HYDROGEL DRUG DELIVERY COMPOSITION

Information

  • Patent Application
  • 20220218571
  • Publication Number
    20220218571
  • Date Filed
    April 20, 2020
    4 years ago
  • Date Published
    July 14, 2022
    a year ago
Abstract
Disclosed is a hydrogel drug delivery composition comprising a therapeutic agent and a poloxamer, wherein the poloxamer is modified to be capable of adhering to a surface of bones or teeth, and wherein the poloxamer is modified by pyrophosphate, bisphosphonates, acidic peptides or tetracycline and its derivatives.
Description
FIELD

The present invention relates to the field of drug delivery systems, specifically to a hydrogel drug delivery composition.


BACKGROUND

Oral disease is a very common disease in human. In most cases, the treatment of oral diseases only requires local medication, not systemic medication. For some patients, systemic medication may cause serious side effects. Moreover, systemic medication can easily lead to drug resistance in the body, reducing the efficacy of drugs. In addition, the oral environment is complex, and many factors affect the health of the oral cavity. For example, long-term use of antibiotics will destroy the normal flora of the oral cavity. Researchers have been looking for a delivery system that can effectively release drugs in the oral cavity in a targeted manner.


Periodontal disease is a prevalent and chronic inflammatory condition that affects the surrounding tissues of the teeth including gingiva, periodontal ligament, and alveolar bone resulting in pocket formation, mobility, bone loss, and eventually lead to the loss of tooth. In addition to the pathogenic bacterial population, the host immune response, which aims at protecting host tissues from bacterial aggression, also acts as a mediator of the periodontal damage[1]. Current treatment strategies aim at reducing bacterial load by mechanical therapy and administration of antimicrobial agents[2, 3]. Additionally, host modulation agents have been utilized to ameliorate inflammation and prevent disease progression[4, 5]. Their efficacy on halting the alveolar bone loss, which is the major destructive and non-reversible hallmark of periodontal disease, is limited. Therefore, there is an unmet clinical need to develop novel therapies that would prevent bone erosion and regenerate the lost alveolar bone.


Glycogen synthase kinase 3 beta (GSK3β) is a multi-tasking serine/threonine kinase with crucial roles in several physiological processes including inflammation and bone homeostasis. It has been shown to play a critical role in the host inflammatory response[6-8] and bone homeostasis[9] as a negative regulator, suggesting that inhibitors of GSK3β may provide therapeutic effects for inflammatory and bone metabolic diseases[10]. Particularly, a GSK3β inhibitor (SB216763) has been studied in periodontal disease and data confirmed its therapeutic benefits in preventing alveolar bone loss associated with periodontal disease[11] During the last decade, several selective GSK-3β inhibitors have been synthesized and tested in clinical trials at various phases[12, 13]. In particular, 6-bromoindirubin-3′-oxime (BIO), a potent GSK3β inhibitor with an enzymatic IC50 of 5 nM, has exhibited anti-inflammatory[7, 8] and strong bone and teeth anabolic effects[14-20]. Due to the involvement of GSK3 in multiple physiological processes, however, systemic administration may cause serious adverse side effects (e.g., diarrhea, hypoglycemia, tumorigenesis)[21-23]. Hence, it is necessary to limit and restrict its biological action primarily at the intended site of action. Local delivery of BIO into the periodontal pocket will permit direct targeting of periodontal tissue, achieving high local concentrations along with minimizing systemic toxicities. However, the poor aqueous solubility and rapid clearance of BIO from the periodontal pockets post major challenges to its effective local delivery.


Statins, which were developed as 3-hydroxy-3-methylglutaryl (HMG)-CoA reductase inhibitors, have been widely used to treat cardiovascular diseases for decades. Mundy et al. reported that two statins, simvastatin (SIM) and lovastatin, have strong bone anabolic effects that were attributed to induction of the bone inducing factor bone morphogenic protein-2 (BMP-2). Later, statins are also noted for their anti-inflammatory effects[56]. Major efforts have been invested since then, attempting to validate this finding. The results remain controversial, however, in part due to the first-pass metabolism in the liver, which limits the amounts that reach bone when given orally and have prevented their clinical application to strengthen skeletal bones or treat periodontitis. Local applications of simvastatin are hindered by its poor water solubility (https://www.drugbank.ca/drugs/DB00641).


Thermoresponsive hydrogel formulations injected into the periodontal pocket would be a promising option for the local delivery of BIO or SIM to prevent the bony defects associated with periodontitis. Poloxamer 407 (Pluronic F127), a frequently used formulation excipient that has been approved by U.S. FDA for pharmaceutical applications, is a triblock amphiphilic copolymer consisting of a center block of polypropylene oxide flanked by two polyethylene oxide blocks (PEO101 PPO56 PEO101)[24-27]. It has been used extensively in controlled drug and cell delivery[28]. Its unique thermoresponsive gelation property in aqueous solutions (20-35% w/v) makes it an appealing carrier material for periodontal drug delivery and other similar applications[29, 30] The amphiphilic F127 polymer enhances the solubility of hydrophobic drugs at room temperature by forming micelles. When exposed to physiological temperature, the polymer solution forms a hydrogel, holding encapsulated drugs in its collapsed micellar structure to provide sustained release kinetics. F127 is also known to be non-toxic and biocompatible[30]. The constant flow of crevicular fluid, the poor bone-adhesion and mechanical properties of F127 hydrogel, however, would significantly limit the bioavailability of the payload drug in the periodontal pocket.


Therefore, there is a need to improve binding of hydrogel to the hard tissues and develop a novel hydrogel drug delivery system for effective drug release in topical site.


SUMMARY

In one aspect, the present disclosure provides a drug delivery composition comprising a therapeutic agent and a poloxamer, wherein the poloxamer is modified to be capable of adhering to a surface of bones or teeth.


Poloxamers are nonionic triblock copolymers composed of a central hydrophobic chain of polyoxypropylene (poly(propylene oxide)) flanked by two hydrophilic chains of polyoxyethylene (poly(ethylene oxide)). Preferably, the poloxamer is modified by pyrophosphate, bisphosphonate, dopamine, acidic peptide, or tetracycline and a derivative thereof.


More preferably, the poloxamer is a pyrophosphorylated poloxamer or a mixture of poloxamer and pyrophosphorylated poloxamer.


In the embodiments of the present disclosure, wherein


the poloxamer is selected from the group consisting of Pluronic L31, L35, F38, L42, L43, L44, L61, L62, L63, L64, P65, F68, L72, P75, F77, L81, P84, P85, F87, F88, L92, F98, L101, P103, P104, P105. F108, L121, L122, L123, F127, 10R5, 10R8, 12R3, 17R1, 17R2, 17R4, 17R8, 22R4, 25R1, 25R2, 25R4, 25R5, 25R8, 31R1, 31R2, 31R, and a mixture thereof; preferably, the poloxamer is poloxamer 407 (Pluronic F127);


the therapeutic agent includes anti-inflammatory compounds, bone anabolic compounds, bone antiresorptive compounds, anti-cancer compounds, statins, and antimicrobial compounds; preferably, the therapeutic agent is a GSK3-beta inhibitor or a statin.


The GSK3-beta inhibitors may be used in this invention are listed in Table 1.










TABLE 1





Inhibitor
Structure







NP 031112


embedded image







GSK-3 Inhibitor XV


embedded image







CHIR98014


embedded image







603281-31-8


embedded image







GSK-3 Inhibitor XXII, Compound A


embedded image







CT20026


embedded image







Ro318220


embedded image







Alsterpaullone


embedded image







AR28






6-Bromoindirubin- 3 oxime


embedded image







GSK-3 Inhibitor IX


embedded image







GSK-3 Inhibitor



XVI






CHIR99021


embedded image







Pyrazolopyridine 34


embedded image







Indirubin-3′-mono xime, 5-Iodo-


embedded image







Hymenialdisine


embedded image







6-Bromoindirubin- 3′-acetoxime


embedded image







CGP60474


embedded image







GSK-3 Inhibitor X


embedded image







CG9






Compound lA


embedded image







Compound 5a


embedded image







AzakenPaullone


embedded image







Pyrazolopyridine 18


embedded image







Indirubin-3-oxime


embedded image







Pyrazolopyridine 9


embedded image







Kenpaullone


embedded image







GSK-3 Inhibitor XIII


embedded image







GSK-3β Inhibitor XI


embedded image







GSK-3β Inhibitor XII, TWS119


embedded image







Compound 29


embedded image







GSK-3beta Inhibitor XXVI


embedded image







Compound 46


embedded image







GSK-3β Peptide
Myr-N-Gly-Lys-Glu-Ala-Pro-Pro-


Inhibitor,
Ala-Pro-Pro-Gln-Ser(PO3H)-Pro-


Cell-permeable
NH2





GSK3β Inhibitor XIX, IM-12


embedded image







AZD2858






Indirubin-3′-mono xime-5-sulphonic Acid


embedded image







Staurosporine


embedded image







GSK-3 Inhibitor IX, Control, MeBIO






ARA014418


embedded image







GSK-3β Inhibitor VIII


embedded image







SB216763


embedded image







SB415286


embedded image







GSK-3β Peptide
H-Lys-Glu-Ala-Pro-Pro-Ala-Pro-


Inhibitor
Pro-G1n-Ser(PO3H)-Pro-NH2





I5


embedded image







GF109203x


embedded image







Compound 8b


embedded image







SU9516


embedded image







GSK-3 Inhibitor II


embedded image







Flavopiridol


embedded image







GSK-3β Inhibitor VII


embedded image







Isogranulatimide


embedded image







Aloisine A


embedded image







TDZD8


embedded image







Aloisine B


embedded image







Compound 12


embedded image







Compound 17


embedded image







Compound 1


embedded image







GSK-3β Inhibitor VI


embedded image







Lithium






GSK-3β Inhibitor I


embedded image







Beryllium



Zinc







text missing or illegible when filed








In one embodiment of the present disclosure, wherein the GSK3-beta inhibitor is selected from the compounds listed in Table 1, such as indirubin and its derivatives, lithium, beryllium, zinc, and a mixture thereof; preferably, the GSK3-beta inhibitor is 6-bromoindirubin-3′-oxime, lithium, or zinc.


The statins may be used in this invention include atorvastatin (Lipitor), fluvastatin (Lescol, Lescol XL), lovastatin (Mevacor, Altoprev), pravastatin (Pravachol), rosuvastatin (Crestor), simvastatin (Zocor), and pitavastatin (Livalo).


Preferably, the composition is in a form of hydrogel and the hydrogel is thermosensitive.


In another aspect, the present disclosure provides a method of manufacturing the drug delivery composition comprising,

    • subjecting the poloxamer to tosylation reaction to synthesize tosylated poloxamer;
    • subjecting the tosylated poloxamer to pyrophosphorylation to obtain pyrophosphorylated poloxamer; and
    • dissolving the therapeutic agent in the pyrophosphorylated poloxamer to obtain a hydrogel containing the drug delivery composition.


In another aspect, the present disclosure provides a method of treating an oral disease comprising administering the drug delivery composition in situ at a topical site to a subject in need thereof.


Preferably, the topical site is a tooth and the oral disease is a dental disease, more preferably periodontitis. The composition is capable of releasing the therapeutic agent to treat the oral disease over a period of at least 2 days, and the composition is adherent only to the topical site.





BRIEF DESCRIPTION OF DRAWINGS


FIG. 1 shows the synthesis of pyrophosphorylated Pluronic F127 (F127-PPi).



FIG. 2 shows in vitro characterization of PF127 hydrogels. (a) Assessment of hydroxyapatite (HA) binding of 25% w/v PF127 hydrogels prepared with different ratio of F127-PPi and F127. **P<0.01, and ****P<0.0001 when compared to F127 hydrogel (ratio of 0); #P<0.05 and #P<0.01 when compared to ratio of 25 (one-way ANOVA with Tukey's multiple comparisons). (b) Solubility of BIO in PF127 vs F127 solutions at 4° C. **P<0.01 and ****P<0.0001 (one-way ANOVA with Tukey's multiple comparisons). (c) Cumulative BIO release from PF127 hydrogels at 37° C. (d) Erosion time of PF127 hydrogels incubated at 37° C. measured by weight remaining (%). Values are presented as the mean±SD, n=3.



FIG. 3 shows effect of different treatments on growth of MC3T3-E1 cells measured by CCK-8 assay following (a) 24 hr and (b) 48 hr exposure. Data are shown as the mean±SD. *P<0.05 when compared to control group. #P<0.05 when compared to BIO group. **P<0.01 when compared to control group. ##P<0.01 when compared to BIO group.



FIG. 4 shows physical properties of hydrogels. (a&d) Storage and loss modulus (G′ and G″) that determined gelation temperature (Tgel) values of PF127-BIO and PF127 hydrogels. The viscosity of (b) PF127 and (c) PF127-BIO hydrogels at different shear rates. The viscosity of (e) F127 and (f) F127-BIO hydrogels at different shear rates.



FIG. 5 shows the in vivo evaluation of PF127-BIO hydrogel's therapeutic efficacy in an experimental periodontitis rat model. (a) Micro-CT sagittal images showing the effect of different treatments on maxillary molar cementoenamel junction (CEJ) to alveolar bone crest (ABC). White vertical lines indicate ABC-CEJ distance. Scale bar=1 mm (b) Measurement of the linear distance between CEJ and ABC. (c-d) Quantitative analysis of alveolar bone quality after different treatments. Bone volume (BV), trabecular thickness (Tb.Th). Values are presented as the mean±SD. *P<0.05, **P<0.01, and ****P<0.0001 (one-way ANOVA with Tukey's multiple comparisons).



FIG. 6 shows histological analysis of papillary connective tissue and alveolar bone between first molar (M1) and second molar (M2) after three weeks of different treatments. (a) Representative images from different treatment groups of H&E stained tissues. Semi-quantitative assessment of (b) inflammatory cells and (c) osteoclasts. Values are presented as the mean±SD. *P<0.05, **P<0.01, ***P<0.001, and ****P<0.0001. #P<0.05, ##P<0.01, ###P<0.001, and ###190 P<0.0001 when compared to Con+ group. (one-way ANOVA). Scale bar=200 μm and 50 μm for 100× and 400×, respectively. Black arrows indicate alveolar bone and asterisks indicate first molar (M1).



FIG. 7 shows immunohistochemistry (IHC) analysis of β-catenin at the treatment area. (a) Representative images from different treatment groups of IHC staining of β-catenin of connective tissue above the alveolar bone crest (ABC) and between first molar (M1) and second molar (M2). (b) Semi-quantitative assessment of β-catenin positive cells interproximally between first molar (M1) and second molar (M2). Values are presented as the mean±SD. *P<0.05, **P<0.01, ***P<0.001, and ****P<0.0001. ###P<0.001 when compared to Con+ group. (one-way ANOVA). Scale bar=200 μm. Black arrows indicate alveolar bone and asterisks indicate first molar (M1) and second molar (M2).



FIG. 8 shows images of the phase transition of SIM-loaded PF127/F127 hydrogel.



FIG. 9 shows the in vivo evaluation of SIM-loaded PF127 hydrogel's therapeutic efficacy in an experimental periodontitis rat model. (a) Micro-CT sagittal images showing the effect of different treatments on maxillary molar cementoenamel junction (CEJ) to alveolar bone crest (ABC). White vertical lines indicate ABC-CEJ distance. Scale bar=1 mm (b) Measurement of the linear distance between CEJ and ABC. (c-d) Quantitative analysis of alveolar bone quality after different treatments. Bone volume (BV), trabecular thickness (Tb.Th). Values are presented as the mean±SD. *P<0.05, **P<0.01 (one-way ANOVA with Tukey's multiple comparisons).





DETAILED DESCRIPTION

The embodiments of the present disclosure will be described in detail below with reference to the embodiments. However, a person having ordinary skill in the art will understand that the following embodiments are merely to illustrate present disclosure and are not intended to limit the scope of the disclosure. For those embodiments in which specific conditions are not specified, they were carried out according to the conventional conditions or the conditions recommended by the manufacturer. For those used reagents or instruments of which the manufacturers are not indicated, they were all commercially available conventional products.


Example 1
Materials and Reagents

Pluronic F127 and pyrophosphate were purchased from Sigma-Aldrich (Saint Louis in Mo., USA). BIO was synthesized according to literature[31]. Dense Ceramic Hydroxyapatite discs (0.5″ diameter and 0.08″ Thick) were obtained from Clarkson Chromatography Products, Inc. (South Williamsport, Pa. USA). Mouse osteoblast MC3T3-L1 cells were acquired from ATCC (Manassas, Va., USA). Fetal bovine serum (FBS, BenchMark™) was acquired from Gemini BenchMark (West Sacramento, Calif.). Minimum Essential Media (alpha-MEM), and trypsin-EDTA were purchased from GIBCO (Grand Island, N.Y., USA). Cell Counting Kit-8 (CCK-8) was bought from Dojindo Molecular Technologies, Inc. (Rockville, Md. USA). All other solvents and reagents, if not specified, were acquired from either Acros Organics (Morris Plains, N.J., USA) or Fisher Scientific (Pittsburgh, Pa., USA).


Synthesis of Pyrophosphorylated Pluronic F127 (F127-PPi)
Synthesis of Tosylated Pluronic F127

Pluronic F127 (1.0 g, 0.079 mmol) and 4-toluenesulfonyl chloride (151 mg, 0.79 mmol) were dissolved in dry dichloromethane. Pyridine (62 μL, 0.79 mmol) was then added and the resulting solution was stirred for 24 hr at 21° C. After adding dichloromethane (80 mL), the resulting solution was washed with HCl (1 M, 20 mL) and brine (80 mL×2). After washing, the organic phase was separated and dried over MgSO4. The solution was filtered, concentrated and the residue was purified on a LH-20 column to produce 922 mg of the final product, yield: 90%. 1H NMR (CDCl3, 500 MHz) δ ppm 7.78 (d, J=3.0 Hz, 4H), 7.33 (d, J=3.0 Hz 4H), 4.14 (t, J=5.0 Hz, 4H), 3.77 (t, J=4.5 Hz, 8H), 3.59-3.55 (m, 872H), 3.50-3.44 (m, 142H), 3.38 (m, 65H), 2.44 (s, 6H), 1.12 (t, J=5.0 Hz, 195H); 13C NMR (CDCl3, 125 MHz) δ ppm 144.7, 133.0, 129.8, 127.9, 75.5, 75.4, 75.1, 73.4, 73.0, 72.9, 72.7, 69.2, 68.7, 21.6, 17.5, 17.3.


Synthesis of Pyrophosphorylated Pluronic F127

The tosylated Pluronic F127 (1.0 g, 0.077 mmol) and tris(tetra-n-butylammonium) hydrogen diphosphate [(n-Bu4N)3(HO)P2O6] (0.31 mmol, 280 mg) were dissolved in dry acetonitrile. The solution was stirred at 21° C. until the starting materials completely disappeared (˜3 hr, monitored by TLC). After removal of the solvents, the residue was dissolved in water (20 mL) and dialyzed (MWCO=12-14 kDa) against NaCl solution (0.1 mol/L) overnight to exchange tetrabutyl ammonium to sodium. The resulting solution was then dialyzed against distilled water to remove the excessive NaCl. The resulting solution was then lyophilized to obtain 859 mg of the final pyrophosphorylated Pluronic F127 (F127-PPi) product, yield: 85%. The synthesis path is shown in FIG. 1.



1H NMR (CDCl3, 500 MHz) δ ppm 4.16 (t, J=5.0 Hz, 4H), 3.78 (t, J=4.5 Hz, 8H), 3.59-3.54 (m, 872H), 3.50-3.44 (m, 142H), 3.39-3.36 (m, 65H), 1.13 (t, J=5.0 Hz, 195H); 13C NMR (CDCl3, 125 MHz) δ ppm 75.5, 75.4, 75.1, 73.4, 73.0, 72.9, 72.8, 72.7, 70.6, 17.5, 17.3; 31P NMR (202.5 MHz, CDCl3): δ (ppm)=−7.70 (d, J=20.2 Hz), −7.91 (d, J=20.2 Hz).


To determine the PPi content, F127-PPi was hydrolyzed in 1 M HCl for 2 hr at 100° C. to release the phosphate. After removal of F127 with chloroform extraction, an equal volume of 1 M HCl solution containing 0.5% (w/v) ammonium molybdate and 2% (w/v) ascorbic acid was added. The samples were incubated at 37° C. for 2 hr then their absorbance at 820 nm was measured using a UV spectrometer[49]. Eighty percent terminal hydroxyl groups of F127 was found to has been pyrophosphorylated.


Preparation of BIO-Loaded Thermoresponsive Hydrogel

PF127 hydrogel formulations with predetermined polymer concentrations (20, 25 and 30% w/v) were prepared by mixing F127-PPi and F127 at different ratio (0:100, 25:75, 50:50, 75:25, 100:0% w/w). Briefly, the desired amount of F127-PPi and F127 was dissolved in PBS (pH 7.4) with stiffing in an ice-water bath (˜4° C., to prevent gelation) until a clear solution (PF127) was obtained and then stored at 4° C. overnight. BIO was then dissolved in the polymer solutions by continuous stirring at 4° C., the obtained solutions were filtered through 0.8 μm filter syringes.


Binding Potential to Model Bone (Hydroxyapatite, or HA)

To assess the binding potential of the formulated hydrogels to hydroxyapatite or HA, which constitute the main inorganic component of bone and teeth, an in vitro binding study was done using HA discs. Briefly, polymer solutions (25% w/v) were prepared, containing 100 μM of BIO, with different ratios of F127-PPi and F127 (0:100, 25:75, 50:50, 75:25, and 100:0% w/w) to optimize binding affinity. Hydrogels (1 mL) were formed on HA disc placed in plastic wells at 37° C., and the hydrogels were allowed to stabilize for 15 min. After that, HA discs, on which hydrogels are formed, were inverted with temperature maintained at 37° C. by keeping the inverted hydrogels inside the water bath (37° C.). The binding time of hydrogel to HA disc was measured. The binding experiment was performed in triplicate. Based on this study, one formulation was selected for all subsequent experiments.


Evaluation of BIO's Solubility in PF127 and F127 Solutions

The aqueous solubility of BIO in PF127 and F127 solutions was assessed in this experiment. The solubility values of BIO were measured at dissolution equilibrium after adding the BIO to the PF127 and F127-containing media for a predetermined period of time, as reported previously[32]. Specifically, an excessive amount of BIO was added to different concentrations of PF127 or F127 solutions in microcentrifuge tubes. The suspensions were mixed on a rotor at 4° C. for 48 hr to achieve dissolution equilibrium. At 4° C., the suspensions were centrifuged at 2,000 rpm for 5 min to settle the undissolved drug to the bottom of the microcentrifuge tube. The supernatants were then filtered through 0.8 μm syringe filters to obtain the saturated BIO solutions. A UV SpectraMax M2 spectrophotometer (Molecular Devices, Sunnyvale, Calif., USA) was used to measure BIO concentrations at 260 nm.


In Vitro Release of BIO from Hydrogel


The release rate of the physically entrapped BIO from PF127 hydrogels (20, 25, and 30% w/v), was studied by a membrane-less experiment, as reported previously[33-35]. Briefly, samples of 1 mL of polymer solutions containing 0.5 mg of BIO were transferred into screw cap tubes and incubated in a water bath at 37° C. until the gels were formed. After gelation, 2 mL of phosphate-buffered saline (PBS; pH 7.4) containing 0.5% Tween 80 pre-equilibrated at 37° C. were gently laid over the surface of the hydrogels and incubated in a water bath at 37° C. with continuous gentle shaking. To measure the release of BIO, release medium (2 mL) were taken at regular time intervals and replaced with an equal volume of pre-equilibrated fresh releasing buffer. The concentration of BIO was determined by measuring the absorbance at 260 nm using a Molecular Device's Spectramax M2 (Sunnyvale, Calif., USA). The release study was performed in triplicate.


In Vitro Hydrogel Erosion

The erosion time of PF127 hydrogel formulations (20, 25, and 30% w/v) was determined by performing the weight remaining (%) experiment, as reported by others[33-35]. Briefly, samples of 1 mL of polymer solutions were transferred into screw cap tubes and incubated in a water bath at 37° C. until the gels were formed. After gelation, the original weight of the hydrogel samples was measured as (W0). Subsequently, 2 mL of PBS (pH 7.4) pre-equilibrated at 37° C. were gently laid over the surface of the hydrogels and incubated in a water bath at 37° C. with continuous gentle shaking. The weight of remaining hydrogels (Wt) was measured at regular time intervals after completely blotting off the buffer. Erosion study was performed in triplicate. Weight remaining (%) was calculated as:







Weight






remaining


(
%
)



=



(


W
0

-

W
t


)


W
0


×
100





Biocompatibility of PF127 Hydrogel

The impact of selected PF127 hydrogel formulation (25% w/v of mixed F127-PPi and F127, 50:50% w/w) and F127 hydrogel (25% w/v) (with or without BIO) on cell viability were assessed using CCK-8 assay. The BIO concentration in all BIO-containing formulations is 100 nM. Briefly, mouse preosteoblast MC3T3-L1 cell line was cultured in cell culture medium (alpha-MEM) with 10% (v/v) FBS and 1% (v/v) Penicillin/Streptomycin. The cells were incubated to 90% confluence under standard conditions at 37° C. in humidified atmosphere with 5% CO2. The hydrogels were extracted in alpha-MEM at 37° C. for 24 hr according to the ISO Standard 10993-12[34, 35]. The ratio between the surface area of the hydrogels and the volume of medium was 1.25 cm2/mL. Undiluted extracts were used for the assay. Cells were grown in 96 well plates (1×104 cells/well) and incubated for 24 hr at 37° C. The cells were then treated with either media only, PF127 extract, PF127-BIO extract, F127 extract, F127-BIO extract, or free BIO, and the plates were incubated at 37° C. for 24 and 48 hr. At each follow-up time point, CCK-8 reagent (10 μL) was added to each well and further incubated for 4 hr at 37° C. The absorbance was measured at 450 nm using a Molecular Device's Spectramax M2 microplate reader (Sunnyvale, Calif., USA).


Gelation and Viscosity Studies

Gelation temperature was determined by measuring the storage modulus (G′) and loss modulus (G″) of samples in the temperature sweep mode[36, 37]. Solutions with w/v concentrations ranging between 20 and 40% were prepared as described above. Each sample was uniformly loaded between the Peltier plate of the rheometer (TA Instruments AR1500ex) and a 60 mm-diameter, 1° cone geometry at 3° C. The linear viscoelastic region of each sample was pre-determined at 45° C. by using the oscillation amplitude function, and the following condition was chosen: strain of 0.1%, and angular frequency of 1 rad/s. Then, the G′ and G″ of each sample were measured for the temperature range from 3° C. to 45° C. by using the oscillation temperature sweep function of the rheometer (heating step: 1° C., soak time: 2 min for each temperature increase). The gelation temperature of each sample was determined as the temperature when G′ and G″ become equal.


The viscosity of gel phase was also investigated at constant temperature by flow sweep[30] (37° C., shear rate from 1 to 100 s−1) using the rheometer and a 25 mm-diameter parallel plate. Samples were dispensed on the Peltier plate of the rheometer, heated to 37° C. and maintained at 37° C. for 15 min to reach thermal stability before isothermally tested.


The Therapeutic Efficacy of BIO-Containing Hydrogels on a Rat Model of Experimental Periodontitis

Ten-month-old female Sprague Dawley rats (retired breeders, Envigo) were acclimated for one week prior to experiment. The animals (n=48) were randomly assigned into the following groups (Table 2): healthy control, experimental periodontitis (EP) treated with saline, EP treated with 25% w/v of mixed F127-PPi and F127 hydrogel (50:50% w/w, containing 100 μM of BIO) (PF127-BIO), EP treated with 25% w/v of F127 hydrogel (containing 100 μM of BIO) (F127-BIO), EP treated with 25% w/v of PF127 hydrogel, and EP treated with free BIO (100 μM). Using silk ligatures, the experimental periodontitis was induced as described previously[32, 38]. Briefly, rats were first anesthetized in an isoflurane chamber with a 1% to 4% isoflurane in 100% O2. After taking the body weight, the rats were positioned with a nose cone supplied with 0.5% to 2% isoflurane and 100% O2 to maintain anesthesia during the entire procedure. To induce experimental periodontitis, a 4-0 silk ligature was gently tightened subgingivally around the maxillary 2nd molars (M2). Following ligature placement, different treatments (10 μL) were locally delivered between the maxillary 1st molar (M1) and 2nd molar (M2) once each week for 3 weeks. After one week, the ligatures were removed. All animals were euthanized at week 4. The entire palate including all three molars was dissected and fixed in 10% formalin prior to micro-CT and histological analyses. All animals-related experiments were approved by the Institutional Animal Care and Use Committee (IACUC) of the University of Nebraska Medical Center (UNMC).














TABLE 2





Group
# of animals
Week 1
Week 2
Week 3
Week 4







Untreated
8



Euthanized


healthy control


(Con+)


EP/saline (Con−)
8
Ligatures,
Remove Ligatures,
Saline
Euthanized




Saline
Saline


EP/PF127
8
Ligatures,
Remove Ligatures,
PF127
Euthanized




PF127
PF127


EP/BIO
8
Ligatures,
Remove Ligatures,
BIO
Euthanized




BIO
BIO


EP/F127-BIO
8
Ligatures,
Remove, Ligatures,
F127-BIO
Euthanized




F127-BIO
F127-BIO


EP/PF127-BIO
8
Ligatures,
Remove Ligatures,
PF127-BIO
Euthanized




PF127-BIO
PF127-BIO





EP: Experimental Periodontitis.






Micro-Computed Tomography (μ-CT) Analysis of Periodontal Bone

All palate samples (including all three molars) were evaluated for alveolar bone quality using a high-resolution X-ray microtomography system (Skyscan 1172, Bruker, Kontich, Belgium), references to previous studies[32, 38]. The X-ray source was set as follows: 70 kV and 141 μA, resolution 12.9 μm, exposure time 1880 ms, and aluminum filter 0.5 mm-thick. To generate 3D images, scanning raw data were reconstructed using NRecon software. Sagittal sections were generated using CT-Analyzer software. To evaluate bone erosion, the distance from cementoenamel junction (CEJ) to alveolar bone crest (ABC) was determined using Skyscan Data viewer software. A longer distance from CEJ to ABC suggests more bone loss. For the analysis of histomorphometric parameters, such as bone volume (BV) and trabecular thickness (Tb.Th), a polygonal region of interest (ROI) between M1 and M2 was identified. The ROI was determined from the distopalatal of M1 to the mesiopalatal of M2 (length), 130 slices below the CEJ of M1 and M2 (height), and from the palatal side to the buccal side of M1 and M2 (width).


As described previously[39], femurs of these tested rats were also collected and scanned to assess the potential systemic effect of BIO. The X-ray source was set as the follows: voltage was 70 kV, current was 141 μA, exposure time was 700 ms, resolution was 8.6 μm, and aluminum filter was 0.5 mm-thick. The 3-D imagine were generated using the Skyscan NRecon and Skyscan DataViewer software. For bone quality analysis, a consistent polygonal ROI of trabecular bone at the distal femur was selected and the ROI was determined from 20 slices to 100 slices proximal to the growth plate. The bone histomorphometric parameters, including mean bone volume (BV), bone volume/tissue volume (BV/TV), trabecular thickness (Tb.Th) and bone mineral density (BMD) were determined using CT-Analyzer software.


Histological Analysis

At the completion of the μ-CT scanning, palates were decalcified for two weeks using 14% EDTA solution. Following decalcification, tissues were embedded in paraffin to obtain 4 μm thick sagittal sections. The slides were hematoxylin and eosin (H&E) stained for microscopic observation. To probe the presence of inflammatory cells between M1 and M2 and osteoclasts in the alveolar crest, a pathologist (SML) blind to experimental group assignment, evaluated the slides semi-quantitatively using an Olympus BX53 microscope. A semi-quantitative scoring system[32, 40, 41] was used to evaluate inflammatory cells, where 0 is negative, 1 is less than 30% of the affected tissues, 2 is some inflammatory cells (30-60%), and 3 is many inflammatory cells (>60%). Similarly, osteoclasts on the alveolar crest were evaluated using a semi-quantitative scoring system[32, 42] where 0 is negative, 1 is a few osteoclasts lining less than 5% of alveolar bone surface, 2 is some osteoclasts (5-25%), and 3 is many osteoclasts (25-50%). Immunohistochemical (IHC) staining of β-catenin was performed using primary antibody (rabbit monoclonal anti-β-catenin antibody, Abcam, ab32572; 1:400 dilution). After deparaffinization and rehydration, sections were incubated in citrate buffer (pH=6.0, 0.1 M) for antigen retrieval, washed, and then incubated in hydrogen peroxide. Sections were then blocked and incubated with the primary antibody, followed by incubation with the secondary antibody. The antibody complexes were visualized using the DAB chromogen. Hematoxylin was used for counterstaining. The staining intensity was independently evaluated by a pathologist (SML) using a scale of from 0 to 3, where 0 is negative, 1 is weak staining, 2 is moderate staining, and 3 is strong staining[19, 43]. SML's scoring was calibrated by another examiner (YA) who evaluated a stack of slides and then compared with the SML's scores.


Statistical Analysis

All the generated data were expressed as the mean±SD (standard deviation). Statistical analyses were carried out using Prism 8.0 software (GraphPad, San Diego, Calif.). The Analysis of Variance (ANOVA) was used to analyze continuous outcomes among more than three groups. Tukey's pairwise post-hoc testing was performed for multiple comparisons. β-value<0.05 was considered statistically significant.


Results
Hydrogel Binding to Hydroxyapatite (HA)

The in vitro binding of the PF127 hydrogel (25% w/v) to HA was analyzed to predict their affinity to bone in vivo. As shown in FIG. 2a, the binding time increased as the F127-PPi content was increased in the hydrogel. Among different ratios of F127-PPi, 50, 75, and 100 (w/w %) showed the longest binding time to HA disc (24.6, 25.6, and 28.2 min), respectively. Their binding time was statistically significant when compared to F127 hydrogel (P<0.0001). The ratio of 25 (w/w %) also exhibited considerable binding time (18.2 min) which was significantly higher than F127 hydrogel (P<0.01). F127 hydrogel had the lowest binding time (8.5 min) among all the formulations. Based on this observation, the formulation of 50:50% w/w ratio of F127-PPi and F127 was used in the subsequent experiments as it showed strong binding affinity with relatively low PPi content.


Solubility of BIO in PF127 vs F127

The solubility of BIO in PF127 and F127-containing solutions of different concentrations was analyzed and compared at 4° C. with solution pH=7. As shown in FIG. 2b, the solubility of BIO appears to be proportionally dependent upon polymer concentration. The solubility of BIO in 20% PF127 was determined to be 0.5 mg/mL when compared to the 0.1 mg/mL in 20% F127. Furthermore, when polymer concentration was increased to 25%, BIO's solubility in PF127 was greatly improved to 2 mg/mL, while the value was only improved to 0.3 mg/mL in F127 solution. When concentration further increased to 30%, solubility in PF127 continued to increase to 3 mg/mL, while in F127 improved to 0.6 mg/mL. Clearly, PF127 can significantly improve BIO's aqueous solubility.


In Vitro Release of BIO from PF127 Hydrogel


The releasing kinetics of BIO from PF127 hydrogels (20, 25, and 30% w/v) was studied at 37° C. As shown in FIG. 2c, the releasing kinetics of BIO from the hydrogels were found to be dependent upon the concentration of polymers: the higher the polymer concentration, the slower the release rate. The total release (˜99%) of BIO from 20% w/v hydrogel occurred in 24 hr with burst release (˜50%) in the first 3 hr. BIO release from 25% w/v hydrogel was shown with a burst release in the first 12 hr followed by a sustained release in the next 48 hr. For 30% w/v hydrogel, the release of BIO was sustained for over 48 hr.


In Vitro Hydrogel Erosion

The PF127 hydrogels erosion behavior was characterized by measuring weight remaining (%) at regular incubation time intervals. The results correlate with the release study and shown to be a function of the concentration of hydrogel. As shown in FIG. 2d, the hydrogel with 20% w/v concentration was completely eroded within 24 hr, ˜50% was eroded in the first 3 hr. While 25% w/v hydrogel was completed eroded in 48 hr, ˜60% weight loss was observed in the first 6 hr (burst erosion) followed by sustained erosion. Hydrogel with 30% w/v concentration was completely eroded in 48 hr in a sustained erosion manner.


Biocompatibility of PF127-Based Hydrogel

To assess the safety of PF127 hydrogel comparing to F127 which is known to be biocompatible, mouse osteoblast MC3T3-L1 cells were treated with PF127, PF127-BIO, F127 and F127-BIO culture media extracts, or free BIO (100 nM) for 24 hr and 48 hr, and cell viability was defined using the CCK-8 assay (FIG. 3). The viability percentage of MC3T3-L1 cells treated with PF127 25% w/v, when compared to media-treated control, was slightly reduced to 84.5% after 24 hr and further decreased to 79.5% after 48 hr. The viability of MC3T3-L1 cells treated with F127 extract was 87.5% and 83.5% after 24 and 48 hr, respectively. When BIO was added to hydrogels, cell viability was not changed significantly. However, when cells treated with 100 nM of free BIO, the viability was determined to be 102% and 104.5% comparing to media-treated control.


Gelation and Viscosity Analysis of the PF127 Hydrogel

Gelation temperature (Tgel) was determined based on the temperature sweep measurements of storage and loss modulus, G′ and G″ (FIG. 4). Samples showed fluid-like behaviors at low temperature (T<Tgel) because G′ was much smaller than G″. As samples were heated, both G′ and G″ increased sharply near Tgel, and G′ became larger than G″ at high temperature (T>Tgel), which shows that samples underwent gelation and demonstrated solid-like behaviors. Therefore, Tgel was determined according to the cross-point between the G′ and G″ curves, and Table 3 summarizes the Tgel values obtained.













TABLE 3





Concentration






(w/v)
F127
PF127
F127-BIO
PF127-BIO







20%
18.0
19.0
19.0
19.0


25%
16.0
15.0
14.0
14.0


30%
13.0
12.1
11.1
11.0









Viscosity measurements were also performed on PF127 and F127 hydrogels with and without BIO at 37° C. to study the effect of the pyrophosphorolation and drug content on their viscous property (FIG. 4). All hydrogels showed a typical shear-thinning behavior (non-Newtonian) and the viscosity decreased as a function of shear rate. No notable differences in viscosity were observed between PF127 and F127 hydrogels within the same concentration, except 20% w/v PF127 which showed low viscosity and forms a very weak hydrogel. BIO addition seemed to have no significant impact on hydrogels' viscous property.


Micro-Computed Tomography (μ-CT) Analysis of the Periodontal Bone

As presented in FIG. 5a, it is evident that the PF127-BIO hydrogel prevented alveolar bone loss compared to the other treatments. The linear measurement (CEJ to ABC) indicated that the PF127-BIO hydrogel-treated group had the shortest distance among all other treated groups. While F127-BIO treated group exhibited statistically significant difference only when compared to the saline-treated group, free BIO-treated group did not show a statistically significant difference comparing to the saline-treated group (FIG. 5b). Bone volume (BV, FIG. 5c) and trabecular thickness (Tb.Th, FIG. 5d) were also quantified to further validate bone quality. The value of BV for the PF127-BIO treated group was significantly higher when compared to all the other treated groups. In contrast, F127-BIO treated group exhibited a statistically significant difference only when compared to the saline-treated group. The values of Tb.Th for the PF127-BIO and F127-BIO treated groups were significantly higher when compared to the saline group. No statistically significant difference between Free BIO treated group and the saline group was observed in the above μ-CT parameters. 3-D movies showing the periodontal bone quality after different treatments can be found in Supporting Information. Femur analyses data did not show any significant difference among all treatment groups.


Histological Evaluation

To study the impacts of different treatment on inflammatory cells, osteoclasts, and β-catenin, images of stained tissue sections (FIGS. 5a, 6a) were assessed semi-quantitatively. Histology scores of inflammatory cells were shown in FIG. 6b. As expected, healthy Con+ group has a score of 0 for both inflammatory cells and osteoclasts, which is significantly lower than all the treatment groups except PF127-BIO group. There is no statistically significant difference between healthy Con+ and PF127-BIO group. PF127-BIO hydrogel-treated group has the lowest inflammatory cell score when compared to all the other groups (P<0.0001 compared to saline Con− group, P<0.01 compared to F127-BIO and free BIO treated groups), while F127-BIO and free BIO treated groups exhibited a statistically significant difference (P<0.01) only when compared to the saline Con− group. PF127-BIO hydrogel treated group had the lowest osteoclast score comparing to all treatment groups except healthy Con+ group and was significantly (P<0.001) lower than that of the saline Con− group (FIG. 6c). Scores of F127-BIO and free BIO treated groups were significantly lower than the saline Con− group. β-Catenin positive cell score of the PF127-BIO hydrogel-treated group was the highest among all treatment groups (FIG. 7b, P<0.0001 compared to saline Con− group, P<0.05 compared to F127-BIO and free BIO treated groups, P<0.001 when compared to healthy Con+ group). Scores of F127-BIO and free BIO-treated groups were not significantly different (P>0.05) from the saline Con−, PF127, and healthy Con+ group.


Example 2

In this example, simvastatin (SIM) was used as the therapeutic agent. The preparation of PF127 was described in Example 1.


Preparation of SIM-Loaded PF127/F127 Hydrogel

Film hydration method: An excess amount of SIM was dissolved in methanol solution of PF127/F127. Methanol was evaporated to form a film and then on the following day hydration was done with iced water for 30 min to prevent formation of gel. Free drug was removed after centrifugation.


Solvent evaporation method: PF127/F127 and SIM were dissolved in acetone and then added dropwise to iced water with stirring. Acetone was evaporated overnight. Free drug was removed after centrifugation.


Direct dissolution method: Iced PF127/F127 solution was prepared first and an excess amount of SIM was added and mixed for 48 hr. Free drug was removed after centrifugation.


Results are shown in Table 4. Based upon this result direct dissolution method was used in the preparation of the SIM-loaded PF127/F127 hydrogel.












TABLE 4





Preparation Method
SIM EE (%)
Particle Size (nm)
PDI


















Direct Dissolution
40.3 ± 5.5
34.32
0.147


Solvent Evaporation
36.2 ± 5.5
53.4
0.136


Film Hydration
66.5 ± 4.8
20.62
0.335









Phase transition of SIM-loaded PF127/F127 hydrogel is shown in FIG. 8. The result suggests that the pyrophosphorylation of F127 does not affect its thermoresponsive behavior.


Animal Study on Therapeutic Efficacy in Periodontitis

The experimental process is described in Example 1. The grouping and treatment are shown in Table 4.


















Group
Number
Side
Week 0
Week 1
Week 2
Week 3







1
5
Left
Place ligature,
Remove
F127 treatment
Euthanasia





F127 treatment
ligature, F127






treatment




Right
Place ligature,
Remove
SIM-loaded





SIM-loaded
ligature,
F127 treatment





F127 treatment
SIM-loaded






F127 treatment


2
5
Left
Place ligature,
Remove
Free SIM
Euthanasia





Free SIM
ligature, Free
treatment





treatment
SIM treatment




Right
Place ligature,
Remove
SIM-loaded





SIM-loaded
ligature,
PF127 treatment





PF127 treatment
SIM-loaded






PF127 treatment


3
5
Left
Place ligature,
Remove
PF127 treatment
Euthanasia





PF127 treatment
ligature, PF127






treatment




Right
Place ligature,
Remove
Saline treatment





Saline treatment
ligature, Saline






treatment









As shown in FIG. 9, SIM-loaded PF127 formulation is the only formulation that can significantly reduce the CEJ-ABC distance when compared to saline control. In addition, the treatment can also significantly increase the periodontal bone volume (BV) and trabecular thickness (Tb.Th).


REFERENCES



  • [1] Pihlstrom B L, Michalowicz B S, Johnson N W. Periodontal diseases. The Lancet. 2005 November; 366(9499):1809-1820.

  • [2] Herrera D, Sanz M, Jepsen S, Needleman I, Roldán S. A systematic review on the effect of systemic antimicrobials as an adjunct to scaling and root planing in periodontitis patients. Journal of Clinical Periodontology. 2002 December; 29: 136-159.

  • [3] Greenstein G. Local drug delivery in the treatment of periodontal diseases: assessing the clinical significance of the results. Journal of Periodontology. 2006 April; 77(4):565-578.

  • [4] Williams R C, Jeffcoat M K, Howell T H, Reddy M S, Johnson H G, Hall C M, Goldhaber P. Ibuprofen: an inhibitor of alveolar bone resorption in beagles. Journal of Periodontal Research. 1988 July; 23 (4): 225-229.

  • [5] Raja S, Byakod G, Pudakalkatti P. Growth factors in periodontal regeneration. International Journal of Dental Hygiene. 2009 May; 7(2):82-89.

  • [6] Wang H, Brown J, Martin M. Glycogen synthase kinase 3: a point of convergence for the host inflammatory response. Cytokine. 2011 February; 53(2):130-140.

  • [7] Huang W C, Lin Y S, Wang C Y, Tsai C C, Tseng H C, Chen C L, Lu P J, Chen P S, Qian L, Hong J S, Lin C F. Glycogen synthase kinase-3 negatively regulates anti-inflammatory interleukin-10 for lipopolysaccharide-induced iNOS/NO biosynthesis and RANTES production in microglial cells. Immunology. 2009 September; 128(1 Pt 2):e275-e286.

  • [8] Whittle B J, Varga C, Pósa A, Molnár A, Collin M, Thiemermann C. Reduction of experimental colitis in the rat by inhibitors of glycogen synthase kinase-3β. British Journal of Pharmacology. 2006 March; 147(5):575-582.

  • [9] Arioka M, Takahashi-Yanaga F, Sasaki M, Yoshihara T, Morimoto S, Hirata M, Mori Y, Sasaguri T. Acceleration of bone regeneration by local application of lithium: Wnt signal-mediated osteoblastogenesis and Wnt signal-independent suppression of osteoclastogenesis. Biochemical Pharmacology. 2014 August; 90(4):397-405.

  • [10] Martinez A, Castro A, Dorronsoro I, Alonso M. Glycogen synthase kinase 3 (GSK-3) inhibitors as new promising drugs for diabetes, neurodegeneration, cancer, and inflammation. Medicinal Research Reviews. 2002 July; 22(4):373-384.

  • [11] Adamowicz K, Wang H, Jotwani R, Zeller I, Potempa J, Scott D A Inhibition of GSK3 abolishes bacterial-induced periodontal bone loss in mice. Molecular Medicine. 2012 August; 18(8):1190-1196.

  • [12] Xie J, Méndez J D, Méndez-Valenzuela V, Aguilar-Hernández M M. Cellular signalling of the receptor for advanced glycation end products (RAGE). Cellular Signaling. 2013 November; 25(11):2185-2197.

  • [13] Sanguineti R, Storace D, Monacelli F, Federici A, Odetti P. Pentosidine effects on human osteoblasts in vitro. Annals of the New York Academy of Sciences. 2008 April; 1126(1):166-172.

  • [14] Wang F S, Ko J Y, Weng L H, Yeh D W, Ke H J, Wu S L Inhibition of glycogen synthase kinase-3β attenuates glucocorticoid-induced bone loss. Life Sciences. 2009 November; 85(19-20):685-692.

  • [15] Krause U, Harris S, Green A, Ylostalo J, Zeitouni S, Lee N, Gregory C A. Pharmaceutical modulation of canonical Wnt signaling in multipotent stromal cells for improved osteoinductive therapy. Proceedings of the National Academy of Sciences. 2010 March; 107(9):4147-4152.

  • [16] Fukuda T, Kokabu S, Ohte S, Sasanuma H, Kanomata K, Yoneyama K, Kato H, Akita M, Oda H, Katagiri T. Canonical Wnts and BMPs cooperatively induce osteoblastic differentiation through a GSK3β-dependent and β-catenin-independent mechanism. Differentiation. 2010 July; 80(1):46-52.

  • [17] Piters E, Boudin E, Van Hul W. Wnt signaling: a win for bone. Archives of Biochemistry and Biophysics. 2008 May; 473(2):112-116.

  • [18] Kwon Y J, Yoon C H, Lee S W, Park Y B, Lee S K, Park M C Inhibition of glycogen synthase kinase-3β suppresses inflammatory responses in rheumatoid arthritis fibroblast-like synoviocytes and collagen-induced arthritis. Joint Bone Spine. 2014 May; 81(3):240-246.

  • [19] Jiang Y Y, Wen J, Gong C, Lin S, Zhang C X, Chen S, Cheng W, Li H. BIO alleviated compressive mechanical force-mediated mandibular cartilage pathological changes through Wnt/β-catenin signaling activation. Journal of Orthopaedic Research. 2018 April; 36(4):1228-1237.

  • [20] Neves V C, Babb R, Chandrasekaran D, Sharpe P T. Promotion of natural tooth repair by small molecule GSK3 antagonists. Scientific Reports. 2017 January; 7:39654.

  • [21] Takahashi-Yanaga F. Activator or inhibitor? GSK-3 as a new drug target. Biochemical Pharmacology. 2013 July; 86(2):191-199.

  • [22] Kahn M. Can we safely target the WNT pathway? Nature Reviews Drug Discovery. 2014 July; 13(7):513-532.

  • [23] Huang P, Yan R, Zhang X, Wang L, Ke X, Qu Y. Activating Wnt/β-catenin signaling pathway for disease therapy: Challenges and opportunities. Pharmacology & Therapeutics. 2019 April; 196:79-90.

  • [24] Giuliano E, Paolino D, Fresta M, Cosco D. Mucosal Applications of Poloxamer 407-Based Hydrogels: An Overview. Pharmaceutics. 2018 September; 10(3):E159.

  • [25] Dumortier G, Grossiord J L, Agnely F, Chaumeil J C. A review of poloxamer 407 pharmaceutical and pharmacological characteristics. Pharmaceutical Research. 2006 December; 23(12):2709-2728.

  • [26] Fakhar-ud-Din, Khan G M. Development and characterisation of levosulpiride-loaded suppositories with improved bioavailability in vivo. Pharmaceutical Development and Technology. 2019 January; 24(1):63-69.

  • [27] Monti D, Burgalassi S, Rossato M S, Albertini B, Passerini N, Rodriguez L, Chetoni P. Poloxamer 407 microspheres for orotransmucosal drug delivery. Part II: In vitro/in vivo evaluation. International Journal of Pharmaceutics. 2010 November; 400(1-2):32-36.

  • [28] Rangabhatla A S, Tantishaiyakul V, Boonrat O, Hirun N, Ouiyangkul P. Novel in situ mucoadhesive gels based on Pluronic F127 and xyloglucan containing metronidazole for treatment of periodontal disease. Iranian Polymer Journal. 2017 November; 26(11):851-859.

  • [29] Akash M S, Rehman K, Li N, Gao J Q, Sun H, Chen S. Sustained delivery of IL-1Ra from pluronic F127-based thermosensitive gel prolongs its therapeutic potentials. Pharmaceutical Research. 2012 Dec. 1:29(12):3475-3485.

  • [30] Diniz I M, Chen C, Xu X, Ansari S, Zadeh H H, Marques M M, Shi S, Moshaverinia A. Pluronic F-127 hydrogel as a promising scaffold for encapsulation of dental-derived mesenchymal stem cells. Journal of Materials Science: Materials in Medicine. 2015 March; 26(3):153.

  • [31] Polychronopoulos P, Magiatis P, Skaltsounis A L, Myrianthopoulos V, Mikros E, Tarricone A, Musacchio A, Roe S M, Pearl L, Leost M, Greengard P. Structural basis for the synthesis of indirubins as potent and selective inhibitors of glycogen synthase kinase-3 and cyclin-dependent kinases Journal of Medicinal Chemistry. 2004 February; 47(4):935-946.

  • [32] Wang X, Jia Z, Almoshari Y, Lele S M, Reinhardt R A, Wang D. Local application of pyrophosphorylated simvastatin prevents experimental periodontitis. Pharmaceutical Research. 2018 August; 35(8):164.

  • [33] Deshmukh M, Singh Y, Gunaseelan S, Gao D, Stein S, Sinko P J. Biodegradable poly (ethylene glycol) hydrogels based on a self-elimination degradation mechanism. Biomaterials. 2010 September; 31(26):6675-6684.

  • [34] Li Y, Cao J, Han S, Liang Y, Zhang T, Zhao H, Wang L, Sun Y. ECM based injectable thermo-sensitive hydrogel on the recovery of injured cartilage induced by osteoarthritis. Artificial Cells, Nanomedicine, and Biotechnology. 2018 November; 46(sup2):152-160.

  • [35] Ma X, He Z, Han F, Zhong Z, Chen L, Li B. Preparation of collagen/hydroxyapatite/alendronate hybrid hydrogels as potential scaffolds for bone regeneration. Colloids and Surfaces B: Biointerfaces. 2016 July; 143:81-87.

  • [36] Winter H H. Can the gel point of a cross-linking polymer be detected by the G′-G″ crossover? Polymer Engineering and Science. 1987 December; 27(22):1698-1702.

  • [37] Bercea M, Darie R N, Nita L E, Morariu S. Temperature Responsive Gels Based on Pluronic F127 and Poly(vinyl alcohol). Industrial & Engineering Chemistry Research. 2011 March; 50(7):4199-4206.

  • [38] Bradley A D, Zhang Y, Jia Z, Zhao G, Wang X, Pranke L, Schmid M J, Wang D, Reinhardt R A. Effect of simvastatin prodrug on experimental periodontitis. Journal of Periodontology. 2016 May; 87(5):577-582.

  • [39] Jia Z, Wang X, Wei X, Zhao G, Foster K W, Qiu F, Gao Y, Yuan F, Yu F, Thiele G M, Bronich T K, O'Dell, J R, Wang D. Micelle-Forming Dexamethasone Prodrug Attenuates Nephritis in Lupus-Prone Mice without Apparent Glucocorticoid Side Effects. ACS Nano. 2018 July; 12(8):7663-7681.

  • [40] Gibson-Corley K N, Olivier A K, Meyerholz D K. Principles for valid histopathologic scoring in research. Veterinary Pathology. 2013 November; 50(6):1007-1015.

  • [41] Schett G, Stolina M, Bolon B, Middleton S, Adlam M, Brown H, Zhu L, Feige U, Zack D J. Analysis of the kinetics of osteoclastogenesis in arthritic rats. Arthritis & Rheumatism: Official Journal of the American College of Rheumatology. 2005 October; 52(10):3192-3201.

  • [42] Bolon B, Morony S, Cheng Y, Hu Y L, Feige U. Osteoclast numbers in Lewis rats with adjuvant-induced arthritis: identification of preferred sites and parameters for rapid quantitative analysis. Veterinary Pathology. 2004 January; 41(1):30-36.

  • [43] Song S, Ajani J A, Honjo S, Maru D M, Chen Q, Scott A W, Heallen T R, Xiao L, Hofstetter W L, Weston B, Lee J H. Hippo coactivator YAP1 upregulates SOX9 and endows esophageal cancer cells with stem-like properties. Cancer Research. 2014 August; 74(15):4170-4182.

  • [44] Georgiou K R, King T J, Scherer M A, Zhou H, Foster B K, Xian C J. Attenuated Wnt/β-catenin signalling mediates methotrexate chemotherapy-induced bone loss and marrow adiposity in rats. Bone. 2012 June; 50(6):1223-1233.

  • [45] Park D W, Jiang S, Liu Y, Siegal G P, Inoki K, Abraham E, Zmijewski J W. GSK3β-dependent inhibition of AMPK potentiates activation of neutrophils and macrophages and enhances severity of acute lung injury. American Journal of Physiology-Lung Cellular and Molecular Physiology. 2014 September; 307(10):L735-L745.

  • [46] Wang Y, Huang W C, Wang C Y, Tsai C C, Chen C L, Chang Y T, Kai J I, Lin C F Inhibiting glycogen synthase kinase-3 reduces endotoxaemic acute renal failure by down-regulating inflammation and renal cell apoptosis. British Journal of Pharmacology. 2009 July; 157(6):1004-1013.

  • [47] Klamer G, Shen S, Song E, Rice A M, Knight R, Lindeman R, O'Brien T A, Dolnikov A. GSK3 inhibition prevents lethal GVHD in mice. Experimental Hematology. 2013 January; 41(1):39-55.

  • [48] Bai X, Gao M, Syed S, Zhuang J, Xu X, Zhang X Q. Bioactive hydrogels for bone regeneration. Bioactive Materials. 2018 December; 3(4):401-417.

  • [49] Chen F, Jia Z, Rice K C, Reinhardt R A, Bayles K W, Wang D. The development of dentotropic micelles with biodegradable tooth-binding moieties. Pharmaceutical Research. 2013 November; 30(11):2808-2817.

  • [50] Wang D, Miller S C, Kopečkovà P, Kopeček J. Bone-targeting macromolecular therapeutics. Advanced Drug Delivery Reviews. 2005 May; 57(7):1049-1076.

  • [51] Purcell P M, Boyd I W. Bisphosphonates and osteonecrosis of the jaw. Medical Journal of Australia. 2005 April; 182(8):417-418.

  • [52] Kennel K A, Drake M T. Adverse effects of bisphosphonates: implications for osteoporosis management. Mayo Clinic Proceedings 2009 July; 84(7):632-638.

  • [53] Shellis R P, Addy M, Rees G D. In vitro studies on the effect of sodium tripolyphosphate on the interactions of stain and salivary protein with hydroxyapatite. Journal of dentistry. 2005 Apr. 1; 33(4):313-24.

  • [54] Sharma P K, Bhatia S R. Effect of anti-inflammatories on Pluronic® F127: micellar assembly, gelation and partitioning. International Journal of Pharmaceutics. 2004 July; 278(2):361-377.

  • [55] Low S A, Galliford C V, Yang J, Low P S, Kopeček J. Biodistribution of fracture-targeted gsk3β inhibitor-loaded micelles for improved fracture healing. Biomacromolecules. 2015 September; 16(10):3145-3153.

  • [56] Jain, M. K. & Ridker, P. M. Anti-Inflammatory Effects of Statins: Clinical Evidence and Basic Mechanisms. Nature Reviews Drug Discovery, 2005.4, p. 977-87.


Claims
  • 1-17. (canceled)
  • 18. A drug delivery composition comprising a therapeutic agent and a poloxamer, wherein the poloxamer is modified to be capable of adhering to a surface of bones or teeth.
  • 19. The drug delivery composition of claim 18, wherein the poloxamer is modified by pyrophosphate, bisphosphonate, dopamine, acidic peptide, or tetracycline, or a derivative thereof.
  • 20. The drug delivery composition of claim 19, wherein the poloxamer is a pyrophosphorylated poloxamer or a mixture of a poloxamer with its pyrophosphorylated derivative.
  • 21. The drug delivery composition of claim 18, wherein the poloxamer is selected from Pluronic L31, L35, F38, L42, L43, L44, L61, L62, L63, L64, P65, F68, L72, P75, F77, L81, P84, P85, F87, F88, L92, F98, L101, P103, P104, P105, F108, L121, L122, L123, F127, 10R5, 10R8, 12R3, 17R1, 17R2, 17R4, 17R8, 22R4, 25R1, 25R2, 25R4, 25R5, 25R8, 31R1, 31R2, 31R, and a mixture thereof.
  • 22. The drug delivery composition of claim 18, wherein the poloxamer is poloxamer 407 (Pluronic F127).
  • 23. The drug delivery composition of claim 18, wherein the therapeutic agent is selected from anti-inflammatory compounds, bone anabolic compounds, bone antiresorptive compounds, anti-cancer compounds, statins, and antimicrobial compounds.
  • 24. The drug delivery composition of claim 18, wherein the therapeutic agent is a GSK3-beta inhibitor or a statin.
  • 25. The drug delivery composition of claim 24, wherein the GSK3-beta inhibitor is selected from indirubin and derivatives thereof, lithium, zinc, beryllium, and mixtures thereof.
  • 26. The drug delivery composition of claim 24, wherein the GSK3-beta inhibitor is 6-bromoindirubin-3′-oxime, lithium, or zinc.
  • 27. The drug delivery composition of claim 24, wherein the statin is selected from atorvastatin, fluvastatin, lovastatin, pravastatin, rosuvastatin, simvastatin, pitavastatin, and mixtures thereof.
  • 28. The drug delivery composition of claim 24, wherein the statin is simvastatin.
  • 29. The drug delivery composition of claim 18, wherein the composition is in a form of thermoresponsive hydrogel.
  • 30. The drug delivery composition of claim 18, wherein the composition is capable of adhering to the surface of teeth and remaining in periodontal pocket for at least 2 days.
  • 31. A method of manufacturing the drug delivery composition of claim 18, the method comprising: subjecting the poloxamer to a tosylation reaction to synthesize a tosylated poloxamer;subjecting the tosylated poloxamer to pyrophosphorylation to obtain a pyrophosphorylated poloxamer; anddissolving the therapeutic agent in the pyrophosphorylated poloxamer to obtain a drug delivery composition.
  • 32. The method according to claim 31, wherein the poloxamer is poloxamer 407 (Pluronic F127), and the therapeutic agent is a GSK3-beta inhibitor or a statin.
  • 33. The method according to claim 31, wherein the drug delivery composition is an injectable solution, which can form a hydrogel upon in vivo instillation.
  • 34. A method of treating an oral disease or condition, comprising administering the drug delivery composition of claim 18 in situ at a topical site to a subject in need thereof.
  • 35. The method of claim 34, wherein the topical site is a periodontal pocket and the oral disease or condition is a dental disease or condition.
  • 36. The method of claim 34, wherein the oral disease or condition is periodontitis.
  • 37. The method of claim 34, wherein the composition is capable of releasing the therapeutic agent to treat the oral disease over a period of at least 2 days, and wherein the composition is only remained at the topical site.
CROSS REFERENCE TO RELATED APPLICATIONS

This application claims the priority to U.S. Provisional Application No. 62/835,542, filed on Apr. 18, 2019, and the disclosures of which are hereby incorporated by reference.

PCT Information
Filing Document Filing Date Country Kind
PCT/CN2020/085549 4/20/2020 WO 00
Provisional Applications (1)
Number Date Country
62835542 Apr 2019 US