Manosree
Chatterjee
*ab and
Nripen
Chanda
*a
aMaterial Processing and Microsystem Laboratory, CSIR – Central Mechanical Engineering Research Institute, Durgapur-713209, India. E-mail: manosree87@gmail.com; n_chanda@cmeri.res.in; Fax: +91-343-2546745; Tel: +91-9474112053 Tel: +91-9933034370
bDepartment of Biotechnology, National Institute of Technology Durgapur, Durgapur-713209, India
First published on 2nd December 2021
The eminence of nano-scale materials prevailed after the invention of high-resolution microscopes. Nowadays, nanoparticles are predominantly found in every application, including biomedical applications. In nanomedicine, the unique properties make nano-scale particles an efficient delivery vehicle to overcome the adverse effects of therapeutic molecules, which are directly administered. The polymeric nanoparticles have gradually gained interest as a nano-carrier over various non-polymeric types of nanoparticles due to their biocompatible nature. PLGA is the most frequently used polymer to synthesize polymeric nano-carriers as it is a clinically approved biodegradable polymer and has a broad scope of modification of its inherent properties. PLGA polymer, before or after nanoparticle formation, can be functionalized using various non-covalent and covalent modification techniques to suit desired applications. Since the beginning of PLGA nanoparticle usage, different synthesis methods have evolved progressively with various advantages and limitations. The present review also discusses the post-surface modification characterization of PLGA nanoparticles and their imaging and drug delivery applications.
These limitations can be addressed by polymer-based biocompatible nano-materials that have emerged as an alternative platform for drug delivery and imaging applications. Among all types of US Food and Drug Administration (FDA) approved nanoparticles used as a drug delivery system, polymeric nanoparticles are the most widely used in clinical practices.15 In addition to the biocompatibility, the polymeric nanoparticles offer a few other advantages, e.g., (a) enhancing the encapsulation of small molecules, (b) preventing degradation or deactivation of the drug in the bloodstream before reaching the target, (c) prolonging blood circulation time, (d) controlling the release of drugs in target tissues or cells, (e) improving the drug loading capacity, (f) increasing the bioavailability of drugs, and finally (g) speeding the passive accumulation of drugs at tumor sites based on the enhanced permeability and retention (EPR) effect.28–32 Besides, there is a flexibility to modify its pristine polymer chain and surface characteristics to enhance drug delivery and imaging efficacy, making them a superior therapeutic nano-platform.
The majority of the polymeric nanoparticles are fabricated using numerous biocompatible polymers to reduce the undesirable systemic toxicity of the drug transporter.33 Among these, FDA and European Medicine Agency (EMA) approved biodegradable PLGA polymer is most widely used as a versatile and clinically proved elemental polymer for the synthesis of efficient nano-carriers.34–36 Simultaneous drug enrichment to the tumor site and minimizing toxicity to normal tissues are indispensable aspects of nanoparticle-mediated drug delivery purposes. The use of biocompatible polymers like PLGA to achieve these aspects through nanoparticle synthesis has been an ever-growing arena in the field of safe drug delivery. PLGA polymer is composed of varying ratios of lactic acid and glycolic acid monomer units that are ester bonded to form the polyester polymer. In an in vivo system, PLGA polymer decomposed upon hydrolysis at the ester bond and eventually metabolized through the kerbs cycle with the nontoxic end products (H2O and CO2), which are eliminated from the body.37–39 The globally accepted and clinically approved PLGA polymer chains randomly orient themselves to form PLGA nanoparticles (Fig. 2). PLGA copolymers with low molecular weights (MW < 10 kDa) are synthesized by the polycondensation reaction of lactic acid and glycolic acid in various ratios and higher molecular weight copolymers (generally used 10–100 kDa) are synthesized by ring-opening polymerization of cyclic dimers (Fig. 2).35,39 Biodegradation of PLGA nanoparticles depends on the integral properties of the PLGA copolymer chain, which include the ratio of lactic acid and glycolic acid monomers in its chain composition and molecular weight.36,40 The drug release from PLGA nanoparticles by the degradation of PLGA copolymer can be regulated by tuning the ratio of lactic acid and glycolic acid monomers in the PLGA chain. If the ratio of lactic acid is increased in the PLGA chain, the hydrophobicity also increases proportionally, resulting in a slow degradation of PLGA as it absorbs less water.35,36,40 On the contrary, faster hydrolysis is observed when the glycolic acid content in PLGA is higher, resulting in the rapid release of drugs from nanoparticles. An acid-terminated PLGA chain with lower molecular weight and equal ratio of lactic acid (LA) and glycolic acid (GA) (50:
50 PLGA) is frequently used to prepare drug delivery vehicles due to its optimum degradation rate (less than two months at 37 °C in an aqueous medium).35,41,42 PLGA copolymer with higher molecular weight causing slower degradation rate of nanoparticles exhibits a slower drug release.40 Various clinically significant biomolecules are efficiently encapsulated inside the PLGA nanoparticle's core through weak covalent interactions for imaging and drug delivery. The physicochemical properties of nanoparticles predominantly depend on the composition and molecular weight of the PLGA polymer.39 The polymer can be dissolved in a wide range of organic solvents like acetone, dichloromethane, tetrahydrofuran, ethyl acetate, and chloroform, which is advantageous for nanoparticle synthesis.38 Moreover, PLGA nanoparticles prove its excellence as a nano-carrier system as it possesses a wide range of degradation rates that provide a desirable formulation opportunity, stability in long-term storage, and high encapsulation efficiency. The first targeting (prostate-specific membrane antigen (PSMA)–targeted) nanomedicine BIND-014 containing docetaxel, which was tested on humans, was prepared with PLGA polymer.43,44 Currently available PLGA polymer-based antitumor drugs approved for clinical practices to treat various types of cancer are listed in Table 1.45–51
![]() | ||
Fig. 1 The review's structure based on PLGA nano-carrier formulation, surface modification, and cancer therapeutic applications. |
Product (manufacturers’ name) | Drug name | Dosage type | PLGA content and LA![]() ![]() |
Drug dosages | Targeted tumor | Approved year |
---|---|---|---|---|---|---|
Decapeptyl® (Ferring Pharmaceuticals Pvt. Ltd.) | Triptorelin | Microsphere | 50![]() ![]() |
3.75 mg in every 28 days for 6 months | Prostate, breast | 1986 |
Zoladex Depot® (AstraZeneca UK Limited) | Goserelin acetate | Implant | 13.3–14.3 mg per dose; 50![]() ![]() |
3.6 mg in every 28 days | Prostate | 1989 |
Lupron Depot® (Abbvie Endocrine Inc) | Leuprolide | Microsphere | 33.1 mg per dose; 75![]() ![]() |
3.75 mg in every month | Prostate | 1989 |
Sandostatin Lar® (Novartis Pharmaceuticals UK Ltd) | Octreotide acetate | Microsphere | 188.8,377.6 and 566.4 mg per dose; 50![]() ![]() |
10 mg, 20 mg or 30 mg in every 4 weeks | Acromegaly | 1998 |
Trelstar® (Ferring Pharmaceuticals Pvt.) | Triptorelin pamoate | Microsphere | 136, 118, 182 mg per dose; 78![]() ![]() |
3.75 mg in every 4 weeks or 11.25 mg in every 12 weeks or 22.5 mg in every 24 weeks | Prostate | 2000 |
Eligard® (Zydus Cadila Healthcare Ltd.) | Leuprolide acetate | In situ forming implant | 82.5 mg per dose of 50![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
1 mg in every day 7.5 mg in a month 22.5 mg in every 3 months 30 mg in every 4 months 45 mg in every 6 months | Prostate | 2002 |
Signifor Lar® (Novartis Pharmaceuticals Corporation) | Pasireotide pamoate | Microsphere | 26.29, 52.58, 78.87 mg per dose of PLGA-50-60![]() ![]() ![]() ![]() |
20 mg, 40 mg, 60 mg in every 28 days | Acromegaly | 2014 |
In the present study, we review the synthesis, surface properties, and superiority of PLGA nano-carriers towards cancer therapeutic applications. It includes a comprehensive discussion on the existing PLGA nanoparticle synthesis methods, surface functionalization processes, and subsequent characterization techniques. It also covers the optimization and limitations of the existing PLGA nano-formulation procedures. Finally, we scrutinized the theranostic efficiency of the PLGA nano-carrier system, a promising alternative to conventional drugs. Then, we have summarized the reported research on the cancer therapeutic applications of the PLGA nano-carrier. The review's complete structure is depicted in Fig. 1. This review article provides an extensive understanding of all the milestones for making a PLGA nano-carrier, which is a potential cancer therapeutic system.
Name of the method | Parameters that affect nanoparticles size | LA![]() ![]() |
Drug loading (%) and encapsulation efficiency (%) | Advantages | Limitations | Ref. |
---|---|---|---|---|---|---|
Conventional methods | ||||||
Emulsion-solvent evaporation | Polymer concentration, speed of agitation, stabilizer concentration | 50![]() ![]() ![]() ![]() ![]() ![]() |
0.022–8 and 20–80 | Spherical morphology, easy and rapid procedure of nanoparticle fabrication, colloidal stability | Heterogeneous in size, residual stabilizer remains, particle agglomeration, drugs may lose activity due to high shear stress | 52–57,59,60 |
Nano-precipitation | Polymer concentration, speed of agitation, stabilizer concentration | 50![]() ![]() ![]() ![]() ![]() ![]() |
1.7–10 and 40–90 | High yield, easy and reproducible | High polydispersity index, use of stabilizer, high speed agitation may degrade drug molecules | 61–66 |
Spray-drying | Spray mesh hole size, concentration of the polymer, density of the spray liquid, flow rate | 50![]() ![]() ![]() ![]() |
1.5–7.4 and 65–90 | It produced powder nanoparticles, stable in storage because free of moisture, nanoparticles produced free from contamination of other chemicals | Degradation of the temperature sensitive drug due to the high heat, high operating cost, and agglomeration of nanoparticles | 67–73 |
Salting out | Polymer concentration, stirring speed, stabilizer concentration, concentration of salting out agent | 50![]() ![]() ![]() ![]() |
5 and 55–80 | Use of nontoxic oil phase, nanoparticles size can be controlled by adjusting different parameters | Purifying the nanoparticles is very hectic due to the presence of salting out agent, high speed agitation may result in loss of drug activity | 74–76 |
Nonconventional methods | ||||||
Microfluidic | Channel geometry, flow rate ratio of the continuous phase and dispersed phase, interfacial tension between two phases, Mixing time | 50![]() ![]() ![]() ![]() |
10–18 and ∼90% | Narrow size distribution, reproducible | Swelling of PDMS polymer alters channel geometry, very low yield | 77–80 |
Template patterns, mold preparation | 50![]() ![]() ![]() ![]() |
1–40 and > 90% | Monodispersed particle, high encapsulation efficiency, reproducible | Low yield, degradation of clinically important fragile molecules during solidification | 81–86 | |
Electrospray | Concentration of the polymer, nature of solvent, needle diameter flow rate, potential difference and distance between needle and grounded electrode | 50![]() ![]() |
5–43 and > 90% | High yield in a short duration of time, single step method, surfactant and high speed agitation free process | 87–89 |
![]() | ||
Fig. 3 Synthesis of PLGA nanoparticles by (a) single emulsion-solvent evaporation method and (b) double emulsion solvent evaporation method. |
In the single emulsion solvent evaporation method, both PLGA and hydrophobic drugs are dissolved in a non-polar organic solvent, which is then added drop-wise to the aqueous solution of surfactants (stabilizers) with continuous agitation (ultra-sonication or homogenization). The high shear stress disperses the oil–water emulsion into nanoparticles, which hardens after evaporation of the organic solvent by continuous stirring (Fig. 3a).56,57 For the encapsulation of hydrophilic drugs, the PLGA solution is drop-wise added to the aqueous drug solution with continuous stirring that produces the first emulsion. Then, the first emulsion is transferred to the aqueous solution of surfactants (water in oil in water) under ultra-sonication or homogenization, which forms the second emulsion that finally transforms into the hydrophilic drug encapsulated PLGA nanoparticles (Fig. 3b).55,57 The nanoparticles are solidified after the evaporation step like the single emulsion method. The vigorous mechanical agitation step in the stabilizer containing aqueous medium disperses the emulsion into nano-droplets of varying sizes with a layer of the stabilizer surrounding it. Nanoparticle size increases with the increase in PLGA concentration. At higher PLGA concentrations, the viscosity of the organic phase obstructs the disruption of the emulsion into very small-sized nano-droplets; as a result, a high number of PLGA polymer resides in the droplet, which eventually produces nanoparticles with a larger size.58,59 Manchanda et al. reported that a higher concentration of the stabilizer in an aqueous medium increases the overall shear force on the emulsion droplets by reducing the organic/aqueous interfacial tension, which eventually helps the formation of nanoparticles with a smaller mean diameter. They also reported that the drug encapsulation efficiency increases with the increase in PLGA concentration as high viscosity resists the diffusion of the drug into the aqueous medium. A high polymer ratio also provides a dense network to trap the drug molecules.60 The agitation speed also has a remarkable effect on the size of the nanoparticles. The high agitation speed produces smaller nanoparticles by rupturing the emulsion droplets into smaller ones containing lesser PLGA polymer. Kadriye Kizilbey conducted a series of experiments to optimize the parameters to encapsulate hydrophobic drugs inside PLGA nanoparticles using the single emulsion solvent evaporation method. He reported that the diameter and encapsulation efficiency of nanoparticles increase with the increase in PLGA concentration. At the same time, the increasing concentration of PVA (stabilizer) has a similar effect on size but has a reverse effect on encapsulation efficiency.56
Every type of polymer with a specific molecular weight has a particular number of chain entanglements in a specific type of solvent, which increases with the increasing polymer concentration. As the concentration of the polymer increases, the viscosity of the polymer solution gets enhanced, producing a higher chain entanglement density that is physical overlapping of the polymer chain. Above the critical chain overlap concentration, the electrospinning that is fiber formation begins by inhibiting Rayleigh disintegration of the droplets. Polymer concentration below the critical chain overlap concentration favors electrospraying. A lower PLGA concentration decreases the nanoparticle's size as there is no or low chain entanglement in the solution.88 The solvent should be highly volatile and conductive in electrospray-mediated polymer nanoparticle synthesis. During the flight of the nano-droplets, the whole solvent must completely evaporate before reaching the collector.88 A specific flow rate of the spray solution through the conductive needle is indispensable to get a stable spray for nanoparticle formation. For the synthesis of monodispersed nanoparticles, a steady flow with a sufficiently low rate is required. At a high flow rate, an intermittent jet is produced instead of an electrospray. The particle size also depends on the flow rate. Smaller particles are formed from the slower flow rate.88 Stable electrospraying depends on the needle diameter when the flow rate, applied voltage, and electrospray setup are stable. A smaller needle diameter produces more stable and smaller spherical nanoparticles.88 The stable coulomb fission can be obtained only above a certain applied voltage at a fixed flow rate and electrospray setup. As the voltage increases, the spray solution's dribbling from the conductive needle starts to be finer and eventually becomes nano-spray.88 The distance between the conductive needle and the grounded electrode decides the electric field strength, which influences the droplets’ rupturing. Shorter distance intensifies the electric field strength, which leads to smaller-sized particle formation when other parameters are fixed. A sufficient distance is also required for the complete evaporation of the solvent. If the distance becomes more, then a higher voltage is required to overcome the spray jet.88
Type of interaction | Conjugating molecule | Coupling agent | Reaction environment | Type of bond formation | Ref. |
---|---|---|---|---|---|
Non-covalent | |||||
Hydrophobic | Hydrophobic biomolecules | — | Depends on the conjugating molecule | Hydrophobic | 92 |
Electrostatic | Positively charged biomolecules | — | Depends on the conjugating molecule | Electrostatic | 92 |
Hydrogen bond | Biomolecules with carboxyl, amine and hydroxyl groups | — | Depends on the conjugating molecule | Hydrogen bond | 92 |
Avidin–Biotin | Any type of biomolecules | Avidin and Biotin | Wide range of pH | Protein and ligand | 93–95 |
Covalent | |||||
Carbodiimide coupling reaction | Biomolecules with primary amine group | Carbodiimides | pH < 7.2 | Amide bond | 91,98–101 |
Thiol-maleimide coupling reaction | Biomolecules with sulfhydryl or thiol group | Maleimide | pH 6.5–7.5 | Thioether bond | 106–109 |
Copper-catalyzed azide–alkyne cycloaddition (CuAAC) reaction | Biomolecules linked with alkyne group | Azide and alkyne | In the presence of copper(I) catalyst at pH 7.2 | 5-Membered heteroatom ring (1,2,3-triazole) | 117,119,121 |
![]() | ||
Fig. 10 Non-covalent interactions for the modification of PLGA nanoparticles (a) hydrophobic interaction, (b) electrostatic interaction, (c) hydrogen bonding, (d) avidin–biotin interaction. |
Hydrodynamic size and charge of the PLGA nano-carrier are analyzed using a particle size analyzer by the dynamic light scattering technique.128 The size and charge (Zeta (ζ) potential) are the determining factors to cross the cell-associated barrier for delivering the drug or tracking through an imaging agent.129 Nano-carrier with size less than 200 nm internalized efficiently by the cell through endocytosis and EPR (enhanced permeability and retention) effect of tumor vasculature. Nanoparticles with a size greater than 200 nm were prematurely eliminated from the body by the reticuloendothelial system (RES).130,131 A higher ζ potential of PLGA nanoparticles provides colloidal stability of the nanoparticles. The higher surface charge of the nanoparticles produces electrostatic repulsion; thus, they remain suspended in solution, which prevents agglomeration and maintains the size.132 Nanoparticles were analyzed using atomic force microscopy (AFM), field emission scanning electron microscopy (FESEM), and transmission electron microscopy (TEM) to study the nanoparticle's shape, increased diameter, and surface texture after surface modification (Fig. 12).90,91 The nanoparticle's shape and morphology depend on the synthesis procedure, which has an effect on the cellular uptake of the nanoparticles. A three-dimensional view and line roughness graph of nanoparticles with the sub-nanometer resolution is observed under AFM in atmospheric or submerged conditions.133 FESEM and TEM analyses also provide information about elemental composition by energy-dispersive X-ray spectroscopy (EDX).134
![]() | ||
Fig. 12 (a) AFM, (b) FESEM and (c) TEM images of 1-pyrenebutyric acid conjugated PLGA nanoparticles and the same after methotrexate conjugation on the nanoparticles surface through amide linkage shown in (d–f), respectively. The inset shows the corresponding particle size distribution histogram where the increase in particle size establishes the conjugation (Reproduced from ref. 91 with permission from the Royal Society of Chemistry). |
Ultraviolet-visible spectroscopy (UV-vis spectroscopy) is used to study the encapsulated and conjugated ligands/biomolecules in PLGA nano-carriers that absorb light in the UV or visible regions of the electromagnetic spectrum.135 In a complementary way of UV-vis spectroscopy, fluorescent molecule conjugated PLGA nanoparticles absorb a specific wavelength of light (usually ultraviolet light), then re-emit the light to return from electronically excited states to the ground state, which is detected by fluorescence spectroscopy (Fig. 13).136 These techniques are used to quantify the conjugates present in modified PLGA nanoparticles to measure the loading and encapsulation efficiency. The wavelength of maximum absorption (λmax) and maximum emission (λem) are utilized to study the release kinetics of the drug molecules carried by the PLGA nanoparticles. These optical characteristics help detect, quantify the concentration, and find the level of degradation of the nano-carrier in biological samples.137
![]() | ||
Fig. 13 Fluorescence spectra of PLGA, 1-pyrenebutyric acid (PBA) conjugated PLGA nanoparticles (PLGA-PBA), and methotrexate (MTX) conjugated PLGA-PBA nanoparticles (PLGA-PBA@MTX). The images of nanoparticles suspensions under UV lamp (λmax = 265 nm) are represented in the inset (Reproduced from ref. 91 with permission from the Royal Society of Chemistry). |
Elemental characterization provides a deeper understanding of physical, chemical, and biological phenomena of the nano-carrier by the Fourier-transform infrared spectroscopy (FTIR), mass spectroscopy, nuclear magnetic resonance (NMR) spectroscopy, X-ray photoelectron spectroscopy (XPS), and Raman spectroscopy.90,91 The most common technique to determine the elemental composition and type of bonds between the molecules in the nano-carrier by their unique stretching frequencies is FTIR analysis (Fig. 14).138 The charge to mass ratio of fragmented ions of the nanoparticles during electrospray ionization in the mass spectroscopic analysis technique helps detect the presence of certain molecules and to predict the structure.139 Proton (1H) and carbon-13 (13C) NMR spectroscopy are generally used to establish the purity, detect the molecular structure, formation of bonds during conjugation, and analyze the molecules’ chemical environment in the nano-conjugate (Fig. 14).127,135 The chemical state and the electronic state of the elements within the nano-carrier are analyzed using the XPS technique.135 Raman spectroscopy is employed to analyze the chemical composition and structure of the nano-carrier through a non-contact and non-destructive way by the molecular fingerprint that is the signature vibrational, rotational, and other low-frequency modes of biomolecules.140
![]() | ||
Fig. 14 (a) FTIR spectrum of PLGA polymer, 1-pyrenebutyric acid (PBA) conjugated PLGA polymer (PLGA-PBA) and methotrexate (MTX) conjugated PLGA-PBA nanoparticles (PLGA-PBA@MTX) showing chemical conjugation of PBA and MTX with PLGA before and after nanoparticle formation. 1H NMR analysis of all the conjugation step i.e. (b) PLGA polymer (c) PLGA-PBA polymer and (d) PLGA@MTX nanoparticles (Reproduced from ref. 91 with permission from the Royal Society of Chemistry). |
Analysis of the nano-carrier through all these characterization procedures makes a complete vision about the complete characteristics, which helps determine their mode of effect in further application phases and their limitations.
The research works carried out in the last decade on PLGA nano-carriers and their applications in cancer therapy are discussed in Table 4.
Nanoparticle formulation | LA![]() ![]() |
Synthesis method | Size (nm) and encapsulation efficiency | Application on and administration route | Ref. |
---|---|---|---|---|---|
Poly(L-lysine)-poly(ethylene glycol)-folate (PLL-PEG-FOL) adsorbed Fe3O4 or CdSe/ZnS and DOXO encapsulated PLGA nanoparticle | Fe3O4 nanocrystals, CdSe/ZnS nanocrystals and Doxorubicin (DOXO) | Single-emulsion solvent evaporation | 100–200;— | KB cancer cells for MR and optical imaging and drug delivery | 162 |
VCR and VRP encapsulated PLGA nanoparticle | 75![]() ![]() |
Combining single emulsion solvent evaporation and salting-out method | 98.8 ± 8.4; 67.86 ± 5.10% for VCR and 80.29 ± 4.55% for VRP | Multidrug resistant breast cancer cells (MCF-7/ADR) for drug delivery | 163 |
POSS-containing conjugated polymer (CP) loaded and Herceptin conjugated PLGA nanoparticle | 50![]() ![]() |
Single-emulsion solvent evaporation | 230 ± 3–243 ± 6; 44% | Breast cancer cells (SKBR-3, MCF-7) for optical imaging | 164 |
Herceptin conjugated, Hydrophobic and hydrophilic drugs encapsulated, MNP embedded PLGA nanoparticle | 65![]() ![]() |
Double emulsion solvent evaporation | 304 ± 4.1; Pac (80.6 ± 2.7%) + Rapa (86.6 ± 3.1%) and 310 ± 3.9; Pac (83.5 ± 3.0%) + Carbo (47.8 ± 1.5%) | Breast cancer cells (MCF-7), pancreatic cancer cells (PANC-1) and rat model for MR imaging and drug delivery; administered through saphenous vein | 165 |
AS1411 aptamer conjugated and paclitaxel-loaded PLGA nanoparticle | 50![]() ![]() |
Single-emulsion solvent evaporation | 200; — | Human glial cancer cells (GI-1 cells) for drug delivery | 166 |
Curcumin and bortezomib co-encapsulated and alendronate (Aln). conjugated PLGA nanoparticles | 50![]() ![]() |
Single-emulsion solvent evaporation | 235 ± 70.30; — | Intraosseous mice model of bone metastasis of breast cancer for drug delivery; administered through tail vein | 167 |
Cyclic peptide (cRGD)-modified monomethoxy (polyethylene glycol)-PLGA-poly (L-lysine) nanoparticle encapsulated either DHAQ or Rb | 50![]() ![]() |
Double emulsion solvent evaporation | 180; 85.3% | Breast cancer cells (MDA-MB-231) for optical imaging or drug delivery | 168 |
siRNA encapsulated and lipid coated PLGA nanoparticle | Small interfering RNAs (siRNA) | 207 ± 4.461; 46% | Human cervical (HeLa), prostate (PC3, DU145, LNCaP) and liver (HepG2) cancer cells for siRNA delivery | 85 | |
Curcumin loaded PLGA nanoparticles | 50![]() ![]() |
Single-emulsion solvent evaporation | 120; 80% | Breast cancer cells (MCF7) for drug delivery | 169 |
DOX and VER combined chitosan shell coated MNPs encapsulated and cRGD peptide functionalized PLGA nanoparticle. | 50![]() ![]() |
Double emulsion solvent evaporation | 144; 74.8% for DOX and 53.2% for VER | Male S-180 sarcoma-bearing mice for drug delivery; administered through tail vein | 170 |
Holo-transferrin conjugated and bortezomib-loaded PLGA nanoparticle | 50![]() ![]() |
Single-emulsion solvent evaporation | 200; 53 ± 6% | Human pancreatic cancer cells (SUIT-2) for drug delivery | 171 |
DOX encapsulated and low-molecular-weight protamine-surface modified PLGA nanoparticle | Doxorubicin (DOX) | Nanoprecipitation | 206.2; 83% | Mice harboring drug-resistant breast tumors for drug delivery; administered through tail vein | 172 |
Tam encapsulated herceptin conjugated polyvinyl-pyrrolidone coated PLGA nanoparticle | 50![]() ![]() |
Double emulsion solvent evaporation | 93.44; 72.4 ± 2.3% | Human breast cancer cells (MCF-7) for drug delivery | 173 |
BSA-Gd complexes and DOX encapsulated and poly(ethylene glycol) conjugated PLGA nanoparticle | Bovine serum albumin gadolinium (BSA-Gd) complexes and Doxorubicin (DOX) | Single-emulsion solvent evaporation | 280; 20.9% | Human cervical cancer cells (HeLa) and female nude mice bearing tumor for MR imaging and drug delivery; administered through tail vein | 174 |
SPION, QDs and the anticancer drug busulfan encapsulated PLGA nanoparticle | 50![]() ![]() ![]() ![]() |
Single-emulsion solvent evaporation | 93; 89 ± 2% | Murine macrophage cells (J774A) and rat model for MR and optical imaging; administered through intravenous mode | 175 |
Antisense-miR-21 and antisense-miR-10b co-loaded urokinase plasminogen activator receptor (uPAR) conjugated PLGA nanoparticle | 50![]() ![]() |
Double emulsion solvent evaporation | 100 to 200; 72.4 ± 6.2% | Triple negative breast cancer (TNBC) cells and TNBC tumor xenografts in nude mice for antisense-miRNAs delivery; administered through intravenous mode | 176 |
Nutlin-3a loaded EpCAM aptamer and quantum dots conjugated PLGA nanoparticle | 50![]() ![]() |
Single-emulsion solvent evaporation | 292 ± 10; 51.24 ± 6.7% | Human breast cancer cells (MCF-7 and ZR751) and ovarian cancer cells (SKOV3) for fluorescence imaging and drug delivery | 177 |
T7-peptide conjugate, MNPs, PTX and CUR co-encapsulated PLGA nanoparticle | 50![]() ![]() |
Single-emulsion solvent evaporation | 130; 68% for PTX and 18% for CUR | Human malignant glioma (U87) and mice bearing orthotopic glioma (U87-Luc) for drug delivery and MRI; administered through tail vein | 178 |
DOX-loaded lipid hybrid PLGA nanoparticles | 75![]() ![]() |
Double emulsion solvent evaporation | 198 ± 12; 86.4 ± 8.5% | Human breast cancer cells (MDA-MB-231/ADR) and human squamous carcinoma cells (KB) for drug delivery | 179 |
Superparamagnetic iron oxide (SPIO3 NPs4) loaded PLGA nanospheres | Oleic acid-coated superparamagnetic iron oxide (SPIO3 NPs4) | Multiple emulsion solvent evaporation method | 130; 90.2 ± 0.3% | T1-weighted MRI scans of C26 colon carcinoma xenograft model; administered through tail vein | 180 |
Rhodamine B encapsulated PLGA nanoparticles | 50![]() ![]() |
Dewetting technique | 80; 93.26% | Human A549lung cancer cells for fluorescence imaging | 181 |
AS1411 aptamer conjugated curcumin and SPIONs encapsulated PLGA nanocapsule | Curcumin and superparamagnetic iron oxide nanoparticles (SPIONs) | Nanoprecipitation | 150; — | Pancreatic cancer cells (PANC-1 and MIA PaCa-2) for optical, MRI, and photoacoustic imaging | 182 |
CN-PPV and NIR dye encapsulate PLGA nanoparticle | 50![]() ![]() |
Single-emulsion solvent evaporation | 50; — | Cervical cancer cells (HeLa) for optical imaging | 183 |
Doxorubicin encapsulated and Cy5.5 labeled PLGA nanoparticle | 50![]() ![]() |
Double emulsion solvent evaporation | 114; ∼80% | U87 human glioma cell line for optical imaging and drug delivery | 184 |
Polyethyleneimine-polyethylene glycol-folic acid functionalized quantum dots, Fe3O4 nanocrystals, and doxorubicin (DOX) encapsulated and (shRNA) adsorbed PLGA nanocomposites | 50![]() ![]() |
Double emulsion solvent evaporation | ∼300 nm; ∼62.97% for DOX | Cervical cancer cells (HeLa) and subcutaneous EMT-6 tumor xenograft mice model for MR and fluorescence imaging and drug delivery; Administered through intratumoral injection | 185 |
Curcumin encapsulated PLGA nanoparticle | 50![]() ![]() |
Microfluidic | 30–70; 67% | Leukemia Jurkat cells for drug delivery | 186 |
Chitosan and PEG-coated curcumin-loaded PLGA nanoparticles (CNPs) | 50![]() ![]() |
Single-emulsion solvent evaporation | 264 nm; 60% | Human pancreatic cancer cell lines PANC-1 and Mia Paca-2 for drug delivery | 187 |
OX26 type monoclonal antibody functionalized TMZ encapsulated PLGA nanoparticle | 50![]() ![]() |
Single-emulsion solvent evaporation | 176; 48 ± 10% | Glioblastoma cells (U215 and U87) for drug delivery | 188 |
DOX encapsulated and Au nanoparticle decorated PLGA nanoparticle | 50![]() ![]() |
Double emulsion solvent evaporation | ∼160; — | Mouse colon cancer cells (CT26), Murine breast cancer cells (4T1) and mice bearing 4T1 tumors for photoacoustic imaging and drug delivery; administered through intravenous mode | 189 |
Transferrin decorated paclitaxel and elacridar co-encapsulated PLGA nanoparticle | Transferrin and elacridar | Nanoprecipitation | 226.9; 76% | Drug-resistant breast cancer cells (EMT6/AR1.0) for drug delivery | 190 |
MTX and CUR co-encapsulated PLGA nanoparticle | Methotrexate (MTX) and curcumin (CUR) | Double emulsion solvent evaporation | 142.3 ± 4.07; MTX 71.32 ± 7.8% and CUR 85.64 ± 6.3% | Breast cancer cells SK-Br-3 cell line and chemically induced mammary tumors in female Sprague Dawley rats for drug delivery; administered through intravenous mode | 191 |
iRGD conjugated and PTX-loaded PLGA nanoparticle | Paclitaxel (PTX) | Single-emulsion solvent evaporation | 147.5 ± 9.5; 88.2% | Human colorectal cancer cells (LS174T, COLO205, HCT116, and SW620) and LS174T tumor-bearing BALB/c (nu/nu) mice for drug delivery; Administered through intravenous mode | 192 |
Curcumin and Niclosamide encapsulated PLGA nanoparticle | 50![]() ![]() |
Nanoprecipitation | 225.9; 58.09% for Curcumin and 85.36% for Niclosamide | Breast cancer cells (MDA-MB-231) for drug delivery | 193 |
Epidermal growth factor functionalized 5Fu and perfluorocarbon (PFC) co-loaded PLGA nanoparticle | 5-Fluorouracil (5Fu) | Double emulsion solvent evaporation | 200; 81.6 ± 5.7% | Human colon cancer cells (SW620) for drug delivery | 194 |
Cannabidiol (CBD) loaded PLGA nanoparticle | Cannabidiol (CBD) | Single-emulsion solvent evaporation | 240; 95% | Epithelial ovarian cancer cells (SKOV-3) for drug delivery | 195 |
Sal and Tam encapsulated (PEG)–PLGA nanoparticle | 75![]() ![]() |
Double emulsion solvent evaporation | 275.3 ± 44.0; Sal 32.63% ± 0.73% and Tam 49.18% ± 3.04% | Mouse breast cancer cell line (4T1) and female BALB/c mice for drug delivery; administered through intraperitoneal mode | 196 |
Tg conjugated PLGA nanoparticle | 50![]() ![]() |
Electrospray | ∼60 nm; 97.22% | Cervical cancer cells (HeLa) for drug delivery and fluorescence imaging | 90 |
PBA conjugated and MTX decorated PLGA nanoparticle | 50![]() ![]() |
Electrospray | ∼105 nm 91.4% | MTX resistant metastatic breast cancer cells (MCF-7 and MDA-MB-231) for fluorescence imaging and drug delivery | 91 |
Molybdenum octahedral clusters encapsulated PLGA nanoparticle | 50![]() ![]() |
Single-emulsion solvent evaporation | 75.7–144.7; 29.2–73.9% | Ovarian cancer cell line (A2780) for photodynamic therapy (PDT) | 197 |
PNAs encapsulated PLGA nanoparticle | Short cationic peptide nucleic acids (PNAs) | Double emulsion solvent evaporation | 145; — | HeLa, A549, HEK-293, SUDHL-5, U2932 cell line and xenograft mouse model for drug delivery; administered through tail vein | 198 |
There is a vast research scope to continue this exploration with modern drug molecules that can be conveyed using PLGA nano-carriers and unveil their applications beyond the cellular model. Their behavior may be studied in in vitro 3D tumor spheroid, which mimics the complex in vivo tumor vasculature on a benchtop. It would be beneficial for the real-time study of drug delivery and imaging efficacy of nano-carriers. This type of study may also provide information regarding tumor vasculature penetration capability, size reducing efficiency of the tumor, cytotoxic efficiency, and monitoring ability of the therapeutic responses. Finally, the nano-carrier efficiency may be studied in clinical trials to understand better the effectiveness as a theranostic nano-system in cancer management.
This journal is © The Royal Society of Chemistry 2022 |