UK 5099

Transdermal Drug Delivery for Hair Regrowth

Guang Yang,* Guojun Chen, and Zhen Gu*

 
ACCESS
Metrics & More
Article Recommendations

 

 

⦁ INTRODUCTION
Hair loss is a common autoimmunity disorder due to multiple reasons, such as aging, diseases, and medications.1−3 According to a recent survey, about 50% of men and 15−30% of women suffer from hair loss.5 Depending on the causes and symptoms, hair loss can be categorized as androgenetic alopecia (AGA), alopecia areata (AA), and other types.1,2,4 AGA, which is also called androgenic alopecia or male pattern baldness, is the most common type of hair loss. Generally, AGA occurs with varying scalp location of hair loss and severity. Typically, the hair loss region in men has a characteristic “horseshoe” pattern. AA is characterized by hair loss due to inflammatory responses in the hair follicles and happens indiscriminately between the sexes with the initial onset age before 30 years old in most cases. Disfiguration in hair brings poor psychological conditions with reduced self-esteem to the sufferers, emphasiz- ing urgent demands for efficient therapies. Until now, the exact development mechanism of AGA is poorly understood, but studies have demonstrated that genetics and immunity are the main contributors to the disease. Readers can refer to recent reviews regarding the link between genetics, the immune response, and other external factors, which result in this disease.6 Generally, the human hair follicle goes through a continuous cyclical round, including anagen, catagen, and telogen.7,8 Anagen is the rapid growth phase, which features keratinocyte proliferation in the follicular epithelium. In the catagen phase, the follicle undergoes involution due to cell apoptosis. Thereafter, the follicle enters the telogen phase, and no hair growth happens.9 To re-enter anagen, a signal that probably comes from the dermal papilla triggers hair follicle
stem cells within the bulge to proliferate and repopulate the follicular epithelium.7,10−13 If the activating signal fails, an alteration in the hair cycle could occur.14 The resting follicles transfer to the kenogen phase, instead of reentering anagen. During this phase, the original hair shaft falls off and no new hair grows, eventually leading to a bald appearance.15
The current management strategies for hair restoration mainly focus on stimulating existing follicles to grow new hair or recreating a fuller head of hair via transplantation.16 Drug therapies and autologous hair transplantation are the two most frequently used treatments for hair loss. Clinically approved drug therapies for hair loss are topical minoxidil and oral finasteride.17,18 Topical formulation of minoxidil is now the first-line treatment of patterned hair loss. Its primary therapeutic effect lies in an increase in the diameter of existing hairs with a less significant increase in hair count. The defined action mechanism of topical minoxidil is still unclear, but it is speculated that minoxidil might act as a potassium channel opener, which relaxes vascular smooth muscle and increases blood flow. The treatment of topical minoxidil is usually associated with side effects, such as skin irritation and unwanted hair growth elsewhere. Finasteride is an inhibitor of 5α-reductase. 5α-Reductase can enzymatically convert

Special Issue: Nanomedicines Beyond Cancer
Received: January 13, 2020
Revised: April 18, 2020
Accepted: April 21, 2020

 

Downloaded via UPPSALA UNIV on May 21, 2020 at 06:33:49 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
© XXXX American Chemical Society
A

https://dx.doi.org/10.1021/acs.molpharmaceut.0c00041

Mol. Pharmaceutics XXXX, XXX, XXX−XXX

testosterone into dihydrotestosterone, which is a hormone related to the miniaturization of hair follicles, giving the main reason for androgenetic alopecia, the most common type of alopecia. The clinical use of oral finasteride usually leads to sexual side effects such as fertility problems.19,20 Notably, both drugs might result in accelerated hair loss when the patients stop the treatments. In terms of hair transplantation, hair follicles from a nonbalding region of the scalp are excised and reimplanted in the balding area to create the illusion of a fuller head of hair.21 However, this treatment is limited by the amount of donor follicles and induces the donor site scar. Additionally, this therapy requires a costly yet time-consuming procedure with only a temporal therapeutic effect.
In view of this, efforts have been made to develop new therapeutic strategies for hair loss.22,23 One is the development of novel drugs with minimized adverse effects and enhanced treatment efficacy by understanding the pathophysiology of hair loss. Several studies revealed that the JAK-STAT signaling pathway plays a major role in initiation and progression of AA. This stimulates the intensive clinical studies on JAK inhibitors, especially for those already approved by Food and Drug Administration (FDA) for the treatment of other diseases, such as myelodysplastic disorders (ruxolitinib, a JAK1/2 inhibitor) and rheumatoid arthritis (tofacitinib, a pan-JAK inhibitor).24,25 Recent research demonstrated that hair follicle stem cells (HFSCs) utilized glycolytic metabolism to generate more lactate, which seemed critical for the activation of HFSCs.

Figure 1. Drug therapy in triggering hair regrowth. The resting follicles transfer from the telogen phase to the anagen via drug therapy, instead of the kenogen phase, which eventually leads to hair loss.
⦁ NANOFORMULATIONS FOR TOPICAL ADMINISTRATION
Development of safe and efficient delivery vehicles is one of the key routes to realize a desirable therapeutic efficacy of drugs. To this end, nanosystems such as lipid nanoparticles, liposomes, and polymeric nanoparticles have been established for hair loss treatment32 (Figure 2). Drugs formulated into

Small molecule UK5099, a potent inhibitor of the mitochon-
drial pyruvate transporter, was proven to induce the hair cycle by increasing lactate production.26 Small molecules α- ketoglutarate and α-ketobutyrate were demonstrated to activate autophagy and reactivate dormant hair follicle stem cells to re-enter new anagen hair growth.27 In addition, a recent study reported that Cas9/sgRNA riboprotein can suppress SRD5A2 protein production via genomic editing in dermal papilla cells in the hair follicle and induce hair regrowth.28 The other route for hair loss therapy concentrates on engineering drug delivery systems to enhance the therapeutic effects while reducing the side effects. Many studies have demonstrated that topical administration can increase local drug concentrations at the target skin region, thus reducing the systemic adverse effects associated with the oral treatment such as hepatotox- icity, hypotension, and gastrointestinal complaints. The therapeutic efficacy of the topical medications largely depends on the transdermal efficiency.29 Stratum corneum, as the outer

skin, is the rate-limiting barrier hindering drug penetration. In general, the drugs with proper molecular characteristics, such as a small molecular mass of 400 to 500 Da and a balanced lipophilicity with an octanol−water partition coefficient of around 2 to 3 are expected to facilitate skin permeability.30 Other drugs that do not possess these physicochemical properties usually have a low transdermal efficiency and thus require a suitable avenue to penetrate the skin. In light of this, this review focuses on the advances of the transdermal drug delivery strategies for hair regrowth reported. Drug delivery systems may interact with skin lipids, facilitate drugs transport to the hair follicles, and act as drug depots for a sustained release.31 These drug delivery systems could supply an effective tool for drugs to achieve a desirable therapy effect (Figure 1). In addition, some physical or chemical penetration enhancers are also introduced, which are often applied in combination with the drug delivery systems to achieve a synergistic effect.
Figure 2. Schematic of representative drug delivery nanoformulations, including solid lipid nanoparticle, nanostructured lipid carrier, liposome, and polymeric nanoparticle, for treatment of diseases associated with hair follicles, due to their natural tendency to accumulate in the follicle regions.

 

nanosystems can facilitate sustained and controllable release and reduce solvent irritation problems related to the conventional topical formulation.33,34 Notably, drug nano- systems are attractive for treating diseases associated with hair follicles because of their inherent tendency to accumulate in the follicle regions, thus producing a high local drug concentration and minimizing side effects.30 Herein, we introduce the progress of nanotechnology in drug formulations for the treatment of hair loss (Figure 2).
⦁ Lipid Nanoparticles. Lipid nanoparticles have been extensively investigated as a targeted transdermal drug delivery
B https://dx.doi.org/10.1021/acs.molpharmaceut.0c00041

carrier to the hair follicles.35,36 In general, nanosystems provide a better skin delivery ability over topical formulations due to the interaction between the particles and the pilosebaceous unit of the follicles.37 For the nanoparticle design, the particle size is often a key parameter. Generally, smaller nanoparticles (<100 nm) are considered to have the ability to facilitate transport to greater depths in the hair follicles compared to larger particles. Besides the particle size, other physicochemical characteristics of lipid nanoparticles such as the chemical composition, shape, and deformability could also affect the penetration behavior.37
The solid lipid nanoparticle (SLN) and nanostructured lipid carrier (NLC) are the two most studied systems among the lipid nanosystems. SLNs can be prepared in a simple procedure by dispersing solid lipids in an aqueous phase and stabilizing them with surfactants. Depending on the character- istics of the lipid matrix, both lipophilic and hydrophilic drugs can be encapsulated into the particles. Minoxidil-loaded SLNs with a diameter of 190 nm showed a better accumulation in porcine skin layers by comparison with commercial minoxidil products.38 Although SLNs have shown good permeation results, limited drug loading and stability problems such as drug expulsion are constant issues, as the solid lipids prepared with highly purified lipids may form crystalline networks, leading to drug expulsion during storage.
To this end, NLCs were developed as a second generation of lipid nanoparticles. NLCs can be prepared using a similar procedure to SLNs, except that NLCs are composed of solid and liquid lipids. The incorporation of the liquid phase in the solid lipid matrix can avert the formation of the crystalline network and increase the drug loading capacity. The weight ratios of solid to liquid lipid matrix range from 70:30 to 90:10. Wang et al. developed a minoxidil incorporated NLC by the hot high pressure homogenization method using 4% stearic acid as a solid lipid, 8% oleic acid as an oil, and 1.5% Tween-80 and 0.5% Span-80 as surfactants.39 This NLC exhibited a spherical shape with a mean diameter of 281 nm and presented high entrapment efficiency of 92.5 ± 0.3% and drug loading of
13.8 ± 0.5%. Compared with SLN, this NLC exhibited better
stability in the particle size and entrapment efficiency for 3 months, and the loading minoxidil showed 10.7 times greater skin permeation due to the presence of oleic acid in the composition that may act as a permeation enhancer. For the design of NLC, the lipid composition is directly related to the entrapment efficiency of drugs. In general, a higher ratio of liquid lipid could bring an amorphous form to the solid lipid
cellulose gel and demonstrated greater permeation than finasteride hydroalcoholic solution or conventional gel containing finasteride.43 Liposomal phospholipids can lead to a rearrangement of intercellular lipids, providing room for drug accumulation and enhancing skin delivery. A vitamin C, ascorbyl palmitate derivative was used to prepare a liposome vesicle (aspasomes) for encapsulation of the antioxidant melatonin.44 This vesicle was applied as a cosmeceutical for clinical treatment of hair loss. Results revealed that melatonin- loaded aspasomes showed a better therapeutic effect than the melatonin solution, reflected by increased hair thickness and density and decreased hair loss in the most of patients.44
2.3. Polymeric Nanoparticles. Polymeric nanoparticles
have the ability to control drug release and increase shelf life.45 Tuning the chemical or physical characters of the polymer matrix can control the drug release and delivery efficiency.34 Minoxidil-loaded polymeric nanoparticles with a diameter of 100−150 nm were fabricated using poly(L-lactide-co-glycolide) (PLGA) as the matrix via W/O/W solvent evaporation and sonication.46 The nanoparticles facilitated the delivery of minoxidil into hair follicles due to their small size as well as a prolonged drug release. Finasteride-loaded PLGA polymeric nanoparticles were prepared via a modified method of the emulsification/solvent diffusion.47 This polymeric nanoparticle had a mean particle size of 300 nm and a high encapsulation efficiency of 79.49 ± 0.47%. In vitro permeation study demonstrated that this nanoparticle extended the residence time of finasteride in the skin. A new delivery system known as phospholipid−polymer hybrid nanoparticles (LPNPs) is composed of a mixture of polymers and phospholipids. This system has integrated characteristics of both liposomes and polymeric nanoparticles and avoids the deficiency associated with their individual usage.48 The polymer provides skeleton structural properties (mechanical strength, narrow size with high surface area, a controlled release, and biodegradability), and the lipid brings cell membrane biomimetic characteristics to this delivery system.

⦁ MICRONEEDLES FOR TRANSDERMAL DELIVERY
Microneedle arrays comprise microneedles that are less than 1000 μm in length.49,50 Microneedles (MNs) can encapsulate various therapeutics, including hydrophilic drugs (peptides or genetic macromolecules). The length of microneedles ensures adequate penetration across the stratum corneum. The loaded drugs can be released upon skin insertion, with the release profiles varied upon adjustment of the MN matrix and

matrix and reduces the crystallinity of the particles, thereby
improving the encapsulation efficiency. A higher proportion of oleic acid in the lipid matrix achieved an entrapment efficiency of above 70%.40
2.2. Liposomes. Besides lipid nanoparticles, liposomes have also been studied for hair loss treatment. Liposomes are spherical lipid bilayer nanostructures with an outer lipid bilayer and an aqueous core. Liposomes consist mainly of cholesterol and phospholipids.41 Therefore, these structures can load hydrophilic drugs in the core and hydrophobic drugs between the bilayers. In addition, the lipophilic character of the
embedded formulations. This brings a high transdermal delivery efficiency of drugs, especially for those with large and hydrophilic cargos, such as proteins, peptides, and cellular particles.51−55 In recent years, MNs have been applied in hair loss treatment. Compared with the conventional topical administration, MNs can improve the therapeutic efficacy and reduce dose frequency.
Jung and coauthors fabricated dissolvable microneedles (DMNs) for transdermal delivery of valproic acid (VPA), which is approved by the FDA as an anticonvulsant drug and is capable to induce hair follicle regrowth.56 Carboxymethyl

liposomes makes them an effective delivery system for
penetrating the hair follicles. Minoxidil-loaded liposomes presented improved skin retention into the pilosebaceous unit and reduced systemic drug permeation by comparison with nonliposomal formulations in a rodent model.42 Lip- osomes containing finasteride were incorporated into methyl-
cellulose (CMC), a biodegradable and biocompatible polymer,
was used as the main matrix of the DMNs, which provides the sufficient mechanical strength for skin penetration. Compared with the currently used topical administration of organic- solvent-based VPA (topical VPA), DMNs increased the VPA delivery efficiency across the skin and induced hair regrowth

 

Figure 3. Microneedle arrays for hair loss treatment. (A) Schematic representation of (a) dissolvable microneedles (DMN) for transdermal delivery of valproic acid (VPA) and (b) the patchless delivery systems.56 Used with permission from ref 56, copyright 2018 Elsevier. (B) Schematic of hair loss therapy through a detachable keratin hydrogel-based microneedle system for transdermal delivery of UK5099-loaded PLGA nanoparticles and mesenchymal stem cell (MSC)-derived exosomes.57 Used with permission from ref 57, copyright 2019 ACS.
significantly in in vivo testing. Additionally, the research demonstrated that the creation of microwounds via DMNs could elevate the expression of Wnt signaling dependent proteins and had a potential of promoting HF regrowth with no involvement of VPA (Figure 3A). Recently, Gu and coauthors described a microneedle-mediated transdermal delivery system for codelivery of a small molecular drug UK5099 and mesenchymal stem cell (MSC)-derived exo- somes57 (Figure 3B). Both MSC-derived exosomes8,58−60 and UK509924 were reported as potential hair follicle stem cell (HFSC) activators. A natural protein derived from human hair, keratin, was used as the matrix material of MNs due to its biodegradation and long-term biocompatibility. Of note, this work constructed a robust keratin hydrogel with a network cross-linking structure consisting of intermolecular disulfide bonds via a mild and simple disulfide shuffling strategy. This strategy made full use of the cysteine-rich characteristic of keratin, with no external chemical cross-linkers involved in the preparation process, thus ensuring the microneedles with high mechanical strength, sustained drug release capacity, and good biocompatibility to the skin tissue. The in vivo test demonstrated that compared with the conventional mode of administration, that is, the subcutaneous injection of exosomes and topical administration of UK5099, this MN-mediated delivery system could greatly enhance the treatment efficiency at a reduced dosage. Promoted hair regrowth was achieved through only two rounds of administration in a mouse model. Of note, unlike the common MNs supported on a patch, both studies mentioned above proposed “patchless” microneedle systems, in view of the practical usage on the scalp or other hairy head regions. The former is achieved via fabricating microneedle arrays over microcavities and performing them using a micropillar application system. The latter designed a detachable microneedle system by building the microneedle arrays over a water-soluble hyaluronic acid (HA)-based patch.
⦁ PENETRATION ENHANCEMENT STRATEGY
Besides the aforementioned drug delivery systems, several penetration enhancement strategies have been established to assist a better penetration of drugs into skin by diminishing the barrier of the skin.61 Such penetration enhancement strategies mainly include physical modality and chemical penetration enhancers.62
⦁ Physical Penetration Strategy. Microneedling and fractional laser therapy have been utilized as a “wounding source” to trigger hair growth and achieve some effects within a small group of patients.63−65 These treatments work by creating an array of microscopic channels cross the outer skin layer. Their mechanism of hair growth stimulation is considered to be associated with wound-induced follicle neogenesis.66 However, no definitive conclusions regarding their efficacy have been made due to the equivocal mechanism and lack of abundant clinical data. However, these treatments are verified potentially to enhance drug penetration efficiency via those “channels”, thus realizing enhanced therapeutic effects compared to topical drug monotherapy. MNs as a drug delivery system were introduced above. Besides this, micro- needling technology is a new treatment modality for stimulating hair growth.66 It can be performed with various instruments, for instance, a barrel-shaped roller studded with fine needles ranging from 0.5 to 2.5 mm in length. By repeated rolling on the head skin, the fine needles could mechanically puncture the stratum corneum, forming diffusion channels for subsequently applied formulations. Evidence has shown that such treatment could trigger the wound healing response and induce collagen deposition, neovascularization, or growth factor production in treated skin, which is believed to be beneficial for hair follicle activation, although the underlying mechanisms need further investigation. Microneedling has also been combined with medications or hair growth activators, such as minoxidil67 and platelet-rich plasma68,69 for hair

This HA patch can rapidly absorb cutaneous tissue fluids after
insertion into the skin and be easily detached at several hours postinsertion.
regrowth. Evidence suggested that microneedling treatment
might enhance the skin penetration of these medications via the diffusion channels across the stratum corneum. Similar to

https://dx.doi.org/10.1021/acs.molpharmaceut.0c00041

Table 1. Summary of Representative Transdermal Drug Delivery Systems for Hair Loss Treatment

 

solid lipid nanoparticle (SLN)
nanostructured lipid carrier (NLC)
solid lipids

solid and liquid lipids
minoxidil38 targeted drug delivery to the hair follicles; prolonged drug release in hair follicles compared to unencapsulated drugs

minoxidil39 targeted drug delivery to the hair follicles; prolonged drug release in hair follicles compared to unencapsulated drugs; increased drug loading levels and stability in storage compared to SLNs

liposome cholesterol and phospholipids
minoxidil42 capacity of follicular penetration
finasteride43

polymeric nanoparticle
lipophilic polymers
minoxidil46 controlled drug release in a sustained manner
finasteride47

microneedles biopolymers or synthetic polymers
valproic acid56
UK5099 and exosomes57
enhanced skin penetration; high bioavailability; controlled drug release in a sustained manner

 

 

microneedling, pretreatment of the skin surface with a fractional laser (10−15 mJ of energy) can also improve the therapeutic effect of topical minoxidil70 or topical finasteride.71 With the advancement of nanotechnology, the physical penetration strategies in combination with nanoparticles greatly improve hair loss therapies. Liao et al. designed a transdermal drug delivery platform for enhancing the delivery of minoxidil to hair follicles by integrating a multifunctional ultrasound (US) contrast agent.72 This US contrast agent was composed of albumin-shelled microbubbles (MBs) that absorbed chitosan oligosaccharide lactate (COL), with a diameter of 4500 nm. Under the US local treatment on the skin, the delivery system resulted in 1.7 and 2.3 times the deposition and penetration, respectively, of minoxidil relative to the minoxidil group. The hair regrowth at days 10 and 14 in vivo increased by 22.6 and 64.7%, respectively. Ryu Jee-Yeon et al. designed a microbubble−nanoliposomal particle for the transdermal delivery of a Cas9/sgRNA ribonucleoprotein complex. Such a particle transferred the protein complex into dermal papilla cells in the hair follicle by US-induced microbubble cavitation.28 In addition, Giorgio et al. described a new therapeutic combination of photodynamic therapy with 5-aminolaevulinic acid and microneedling for alopecia areata. The therapy effect was evaluated on 41 patients with moderate to severe AA. The results demonstrated that microneedling facilitated better skin penetration of 5-aminolaevulinic acid and
subsequent indirect immunosuppression.73
⦁ Chemical Penetration Enhancers. Chemical pene- tration enhancers are the agents that can chemically disturb the stratum corneums and thus enhance drug penetration
minoxidil to penetrate the skin but also realized a sustained drug release, leading to enhanced therapeutic outcomes. Due to a lack of organic solvents, no adverse effects associated with the hydrogel formulation were triggered.

⦁ CONCLUSIONS AND OUTLOOK
Topical application of drugs offers advantages for hair loss treatment. However, the stratum corneum of the skin poses a major barrier to transdermal delivery of drugs, especially for large and hydrophilic molecules. To date, many transdermal drug delivery systems for hair loss treatment, including drug nanocarriers, microneedle arrays, and physical or chemical penetration strategies have been developed for enhanced transdermal delivery and therapeutic efficacy (Table 1). However, further improvement of the existing therapy strategies and exploring novel treatment modalities are still highly valuable.
First, better understanding of the mechanisms of hair loss is
needed to fully explain the genetic basis of hair loss via suitable animal models, such as C3H/HeJ mice and human skin xenotransplants on SCID mice,76 which may direct the screening of more effective therapeutic agents as well as the exploration of new targeted therapeutic alternatives.5,18,77−83 Second, it is essential to thoroughly evaluate the efficacy and safety of currently developed drug delivery systems in hair loss therapy.84−86 Third, synergistic combinations of penetration enhancers and drug delivery systems might be an efficient way to achieve the full potential of topical treatment.87
It should be noted that cell therapy represents a promising

efficiency.62,74 Ethanol, dimethyl sulfoxide, dimethyl isosor- 88

bide, propylene glycol, and isopropyl myristate have been
alternative for hair loss treatment. Dermal papilla cells

investigated or used in first-line drugs as chemical penetration enhancers. The clinically used topical formulation of minoxidil contains a significant amount of ethanol and propylene glycol as cosolvents.62 These cosolvents greatly improve the water solubility of minoxidil and enhance the drug penetration efficiency across the skin. For example, Rogaine Extra Strength (5% w/v Minoxidil) contains 50% w/v propylene glycol and 30% w/v ethanol. However, a certain amount of patients present clinical complaints like irritation, pruritus, and dermatitis after repeated applications, mainly due to these cosolvents. Efforts have been made to resolve these limitations. Lopedota et al. developed an alginate hydrogel loaded with a minoxidil/hydroxypropyl-β-cyclodextrin inclusion complex as a topical formulation.75 This formulation not only facilitated
(DPCs) are highly specialized mesenchymal cells found in hair
follicles. DPCs retain a remarkable capacity to induce hair regeneration, due to their role in the epithelial−mesenchymal interaction that governs hair follicle morphogenesis and hair formation.7 This provides a possible solution of cell therapy for hair loss, in which DPCs from donor follicles can be expanded in culture and thereafter be implanted into a bald region of the skin.89 Many new follicles can then be formed via in situ cell expansion. According to the literature, the key challenge for cell therapy lies in the retention of inherent inducing capabilities and productivity of the implanted cells in situ.90−93 In view of this, strategies can be developed for remodeling the hair microenvironment and improving the retention of cellular functions.94,95

https://dx.doi.org/10.1021/acs.molpharmaceut.0c00041

⦁ AUTHOR INFORMATION
Corresponding Authors
Zhen Gu − Department of Bioengineering and California NanoSystems Institute, Jonsson Comprehensive Cancer Center, and Center for Minimally Invasive Therapeutics, University of California, Los Angeles, Los Angeles, California 90095, United States; orcid.org/0000-0003-2947-4456; Email: guzhen@ ucla.edu
Guang Yang − Key Laboratory of Science & Technology of Eco- Textile, Donghua University, Ministry of Education, Shanghai 201620, China; College of Chemistry, Chemical Engineering and Biotechnology, Donghua University, Shanghai 201620, China; orcid.org/0000-0002-5889-3978; Email: gyang@ dhu.edu.cn
Author
Guojun Chen − Department of Bioengineering and California NanoSystems Institute, Jonsson Comprehensive Cancer Center, and Center for Minimally Invasive Therapeutics, University of California, Los Angeles, Los Angeles, California 90095, United States
Complete contact information is available at: https://pubs.acs.org/10.1021/acs.molpharmaceut.0c00041

Notes
The authors declare no competing financial interest.
⦁ ACKNOWLEDGMENTS
This work was supported by the Fundamental Research Funds
for the Central Universities (2232019G-04) and Shanghai Pujiang Program (2019PJD002) to G.Y. and the start-up packages of UCLA to Z.G.
⦁ REFERENCES
⦁ Lolli, F.; Pallotti, F.; Rossi, A.; Fortuna, M. C.; Caro, G.; Lenzi,
A.; Sansone, A.; Lombardo, F. Androgenetic alopecia: a review.
Endocrine 2017, 57 (1), 9−17.
⦁ Simakou, T.; Butcher, J. P.; Reid, S.; Henriquez, F. L. ⦁ Alopecia ⦁ areata: A multifactorial autoimmune condition. J. Autoimmun. 2019, 98, 74−85.
⦁ Rajabi, F.; Drake, L. A.; Senna, M. M.; Rezaei, N. ⦁ Alopecia ⦁ areata:⦁ ⦁ a⦁ ⦁ review⦁ ⦁ of⦁ ⦁ disease⦁ ⦁ pathogenesis. Br. J. Dermatol. 2018, 179 (5), 1033−1048.
⦁ Gilhar, A.; Kalish, R. S. ⦁ Alopecia⦁ ⦁ areata:⦁ ⦁ a⦁ ⦁ tissue⦁ ⦁ specific ⦁ autoimmune⦁ ⦁ disease⦁ ⦁ of⦁ ⦁ the⦁ ⦁ hair⦁ ⦁ follicle.⦁ Autoimmun. Rev. 2006, 5 (1), 64−9.
⦁ Son, M. J.; Jeong, J. K.; Kwon, Y.; Ryu, J. S.; Mun, S. J.; Kim, H. J.; Kim, S. W.; Yoo, S.; Kook, J.; Lee, H.; Kim, J.; Chung, K. S. ⦁ A⦁ ⦁ novel ⦁ and safe small molecule enhances hair follicle regeneration ⦁ by ⦁ facilitating metabolic reprogramming. Exp. Mol. Med. 2018, 50 (12), 160.
⦁ Simakou, T.; Butcher, J.; Reid, S.; Henriquez, F. ⦁ Alopecia⦁ ⦁ areata: ⦁ A⦁ ⦁ multifactorial⦁ ⦁ autoimmune⦁ ⦁ condition.⦁ J. Autoimmun. 2019, 98, 74− 85.
⦁ Hsu, Y. C.; Pasolli, H. A.; Fuchs, E. ⦁ Dynamics⦁ ⦁ between⦁ ⦁ stem ⦁ cells,⦁ ⦁ niche,⦁ ⦁ and⦁ ⦁ progeny⦁ ⦁ in⦁ ⦁ the⦁ ⦁ hair⦁ ⦁ follicle.⦁ Cell 2011, 144 (1), 92− 105.
⦁ Hsu, Y. C.; Li, L.; Fuchs, E. ⦁ Emerging⦁ ⦁ interactions⦁ ⦁ between⦁ ⦁ skin ⦁ stem⦁ ⦁ cells⦁ ⦁ and⦁ ⦁ their⦁ ⦁ niches.⦁ Nat. Med. 2014, 20 (8), 847−56.
⦁ Oh, J. W.; Kloepper, J.; Langan, E. A.; Kim, Y.; Yeo, J.; Kim, M. J.; Hsi, T. C.; Rose, C.; Yoon, G. S.; Lee, S. J.; Seykora, J.; Kim, J. C.; Sung, Y. K.; Kim, M.; Paus, R.; Plikus, M. V. ⦁ A⦁ ⦁ Guide⦁ ⦁ to⦁ ⦁ Studying ⦁ Human⦁ ⦁ Hair⦁ ⦁ Follicle⦁ ⦁ Cycling⦁ ⦁ In⦁ ⦁ Vivo.⦁ J. Invest. Dermatol. 2016, 136 (1), 34−44.

⦁ Ohn, J.; Kim, K. H.; Kwon, O. ⦁ Evaluating hair growth ⦁ promoting effects of candidate substance: A review of research ⦁ methods.⦁ J. Dermatol. Sci. 2019, 93 (3), 144−149.
⦁ Purba, T. S.; Haslam, I. S.; Poblet, E.; Jimenez, F.; Gandarillas, A.; Izeta, A.; Paus, R. ⦁ Human⦁ ⦁ epithelial⦁ ⦁ hair⦁ ⦁ follicle⦁ ⦁ stem⦁ ⦁ cells⦁ ⦁ and ⦁ their progeny: current state of knowledge, the widening gap ⦁ in ⦁ translational research and future challenges. BioEssays 2014, 36 (5), 513−25.
⦁ Fan, S. M.; Tsai, C. F.; Yen, C. M.; Lin, M. H.; Wang, W. H.; Chan, C. C.; Chen, C. L.; Phua, K. K. L.; Pan, S. H.; Plikus, M. V.; Yu,
S. L.; Chen, Y. J.; Lin, S. J. Inducing hair follicle neogenesis with secreted proteins enriched in embryonic skin. Biomaterials 2018, 167, 121−131.
⦁ Zhang, P.; Kling, R. E.; Ravuri, S. K.; Kokai, L. E.; Rubin, J. P.; Chai, J. K.; Marra, K. G. ⦁ A review of adipocyte lineage cells and ⦁ dermal⦁ ⦁ papilla⦁ ⦁ cells⦁ ⦁ in⦁ ⦁ hair⦁ ⦁ follicle⦁ ⦁ regeneration.⦁ J. Tissue Eng. 2014, 5, 204173141455685.
⦁ Lavker, R. M.; Sun, T. T.; Oshima, H.; Barrandon, Y.; Akiyama, M.; Ferraris, C.; Chevalier, G.; Favier, B.; Jahoda, C. A.; Dhouailly, D.; Panteleyev, A. A.; Christiano, A. M. ⦁ Hair⦁ ⦁ follicle⦁ ⦁ stem⦁ ⦁ cells.⦁ J. Invest. Dermatol. Symp. Proc. 2003, 8 (1), 28−38.
⦁ Pantelireis, N.; Higgins, C. A. ⦁ A bald statement – Current ⦁ approaches to manipulate miniaturisation focus only on⦁ ⦁ promoting ⦁ hair growth. Experimental dermatology 2018, 27 (9), 959−965.
⦁ Liu, Y.-C. S.; Jee, S.-H.; Chan, J.-Y. L. ⦁ Hair⦁ ⦁ transplantation⦁ ⦁ for ⦁ the⦁ ⦁ treatment⦁ ⦁ of⦁ ⦁ lichen⦁ ⦁ planopilaris⦁ ⦁ and⦁ ⦁ frontal⦁ ⦁ fibrosing⦁ ⦁ alopecia:⦁ ⦁ A ⦁ report⦁ ⦁ of⦁ ⦁ two⦁ ⦁ cases.⦁ Australas. J. Dermatol. 2018, 59 (2), No. e118.
⦁ Goren, A.; Naccarato, T. ⦁ Mino⦁ x⦁ idil in the treatment ⦁ of ⦁ androgenetic alopecia. Dermatologic therapy 2018, 31 (5), No. e12686.
⦁ Guo, H.; Gao, W. V.; Endo, H.; McElwee, K. J. ⦁ E⦁ x⦁ perimental ⦁ and early investigational drugs for androgenetic alopecia. Expert Opin. Invest. Drugs 2017, 26 (8), 917−932.
⦁ Gupta, A. K.; Carviel, J.; MacLeod, M. A.; Shear, N. ⦁ Assessing ⦁ finasteride-associated sexual dysfunction using the FAERS database. Journal of the European Academy of Dermatology and Venereology: JEADV 2017, 31 (6), 1069−1075.
⦁ Guo, M.; Heran, B.; Flannigan, R.; Kezouh, A.; Etminan, M. ⦁ Persistent⦁ ⦁ Sexual⦁ ⦁ Dysfunction⦁ ⦁ with⦁ ⦁ Finasteride⦁ ⦁ 1⦁ ⦁ mg⦁ ⦁ Taken⦁ ⦁ for⦁ Hair ⦁ Loss. Pharmacotherapy 2016, 36 (11), 1180−1184.
⦁ Ekelem, C.; Pham, C.; Atanaskova Mesinkovska, N. ⦁ A ⦁ Systematic Review of the Outcome of Hair Transplantation ⦁ in ⦁ Primary⦁ ⦁ Scarring⦁ ⦁ Alopecia.⦁ Skin appendage disorders 2019, 5 (2), 65− 71.
⦁ Smith, A. A.; Li, J.; Liu, B.; Hunter, D.; Pyles, M.; Gillette, M.; Dhamdhere, G. R.; Abo, A.; Oro, A.; Helms, J. A. ⦁ Activating⦁ ⦁ Hair ⦁ Follicle⦁ ⦁ Stem⦁ ⦁ Cells⦁ ⦁ via⦁ ⦁ R-spondin2⦁ ⦁ to⦁ ⦁ Stimulate⦁ ⦁ Hair⦁ ⦁ Growth.⦁ J. Invest. Dermatol. 2016, 136 (8), 1549−1558.
⦁ Gross, J. C.; Chaudhary, V.; Bartscherer, K.; Boutros, M. ⦁ Active ⦁ Wnt⦁ ⦁ proteins⦁ ⦁ are⦁ ⦁ secreted⦁ ⦁ on⦁ ⦁ exosomes.⦁ Nat. Cell Biol. 2012, 14 (10), 1036−45.
⦁ Iorizzo, M.; Tosti, A. ⦁ Emerging⦁ ⦁ drugs⦁ ⦁ for⦁ ⦁ alopecia⦁ ⦁ areata:⦁ ⦁ JAK ⦁ inhibitors.⦁ Expert Opin. Emerging Drugs 2018, 23 (1), 77−81.
⦁ Wang, E. H. C.; Sallee, B. N.; Tejeda, C. I.; Christiano, A. M. ⦁ JAK Inhibitors for Treatment of Alopecia Areata. J. Invest. Dermatol. 2018, 138 (9), 1911−1916.
⦁ Flores, A.; Schell, J.; Krall, A. S.; Jelinek, D.; Miranda, M.; Grigorian, M.; Braas, D.; White, A. C.; Zhou, J. L.; Graham, N. A.; Graeber, T.; Seth, P.; Evseenko, D.; Coller, H. A.; Rutter, J.; Christofk, H. R.; Lowry, W. E. ⦁ Lactate⦁ ⦁ dehydrogenase⦁ ⦁ activity⦁ ⦁ drives ⦁ hair follicle stem cell activation. Nat. Cell Biol. 2017, 19 (9), 1017− 1026.
⦁ Chai, M.; Jiang, M.; Vergnes, L.; Fu, X.; de Barros, S. C.; Doan,
N. B.; Huang, W.; Chu, J.; Jiao, J.; Herschman, H.; Crooks, G. M.; Reue, K.; Huang, J. Stimulation of Hair Growth by Small Molecules that Activate Autophagy. Cell Rep. 2019, 27 (12), 3413−3421.
⦁ Ryu, J. Y.; Won, E. J.; Lee, H. A. R.; Kim, J. H.; Hui, E.; Kim, H. P.; Yoon, T. J. ⦁ Ultrasound-activated⦁ ⦁ particles⦁ ⦁ as⦁ ⦁ CRISPR/Cas9

https://dx.doi.org/10.1021/acs.molpharmaceut.0c00041

delivery system for androgenic alopecia therapy. Biomaterials 2020,
232, 119736.
⦁ Prausnitz, M. R.; Langer, R. ⦁ Transdermal drug delivery. Nat. Biotechnol. 2008, 26 (11), 1261−8.
⦁ Vogt, A.; Wischke, C.; Neffe, A. T.; Ma, N.; Alexiev, U.; Lendlein, A. ⦁ Nanocarriers⦁ ⦁ for⦁ ⦁ drug⦁ ⦁ delivery⦁ ⦁ into⦁ ⦁ and⦁ ⦁ through⦁ ⦁ the⦁ ⦁ skin⦁ ⦁ – ⦁ Do existing technologies match clinical challenges? J. Controlled Release 2016, 242, 3−15.
⦁ Patzelt, A.; Lademann, J. ⦁ Drug⦁ ⦁ delivery⦁ ⦁ to⦁ ⦁ hair⦁ ⦁ follicles.⦁ Expert Opin. Drug Delivery 2013, 10 (6), 787−797.
⦁ Carter, P.; Narasimhan, B.; Wang, Q. ⦁ Biocompatible⦁ ⦁ nano- ⦁ particles⦁ ⦁ and⦁ ⦁ vesicular⦁ ⦁ systems⦁ ⦁ in⦁ ⦁ transdermal⦁ ⦁ drug⦁ ⦁ delivery⦁ ⦁ for⦁ ⦁ various ⦁ skin⦁ ⦁ diseases.⦁ Int. J. Pharm. 2019, 555, 49−62.
⦁ Roberts, M. S.; Mohammed, Y.; Pastore, M. N.; Namjoshi, S.; Yousef, S.; Alinaghi, A.; Haridass, I. N.; Abd, E.; Leite-Silva, V. R.; Benson, H.; Grice, J. E. ⦁ Topical and cutaneous delivery ⦁ using ⦁ nanosystems. J. Controlled Release 2017, 247, 86−105.
⦁ Palmer, B. C.; DeLouise, L. A. ⦁ Nanoparticle-Enabled ⦁ Transdermal⦁ ⦁ Drug⦁ ⦁ Delivery⦁ ⦁ Systems⦁ ⦁ for⦁ ⦁ Enhanced⦁ ⦁ Dose⦁ ⦁ Control ⦁ and⦁ ⦁ Tissue⦁ ⦁ Targeting.⦁ Molecules 2016, 21 (12), 1719.
⦁ Sala, M.; Diab, R.; Elaissari, A.; Fessi, H. ⦁ Lipid⦁ ⦁ nanocarriers⦁ ⦁ as ⦁ skin drug delivery systems: Properties, mechanisms ⦁ of ⦁ skin ⦁ interactions⦁ ⦁ and⦁ ⦁ medical⦁ ⦁ applications. Int. J. Pharm. 2018, 535 (1− 2), 1−17.
⦁ Mishra, D. K.; Shandilya, R.; Mishra, P. K. ⦁ Lipid ⦁ based ⦁ nanocarriers: a translational perspective. Nanomedicine 2018, 14 (7), 2023−2050.
⦁ Garces, A.; Amaral, M. H.; Sousa Lobo, J. M.; Silva, A. C. ⦁ Formulations ⦁ based ⦁ on ⦁ solid lipid nanoparticles (SLN) ⦁ and ⦁ nanostructured⦁ ⦁ lipid⦁ ⦁ carriers⦁ ⦁ (NLC)⦁ ⦁ for⦁ ⦁ cutaneous⦁ ⦁ use:⦁ ⦁ A⦁ ⦁ review. Eur. J. Pharm. Sci. 2018, 112, 159−167.
⦁ Padois, K.; Cantieni, C.; Bertholle, V.; Bardel, C.; Pirot, F.; Falson, F. ⦁ Solid lipid nanoparticles suspension versus commercial ⦁ solutions⦁ ⦁ for⦁ ⦁ dermal⦁ ⦁ delivery⦁ ⦁ of⦁ ⦁ mino⦁ x⦁ idil. Int. J. Pharm. 2011, 416 (1), 300−304.
⦁ Wang, W.; Chen, L.; Huang, X.; Shao, A. ⦁ Preparation ⦁ and ⦁ Characterization of Mino⦁ x⦁ idil Loaded Nanostructured Lipid Carriers. AAPS PharmSciTech 2017, 18 (2), 509−516.
⦁ Nagaich, U.; Gulati, N. ⦁ Nanostructured lipid carriers ⦁ (NLC) ⦁ based controlled release topical gel of clobetasol propionate:⦁ ⦁ design ⦁ and⦁ ⦁ in⦁ ⦁ vivo⦁ ⦁ characterization. Drug Delivery Transl. Res. 2016, 6 (3), 289−98.
⦁ Rideau, E.; Dimova, R.; Schwille, P.; Wurm, F. R.; Landfester,
K. Liposomes and polymersomes: a comparative review towards cell mimicking. Chem. Soc. Rev. 2018, 47 (23), 8572−8610.
⦁ Khandelia, R.; Bhandari, S.; Pan, U. N.; Ghosh, S. S.; Chattopadhyay, A. ⦁ Gold Nanocluster Embedded Albumin ⦁ Nano- ⦁ particles for Two-Photon Imaging of Cancer Cells ⦁ Accompanying ⦁ Drug⦁ ⦁ Delivery.⦁ Small 2015, 11 (33), 4075−81.
⦁ Lai, X.; Tan, L.; Deng, X.; Liu, J.; Li, A.; Liu, J.; Hu, J. ⦁ Coordinatively Self-Assembled Luminescent Gold Nanoparticles: ⦁ Fluorescence Turn-On System for High-Efficiency Passive ⦁ Tumor ⦁ Imaging. ACS Appl. Mater. Interfaces 2017, 9 (6), 5118−5127.
⦁ Hatem, S.; Nasr, M.; Moftah, N. H.; Ragai, M. H.; Geneidi, A. S.; Elkheshen, S. A. ⦁ Melatonin vitamin C-based nanovesicles ⦁ for ⦁ treatment of androgenic alopecia: Design, characterization and clinical ⦁ appraisal.⦁ Eur. J. Pharm. Sci. 2018, 122, 246−253.
⦁ El-Say, K. M.; El-Sawy, H. S. ⦁ Polymeric nanoparticles: ⦁ Promising⦁ ⦁ platform⦁ ⦁ for⦁ ⦁ drug⦁ ⦁ delivery. Int. J. Pharm. 2017, 528 (1− 2), 675−691.
⦁ Takeuchi, I.; Hida, Y.; Makino, K. ⦁ Mino⦁ x⦁ idil-encapsulated ⦁ poly(L-lactide-co-glycolide)⦁ ⦁ nanoparticles⦁ ⦁ with⦁ ⦁ hair⦁ ⦁ follicle⦁ ⦁ delivery ⦁ properties prepared using W/O/W solvent evaporation ⦁ and ⦁ sonication.⦁ Bio-Med. Mater. Eng. 2018, 29 (2), 217−228.
⦁ Roque, L. V.; Dias, I. S.; Cruz, N.; Rebelo, A.; Roberto, A.; Rijo, P.; Reis, C. P. ⦁ Design of Finasteride-Loaded Nanoparticles ⦁ for ⦁ Potential⦁ ⦁ Treatment⦁ ⦁ of⦁ ⦁ Alopecia. Skin pharmacology and physiology 2017, 30 (4), 197−204.
⦁ Das, L.; Kaurav, M.; Pandey, R. S. ⦁ Phospholipid-polymer ⦁ hybrid nanoparticle-mediated transfollicular delivery of quercetin: ⦁ prospective⦁ ⦁ implement⦁ ⦁ for⦁ ⦁ the⦁ ⦁ treatment⦁ ⦁ of⦁ ⦁ androgenic⦁ ⦁ alopecia.⦁ Drug Dev. Ind. Pharm. 2019, 45 (10), 1654−1663.
⦁ Kim, Y. C.; Park, J. H.; Prausnitz, M. R. ⦁ Microneedles⦁ ⦁ for⦁ ⦁ drug ⦁ and⦁ ⦁ vaccine⦁ ⦁ delivery.⦁ Adv. Drug Delivery Rev. 2012, 64 (14), 1547−68.
⦁ Ye, Y.; Yu, J.; Wen, D.; Kahkoska, A. R.; Gu, Z. ⦁ Polymeric ⦁ microneedles for transdermal protein delivery. Adv. Drug Delivery Rev. 2018, 127, 106−118.
⦁ Yu, J.; Wang, J.; Zhang, Y.; Chen, G. J.; Mao, W. W.; Ye, Y. Q.; Kahkoska, A. R.; Buse, J. B.; Langer, R.; Gu, Z. ⦁ Glucose-responsive ⦁ insulin⦁ ⦁ patch⦁ ⦁ for⦁ ⦁ the⦁ ⦁ regulation⦁ ⦁ of⦁ ⦁ blood⦁ ⦁ glucose⦁ ⦁ in⦁ ⦁ mice⦁ ⦁ and⦁ ⦁ minipigs. Nature Biomedical Engineering 2020. .
⦁ Chen, G. J.; Chen, Z. T.; Wen, D.; Wang, Z. J.; Li, H. J.; Zeng, Y.; Dotti, G.; Wirz, R. E.; Gu, Z. ⦁ Transdermal⦁ ⦁ cold⦁ ⦁ atmospheric ⦁ plasma-mediated immune checkpoint blockade therapy. Proc. Natl. Acad. Sci. U. S. A. 2020, 117 (7), 3687−3692.
⦁ Lee, K.; Goudie, M. J.; Tebon, P.; Sun, W. J.; Luo, Z. M.; Lee, J. M.; Zhang, S. M.; Fetah, K.; Kim, H.; Xue, Y.; Darabi, M. A.; Ahadian, S.; Sarikhani, E.; Ryu, W.; Gu, Z.; Weiss, P. S.; Dokmeci, M. R.; Ashammakhi, N.; Khademhosseini, A. ⦁ Non-transdermal ⦁ microneedles ⦁ for⦁ ⦁ advanced⦁ ⦁ drug⦁ ⦁ delivery.⦁ Adv. Drug Delivery Rev. 2019. In press.
⦁ Prausnitz, M. R. ⦁ Microneedles for transdermal drug⦁ ⦁ delivery.
Adv. Drug Delivery Rev. 2004, 56 (5), 581−7.
⦁ Jin, X.; Zhu, D. D.; Chen, B. Z.; Ashfaq, M.; Guo, X. D. ⦁ Insulin ⦁ delivery⦁ ⦁ systems⦁ ⦁ combined⦁ ⦁ with⦁ ⦁ microneedle⦁ ⦁ technology. Adv. Drug Delivery Rev. 2018, 127, 119−137.
⦁ Fakhraei Lahiji, S.; Seo, S. H.; Kim, S.; Dangol, M.; Shim, J.; Li,
C. G.; Ma, Y.; Lee, C.; Kang, G.; Yang, H.; Choi, K.-Y.; Jung, H. Transcutaneous implantation of valproic acid-encapsulated dissolving microneedles induces hair regrowth. Biomaterials 2018, 167, 69−79.
⦁ Yang, G.; Chen, Q.; Wen, D.; Chen, Z.; Wang, J.; Chen, G.; Wang, Z.; Zhang, X.; Zhang, Y.; Hu, Q.; Zhang, L.; Gu, Z. ⦁ A ⦁ Therapeutic⦁ ⦁ Microneedle⦁ ⦁ Patch⦁ ⦁ Made⦁ ⦁ from⦁ ⦁ Hair-Derived⦁ ⦁ Keratin⦁ ⦁ for ⦁ Promoting Hair Regrowth. ACS Nano 2019, 13 (4), 4354−4360.
⦁ Carrasco, E.; Soto-Heredero, G.; Mittelbrunn, M. ⦁ The Role ⦁ of ⦁ E⦁ x⦁ tracellular Vesicles in Cutaneous Remodeling and Hair Follicle ⦁ Dynamics. Int. J. Mol. Sci. 2019, 20 (11), 2758.
⦁ Zhang, B.; Wang, M.; Gong, A.; Zhang, X.; Wu, X.; Zhu, Y.; Shi, H.; Wu, L.; Zhu, W.; Qian, H.; Xu, W. ⦁ HucMSC-E⦁ x⦁ osome ⦁ Mediated-Wnt4 Signaling Is Required for Cutaneous Wound Healing. Stem Cells 2015, 33 (7), 2158−68.
⦁ Zhou, L.; Wang, H.; Jing, J.; Yu, L.; Wu, X.; Lu, Z. ⦁ Regulation ⦁ of hair follicle development by exosomes derived from dermal papilla ⦁ cells. Biochem. Biophys. Res. Commun. 2018, 500 (2), 325−332.
⦁ Lee, H.; Song, C.; Baik, S.; Kim, D.; Hyeon, T.; Kim, D. H. ⦁ Device-assisted transdermal drug delivery. Adv. Drug Delivery Rev. 2018, 127, 35−45.
⦁ Marwah, H.; Garg, T.; Goyal, A. K.; Rath, G. ⦁ Permeation ⦁ enhancer⦁ ⦁ strategies⦁ ⦁ in⦁ ⦁ transdermal⦁ ⦁ drug⦁ ⦁ delivery.⦁ Drug Delivery 2016, 23 (2), 564−78.
⦁ Suchonwanit, P.; Chalermroj, N.; Khunkhet, S. ⦁ Low-level⦁ ⦁ laser ⦁ therapy for the treatment of androgenetic alopecia in Thai men ⦁ and ⦁ women: a 24-week, randomized, double-blind, sham device-controlled ⦁ trial.⦁ Lasers in medical science 2019, 34 (6), 1107−1114.
⦁ Dabek, R. J.; Roh, D. S.; Ozdemir, D.; Austen, W. G., Jr.; Bojovic, B. ⦁ Fractional Laser-assisted Hair Regrowth and ⦁ Micro- ⦁ needling⦁ ⦁ for⦁ ⦁ the⦁ ⦁ Treatment⦁ ⦁ of⦁ ⦁ Alopecia⦁ ⦁ Areata:⦁ ⦁ A⦁ ⦁ Review. Cureus 2019, 11 (6), No. e4943.
⦁ Darwin, E.; Heyes, A.; Hirt, P. A.; Wikramanayake, T. C.; Jimenez, J. J. ⦁ Low-level⦁ ⦁ laser⦁ ⦁ therapy⦁ ⦁ for⦁ ⦁ the⦁ ⦁ treatment⦁ ⦁ of⦁ ⦁ androgenic ⦁ alopecia:⦁ ⦁ a⦁ ⦁ review.⦁ Lasers in medical science 2018, 33 (2), 425−434.
⦁ Kim, Y. S.; Jeong, K. H.; Kim, J. E.; Woo, Y. J.; Kim, B. J.; Kang,
H. Repeated Microneedle Stimulation Induces Enhanced Hair Growth in a Murine Model. Ann. Dermatol. 2016, 28 (5), 586−592.
⦁ Yu, A. J.; Luo, Y. J.; Xu, X. G.; Bao, L. L.; Tian, T.; Li, Z. X.; Dong, Y. X.; Li, Y. H. ⦁ A⦁ ⦁ pilot⦁ ⦁ split-scalp⦁ ⦁ study⦁ ⦁ of⦁ ⦁ combined⦁ ⦁ fractional ⦁ radiofrequency microneedling and 5% topical mino⦁ x⦁ idil in treating ⦁ male⦁ ⦁ pattern⦁ ⦁ hair⦁ ⦁ loss.⦁ Clin. Exp. Dermatol. 2018, 43 (7), 775−781.

https://dx.doi.org/10.1021/acs.molpharmaceut.0c00041

⦁ Xiao, S. E.; Miao, Y.; Wang, J.; Jiang, W.; Fan, Z. X.; Liu, X. M.; Hu, Z. Q. ⦁ As⦁ ⦁ a⦁ ⦁ carrier-transporter⦁ ⦁ for⦁ ⦁ hair⦁ ⦁ follicle⦁ ⦁ reconstitution, ⦁ platelet-rich plasma promotes proliferation and induction of⦁ ⦁ mouse ⦁ dermal⦁ ⦁ papilla⦁ ⦁ cells.⦁ Sci. Rep. 2017, 7 (1), 1125.
⦁ Dervishi, G.; Liu, H.; Peternel, S.; Labeit, A.; Peinemann, F. ⦁ Autologous platelet-rich plasma therapy for pattern hair loss:⦁ ⦁ A ⦁ systematic⦁ ⦁ review.⦁ Journal of cosmetic dermatology 2020, 19 (4), 827.
⦁ Wang, W.; Gegentana; Tonglaga; Bagenna; Li, Y. ⦁ Treatment⦁ ⦁ of ⦁ alopecia areata with nonablative fractional laser combined with ⦁ topical ⦁ mino⦁ x⦁ idil.⦁ Journal of cosmetic dermatology 2019, 18 (4), 1009−1013.
⦁ Ahmed, O. A.; Rizg, W. Y. ⦁ Finasteride⦁ ⦁ nano-transferosomal⦁ ⦁ gel ⦁ formula⦁ ⦁ for⦁ ⦁ management⦁ ⦁ of⦁ ⦁ androgenetic⦁ ⦁ alopecia:⦁ ⦁ ex⦁ ⦁ vivo⦁ ⦁ investiga- ⦁ tional approach. Drug Des., Dev. Ther. 2018, 12, 2259−2265.
⦁ Liao, A. H.; Lu, Y. J.; Lin, Y. C.; Chen, H. K.; Sytwu, H. K.; Wang, C. H. ⦁ Effectiveness of a Layer-by-Layer⦁ ⦁ Microbubbles-Based ⦁ Delivery System for Applying Mino⦁ x⦁ idil to Enhance Hair ⦁ Growth. Theranostics 2016, 6 (6), 817−27.
⦁ Giorgio, C. M.; Babino, G.; Caccavale, S.; Russo, T.; De Rosa,
A. B.; Alfano, R.; Fulgione, E.; Argenziano, G. Combination of photodynamic therapy with 5-aminolaevulinic acid and microneedling in the treatment of alopecia areata resistant to conventional therapies: our experience with 41 patients. Clin. Exp. Dermatol. 2020, 45, 323.
⦁ Pham, Q. D.; Bjorklund, S.; Engblom, J.; Topgaard, D.; Sparr,
E. Chemical penetration enhancers in stratum corneum – Relation between molecular effects and barrier function. J. Controlled Release 2016, 232, 175−87.
⦁ Lopedota, A.; Denora, N.; Laquintana, V.; Cutrignelli, A.; Lopalco, A.; Tricarico, D.; Maqoud, F.; Curci, A.; Mastrodonato, M.; la Forgia, F.; Fontana, S.; Franco, M. ⦁ Alginate-Based Hydrogel ⦁ Containing Mino⦁ x⦁ idil/Hydro⦁ x⦁ ypropyl-beta-Cyclodextrin Inclusion ⦁ Complex⦁ ⦁ for⦁ ⦁ Topical⦁ ⦁ Alopecia⦁ ⦁ Treatment. J. Pharm. Sci. 2018, 107 (4), 1046−1054.
⦁ Simakou, T.; Butcher, J. P.; Reid, S.; Henriquez, F. L. ⦁ Alopecia
areata: A multifactorial autoimmune condition. J. Autoimmun. 2019,
98, 74−85.
⦁ Choi, Y. S.; Zhang, Y.; Xu, M.; Yang, Y.; Ito, M.; Peng, T.; Cui, Z.; Nagy, A.; Hadjantonakis, A. K.; Lang, R. A.; Cotsarelis, G.; Andl, T.; Morrisey, E. E.; Millar, S. E. ⦁ Distinct functions for Wnt/beta- ⦁ catenin in hair follicle stem cell proliferation and survival ⦁ and ⦁ interfollicular epidermal homeostasis. Cell stem cell 2013, 13 (6), 720−33.
⦁ Wang, X.; Chen, H.; Tian, R.; Zhang, Y.; Drutskaya, M. S.; Wang, C.; Ge, J.; Fan, Z.; Kong, D.; Wang, X.; Cai, T.; Zhou, Y.; Wang, J.; Wang, J.; Wang, S.; Qin, Z.; Jia, H.; Wu, Y.; Liu, J.; Nedospasov, S. A.; Tredget, E. E.; Lin, M.; Liu, J.; Jiang, Y.; Wu, Y. ⦁ Macrophages⦁ ⦁ induce⦁ ⦁ AKT/beta-catenin-dependent⦁ ⦁ Lgr5(+)⦁ ⦁ stem⦁ ⦁ cell ⦁ activation⦁ ⦁ and⦁ ⦁ hair⦁ ⦁ follicle⦁ ⦁ regeneration⦁ ⦁ through⦁ ⦁ TNF.⦁ Nat. Commun. 2017, 8, 14091.
⦁ Ali, N.; Zirak, B.; Rodriguez, R. S.; Pauli, M. L.; Truong, H. A.;
Lai, K.; Ahn, R.; Corbin, K.; Lowe, M. M.; Scharschmidt, T. C.; Taravati, K.; Tan, M. R.; Ricardo-Gonzalez, R. R.; Nosbaum, A.; Bertolini, M.; Liao, W.; Nestle, F. O.; Paus, R.; Cotsarelis, G.; Abbas,
A. K.; Rosenblum, M. D. Regulatory T Cells in Skin Facilitate Epithelial Stem Cell Differentiation. Cell 2017, 169 (6), 1119−1129.
⦁ Liu, X.; Zhang, P.; Zhang, X.; Li, X.; Bai, Y.; Ao, Y.; Hexig, B.; Guo, X.; Liu, D. ⦁ Fgf21⦁ ⦁ knockout⦁ ⦁ mice⦁ ⦁ generated⦁ ⦁ using⦁ ⦁ CRISPR/Cas9 ⦁ reveal⦁ ⦁ genetic⦁ ⦁ alterations⦁ ⦁ that⦁ ⦁ may⦁ ⦁ affect⦁ ⦁ hair⦁ ⦁ growth.⦁ Gene 2020, 733, 144242.
⦁ Bak, D. H.; Lee, E.; Lee, B. C.; Choi, M. J.; Kwon, T. R.; Hong, J.; Mun, S.-k.; Lee, K.; Kim, S.; Na, J.; Kim, B. J. ⦁ Boehmite⦁ ⦁ enhances ⦁ hair follicle growth via stimulation of dermal papilla cells⦁ ⦁ by ⦁ upregulating beta-catenin signalling. Experimental dermatology 2019, 28, 169.
⦁ Cervantes, J.; Perper, M.; Wong, L. L.; Eber, A. E.; Villasante
Fricke, A. C.; Wikramanayake, T. C.; Jimenez, J. J. Effectiveness of Platelet-Rich Plasma for Androgenetic Alopecia: A Review of the Literature. Skin appendage disorders 2018, 4 (1), 1−11.
⦁ Renert-Yuval, Y.; Guttman-Yassky, E. ⦁ The Changing Landscape ⦁ of⦁ ⦁ Alopecia⦁ ⦁ Areata:⦁ ⦁ The⦁ ⦁ Therapeutic⦁ ⦁ Paradigm.⦁ Adv. Ther. 2017, 34 (7), 1594−1609.
⦁ Browne, R.; Stewart, L.; Williams, H. C. ⦁ Is⦁ ⦁ methotrexate⦁ ⦁ an ⦁ effective⦁ ⦁ and⦁ ⦁ safe⦁ ⦁ treatment⦁ ⦁ for⦁ ⦁ maintaining⦁ ⦁ hair⦁ ⦁ regrowth⦁ ⦁ in⦁ ⦁ people ⦁ with alopecia totalis? A Critically Appraised Topic. Br. J. Dermatol. 2018, 179 (3), 609−614.
⦁ Guttman-Yassky, E.; Nia, J. K.; Hashim, P. W.; Mansouri, Y.; Alia, E.; Taliercio, M.; Desai, P. N.; Lebwohl, M. G. ⦁ Efficacy ⦁ and ⦁ safety of secukinumab treatment in adults with extensive alopecia ⦁ areata.⦁ Arch. Dermatol. Res. 2018, 310 (8), 607−614.
⦁ Murad, A.; Bergfeld, W. ⦁ 5-alpha-reductase⦁ ⦁ inhibitor⦁ ⦁ treatment ⦁ for frontal fibrosing alopecia: an evidence-based treatment update. Journal of the European Academy of Dermatology and Venereology: JEADV 2018, 32 (8), 1385−1390.
⦁ Esmat, S. M.; Hegazy, R. A.; Gawdat, H. I.; Abdel Hay, R. M.; Allam, R. S.; El Naggar, R.; Moneib, H. ⦁ Low⦁ ⦁ level⦁ ⦁ light-mino⦁ x⦁ idil⦁ ⦁ 5% ⦁ combination versus either therapeutic modality alone in management ⦁ of⦁ ⦁ female⦁ ⦁ patterned⦁ ⦁ hair⦁ ⦁ loss:⦁ ⦁ A⦁ ⦁ randomized⦁ ⦁ controlled⦁ ⦁ study.⦁ Lasers Surg. Med. 2017, 49 (9), 835−843.
⦁ Gentile, P.; Garcovich, S. ⦁ Advances⦁ ⦁ in⦁ ⦁ Regenerative⦁ ⦁ Stem⦁ ⦁ Cell ⦁ Therapy in Androgenic Alopecia and Hair Loss: Wnt pathway, ⦁ Growth-Factor, and Mesenchymal Stem Cell Signaling Impact ⦁ Analysis on Cell Growth and Hair Follicle Development. Cells 2019, 8 (5), 466.
⦁ Balana, M. E.; Charreau, H. E.; Leiros, G. J. ⦁ Epidermal⦁ ⦁ stem ⦁ cells and skin tissue engineering in hair follicle regeneration. World journal of stem cells 2015, 7 (4), 711−27.
⦁ Kageyama, T.; Yan, L.; Shimizu, A.; Maruo, S.; Fukuda, J. ⦁ Preparation⦁ ⦁ of⦁ ⦁ hair⦁ ⦁ beads⦁ ⦁ and⦁ ⦁ hair⦁ ⦁ follicle⦁ ⦁ germs⦁ ⦁ for⦁ ⦁ regenerative ⦁ medicine. Biomaterials 2019, 212, 55−63.
⦁ Kageyama, T.; Yoshimura, C.; Myasnikova, D.; Kataoka, K.; Nittami, T.; Maruo, S.; Fukuda, J. ⦁ Spontaneous⦁ ⦁ hair⦁ ⦁ follicle⦁ ⦁ germ ⦁ (HFG) formation in vitro, enabling the large-scale production⦁ ⦁ of ⦁ HFGs⦁ ⦁ for⦁ ⦁ regenerative⦁ ⦁ medicine.⦁ Biomaterials 2018, 154, 291−300.
⦁ Wang, J.; Miao, Y.; Huang, Y.; Lin, B.; Liu, X.; Xiao, S.; Du, L.; Hu, Z.; Xing, M. ⦁ Bottom-up⦁ ⦁ Nanoencapsulation⦁ ⦁ from⦁ ⦁ Single⦁ ⦁ Cells⦁ ⦁ to ⦁ Tunable and Scalable Cellular Spheroids for Hair Follicle Regener- ⦁ ation.⦁ Adv. Healthcare Mater. 2018, 7 (3), 1700447.
⦁ Abaci, H. E.; Coffman, A.; Doucet, Y.; Chen, J.; Jackow, J.; Wang, E.; Guo, Z.; Shin, J. U.; Jahoda, C. A.; Christiano, A. M. ⦁ Tissue ⦁ engineering of human hair follicles using a biomimetic ⦁ developmental ⦁ approach.⦁ Nat. Commun. 2018, 9 (1), 5301.
⦁ Liu, T.; Wang, Y.; Zhong, W.; Li, B.; Mequanint, K.; Luo, G.; Xing, M. ⦁ Biomedical Applications of Layer-by-Layer Self-Assembly ⦁ for ⦁ Cell Encapsulation: Current Status and Future Perspectives. Adv. Healthcare Mater. 2019, 8 (1), No. 1800939.
⦁ Lin, B.-j.; Wang, J.; Miao, Y.; Liu, Y.-q.; Jiang, W.; Fan, Z.-x.; Darabi, M.-A.; Hu, Z.-q.; Xing, M. ⦁ Cytokine loaded layer-by-layer ⦁ ultrathin matrices to deliver single dermal papilla cells for ⦁ spot-by-spot ⦁ hair follicle regeneration. J. Mater. Chem. B 2016, 4 (3), 489−504.UK 5099