research and clinical development trends of liposome and lipid nanoparticle drug delivery systems. J Pharmaceut Sci. 2014;103(1):29–52. doi:10.1002/jps.23773
59. Tejera-Garcia R, Ranjan S, Zamotin V, Sood R, Kinnunen PK. Making unilamellar liposomes using focused ultrasound. Langmuir. 2011;27(16):10088–10097. doi:10.1021/la201708x
60. Olson F, Hunt C, Szoka F, Vail W, Papahadjopoulos D. Preparation of liposomes of defined size distribution by extrusion through polycarbonate membranes. Biochimica et Biophysica Acta (BBA)-Biomembranes. 1979;557(1):9–23. doi:10.1016/0005-2736(79)90085-3
61. Çağdaş M, Sezer AD, Bucak S. Liposomes as potential drug carrier systems for drug delivery. Appl Nanotechnol Drug Del. 2014;1:1–50.
62. Elsana H, Olusanya TO, Carr-Wilkinson J, Darby S, Faheem A, Elkordy AA. Evaluation of novel cationic gene based liposomes with cyclodextrin prepared by thin film hydration and microfluidic systems. Sci Rep. 2019;9(1):1–17. doi:10.1038/s41598-019-51065-4
63. Sercombe L, Veerati T, Moheimani F, Wu SY, Sood AK, Hua S. Advances and challenges of liposome assisted drug delivery. Front Pharmacol. 2015;6:286. doi:10.3389/fphar.2015.00286
64. Danaei M, Dehghankhold M, Ataei S, et al. Impact of particle size and polydispersity index on the clinical applications of lipidic nanocarrier systems. Pharmaceutics. 2018;10(2):57. doi:10.3390/pharmaceutics10020057
65. Gaumet M, Vargas A, Gurny R, Delie F. Nanoparticles for drug delivery: the need for precision in reporting particle size parameters. Eur J Pharm Biopharm. 2008;69(1):1–9. doi:10.1016/j.ejpb.2007.08.001
66. Koppel DE. Analysis of macromolecular polydispersity in intensity correlation spectroscopy: the method of cumulants. J Chem Phys. 1972;57(11):4814–4820. doi:10.1063/1.1678153
67. Fissan H, Ristig S, Kaminski H, Asbach C, Epple M. Comparison of different characterization methods for nanoparticle dispersions before and after aerosolization. Anal Methods. 2014;6(18):7324–7334. doi:10.1039/C4AY01203H
68. Kim A, Ng WB, Bernt W, Cho N-J. Validation of size estimation of nanoparticle tracking analysis on polydisperse macromolecule assembly. Sci Rep. 2019;9(1):1–14. doi:10.1038/s41598-018-37186-2
69. Filipe V, Hawe A, Jiskoot W. Critical evaluation of Nanoparticle Tracking Analysis (NTA) by NanoSight for the measurement of nanoparticles and protein aggregates. Pharm Res. 2010;27(5):796–810. doi:10.1007/s11095-010-0073-2
70. Malloy A, Carr B. NanoParticle tracking analysis–The halo™ system. Part Part Syst Charact. 2006;23(2):197–204. doi:10.1002/ppsc.200601031
71. Elizondo E, Moreno E, Cabrera I, et al. Liposomes and other vesicular systems: structural characteristics, methods of preparation, and use in nanomedicine. Progress Mol Biol Transl Sci. 2011;104:1–52.
72. Hunter RJ, Midmore BR, Zhang H. Zeta potential of highly charged thin double-layer systems. J Colloid Interface Sci. 2001;237(1):147–149. doi:10.1006/jcis.2001.7423
73. Smith MC, Crist RM, Clogston JD, McNeil SE. Zeta potential: a case study of cationic, anionic, and neutral liposomes. Anal Bioanal Chem. 2017;409(24):5779–5787. doi:10.1007/s00216-017-0527-z
74. Clogston JD, Patri AK. Importance of physicochemical characterization prior to immunological studies. In: Handbook of Immunological Properties of Engineered Nanomaterials. 2013. 2013. World Scientific; 25–52.
75. Kaszuba M, Corbett J, Watson FM, Jones A. High-concentration zeta potential measurements using light-scattering techniques. Philos Trans Royal Soc A. 2010;368(1927):4439–4451. doi:10.1098/rsta.2010.0175
76. Manconi M, Aparicio J, Vila A, Pendás J, Figueruelo J, Molina F. Viscoelastic properties of concentrated dispersions in water of soy lecithin. Colloids Surf A. 2003;222(1–3):141–145. doi:10.1016/S0927-7757(03)00249-8
77. Kostarelos K, Emfietzoglou D, Papakostas A, Yang WH, Ballangrud Å, Sgouros G. Binding and interstitial penetration of liposomes within avascular tumor spheroids. Int J Cancer. 2004;112(4):713–721. doi:10.1002/ijc.20457
78. Kostarelos K, Emfietzoglou D, Papakostas A, Yang W-H, Ballangrud ÅM, Sgouros G. Engineering lipid vesicles of enhanced intratumoral transport capabilities: correlating liposome characteristics with penetration into human prostate tumor spheroids. J Liposome Res. 2005;15(1–2):15–27. doi:10.1081/LPR-64953
79. Krasnici S, Werner A, Eichhorn ME, et al. Effect of the surface charge of liposomes on their uptake by angiogenic tumor vessels. Int J Cancer. 2003;105(4):561–567. doi:10.1002/ijc.11108
80. AS AL, Kizuki S, Ishida T, Ishida T, Kiwada H, Kiwada H. Oxaliplatin encapsulated in PEG-coated cationic liposomes induces significant tumor growth suppression via a dual-targeting approach in a murine solid tumor model. J Control Release. 2009;137(1):8–14. doi:10.1016/j.jconrel.2009.02.023
81. Waite CL, Roth CM. Nanoscale drug delivery systems for enhanced drug penetration into solid tumors: current progress and opportunities. Crit Rev Biomed Eng. 2012;40(1):21–41. doi:10.1615/critrevbiomedeng.v40.i1.20
82. Spyratou E, Mourelatou EA, Makropoulou M, Demetzos C. Atomic force microscopy: a tool to study the structure, dynamics and stability of liposomal drug delivery systems. Expert Opin Drug Delivery. 2009;6(3):305–317. doi:10.1517/17425240902828312
83. Fröhlich M, Brecht V, Peschka-Süss R. Parameters influencing the determination of liposome lamellarity by 31P-NMR. Chem Phys Lipids. 2001;109(1):103–112. doi:10.1016/S0009-3084(00)00220-6
84. Mayer L, Hope M, Cullis P, Janoff A. Solute distributions and trapping efficiencies observed in freeze-thawed multilamellar vesicles. Biochimica et Biophysica Acta (BBA)-Biomembranes. 1985;817(1):193–196. doi:10.1016/0005-2736(85)90084-7
85. Craig D, Taylor K, Barker S. Calorimetric investigations of liposome formation. J Pharm Pharmacol. 1990;42(Supplement_1):29P. doi:10.1111/j.2042-7158.1990.tb14402.x
86. Pentak D. Alternative methods of determining phase transition temperatures of phospholipids that constitute liposomes on the example of DPPC and DMPC. Thermochim Acta. 2014;584:36–44. doi:10.1016/j.tca.2014.03.020
87. Sot J, Aranda FJ, Collado M-I, Goni FM, Alonso A. Different effects of long-and short-chain ceramides on the gel-fluid and lamellar-hexagonal transitions of phospholipids: a calorimetric, NMR, and x-ray diffraction study. Biophys J. 2005;88(5):3368–3380. doi:10.1529/biophysj.104.057851
88. Youssefian S, Rahbar N, Lambert CR, Van Dessel S. Variation of thermal conductivity of DPPC lipid bilayer membranes around the phase transition temperature. J Royal Soc Interface. 2017;14(130):20170127. doi:10.1098/rsif.2017.0127
89. 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 Controlled Release. 2009;139(1):73–80. doi:10.1016/j.jconrel.2009.05.036
90. Bakonyi M, Berkó S, Budai-Szűcs M, Kovács A, Csányi E. DSC for evaluating the encapsulation efficiency of lidocaine-loaded liposomes compared to the ultracentrifugation method. J Therm Analysis Calorimetry. 2017;130(3):1619–1625. doi:10.1007/s10973-017-6394-1
91. Edwards KA, Baeumner AJ. Analysis of liposomes. Talanta. 2006;68(5):1432–1441. doi:10.1016/j.talanta.2005.08.031
92. Anzai K, Yoshida M, Kirino Y. Change in intravesicular volume of liposomes by freeze-thaw treatment as studied by the ESR stopped-flow technique. Biochimica et Biophysica Acta (BBA)-Biomembranes. 1990;1021(1):21–26. doi:10.1016/0005-2736(90)90378-2
93. Zhang X-M, Patel AB, de Graaf RA, Behar KL. Determination of liposomal encapsulation efficiency using proton NMR spectroscopy. Chem Phys Lipids. 2004;127(1):113–120. doi:10.1016/j.chemphyslip.2003.09.013
94. Sharma A, Sharma US. Liposomes in drug delivery: progress and limitations. Int J Pharm. 1997;154(2):123–140. doi:10.1016/S0378-5173(97)00135-X
95. SASAKI H, TAKAKURA Y, Hashida M, Kimura T, SEZAKI H. Antitumor activity of lipophilic prodrugs of mitomycin C entrapped in liposome or o/w emulsion. J Pharmacobio Dyn. 1984;7(2):120–130. doi:10.1248/bpb1978.7.120
96. Gulati M, Grover M, Singh S, Singh M. Lipophilic drug derivatives in liposomes. Int J Pharm. 1998;165(2):129–168. doi:10.1016/S0378-5173(98)00006-4
97. Nii T, Ishii F. Encapsulation efficiency of water-soluble and insoluble drugs in liposomes prepared by the microencapsulation vesicle method. Int J Pharm. 2005;298(1):198–205. doi:10.1016/j.ijpharm.2005.04.029
98. Balon K, Riebesehl BU, Müller BW. Determination of liposome partitioning of ionizable drugs by titration. J Pharmaceut Sci. 1999;88(8):802–806. doi:10.1021/js9804213
99. Mayer LD, Bally MB, Hope MJ, Cullis PR. Techniques for encapsulating bioactive agents into liposomes. Chem Phys Lipids. 1986;40(2–4):333–345. doi:10.1016/0009-3084(86)90077-0
100. Pauli G, Tang W-L, Li S-D. Development and characterization of the solvent-assisted active loading technology (SALT) for liposomal loading of poorly water-soluble compounds. Pharmaceutics. 2019;11(9):465. doi:10.3390/pharmaceutics11090465
101. Li T, Cipolla D, Rades T, Boyd BJ. Drug nanocrystallisation within liposomes. J Control Release. 2018;288:96–110. doi:10.1016/j.jconrel.2018.09.001
102. Dash S, Murthy PN, Nath L, Chowdhury P. Kinetic modeling on drug release from controlled drug delivery systems. Acta Pol Pharm. 2010;67(3):217–223.
103. Elhissi A. Liposomes for pulmonary drug delivery: the role of formulation and inhalation device design. Curr Pharm Des. 2017;23(3):362–372. doi:10.2174/1381612823666161116114732
104. Cipolla D, Gonda I, Chan H-K. Liposomal formulations for inhalation. Therapeutic Delivery. 2013;4(8):1047–1072. doi:10.4155/tde.13.71
105. Lamichhane N, Udayakumar TS, D’Souza WD, et al. Liposomes: clinical applications and potential for image-guided drug delivery. Molecules. 2018;23(2):288. doi:10.3390/molecules23020288
106. Nisini R, Poerio N, Mariotti S, De Santis F, Fraziano M. The multirole of liposomes in therapy and prevention of infectious diseases. Front Immunol. 2018;9:155. doi:10.3389/fimmu.2018.00155
107. Zylberberg C, Matosevic S. Pharmaceutical liposomal drug delivery: a review of new delivery systems and a look at the regulatory landscape. Drug Delivery. 2016;23(9):3319–3329. doi:10.1080/10717544.2016.1177136
108. Meers P, Neville M, Malinin V, et al. Biofilm penetration, triggered release and in vivo activity of inhaled liposomal amikacin in chronic Pseudomonas aeruginosa lung infections. J Antimicrob Chemother. 2008;61(4):859–868. doi:10.1093/jac/dkn059
109. Rose SJ, Neville ME, Gupta R, Bermudez LE. Delivery of aerosolized liposomal amikacin as a novel approach for the treatment of nontuberculous mycobacteria in an experimental model of pulmonary infection. PLoS One. 2014;9(9):e108703. doi:10.1371/journal.pone.0108703
110. Zhang J, Leifer F, Rose S, et al. Amikacin liposome inhalation suspension (ALIS) penetrates non-tuberculous mycobacterial biofilms and enhances amikacin uptake into macrophages. Front Microbiol. 2018;9:915. doi:10.3389/fmicb.2018.00915
111. Hong K, Drummond DC, Kirpotin DB. Liposomes for drug delivery. Google Patents. 2017.
112. Allen TM, Cullis PR. Liposomal drug delivery systems: from concept to clinical applications. Adv Drug Delivery Rev. 2013;65(1):36–48. doi:10.1016/j.addr.2012.09.037
113. Abra R, Bankert R, Chen F, et al. The next generation of liposome delivery systems: recent experience with tumor-targeted, sterically-stabilized immunoliposomes and active-loading gradients. J Liposome Res. 2002;12(1–2):1–3. doi:10.1081/LPR-120004770
114. Cattel L, Ceruti M, Dosio F. From conventional to stealth liposomes a new frontier in cancer chemotherapy. Tumori J. 2003;89(3):237–249. doi:10.1177/030089160308900302
115. Immordino ML, Dosio F, Cattel L. Stealth liposomes: review of the basic science, rationale, and clinical applications, existing and potential. Int j Nanomed. 2006;1(3):297.
116. Monteiro N, Martins A, Reis RL, Neves NM. Liposomes in tissue engineering and regenerative medicine. J Royal Soc Interface. 2014;11(101):20140459. doi:10.1098/rsif.2014.0459
117. Saraf S, Jain A, Tiwari A, Verma A, Panda PK, Jain SK. Advances in liposomal drug delivery to cancer: an overview. J Drug Delivery Sci Technol. 2020;56:101549. doi:10.1016/j.jddst.2020.101549
118. Hatakeyama H, Akita H, Harashima H. The polyethyleneglycol dilemma: advantage and disadvantage of PEGylation of liposomes for systemic genes and nucleic acids delivery to tumors. Biol Pharm Bull. 2013;36(6):892–899. doi:10.1248/bpb.b13-00059
119. Madni A, Sarfraz M, Rehman M, et al. Liposomal drug delivery: a versatile platform for challenging clinical applications. J Pharm Pharm Sci. 2014;17(3):401–426. doi:10.18433/J3CP55
120. Fathi S, Oyelere AK. Liposomal drug delivery systems for targeted cancer therapy: is active targeting the best choice? Future Med Chem. 2016;8(17):2091–2112. doi:10.4155/fmc-2016-0135
121. NTT L, Cao VD, Nguyen TNQ, TTH L, Tran TT, Hoang Thi TT. Soy lecithin-derived liposomal delivery systems: surface modification and current applications. Int J Mol Sci. 2019;20(19):4706. doi:10.3390/ijms20194706
122. Eloy JO, Petrilli R, Trevizan LNF, Chorilli M. Immunoliposomes: a review on functionalization strategies and targets for drug delivery. Colloids Surf B. 2017;159:454–467. doi:10.1016/j.colsurfb.2017.07.085
123. Wallis JG, Browse J. Mutants of Arabidopsis reveal many roles for membrane lipids. Prog lipid res. 2002;41(3):254–278. doi:10.1016/S0163-7827(01)00027-3
124. Karanth H, Murthy R. pH‐Sensitive liposomes‐principle and application in cancer therapy. J Pharm Pharmacol. 2007;59(4):469–483. doi:10.1211/jpp.59.4.0001
125. Li L, ten Hagen TL, Schipper D, et al. Triggered content release from optimized stealth thermosensitive liposomes using mild hyperthermia. J Control Release. 2010;143(2):274–279. doi:10.1016/j.jconrel.2010.01.006
126. Lu Y, Sun W, Gu Z. Stimuli-responsive nanomaterials for therapeutic protein delivery. J Controlled Release. 2014;194:1–19. doi:10.1016/j.jconrel.2014.08.015
127. Li S, Goins B, Zhang L, Bao A. Novel multifunctional theranostic liposome drug delivery system: construction, characterization, and multimodality MR, near-infrared fluorescent, and nuclear imaging. Bioconjugate Chem. 2012;23(6):1322–1332. doi:10.1021/bc300175d
128. Fay F, Scott CJ. Antibody-targeted nanoparticles for cancer therapy. Immunotherapy. 2011;3(3):381–394. doi:10.2217/imt.11.5
129. Attia MF, Anton N, Wallyn J, Omran Z, Vandamme, TF. An overview of active and passive targeting strategies to improve the nanocarriers efficiency to tumour sites. J Pharm Pharmacol. 2019;71(8):1185–1198. doi:10.1111/jphp.13098
130. Wicki A, Witzigmann D, Balasubramanian V, Huwyler J. Nanomedicine in cancer therapy: challenges, opportunities, and clinical applications. J Controlled Release. 2015;200:138–157. doi:10.1016/j.jconrel.2014.12.030
131. Gogoi M, Kumar N, Patra S. Multifunctional magnetic liposomes for cancer imaging and therapeutic applications. Nanoarchitectonics Smart Delivery Drug Targeting. 2016;2016743–782.
132. Biswas S, Torchilin VP. Nanopreparations for organelle-specific delivery in cancer. Adv Drug Delivery Rev. 2014;66:26–41. doi:10.1016/j.addr.2013.11.004
133. 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 Delivery Rev. 2011;63(3):136–151. doi:10.1016/j.addr.2010.04.009
134. Byrne JD, Betancourt T, Brannon-Peppas L. Active targeting schemes for nanoparticle systems in cancer therapeutics. Adv Drug Delivery Rev. 2008;60(15):1615–1626. doi:10.1016/j.addr.2008.08.005
135. Park J, Choi Y, Chang H, Um W, Ryu JH, Kwon IC. Alliance with EPR effect: combined strategies to improve the EPR effect in the tumor microenvironment. Theranostics. 2019;9(26):8073. doi:10.7150/thno.37198
136. Strebhardt K, Ullrich A. Paul Ehrlich’s magic bullet concept: 100 years of progress. Nat Rev Cancer. 2008;8(6):473–480. doi:10.1038/nrc2394
137. Bazak R, Houri M, El Achy S, Kamel S, Refaat T. Cancer active targeting by nanoparticles: a comprehensive review of literature. J Cancer Res Clin Oncol. 2015;141(5):769–784. doi:10.1007/s00432-014-1767-3
138. Noble GT, Stefanick JF, Ashley JD, Kiziltepe T, Bilgicer B. Ligand-targeted liposome design: challenges and fundamental considerations. Trends Biotechnol. 2014;32(1):32–45. doi:10.1016/j.tibtech.2013.09.007
139. Sawant RR, Torchilin VP. Challenges in development of targeted liposomal therapeutics. AAPS J. 2012;14(2):303–315. doi:10.1208/s12248-012-9330-0
140. Marqués-Gallego P, de Kroon AI. Ligation strategies for targeting liposomal nanocarriers. Biomed Res Int. 2014;2014:1–12. doi:10.1155/2014/129458
141. Conde J, Dias JT, Grazú V, Moros M, Baptista PV, de la Fuente JM. Revisiting 30 years of biofunctionalization and surface chemistry of inorganic nanoparticles for nanomedicine. Front Chem. 2014;2:48. doi:10.3389/fchem.2014.00048
142. Steenpaß T, Lung A, Schubert R. Tresylated PEG-sterols for coupling of proteins to preformed plain or PEGylated liposomes. Biochimica et Biophysica Acta (BBA)-Biomembranes. 2006;1758(1):20–28. doi:10.1016/j.bbamem.2005.12.010
143. Drummond DC, Meyer O, Hong K, Kirpotin DB, Papahadjopoulos D. Optimizing liposomes for delivery of chemotherapeutic agents to solid tumors. Pharmacol Rev. 1999;51(4):691–744.
144. Gabizon AA. Pegylated liposomal doxorubicin: metamorphosis of an old drug into a new form of chemotherapy. Cancer Invest. 2001;19(4):424–436. doi:10.1081/CNV-100103136
145. Adams GP, Schier R, McCall AM, et al. High affinity restricts the localization and tumor penetration of single-chain fv antibody molecules. Cancer Res. 2001;61(12):4750–4755.
146. Gosk S, Moos T, Gottstein C, Bendas G. VCAM-1 directed immunoliposomes selectively target tumor vasculature in vivo. Biochimica et Biophysica Acta (BBA)-Biomembranes. 2008;1778(4):854–863. doi:10.1016/j.bbamem.2007.12.021
147. Allen TM. Ligand-targeted therapeutics in anticancer therapy. Nat Rev Cancer. 2002;2(10):750–763. doi:10.1038/nrc903
148. Low PS, Kularatne SA. Folate-targeted therapeutic and imaging agents for cancer. Curr Opin Chem Biol. 2009;13(3):256–262. doi:10.1016/j.cbpa.2009.03.022
149. Ni S, Stephenson SM, Lee RJ. Folate receptor targeted delivery of liposomal daunorubicin into tumor cells. Anticancer Res. 2002;22(4):2131–2135.
150. Pan XQ, Wang H, Lee RJ. Antitumor activity of folate receptor-targeted liposomal doxorubicin in a KB oral carcinoma murine xenograft model. Pharm Res. 2003;20(3):417–422. doi:10.1023/A:1022656105022
151. Pan XQ, Zheng X, Shi G, Wang H, Ratnam M, Lee RJ. Strategy for the treatment of acute myelogenous leukemia based on folate receptor β–targeted liposomal doxorubicin combined with receptor induction using all-trans retinoic acid. Blood J Am Soc Hematol. 2002;100(2):594–602.
152. Liu M-C, Liu L, Wang X-R, et al. Folate receptor-targeted liposomes loaded with a diacid metabolite of norcantharidin enhance antitumor potency for H22 hepatocellular carcinoma both in vitro and in vivo. Int j Nanomed;2016. 1395–1412. doi:10.2147/IJN.S96862
153. Wu J, Liu Q, Lee RJ. A folate receptor-targeted liposomal formulation for paclitaxel. Int J Pharm. 2006;316(1–2):148–153. doi:10.1016/j.ijpharm.2006.02.027
154. Soni V, Kohli D, Jain S. Transferrin-conjugated liposomal system for improved delivery of 5-fluorouracil to brain. J Drug Targeting. 2008;16(1):73–78. doi:10.1080/10611860701725381
155. van Rooy I, Mastrobattista E, Storm G, Hennink WE, Schiffelers RM. Comparison of five different targeting ligands to enhance accumulation of liposomes into the brain. J Controlled Release. 2011;150(1):30–36. doi:10.1016/j.jconrel.2010.11.014
156. Zolnik BS, Stern ST, Kaiser JM, et al. Rapid distribution of liposomal short-chain ceramide in vitro and in vivo. Drug Metab Dispos. 2008;36(8):1709–1715. doi:10.1124/dmd.107.019679
157. Li X, Ding L, Xu Y, Wang Y, Ping Q. Targeted delivery of doxorubicin using stealth liposomes modified with transferrin. Int J Pharm. 2009;373(1–2):116–123. doi:10.1016/j.ijpharm.2009.01.023
158. Sharma G, Modgil A, Sun C, Singh J. Grafting of cell-penetrating peptide to receptor-targeted liposomes improves their transfection efficiency and transport across blood–brain barrier model. J Pharmaceut Sci. 2012;101(7):2468–2478. doi:10.1002/jps.23152
159. Gao J-Q, Lv Q, Li L-M, et al. Glioma targeting and blood–brain barrier penetration by dual-targeting doxorubincin liposomes. Biomaterials. 2013;34(22):5628–5639. doi:10.1016/j.biomaterials.2013.03.097
160. Lehtinen J, Raki M, Bergström KA, et al. Pre-targeting and direct immunotargeting of liposomal drug carriers to ovarian carcinoma. PLoS One. 2012;7(7):e41410. doi:10.1371/journal.pone.0041410
161. Kim SK, Huang L. Nanoparticle delivery of a peptide targeting EGFR signaling. J Controlled Release. 2012;157(2):279–286. doi:10.1016/j.jconrel.2011.08.014
162. Danhier F, Feron O, Préat V. To exploit the tumor microenvironment: passive and active tumor targeting of nanocarriers for anti-cancer drug delivery. J Controlled Release. 2010;148(2):135–146. doi:10.1016/j.jconrel.2010.08.027
163. Mamot C, Drummond DC, Greiser U, et al. Epidermal growth factor receptor (EGFR)-targeted immunoliposomes mediate specific and efficient drug delivery to EGFR-and EGFRvIII-overexpressing tumor cells. Cancer Res. 2003;63(12):3154–3161.
164. Mamot C, Ritschard R, Küng W, Park JW, Herrmann R, Rochlitz CF. EGFR-targeted immunoliposomes derived from the monoclonal antibody EMD72000 mediate specific and efficient drug delivery to a variety of colorectal cancer cells. J Drug Targeting. 2006;14(4):215–223. doi:10.1080/10611860600691049
165. Dagar S, Krishnadas A, Rubinstein I, Blend MJ, Önyüksel H. VIP grafted sterically stabilized liposomes for targeted imaging of breast cancer: in vivo studies. J Control Release. 2003;91(1–2):123–133. doi:10.1016/S0168-3659(03)00242-6
166. Peer D, Margalit R. Loading mitomycin C inside long circulating hyaluronan targeted nano‐liposomes increases its antitumor activity in three mice tumor models. Int J Cancer. 2004;108(5):780–789. doi:10.1002/ijc.11615
167. Lee CM, Tanaka T, Murai T, et al. Novel chondroitin sulfate-binding cationic liposomes loaded with cisplatin efficiently suppress the local growth and liver metastasis of tumor cells in vivo. Cancer Res. 2002;62(15):4282–4288.
168. Hashida M, Nishikawa M, Yamashita F, Takakura Y. Cell-specific delivery of genes with glycosylated carriers. Adv Drug Delivery Rev. 2001;52(3):187–196. doi:10.1016/S0169-409X(01)00209-5
169. Tu R, Mohanty K, Tirrell M. Liposomal targeting through peptide-amphiphile functionalization. Am Pharm Rev. 2004;7:36–48.
170. Kang DI, Lee S, Lee JT, et al. Preparation and in vitro evaluation of anti-VCAM-1-Fab′-conjugated liposomes for the targeted delivery of the poorly water-soluble drug celecoxib. J Microencapsulation. 2011;28(3):220–227. doi:10.3109/02652048.2011.552989
171. Du H, Cui C, Wang L, Liu H, Cui G. Novel tetrapeptide, RGDF, mediated tumor specific liposomal doxorubicin (DOX) preparations. Mol Pharmaceut. 2011;8(4):1224–1232. doi:10.1021/mp200039s
172. Chen Z, Deng J, Zhao Y, Tao T. Cyclic RGD peptide-modified liposomal drug delivery system: enhanced cellular uptake in vitro and improved pharmacokinetics in rats. Int j Nanomed. 2012;3803–3811. doi:10.2147/IJN.S33541
173. Sriraman SK, Torchilin VP. Recent advances with liposomes as drug carriers. Adv Biomater Biodevice. 2014;2014;79–119.
174. Torchilin VP. Targeted pharmaceutical nanocarriers for cancer therapy and imaging. AAPS J. 2007;9(2):E128–E147. doi:10.1208/aapsj0902015
175. Cardone RA, Casavola V, Reshkin SJ. The role of disturbed pH dynamics and the Na+/H+ exchanger in metastasis. Nat Rev Cancer. 2005;5(10):786–795. doi:10.1038/nrc1713
176. Arias J L. Drug targeting strategies in cancer treatment: an overview. Mini Rev Med Chem. 2011;11(1):1–17. doi:10.2174/138955711793564024
177. Wang L, Geng D, Su H. Safe and efficient pH sensitive tumor targeting modified liposomes with minimal cytotoxicity. Colloids Surf B. 2014;123:395–402. doi:10.1016/j.colsurfb.2014.09.003
178. Júnior ÁD, Mota LG, Nunan EA, et al. Tissue distribution evaluation of stealth pH-sensitive liposomal cisplatin versus free cisplatin in Ehrlich tumor-bearing mice. Life Sci. 2007;80(7):659–664. doi:10.1016/j.lfs.2006.10.011
179. Hu Y, Zeng F, Ju R, Lu W. Advances in liposomal drug delivery system in the field of chemotherapy. Clin Oncol. 2016;1:1092.
180. May JP, Li S-D. Hyperthermia-induced drug targeting. Expert Opin Drug Delivery. 2013;10(4):511–527. doi:10.1517/17425247.2013.758631
181. Kakinuma K, Tanaka R, Takahashi H, Sekihara Y, Watanabe M, Kuroki M. Drug delivery to the brain using thermosensitive liposome and local hyperthermia. Int j Hyperthermia. 1996;12(1):157–165. doi:10.3109/02656739609023698
182. Yatvin MB, Weinstein JN, Dennis WH, Blumenthal R. Design of liposomes for enhanced local release of drugs by hyperthermia. Science. 1978;202(4374):1290–1293. doi:10.1126/science.364652
183. Mura S, Nicolas J, Couvreur P. Stimuli-responsive nanocarriers for drug delivery. Nat Mater. 2013;12(11):991–1003. doi:10.1038/nmat3776
184. Wang -X-X, Li Y-B, Yao H-J, et al. The use of mitochondrial targeting resveratrol liposomes modified with a dequalinium polyethylene glycol-distearoylphosphatidyl ethanolamine conjugate to induce apoptosis in resistant lung cancer cells. Biomaterials. 2011;32(24):5673–5687. doi:10.1016/j.biomaterials.2011.04.029
185. Ma X, Zhou J, Zhang C-X, et al. Modulation of drug-resistant membrane and apoptosis proteins of breast cancer stem cells by targeting berberine liposomes. Biomaterials. 2013;34(18):4452–4465. doi:10.1016/j.biomaterials.2013.02.066
186. Kulshrestha P, Gogoi M, Bahadur D, Banerjee R. In vitro application of paclitaxel loaded magnetoliposomes for combined chemotherapy and hyperthermia. Colloids Surf B. 2012;96:1–7. doi:10.1016/j.colsurfb.2012.02.029
187. Eloy JO, de Souza MC, Petrilli R, Barcellos JPA, Lee RJ, Marchetti JM. Liposomes as carriers of hydrophilic small molecule drugs: strategies to enhance encapsulation and delivery. Colloids Surf B. 2014;123:345–363. doi:10.1016/j.colsurfb.2014.09.029
188. Fattahi H, Laurent S, Liu F, Arsalani N, Elst LV, Muller RN. Magnetoliposomes as multimodal contrast agents for molecular imaging and cancer nanotheragnostics. Nanomedicine. 2011;6(3):529–544. doi:10.2217/nnm.11.14
189. Al-Jamal WT, Kostarelos K. Liposome–nanoparticle hybrids for multimodal diagnostic and therapeutic applications. Nanomedicine. 2007;2(1):85–98. doi:10.2217/17435889.2.1.85
190. Saiyed ZM, Gandhi NH, Nair MP. Magnetic nanoformulation of azidothymidine 5’-triphosphate for targeted delivery across the blood–brain barrier. Int j Nanomed. 2010;5:157–166. doi:10.2147/ijn.s8905
191. Riviere C, Martina M-S, Tomita Y, et al. Magnetic targeting of nanometric magnetic fluid–loaded liposomes to specific brain intravascular areas: a dynamic imaging study in mice. Radiology. 2007;244(2):439–448. doi:10.1148/radiol.2442060912
192. Clares B, Biedma-Ortiz RA, Sáez-Fernández E, et al. Nano-engineering of 5-fluorouracil-loaded magnetoliposomes for combined hyperthermia and chemotherapy against colon cancer. Eur J Pharm Biopharm. 2013;85(3):329–338. doi:10.1016/j.ejpb.2013.01.028
193. Andresen TL, Jensen SS, Jørgensen K. Advanced strategies in liposomal cancer therapy: problems and prospects of active and tumor specific drug release. Prog lipid res. 2005;44(1):68–97. doi:10.1016/j.plipres.2004.12.001
194. Huang S-L, MacDonald RC. Acoustically active liposomes for drug encapsulation and ultrasound-triggered release. Biochimica et Biophysica Acta (BBA)-Biomembranes. 2004;1665(1–2):134–141. doi:10.1016/j.bbamem.2004.07.003
195. Buchanan KD, Huang S-L, Kim H, McPherson DD, MacDonald RC. Encapsulation of NF-κB decoy oligonucleotides within echogenic liposomes and ultrasound-triggered release. J Controlled Release. 2010;141(2):193–198. doi:10.1016/j.jconrel.2009.09.017
196. Lin C-Y, Javadi M, Belnap DM, Barrow JR, Pitt WG. Ultrasound sensitive eLiposomes containing doxorubicin for drug targeting therapy. Nanomed Nanotechnol Biol Med. 2014;10(1):67–76. doi:10.1016/j.nano.2013.06.011
197. Menon JU, Jadeja P, Tambe P, Vu K, Yuan B, Nguyen KT. Nanomaterials for photo-based diagnostic and therapeutic applications. Theranostics. 2013;3(3):152. doi:10.7150/thno.5327
198. Ashrafizadeh M, Delfi M. Stimuli-responsive liposomal nanoformulations in cancer therapy: Pre-clinical & clinical approaches. J Control Release. 2022;351():50–80. doi:10.1016/j.jconrel.2022.08.001
199. de Visscher SA, Kaščáková S, de Bruijn HS, et al. Fluorescence localization and kinetics of mTHPC and liposomal formulations of mTHPC in the window‐chamber tumor model. Lasers Surg Med. 2011;43(6):528–536. doi:10.1002/lsm.21082
200. You J, Zhang P, Hu F, et al. Near-infrared light-sensitive liposomes for the enhanced photothermal tumor treatment by the combination with chemotherapy. Pharm Res. 2014;31(3):554–565. doi:10.1007/s11095-013-1180-7
201. Qiu Y, Zhu Z, Miao Y, et al. Polymerization of dopamine accompanying its coupling to induce self-assembly of block copolymer and application in drug delivery. Polym Chem. 2020;11(16):2811–2821. doi:10.1039/D0PY00085J
202. Zalipsky S. Chemistry of polyethylene glycol conjugates with biologically active molecules. Adv Drug Delivery Rev. 1995;16(2–3):157–182. doi:10.1016/0169-409X(95)00023-Z
203. Jain RK, Stylianopoulos T. Delivering nanomedicine to solid tumors. Nat Rev Clin Oncol. 2010;7(11):653–664. doi:10.1038/nrclinonc.2010.139
204. Simoes S, Slepushkin V, Düzgünes N, de Lima MCP. On the mechanisms of internalization and intracellular delivery mediated by pH-sensitive liposomes. Biochimica Et Biophysica Acta (BBA)-Biomembranes. 2001;1515(1):23–37. doi:10.1016/S0005-2736(01)00389-3
205. Maurer N, Fenske DB, Cullis PR. Developments in liposomal drug delivery systems. Expert opin biol ther. 2001;1(6):923–947. doi:10.1517/14712598.1.6.923
206. Tang Y, Wang Y, Kiani MF, Wang B. Classification, treatment strategy, and associated drug resistance in breast cancer. Clin Breast Cancer. 2016;16(5):335–343. doi:10.1016/j.clbc.2016.05.012
207. Zhou L, Lu R, Liu Q, et al. Two branched fructose modification improves tumor targeting delivery of liposomes to breast cancer in intro and in vivo. J Drug Delivery Sci Technol. 2021;61:102312. doi:10.1016/j.jddst.2020.102312
208. Hossen S, Hossain MK, Basher M, Mia M, Rahman M, Uddin MJ. Smart nanocarrier-based drug delivery systems for cancer therapy and toxicity studies: a review. J Adv Res. 2019;15:1–18. doi:10.1016/j.jare.2018.06.005
209. Xiao W, Ruan S, Yu W, et al. Normalizing tumor vessels to increase the enzyme-induced retention and targeting of gold nanoparticle for breast cancer imaging and treatment. Mol Pharmaceut. 2017;14(10):3489–3498. doi:10.1021/acs.molpharmaceut.7b00475
210. Shafei A, El-Bakly W, Sobhy A, et al. A review on the efficacy and toxicity of different doxorubicin nanoparticles for targeted therapy in metastatic breast cancer. Biomed Pharmacother. 2017;95:1209–1218. doi:10.1016/j.biopha.2017.09.059
211. Rip J, Chen L, Hartman R, et al. Glutathione PEGylated liposomes: pharmacokinetics and delivery of cargo across the blood–brain barrier in rats. J Drug Targeting. 2014;22(5):460–467. doi:10.3109/1061186X.2014.888070
212. Gkionis L, Campbell RA, Aojula H, Harris LK, Tirella A. Manufacturing drug co-loaded liposomal formulations targeting breast cancer: influence of preparative method on liposomes characteristics and in vitro toxicity. Int J Pharm. 2020;590:119926. doi:10.1016/j.ijpharm.2020.119926
213. d’Avanzo N, Torrieri G, Figueiredo P, et al. LinTT1 peptide-functionalized liposomes for targeted breast cancer therapy. Int J Pharm. 2021;597:120346. doi:10.1016/j.ijpharm.2021.120346
214. Sharma P, Brown S, Walter G, Santra S, Moudgil B. Nanoparticles for bioimaging. Adv Colloid Interface Sci. 2006;123:471–485. doi:10.1016/j.cis.2006.05.026
215. El-Senduny FF, Altouhamy M, Zayed G, et al. Azadiradione-loaded liposomes with improved bioavailability and anticancer efficacy against triple negative breast cancer. J Drug Delivery Sci Technol. 2021;65:102665. doi:10.1016/j.jddst.2021.102665
216. Haun JB, Yoon TJ, Lee H, Weissleder R. Magnetic nanoparticle biosensors. Wiley Interdiscip Rev Nanomed Nanobiotechnol. 2010;2(3):291–304. doi:10.1002/wnan.84
217. Ağardan NM, Değim Z, Yılmaz Ş, Altıntaş L, Topal T. Tamoxifen/raloxifene loaded liposomes for oral treatment of breast cancer. J Drug Delivery Sci Technol. 2020;57:101612. doi:10.1016/j.jddst.2020.101612
218. Caravan P. Strategies for increasing the sensitivity of gadolinium based MRI contrast agents. Chem Soc Rev. 2006;35(6):512–523. doi:10.1039/b510982p
219. Bulbake U, Kommineni N, Bryszewska M, Ionov M, Khan W. Cationic liposomes for co-delivery of paclitaxel and anti-Plk1 siRNA to achieve enhanced efficacy in breast cancer. J Drug Delivery Sci Technol. 2018;48:253–265. doi:10.1016/j.jddst.2018.09.017
220. Michalet X, Pinaud FF, Bentolila LA, et al. Quantum dots for live cells, in vivo imaging, and diagnostics. science. 2005;307(5709):538–544. doi:10.1126/science.1104274
221. Mielczarek L, Krug P, Mazur M, Milczarek M, Chilmonczyk Z, Wiktorska K. In the triple-negative breast cancer MDA-MB-231 cell line, sulforaphane enhances the intracellular accumulation and anticancer action of doxorubicin encapsulated in liposomes. Int J Pharm. 2019;558:311–318. doi:10.1016/j.ijpharm.2019.01.008
222. Wu W, He Q, Jiang C. Magnetic iron oxide nanoparticles: synthesis and surface functionalization strategies. Nanoscale Res Lett. 2008;3(11):397–415. doi:10.1007/s11671-008-9174-9
223. Han S-M, Baek J-S, Kim M-S, Hwang S-J, Cho C-W. Surface modification of paclitaxel-loaded liposomes using d-α-tocopheryl polyethylene glycol 1000 succinate: enhanced cellular uptake and cytotoxicity in multidrug resistant breast cancer cells. Chem Phys Lipids. 2018;213:39–47. doi:10.1016/j.chemphyslip.2018.03.005
224. Huang X, El-Sayed IH, Qian W, El-Sayed MA. Cancer cell imaging and photothermal therapy in the near-infrared region by using gold nanorods. J Am Chem Soc. 2006;128(6):2115–2120. doi:10.1021/ja057254a
225. Feuser PE, Cordeiro AP, de Bem Silveira G, et al. Co-encapsulation of sodium diethyldithiocarbamate (DETC) and zinc phthalocyanine (ZnPc) in liposomes promotes increases phototoxic activity against (MDA-MB 231) human breast cancer cells. Colloids Surf B. 2021;197:111434. doi:10.1016/j.colsurfb.2020.111434
226. Ulker D, Barut I, Şener E, Bütün V. Advanced liposome based PEGylated microgel as a novel release system for 5-fluorouracil against MCF-7 cancer cell. Eur Polym J. 2021;146:110270. doi:10.1016/j.eurpolymj.2021.110270
227. Smith AM, Duan H, Rhyner MN, Ruan G, Nie S. A systematic examination of surface coatings on the optical and chemical properties of semiconductor quantum dots. Phys Chem Chem Phys. 2006;8(33):3895–3903. doi:10.1039/b606572b
228. Zhao C-Y, Cheng R, Yang Z, Tian Z-M. Nanotechnology for cancer therapy based on chemotherapy. Molecules. 2018;23(4):826. doi:10.3390/molecules23040826
229. Almurshedi AS, Radwan M, Omar S, et al. A novel pH-sensitive liposome to trigger delivery of Afatinib to cancer cells: impact on lung cancer therapy. J Mol Liq. 2018;259:154–166. doi:10.1016/j.molliq.2018.03.024
230. Davidson MR, Gazdar AF, Clarke BE. The pivotal role of pathology in the management of lung cancer. J Thoracic Dis. 2013;5(Suppl 5):S463. doi:10.3978/j.issn.2072-1439.2013.08.43
231. Price N, Belani CP, Jain VK. Bisphosphonates to prevent skeletal morbidity in patients with lung cancer with bone metastases. Clin Lung Cancer. 2004;5(5):267–269. doi:10.1016/S1525-7304(11)70347-3
232. Xiao Z, Zhuang B, Zhang G, Li M, Jin Y. Pulmonary delivery of cationic liposomal hydroxycamptothecin and 5-aminolevulinic acid for chemo-sonodynamic therapy of metastatic lung cancer. Int J Pharm. 2021;601:120572. doi:10.1016/j.ijpharm.2021.120572
233. Goel A, Baboota S, Sahni JK, Ali J. Exploring targeted pulmonary delivery for treatment of lung cancer. Int J Pharm Invest. 2013;3(1):8. doi:10.4103/2230-973X.108959
234. Carvalho TC, Carvalho SR, McConville JT. Formulations for pulmonary administration of anticancer agents to treat lung malignancies. J Aerosol Med Pulmonary Drug Delivery. 2011;24(2):61–80. doi:10.1089/jamp.2009.0794
235. Anabousi S, Bakowsky U, Schneider M, Huwer H, Lehr C-M, Ehrhardt C. In vitro assessment of transferrin-conjugated liposomes as drug delivery systems for inhalation therapy of lung cancer. Eur J Pharm Sci. 2006;29(5):367–374. doi:10.1016/j.ejps.2006.07.004
236. Crous A, Abrahamse H. Effective gold nanoparticle-antibody-mediated drug delivery for photodynamic therapy of lung cancer stem cells. Int J Mol Sci. 2020;21(11):3742. doi:10.3390/ijms21113742
237. Arthur P, Patel N, Surapaneni SK, et al. Targeting lung cancer stem cells using combination of Tel and Docetaxel liposomes in 3D cultures and tumor xenografts. Toxicol Appl Pharmacol. 2020;401:115112. doi:10.1016/j.taap.2020.115112
238. Zhang M, Li M, Du L, Zeng J, Yao T, Jin Y. Paclitaxel-in-liposome-in-bacteria for inhalation treatment of primary lung cancer. Int J Pharm. 2020;578:119177. doi:10.1016/j.ijpharm.2020.119177
239. Kia P, Ruman U, Pratiwi AR, Hussein MZ. Innovative Therapeutic Approaches Based on Nanotechnology for the Treatment and Management of Tuberculosis. Int J Nanomed. 2023;Volume 18:1159–1191. doi:10.2147/IJN.S364634
240. Kumar A, Ruokolainen J, Kesari KK, Kashyap BK, Singh VV, Solanki MK. Smart Nanomaterials in Cancer Theranostics: challenges and Opportunities. ACS omega. 2023;8(16):14290–14320. doi:10.1021/acsomega.2c07840
241. Zhu X, Kong Y, Liu Q, et al. Inhalable dry powder prepared from folic acid-conjugated docetaxel liposomes alters pharmacodynamic and pharmacokinetic properties relevant to lung cancer chemotherapy. Pulmonary Pharmacol Therap. 2019;55:50–61. doi:10.1016/j.pupt.2019.02.001
242. Peng J, He X, Wang K, et al. An antisense oligonucleotide carrier based on amino silica nanoparticles for antisense inhibition of cancer cells. Nanomed Nanotechnol Biol Med. 2006;2(2):113–120. doi:10.1016/j.nano.2006.04.003
243. Jiménez-López J, Bravo-Caparrós I, Cabeza L, et al. Paclitaxel antitumor effect improvement in lung cancer and prevention of the painful neuropathy using large pegylated cationic liposomes. Biomed Pharmacother. 2021;133:111059. doi:10.1016/j.biopha.2020.111059
244. Shanmugam M, Kuthala N, Kong X, Chiang C-S, Hwang KC. Combined Gadolinium and Boron Neutron Capture Therapies for Eradication of Head-and-Neck Tumor Using Gd10B6 Nanoparticles under MRI/CT Image Guidance. JACS Au. 2023;3(8):2192–2205. doi:10.1021/jacsau.3c00250
245. Cano ME, Lesur D, Bincoletto V, et al. Synthesis of defined oligohyaluronates-decorated liposomes and interaction with lung cancer cells. Carbohydr Polym. 2020;248:116798. doi:10.1016/j.carbpol.2020.116798
246. Lindberg HK, Falck GC-M, Singh R, et al. Genotoxicity of short single-wall and multi-wall carbon nanotubes in human bronchial epithelial and mesothelial cells in vitro. Toxicology. 2013;313(1):24–37. doi:10.1016/j.tox.2012.12.008
247. Karpuz M, Silindir-Gunay M, Kursunel MA, Esendagli G, Dogan A, Ozer AY. Design and in vitro evaluation of folate-targeted, co-drug encapsulated theranostic liposomes for non-small cell lung cancer. J Drug Delivery Sci Technol. 2020;57:101707. doi:10.1016/j.jddst.2020.101707
248. Mukherjee A, Paul M, Mukherjee S. Recent progress in the theranostics application of nanomedicine in lung cancer. Cancers. 2019;11(5):597. doi:10.3390/cancers11050597
249. Çoban Ö, Barut B, Yalçın CÖ, Özel A, Bıyıklıoğlu Z. Development and in vitro evaluation of BSA-coated liposomes containing Zn (II) phthalocyanine-containing ferrocene groups for photodynamic therapy of lung cancer. J Organomet Chem. 2020;925:121469. doi:10.1016/j.jorganchem.2020.121469
250. Liu J, Cheng H, Han L, et al. Synergistic combination therapy of lung cancer using paclitaxel- and triptolide-coloaded lipid–polymer hybrid nanoparticles. Drug Des Devel Ther. 2018;Volume 12:3199–3209. doi:10.2147/DDDT.S172199
251. Zhang T, Chen Y, Ge Y, Hu Y, Li M, Jin Y. Inhalation treatment of primary lung cancer using liposomal curcumin dry powder inhalers. Acta Pharmaceutica Sinica B. 2018;8(3):440–448. doi:10.1016/j.apsb.2018.03.004
252. Fahmy HM. In vitro study of the cytotoxicity of thymoquinone/curcumin fluorescent liposomes. Naunyn-Schmiedeberg’s Arch Pharmacol. 2019;392(11):1465–1476. doi:10.1007/s00210-019-01688-1
253. Wang X, Cai H, Huang X, et al. Formulation and evaluation of a two-stage targeted liposome coated with hyaluronic acid for improving lung cancer chemotherapy and overcoming multidrug resistance. J biomater sci Poly ed;2023. 1–24. doi:10.1080/09205063.2022.2105103
254. Rybak AP, He L, Kapoor A, Cutz J-C, Tang D. Characterization of sphere-propagating cells with stem-like properties from DU145 prostate cancer cells. Biochimica Et Biophysica Acta (BBA)-Molecular Cell Research. 2011;1813(5):683–694. doi:10.1016/j.bbamcr.2011.01.018
255. Qu H, Liu H, Jin Y, Cui Z, Han G. HUWE1 upregulation has tumor suppressive effect in human prostate cancer cell lines through c-Myc. Biomed Pharmacother. 2018;106:309–315. doi:10.1016/j.biopha.2018.06.058
256. Le Broc-Ryckewaert D, Carpentier R, Lipka E, et al. Development of innovative paclitaxel-loaded small PLGA nanoparticles: study of their antiproliferative activity and their molecular interactions on prostatic cancer cells. Int J Pharm. 2013;454(2):712–719. doi:10.1016/j.ijpharm.2013.05.018
257. Pützer BM, Solanki M, Herchenröder O. Advances in cancer stem cell targeting: how to strike the evil at its root. Adv Drug Delivery Rev. 2017;120:89–107. doi:10.1016/j.addr.2017.07.013
258. Kim YJ, Liu Y, Li S, et al. Co-eradication of breast cancer cells and cancer stem cells by cross-linked multilamellar liposomes enhances tumor treatment. Mol Pharmaceut. 2015;12(8):2811–2822. doi:10.1021/mp500754r
259. Ramasamy T, Ruttala HB, Chitrapriya N, et al. Engineering of cell microenvironment-responsive polypeptide nanovehicle co-encapsulating a synergistic combination of small molecules for effective chemotherapy in solid tumors. Acta Biomater. 2017;48:131–143. doi:10.1016/j.actbio.2016.10.034
260. Duan X, Xiao J, Yin Q, et al. Smart pH-sensitive and temporal-controlled polymeric micelles for effective combination therapy of doxorubicin and disulfiram. ACS nano. 2013;7(7):5858–5869. doi:10.1021/nn4010796
261. Crain ML. Daunorubicin & Cytarabine liposome (vyxeos™). Oncol Times. 2018;40(10):30. doi:10.1097/01.COT.0000534146.30839.ec
262. Nel J, Elkhoury K, Velot É, Bianchi A, Acherar S, Francius G, Tamayol A, Grandemange S and Arab-Tehrany E. (2023). Functionalized liposomes for targeted breast cancer drug delivery. Bioactive Materials, 24 401–437. 10.1016/j.bioactmat.2022.12.027
263. Zhang RX, Wong HL, Xue HY, Eoh JY, Wu XY. Nanomedicine of synergistic drug combinations for cancer therapy–Strategies and perspectives. J Control Release. 2016;240:489–503. doi:10.1016/j.jconrel.2016.06.012
264. Tian J, Guo F, Chen Y, Li Y, Yu B, Li Y. Nanoliposomal formulation encapsulating celecoxib and genistein inhibiting COX-2 pathway and Glut-1 receptors to prevent prostate cancer cell proliferation. Cancer Lett. 2019;448:1–10. doi:10.1016/j.canlet.2019.01.002
265. Patil Y, Shmeeda H, Amitay Y, Ohana P, Kumar S, Gabizon A. Targeting of folate-conjugated liposomes with co-entrapped drugs to prostate cancer cells via prostate-specific membrane antigen (PSMA). Nanomed Nanotechnol Biol Med. 2018;14(4):1407–1416. doi:10.1016/j.nano.2018.04.011
266. Wadajkar AS, Menon JU, Tsai Y-S, et al. Prostate cancer-specific thermo-responsive polymer-coated iron oxide nanoparticles. Biomaterials. 2013;34(14):3618–3625. doi:10.1016/j.biomaterials.2013.01.062
267. Fernandes MA, Eloy JO, Luiz MT, et al. Transferrin-functionalized liposomes for docetaxel delivery to prostate cancer cells. Colloids Surf A. 2021;611:125806. doi:10.1016/j.colsurfa.2020.125806
268. Jurczyk M, Kasperczyk J, Wrześniok D, Beberok A, Jelonek K. Nanoparticles loaded with docetaxel and resveratrol as an advanced tool for cancer therapy. Biomedicines. 2022;10(5):1187. doi:10.3390/biomedicines10051187
269. Nandi U, Onyesom I, Douroumis D. Anti-cancer activity of sirolimus loaded liposomes in prostate cancer cell lines. J Drug Delivery Sci Technol. 2021;61:102200. doi:10.1016/j.jddst.2020.102200
270. Yeh C-Y, Hsiao J-K, Wang Y-P, Lan C-H, Wu H-C. Peptide-conjugated nanoparticles for targeted imaging and therapy of prostate cancer. Biomaterials. 2016;99:1–15. doi:10.1016/j.biomaterials.2016.05.015
271. Nassir AM, Ibrahim IA, Md S, et al. Surface functionalized folate targeted oleuropein nano-liposomes for prostate tumor targeting: in vitro and in vivo activity. Life Sci. 2019;220:136–146. doi:10.1016/j.lfs.2019.01.053
272. Kopel P, Wawrzak D, Moulick A, Milosavljevic V, Kizek R. Nanotransporters for anticancer drugs, modifications, target molecules. J Metallomics Nanotechnol. 2015;2:32–38.
273. Al-Azayzih A, Missaoui WN, Cummings BS, Somanath PR. Liposome-mediated delivery of the p21 activated kinase-1 (PAK-1) inhibitor IPA-3 limits prostate tumor growth in vivo. Nanomed Nanotechnol Biol Med. 2016;12(5):1231–1239. doi:10.1016/j.nano.2016.01.003
274. Sauvage F, Franzè S, Bruneau A, et al. Formulation and in vitro efficacy of liposomes containing the Hsp90 inhibitor 6BrCaQ in prostate cancer cells. Int J Pharm. 2016;499(1–2):101–109. doi:10.1016/j.ijpharm.2015.12.053
275. Pavlov R, Gaynanova G, Kuznetsov D, et al. A study involving PC-3 cancer cells and novel carbamate gemini surfactants: is zeta potential the key to control adhesion to cells? Smart Mater Med. 2023;4:123–133. doi:10.1016/j.smaim.2022.09.001
276. Kroon J, Buijs JT, Van Der Horst G, et al. Liposomal delivery of dexamethasone attenuates prostate cancer bone metastatic tumor growth in vivo. Prostate. 2015;75(8):815–824. doi:10.1002/pros.22963
277. Sutherland M, Gordon A, Shnyder SD, Patterson LH, Sheldrake HM. RGD-binding integrins in prostate cancer: expression patterns and therapeutic prospects against bone metastasis. Cancers. 2012;4(4):1106–1145. doi:10.3390/cancers4041106
278. Xiang B, Dong D-W, Shi N-Q, et al. PSA-responsive and PSMA-mediated multifunctional liposomes for targeted therapy of prostate cancer. Biomaterials. 2013;34(28):6976–6991. doi:10.1016/j.biomaterials.2013.05.055
279. Du C-X, Zhang T-B, Dong S-L, et al. A magnetic gene delivery nanosystem based on cationic liposomes. J Mater Sci. 2016;51(18):8461–8470. doi:10.1007/s10853-016-0106-2
280. Laskar P, Jaggi M, Chauhan SC, Yallapu MM, Yallapu MM. Biomolecule-functionalized nanoformulations for prostate cancer theranostics. J Adv Res. 2023;51:197–217. doi:10.1016/j.jare.2022.11.001
281. Moreira T, Silva ADO, Vasconcelos BRF, et al. DOPE/CHEMS-Based EGFR-Targeted Immunoliposomes for Docetaxel Delivery: formulation Development, Physicochemical Characterization and Biological Evaluation on Prostate Cancer Cells. Pharmaceutics. 2023;15(3):915. doi:10.3390/pharmaceutics15030915
282. Sesarman A, Tefas L, Sylvester B, et al. Anti-angiogenic and anti-inflammatory effects of long-circulating liposomes co-encapsulating curcumin and doxorubicin on C26 murine colon cancer cells. Pharmacol Rep. 2018;70(2):331–339. doi:10.1016/j.pharep.2017.10.004
283. Kenidra B, Benmohammed M. An ultra-fast method for clustering of big genomic data. IJAMC. 2020;11(1):45–60. doi:10.4018/IJAMC.2020010104
284. Palmer-Wackerly AL, Dailey PM, Krok-Schoen JL, Rhodes ND, Krieger JL. Patient perceptions of illness identity in cancer clinical trial decision-making. Health Commun. 2018;33(8):1045–1054. doi:10.1080/10410236.2017.1331189
285. Ogunwobi OO, Mahmood F, Akingboye A. Biomarkers in colorectal cancer: current research and future prospects. Int J Mol Sci. 2020;21(15):5311. doi:10.3390/ijms21155311
286. Shussman N, Wexner SD. Colorectal polyps and polyposis syndromes. Gastroenterol Rep. 2014;2(1):1–15. doi:10.1093/gastro/got041
287. Wahab S, Alshahrani MY, Ahmad MF, Abbas H. Current trends and future perspectives of nanomedicine for the management of colon cancer. Eur J Pharmacol. 2021;910:174464. doi:10.1016/j.ejphar.2021.174464
288. Yang G, Zhao Y, Zhang Y, Dang B, Liu Y, Feng N. Enhanced oral bioavailability of silymarin using liposomes containing a bile salt: preparation by supercritical fluid technology and evaluation in vitro and in vivo. Int j Nanomed. 2015;6633–6644. doi:10.2147/IJN.S92665
289. Yu Y, Lu Y, Bo R, et al. The preparation of gypenosides liposomes and its effects on the peritoneal macrophages function in vitro. Int J Pharm. 2014;460(1–2):248–254. doi:10.1016/j.ijpharm.2013.11.018
290. Liu H, Zhang Y, Han Y, et al. Characterization and cytotoxicity studies of DPPC: M2+ novel delivery system for cisplatin thermosensitivity liposome with improving loading efficiency. Colloids Surf B. 2015;131:12–20. doi:10.1016/j.colsurfb.2015.04.029
291. Kan S, Lu J, Liu J, Wang J, Zhao Y. A quality by design (QbD) case study on enteric-coated pellets: screening of critical variables and establishment of design space at laboratory scale. Asian J Pharm Sci. 2014;9(5):268–278. doi:10.1016/j.ajps.2014.07.005
292. Shamshiri MK, Jaafari MR, Badiee A. Preparation of liposomes containing IFN-gamma and their potentials in cancer immunotherapy: in vitro and in vivo studies in a colon cancer mouse model. Life Sci. 2021;264:118605. doi:10.1016/j.lfs.2020.118605
293. Handali S, Moghimipour E, Rezaei M, et al. A novel 5-Fluorouracil targeted delivery to colon cancer using folic acid conjugated liposomes. Biomed Pharmacother. 2018;108:1259–1273. doi:10.1016/j.biopha.2018.09.128
294. Sinha V, Mittal B, Bhutani K, Kumria R. Colonic drug delivery of 5-fluorouracil: an in vitro evaluation. Int J Pharm. 2004;269(1):101–108. doi:10.1016/j.ijpharm.2003.09.036
295. Gupta Y, Jain A, Jain P, Jain SK. Design and development of folate appended liposomes for enhanced delivery of 5-FU to tumor cells. J Drug Targeting. 2007;15(3):231–240. doi:10.1080/10611860701289719
296. Gobbo OL, Sjaastad K, Radomski MW, Volkov Y, Prina-Mello A. Magnetic nanoparticles in cancer theranostics. Theranostics. 2015;5(11):1249. doi:10.7150/thno.11544
297. Amin M, Badiee A, Jaafari MR. Improvement of pharmacokinetic and antitumor activity of PEGylated liposomal doxorubicin by targeting with N-methylated cyclic RGD peptide in mice bearing C-26 colon carcinomas. Int J Pharm. 2013;458(2):324–333. doi:10.1016/j.ijpharm.2013.10.018
298. Teymouri M, Farzaneh H, Badiee A, Golmohammadzadeh S, Sadri K, Jaafari MR. Investigation of Hexadecylphosphocholine (miltefosine) usage in Pegylated liposomal doxorubicin as a synergistic ingredient: in vitro and in vivo evaluation in mice bearing C26 colon carcinoma and B16F0 melanoma. Eur J Pharm Sci. 2015;80:66–73. doi:10.1016/j.ejps.2015.08.011
299. Amin M, Mansourian M, Burgers PC, Amin B, Jaafari MR, Ten Hagen TL. Increased Targeting Area in Tumors by Dual-Ligand Modification of Liposomes with RGD and TAT Peptides. Pharmaceutics. 2022;14(2):458. doi:10.3390/pharmaceutics14020458
300. Shahraki N, Mehrabian A, Amiri-Darban S, Moosavian SA, Jaafari MR. Preparation and characterization of PEGylated liposomal Doxorubicin targeted with leptin-derived peptide and evaluation of their anti-tumor effects, in vitro and in vivo in mice bearing C26 colon carcinoma. Colloids Surf B. 2021;200:111589. doi:10.1016/j.colsurfb.2021.111589
301. Neuberger K, Boddupalli A, Bratlie KM. Effects of arginine-based surface modifications of liposomes for drug delivery in Caco-2 colon carcinoma cells. Biochem Eng J. 2018;139:8–14. doi:10.1016/j.bej.2018.08.003
302. Yang S-J, Lin F-H, Tsai K-C, et al. Folic acid-conjugated chitosan nanoparticles enhanced protoporphyrin IX accumulation in colorectal cancer cells. Bioconjugate Chem. 2010;21(4):679–689. doi:10.1021/bc9004798
303. Cortese K, Marconi S, Aiello C, et al. Liposomes loaded with the proteasome inhibitor z-leucinyl-leucinyl-norleucinal are effective in inducing apoptosis in colorectal cancer cell lines. Membranes. 2020;10(5):91. doi:10.3390/membranes10050091
304. Udofot O, Affram K, Smith T, et al. Pharmacokinetic, biodistribution and therapeutic efficacy of 5-fluorouracil-loaded pH-sensitive PEGylated liposomal nanoparticles in HCT-116 tumor bearing mouse. J Nat Sci. 2016;2(1):1.
305. Azarifar Z, Amini R, Tanzadehpanah H, Afshar S, Najafi R. In vitro co-delivery of 5-fluorouracil and all-trans retinoic acid by PEGylated liposomes for colorectal cancer treatment. Mol Biol Rep. 2023;50:1–13.
306. VanOsdol J, Ektate K, Ramasamy S, et al. Sequential HIFU heating and nanobubble encapsulation provide efficient drug penetration from stealth and temperature sensitive liposomes in colon cancer. J Control Release. 2017;247:55–63. doi:10.1016/j.jconrel.2016.12.033
307. Simón M, Jørgensen JT, Norregaard K, et al. Neoadjuvant Gold Nanoshell-Based Photothermal Therapy Combined with Liposomal Doxorubicin in a Mouse Model of Colorectal Cancer. Int J Nanomed. 2023;Volume 18:829–841. doi:10.2147/IJN.S389260
308. Kim MS, Lee E-J, Kim J-W, et al. Gold nanoparticles enhance anti-tumor effect of radiotherapy to hypoxic tumor. Radiation Oncology Journal. 2016;34(3):230. doi:10.3857/roj.2016.01788
309. Tiwari A, Gajbhiye V, Jain A, et al. Hyaluronic acid functionalized liposomes embedded in biodegradable beads for duo drugs delivery to oxaliplatin-resistant colon cancer. J Drug Delivery Sci Technol. 2022;77:103891. doi:10.1016/j.jddst.2022.103891
310. Chen J, Hu S, Sun M, et al. Recent advances and clinical translation of liposomal delivery systems in cancer therapy. Eur J Pharm Sci. 2024;193:106688. doi:10.1016/j.ejps.2023.106688
311. Tejada-Berges T, Granai C, Gordinier M, Gajewski W. Caelyx/Doxil for the treatment of metastatic ovarian and breast cancer. Expert Rev Anticancer Ther. 2002;2(2):143–150. doi:10.1586/14737140.2.2.143
312. Bulbake U, Doppalapudi S, Kommineni N, Khan W. Liposomal formulations in clinical use: an updated review. Pharmaceutics. 2017;9(2):12. doi:10.3390/pharmaceutics9020012
313. Batist G, Barton J, Chaikin P, Swenson C, Welles L. Myocet (liposome-encapsulated doxorubicin citrate): a new approach in breast cancer therapy. Expert Opinion Pharmacother. 2002;3(12):1739–1751. doi:10.1517/14656566.3.12.1739
314. Smith JA, Costales AB, Jaffari M, et al. Is it equivalent? Evaluation of the clinical activity of single agent Lipodox® compared to single agent Doxil® in ovarian cancer treatment. J Oncol Pharm Pract. 2016;22(4):599–604. doi:10.1177/1078155215594415
315. Rosenthal E, Poizot-Martin I, Saint-Marc T, Spano J-P, Cacoub P, Group DS. Phase IV study of liposomal daunorubicin (DaunoXome) in AIDS-related Kaposi sarcoma. Am J Clin Oncol. 2002;25(1):57–59. doi:10.1097/00000421-200202000-00012
316. Passero Jr FC, Grapsa D, Syrigos KN, Saif MW. The safety and efficacy of Onivyde (irinotecan liposome injection) for the treatment of metastatic pancreatic cancer following gemcitabine-based therapy. Expert Rev Anticancer Ther. 2016;16(7):697–703. doi:10.1080/14737140.2016.1192471
317. Phuphanich S, Maria B, Braeckman R, Chamberlain M. A pharmacokinetic study of intra-CSF administered encapsulated cytarabine (DepoCyt®) for the treatment of neoplastic meningitis in patients with leukemia, lymphoma, or solid tumors as part of a phase III study. J Neuro-Oncol. 2007;81(2):201–208. doi:10.1007/s11060-006-9218-x
318. Rodriguez MA, Pytlik R, Kozak T, et al. Vincristine sulfate liposomes injection (Marqibo) in heavily pretreated patients with refractory aggressive non‐Hodgkin lymphoma: report of the pivotal Phase 2 study. Cancer: Interdiscip Int J Am Cancer Soc. 2009;115(15):3475–3482. doi:10.1002/cncr.24359
319. Akinc A, Maier MA, Manoharan M, et al. The Onpattro story and the clinical translation of nanomedicines containing nucleic acid-based drugs. Nature Nanotechnol. 2019;14(12):1084–1087. doi:10.1038/s41565-019-0591-y
320. Long HJ. Paclitaxel (Taxol): a novel anticancer chemotherapeutic drug. Mayo Clin Proc. 1994;69(4):341–345. doi:10.1016/S0025-6196(12)62219-8
321. Zhang Q, Huang X-E, Gao -L-L. A clinical study on the premedication of paclitaxel liposome in the treatment of solid tumors. Biomed Pharmacother. 2009;63(8):603–607. doi:10.1016/j.biopha.2008.10.001
© 2025 The Author(s). This work is published and licensed by Dove Medical Press Limited. The
full terms of this license are available at https://www.dovepress.com/terms.php
and incorporate the Creative Commons Attribution
- Non Commercial (unported, 3.0) License.
By accessing the work you hereby accept the Terms. Non-commercial uses of the work are permitted
without any further permission from Dove Medical Press Limited, provided the work is properly
attributed. For permission for commercial use of this work, please see paragraphs 4.2 and 5 of our Terms.
Recommended articles
Folate-Functionalized CS/rGO/NiO Nanocomposites as a Multifunctional Drug Carrier with Anti-Microbial, Target-Specific, and Stimuli-Responsive Capacities
Obireddy SR, Ayyakannu A, Kasi G, Sridharan B, Lai WF, Viswanathan K
International Journal of Nanomedicine 2025, 20:1965-1981
Published Date: 14 February 2025