Open access peer-reviewed chapter

Selection and Role of Polymers for Designing of a Drug Carrier

Written By

Pankaj Sharma, Vinay Jain and Mukul Tailang

Submitted: 20 January 2022 Reviewed: 08 February 2022 Published: 19 March 2022

DOI: 10.5772/intechopen.103125

From the Edited Volume

Drug Carriers

Edited by Luis Jesús Villarreal-Gómez

Chapter metrics overview

251 Chapter Downloads

View Full Metrics

Abstract

Polymers have helped to develop drug carrier technologies by allowing for the regulated release of bioactive molecules in consistent dosages over extended periods of time, cyclic dosing, and adjustable delivery of both hydrophobic and hydrophilic medicines. Formulations are released in a coordinated and consistent fashion over long periods of time. Polymers going to act as just an inert carrier whereby a substance can be conjugated having significant advantages. For instance, the polymer enhances the pharmacodynamic and pharmacokinetic characteristics of biopharmaceuticals in a variety of ways, such as plasma half-life, reduces immunogenicity, increases biopharmaceutical consistency, enhances the solubilization of low-molecular-weight substances, and has the prospects for targeted delivery. Smart polymeric delivery systems, in instance, have been investigated as “smart” delivery methods capable of releasing encapsulated pharmaceuticals at the right time and place of activity with respect to certain physiological stimuli. The development of novel polymeric materials and cross-linkers that are more biocompatible and biodegradable would expand and improve present uses. Polymer sensitivity to a particular stimulus may be tuned within a limited range because of the diversity of polymer substrates and their sequential production. The methods through which polymer frameworks are formed in situ to construct implanted systems for continuous release of medicinal macromolecules are discussed in this chapter, as well as numerous applicability of enhanced drug delivery.

Keywords

  • polymeric material
  • drug delivery
  • thermally responsive
  • smart polymer
  • glucose
  • enzyme
  • oxidation-reduction

1. Introduction

Since the 1980s, researchers have been working on polymeric drug delivery systems [1, 2, 3, 4]. Several of the frontier scientific fields are the hunt for novel medication delivery mechanisms and novel action mechanisms. These include multidisciplinary research techniques that aim to make significant improvements in therapeutic efficacy and bioavailability just at point of medication administration [5, 6]. One or more traditional medication delivery mechanisms are combined with engineering technologies in a drug delivery system. The technologies allow for precise targeting of the place in the body where a medicine has been delivered and/or the pace at which it has been released.

Short half-lives, low bioavailability, and physicochemical instability are all common limitations of biopharmaceutical therapies. Physiological instability is characterized by changes in the highly organized structure of proteins, which can result in undesired events including denaturation, aggregation, and precipitation. The chemical instability of pharmaceuticals is exacerbated by processes including such oxidation, deamidation, hydrolysis, and racemization. Stimulus-responsive polymers provide a pharmaceutical delivery mechanism for delivering pharmaceuticals at a regulated pace and in a durable and physiologically functional state. Research in stimuli-responsive polymers has grown over the years, and a lot of effort has gone into designing eco-friendly macromolecules that may be molded into novel smart polymers [7]. A composition or platform that allows the administration of a medicinal chemical into the body is known as just polymeric drug carriers. By regulating the pace, duration, and location of medication distribution in the system, it increases its effectiveness and safety. There in previous two decades, delivery of drugs has progressed significantly, but regulating medication entrance into the system, particularly the brain, has remained a tough challenge. Recent development in investigations of nano-drug delivery system distribution across the blood-brain membrane via carrier-mediated carriage is starting to give a reasonable basis for directing medication delivery to the brain. Natural materials such as amino acids, hexose, peptides, monocarboxylate, and stem cells are transported over the blood-brain membrane via ingestion transporters [8, 9, 10]. In the type of biomaterials with liposomes, polymers in reservoir-containing drug delivery applications have made tremendous development. Additional applications of the polymers include diffusion-based drug delivery systems and solvent-triggered/activated drug delivery systems. Drugs are dissolved in a non-swellable solution or a completely inflated matrix that does not breakdown throughout their engagement period in diffusion-based drug delivery applications. Whenever subjected to an aquatic media, solvent triggered materials such as hydrogels expand and release the medication. They are naturally hydrophilic. Because of their well-engineered polymeric by the changes in the underlying reasons of the biological function, biocompatible polymers provide a safe pathway for medication transport. Biodegradable polymers disintegrate owing to the breakage of covalent bonds among them, whereas bioerodible polymers cause degradation of the polymer owing to the dissolving of connecting strands without causing any changes in the molecule’s chemical properties. Aqueous soluble, safe, as well as non-immunogenic polymers are being used as therapeutic carriers. They act in the background to reduce medication breakdown and increase circulation time. Another crucial consideration is the drug’s appropriate elimination. If indeed the polymer is non-degradable, it really should be avoided accumulating in the body, and if it is biodegradable, the fragmented elements should be safe and not cause an immune reaction. Polymers that resemble important biological respond to environmental stimuli such as changes in pH or thermal by altering features such as solubility, hydrophobic/hydrophilic equilibrium, biomolecule (pharmaceutical component) releases, as well as configuration [11, 12].

The polymeric medicinal delivery compositions are classified into several classes, such as, biodegradable (chemically-controlled), diffusion-controlled, externally-responsive systems (e.g., temperature pH,) [13], solvent-actuated [14] and nanosized polymeic delivery platform that accomplish in three prime technologies [15]: (i) PEGylation [16, 17], (ii) active targeting of certain cells and organs [18, 19, 20] and (iii) Increased permeability and retaining allows for passive targeting effect [20, 21]. The more sophisticated polymeric therapeutic delivery technologies are indeed being anticipated as multidimensional fully – featured systems that will enable instantly improved pharmacokinetics, decreased toxicity, faster targeting, as well as a programmable drug release pattern. Furthermore, greater appropriate therapy might be provided by combination treatment, which involves the simultaneous administration of 2 or more medicaments/diagnostics substances [22, 23, 24]. In reaction to a modest external/internal stimulation, a stimuli-reactive or smart polymer changes its physical characteristics abruptly. Although minor changes take place in subjected to external/internal stimuli stimulus until a crucial limit is found, and they have the potential to revert to their original form when the stimulus is withdrawn, those polymers are indeed known as smart polymers [25, 26, 27]. The uniqueness of these polymers resides in their unpredictable reaction, which is initiated by a really tiny stimulus and results in enormous structural changes. Different triggers responsible for modulating the release of the drug using innovative polymeric drug delivery compositions are depicted in Figure 1. Modifications in physical state, structure, solubility, solvent interactions, aqueous soluble and lipid soluble equilibrium, and conductance are all reversible transitions. The introduction of oppositely charged polymers or a pH change to neutralize charged groups, as well as variations in the water-loving/lipid-loving balance or hydrogen bonding owing to temperature differences, are the driving factors underlying such transitions. Fewer dosage periodicities, simplicity of preparation, preservation of optimal therapeutic level at a single dose, longer delivery of integrated medication, decreased adverse effects, and increased stability are all advantages of innovative polymer-based medicaments delivery systems [28, 29, 30].

Figure 1.

Stimuli and materials that respond to them.

A dynamic polymeric material can respond in a variety of ways. The breakdown and development of numerous secondary interactions such as hydrogen bonding, van der Waals forces, hydrophobic forces, and electrostatic interaction [31, 32] restrict the responsiveness of such a polymeric solution induced by physicochemical stimuli. Fundamental processes including acid-base reaction, reduction, oxidation, and hydrolysis of components linked to the polymer chain are examples of chemical processes. Destruction of a polymeric structure owing to irreversible bond breaking in response to external stimuli is one example of the significant conformational shift in the polymer backbone. Biodegradability and biocompatibility; sustained-release characteristics; drug-loading potential; the dearth of deleterious characteristics including systemic toxicity, carcinogenic effects, immunogenicity, and reproductive toxicity; as well as outstanding stability characteristics are all important characteristics of a smart polymer.

Advertisement

2. Criteria for choosing a polymeric system

2.1 Polymers with temperature sensitivity

These are polymeric frameworks that are susceptible to thermal fluctuations. These polymers exhibit a gel-to-gel shift as temperature dependent, and can be used to deliver medicinal compounds in vivo. This sort of system seems to have a crucial temperature of the solution (usually in aqua) where the polymer and solution phases shift according to respective content. The solubility of several polymers varies dramatically as a result of ambient temperature. This characteristic was used to create aqueous solutions of these polymeric materials that go through a sol-gel changeover when the temperature varies. A maximum crucial solution temperature (MaxCST) exists for thermally sensitive polymer blends that display one component above a specific temperature with phase separation underneath it (MaxCST). Polymeric solutions that seem to be monophasic under a certain temperature but biphasic beyond that temperature are said to have a minimum crucial solution temperature (MinCST) [33, 34]. The MinCST seems to be the temperature where a polymer solution divides into two portions (anisotropic and isotropic states), abundant and deficient in the polymer. Such solution also is monophasic under a certain temperature but biphasic beyond that degree. The enthalpy parameter, which is connected to hydrogen bonding here between polymer and the water molecules, is accountable for polymer breakdown underneath the MinCST. When temperatures are raised just above MinCST, the entropy component (lipophilic contacts) takes precedence, resulting in polymer deposition. Among the most biocompatible polymers with MinCST characteristics includes poly (ethylene oxide). Nevertheless, based on the molecular mass, the MinCST transition of poly (ethylene oxide) aqueous solutions happens at ambient temperature, spanning between 100° C to 150° C. At minimum temperatures than just the poly (ethylene oxide) MinCST, a polymer with ethylene oxide components and hydrophilic sections (e.g. ethanol) would show phase changes. When a linear polymer with small sufficient Ethylene oxide sections is utilized to avoid micelle production, the precipitating from the aqueous phase can be thought about as a rapid MinCST changeover. Furthermore, in the lack of intermolecular and intramolecular hydrogen bonding, a continuous alternation of ethylene oxide-ethylene monomer copolymer pattern throughout the polymer would result in a MinCST defined either by lipophilic/hydrophilic equilibrium.

Poly(N-alkylacrylamide)s, Poloxamers, Poly(N-vinylcaprolactam)s, Chitosan, poly (ethylene oxide)- poly (propyleneoxide)- poly (ethylene oxide), Cellulose, xyloglucan, etc. are instances of thermally sensitive polymers (lactic acid) – tri blocks of poly (ethylene glycol). Poly (N-isopropyl acrylamide) and Poly (N-alkyl substituted acrylamides) with an annealing temperature of 32° C as well as poly (Nvinylalkylamides) like poly (N-vinyliso-butyramide) with just an annealing temperature of 39° C are perhaps the most extensively utilized thermally sensitive polymers [7, 35].

2.1.1 Thermally responsive smart polymers’ mechanisms of action

The occurrence of a minimum crucial solution temperature (MinCST) above which the polymer turns aqueous insoluble is generally the source of thermally-responsive smart polymeric solubility. This is characteristic of polymers that create hydrogen bonds with aqua, and it also has a wide spectrum of biological possibilities, including cell mapping, smart medication delivery, DNA sequencing, and so on. The chemical makeup of the monomers is varied throughout this strategy to regulate the polymer thermal sensitivity in aqua. To accomplish this, a variety of polymers centered on ethyleneoxide/ethylene monomer were developed and produced via multiple condensation processes of polyfunctional ethyleneoxide/ethylene monomer oligomers. The cloud point reflects the hydrophobicity/hydrophobicity balance continuously and may be customized in the spectrum of 7–70°C by adjusting the composition and polymer type.

The lack of organic solvents is an important benefit of such compositions. The shrinking in the volume that emits a considerable quantity of an encapsulating medication has been linked to the strong initial bursting impact of such approaches. The solubility behavior of polymer grafted onto the silicon surface is identical. The solubility cloud levels of grafting polymers are similar to those of bulk polymer solutions, according to binding energy studies.

Thermally responsive smart polymers’ dynamic solubility is generated by variations in the lipophilic/hydrophilic balance of the electron-deficient polymer, which are triggered by rising temperature or ionic intensity. Because of hydrogen bonds between aqueous molecules, electron-deficient polymers are soluble in aqua. The efficacy of hydrogen bonding decreases even as the temperature goes up. Whenever the effectiveness of hydrogen bonding is inadequate for macromolecule immersion, a polymer phase transition occurs. A phase transition occurs whenever the temperature of the water solution of innovative polymers is raised beyond a particularly critical point. There is a formation of an aqueous phase with almost minimal polymer and a polymer richer phase. The temperature at which a phase transformation occurs is determined by the amount of polymer present as well as the molecular mass of a polymer [35, 36].

2.2 Polymers with pH sensitivity

pH-Adaptive polymers are a class of stimuli-sensitive polymers that may alter their structural and physical properties in reaction to variations in solution pH, including surface properties, chain conformation, solubility, and arrangement. The phrase “pH-reactive polymers” refers to polymers containing ionizable basic or acidic groups where ionization is affected by the pH of the solution. In the latest days, the topic of pH-reactive polymers has grown in popularity, with scientific research being published year after year. As either a result, pH-sensitive polymer systems are extremely helpful in a broad array of applications, including gene delivery, drug administration, surfaces, membranes receptors, and chromatography [37, 38, 39].

Polymers that respond to pH might be linear, branching, or networked. According to their architectures, polymers may have varied sensitivities to solution circumstances and variable self-assembly tendencies. A pH shift, for instance, might result in the (de)protonation of functional moiety in the polymeric chain. It can produce flocculation, strand collapse-extension, including deposition in homopolymers in certain situations. It also might produce self-assembly in the forms of micelles, unimers, gels, vesicles, swelling, and deswelling, among other things. Surface active behaviors are demonstrated by pH alteration in block (co)polymers, branching (co)polymers, and starry (co)polymers with pH-sensitive block(s). Furthermore, pH changes cause (de)swelling in hydrogel as well as dendrimer-like formations. Surfaces altered with polymers allow for the creation of ionic interfaces with thin/thick layers as a result of pH changes. Figure 2 depicts the variations in polymers of various topologies caused by pH changes.

Figure 2.

Polymers that respond to pH in a variety of ways.

pH Adaptive polymers are polyelectrolytes with weakly basic or acidic moieties in their architecture that receive or liberate protons in reaction to variations in the pH of the surroundings. Polymers containing acidic or basic groups, such as carboxyl, pyridine, sulfonic, phosphate, and tertiary amines, are commonly referred to as pH adaptive polymers because of ionization of the molecules with pH variation causes a structural change. Their pH sensibility or ionization allows us to modify its self-assembly behavior, wettability phase segregation, polyelectrolyte character, and other properties, in complement to their biotechnological uses. It is feasible to make a polymer with a pKa ranging from 1 to 14. pH Reactive polymers having basic monomers behave like cationic polymers in acidic conditions, whereas polymers having acidic monomers behave like anionic polymers in basic conditions. Depending on the requirements, a few of these two types or a combination of the two with the appropriate composition is necessary. Natural polymers, as well as manmade polymers, have indeed been thoroughly investigated. Biopolymers are by far the most widely investigated because of their richness in ecology, rapid degradation, bio-compatibility, their potential to be modified. Polypeptides such as poly(histidine), poly(L-glutamic acid), and poly(aspartic acid) can be used to synthesize pH-reactive polymers. Such polymers are biodegradable and biocompatible, just like biopolymers. These biopolymers are quite significant among pH-sensitive polymers [38, 40, 41].

2.3 Polymers sensitive to two impulses (pH and temperature)

These really are polymeric frameworks that are thermal and pH-sensitive, but they are created by combining ionizable with lipophilic (inverse thermo-sensitive) moiety in a straightforward way [42]. Chitosan, acrylic acid, N,Ndimethylaminoethylmethacrylate, and other polymers that really are thermal and pH-responsive are instances. This is accomplished mostly by copolymerizing monomers with all these molecules, integrating thermally responsive polymers with polyelectrolytes, or developing novel monomers that adapt to both stimulation concurrently [43, 44].

2.4 Polymers that respond to glucose

Glucose-sensitive polymers can imitate typical internal insulin production, reducing diabetes problems and allowing for regulated delivery of the bioactive chemical. These really are sugar responsive and exhibit a wide range of responses to glucose. Although their applicability for both glucose monitoring and insulin administration, such polymers have gotten a lot of interest. Despite these benefits, the main drawbacks are the quick reaction time as well as the possibility of non-biocompatibility. The following techniques have been used to build glucose-sensitive polymeric-based formulations: enzymatic oxidation of glucose using glucose oxidase, glucose binding using lectin, or reversible covalent bond creation using phenylboronic acid molecules. Glucose responsiveness is caused by the polymer’s reaction to the by-products produced either by oxidation (enzymatic) of glucose. Glucose oxidase (GOx) is oxidized to form glucose to produce gluconic acid with hydrogen peroxide (H2O2). Within the instance of poly (acrylicacid) coupled with GOx mechanism, for instance, when blood glucose levels rise, conversion of glucose to gluconic acid, causing a drop in pH enabling hydrogenation of PAA carboxylate groups, allowing insulin to be released more quickly. Because its release profile closely resembles that of internal insulin, this approach is gaining popularity [45, 46].

Another technique makes use of lectin’s specific carbohydrate-binding characteristics to create a glucose-responsive system. Lectins are bifunctional proteins, and their glucose-biding function allows them to produce a variety of glucose-sensitive materials. The responses of these mechanisms were unique to glucose and mannose, with no reaction to certain other sugars. Concanavalin A is a 4 binding-site lectin that has been widely employed in insulin-containing medication delivery. The insulin component is chemically changed by inserting functional moieties (or glucose molecule) and afterward connected to a transporter or support via particular interactions that can only be disrupted by the glucose it in this sort of system. Concanavalin A competitive binding characteristic to glucose as well as glycosylated insulin is exploited in the glycosylated insulin-Concanavalin A combination. The bioactive unbound glucose moieties cause glycosylated Concanavalin A-insulin complex to be displaced inside the surrounding structures. The production of single-substituted glucosyl terminal PEG with insulin complex was also described in other investigations. The G-PEG–insulin complex was covalently coupled to Concanavalin A, which was connected to a PEG–poly(vinylpyrrolidone-co-acrylic acid) framework, and when the levels of sugar grew, the competitive attachment of glucose to Con A caused the G-PEG insulin complex to be displaced and released (Figure 3) [47].

Figure 3.

Polymers that respond to glucose in a variety of ways.

2.5 Smart polymers with photo responsiveness

Photo-sensitive polymeric materials are useful in that they could transport bioactive substances in reaction to light, including drug release happening nearly instantly and with excellent precision due to photo-induced restructuring in nano-carriers [48]. Three primary strategies were used to do this: This non-invasive form of drug administration reacts to the lighting of a certain wavelength and depends on either a single or multiple on-off drug release patterns [49]: (1) photo-generated change of hydrophobic nature to hydrophilic nature, (2) photo splitting reaction, and (3) photo-induced warming. Whilst also electromagnetic radiations with wavelengths in the range from 250 to 380 nm (ultraviolet region) and 700–900 nm (near-infrared region) are being used to stimulate photo-sensitive responses, light with wavelengths greater than 900 nm is inappropriate for delivery of drugs to certain parts of the human body, including the posterior section of the ocular system, because it cannot permeate the ocular soft tissue. Despite the fact that various polymers have been explored for ocular administration, several have been ruled out owing to chromophore intolerance and tissue destruction from photostimulation [50]. In order to establish an osmolality of a gel system, UV-responsiveness polymeric materials have been used in the eye to trigger an ionization process in the exposed to UV light, culminating in drug release through an inflow of solvent [51]. In another study, Viger et al. [52] used light thermally release of drugs to show the liberation of aqueous nano-platforms from watered poly(lactic-co-glycolic acid) (PLGA) micro-particulate system. Whenever moisture was subjected to NIR light with a wavelength of 980 nm, the photo-energy was quickly converted into thermal energy. The PLGA changed to a rubbery condition as a result of the warming, allowing the Nile red or Nile blue to be released from the micro-particulate system more easily. When compared with untreated particulates, the substantial release was achieved, which was also shown in vitro [52].

At the minimum one aqueous soluble area, at minimum one biodegradable part, as well as at least minimum of two free radical polymerizable portions are included in the macromers. Free radical activators polymerize macromers in presence of UV irradiation, visible light stimulation, or heat energy. Poly (vinyl alcohol), PEG, polysaccharides like hyaluronan, or peptides like albumin can make up the core aqueous soluble area. Polymers consisting of polyglycolic acid, polylactic acid, poly(anhydrides), polylactones, and poly(amino acids), may be used in the biodegradable zones. Acrylates, methacrylates, diacrylates, and other physiologically acceptable polymerizable units are favored polymerizable areas. Ethyl eosin, camphorquinone, and acetophenone analogs, are examples of promoters that can be employed to generate free radicals [53].

2.6 Enzyme sensitive polymeric material

Some fundamental guidelines should be followed while synthesizing enzyme-sensitive polymers with biomedical utilization. Enzymes must work in certain settings (e.g., an aquatic milieu having multiple ions with a pH of 7.4 or mildly basic or acid), while enzyme-sensitive polymers must withstand these circumstances. Apart from the availability of a substrate/substrate-mimic molecule for such focused enzyme to respond, the focused enzymes’ operations must cause a variation in the polymers’ characteristics for the particular activities to occur. The activity of the enzyme and the reaction of the final substance can be performed concurrently or in a step-by-step manner. For instance, proteins were used as a crosslinking agent in the DNA nanoparticles, and proteases quickly degrade the protein, destroying the nanoparticles [54]. In some other cases, enzymatic dissociation of a protective moiety causes peptides generated from amyloid to fold, reorganize, and self-assemble forming fibrillar clumps [55].

In live organisms, enzymes govern the bond generation and breakage, substrate oxidation/reduction, as well as isomerization processes, with the first two chemical reactions being exploited in the development of enzyme-sensitive materials. The bond breakage process has been utilized to cleave protein as well as ester bonds with polymers and/or tiny moiety, which really is important in controlled medication delivery with implant biodegradation. The kinase/phosphatase combination, which catalyzes the dephosphorylation/phosphorylation events here on substrates, might be employed to build reversibly sensitive materials through enzymatic bond creation and breaking.

Chitosan, alginate, dextran, polyethylene glycol, polyacrylamide, and polyethylene oxide have all been investigated as polymer matrices for the creation of enzyme-sensitive systems (butyl methacrylate) (Figure 4) [56, 57, 58].

Figure 4.

Polymers that respond to enzymes in a variety of ways.

2.7 Oxidation and reduction sensitive polymeric material

Redox polymeric materials can be separated into reduction reactive systems and oxidation reactive processes depending on the nature of a reductive stimulus. Disulfide and diselenide connections are typically seen in the reduction reactive system, which will be disrupted by considerable growth in the cost of nearby reducing molecules such as GSH. Direct production of disulfide coupling and bridging with a disulphide-containing crosslinking agent are basically two strategies for incorporating disulfide coupling in the process. By live or regulated polymerization, disulfide could be incorporated into the polymer as the oligomer (e.g. Reversible addition-cleavage crosslinking polymerization and atomic transfer radical polymerization) [59]. The thiol-disulfide interchange process, which is commonly utilized to create reduction reactive prodrugs including genetic transporters, is another viable method with gentler circumstances (e.g. at room temperature) than the controlled/living polymerization [60]. To stop the drug leaking, polymeric micelles comprising substances can indeed be crosslinked with covalently crosslinking agents (using bis(2,2′-hydroxyethyl)disulfide, dithiodipropionic acid, and their derived products) and afterward the disulfide conduits split to discharge the substances after the micelles meet the goal [61]. Since the bond-breaking energies of the C-Se (244 kJ mol1) and Se-Se (172 kJ mol1) bond formation are lesser than those from the C-S (272 kJ mol1) and S-S (251 kJ mol1) bond formation, replacing the disulfide interconnection with the diselenide connection is a simple approach to strengthen the responsiveness of the redox-sensitive system. However, diselenide link insertion into a polymeric matrix is more difficult than disulfide link formation, and more research into effective synthetic techniques is needed (Table 1) [62].

ResponsivenessMeritsConstraints
ThermalIntroduction of active moieties is simple.
Manufacturing and composition are easy.
Issues with injectability during application criteria.
Weak mechanical sturdiness, biocompatibility problems, and thermolablile medicine instabilities.
pHDrugs that are thermolabile will benefit from this.There is a scarcity of data about toxicity.
Mechanical strength is low.
LightManaging the trigger procedure is simple.
Controlling the stimuli with precision.
Gel has a poor mechanical strength, which increases the risk of noncovalently bound chromophores seeping off.
Electric fieldVariations in electrical charge cause pulsative releasing.Unpredictable behaviors to light.
Implantation via surgery is necessary.
External stimulus delivery necessitates the use of extra equipment.
Perfecting the size of electric charge is challenging.
UltrasoundProtein release that can be controlled.Regulating the release using specialized equipment.
Non-biodegradable delivery systems necessitate surgical implantation.
Mechanical abrasionPossibility of obtaining medication releaseManaging the release pattern is difficult.

Table 1.

Several smart polymeric drug delivery technologies are available [7].

The reactive oxygen system, the result of aerobic metabolism, is what activates oxidation reactive systems. Oxidation reactive materials include sulfur-based compounds. To accomplish the lipophilic-hydrophilic shift, reactive oxygen moiety can oxidize poly(propylene sulfide) (PPS) to generate sulphoxide [63]. The comparatively higher stability of sulfur in sulfur-containing substances is a fundamental restriction, and the reaction to reactive oxygen may not be as sensitive. The addition of selenium to the polymeric materials, which will be more sensitive than sulfur, improves the susceptibility of the reaction to reactive oxygen [64]. Owing to its own oxidation responsiveness, ferrocene-containing polymeric materials are another prominent family of oxidation reactive polymers [65, 66]. Ferrocene can be inserted in the framework, side chain, and terminal unit of the polymeric materials. Developing themes such as boronic ester moieties, oligoproline, and tetrathiafulvalene have been studied for the development of new oxidation-sensitive polymeric substances to broaden the uses [67, 68].

Advertisement

3. Applications of specialized polymeric system

3.1 Drug delivery

The majority of bio-sensitive systems, notably those used in cancer therapy, rely on regulated medication release. While significant advancements in chemotherapy have resulted in the development of a number of novel medications for treating cancer that has significantly improved patients’ prognoses and standard of living, a key obstacle remains the treatments’ lack of compassion for neoplastic cells [69]. The treatment impact of the anticancer treatment is harmed by the possibility of a deadly systemic adverse effect and the development of resistant strains [70]. Continued improvement of chemotherapy necessitates adequate drug release at the tumor site as well as the avoidance of drug-carrier endosomal sequestration, and the development of suitable stimulus-sensitive systems has shown tremendous promise in both areas. This is attributable to the fact that tumor tissues’ milieu can produce a variety of natural signals. For instance, tumor cells contain moderate acidity, significant GSH (glutathione) levels, as well as a top-level of hyaluronidase [71], therefore pH-, redox-, and enzyme sensitive drug carriers, as well as their combinations (to optimize the release of drug efficiency), have been extensively studied. Blood serum albumin (HSA)-coatedMnO2 nanomaterial’s as an adaptive transporter of cis-platinum is a fresh example. The MnO2 combines with internal H2O2 just at tumor site to produce O2 in vivo, overcoming medicaments resistance caused by local hypoxic, while the nanoparticles disintegrate in an acidic medium, releasing cis-platinum [72]. In another layout, the water-soluble rhodamine B was covalently conjugated to the PDMAEMA (Poly((2-dimethylamino)ethyl methacrylate) and via disulphide bond formation with the lipophilic coumarin 102 physiologically encapsulated inside the nanogel, and the hydrophilic rhodamine B has been covalently linked to the PDMAEMA via disulfide bond formation with the lipophilic coumarin 102. The nano gel is swollen in an acidic medium and shrinks at increased temperature to liberate the coumarin 102, whereas decreasing DL-dithiothreitol cleaves the disulfide bridges to liberate the aqueous cargo medication [73]. The development of bio-sensitive drug carriers for controlled release has exploded in the past few decades, and additional improvements in release effectiveness have resulted in dual and numerous systems that can carry several medicines for programmable site-specific delivery of drugs. Pharmaceutical loading, persistence in a microenvironment, tumor-targetability, effective absorption of cancerous cells, and controlled intracellular release of the drug are among the fundamental difficulties in the delivery of drugs addressed by the many configurations of the bio-sensitive delivery mechanism. Even though there are a lot of good studies, most of it makes a specialty of the difficulties and still in the concept-proofing phase [74]. The challenges are associated with most existing bio-sensitive drug delivery mechanisms, such as poor drug loading efficiency, biodegradability, as well as the ability to remain circulatory and concentrate in the target organs, must be overcome in order to convert the study into practical practice (e.g. tumor). In contrast, more research into the subatomic scale in vivo behavior of bio-sensitive systems, as well as the influence of systemic physiological parameters on the release of the drug, is needed [74].

3.2 Biomaterials actuators and micro-fabrication

Designing microfluidic technologies for biochemical applications has proven to be a difficult task, and a properly working valve is a critical component in these technologies. Traditional micro-actuators are somewhat sophisticated components that needed additional electricity to operate. The use of sensitive smart polymer composites to govern flow eradicates this need for external power, output control, and complicated fabrication ploys, allowing them to be integrated within microfluidics streams and dwindle or perk up in response to an external stimulus, causing streams to open or close. Photo triggered polymerization inside the stream of a microfluidic chip that may be employed as a gate for changing; transmission, measuring, and closing of a PCR reaction vessel produced monolithic plugging PNiPAAm complexes using 5% methylenebisacrylamide. Because of their simple construction of sensors, the kinetic studies of the volume phase change process as a feature of gel structure and shape, the capacity of the sensors to thwart and supplant the transition between two fluids, anisotropic bulging of a polymer, as well as the ability to adapt to changing stimuli, responsive smart polymeric materials are the structural elements for microfluidic devices. Thermally sensitive smart polymeric materials have also been utilized to create “smart” affinities beads which can be transiently mounted on microfluidic walls of the channel just above MinCST in order to acquire the target biomaterials via its friendliness component. Proteomic functionalities, such as pre-concentration and isolation of soluble proteins on an embedded fluidics device, have been enabled by this technology. Many efforts were made to emulate live creatures’ effective transition of chemical energy to mechanical energy. The bio-inspired actuators might be employed in future ‘soft’ technologies that are based on biological concepts rather than mechanical ones. Because bio-inspired actuators can tolerate extremely hostile conditions, they can also be utilized to pick up extremely small items in watery solutions. By contorting a barrier that subsequently occludes an opening, a system built on pH-sensitive smart polymeric discs of polymethacrylic acid-triethylene glycol dimethacrylate (PMAA-EG) has indeed been utilized to control medication delivery. The electronegative interpenetrating matrix (IPM) made of PVA with PNiPAAm was studied in aquatic NaCl solution for its moisture content and carrying behavior with electromagnetic current, with the goal of using it in bio-inspired sensors and devices that respond quickly to exterior electric fields. The immobilized smart polymer’s prompted manipulation of interfacial characteristics at the solid-liquid interface has benefits in the development of microfluidics bio-analytical systems since they supply the actuation pressure necessary for both valving and dispensing functionalities in micro-dispensing gadgets [75, 76, 77].

3.3 Diagnostic uses

Biomedicine research involves advancing our understanding of biology and the processes behind physiological activity and disorders. As a result, in addition to illness therapy, one of the most significant goals is diagnostics, wherein bio-sensitive materials have shown promising potential in detecting low concentrations of biochemical, proteins, and genes that act as sickness-specific indicators. Those indicators are typically tested using high-cost chromatography techniques like high-performance liquid chromatography and gas chromatography-mass spectrometry, but using stimuli-sensitive systems, easy, rapid, precise, and low-cost detection procedures may be established.

For instance, metallic nanoparticles with a size of 4 nm may greatly boost T1 distinction in magnetic resonance imaging; however, their aggregation led in T2 contrasting augmentation owing to in uniform magnetic field around the aggregates. As a result, IONs like these have been employed as a T2 contrast media to diagnose liver disorders. They are, nevertheless, unsuitable for the identification of smaller hepatocellular carcinomas that requires a good detection to improve the individuals’ average five-year rate of survival [78]. The fall in pH dispersed the aggregation of the functional metallic nanoparticles when they were treated using i-Motif DNAs that really can convert from unistranded to fused quadruple-helical structure in an acidic medium. Because acidification of the tumor encouraged the breakdown of the metallic nanoparticles aggregates and shifted the MRI signal between T2 to T1 augmentation to better the differentiation between hepatocyte and tiny hepatocytic carcinoma tissues, tiny hepatocytic carcinoma may be diagnosed with these bifunctional metallic nanoparticles [79]. pH-sensitive surfaces made comprised of nanoparticles with just an amino group having a silane layer are another intriguing instance. In an acidic medium, the amino groups are protonated, making the surfaces highly hydrophilic, whereas in a highly alkaline, the surfaces become really hydrophobic. The amount of glucose in the mouth and pee may be reliably determined in one second using this surface via measuring the contact area of the liquid specimen, which is dependent on the created gluconic acid following adding glucose oxidase to the specimen [80]. This non-invasive, economic approach of fast glucose measurement is useful for overcoming the drawbacks of standard intrusive diagnosis of diabetes, including such discomfort and infection hazard. While contemporary research has demonstrated the stimuli-sensitive system’s potential and performance in preclinical testing for diagnostic uses, the majority of the built systems do not fulfill the standards for clinical usage. This is owing to the large variety of chemicals found in real specimens collected from individuals with varying situations (e.g., various diets, ethnicities, and lifestyles), which considerably affects the measurement’s specificity and stability [81]. Aside from identifying biochemical levels, constant monitoring and distribution centres in human, both of which are challenging to perform, may be required. As a result, motivated monitoring technologies are still in the early stages of development, and more investigation is necessary before they can be used in clinical illness treatment.

3.4 Implants for cardiology

Creating actuators including such valves and levers out of the material, which could be utilized as blood artery implants, is one potential application. To modulate blood flow, the artery might be enlarged or constricted, also utilizing internal biochemical impulses. The valves would’ve been placed into the blood channels of the heart, or prosthetic muscular implants may be created [82].

3.5 Mucoadhesive delivery using polymers

Hydrophilic polymers must be employed to construct the liquid ophthalmic delivery mechanism since they may serve as a useful viscosity altering or boosting agent. In the ophthalmic mucoadhesive delivery method, polysaccharides are often employed. Hyaluronic acid, methylcellulose, hydroxypropyl methylcellulose, chitosan, gellan gum, carrageenan, xanthan gum, and guar gum are some of its variants. Chitosan is a polysaccharide polymer made up of polysaccharides. It is appropriate for usage in medication compositions due to its biodegradability, low toxicity, and biocompatibility [83]. Polyvinylpyrrolidone, poloxamer, and polyvinyl alcohol are among additional non-ionic polymers utilized for mucoadhesive characteristics [84].

3.6 Cancer treatment using a polymer drug combination

The medications as well as the polymer have a physiologically labile connection. Paclitaxel [poly(L-glutamic acid)] is a chemotherapeutic medication employed to treat cancers of the ovary, breast, as well as lung. Phase III studies have been conducted on it. Among its 2’hydroxyl unit and the carboxylic acid of poly(L-glutamic acid), it possesses an ester bond [83]. To improve its efficacy as an antitumor targeted drug delivery, Poly (amidoamine) and PEG is covalently attached with the chemotherapy medication Paclitaxel. Both improve the solubility of the substance. In an in vitro investigation of mankind ovarian cancer cells, it was discovered that PEG-based conjugates lowered paclitaxel activity by 25-fold, but the Poly (amidoamine)]-G4 dendrimer increased its efficacy by more than ten times [12]. The medication 5-fluorouracil induces cell death. Some researchers created PLA nanospheres as an encapsulating reagent for 5-fluorouracil [12].

3.7 Medicine and biotechnology

Smart polymeric materials may be chemically attached to bio-substances or physically combined with them to create a vast variety of polymeric materials and bio-molecular systems that really can adapt to physiological and chemical stimuli. Oligosaccharides, Polypeptides, glucose and polysaccharides, solitary as well as double-sided oligonucleotides, DNA plasmid, basic lipids and ligands, phospholipids, as well as synthesized medicine compounds are examples of bio-substances that can be polymer linked. Smart polymeric materials and sensitive surfaces that cope with environmental stimuli are made with these materials. Smart polymers with size-specific switches for turning proteins on and off were also studied. When a sensible polymer chain is connected to a protein complex that is further away from active site, the expanded polymer chain shields the active-locations, preventing bigger molecules from attaching. These polymers operate as a molecular gatekeeper, limiting the types of molecules that really can attach to proteins depending on their size [85].

3.8 Glucose level monitoring

The manufacture of insulin administration devices for the management of diabetic individuals is a prime utilization of smart polymeric materials. Several technologies have been used to give precise amounts of insulin at precisely the right moment, and all of them include a glucose sensor, sometimes known as a “biosensor,” incorporated into the mechanism. The word ‘biosensor’ refers to sensing devices that are used to detect the number of chemicals and other biologically relevant analytes. mGlucose oxidase (GluOx) is primarily employed in glucose monitoring and enables the use of various pH-sensitive smart polymeric materials for regulated insulin administration [86, 87].

3.9 Surfaces that react to stimuli

Tissue culture techniques have leveraged the shift in surface characteristics of thermally sensitive smart polymeric materials from hydrophilic well above threshold temperature to hydrophobic underneath it. Human cells are grown on hydrophobic solid culture plates and are normally separated from them using a protease therapy that causes the cells to be damaged. Because of the close connectivity among cells and cells, this allows for a high level of effectiveness whenever transplanted into individuals. The intensity of each molecule’s reaction to variations in stimuli is a combination of single monomer unit modifications that are weak on their own, and these modest reactions combine to generate a force that drives biochemical mechanisms. Likewise, the chromatographic matrix has been modified using surfaces with thermally responsive hydrophobic/hydrophilic qualities. For protein-rich selectivity with minimal non-specific couplings, thermally responsive size-exclusion chromatography is utilized. Smart polymeric mats are distinguished by their non-linear behavior [88, 89]. A minor stimulus can cause a substantial change in structure and characteristics (Table 2). Once that shift happens, the polymer exhibits a predictable all-or-nothing reaction with full homogeneity throughout [96].

StimuliDrugPolymerUsesGoal/outcome of the research
Thermal responsiveExenatidePLGA–PEG–PLGADiabetic type 2 treatmentTo create an injectable composition with a long-acting effect [90].
LeuprolidePolybenzofulveneFor treatment of tumorsExternal warmth is used to preserve the oligopeptide medication and modulate the release rate [91].
pH responsiveKetoprofenPoly(acrylamide)-g-carrageenan and sodium alginateTargeted distribution to the colonWhenever the ph of the sample was changed from acidic to basic, ketoprofen release rose considerably [92].
Dauxorubicin and paclitaxelPoly(ethylene glycol)-block-poly (propylene glycol)-poly(ethylene glycol)Survival time is extended as compared to single-drug treatment.The rate of release can be enhanced by lowering the pH of the external surroundings from acidic to basic [93].
Glucose responsiveInsulin-Con A complexMethacrylate derivatives
of dextran and concanavellin
Insulin delivery that is self-controlledThe findings showed that insulin release was bidirectional in reaction to varying glucose level, and also that the insulin produced was effective [94].
SulphonamideN,N(dimethylacrylamide) and sulfadimethoxine monomerGlucose-responsive hydrogel made of sulphonamideIn such a buffered salt solution at pH 7.4, the gelatin displayed bidirectional expansion as a result of glucose content from 0 and 300 mg/dL [95].
Enzyme responsiveAmyloidChitosanThe amyloid-derived proteins are rearranged as a result of this.In live organisms, enzymes govern binding and breakage, substrates oxidation / reduction, including isomerization processes, with the first two chemical reactions being exploited in the development of enzyme responsive substances [57].
Photo responsiveCross linked hyalouronic acid hydrogelTrisodium salt of copper chlorophyllinThe enzymatic reaction is what drives the prospective application of visible light-responsive hydrogels for temporal delivery of drugs.Photosensitive compounds including such chromophores are used to make visible light-sensitive hydrogels [94].

Table 2.

Several uses of advanced drug delivery systems.

Advertisement

4. Conclusion and future trends

The topic of stimuli-sensitive polymeric materials is quickly evolving, with evidence of its pharmacokinetic and therapeutic usefulness in the delivery of drugs. Stimuli-sensitive polymeric materials, which constitute an adaptive delivery strategy, have been proven to effectively respond in the required manner to the associated stimuli and are interesting prospects as a specific drug delivery option. Despite the fact that most of these polymeric’ multifunctionality qualities enable for a variety of health uses, stimuli sensitive polymeric substances have limited therapeutic potential due to cytocompatibility and toxicity concerns. The enormous progress and different benefits and prospects of stimuli-sensitive polymeric materials have been highlighted in this chapter. Despite the various options that are already available, additional research into healthier and much more biocompatible delivery mechanisms has still been needed.

Advertisement

Consent for publication

Not applicable.

Advertisement

Availability of data and materials

All the information in the manuscript has been referred from the included references and is available upon request from the corresponding author.

Advertisement

Competing interests

The authors declare no competing interest.

Advertisement

Authors’ contributions

I declare that this work was done by the author named in this article. PS conceived, designed the study, carried out the literature collection of the data, writing, and corrected the manuscript. PS, MT, and VJ read and approved the final manuscript.

Advertisement

Abbreviations

MaxCSTMaximum crucial solution temperature
MinCSTMinimum crucial solution temperature
GOxGlucose oxidase
PEGPolyethylene Glycol
PPSpoly(propylene sulphide)
GSHGlutathione
PDMAEMA(Poly((2-dimethylamino)ethyl methacrylate)
PMAA-EGPolymethacrylic acid-triethylene glycol dimethacrylate
IPMInterpenetrating matrix

References

  1. 1. Anderson JM, Kim SW. Advances in drug delivery systems (3), book review. Journal of Pharmaceutical Sciences. 1989;78:608-609
  2. 2. Langer R, Peppas NA. Advances in biomaterials, drug delivery, and bio nanotechnology. AICHE Journal. 2003;49:2990-3006
  3. 3. Heller A. Integrated medical feedback systems for drug delivery. AICHE Journal. 2005;51:1054-1066
  4. 4. Martinho N, Damge C, Reis PC. Recent advances in drug delivery systems. Journal of Biomaterials and Nanobiotechnology. 2011;2:510-526
  5. 5. Din F, Aman W, Ullah I, Quereshi OS, Mustapha O, Shafique S, et al. Effective use of nano-carriers as drug delivery systems for the treatment of selected tumors. International Journal of Nanomedicine. 2017;12:7291-72309
  6. 6. Tiwari G, Tiwari R, Sriwastawa B, Bhati L, Pandey S, Pandey P, et al. Drug delivery systems: An updated review. International Journal of Pharmaceutical Investigation. 2012;2:2-11
  7. 7. James HP, John R, Alex A, Anoop K. Smart polymers for the controlled delivery of drugs–a concise overview. Acta Pharmaceutica Sinica B. 2014;4(2):120-127
  8. 8. Teleanu DM, Chircov C, Grumezescu AM, Volceanov A, Teleanu RI. Blood brain delivery methods using nanotechnology. Pharmaceutics. 2018;10:298-305
  9. 9. Cacciatore I, Ciulla M, Fornasari E, Marinelli L, Di Stefano A. Solid lipid nanoparticles as a drug delivery system for the treatment of neurodegenerative diseases. Expert Opinion on Drug Delivery. 2016;13(8):1121-1131
  10. 10. Patel M, Souto EB, Singh KK. Advances in brain drug targeting and delivery: Limitations and challenges of solid lipid nanoparticles. Expert Opinion on Drug Delivery. 2013;10:889-905
  11. 11. Schmaljohann D. Thermo and pH responsive polymers in drug delivery. Advanced Drug Delivery Reviews. 2006;58:1655-1670
  12. 12. Liechty WB et al. Polymers for drug delivery systems. Annual Review of Chemical and Biomolecular Engineering. 2010;1:149-173
  13. 13. Langer RS, Peppas NA. Present and future applications of biomaterials in controlled drug delivery systems. Biomaterials. 1981;2:201-214
  14. 14. Verma RK, Mishra B, Garg S. Osmotically controlled oral drug delivery. Drug Development and Industrial Pharmacy. 2000;26:695-708
  15. 15. Hoffman AS. The origins and evolution of “controlled” drug delivery systems. Journal of Controlled Release. 2008;132(3):153-163
  16. 16. Howard MD, Jay M, Dziubla TD, Lu X. PEGylation of nanocarrier drug delivery systems: State of the art. Journal of Biomedical Nanotechnology. 2008;4(2):133-148
  17. 17. Knop K, Hoogenboom R, Fischer D, Schubert US. Poly (ethylene glycol) in drug delivery: Pros and cons as well as potential alternatives. Angewandte Chemie – International Edition is Angew. 2010;49:6288
  18. 18. Bae Y, Kataoka. Intelligent polymeric micelles from functional poly(ethylene glycol)-poly(amino acid) block copolymers. Advanced Drug Delivery Reviews. 2009;61:768
  19. 19. Marcucci F, Lefoulon F. Active targeting with particulate drug carriers in tumor therapy: Fundamentals and recent progress. Drug Discovery Today. 2004;9:219
  20. 20. Torchilin VP. Passive and active drug targeting: Drug delivery to tumors as an example. Handbook of Experimental Pharmacology. 2010;197:3
  21. 21. Bhadra D, Bhadra S, Jain P, Jain NK. Pegnology: A review of PEGylated systems. Die Pharmazie. 2002;57:65
  22. 22. Lammers T, Kiessling F, Hennink WE, Storm G. Nanotheranostics and image-guided drug delivery: Current concepts. Molecular Pharmaceutics. 2010;7:1899
  23. 23. Ahmed F, Pakunlu RI, Srinivas G, Brannan A, Bates F, Klein ML, et al. Shrinkage of a rapidly growing tumor by drug-loaded polymersomes: pH-triggered release through copolymer degradation. Molecular Pharmaceutics. 2006;3(3):340-350
  24. 24. Ahmed F, Pakunlu RI, Brannan A, Bates F, Minko T, Discher DE. Biodegradable polymersomes loaded with both paclitaxel and doxorubicin permeate and shrink tumors, inducing apoptosis in proportion to accumulated drug. Journal of Controlled Release. 2006;116(2):150-158
  25. 25. Ward MA, Georgiou TK. Thermoresponsive polymers for biomedical applications. Polymers. 2011;3:1215-1242
  26. 26. Kikuchi A, Okano T. Intelligent thermoresponsive polymeric stationary phases for aqueous chromatography of biological compounds. Progress in Polymer Science. 2002;27:1165-1193
  27. 27. Hoffman AS, Stayton PS, Bulmus V, Chen G, Chen J, Cheung C, et al. Really smart bioconjugates of smart polymers and receptor proteins. Journal of Biomedical Materials Research. 2000;52(4):577-586
  28. 28. Al-Tahami K, Singh J. Smart polymer based delivery systems for peptides and proteins. Recent Patents on Drug Delivery & Formulation. 2007;1:65-71
  29. 29. Kumar A, Srivasthava A, Galevey IY, Mattiasson B. Smart polymers: Physical forms and bioengineering applications. Progress in Polymer Science. 2007;32:1205-1237
  30. 30. Bawa P, Viness B, Yahya EC, Lisa C. Stimuli-responsive polymers and their applications in drug delivery. Biomedical Materials. 2009;4:022001
  31. 31. Diez-Pena E, Quijada-Garrido I, Barrales-Rienda JM. On the water swelling behaviour of poly(N-isopropylacrylamide) [P(N-iPAAm)], poly(methacrylic acid) [P(MAA)], their random copolymers and sequential interpenetrating polymer networks (IPNs). Polymer. 2002;43:4341-4348
  32. 32. Varga I, Gilanyi T, Meszaros R, Filipcsei G, Zrinyi M. Effect of crosslink density on the internal structure of poly(N isopropylacrylamide) microgels. The Journal of Physical Chemistry. B. 2001;105:9071-9076
  33. 33. Yan L, Zhu Q, Kenkare PU. Lower critical solution temperature of linear PNIPA obtained from a yukana potential chains. Journal of Applied Polymer Science. 2007;78:1971-1976
  34. 34. Jones MS. Effect of pH on the LCST of random copolymers of N-isopropylacrylamide and acrylic acid. European Polymer Journal. 1999;35:795-801
  35. 35. Cao YL, Ibarra C, Vacanti C. Preparation and use of thermo responsive polymers. In: Morgan JR, Yarmush ML, editors. Tissue Engineering: Methods and Protocols. Totowa: Humana Press; 2006. pp. 75-84
  36. 36. Spohr R, Reber N. Thermal control of drug release by a responsive ion track membrane observed by radio tracer flow dialysis. Journal of Controlled Release. 1988;50:1-11
  37. 37. Hu J, Zhang G, Ge Z, Liu S. Stimuli-responsive tertiary amine methacrylate-based block copolymers: Synthesis, supramolecular self-assembly and functional applications. Progress in Polymer Science. 2014;39(6):1096-1143
  38. 38. Hoffman AS. Hydrogels for biomedical applications. Progress in Polymer Science. 2012;64:18-23
  39. 39. Dai S, Ravi P, Tam KC. pH-responsive polymers: Synthesis, properties and applications. Soft Matter. 2008;4(3):435-449
  40. 40. Alvarez-Lorenzo C, Blanco-Fernandez B, Puga AM, Concheiro A. Crosslinked ionic polysaccharides for stimuli-sensitive drug delivery. Advanced Drug Delivery Reviews. 2013;65:1148-1171
  41. 41. Jiang TY, Wang ZY, Tang LX, Mo FK, Chen C. Polymer micellar aggregates of novel amphiphilic biodegradable graft copolymer composed of poly (aspartic acid) derivatives: Preparation, characterization, and effect of pH on aggregation. Journal of Applied Polymer Science. 2006;99(5):2702-2709
  42. 42. Gil ES, Hudson SM. Stimuli-responsive polymers and their bioconjugates. Progress in Polymer Science. 2004;29:1173-1222
  43. 43. Lou L, Kato M, Tsuruta T, Kataoka K, Nagasaki Y. Stimuli-sensitive polymer gels that stiffen upon swelling. Macromolecules. 2000;33:4992-4994
  44. 44. Gonalez N, Elvira C, San RJ. Novel dualstimuli- responsive polymers derived from ethylpyrrolidine. Macromolecules. 2005;38:9298-9303
  45. 45. Ravaine V, Ancla C, Catargi B. Chemically controlled closed-loop insulin delivery. Journal of Controlled Release. 2008;132:2-11
  46. 46. Roy VD, Cambre JN, Sumerlin BS. Future perspectives and recent advances in stimuli-responsive materials. Progress in Polymer Science. 2010;35:278-301
  47. 47. Takemoto Y, Ajiro H, Asoh TA, Akashi M. Fabrication of surface modified. Hydrogels with polyion complex for controlled release. Chemistry of Materials. 2010;22:2923-2929
  48. 48. Cabane E, Zhang X, Langowska K, Palivan CG, Meier W. Stimuli-responsive polymers and their applications in nanomedicine. Biointerphases. 2012;7(1):9
  49. 49. Mura S, Nicolas J, Couvreur P. Stimuli-responsive nanocarriers for drug delivery. Nature Materials. 2013;12(11):991-1003
  50. 50. Yasin MN, Svirskis D, Seyfoddin A, Rupenthal ID. Implants for drug delivery to the posterior segment of the eye: A focus on stimuli-responsive and tunable release systems. Journal of Controlled Release. 2014;196:208-221
  51. 51. Mamada A, Tanaka T, Kungwatchakun D, Irie M. Photoinduced phase transition of gels. Macromolecules. 1990;23(5):1517-1519
  52. 52. Viger ML, Sheng W, Doré K, Alhasan AH, Carling CJ, Lux J, et al. Near-infrared-induced heating of confined water in polymeric particles for efficient payload release. ACS Nano. 2014;8(5):4815-4826
  53. 53. Gan LH, Gan YY, Roshan DG. Poly (N-acrylol-N propylpiperazine): A new stimuli responsive polymer. Macromolecules. 2000;33:7893-7897
  54. 54. Santiana JJ, Sui B, GomezNand Rouge JL. Programmable peptide-cross-linked nucleic acid nanocapsules as a modular platform for enzyme specific cargo release. Bioconjugate Chemistry. 2017;28:2910-2914
  55. 55. Dos Santos S, Chandravarkar A, Mandal B, Mimna R, Murat K, Saucède L, et al. Switch-peptides: Controlling self-assembly of amyloid β-derived peptides in vitro by consecutive triggering of acyl migrations. JACS. 2005;127:11888-11889
  56. 56. Hu Q, Katti PS, Gu Z. Enzyme-responsive nanomaterials for controlled drug delivery. Nanoscale. 2014;6:12273-12286
  57. 57. Hu J, Gand Z, Liu S. Enzyme-responsive polymeric assemblies, nanoparticles and hydrogels. Chemical Society Reviews. 2012;41:5933-5949
  58. 58. Zelzer M, Todd SJ, Hirst AR, McDonald TO, Ulijn RV. Enzyme responsive materials: Design strategies and future developments. Biomaterials Science. 2013;1:11-39
  59. 59. Siegwart DJ, Oh JK, Matyjaszewski K. ATRP in the design of functional materials for biomedical applications. Progress in Polymer Science. 2012;37:18-37
  60. 60. Huo M, Yuan J, Tao L, Wei Y. Redox-responsive polymers for drug delivery: From molecular design to applications. Polymer Chemistry. 2014;5:1519-1528
  61. 61. Zhang S, Zhao Y. Rapid release of entrapped contents from multi-functionalizable, surface cross-linked micelles upon different stimulation. JACS. 2010;132:10642-10644
  62. 62. Xu H, Cao W, Zhang X. Selenium-containing polymers: Promising biomaterials for controlled release and enzyme mimics. Accounts of Chemical Research. 2013;46:1647-1658
  63. 63. Napoli A, Valentini M, Tirelli N, Muller M, Hubbell JA. Oxidation-responsive polymeric vesicles. Nature Materials. 2004;3:183-189
  64. 64. Liu J, Pang Y, Zhu Z, Wang D, Li C, Huang W, et al. Therapeutic nanocarriers with hydrogen peroxide-triggered drug release for cancer treatment. Biomacromolecules. 2013, 2013;14:1627-1636
  65. 65. Zhang X, Han L, Liu M, Wang K, Tao L, Wan Q, et al. Recent progress and advances in redox-responsive polymers as controlled delivery nanoplatforms. Materials Chemistry Frontiers. 2017;1:807-822
  66. 66. Staff RH, Gallei M, Mazurowski M, Rehahn M, Berger R, Landfester K, et al. Patchy nanocapsules of poly (vinylferrocene)-based block copolymers for redox-responsive release. ACS Nano. 2012;6:9042-9049
  67. 67. Broaders KE, Grandhe S, Frechet JM. A biocompatible oxidation-triggered carrier polymer with potential in therapeutics. JACS. 2011;133:756-758
  68. 68. de Gracia LC, Joshi-Barr S, Nguyen T, Mahmoud E, Schopf E, Fomina N, et al. Biocompatible polymeric nanoparticles degrade and release cargo in response to biologically relevant levels of hydrogen peroxide. JACS. 2012;134:15758-15764
  69. 69. Cao Y, DePinho RA, Ernst M, Vousden K. Cancer research: Past, present and future. Nature Reviews. Cancer. 2011;11:749-754
  70. 70. Patel NR, Pattni BS, Abouzeid AH, Torchilin VP. Nanopreparations to overcome multidrug resistance in cancer. Advanced Drug Delivery Reviews. 2013;65:1748-1762
  71. 71. Lu Y, Aimetti AA, Langer R, Gu Z. Bioresponsive materials. Nature Reviews Materials. 2016;2:16075
  72. 72. Chen Q, Feng L, Liu J, Zhu W, Dong Z, Wu Y, et al. Intelligent albumin–MnO2 nanoparticles as pH-/H2O2-responsive dissociable nanocarriers to modulate tumor hypoxia for effective combination therapy. Advanced Materials. 2016;28:7129-7136
  73. 73. Cao Z, Zhou X, Wang G. Selective release of hydrophobic and hydrophilic cargos from multi-stimuli-responsive nanogels. ACS Applied Materials & Interfaces. 2016;8:28888-28896
  74. 74. Cheng R, Meng F, Deng C, Klok HA, Zhong Z. Dual and multi-stimuli responsive polymeric nanoparticles for programmed site-specific drug delivery. Biomaterials. 2013;34:3647-3657
  75. 75. Zangmeister RA, Tarlov MJ. DNA displacement assay integrated into microfluidic channels. Analytical Chemistry. 2004;76:3655-3659
  76. 76. Harmon ME, Tang M, Frank CW. A microfluidic actuator based on thermo responsive hydrogel. Polymer. 2003;44:4547-4556
  77. 77. Huber DL, Manginell RP, Samara MA, Kim B, Bunker BC. Programmed adsorption and release of proteins in a microfluidic device. Science. 2003;301:352-354
  78. 78. Llovet JM, Burroughs A, Bruix J. Hepatocellular carcinoma. Lancet. 2013;362:1907-1917
  79. 79. Lu J, Sun J, Li F, Wang J, Liu J, Kim D, et al. Highly sensitive diagnosis of small hepatocellular carcinoma using pH-responsive iron oxide nanocluster assemblies. JACS. 2018;140:10071-10074
  80. 80. Gao ZF, Sann EE, Lou X, Liu R, Dai J, Zuo X, et al. Naked-eye point-of-care testing platform based on a pHresponsive superwetting surface: Toward the non-invasive detection of glucose NPG. Asia Mater. 2018;10:177-189
  81. 81. Jung IY, Kim JS, Choi BR, Lee K, Lee H. Hydrogel based biosensors for in vitro diagnostics of biochemicals, proteins, and genes. Advanced Healthcare Materials. 2017;6:1601475
  82. 82. Mahajan A, Aggarwal G. Smart polymers: Innovations in novel drug delivery. International Journal of Drug Development and Research. 2011;3(3):16-30
  83. 83. Vilar G, Tulla-Puche J, Albericio F. Polymers and drug delivery systems. Current Drug Delivery. 2012;9:367-394
  84. 84. Javadzadeh Y, Hamedeyazdan S. Novel drug delivery systems for modulation of gastrointestinal transit time. In: Recent Advances in Novel Drug Carrier Systems. Rijeka: IntechOpen; 2012
  85. 85. Cao X, Lai S, Lee LJ. Design of a self-regulated drug delivery device. Biomedical Microdevices. 2001;3:109-118
  86. 86. Brahmin S, Narinesingh D, Gniseppi-elie A. Biosmart hydrogels: Conjoined molecular recognition and signal transduction in biosensor fabrication and drug delivery. Biosensors & Bioelectronics. 2002;17:973-981
  87. 87. Sheppard NF, Lesho MJ, Mc Nally P, Shaun FA. Microfabricated conductimetric pH sensor. Sensors & Actuators, B: Chemical. 1995;28:95-102
  88. 88. Sharma P, Tailang M. Design, optimization, and evaluation of hydrogel of primaquine loaded nanoemulsion for malaria therapy. Future Journal of Pharmaceutical Sciences. 2020;6:26
  89. 89. Sharma P, Tailang M. In-vivo study of orodispersible tablet of primaquine. International Journal of Pharmaceutical Sciences and Research. 2018;9(8):3506-3510
  90. 90. Li K, Yu L, Liu XJ, Chen C, Chen Q, Ding J. A long-acting formulation of a polypeptide drug exenatide in treatment of diabetes using an injectable block copolymer hydrogel. Biomaterials. 2013;34:2834-2842
  91. 91. Licciardi M, Amato G, Cappelli A, Paolino M, Giuliani G, Belmonte B, et al. Evaluation of thermo-responsive properties and biocompatibility of polybenzofulvene aggregates for leuprolide delivery. International Journal of Pharmaceutics. 2012;438:279-286
  92. 92. Kulkarni RV, Boppana R, Krishna MG, Mutalik S, Kalyane NV. pHresponsive interpenetrating network hydrogel beads of poly(acrylamide)- g-carrageenan and sodium alginate for intestinal targeted drug delivery: Synthesis, in vitro and in vivo evaluation. Journal of Colloid and Interface Science. 2012;367:509-517
  93. 93. Zhao LL, Zhu L, Liu FY, Liu CY, Shan-Dan WQ, et al. pH triggered injectable amphiphilic hydrogel containing doxorubicin and paclitaxel. International Journal of Pharmaceutics. 2011;410(83-91):26
  94. 94. Yin R, Tong Z, Yang D, Nie J. Glucose and pH dual-responsive concanavalin A based microhydrogels for insulin delivery. International Journal of Biological Macromolecules. 2011;49:1137-1142
  95. 95. Kang SI, Bae YH. A sulfonamide based glucose-responsive hydrogel with covalently immobilized glucose oxidase and catalase. Journal of Controlled Release. 2003;86:115-121
  96. 96. Ruan C, Zeng K, Grimes CA. A mass-sensitive pH sensor based on a stimuli-responsive polymer. Analytica Chimica Acta. 2003;497:123-131

Written By

Pankaj Sharma, Vinay Jain and Mukul Tailang

Submitted: 20 January 2022 Reviewed: 08 February 2022 Published: 19 March 2022