Mini Review | Volume: 10, Issue: 10, October, 2020

Non-invasive strategies for protein drug delivery: Oral, transdermal, and pulmonary

Marina Ika Irianti Ratika Rahmasari Ayun Erwina Arifianti Raditya Iswandana   

Open Access   

Published:  Oct 05, 2020

DOI: 10.7324/JAPS.2020.1010017
Abstract

Proteins are the building blocks of human life which involve physiological processes such as growth, development, metabolism, and reproduction. Despite its role in various biological processes, recently, the protein’s function has been evolving as a promising therapy. The use of protein and peptide as therapeutic agents has several advantages upon small-molecule drugs, such as high specific interaction with its target that is less likely to elicit immune response. Currently, hundreds of protein drugs are available in the market, and this number is expected to increase each year. Consequently, the growth of protein therapeutics requires several improved strategies for drug delivery processes. Generally, protein and peptide drugs are administrated parenterally by conventional injections due to its poor oral bioavailability and limited permeability across epithelial cells in the gastrointestinal tract. However, a high frequency of injections results in decreased patient compliance because of the pain and skin wound. Therefore, a lot of research has been conducted in order to study the non-parenteral route of protein and peptide drug. In this review, we discuss recent findings for non-parenteral administration of protein drugs, for instance, oral, transdermal, and pulmonary route. The recent advancements in protein drug delivery make the non-parenteral route a promising method for protein drug delivery because of the ease of use among patients.


Keyword:     Protein peptide therapeutic agent drug delivery system non-parenteral oral transdermal pulmonary.


Citation:

Irianti MI, Rahmasari R, Arifianti AE, Iswandana R. Non-invasive strategies for protein drug delivery: Oral, transdermal, and pulmonary. J Appl Pharm Sci, 2020; 10(10):166-179.

Copyright: © The Author(s). This is an open-access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

HTML Full Text

INTRODUCTION

Proteins are complex amino acids, usually containing more than 50 different amino acids, while peptides consist of less than 20 amino acids (Ratnaparkhi et al., 2011). For the last 10 years, the use of therapeutic protein has been increased, and there are hundreds of approved protein therapeutic products in the market (Fosgerau and Hoffmann, 2015). The use of protein and peptide drugs has been attracted to its high selectivity toward the target. However, peptide or protein-based drugs have very low stability and bioavailability, which also require a very high production cost (Craik et al., 2013). Generally, protein and peptide drugs are administrated parenterally (Jitendra et al., 2011) because of their short half-life and low bioavailability issue (Hamman et al., 2005). Moreover, Peptide-based drug formulation is quite challenging because they are unstable and susceptible to aggregation or oxidation reactions, which subsequently affect the activity of protein-based drugs (Torosantucci et al., 2014). As peptides contain smaller polypeptides, it is difficult for them to form globular structures and tertiary structures. Therefore, peptides tend to be more susceptible to degradation particularly in solution. Compared to protein, the peptide-based drug formulation is harder because of the chemical and physical instability, also because of the tendency of peptides to shape different conformations (Payne and Manning, 2009).

The foremost challenging problem of the parenteral administration is its short half-life, which is related to enzymatic degradation and rapid renal clearance. Besides, this route is administrated frequently, leading to patients’ inconvenience and incompliance (Cleland et al., 2001; Schiffter et al., 2011). The challenges that protein and peptide drugs present have encouraged strategies to focus on improving the bioavailability through the delivery system and innovative formulation strategies. These strategies aim to improve protein stability during manufacturing and storage until the drug reaches the intended biological target. In this review, we will discuss the recent findings on the non-parenteral administration of protein or peptide drugs, which cover oral, transdermal, and pulmonary routes, and discuss the recent formulation technology in improving protein or peptide drug bioavailability.


PHYSICOCHEMICAL DRUG PROPERTIES OF PROTEINS AND PEPTIDES

Recent advances in the field of genetic engineering and pharmaceutical biotechnology have made it possible to treat various life-threatening diseases using therapeutic proteins. The U.S. Food and Drug Administration Center for Drug Evaluation and Review and the Center for Biologics Evaluation and Review have approved many recombinant therapeutic proteins, consisting of monoclonal antibodies (Mab), coagulation factors, and replacement enzymes, as well as fusion proteins, hormones, growth factors, and plasma proteins (Food and Drug Administration, 2020). Recombinant protein therapy was developed to treat various clinical indications, such as cancer, inflammation, exposure to infectious agents, and genetic disorders (Sauna et al., 2017). In addition, protein has also been proven effective as a vaccine that helps stimulate the body’s natural defense mechanism against immunogenic responses (Akash et al., 2015).

Most of the therapeutic proteins are given by parenteral route because of its instability, size, and poor transport (Wagner et al., 2018). Proteins also have a short half-life and high elimination rate; therefore, (Martins et al., 2007). This can be a burden on patients with increased costs and decreased comfort. Giving therapeutic proteins via the oral route can be an alternative to improve patient compliance. However, the unstable nature of proteins in an acidic environment and susceptibility to proteolytic enzymes in the gastrointestinal tract become a challenge in oral administration (Wagner et al., 2018).

Protein solubility is strongly influenced by pH, ions and temperature. At the isoelectric point, the solubility of the protein is very low. Proteins are very hydrophilic with very small partition coefficients in octanol-water solvents. Therefore, protein absorption by passive diffusion needs to be increased by increasing lipophilicity (Ratnaparkhi et al., 2011).

Although generally in the solid phase, peptides and proteins still undergo various degradation reactions, namely chemical and physical degradation (Capelle et al., 2007). Chemical degradation involves a covalent modification of the primary structure of proteins by breaking or forming bonds. While, physical degradation refers to changes in structure that are more structured due to denaturation and aggregation or noncovalent precipitation (Feridooni et al., 2016). The mechanisms of chemical degradation that often occur in the solid phase include deamidation, oxidation, and the Maillard reaction.

Deamidation is the process where amide side chain hydrolysis of glutamine or asparagine residues produce carboxyl acids (Chang and Pikal, 2009). The deamidation reaction that often occurs in proteins in drug formulations is the nonenzymatic intramolecular deamidation reaction of Asn residues. In contrast to the deamidation reaction, the potential for a degradative oxidation reaction can be found at various stages of production, packaging, and storage (Feridooni et al., 2016). For example, peroxide contamination that has been found in formulation excipients, such as polyethylene glycol and surfactants, which causes oxidation of these products. Activation of molecular oxygen into more reactive species requires light or reducing agents and trace levels of transition metal ions, which can then convert molecular oxygen to more reactive oxidizing species, such as superoxide radicals (O2-*), hydroxyl radicals (*OH), or hydrogen peroxide (H2O2). Transition metal ions are often present in excipients and production processes with stainless steel equipment can cause significant iron contamination (Chang and Pikal, 2009). This potential source can contribute to the degradative oxidation of protein drugs.

After formulation, the selection of the final packaging can also have an effect on drug stability. Research results have shown that overall oxidation in certain products can still occur despite low oxygen levels (1%) on the vial's head (Chang and Pikal, 2009). By understanding the potential sources of contaminants that cause oxidation at all stages of drug production and the mechanism by which oxidation reactions occur, formulation strategies can be designed to minimize these events.

Protein drugs often experience physical changes that can cause changes in pharmacological effects and potential. Physical instability includes changes in the integrity of the three-dimensional conformation of proteins and does not always involve covalent modification (Feridooni et al., 2016). The physical process includes denaturation, aggregation, precipitation, and adsorption on the surface (Lai and Topp, 1999). Protein drugs can undergo these changes during manufacturing, shipping, storage and administration. In recent years, aggregation has become a major problem in therapeutic proteins (Ameri et al., 2009). Protein aggregation is a multilevel process that involves unfolding or misfolding units of protein monomers together with one or more steps of assembling protein monomers to form soluble or insoluble oligomers or aggregates with higher molecular weight (Li et al., 1996). Protein aggregation can be a problem during the drug manufacturing process, especially if the drug is insoluble and tends to experience precipitation. This process usually reduces drug stability and half-life (Weiss et al., 2009). Shear stress, high temperatures, pH changes, and high protein concentrations are factors that trigger protein aggregation (Frokjaer and Otzen, 2005).


BIOLOGICAL BARRIERS OF PROTEIN AND PEPTIDE DRUG DELIVERY

Protein and peptide drug delivery remains challenging due to the biological barriers in human body. The protein and peptide drugs should be sustained under enzymatic, pH changes and the mucosal barrier in the gastrointestinal tract. The presence of proteolytic enzymes in human biological process, such as proteinases, peptidases, and proteases, has an effect on the delivery of protein and peptide drugs. The proteases, generated by human cells, consist of aspartic proteases, threonine, cysteine, serine, and metalloproteinases (Choi et al., 2012). When the protein and peptide drugs reach the colon, the microorganisms in colon may generate peptidases to hydrolyze the peptide bonds. According to the site of actions, proteases are categorized as exopeptidases and endopeptidases. The cleavage of these peptidases through the hydrolysis process is irreversible and leads to protein degradation (Mahato et al., 2003).

Besides the presence of enzymatic reactions, the pH across the gastrointestinal tract also changes the stability of the protein and peptide drugs, for instance, the extreme pH in the stomach may lead to protein hydrolysis. The alteration of pH affects the ionic and hydrogen interaction in the protein, which subsequently transforms the protein conformation and folding (Mahato et al., 2003). The protein folding is important for the biological activity and its misfolding, which causes dysregulation of the protein function (Dobson, 2003). Therefore, the delivery of protein and peptide drugs through oral route faces difficulties due to the different pH conditions in the gastrointestinal tract. The pH in the GI tract is varied from acidic condition (pH 1.2–3.0) to alkaline condition (pH 6.5–8.0) (He et al., 2019).

The most notable biological barriers for protein and peptide drug delivery are presented in the brain and intestine. In the intestine, the barrier for protein and peptide drugs delivery is comprised of epithelial cells, lamina propria, and the muscularis mucosae, while the blood–brain barrier (BBB) is composed of luminal and abluminal membranes of the brain endothelium (Ulapane et al., 2017). Once the protein enters the GI tract, it is destructed into amino acids and is absorbed through the intestinal epithelium. However, the amino acids are not easily absorbed since the brush border (microvilli) in the epithelium contains digestive enzymes. In addition, the existence of glycocalyx and mucus made the absorption through intestinal epithelium even more difficult (Carino and Mathiowitz, 1999). It should be taken into account that there are differences for drug metabolism in different parts of the intestines for both men and women (Iswandana et al., 2018). Subsequently, the regional differences of drug metabolism also affect the protein and peptide drug absorption in different parts of human intestines.

Besides the intestinal epithelium, the delivery of protein and peptide drugs across the BBB drug delivery is hardly successful. In the brain, the protein and peptide drugs are transported across the cerebrovascular endothelium. The endothelium is characterized by very tight junctions. It is challenging for the hydrophilic molecules to pass the tight junction; in contrast, hormones and peptides were allowed to cross the BBB through the receptor-mediated transcytosis (van Bree et al., 1990). Large molecules, such as proteins, need to be reengineered structurally for crossing the BBB. One of the approaches for BBB delivery is utilizing antibodies that bind to the transferrin receptor (TfR). This molecular tool, namely “Trojan horse” technology, is mainly a peptidomimetic monoclonal antibody or endogenous peptide that passes the BBB through receptor-mediated transport. The transferrin receptor monoclonal antibody a fusion of TfR and Mab, allows the penetration to BBB and acts like the lipophilic small molecules (Pardridge, 2015). However, the weakness of this method is the short half-life of the anti-TfR antibodies. To overcome this limitation, another technology is reported to lengthen the antibodies half-life, called AccumuBrain. The AccumuBrain helps to raise the antibody concentration in the blood by binding to myelin oligodendrocyte glycoprotein, which is present in oligodendrocytes. It was shown that AccumuBrain stimulates the antibody levels ten times higher than the anti-TfR antibodies (Nakano et al., 2019).

Figure 1. Barriers to peptide delivery in epithelial cells (Adapted from Bruno et al., 2013).

[Click here to view]


NON-PARENTERAL DELIVERY ROUTE

Oral

The oral route of drug administration is desired due to its convenience for the patients (Park et al., 2011). However, one of the problems for oral delivery drugs is the absorption in the gastrointestinal tract, particularly for protein-based drugs that are instable (Schiffter et al., 2011). The degradation of the peptide may be due to biochemical factors, such as the acidic condition in the gastrointestinal tract and the presence of microorganisms that contribute to the metabolism of the peptides (Hamman et al., 2005). Besides biochemical barriers, there are also physical barriers consisting of an unstirred water layer that restricts the transportation of peptides to the epithelial cells, the columnar epithelial cells in intestines, tight junctions, and efflux systems as shown in Figure 1.

As shown in Figure 2, some mechanisms are possibly involved in the absorption of proteins and peptides in the gastrointestinal tract, such as passive transport through diffusion, active transport, and endocytosis (Bruno et al., 2013). Thus, the protein is transported through transcellular and paracellular pathways (Mnard et al., 2012). In the transcellular pathway, the transport of protein molecules across the cell can be by either passive diffusion or utilizing specific carriers (Mnard et al., 2012). The passive diffusion depends on the characteristics of the drug molecule, including molecular weight and charge, which allows the molecules to travel from a high concentration in the intestinal lumen to a lower concentration in the blood (Zhu et al., 2017). On the other hand, the drug molecules can also be transported using certain carriers, such as a peptide or amino acid transporters, to facilitate the drug across the cells (Zhu et al., 2017). The paracellular pathway requires the transport of small hydrophilic molecules (<200 Da) between the adjacent cells (Antunes et al., 2013).

Some strategies have been developed to obtain the optimal delivery of protein or peptide drugs through oral administration, such as modifying physicochemical structures of the drugs and creating site-specific delivery of peptides (Kumar et al., 2007). Another approach is designing prodrugs, which can be converted to the parent molecules due to the metabolism process (Gangwar et al., 1997). Mainly, formulation technologies have been initiated for enhancing protein and peptide drug bioavailability, including the use of permeation enhancer and protease inhibitors (Choonara et al., 2014). Several small-molecules categories may enhance the permeation of protein and peptide drugs, such as acids and surfactants; also, human bile salts help for the protein and peptide permeation (Brown et al., 2020). Acids, like citric, fumaric, and tartaric acids, help to decrease the adjacent cells integrity by binding to Ca2+ which initiate the protein kinase C activation (Tomita et al., 1996; Brown et al., 2020). Surfactants and bile salts are amphiphilic molecules which facilitate the permeation of protein and peptide-based drugs across the paracellular and transcellular routes by altering the membrane integrity (Brown et al., 2020; Maher et al., 2019)

Figure 2. The pathways of protein and peptide drugs in intestinal epithelium: (a) transcellular pathway, (b) paracellular pathway, (c) receptor-mediated endocytosis and transcytosis, and (d) transportation to lymphatic circulation through M-cells of Peyer’s patches (Adapted from Goldberg and Gomez-Orellana, 2003).

[Click here to view]

The use of plant cells is considered as one of the innovative approaches that has been studied to enable the oral administration of protein drugs for several diseases, such as Gaucher’s disease, diabetes, Alzheimer’s disease, ocular disease, and hypertension (Kwon and Daniell, 2016). The plant cells are used to encapsulate the protein drugs, so they are protected from the acids in the stomach and then digested by the microorganism in the intestines (Kwon and Daniell, 2016). Several examples of technologies that have been developed by pharmaceutical companies for oral protein drug delivery are presented in Table 1.

Transdermal

Transdermal drug administration has been generated recently because it allows the prevention of first-pass metabolism, especially for short half-life drugs. In addition, this type of administration is noninvasive and more convenient for patients (Kalluri and Banga, 2011). Some existing drugs have been delivered by using transdermal route, i.e., nicotine, estrogen, and scopolamine.

Anatomically, the skin is composed of four layers: non-viable epidermis (stratum corneum), viable epidermis, viable dermis (corium), and subcutaneous connective tissue (hypodermis) (Kanikkannan et al., 2012). Among these layers, the stratum corneum is the outermost part of the skin, which contains dead keratinocytes (mainly 75%–85%) and lipids (5%–15%) (Kanikkannan et al., 2012). This layer acts as a barrier for peptide drugs because it limits the absorption of large molecular weight and hydrophilic molecules (Kalluri and Banga, 2011). The viable epidermis is located beneath the stratum corneum, with 50–100 µm thickness and is composed of 90% water (Pathan and Setty, 2009). Just below the viable epidermis, there is the dermis, which is approximately 2–3 mm thick, and contains fibrous protein (Pathan and Setty, 2009). The lower layer of the skin is namely hypodermis that is composed of fibrous connective tissue, sweat gland, and cutaneous nerves where the drug is initiated to enter the circulation system (Pathan and Setty, 2009). When the drug is administrated parenterally, it first enters the outer layer of the skin and penetrates across the stratum corneum. The drug partition is continued to the viable epidermis and then available for systemic absorption when reaching the dermis (Alkilani et al., 2015).

Technology and formulation approaches are initiated to increase the penetration of drugs and overcome the stratum corneum barrier. Some strategies are explored in the formulation of transdermal delivery of protein drugs, such as using chemical enhancers, nanocarriers, and prodrugs (Chaulagain et al., 2018). Generally, the use of a chemical enhancer is mostly used as a formulation approach (Banga et al., 2013). However, more technology techniques are involved in transdermal delivery of protein and peptide drugs, such as microneedles, electroporation, thermal and radiofrequency ablation, sonophoresis, and iontophoresis (Kalluri and Banga, 2011).

Microneedle

Microneedle is one of the major technologies developed to increase penetration. Various types of microneedles are used in protein drug delivery, as shown in Figure 3. The first method is initiated by perforating the skin to make pores and is continued by the application of the drug-loaded patch. The pores allow the diffusion of drugs to the lower layer of the skin (Li et al., 2010). The second method is the insertion of microneedles that are covered with drugs (Verbaan et al., 2007). Upon the insertion of microneedles, the drugs are released and dissolved in the skin. The drug-coated microneedles have limitations for the amount of protein that can be used. The third method is called soluble microneedles and they are usually made from biodegradable excipients, such as carboxymethyl cellulose (Lee et al., 2008). The fourth method is hollow microneedles, where the liquid drugs can be infused from the reservoir. For the last few years, microneedle arrays have gained a lot of interests in their application in delivering proteins or peptides. However, there is truly no microneedle array of products in the market yet (Larrañeta et al., 2016). Several microneedle devices have been developed by companies, such as Solid Microstructured Transdermal System (3M), Microinfusor (BD technologies), microinjection patch Macroflux® (Alza), Micro-Trans™ Microneedle Array Patch and h-Patch™ developed by Valeritas, and also Soluvia® and Microinjet® which are available in the market (Larrañeta et al., 2016).

Table 1. Examples of technologies in various administrations of protein drug delivery.

[Click here to view]

Figure 3. Different types of microneedle technologies: (a) porous formation by solid microneedle before the application of drug patch, (b) coated microneedle, (c) dissolving microneedle, and (d) hollow microneedle (Adapted from Kalluri and Banga, 2011).

[Click here to view]

Iontophoresis

Iontophoresis technique, in principle, is the use of mild electric current (approximately less than 0.5 mA/cm2) to drive the charged drug molecules into the skin (Gujjar and Banga, 2014). Mainly, the charged molecules are transported through the electro-migration mechanism, while the neutral molecules are delivered by electro-osmosis (Kalluri and Banga, 2011). Besides the current strength and density, other parameters also affect the iontophoretic drug delivery, like the drug features and formulation, patient biological condition, and experimental factors related to the current and electrode material pH, electro-osmosis transport, and patient anatomical factor (Khan et al., 2011). This method has limitations for proteins with a size more than 15 kDa. Thus, the liability of protein to form aggregates should also be considered. The administration of Interferon alpha-2B (hIFN-2b) in hairless rats was improved by using iontophoresis (Badkar et al., 2007).

Electroporation

Electroporation technique is generally used in the transformation method for bacteria cells. The high voltage aims to make the bacteria cell membrane become more permeable for DNA insertion (Miller et al., 1988). The principle in electroporation is used for transdermal drug delivery, where the electric field helps to improve the skin permeability and allows the penetration of protein drugs into the skin (Kalluri and Banga, 2011). The voltage (around 50–500 V) is required to create pores on the skin so that the large molecules will penetrate the skin (Szunerits and Boukherroub, 2018). Electroporation helps the diffusion of insulin delivery on rabbits’ skin (Mohammad et al., 2016). The application of both electroporation and iontophoresis resulted in a synergistic effect in transdermal administration of human parathyroid hormone (Medi and Singh, 2003). Moreover, electroporation has been used in combination with a microneedle roller and a flexible interdigitated electroporation array to deliver nucleic acid-based drugs (DNA and siRNA) onto the mouse skin (Huang et al., 2020).

Thermal and radiofrequency ablation

Thermal and radiofrequency ablation utilizes high temperatures to deliver protein or peptide drugs through the disruption of stratum corneum (Aljuffali et al., 2014). The heat creates pores and ablation which help the protein drugs to enter the skin (Szunerits and Boukherroub, 2018). One of the patented products was initiated by Altea Therapeutics (Atlanta, GA), namely PassPort™ patch. This device is ideally applied for proteins and peptides with a molecular weight less than 10 kDa (Banga, 2006).

Pulmonary

Advanced technologies for pulmonary delivery are widely studied in the last two decades since the lungs can be used as a portal for systemic drug delivery. The pulmonary delivery, such as aerosol, has the advantage of high drug concentration to the airway, which decreases the adverse effect and is painless (Hess, 2008). Moreover, the lack of the first-pass metabolism in the pulmonary route gives a higher possibility for the lung to become an advantageous route of entry for peptide and protein drugs to the body (Agu et al., 2001; Ibrahim et al., 2015). Two different technologies that have been used to deliver drugs through the pulmonary route are pressurized metered-dose inhaler (pMDI) and dry powder inhalers (DPIs) (Ibrahim et al., 2015). Nebulizers consist of jet and ultrasonic nebulizers, which can be differentiated based on the force used to atomize the liquid (Ibrahim et al., 2015). According to Venturi’s principle, the fluid pressure in aerosol declines as it moves through a diminishing area (Watts et al., 2008). The challenges in jet nebulizers are the need for compressors to produce the aerosol, the sound that it generates, and the temperature fall because of the liquid evaporation in the nebulized globules (Rubin and Williams, 2014). The sound waves in ultrasonic nebulizers are produced because of the high-frequency vibration from piezoelectric crystals, resulting in crests that split the liquid into small droplets. Ultrasonic nebulizers are costlier compared to jet nebulizers (Dolovich and Dhand, 2011; Ibrahim et al., 2015).

Moreover, this type is not efficient anymore in nebulizing viscous liquids and suspensions because it is less portable due to the need for electricity and it tends to raise the temperature of the nebulized drug solution. Therefore, they are considered inappropriate to nebulize thermolabile peptides or DNA. Generally, nebulizers generate 1–5 µm droplets according to the model and the manufacturer. Nebulizers have advantages over pediatric, geriatric, ventilated, non-conscious patients, or those who cannot use pMDIs or DPIs. Also, nebulizers are potential for administrating larger doses than other aerosol devices. However, this will need longer delivery times (Ibrahim et al., 2015).

The accumulation of aerosolized particles in the oropharyngeal domain and upper airways and the lack of synchronization between the device activation and inhalation are the main problems with the use of inhaler devices. In general, pMDIs generate aerosol faster than the patient can inhale. Therefore, children and elderly find it difficult to make a coordination between device actuation and inhalation. On the other hand, the use of DPIs requires the inhalation of the patient to be at maximum power in order to disperse and inhale the powder. However, this requirement is rarely achieved if the patient is not properly trained (Ibrahim et al., 2015). The volume of the drug solution, the viscosity, the airflow and pressure, the tubing, mask, or mouthpiece utilized in the device are some factors that must be considered to get a precise and uniform dose with the nebulizer (Ibrahim et al., 2015). The limited optimization of these variables causes dose variability among the patients. A drawback for nebulizer users is the need for assembling and loading the medication before usage. Also, the users have to de-assemble and clean the device for another usage (Hess, 2008). Insulin and interferon are two examples of protein-based drugs that have been widely studied for pulmonary delivery (Agu et al., 2001; Oleck et al., 2016). When the drug is administrated through the pulmonary route, it can be absorbed through the membrane pores, vesicular, intracellular tight junction, and transporter-mediated transport (Ibrahim and Garcia-Contreras, 2013). However, there are some challenges in protein and peptide-based drug administration, such as the degradation of protein by protease or macrophage; also, the presence of both mucus and surfactant in the alveoli can limit protein absorption (Agu et al., 2001).

Insulin

Insulin is extremely needed in diabetes management, but unfortunately insulin absorption through oral administration is poor. A market innovation in delivering insulin through pulmonary delivery was achieved by the first two rapid-acting inhaled insulins in the market, which are Exubera® in 2006 and Afrezza® in 2014. Inhaled insulin is an advantage for people who have incorrect injection methods or needle phobia. However, that inhaled insulin was withdrawn due to a poor sales volume from low insurance coverage, finding new concern about the adverse effects, and competition from other insulin alternatives. As a pulmonary delivery, contraindications would included smokers and respiratory diseases like asthma due to a change in pulmonary lung function. Besides, the risk of respiratory adverse effects was also increased, such as cough, pharyngitis, rhinitis, and respiratory infection (Banga, 2015).

Interferon

Interferon is another type of drug that is potentially administered through the lungs. A study from Jaffe et al. (1991) compared aerosol and subcutaneous injection in humans delivering recombinant interferon-γ (rINF-γ) to activate alveolar macrophages for cytokine therapy (Jaffe et al., 1991). Compared to parenteral drugs that usually have a systemic side effect, the inhalation drug delivery had a better acceptance with no side effects. The lower significant systemic concentrations rINF-γ may be due to high drug deposition in the lung and reduced inhalation absorption rate. Another interferon study from Dai et al. (1987) with 7 years of clinical study in China showed INF-α aerosol treatment as an effective and safe therapy for viral diseases, including asthma, asthmatic bronchitis influenza, bronchiolitis, mumps, and recurrent upper respiratory tract infections in children (Banga, 2015, Dai et al., 1987).


FORMULATION TECHNOLOGY APPROACHES

Despite the efforts to administrate peptides and proteins through a noninvasive delivery system, the instability of protein drugs due to enzyme degradation, pH, low bioavailability, and toxicity are still the foremost challenging problem. Several novel strategies in the formulation of protein or peptide drugs have been developed to face these challenges. The examples of technologies that have been developed by the pharmaceutical companies for oral protein drug delivery are presented in Table 1. Chemical modification and development of colloidal carriers are the formulation approaches that can be applied for nasal, transdermal, and pulmonary delivery (Bajracharya et al., 2019), and these approaches are discussed in this article.

Chemical modification

Since the manipulation of the peptide and protein structure is less feasible to increase half-life time, the current approach in chemical modification goes to the addition of covalent conjugation of the polymer, such as mannosylation, PEGylation, and hyperglycosylation. Mannosylation of protein is endowed to target the mannose receptor cells, which are highly expressed by macrophages, dendritic cells, hepatic, and lymphatic endothelial cells. The in-vivo study showed that mannosylated protein therapeutics result in a better therapeutic outcome as reported in the enhancement of Antigen-specific antibody and T lymphocyte response after the administration of mannosylated mucin-type immunoglobulin fusion protein (Ahlén et al., 2012)

PEGylation is a protein modification that conjugates to polyethylene glycol (PEG) in order to enhance protein delivery. To provide a suitable conjugation site, PEG is usually conjugated to amine terminal which allows the suitable conjugation site (Parveen and Sahoo, 2006; Ryan et al., 2008). The addition of PEG alters the solubility and steric hindrance of proteins, resulting in better stability, increased half-life time, and optimal pharmacokinetic. The steric hindrance ability of higher PEGylation causes the reduction of contact with the active site. However, it shields the protein from enzymatic degradation and reduces contact with the antigen-presenting cell. Consequently, it increases the systemic circulation and therapeutic outcome (Patel et al., 2014).

The physicochemical alteration of PEGylated protein and peptide because of PEG characteristic can improve the systemic circulation and reduce renal filtration (Parveen and Sahoo, 2006). PEG has been successful in generating a market for protein therapeutics as seen in the first Food and Drug Administration (FDA) approved PEGylated-protein drug in 1990, namely adenosine deaminase, Adagen®, which contains an enzyme for severe combined immunodeficiency disease. This successful production was followed by other drugs such as doxorubicin liposomal, Doxil® as an antineoplastic drug, PEGinterferon alfa-2a PEGasys®, as anti-hepatitis B and C, an opioid antagonist for opioid addiction Movantik® (Bailon et al., 2001).

The success of previous PEGylated-protein drugs has proved to increase half-life, the stability of protein-based therapeutics and enhance peptide, and protein delivery (AlQahtani et al., 2019). They are thus triggering further development of protein therapeutics through PEGylation, such as for filgrastim (methionyl human granulocyte colony-stimulating factor, rh-met-G-CSF) produced by recombinant DNA technology. Filgrastim has a function to regulate the production of neutrophils within the bone marrow (Welte et al., 1987). In order to treat neutropenia, filgrastim has to be administered every 24 hours continuously for 11–20 days to maintain steady-state serum concentrations. The current research showed that enzymatic and nonenzymatic PEGylated of filgrastim prolonged the stability and plasma half-life in vitro (Scaramuzza et al., 2012). However, further preclinical and clinical trials of this enzymatic PEGylated filgrastim are needed.

Hyperglycosylation is a co- or post-enzymatic proses which conjugate protein, or other organic molecules with the polysaccharide. Hyperglycosylation improves the pharmacokinetic profile of peptide and protein therapeutics. There are two types of hyperglycosylations: in-situ chemical reaction and site-directed mutagenesis, which can result in N-linked or O-linked protein glycosylation. The N-linked is specifically attached to asparagine. Meanwhile, O-linked oligosaccharide is not site-specific, but is generally found binding to serine or threonine (Pisal et al., 2010). The addition of carbohydrates may stabilize protein by the formation of hydrogen bonds with polypeptide backbone or surface hydrophilic amino acid and steric interaction with the adjacent residues (Patel et al., 2014). Hyperglycosylation may also work to hinder the human immune system, such as polysialic acid (PSA), which are available at different sizes of molecules. Also, PSA is also able to control the clearance rate of conjugated proteins or peptides. The terminal ends of every glycan added for hyperglycosylation usually contains a functional structure, such as phosphate, sulfates, and carboxylic acids, which can alter the protein surface charge, isoelectric point, and increase half-lifetime of circular hyperglycosylated protein (Solá and Griebenow, 2010). Notably, another advantage of hyperglycosylation is the nature of its biodegradability in the human body. Examples of FDA-approved hyperglycosylated proteins are Cerezyme®, Fabrazyme®, and Naglazyme®.

Mannosylation of protein is endowed to target the mannose receptor cells, which are highly expressed by macrophages, dendritic cells, hepatic, and lymphatic endothelial cells. The in-vivo study showed that mannosylated protein therapeutics result in a better therapeutic outcome as reported in the enhancement of Antigen-specific antibody and T lymphocyte response after the administration of mannosylated mucin-type immunoglobulin fusion protein (Ahlén et al., 2012).

Colloidal carrier

Colloidal carrier, a lipid-based formulation, has been widely used to overcome delivery problems of peptides and protein drug. This carrier can protect the drug against degradation in vitro and in vivo, modify the release rate, as well as target specifically in the body (Martins et al., 2007). The simple approaches of colloidal carrier are nanoemulsions (NEs), microemulsions (MEs), and nanogels (NGs).

Figure 4. Different types of microneedle technologies: (a) porous formation by solid microneedle before the application of drug patch, (b) coated microneedle, (c) dissolving microneedle, and (d) hollow microneedle (Adapted from Kalluri and Banga, 2011).

[Click here to view]

NEs are a colloidal dispersion system consisting of two immiscible liquids (water and oil), in which one liquid is dispersed in the other by means of an appropriate surfactant/co-surfactant mixture, forming oil-in-water (o/w) or water-in-oil (w/o) nanodroplet systems, with droplets of 20–200 nm in size. NEs are a very cost-effective technique due to high storage stability and ease of preparation. Despite the similarities of NEs and MEs in terms of their physical appearance, components, and preparation techniques, NEs are kinetically stable and thermodynamically metastable, while MEs are thermodynamically stable (Shaker et al., 2019).

NGs are nanosized, three-dimensionally, cross-linked, hydrophilic polymeric networks that are composed of hydrogel particulate entities with a nanometer-sized space; so it has the features of hydrogel (high water content and versatile mechanical properties) and nanoparticles at the same time. Dimensions, less than 200 nm in diameter, facilitate cellular uptake through receptor-mediated endocytosis, making NGs suitable carriers for the peptide drugs. NGs may have a role as chaperones in preventing denaturation or aggregation of proteins, promoting refolding, and controlling the release rate. When proteins are encapsulated within the cross-linked polymer matrix of NGs, higher stability is reported even at temperatures above the physiological values and in the presence of organic solvents (Grimaudo et al., 2019; Zhang et al., 2016).

Other colloidal carriers that can be used to deliver protein and peptide drugs are liposomes, microparticles carbon nanotubes (CNTs), and nanoparticles (polymeric nanoparticles, solid lipid nanoparticles (SLNs), micelle, and CNTs (Bajracharya et al., 2019; Patel et al., 2014). The structures of each colloidal carrier are shown in Figure 4. Compared to chemical modification approaches, colloidal carriers exhibit the ability to protect sensitive proteins and prolonged release.

Compared to the liposome, which are vesicular nanostructures made of phospholipid and amphipathic lipid, microparticles and nanoparticles have a better kinetic morphology and rigid structure (Battaglia and Ugazio, 2019). Microparticle biodegradable polymers are extensively studied to provide controlled release over months (Shi and Li, 2005). Polymer types as a coating material, fabrications method, and formulation are identified as important factors for microparticles. The coating material needs to have the biodegradable ability as they can break into nontoxic material in the body and can easily (Fredenberg et al., 2011). Three major subsets of polymers are natural, semi-synthetic, and synthetic, which are known to be used as coating material such as starch, alginate, collagen, chitosan, lecithin, ethyl-cellulose, cellulose acetophthalate, polyesters poly(glycolic acid), poly(D,L-lactic acid), and poly(D,L-lactic co-glycolic acid) (PLGA) (Kamaly et al., 2016; Saez et al., 2007). Due to the biocompatibility and capacity to achieve different drug release, PLGA and lactic acid homopolymers are mostly employed as coating material (Saez et al., 2007). The success of PLGA as a microencapsulated polymer was shown with the availability of marketed drugs, such as Nutropin® Depot, as a treatment for growth disorder pediatric patients with monthly administration, Trelstar® Depot and Plenaxi®, as a treatment for advanced prostatic cancer (Patel et al., 2014; Saez et al., 2007). Even though it advances in control drug release, there is a major concern related to the possible initiation of immunological response, which can be triggered by degraded protein because of the fabrication process (Van De Weert et al., 2000). Thus, fabrication and formulation methods become important. Although the development of better methods and formulations are ongoing, solvent evaporation (single and double emulsion process), phase separation (coacervation), spray-drying emulsion techniques are being used to enhance protein stability (Makadia and Siegel, 2011). Meanwhile, various additions of excipients can be added to sustain the release of the drug, such as the addition of alginate, chitosan, and caffeic acid-grafted PLGA (Han et al., 2016; Selmin et al., 2015, Zheng and Liang, 2010).

Two types of nanoparticles – polymeric and SLNs – have been widely investigated. Both of them are alternative carrier systems, not only for hydrophilic but also insoluble and labile compounds. In addition, they are also able to deliver the drugs in a sustained release manner and reduce the degradation of labile compounds. These systems have less toxicity compared to other because the matrix is biodegradable and well tolerated in the human body (Campos et al., 2015).

Polymeric nanoparticles are colloidal carriers that have 1–1,000 nm in size. The extreme biocompatibility of polymeric nanoparticle makes it an efficient nanocarriers in the medical field. Two types of polymeric nanoparticles are nanocapsules (a polymeric membrane containing protein/peptide), and nanospheres (protein/peptide are well distributed into the polymeric matrix). Both of these can be generated via fabrication methods. The above-described preparation of microparticles can be employed for polymer nanoparticles preparation. The most widely used technique for hydrophobic encapsulation is solvent evaporation techniques (Makadia and Siegel, 2011; Mao et al., 2007). Nanoparticles and salting-out (w/o/w) are the alternative techniques for nanoparticle encapsulation (Hans and Lowman, 2002; Kwon and Daniell, 2016; Lamprecht et al., 2000). On the other hand, the active compound loading into nanoparticles can be done by two methods, by incorporating during nanoparticles production or incubating the nanoparticles with concentrated active compounds (Yih et al., 2006). Since the active component was encapsulated or well distributed in its polymeric matrix, the release of active compounds follow the diffusion process with three step, which are matrix swelling, the rubbery matrix formation and active compound diffusion through rubbery matrix (Jawahar and Meyyanathan, 2012).

Although polymeric nanoparticles are the site-specific target and control the drug release, polymer internalization in the cells gives the possibility of cytotoxic induction (Smith and Hunneyball, 1986; Wissing et al., 2004). Therefore, SLNs can be an alternative for nanoparticles with better cell tolerability. SLNs are lipid nanoparticle systems that have a size less than 1,000 nm. They are stabilized by an emulsifier together with tolerated lipid contents, such as triglycerides, diglycerides, monoglycerides and fatty acids (Abhishek et al., 2019). SLNs are very useful carriers for active compounds and the mobility was restricted by lipid; therefore, it can lead a modified release profile (Wissing et al., 2004).

However, SLNs still have disadvantages related to the loading capacity of the drug (25%) and the formation of aggregates (Wissing et al., 2014). A combination of suitable preparation techniques might help in reducing those disadvantages (Abhishek et al., 2019). A derivate of SLNs is introduced to overcome the disadvantages of conventional SLNs, such as nanostructured lipid carriers and Lipid protein conjugate. Some of the techniques can be used for SLNs and its derivative fabrication, such as High-pressure homogenization, MEs, solvent emulsification–evaporation or diffusion, high-speed stirring, ultrasonication, and double emulsion method (Cortesi et al., 2002).

Furthermore, another nanopolymeric carrier, called dendrimers, has attracted the interest of scientists in delivering therapeutics, targeting, and diagnostic agents together in a single system. Dendrimers are interesting for biomedical applications because of their properties, including hyperbranching, well-defined globular structures, excellent structural uniformity, multivalency, variable chemical composition, and high biological compatibility (Noriega-Luna et al., 2014). Previously, Ciolkowski et al. (2012) showed the influence of dendrimers’ surface modification on the strength of interaction with proteins. This study was performed using poly (propylene imine) G4 and G3.5 polyamidoamine dendrimers as a drug carrier and a model protein from hen egg white lysozyme. Moreover, Liu et al. (2019) reported a boronic acid-rich dendrimer for cytosolic delivery of native proteins. This system could deliver 13 cargo proteins into the cytosol of living cells and maintained their bioactivities after cytosolic delivery.

Other colloidal carriers that contribute to peptide delivery are micelles and CNTs. Micelles have advantages over others on its particle size which ranges from 10 to 100 nm (Zhang et al., 2014). PEG as a hydrophilic segment and lipid as core segment is usually used for its amphiphilic block. In protein delivery, water-in-oil-in-water micelles are more preferable, since the proteins will be entrapped in an aqueous chamber of micelles. Recently, micelles have been known as first-line drug delivery because of their size and large manufacturing feasibility (Kim et al., 2010). The stability of micelles has become one of the factors that make it feasible for manufacturing. In order to maintain its stability, some techniques, such as shell cross-linking (covalent bond between shellcore) and noncovalent cross-linking (static electric interaction), have been used widely (Lu et al., 2018). For example, micelles technology has been applied for insulin to make slower degradation and controlled release (Li et al., 2016).

One type of micelles technology is polyion complex micelles (PIC) that can also be used to deliver organic solvent-sensitive therapeutic agents, such as proteins and nucleic acid, which are naturally occurring polyelectrolytes (Chen and Stenzel, 2018). Harada and Kataoka (1998) prepared a PIC from chicken egg white lysozyme and poly(ethylene glycol)–poly(aspartic acid) block copolymer through electrostatic interaction in aqueous medium. Their study showed that the PIC was expected to be useful as functional materials including carrier systems in drug delivery applications and a nanometric-scale reactor for enzymes. Furthermore, Wakebayashi et al. (2004) produced PIC in an aqueous solution by using an acetal-poly(ethylene glycol)-poly(2-(dimethylamino)ethyl methacrylate) (acetal-PEG-PAMA) block copolymer spontaneously associated with plasmid DNA (pDNA). They showed that the pDNA in the micelle was adequately protected from DNase I attack. The transfection ability of the PIC micelles toward 293 cells was remarkably enhanced with an increasing the residual molar mixing ratio as high as 25.

CNTs, large molecules with a cylindrical shape, have been known to not only carry small molecules but also large molecules such as protein (Elhissi et al., 2012; Zhang et al., 2011). The walls of the CNTs in graphene sheets have two types, single-walled and multi-walled. These sheets impact their size length. Despite their variation in physical length (hundred nanometers to micrometer), CNTs is also be used not only in noninvasive administration but also in invasive route (Yang et al., 2010). A large surface area of CNTs is important for their ability to conjugate with various molecules and also to penetrate the target area; thus, surface modification is usually performed for the effective delivery (Elhissi et al., 2012). Previous research has showed that bovine serum albumin (Huang et al., 2002), DNA (Singh et al., 2005), and other proteins can be bind to CNTs. Some of the proteins that have been immobilized onto CNTs through covalent linkages include chymotrypsin, ferritin (Lin et al., 2004), fibrinogen, hemoglobin (Wei et al., 2010), and streptavidin (Elhissi et al., 2012).

Protein-functionalized CNTs have shown their advantages in drug delivery, i.e., by having high pay loads, long release rates, retaining their biological function, and relatively easily enter the cells than free proteins (Nagaraju et al., 2015). Despite its beneficial properties, further research is needed to understand the safety aspect of CNTs before and after functionalization with proteins since there were some specific studies which showed cytotoxic effects of CNTs (Sun et al., 2011).


FUTURE PROSPECTIVE

The drawbacks of invasive delivery of protein drugs, like the inconvenience, high price, hydrophilic properties, and high molecular weight, have encouraged more studies to invent non-parenteral administration. Furthermore, the high inventions of protein-based drugs require more research on designing the appropriate delivery system and technologies. In the last decades, some advanced technologies have been developed by pharmaceutical companies to assist protein drug delivery through oral, transdermal, and pulmonary routes. With these new technologies, the advancement of protein drug delivery systems will increase shortly.

In addition to advancing the technology aspects, formulation modification was generated to improve drug carrying due to the limitation of technologies to tackle several barriers to protein drug delivery. Therefore, both technologies and formulation approaches for protein-based drugs complement each other and address patient questions and concerns in drug administration. Challenges in the future will be to find a better formulation or specific dosage form to obtain an effective therapy and safety.

Furthermore, in vitro and in vivo evaluation of protein or peptide drug delivery also play a prominent role in the development of effective therapy. To establish in vitro/in vivo correlations, it seems that more than one testing method should be applied. It is necessary to characterize the drug release and justify the system design according to the real condition. This limitation gives challenges to pharmaceutical scientists to develop a method that brings about a better understanding for evaluation of protein or peptide drug delivery and also which is feasible to be used routinely in the industry setting.

Taken together, noninvasive strategies for protein drug delivery will attract more attention from the public in the near future. New technologies, formulation modification, and better understanding for evaluations are required in order to develop a dosage form that is therapeutically viable in the market.


ACKNOWLEDGEMENTS

None


AUTHORS CONTRIBUTIONS

M.I.I.: designed the review, wrote the review, drafted and revised the manuscript. R.R. and A.E.A.: wrote the review and revised the manuscript. R.I.: designed the review, drafted and revised the critical content of the manuscript, and gave final approval before submission.


CONFLICT OF INTEREST

The authors declare no conflict of interest, financial or otherwise.


FUNDING

The authors received no financial support for the research, authorship, and/or publication of this article.


REFERENCES

Abhishek S, Vedamurthy J, Nagesh C, Rajeshwari AG. A review on solid lipid nanoparticles. Eur J Pharm Med Res, 2019; 6:229–34.

Agu RU, Ugwoke MI, Armand M, Kinget R, Verbeke N. The lung as a route for systemic delivery of therapeutic proteins and peptides. Respir Res, 2001; 2:198–209. CrossRef

Ahlén G, Strindelius L, Johansson T, Nilsson A, Chatzissavidou N, Sjöblom M, Rova U, Holgersson J. Mannosylated mucin-type immunoglobulin fusion proteins enhance antigen-specific antibody and T lymphocyte responses. PLoS One, 2012; 7:e46959. CrossRef

Akash MSH, Rehman K, Tariq M, Chen S. Development of therapeutic proteins: advances and challenges. Turk J Biol, 2015; 39:343–58. CrossRef

Aljuffali IA, Lin CF, Fang JY. Skin ablation by physical techniques for enhancing dermal/transdermal drug delivery. J Drug Deliv Sci Technol, 2014; 24:277–87. CrossRef

Alkilani AZ, McCrudden MTC, Donnelly RF. Transdermal drug delivery: innovative pharmaceutical developments based on disruption of the barrier properties of the stratum corneum. Pharmaceutics 2015; 7:438–70. CrossRef

AlQahtani AD, O’Connor D, Domling A, Goda SK. Strategies for the production of long-acting therapeutics and efficient drug delivery for cancer treatment. Biomed Pharmacother, 2019; 113:108750. CrossRef

Ameri M, Daddona PE, Maa YF. Demonstrated solid-state stability of parathyroid hormone PTH(1-34) coated on a novel transdermal microprojection delivery system. Pharm Res 2009; 26:2454–63. CrossRef

Anhalt H, Bohannon NJ. Insulin patch pumps: their development and future in closed-loop systems. Diabetes Technol Ther, 2010;12:51–8. CrossRef

Antunes F, Andrade F, Ferreira D, Mørck Nielsen H, Sarmento B. Models to predict intestinal absorption of therapeutic peptides and proteins. Curr Drug Metab, 2013; 14:4–20. CrossRef

Badkar AV, Smith AM, Eppstein JA, Banga AK. Transdermal delivery of interferon alpha-2b using microporation and iontophoresis in hairless rats. Pharm Res, 2007; 24:1389–95. CrossRef

Bailon P, Palleroni A, Schaffer CA, Spence CL, Fung WJ, Porter JE, Ehrlich GK, Pan W, Xu ZX, Modi MW, Farid A, Berthold W, Graves M. Rational design of a potent, long-lasting form of interferon: a 40 kDa branched polyethylene glycol-conjugated interferon α-2a for the treatment of hepatitis C. Bioconjug Chem, 2001; 12:195–202. CrossRef

Bajracharya R, Song JG, Back SY, Han HK. Recent advancements in non-invasive formulations for protein drug delivery. Comput Struct Biotechnol J, 2019; 17:1290–308. CrossRef

Banga AK. New technologies to allow transdermal delivery of therapeutic proteins and small water-soluble drugs. Am J Drug Deliv, 2006; 4:221–30. CrossRef

Banga AK. Pulmonary and other mucosal delivery of therapeutic peptides and proteins. Ther Pept Proteins, 2015; 29:301–38. CrossRef

Banga AK, Donnelly R, Stinchcomb AL. Transdermal drug delivery. Ther Deliv, 2013; 4:1235–8. CrossRef

Battaglia L, Ugazio E. Lipid nano- and microparticles: an overview of patent-related research. J Nanomater, 2019; 2019:1–22. CrossRef

Brown TD, Whitehead KA, Mitragotri S. Materials for oral delivery of proteins and peptides. Nat Rev, 2020; 5:127–8. CrossRef

Bruno BJ, Miller GD, Lim CS. Basics and recent advances in peptide and protein drug delivery. Ther Deliv, 2013; 4:1443–67. CrossRef

Campos EVR, De Oliveira JL, Da Silva CMG, Pascoli M, Pasquoto T, Lima R, Abhilash PC, Leonardo FF. Polymeric and solid lipid nanoparticles for sustained release of carbendazim and tebuconazole in agricultural applications. Sci Rep, 2015; 5:13809. CrossRef

Capelle MAH, Gurny R, Arvinte T. High throughput screening of protein formulation stability: practical considerations. Eur J Pharm Biopharm, 2007; 65:131–48. CrossRef

Carino GP, Mathiowitz E. Oral insulin delivery. Adv Drug Deliv Rev, 1999; 35:249–57. CrossRef

Chang LL, Pikal MJ. Mechanisms of protein stabilization in the solid state. J Pharm Sci, 2009; 98(9):2886–908. CrossRef

Chaulagain B, Jain A, Tiwari A, Verma A, Jain SK. Passive delivery of protein drugs through transdermal route. Artif Cells Nanomed Biotechnol, 2018; 46:472–87. CrossRef

Chen F, Stenzel MH. Polyion complex micelles for protein delivery. Aust J Chem, 2018; 71:768–80. CrossRef

Choi KY, Swierczewska M, Lee S, Chen X. Protease-activated drug development. Theranostics, 2012; 2:156–78. CrossRef

Choonara BF, Choonara YE, Kumar P, Bijukumar D, du Toit LC, Pillay V. A review of advanced oral drug delivery technologies facilitating the protection and absorption of protein and peptide molecules. Biotechnol Adv, 2014; 32:1269–82. CrossRef

Ciolkowski M, PaÅ‚ecz B, Appelhans D, Voit B, Klajnert B, Bryszewska M. The influence of maltose modified poly(propylene imine) dendrimers on hen egg white lysozyme structure and thermal stability. Colloids Surf B Biointerfaces, 2012; 95:103–8. CrossRef

Cleland JL, Daugherty A, Mrsny R. Emerging protein delivery methods. Curr Opin Biotechnol, 2001; 12:212–9. CrossRef

Clement S, Still JG, Kosutic G, McAllister RG. Oral insulin product hexyl-insulin monoconjugate 2 (HIM2) in type 1 diabetes mellitus: the glucose stabilization effects of HIM2. Diabetes Technol Ther, 2002; 4:459–66. CrossRef

Cortesi R, Esposito E, Luca G, Nastruzzi C. Production of lipospheres as carriers for bioactive compounds. Biomaterials, 2002; 23:2283–94. CrossRef

Craik DJ, Fairlie DP, Liras S, Price D. The future of peptide-based drugs. Chem Biol Drug Des, 2013; 81:136–47. CrossRef

Dai JX, You CH, Qi ZT, Wang XM, Sun PQ, Bi WS, Qian Y, Ding RL, Du P, He Y. Children’s respiratory viral diseases treated with interferon aerosol. Chin Med J (Engl), 1987; 100:162–6.

Dobson CM. Protein folding and misfolding. Nature, 2003; 426:884–90. CrossRef

Dolovich MB, Dhand R. Aerosol drug delivery: developments in device design and clinical use. Lancet, 2011; 377:1032–45. CrossRef

Eldor R, Arbit E, Corcos A, Kidron M. Glucose-reducing effect of the ORMD-0801 oral insulin preparation in patients with uncontrolled type 1 diabetes: a pilot study. PLoS One, 2013; 8:e59524. CrossRef

Elhissi AMA, Ahmed W, Hassan IU, Dhanak VR, D’Emanuele A. Carbon nanotubes in cancer therapy and drug delivery. J Drug Deliv, 2012; 2012:837327. CrossRef

Feridooni T, Hotchkiss A, Agu RU. Noninvasive strategies for systemic delivery of therapeutic proteins - prospects and challenges. Intech, London, UK, pp 197–218, 2016. CrossRef

Food and Drug Administration. Fortical calcitonin-salmon. FDA, Silver Spring, MD, 2005. Available via https://www.accessdata.fda.gov/drugsatfda_docs/label/2005/021406lbl.pdf (Accessed 10 August 2020).

Food and Drug Administration. Exubera. FDA, Silver Spring, MD, 2006. Available via https://www.accessdata.fda.gov/drugsatfda_docs/label/2006/021868lbl.pdf (Accessed 10 August 2020).

Food and Drug Administration. Miacalcin. FDA, Silver Spring, MD, 2011. Reference ID: 3118288. Available via https://www.accessdata.fda.gov/drugsatfda_docs/label/2012/020313s033lbl.pdf (Accessed 10 August 2020).

Food and Drug Administration. Afrezza Inhalation Powder. FDA, Silver Spring, MD, 2014. Reference ID: 3533688. Available via https://www.accessdata.fda.gov/drugsatfda_docs/label/2014/022472lbl.pdf (Accessed 10 August 2020).

Food and Drug Administration. Purple book: lists of licensed biological products with reference product exclusivity and biosimilarity or interchangeability evaluations. FDA, Silver Spring, MD, 2020.

Fosgerau K, Hoffmann T. Peptide therapeutics: current status and future directions. Drug Discov Today, 2015; 20:122–8. CrossRef

Fredenberg S, Wahlgren M, Reslow M, Axelsson A. The mechanisms of drug release in poly(lactic-co-glycolic acid)-based drug delivery systems - a review. Int J Pharm, 2011; 415:34–52. CrossRef

Frokjaer S, Otzen DE. Protein drug stability: a formulation challenge. Nat Rev Drug Discov, 2005; 4:298–306. CrossRef

Gangwar S, Pauletti GM, Wang B, Siahaan TJ, Stella VJ, Borchardt RT. Prodrug strategies to enhance the intestinal absorption of peptides. Drug Discov Today, 1997; 2:148–55. CrossRef

Goldberg M, Gomez-Orellana I. Challenges for the oral delivery of macromolecules. Nat Rev Drug Discov, 2003; 2:289–95. CrossRef

Grimaudo MA, Concheiro A, Alvarez-Lorenzo C. Nanogels for regenerative medicine. J Control Release, 2019; 313:148–60. CrossRef

Gujjar M, Banga AK. Iontophoretic and microneedle mediated transdermal delivery of glycopyrrolate. Pharmaceutics, 2014; 6:663–71. CrossRef

Hamman JH, Enslin GM, Kotzé AF. Oral delivery of peptide drugs: barriers and developments. BioDrugs, 2005; 19:165–77. CrossRef

Han FY, Thurecht KJ, Whittaker AK, Smith MT. Bioerodable PLGA-based microparticles for producing sustained-release drug formulations and strategies for improving drug loading. Front Pharmacol, 2016; 7:185. CrossRef

Hans ML, Lowman AM. Biodegradable nanoparticles for drug delivery and targeting. Curr Opin Solid State Mater Sci, 2002; 334:137–48.

Harada A, Kataoka K. Novel polyion complex micelles entrapping enzyme molecules in the core: preparation of narrowly-distributed micelles from lysozyme and poly(ethylene glycol)-poly(aspartic acid) block copolymer in aqueous medium. Macromolecules, 1998; 31:288–94.

He S, Liu Z, Xu D. Advance in oral delivery systems for therapeutic protein. J Drug Target, 2019; 27:283–91. CrossRef

Hess DR. Aerosol delivery devices in the treatment of asthma. Respir Care, 2008; 53:699–723.

Huang D, Huang Y, Li Z. Transdermal delivery of nucleic acid mediated by punching and electroporation. Methods Mol Biol, 2020; 2050:101–12. CrossRef

Huang W, Taylor S, Fu K, Lin Y, Zhang D, Hanks TW, Rao AM, Sun YP. Attaching proteins to carbon nanotubes via diimide-activated amidation. Nano Lett, 2002; 2:311–4. CrossRef

Ibrahim M, Garcia-Contreras L. Mechanisms of absorption and elimination of drugs administered by inhalation. Ther Deliv, 2013; 4:1027–45. CrossRef

Ibrahim M, Verma R, Garcia-Contreras L. Inhalation drug delivery devices: technology update. Med Devices Evid Res, 2015; 8:131–9. CrossRef

Iswandana R, Irianti MI, Oosterhuis D, Hofker HS, Merema MT, De Jager MH, Mutsaers HAM, Olinga P. Regional differences in human intestinal drug metabolism. Drug Metab Dispos, 2018; 46:1879–85. CrossRef

Jaffe HA, Buhl R, Mastrangeli A, Holroyd KJ, Saltini C, Czerski D, Jaffe HS, Kramer S, Sherwin S, Crystal RG. Organ specific cytokine therapy: local activation of mononuclear phagocytes by delivery of an aerosol of recombinant interferon-γ to the human lung. J Clin Invest, 1991; 88:297–302. CrossRef

Jawahar N, Meyyanathan S. Polymeric nanoparticles for drug delivery and targeting: a comprehensive review. Int J Health Allied Sci, 2012; 1:217–23. CrossRef

Jitendra, Sharma PK, Bansal S, Banik A. Noninvasive routes of proteins and peptides drug delivery. Indian J Pharm Sci, 2011; 73:367–75.

Kalluri H, Banga AK. Transdermal delivery of proteins. AAPS PharmSciTech, 2011; 12:431–41. CrossRef

Kamaly N, Yameen B, Wu J, Farokhzad OC. Degradable controlled-release polymers and polymeric nanoparticles: mechanisms of controlling drug release. Chem Rev, 2016; 116:2602–63. CrossRef

Kanikkannan N, Kandimalla K, Lamba S, Singh M. Structure-activity relationship of chemical penetration enhan-cers in transdermal drug delivery. Curr Med Chem, 2012; 7:593–608. CrossRef

Katsuma M, Watanabe S, Takemura S, Sako K, Sawada T, Masuda Y, Nakamura K, Fukui M, Connor AL, Wilding IR. Scintigraphic evaluation of a novel colon-targeted delivery system (CODESTM) in healthy volunteers. J Pharm Sci, 2004; 93:1287–99. CrossRef

Khan A, Yasir M, Asif M, Chauhan I, Singh AP, Sharma R, Singh P, Rai S. Iontophoretic drug delivery: history and applications. J Appl Pharm Sci, 2011; 1:11–24.

Kim S, Shi Y, Kim JY, Park K, Cheng JX. Overcoming the barriers in micellar drug delivery: loading efficiency, in vivo stability, and micelle-cell interaction. Expert Opin Drug Deliv, 2010; 7:49–62. CrossRef

Kumar MD, Baboota S, Ahuja A, Hasan S, Ali J. Recent advances in protein and peptide drug delivery systems. Curr Drug Deliv, 2007; 4:141–51. CrossRef

Kwon KC, Daniell H. Oral delivery of protein drugs bioencapsulated in plant cells. Mol Ther, 2016; 24:1342–50. CrossRef

Lai MC, Topp EM. Solid-state chemical stability of proteins and peptides. J Pharm Sci, 1999; 88:489–500. CrossRef

Lamprecht A, Ubrich N, Hombreiro Pérez M, Lehr CM, Hoffman M, Maincent P. Influences of process parameters on nanoparticle preparation performed by a double emulsion pressure homogenization technique. Int J Pharm, 2000; 196:177–82. CrossRef

Larrañeta E, Lutton REM, Woolfson AD, Donnelly RF. Microneedle arrays as transdermal and intradermal drug delivery systems: materials science, manufacture and commercial development. Mater Sci Eng R Rep, 2016; 104:1–32. CrossRef

Lee JW, Park J-H, Prausnitz MR. Dissolving microneedles for transdermal drug delivery. Biomaterials, 2008; 117:238–45. CrossRef

Li G, Badkar A, Kalluri H, Banga AK. Microchannels created by sugar and metal microneedles: characterization by microscopy, macromolecular flux and other techniques. J Pharm Sci, 2010; 99:1931–41. CrossRef

Li S, Patapoff TW, Overcashier D, Hsu C, Nguyen TH, Borchardt RT. Effects of reducing sugars on the chemical stability of human relaxin in the lyophilized state. J Pharm Sci, 1996; 85:873–7. CrossRef

Li X, Shang H, Wu W, Li S, Lin Z, Duan J, Xu L, Li J. Glucose-responsive micelles for controlled insulin release based on transformation from amphiphilic to double hydrophilic. J Nanosci Nanotechnol, 2016; 16:5457–63. CrossRef

Lin Y, Allard LF, Sun YP. Protein-affinity of single-walled carbon nanotubes in water. J Phys Chem B, 2004; 108:3760–4. CrossRef

Liu C, Wan T, Wang H, Zhang S, Ping Y, Cheng Y. A boronic acid–rich dendrimer with robust and unprecedented efficiency for cytosolic protein delivery and CRISPR-Cas9 gene editing. Sci Adv, 2019; 5:eaaw8922. CrossRef

Lu Y, Zhang E, Yang J, Cao Z. Strategies to improve micelle stability for drug delivery. Nano Res, 2018; 11:4985–98. CrossRef

Mahato RI, Narang AS, Thoma L, Miller DD. Emerging trends in oral delivery of peptide and protein drugs. Crit Rev Ther Drug Carrier Syst, 2003; 20:153–214. CrossRef

Maher S, Brayden DJ, Casettari L, Illum L. Application of permeation enhancers in oral delivery of macromolecules: an update. Pharmaceutics, 2019; 11:41. CrossRef

Makadia HK, Siegel SJ. Poly lactic-co-glycolic acid (PLGA) as biodegradable controlled drug delivery carrier. Polymers (Basel), 2011; 3:1377–97. CrossRef

Mao S, Xu J, Cai C, Germershaus O, Schaper A, Kissel T. Effect of WOW process parameters on morphology and burst release of FITC-dextran loaded PLGA microspheres. Int J Pharm, 2007; 334:137–48 CrossRef

Martins S, Sarmento B, Ferreira DC, Souto EB. Lipid-based colloidal carriers for peptide and protein delivery-Liposomes versus lipid nanoparticles. Int J Nanomedicine, 2007; 2:595–607.

Medi BM, Singh J. Electronically facilitated transdermal delivery of human parathyroid hormone (1-34). Int J Pharm, 2003; 263:25–33. CrossRef

Miller JF, Dower WJ, Tompkins LS. High-voltage electroporation of bacteria: genetic transformation of campylobacter jejuni with plasmid DNA. Proc Natl Acad Sci U S A, 1988; 85:856–60. CrossRef

Mnard S, Lebreton C, Schumann M, Matysiak-Budnik T, Dugave C, Bouhnik Y, Malamut G, Cellier C, Allez M, Crenn P, Schulzke JD, Cerf-Bensussan N, Heyman M. Paracellular versus transcellular intestinal permeability to gliadin peptides in active celiac disease. Am J Pathol, 2012; 180:608–15. CrossRef

Mohammad EA, Elshemey WM, Elsayed AA, Abd-Elghany AA. Electroporation parameters for successful transdermal delivery of insulin. Am J Ther, 2016; 23:1560–67. CrossRef

Nagaraju K, Reddy R, Reddy N. A review on protein functionalized carbon nanotubes. J Appl Biomater Funct Mater, 2015; 13:e301–12. CrossRef

Nakano R, Takagi-Maeda S, Ito Y, Kishimoto S, Osato T, Noguchi K, Kurihara-Suda K, Takahashi N. A new technology for increasing therapeutic protein levels in the brain over extended periods. PloS One, 2019; 14:e0214404. CrossRef

Noriega-Luna B, Godínez LA, Rodríguez FJ, Rodríguez A, Zaldívar-Lelo De Larrea G, Sosa-Ferreyra CF, Mercado-Curiel RF, Manríquez J, Bustos E. Applications of dendrimers in drug delivery agents, diagnosis, therapy, and detection. J Nanomater, 2014; 2014:1–19. CrossRef

Oleck J, Kassam S, Goldman JD. Commentary: why was inhaled insulin a failure in the market? Diabetes Spectr, 2016; 29:180–4. CrossRef

Pardridge WM. Blood-brain barrier drug delivery of IgG fusion proteins with a transferrin receptor monoclonal antibody. Expert Opin Drug Deliv, 2015; 12:207–22. CrossRef

Park K, Kwon IC, Park K. Oral protein delivery: current status and future prospect. React Funct Polym, 2011; 71:280–7. CrossRef

Parveen S, Sahoo SK. Nanomedicine: clinical applications of polyethylene glycol conjugated proteins and drugs. Clin Pharmacokinet, 2006; 45:965–88. CrossRef

Patel A, Cholkar K, Mitra AK. Recent developments in protein and peptide parenteral delivery approaches. Ther Deliv, 2014; 5:337–65. CrossRef

Pathan IB, Setty CM. Chemical penetration enhancers for transdermal drug delivery systems. Trop J Pharm Res, 2009; 8:173–9. CrossRef

Payne RW, Manning MC. Peptide formulation: challenges and strategies. Innov Pharm Technol, 2009;28:64–8.

Pisal DS, Kosloski MP, Balu-lyer SV. Delivery of theurapeutic proteins. J Pharm Sci, 2010; 99:2557–75. CrossRef

Ratnaparkhi MP, Chaudhari SP, Pandya VA. Peptides and proteins in pharmaceuticals. Int J Curr Pharm Res, 2011; 3:1–9.

Rubin BK, Williams RW. Emerging aerosol drug delivery strategies: from bench to clinic. Adv Drug Deliv Rev, 2014; 75:141–8. CrossRef

Ryan SM, Mantovani G, Wang X, Haddleton DM, Brayden DJ. Advances in PEGylation of important biotech molecules: delivery aspects. Expert Opin Drug Deliv, 2008; 5:371–83. CrossRef

Saez V, Hernández JR, Peniche C. Microspheres as delivery systems for the controlled release of peptides and proteins. Biotecnol Apl, 2007; 24:108–16.

Sandborn WJ, Travis S, Moro L, Jones R, Gautille T, Bagin R, Huang M, Yeung P, Ballard ED. Once-daily budesonide MMX® extended-release tablets induce remission in patients with mild to moderate ulcerative colitis: results from the CORE I study. Gastroenterology, 2012; 143:1218–26. CrossRef

Sauna ZE, Lagassé HAD, Alexaki A, Simhadri VL, Katagiri NH, Jankowski W, Kimchi-Sarfaty C. Recent advances in (therapeutic protein) drug development. F1000Res, 2017; 6:113. CrossRef

Scaramuzza S, Tonon G, Olianas A, Messana I, Schrepfer R, Orsini G, Caliceti P. A new site-specific monoPEGylated filgrastim derivative prepared by enzymatic conjugation: production and physicochemical characterization. J Control Release, 2012; 164:355–63. CrossRef

Schiffter HA. The delivery of drugs - peptides and proteins. In: Moo-Young M (ed.). Comprehensive biotechnology. 2nd edition, Elsevier, Amsterdam, Netherlands, p 5320, 2011. CrossRef

Selmin F, Puoci F, Parisi O, Franzé S, Musazzi U, Cilurzo F. Caffeic acid-PLGA conjugate to design protein drug delivery systems stable to irradiation. J Funct Biomater, 2015; 6:1–13. CrossRef

Shaker DS, Ishak RAH, Ghoneim A, Elhuoni MA. Nanoemulsion: a review on mechanisms for the transdermal delivery of hydrophobic and hydrophilic drugs. Sci Pharm, 2019; 87:17. CrossRef

Sheldon RA. Characteristic features and biotechnological applications of cross-linked enzyme aggregates (CLEAs). Appl Microbiol Biotechnol, 2011; 92:467–77. CrossRef

Shi Y, Li LC. Current advances in sustained-release systems for parenteral drug delivery. Expert Opin Drug Deliv, 2005; 2:1039–58. CrossRef

Singh R, Pantarotto D, McCarthy D, Chaloin O, Hoebeke J, Partidos CD, Briand J, Prato M, Bianco A, Kostarelos K. Binding and condensation of plasmid DNA onto functionalized carbon nanotubes: toward the construction of nanotube-based gene delivery vectors. J Am Chem Soc, 2005; 127:4388–96. CrossRef

Smith A, Hunneyball IM. Evaluation of poly(lactic acid) as a biodegradable drug delivery system for parenteral administration. Int J Pharm, 1986; 30:215–20. CrossRef

Solá RJ, Griebenow K. Glycosylation of theurapeutic proteins: an effective strategy to optimize efficacy. BioDrugs, 2010; 24:9–21. CrossRef

Stern W, Mehta N, Carl SM. Peptide delivery - oral delivery of peptides by peptelligence technology. Drug Dev Deliv, 2013; 143:1218–26.

Sun Z, Liu Z, Meng J, Meng J, Duan J, Xie S, Lu X, Zhu Z, Wang C, Chen S, Xu H, Yang X. Carbon nanotubes enhance cytotoxicity mediated by human lymphocytes in vitro. PloS One, 2011; 6:e21073. CrossRef

Szunerits S, Boukherroub R. Heat: a highly efficient skin enhancer for transdermal drug delivery. Front Bioeng Biotechnol, 2018; 6:15. CrossRef

Tomita M, Hayashi M, Awazu S. Absorption-enhancing mechanism of EDTA, caprate, and decanoylcarnitine in Caco-2 cells. J Pharm Sci, 1996; 85:608–11. CrossRef

Torosantucci R, Schöneich C, Jiskoot W. Oxidation of therapeutic proteins and peptides: structural and biological consequences. Pharm Res, 2014; 31:541–53. CrossRef

Ulapane KR, Kopec BM, Moral MEG, Siahaan TJ. Peptides and drug delivery. In: Sunna A, Care A, Bergquist P. (eds.). Peptides and peptide-based biomaterials and their biomedical applications. Advances in experimental medicine and biology 1030. Springer International Publishing, New York, NY, 2017. CrossRef

van Bree JBMM, de Boer AG, Verhoef J, Danhof M, Breimer DD. Peptide transport across the blood-brain barrier. J Control Release, 1990; 13:175–84. CrossRef

van De Weert M, Hennink WE, Jiskoot W. Protein instability in poly(lactic-co-glycolic acid) microparticles. Pharm Res, 2000; 17:1159–67. CrossRef

Verbaan FJ, Bal SM, van den Berg DJ, Groenink WHH, Verpoorten H, Lüttge R, Bouwstra JA. Assembled microneedle arrays enhance the transport of compounds varying over a large range of molecular weight across human dermatomed skin. J Control Release, 2007; 117:238–45. CrossRef

Victor S, Paul W, Sharma C. Eligen technology for oral delivery of proteins and peptides. In: das Neves J, Sarmento B (eds.). Mucosal delivery of biopharmaceuticals. Springer, Boston, MA, pp 407–22, 2014. CrossRef

Wagner AM, Gran MP, Peppas NA. Designing the new generation of intelligent biocompatible carriers for protein and peptide delivery. Acta Pharm Sin B, 2018; 8:147–64. CrossRef

Wakebayashi D, Nishiyama N, Itaka K, Miyata K, Yamasaki Y, Harada A, Koyama H, Nagasaki Y, Kataoka K. Polyion complex micelles of pDNA with acetal-poly(ethylene glycol)-poly(2-(dimethylamino)ethyl methacrylate) block copolymer as the gene carrier system: physicochemical properties of micelles relevant to gene transfection efficacy. Biomacromolecules, 2004; 5:2128–36. CrossRef

Watts AB, McConville JT, Williams RO. Current therapies and technological advances in aqueous aerosol drug delivery. Drug Dev Ind Pharm, 2008; 34:913–22. CrossRef

Wei G, Pan C, Reichert J, Jandt KD. Controlled assembly of protein-protected gold nanoparticles on noncovalent functionalized carbon nanotubes. Carbon, 2010; 48:645–53. CrossRef

Weiss WF 4th, Young TM, Roberts CJ. Principles, approaches, and challenges for predicting protein aggregation rates and shelf life. J Pharm Sci, 2009; 98:1246–77. CrossRef

Welte K, Ann Bonilla M, Gillio AP, Boone TC, Potter GK, Gabrilove JL, Moore MA, O'Reilly RJ, Souza LM. Recombinant human granulocyte colony-stimulating factor: effects on hematopoiesis in normal and cyclophosphamide-treated primates. J Exp Med, 1987; 165:941–8. CrossRef

Wissing SA, Kayser O, Müller RH. Solid lipid nanoparticles for parenteral drug delivery. Adv Drug Deliv Rev, 2004; 56:1257–72.

Yang Z, Zhang Y, Yang Y, Sun L, Han D, Li H, Wang C. Pharmacological and toxicological target organelles and safe use of single-walled carbon nanotubes as drug carriers in treating Alzheimer disease. Nanomedicine, 2010; 6:427–41.

Yih TC, Al-Fandi M. Engineered nanoparticles as precise drug delivery systems. J Cell Biochem, 2006; 97:1184–90.

Zhang H, Zhai Y, Wang J, Zhai G. New progress and prospects: the application of nanogel in drug delivery. Mater Sci Eng C Mater Biol Appl, 2016; 60:560–8.

Zhang W, Zhang Z, Zhang Y. The application of carbon nanotubes in target drug delivery systems for cancer therapies. Nanoscale Res Lett, 2011; 6:555.

Zhang Y, Huang Y, Li S. Polymeric micelles: Nanocarriers for cancer-targeted drug delivery. AAPS PharmSciTech, 2014; 15:862–71.

Zheng C, Liang W. A one-step modified method to reduce the burst initial release from PLGA microspheres. Drug Deliv, 2010; 17:77–82.

Zhu L, Lu L, Wang S, Wu J, Shi J, Yan T, Xie C, Li Q, Hu M, Liu Z. Oral absorption basics: pathways and physicochemical and biological factors affecting absorption. In: Qiu Y, Zhang GGZ, Mantri RV, Chen Y, Yu L (ed.). Developing solid oral dosage forms: pharmaceutical theory and practice: 2nd edition, Academic Press, Cambridge, MA, pp 297–329, 2017.

Reference

Abhishek S, Vedamurthy J, Nagesh C, Rajeshwari AG. A review on solid lipid nanoparticles. Eur J Pharm Med Res, 2019; 6:229-34.

Agu RU, Ugwoke MI, Armand M, Kinget R, Verbeke N. The lung as a route for systemic delivery of therapeutic proteins and peptides. Respir Res, 2001; 2:198-209. https://doi.org/10.1186/rr58

Ahlén G, Strindelius L, Johansson T, Nilsson A, Chatzissavidou N, Sjöblom M, Rova U, Holgersson J. Mannosylated mucin-type immunoglobulin fusion proteins enhance antigen-specific antibody and T lymphocyte responses. PLoS One, 2012; 7:e46959. https://doi.org/10.1371/journal.pone.0046959

Akash MSH, Rehman K, Tariq M, Chen S. Development of therapeutic proteins: advances and challenges. Turk J Biol, 2015; 39:343-58. https://doi.org/10.3906/biy-1411-8

Aljuffali IA, Lin CF, Fang JY. Skin ablation by physical techniques for enhancing dermal/transdermal drug delivery. J Drug Deliv Sci Technol, 2014; 24:277-87. https://doi.org/10.1016/S1773-2247(14)50046-9

Alkilani AZ, McCrudden MTC, Donnelly RF. Transdermal drug delivery: innovative pharmaceutical developments based on disruption of the barrier properties of the stratum corneum. Pharmaceutics 2015; 7:438-70. https://doi.org/10.3390/pharmaceutics7040438

AlQahtani AD, O'Connor D, Domling A, Goda SK. Strategies for the production of long-acting therapeutics and efficient drug delivery for cancer treatment. Biomed Pharmacother, 2019; 113:108750. https://doi.org/10.1016/j.biopha.2019.108750

Ameri M, Daddona PE, Maa YF. Demonstrated solid-state stability of parathyroid hormone PTH(1-34) coated on a novel transdermal microprojection delivery system. Pharm Res 2009; 26:2454-63. https://doi.org/10.1007/s11095-009-9960-9

Anhalt H, Bohannon NJ. Insulin patch pumps: their development and future in closed-loop systems. Diabetes Technol Ther, 2010;12:51-8. https://doi.org/10.1089/dia.2010.0016

Antunes F, Andrade F, Ferreira D, Mørck Nielsen H, Sarmento B. Models to predict intestinal absorption of therapeutic peptides and proteins. Curr Drug Metab, 2013; 14:4-20. https://doi.org/10.2174/138920013804545160

Badkar AV, Smith AM, Eppstein JA, Banga AK. Transdermal delivery of interferon alpha-2b using microporation and iontophoresis in hairless rats. Pharm Res, 2007; 24:1389-95. https://doi.org/10.1007/s11095-007-9308-2

Bailon P, Palleroni A, Schaffer CA, Spence CL, Fung WJ, Porter JE, Ehrlich GK, Pan W, Xu ZX, Modi MW, Farid A, Berthold W, Graves M. Rational design of a potent, long-lasting form of interferon: a 40 kDa branched polyethylene glycol-conjugated interferon α-2a for the treatment of hepatitis C. Bioconjug Chem, 2001; 12:195-202. https://doi.org/10.1021/bc000082g

Bajracharya R, Song JG, Back SY, Han HK. Recent advancements in non-invasive formulations for protein drug delivery. Comput Struct Biotechnol J, 2019; 17:1290-308. https://doi.org/10.1016/j.csbj.2019.09.004

Banga AK. New technologies to allow transdermal delivery of therapeutic proteins and small water-soluble drugs. Am J Drug Deliv, 2006; 4:221-30. https://doi.org/10.2165/00137696-200604040-00005

Banga AK. Pulmonary and other mucosal delivery of therapeutic peptides and proteins. Ther Pept Proteins, 2015; 29:301-38. https://doi.org/10.1201/b18392-12

Banga AK, Donnelly R, Stinchcomb AL. Transdermal drug delivery. Ther Deliv, 2013; 4:1235-8. https://doi.org/10.4155/tde.13.100

Battaglia L, Ugazio E. Lipid nano- and microparticles: an overview of patent-related research. J Nanomater, 2019; 2019:1-22. https://doi.org/10.1155/2019/2834941

Brown TD, Whitehead KA, Mitragotri S. Materials for oral delivery of proteins and peptides. Nat Rev, 2020; 5:127-8. https://doi.org/10.1038/s41578-019-0156-6

Bruno BJ, Miller GD, Lim CS. Basics and recent advances in peptide and protein drug delivery. Ther Deliv, 2013; 4:1443-67. https://doi.org/10.4155/tde.13.104

Campos EVR, De Oliveira JL, Da Silva CMG, Pascoli M, Pasquoto T, Lima R, Abhilash PC, Leonardo FF. Polymeric and solid lipid nanoparticles for sustained release of carbendazim and tebuconazole in agricultural applications. Sci Rep, 2015; 5:13809. https://doi.org/10.1038/srep13809

Capelle MAH, Gurny R, Arvinte T. High throughput screening of protein formulation stability: practical considerations. Eur J Pharm Biopharm, 2007; 65:131-48. https://doi.org/10.1016/j.ejpb.2006.09.009

Carino GP, Mathiowitz E. Oral insulin delivery. Adv Drug Deliv Rev, 1999; 35:249-57. https://doi.org/10.1016/S0169-409X(98)00075-1

Chang LL, Pikal MJ. Mechanisms of protein stabilization in the solid state. J Pharm Sci, 2009; 98(9):2886-908. https://doi.org/10.1002/jps.21825

Chaulagain B, Jain A, Tiwari A, Verma A, Jain SK. Passive delivery of protein drugs through transdermal route. Artif Cells Nanomed Biotechnol, 2018; 46:472-87. https://doi.org/10.1080/21691401.2018.1430695

Chen F, Stenzel MH. Polyion complex micelles for protein delivery. Aust J Chem, 2018; 71:768-80. https://doi.org/10.1071/CH18219

Choi KY, Swierczewska M, Lee S, Chen X. Protease-activated drug development. Theranostics, 2012; 2:156-78. https://doi.org/10.7150/thno.4068

Choonara BF, Choonara YE, Kumar P, Bijukumar D, du Toit LC, Pillay V. A review of advanced oral drug delivery technologies facilitating the protection and absorption of protein and peptide molecules. Biotechnol Adv, 2014; 32:1269-82. https://doi.org/10.1016/j.biotechadv.2014.07.006

Ciolkowski M, Pałecz B, Appelhans D, Voit B, Klajnert B, Bryszewska M. The influence of maltose modified poly(propylene imine) dendrimers on hen egg white lysozyme structure and thermal stability. Colloids Surf B Biointerfaces, 2012; 95:103-8. https://doi.org/10.1016/j.colsurfb.2012.02.021

Cleland JL, Daugherty A, Mrsny R. Emerging protein delivery methods. Curr Opin Biotechnol, 2001; 12:212-9. https://doi.org/10.1016/S0958-1669(00)00202-0

Clement S, Still JG, Kosutic G, McAllister RG. Oral insulin product hexyl-insulin monoconjugate 2 (HIM2) in type 1 diabetes mellitus: the glucose stabilization effects of HIM2. Diabetes Technol Ther, 2002; 4:459-66. https://doi.org/10.1089/152091502760306544

Cortesi R, Esposito E, Luca G, Nastruzzi C. Production of lipospheres as carriers for bioactive compounds. Biomaterials, 2002; 23:2283-94. https://doi.org/10.1016/S0142-9612(01)00362-3

Craik DJ, Fairlie DP, Liras S, Price D. The future of peptide-based drugs. Chem Biol Drug Des, 2013; 81:136-47. https://doi.org/10.1111/cbdd.12055

Dai JX, You CH, Qi ZT, Wang XM, Sun PQ, Bi WS, Qian Y, Ding RL, Du P, He Y. Children's respiratory viral diseases treated with interferon aerosol. Chin Med J (Engl), 1987; 100:162-6.

Dobson CM. Protein folding and misfolding. Nature, 2003; 426:884-90. https://doi.org/10.1038/nature02261

Dolovich MB, Dhand R. Aerosol drug delivery: developments in device design and clinical use. Lancet, 2011; 377:1032-45. https://doi.org/10.1016/S0140-6736(10)60926-9

Eldor R, Arbit E, Corcos A, Kidron M. Glucose-reducing effect of the ORMD-0801 oral insulin preparation in patients with uncontrolled type 1 diabetes: a pilot study. PLoS One, 2013; 8:e59524. https://doi.org/10.1371/journal.pone.0059524

Elhissi AMA, Ahmed W, Hassan IU, Dhanak VR, D'Emanuele A. Carbon nanotubes in cancer therapy and drug delivery. J Drug Deliv, 2012; 2012:837327. https://doi.org/10.1155/2012/837327

Feridooni T, Hotchkiss A, Agu RU. Noninvasive strategies for systemic delivery of therapeutic proteins - prospects and challenges. Intech, London, UK, pp 197-218, 2016. https://doi.org/10.5772/61266

Food and Drug Administration. Fortical calcitonin-salmon. FDA, Silver Spring, MD, 2005. Available via https://www.accessdata.fda. gov/drugsatfda_docs/label/2005/021406lbl.pdf (Accessed 10 August 2020).

Food and Drug Administration. Exubera. FDA, Silver Spring, MD, 2006. Available via https://www.accessdata.fda.gov/drugsatfda_docs/ label/2006/021868lbl.pdf (Accessed 10 August 2020).

Food and Drug Administration. Miacalcin. FDA, Silver Spring, MD, 2011. Reference ID: 3118288. Available via https://www.accessdata. fda.gov/drugsatfda_docs/label/2012/020313s033lbl.pdf (Accessed 10 August 2020).

Food and Drug Administration. Afrezza Inhalation Powder. FDA, Silver Spring, MD, 2014. Reference ID: 3533688. Available via https://www.accessdata.fda.gov/drugsatfda_docs/label/2014/022472lbl.pdf (Accessed 10 August 2020).

Food and Drug Administration. Purple book: lists of licensed biological products with reference product exclusivity and biosimilarity or interchangeability evaluations. FDA, Silver Spring, MD, 2020.

Fosgerau K, Hoffmann T. Peptide therapeutics: current status and future directions. Drug Discov Today, 2015; 20:122-8. https://doi.org/10.1016/j.drudis.2014.10.003

Fredenberg S, Wahlgren M, Reslow M, Axelsson A. The mechanisms of drug release in poly(lactic-co-glycolic acid)-based drug delivery systems - a review. Int J Pharm, 2011; 415:34-52. https://doi.org/10.1016/j.ijpharm.2011.05.049

Frokjaer S, Otzen DE. Protein drug stability: a formulation challenge. Nat Rev Drug Discov, 2005; 4:298-306. https://doi.org/10.1038/nrd1695

Gangwar S, Pauletti GM, Wang B, Siahaan TJ, Stella VJ, Borchardt RT. Prodrug strategies to enhance the intestinal absorption of peptides. Drug Discov Today, 1997; 2:148-55. https://doi.org/10.1016/S1359-6446(97)01011-8

Goldberg M, Gomez-Orellana I. Challenges for the oral delivery of macromolecules. Nat Rev Drug Discov, 2003; 2:289-95. https://doi.org/10.1038/nrd1067

Grimaudo MA, Concheiro A, Alvarez-Lorenzo C. Nanogels for regenerative medicine. J Control Release, 2019; 313:148-60. https://doi.org/10.1016/j.jconrel.2019.09.015

Gujjar M, Banga AK. Iontophoretic and microneedle mediated transdermal delivery of glycopyrrolate. Pharmaceutics, 2014; 6:663-71. https://doi.org/10.3390/pharmaceutics6040663

Hamman JH, Enslin GM, Kotzé AF. Oral delivery of peptide drugs: barriers and developments. BioDrugs, 2005; 19:165-77. https://doi.org/10.2165/00063030-200519030-00003

Han FY, Thurecht KJ, Whittaker AK, Smith MT. Bioerodable PLGA-based microparticles for producing sustained-release drug formulations and strategies for improving drug loading. Front Pharmacol, 2016; 7:185. https://doi.org/10.3389/fphar.2016.00185

Hans ML, Lowman AM. Biodegradable nanoparticles for drug delivery and targeting. Curr Opin Solid State Mater Sci, 2002; 334:137-48.

Harada A, Kataoka K. Novel polyion complex micelles entrapping enzyme molecules in the core: preparation of narrowly-distributed micelles from lysozyme and poly(ethylene glycol)-poly(aspartic 177

acid) block copolymer in aqueous medium. Macromolecules, 1998; 31:288-94. https://doi.org/10.1021/ma971277v

He S, Liu Z, Xu D. Advance in oral delivery systems for therapeutic protein. J Drug Target, 2019; 27:283-91. https://doi.org/10.1080/1061186X.2018.1486406

Hess DR. Aerosol delivery devices in the treatment of asthma. Respir Care, 2008; 53:699-723.

Huang D, Huang Y, Li Z. Transdermal delivery of nucleic acid mediated by punching and electroporation. Methods Mol Biol, 2020; 2050:101-12. https://doi.org/10.1007/978-1-4939-9740-4_11

Huang W, Taylor S, Fu K, Lin Y, Zhang D, Hanks TW, Rao AM, Sun YP. Attaching proteins to carbon nanotubes via diimide-activated amidation. Nano Lett, 2002; 2:311-4. https://doi.org/10.1021/nl010095i

Ibrahim M, Garcia-Contreras L. Mechanisms of absorption and elimination of drugs administered by inhalation. Ther Deliv, 2013; 4:1027-45. https://doi.org/10.4155/tde.13.67

Ibrahim M, Verma R, Garcia-Contreras L. Inhalation drug delivery devices: technology update. Med Devices Evid Res, 2015; 8:131-9. https://doi.org/10.2147/MDER.S48888

Iswandana R, Irianti MI, Oosterhuis D, Hofker HS, Merema MT, De Jager MH, Mutsaers HAM, Olinga P. Regional differences in human intestinal drug metabolism. Drug Metab Dispos, 2018; 46:1879-85. https://doi.org/10.1124/dmd.118.083428

Jaffe HA, Buhl R, Mastrangeli A, Holroyd KJ, Saltini C, Czerski D, Jaffe HS, Kramer S, Sherwin S, Crystal RG. Organ specific cytokine therapy: local activation of mononuclear phagocytes by delivery of an aerosol of recombinant interferon-γ to the human lung. J Clin Invest, 1991; 88:297-302. https://doi.org/10.1172/JCI115291

Jawahar N, Meyyanathan S. Polymeric nanoparticles for drug delivery and targeting: a comprehensive review. Int J Health Allied Sci, 2012; 1:217-23. https://doi.org/10.4103/2278-344X.107832

Jitendra, Sharma PK, Bansal S, Banik A. Noninvasive routes of proteins and peptides drug delivery. Indian J Pharm Sci, 2011; 73:367-75.

Kalluri H, Banga AK. Transdermal delivery of proteins. AAPS PharmSciTech, 2011; 12:431-41. https://doi.org/10.1208/s12249-011-9601-6

Kamaly N, Yameen B, Wu J, Farokhzad OC. Degradable controlled-release polymers and polymeric nanoparticles: mechanisms of controlling drug release. Chem Rev, 2016; 116:2602-63. https://doi.org/10.1021/acs.chemrev.5b00346

Kanikkannan N, Kandimalla K, Lamba S, Singh M. Structure-activity relationship of chemical penetration enhan-cers in transdermal drug delivery. Curr Med Chem, 2012; 7:593-608. https://doi.org/10.2174/0929867003374840

Katsuma M, Watanabe S, Takemura S, Sako K, Sawada T, Masuda Y, Nakamura K, Fukui M, Connor AL, Wilding IR. Scintigraphic evaluation of a novel colon-targeted delivery system (CODESTM) in healthy volunteers. J Pharm Sci, 2004; 93:1287-99. https://doi.org/10.1002/jps.20063

Khan A, Yasir M, Asif M, Chauhan I, Singh AP, Sharma R, Singh P, Rai S. Iontophoretic drug delivery: history and applications. J Appl Pharm Sci, 2011; 1:11-24.

Kim S, Shi Y, Kim JY, Park K, Cheng JX. Overcoming the barriers in micellar drug delivery: loading efficiency, in vivo stability, and micelle-cell interaction. Expert Opin Drug Deliv, 2010; 7:49-62. https://doi.org/10.1517/17425240903380446

Kumar MD, Baboota S, Ahuja A, Hasan S, Ali J. Recent advances in protein and peptide drug delivery systems. Curr Drug Deliv, 2007; 4:141-51. https://doi.org/10.2174/156720107780362339

Kwon KC, Daniell H. Oral delivery of protein drugs bioencapsulated in plant cells. Mol Ther, 2016; 24:1342-50. https://doi.org/10.1038/mt.2016.115

Lai MC, Topp EM. Solid-state chemical stability of proteins and peptides. J Pharm Sci, 1999; 88:489-500. https://doi.org/10.1021/js980374e

Lamprecht A, Ubrich N, Hombreiro Pérez M, Lehr CM, Hoffman M, Maincent P. Influences of process parameters on nanoparticle preparation performed by a double emulsion pressure homogenization technique. Int J Pharm, 2000; 196:177-82. https://doi.org/10.1016/S0378-5173(99)00422-6

Larrañeta E, Lutton REM, Woolfson AD, Donnelly RF. Microneedle arrays as transdermal and intradermal drug delivery systems: materials science, manufacture and commercial development. Mater Sci Eng R Rep, 2016; 104:1-32. https://doi.org/10.1016/j.mser.2016.03.001

Lee JW, Park J-H, Prausnitz MR. Dissolving microneedles for transdermal drug delivery. Biomaterials, 2008; 117:238-45. https://doi.org/10.1016/j.biomaterials.2007.12.048

Li G, Badkar A, Kalluri H, Banga AK. Microchannels created by sugar and metal microneedles: characterization by microscopy, macromolecular flux and other techniques. J Pharm Sci, 2010; 99:1931-41. https://doi.org/10.1002/jps.21981

Li S, Patapoff TW, Overcashier D, Hsu C, Nguyen TH, Borchardt RT. Effects of reducing sugars on the chemical stability of human relaxin in the lyophilized state. J Pharm Sci, 1996; 85:873-7. https://doi.org/10.1021/js950456s

Li X, Shang H, Wu W, Li S, Lin Z, Duan J, Xu L, Li J. Glucose-responsive micelles for controlled insulin release based on transformation from amphiphilic to double hydrophilic. J Nanosci Nanotechnol, 2016; 16:5457-63. https://doi.org/10.1166/jnn.2016.11655

Lin Y, Allard LF, Sun YP. Protein-affinity of single-walled carbon nanotubes in water. J Phys Chem B, 2004; 108:3760-4. https://doi.org/10.1021/jp031248o

Liu C, Wan T, Wang H, Zhang S, Ping Y, Cheng Y. A boronic acid-rich dendrimer with robust and unprecedented efficiency for cytosolic protein delivery and CRISPR-Cas9 gene editing. Sci Adv, 2019; 5:eaaw8922. https://doi.org/10.1126/sciadv.aaw8922

Lu Y, Zhang E, Yang J, Cao Z. Strategies to improve micelle stability for drug delivery. Nano Res, 2018; 11:4985-98. https://doi.org/10.1007/s12274-018-2152-3

Mahato RI, Narang AS, Thoma L, Miller DD. Emerging trends in oral delivery of peptide and protein drugs. Crit Rev Ther Drug Carrier Syst, 2003; 20:153-214. https://doi.org/10.1615/CritRevTherDrugCarrierSyst.v20.i23.30

Maher S, Brayden DJ, Casettari L, Illum L. Application of permeation enhancers in oral delivery of macromolecules: an update. Pharmaceutics, 2019; 11:41. https://doi.org/10.3390/pharmaceutics11010041

Makadia HK, Siegel SJ. Poly lactic-co-glycolic acid (PLGA) as biodegradable controlled drug delivery carrier. Polymers (Basel), 2011; 3:1377-97. https://doi.org/10.3390/polym3031377

Mao S, Xu J, Cai C, Germershaus O, Schaper A, Kissel T. Effect of WOW process parameters on morphology and burst release of FITC-dextran loaded PLGA microspheres. Int J Pharm, 2007; 334: 137-48 https://doi.org/10.1016/j.ijpharm.2006.10.036

Martins S, Sarmento B, Ferreira DC, Souto EB. Lipid-based colloidal carriers for peptide and protein delivery-Liposomes versus lipid nanoparticles. Int J Nanomedicine, 2007; 2:595-607.

Medi BM, Singh J. Electronically facilitated transdermal delivery of human parathyroid hormone (1-34). Int J Pharm, 2003; 263:25-33. https://doi.org/10.1016/S0378-5173(03)00337-5

Miller JF, Dower WJ, Tompkins LS. High-voltage electroporation of bacteria: genetic transformation of campylobacter jejuni with plasmid DNA. Proc Natl Acad Sci U S A, 1988; 85:856-60. https://doi.org/10.1073/pnas.85.3.856

Mnard S, Lebreton C, Schumann M, Matysiak-Budnik T, Dugave C, Bouhnik Y, Malamut G, Cellier C, Allez M, Crenn P, Schulzke JD, Cerf-Bensussan N, Heyman M. Paracellular versus transcellular intestinal permeability to gliadin peptides in active celiac disease. Am J Pathol, 2012; 180:608-15. https://doi.org/10.1016/j.ajpath.2011.10.019

Mohammad EA, Elshemey WM, Elsayed AA, Abd-Elghany AA. Electroporation parameters for successful transdermal delivery of insulin. Am J Ther, 2016; 23:1560-67. https://doi.org/10.1097/MJT.0000000000000198

Nagaraju K, Reddy R, Reddy N. A review on protein functionalized carbon nanotubes. J Appl Biomater Funct Mater, 2015; 13:e301-12. https://doi.org/10.5301/jabfm.5000231

Nakano R, Takagi-Maeda S, Ito Y, Kishimoto S, Osato T, Noguchi K, Kurihara-Suda K, Takahashi N. A new technology for increasing therapeutic protein levels in the brain over extended periods. PloS One, 2019; 14:e0214404. https://doi.org/10.1371/journal.pone.0214404

Noriega-Luna B, Godínez LA, Rodríguez FJ, Rodríguez A, Zaldívar-Lelo De Larrea G, Sosa-Ferreyra CF, Mercado-Curiel RF, Manríquez J, Bustos E. Applications of dendrimers in drug delivery agents, diagnosis, therapy, and detection. J Nanomater, 2014; 2014:1-19. https://doi.org/10.1155/2014/507273

Oleck J, Kassam S, Goldman JD. Commentary: why was inhaled insulin a failure in the market? Diabetes Spectr, 2016; 29:180-4. https://doi.org/10.2337/diaspect.29.3.180

Pardridge WM. Blood-brain barrier drug delivery of IgG fusion proteins with a transferrin receptor monoclonal antibody. Expert Opin Drug Deliv, 2015; 12:207-22. https://doi.org/10.1517/17425247.2014.952627

Park K, Kwon IC, Park K. Oral protein delivery: current status and future prospect. React Funct Polym, 2011; 71:280-7. https://doi.org/10.1016/j.reactfunctpolym.2010.10.002

Parveen S, Sahoo SK. Nanomedicine: clinical applications of polyethylene glycol conjugated proteins and drugs. Clin Pharmacokinet, 2006; 45:965-88. https://doi.org/10.2165/00003088-200645100-00002

Patel A, Cholkar K, Mitra AK. Recent developments in protein and peptide parenteral delivery approaches. Ther Deliv, 2014; 5:337-65. https://doi.org/10.4155/tde.14.5

Pathan IB, Setty CM. Chemical penetration enhancers for transdermal drug delivery systems. Trop J Pharm Res, 2009; 8:173-9.178 https://doi.org/10.4314/tjpr.v8i2.44527

Payne RW, Manning MC. Peptide formulation: challenges and strategies. Innov Pharm Technol, 2009;28:64-8.

Pisal DS, Kosloski MP, Balu-lyer SV. Delivery of theurapeutic proteins. J Pharm Sci, 2010; 99:2557-75. https://doi.org/10.1002/jps.22054

Ratnaparkhi MP, Chaudhari SP, Pandya VA. Peptides and proteins in pharmaceuticals. Int J Curr Pharm Res, 2011; 3:1-9.

Rubin BK, Williams RW. Emerging aerosol drug delivery strategies: from bench to clinic. Adv Drug Deliv Rev, 2014; 75:141-8. https://doi.org/10.1016/j.addr.2014.06.008

Ryan SM, Mantovani G, Wang X, Haddleton DM, Brayden DJ. Advances in PEGylation of important biotech molecules: delivery aspects. Expert Opin Drug Deliv, 2008; 5:371-83. https://doi.org/10.1517/17425247.5.4.371

Saez V, Hernández JR, Peniche C. Microspheres as delivery systems for the controlled release of peptides and proteins. Biotecnol Apl, 2007; 24:108-16.

Sandborn WJ, Travis S, Moro L, Jones R, Gautille T, Bagin R, Huang M, Yeung P, Ballard ED. Once-daily budesonide MMX® extended-release tablets induce remission in patients with mild to moderate ulcerative colitis: results from the CORE I study. Gastroenterology, 2012; 143:1218-26. https://doi.org/10.1053/j.gastro.2012.08.003

Sauna ZE, Lagassé HAD, Alexaki A, Simhadri VL, Katagiri NH, Jankowski W, Kimchi-Sarfaty C. Recent advances in (therapeutic protein) drug development. F1000Res, 2017; 6:113. https://doi.org/10.12688/f1000research.9970.1

Scaramuzza S, Tonon G, Olianas A, Messana I, Schrepfer R, Orsini G, Caliceti P. A new site-specific monoPEGylated filgrastim derivative prepared by enzymatic conjugation: production and physicochemical characterization. J Control Release, 2012; 164:355-63. https://doi.org/10.1016/j.jconrel.2012.06.026

Schiffter HA. The delivery of drugs - peptides and proteins. In: Moo-Young M (ed.). Comprehensive biotechnology. 2nd edition, Elsevier, Amsterdam, Netherlands, p 5320, 2011. https://doi.org/10.1016/B978-0-08-088504-9.00338-X

Selmin F, Puoci F, Parisi O, Franzé S, Musazzi U, Cilurzo F. Caffeic acid-PLGA conjugate to design protein drug delivery systems stable to irradiation. J Funct Biomater, 2015; 6:1-13. https://doi.org/10.3390/jfb6010001

Shaker DS, Ishak RAH, Ghoneim A, Elhuoni MA. Nanoemulsion: a review on mechanisms for the transdermal delivery of hydrophobic and hydrophilic drugs. Sci Pharm, 2019; 87:17. https://doi.org/10.3390/scipharm87030017

Sheldon RA. Characteristic features and biotechnological applications of cross-linked enzyme aggregates (CLEAs). Appl Microbiol Biotechnol, 2011; 92:467-77. https://doi.org/10.1007/s00253-011-3554-2

Shi Y, Li LC. Current advances in sustained-release systems for parenteral drug delivery. Expert Opin Drug Deliv, 2005; 2:1039-58. https://doi.org/10.1517/17425247.2.6.1039

Singh R, Pantarotto D, McCarthy D, Chaloin O, Hoebeke J, Partidos CD, Briand J, Prato M, Bianco A, Kostarelos K. Binding and condensation of plasmid DNA onto functionalized carbon nanotubes: toward the construction of nanotube-based gene delivery vectors. J Am Chem Soc, 2005; 127:4388-96. https://doi.org/10.1021/ja0441561

Smith A, Hunneyball IM. Evaluation of poly(lactic acid) as a biodegradable drug delivery system for parenteral administration. Int J Pharm, 1986; 30:215-20. https://doi.org/10.1016/0378-5173(86)90081-5

Solá RJ, Griebenow K. Glycosylation of theurapeutic proteins: an effective strategy to optimize efficacy. BioDrugs, 2010; 24:9-21. https://doi.org/10.2165/11530550-000000000-00000

Stern W, Mehta N, Carl SM. Peptide delivery - oral delivery of peptides by peptelligence technology. Drug Dev Deliv, 2013; 143:1218-26.

Sun Z, Liu Z, Meng J, Meng J, Duan J, Xie S, Lu X, Zhu Z, Wang C, Chen S, Xu H, Yang X. Carbon nanotubes enhance cytotoxicity mediated by human lymphocytes in vitro. PloS One, 2011; 6:e21073. https://doi.org/10.1371/journal.pone.0021073

Szunerits S, Boukherroub R. Heat: a highly efficient skin enhancer for transdermal drug delivery. Front Bioeng Biotechnol, 2018; 6:15. https://doi.org/10.3389/fbioe.2018.00015

Tomita M, Hayashi M, Awazu S. Absorption-enhancing mechanism of EDTA, caprate, and decanoylcarnitine in Caco-2 cells. J Pharm Sci, 1996; 85:608-11. https://doi.org/10.1021/js9504604

Torosantucci R, Schöneich C, Jiskoot W. Oxidation of therapeutic proteins and peptides: structural and biological consequences. Pharm Res, 2014; 31:541-53. https://doi.org/10.1007/s11095-013-1199-9

Ulapane KR, Kopec BM, Moral MEG, Siahaan TJ. Peptides and drug delivery. In: Sunna A, Care A, Bergquist P. (eds.). Peptides and peptide-based biomaterials and their biomedical applications. Advances in experimental medicine and biology 1030. Springer International Publishing, New York, NY, 2017. https://doi.org/10.1007/978-3-319-66095-0_8

van Bree JBMM, de Boer AG, Verhoef J, Danhof M, Breimer DD. Peptide transport across the blood-brain barrier. J Control Release, 1990; 13:175-84. https://doi.org/10.1016/0168-3659(90)90008-H

van De Weert M, Hennink WE, Jiskoot W. Protein instability in poly(lactic-co-glycolic acid) microparticles. Pharm Res, 2000; 17:1159-67. https://doi.org/10.1023/A:1026498209874

Verbaan FJ, Bal SM, van den Berg DJ, Groenink WHH, Verpoorten H, Lüttge R, Bouwstra JA. Assembled microneedle arrays enhance the transport of compounds varying over a large range of molecular weight across human dermatomed skin. J Control Release, 2007; 117:238-45. https://doi.org/10.1016/j.jconrel.2006.11.009

Victor S, Paul W, Sharma C. Eligen technology for oral delivery of proteins and peptides. In: das Neves J, Sarmento B (eds.). Mucosal delivery of biopharmaceuticals. Springer, Boston, MA, pp 407-22, 2014. https://doi.org/10.1007/978-1-4614-9524-6_18

Wagner AM, Gran MP, Peppas NA. Designing the new generation of intelligent biocompatible carriers for protein and peptide delivery. Acta Pharm Sin B, 2018; 8:147-64. https://doi.org/10.1016/j.apsb.2018.01.013

Wakebayashi D, Nishiyama N, Itaka K, Miyata K, Yamasaki Y, Harada A, Koyama H, Nagasaki Y, Kataoka K. Polyion complex micelles of pDNA with acetal-poly(ethylene glycol)-poly(2-(dimethylamino) ethyl methacrylate) block copolymer as the gene carrier system: physicochemical properties of micelles relevant to gene transfection efficacy. Biomacromolecules, 2004; 5:2128-36. https://doi.org/10.1021/bm040009j

Watts AB, McConville JT, Williams RO. Current therapies and technological advances in aqueous aerosol drug delivery. Drug Dev Ind Pharm, 2008; 34:913-22. https://doi.org/10.1080/03639040802144211

Wei G, Pan C, Reichert J, Jandt KD. Controlled assembly of protein-protected gold nanoparticles on noncovalent functionalized carbon nanotubes. Carbon, 2010; 48:645-53. https://doi.org/10.1016/j.carbon.2009.10.006

Weiss WF 4th, Young TM, Roberts CJ. Principles, approaches, and challenges for predicting protein aggregation rates and shelf life. J Pharm Sci, 2009; 98:1246-77. https://doi.org/10.1002/jps.21521

Welte K, Ann Bonilla M, Gillio AP, Boone TC, Potter GK, Gabrilove JL, Moore MA, O'Reilly RJ, Souza LM. Recombinant human granulocyte colony-stimulating factor: effects on hematopoiesis in normal and cyclophosphamide-treated primates. J Exp Med, 1987; 165:941-8. https://doi.org/10.1084/jem.165.4.941&

Article Metrics
751 Views 91 Downloads 842 Total

Year

Month

Related Search

By author names