design and characterization of three-dimensional twist-braid … · 2019-04-30 · design and...

9
TECHNOLOGY ARTICLE 98 TECHNOLOGY l VOLUME 5 NUMBER 2 JUNE 2017 © World Scientific Publishing Co. Design and characterization of three-dimensional twist-braid scaffolds for anterior cruciate ligament regeneration Shreya Madhavarapu 1 , Rohit Rao 2 , Sarah Libring 1 , Emma Fleisher 1 , Yasonia Yankannah 1 & Joseph W. Freeman 1 The anterior cruciate ligament (ACL) is the most commonly injured ligament in the knee, with more than 350,000 ACL injuries reported annually in the US. Current treatments include the use of autografts and allografts, which have a number of disadvantages. Previous attempts to use synthetic materials in ligament replacement have been unsuccessful due to their inability to replicate the long-term mechanical properties of the native ligament. The focus of this study was to develop twist-braid poly(L-lactic acid) (PLLA) scaffolds for ACL regeneration. Poly(ethylene glycol) diacrylate (PEGDA) was incorporated into the twist-braid scaffolds to evaluate its impact on their mechanical behavior. The twist-braid scaffolds were also compared with braided scaffolds. Scaffold mechanical properties were evaluated based on stress-relaxation, tensile and fatigue properties of the braided-only, twist-braid, and the twist-braid scaffolds with PEGDA. All the scaffolds exhibited properties comparable to the native human ACL with the twist-braid scaffolds displaying resistance to fatigue. Scaffolds were seeded with rat patellar tendon broblasts. Cell viability and the amount of protein released were studied over a course of 8 weeks. The scaffolds were stained with Picrosirius red after 8 weeks to show the deposition of extracellular matrix by the cells. The results from this study showed that the twist-braid scaffolds have properties most suitable for ligament regeneration. Keywords: Anterior Cruciate Ligament (ACL); Ligament Tissue Engineering; Polymer Scaffolds; Mechanical Properties; Cell Biocom- patibility. INNOVATION The development of scaffolds for anterior cruciate ligament (ACL) regeneration is challenging. An ideal scaffold is one that promotes tissue integration while also being strong enough to maintain knee joint stability and promote locomotion. Poly(L-lactic acid) (PLLA) is an FDA-approved bioresorbable material that has been widely used to develop such scaf- folds. Our approach uses three-dimensional (3D) braiding technology to develop scaffolds made of PLLA fibers whose sequential arrangement mimics the structure of ligaments. An element of fiber twisting was incorporated to provide fatigue resistance. Scaffold mechanical proper- ties and biocompatibility were tested to show that these scaffolds have potential as graſts for ACL regeneration. INTRODUCTION Ligaments can be described as dense connective tissues that connect bones to other bones to form joints. e knee joint is held together by four very important ligaments: the ACL, the posterior cruciate ligament (PCL), the medial collateral ligament (MCL) and the lateral collateral ligament (LCL). ese ligaments connect the femur and the tibia and play a key role in normal knee kinematics and maintaining joint stability 1–5 . Ligaments are composed of collagen (about 75% of the dry weight), elas- tin, proteoglycans (<1% of the dry weight), glycoproteins, other proteins, water and cells 2,3,6–8 . ey display a hierarchal structure comprising of collagen molecules, collagen fibrils, fibril bundles and fascicles, which are arranged parallel to the long-axis of the ligament 4,9 . e collagen fibrils are responsible for the crimp pattern (periodic change in direction) that is seen with ligaments. In the ACL, this is observed every 45–60 μm 10 . Ligaments exhibit a non-linear stress–strain relationship. When the ACL is subjected to a force, it is first transferred to the collagen fibrils resulting in their lateral contraction and straightening of the crimp pattern. Once it is straightened, the force is translated as collagen molecular strain 11 . As the force applied increases, the collagen fibrils extend and begin to slide over one another. With further increase in force, the collagen fibrils extend completely and start rupturing, eventually leading to failure of the ACL. e ACL is the most commonly injured ligament in the knee with over 350,000 injuries reported in the US every year 12,13 . ACL reconstruction surgery is the standard care procedure which is used to treat such injuries, and more than 200,000 such procedures were performed in 2011 in the US 13,14 . ACL injuries lead to a significant amount of joint dysfunction which over time can cause injury to other tissues in the joint and lead to the development of degenerative joint diseases like osteoarthritis and 1 Department of Biomedical Engineering, 2 Department of Chemical and Biochemical Engineering, Rutgers University, 599 Taylor Road, Piscataway, NJ 08854, USA. Correspondence should be addressed to: J.W.F. ([email protected]). Received 12 December 2016; accepted 3 May 2017; published online 31 March 2017; doi:10.1142/S2339547817500066

Upload: others

Post on 27-May-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Design and characterization of three-dimensional twist-braid … · 2019-04-30 · Design and characterization of three-dimensional twist-braid scaffolds for anterior cruciate ligament

TECHNOLOGYARTICLE

98 TECHNOLOGY l VOLUME 5 • NUMBER 2 • JUNE 2017© World Scientific Publishing Co.

Design and characterization of three-dimensional twist-braid scaffolds for anterior cruciate ligament regenerationShreya Madhavarapu1, Rohit Rao2, Sarah Libring1, Emma Fleisher1, Yasonia Yankannah1 & Joseph W. Freeman1

The anterior cruciate ligament (ACL) is the most commonly injured ligament in the knee, with more than 350,000 ACL injuries reported annually in the US. Current treatments include the use of autografts and allografts, which have a number of disadvantages. Previous attempts to use synthetic materials in ligament replacement have been unsuccessful due to their inability to replicate the long-term mechanical properties of the native ligament. The focus of this study was to develop twist-braid poly(L-lactic acid) (PLLA) scaffolds for ACL regeneration. Poly(ethylene glycol) diacrylate (PEGDA) was incorporated into the twist-braid scaffolds to evaluate its impact on their mechanical behavior. The twist-braid scaffolds were also compared with braided scaffolds. Scaffold mechanical properties were evaluated based on stress-relaxation, tensile and fatigue properties of the braided-only, twist-braid, and the twist-braid scaffolds with PEGDA. All the scaffolds exhibited properties comparable to the native human ACL with the twist-braid scaffolds displaying resistance to fatigue. Scaffolds were seeded with rat patellar tendon fi broblasts. Cell viability and the amount of protein released were studied over a course of 8 weeks. The scaffolds were stained with Picrosirius red after 8 weeks to show the deposition of extracellular matrix by the cells. The results from this study showed that the twist-braid scaffolds have properties most suitable for ligament regeneration.

Keywords: Anterior Cruciate Ligament (ACL); Ligament Tissue Engineering; Polymer Scaffolds; Mechanical Properties; Cell Biocom-patibility.

INNOVATIONThe development of scaffolds for anterior cruciate ligament (ACL) regeneration is challenging. An ideal scaff old is one that promotes tissue integration while also being strong enough to maintain knee joint stability and promote locomotion. Poly(L-lactic acid) (PLLA) is an FDA-approved bioresorbable material that has been widely used to develop such scaf-folds. Our approach uses three-dimensional (3D) braiding technology to develop scaff olds made of PLLA fi bers whose sequential arrangement mimics the structure of ligaments. An element of fi ber twisting was incorporated to provide fatigue resistance. Scaff old mechanical proper-ties and biocompatibility were tested to show that these scaff olds have potential as graft s for ACL regeneration.

INTRODUCTIONLigaments can be described as dense connective tissues that connect bones to other bones to form joints. Th e knee joint is held together by four very important ligaments: the ACL, the posterior cruciate ligament (PCL), the medial collateral ligament (MCL) and the lateral collateral ligament (LCL). Th ese ligaments connect the femur and the tibia and play a key role in normal knee kinematics and maintaining joint stability1–5.

Ligaments are composed of collagen (about 75% of the dry weight), elas-tin, proteoglycans (<1% of the dry weight), glycoproteins, other proteins, water and cells2,3,6–8. Th ey display a hierarchal structure comprising of collagen molecules, collagen fi brils, fi bril bundles and fascicles, which are arranged parallel to the long-axis of the ligament4,9. Th e collagen fi brils are responsible for the crimp pattern (periodic change in direction) that is seen with ligaments. In the ACL, this is observed every 45–60 μm10. Ligaments exhibit a non-linear stress–strain relationship. When the ACL is subjected to a force, it is fi rst transferred to the collagen fi brils resulting in their lateral contraction and straightening of the crimp pattern. Once it is straightened, the force is translated as collagen molecular strain11. As the force applied increases, the collagen fi brils extend and begin to slide over one another. With further increase in force, the collagen fi brils extend completely and start rupturing, eventually leading to failure of the ACL.

Th e ACL is the most commonly injured ligament in the knee with over 350,000 injuries reported in the US every year12,13. ACL reconstruction surgery is the standard care procedure which is used to treat such injuries, and more than 200,000 such procedures were performed in 2011 in the US13,14. ACL injuries lead to a signifi cant amount of joint dysfunction which over time can cause injury to other tissues in the joint and lead to the development of degenerative joint diseases like osteoarthritis and

1Department of Biomedical Engineering, 2Department of Chemical and Biochemical Engineering, Rutgers University, 599 Taylor Road, Piscataway, NJ 08854, USA. Correspondence should be addressed to: J.W.F. ([email protected]).

Received 12 December 2016; accepted 3 May 2017; published online 31 March 2017; doi:10.1142/S2339547817500066

1750006.indd 981750006.indd 98 5/31/2017 10:52:55 AM5/31/2017 10:52:55 AM

Page 2: Design and characterization of three-dimensional twist-braid … · 2019-04-30 · Design and characterization of three-dimensional twist-braid scaffolds for anterior cruciate ligament

ARTICLE

99TECHNOLOGY l VOLUME 5 • NUMBER 2 • JUNE 2017© World Scientific Publishing Co.

osteoporosis15. Th e ACL is completely surrounded by synovial fl uid and lacks vascularity which is why it is unable to heal completely following an injury16. Hence, an ACL reconstruction is performed which requires an ACL graft . Traditionally, biological graft s have been used to treat ACL injuries11. Th ese include autograft s, allograft s and xenograft s17. Autograft s and allograft s most commonly used are derived from the hamstring tendon, quadriceps tendon, and bone–patellar tendon–bone (BPTB) with the BPTB being the gold standard for ACL reconstruction surgery1,18–21. Long-term evaluation has shown that biological graft s have a 85–90% success rate in terms of patient satisfaction and restoration of knee stability12,22. Autograft s signifi cantly promote bone-to-bone healing, fi xation securing and return to pre-injury level of activity12. Th e graft used for reconstruction is harvested during the same surgery resulting in a longer procedure which contributes to increased pain and a longer healing time17,23–25. Other complications associated with autograft s include anterior knee pain, kneeling pain, patellar fracture, nerve injury, graft failure, patellofemoral pain, decreased range-of-motion and muscle atrophy12,22,25,26. Unlike autograft s, patients undergoing ACL reconstruction with allograft s require shorter surgical procedures. Th ey are associated with a lower incidence of postoperative arthrofi brosis and knee stiff ness12,27. Th ey have a better short-term function when compared to autograft s. Despite the advantages, they are associated with the risk of disease transmission, infection, and possible immune rejection23,26,28. A number of sterilization techniques are applied to overcome these complications but this leads to inferior mechanical properties in terms of knee laxity, stress to failure and post-operative traumatic rupture rate12,24. Th ere is also delayed incorporation, bone tunnel enlargement and slower remodeling activity when compared to autograft s due to lower cellular-ity and the absence of revascularization12. Another important concern regarding allograft s is the limited supply of donors which is unable to meet the escalating demand for allograft tissue17,22.

Over the last few decades, various synthetic graft s have made their way into the commercial space. Th ey off er a simple and easily reproduc-ible option, require a shorter surgery, show signifi cant strength and are associated with an accelerated rehabilitation period24. Synthetic graft s can either be prosthetics or augmentation devices. While these graft s gained some popularity in the 1970s and 1980s, it was under very limited conditions1,24,29. Most of these graft s failed due to their inability to mimic the mechanical properties of the human ACL. Th e Gore-Tex (braided polytetrafl uoroethylene fi bers) and the Stryker–Dacron ligament (woven polyethylene terephthalate) were approved by the FDA for use as liga-ment prosthesis30. Th e Leeds–Keio prosthesis was another such device which was popular outside the US24. Results from ACL reconstructions using these prosthesis showed foreign-body reactions and infl amma-tions, limited integration between the graft and the host tissue, material degradation, poor abrasion resistance, and fatigue failure8,11,12,26,29,31. Th ese ligament prosthesis were withdrawn from the market due to their long-term drawbacks which included premature graft failure, synovitis, osteoporosis and osteoarthritis1,11,12,22,24,29,31,32. Augmentation devices such as the Kennedy Ligament Augmentation Device (LAD) were de-signed to promote integration of the graft and the host30. However, these were also fraught with complications such as infl ammatory reactions and graft failure22. Th e most successful augmentation device to date is the LARS (Ligament Advanced Reinforcement System) which is made from polyethylene terephthalate (PET), a non-biodegradable material21,33. Th e two ends of this graft are composed of woven fi bers that provide strength while the intra-articular portion consists of multiple longitudinal fi bers twisted at an angle of 90° to resist fatigue and allow tissue ingrowth33. While the LARS has shown promising results and shorter rehabilitation times, tissue regeneration is not possible with non-degradable materials like PET. To achieve this, biodegradable materials have to be used which has prompted an interest in tissue engineering for alternate options.

Tissue engineering has been described as a multidisciplinary fi eld that incorporates the principles of biochemistry, engineering, and

material science to develop substitutes for replacing injured or diseased tissues4,22,34,35. Tissue-engineered woven and braided scaff olds are being examined as options for biodegradable ACL graft s. Materials like silk, poly(glycolic acid) (PGA), poly(lactic-co-glycolic acid) (PLGA), PLLA and polycaprolactone (PCL) are being used for this purpose. In a study by Lu et al., PGA, PLGA and PLLA fi bers were used to make 3D braided scaff olds31. Comparing them, it was shown that PLLA scaff olds promoted greater cell adhesion and proliferation and had a slower degradation rate, which led to superior mechanical properties over the same duration in culture as the PGA and PLGA. Th e study also showed that a fi bronectin coating on such scaff olds allows for improved tissue integration. Cooper et al. developed a method for using 3D braiding technology to produce biodegradable ACL replacements. PLGA-based braided scaff olds were fabricated comprising of bony attachment and intra-articular regions to mimic the native anatomy of a human ACL, showing that such scaff olds promoted tissue regeneration and that the 3D braiding technique can be used to control the structural and mechanical properties of the scaff olds26. Interest in PLLA as a material for ACL regeneration is strengthened by the fact that it takes approximately 2 years to degrade, hence allowing tissue integration while maintaining mechanical stability21. Owing to its advantages, Freeman et al. studied two-dimensional (2D) braided, twisted and braid-twist PLLA scaff olds11. Th ey showed that among the three, the braid-twist scaff olds aff ord the best mechanical stability. PLLA woven braids wrapped with gelatin hydrogels were studied in rabbits by Kimura et al.36 Th ey showed that there was as gradual degradation of the scaff olds leading to a decrease in their mechanical strength, but this was aided by collagen production and osteointegration, both of which helped maintain mechanical stability. Th ey also showed that hydrogels, like gelatin, can be used as media for the delivery of growth factors that are necessary for cell integration and ECM (extracellular matrix) production.

Given the advantages of using PLLA for ligament regeneration, and the ease with which 3D braiding technology can be used to tailor mechanical properties, PLLA was used in this study to fabricate 3D braid-twist scaff olds. Previous studies with 2D scaff olds had shown that combining the techniques of fi ber twisting and braiding will enable the development of scaff olds that more closely mimic the hierarchal structure and mechanical properties of a human ACL. In this scaff old, the microfi bers are grouped into fi bers, fi bers into bundles and bundles into yarns. Th e twisting mimics aspects of the crimp pattern in ligaments. Th e combination of braiding and twisting gives the scaff olds a stress–strain profi le that is similar to the natural ligament. Scaff olds were mechanically characterized based on their tensile, viscoelastic and fatigue properties, and their cell biocompatibility was studied based on cell viability and protein production.

MATERIALS AND METHODSTh e PLLA fi bers (120 denier/30 microfi laments [60 gm]) used to fabricate the diff erent types of scaff olds were purchased from Biomedical Struc-tures, RI. Two types of scaff olds were made: the twist-braid scaff olds and the braided-only scaff olds. Each of these scaff olds were composed of 288 PLLA fi bers. Braided structures were fabricated using our custom-built braiding machine.

Fabrication of twist-braid (TB) scaffoldsFiber yarns were made using a previously developed method using a Conair device (Model QB3ECS, Conair Corporation, East Windsor, NJ)11. Briefl y, PLLA fi bers were arranged in groups of four and each group was fi rst twisted in a counterclockwise direction at a twisting angle of 60°. Th e three twisted bundles were then twisted together at an angle of 72° to form a yarn. Twisting times to obtain these angles were previously optimized by Freeman et al., who showed that these resulted in mechanical properties optimal for ligament regeneration11. Twisting angles are defi ned as the angle between the long axis of the fi ber and the

1750006.indd 991750006.indd 99 5/31/2017 10:52:56 AM5/31/2017 10:52:56 AM

Page 3: Design and characterization of three-dimensional twist-braid … · 2019-04-30 · Design and characterization of three-dimensional twist-braid scaffolds for anterior cruciate ligament

ARTICLE

100 TECHNOLOGY l VOLUME 5 • NUMBER 2 • JUNE 2017© World Scientific Publishing Co.

line perpendicular to the long axis of the yarn. Twenty-four yarns were braided using the MoTR Lab’s Custom 3D Braiding Machine at a height of 31 cm to obtain a braiding angle of 72°. Th is angle was chosen based on previous experiments which showed that braids with this angle have tensile properties most comparable to the native human ACL37. Th e braiding angle is defi ned as the angle made by the long axis of the yarns incorporated into middle of the braid and the line perpendicular to the long axis of the scaff old.

Fabrication of braided-only (BO) scaffoldsTwelve fi bers were cut and bundled together. Twenty-four such bundles were attached to the braiding machine and braided together to form the BO scaff olds.

Incorporation of hydrogelA 10% w/v solution of poly(ethylene glycol) diacrylate (PEGDA, mo-lecular weight 4,000 Da, Monomer-Polymer and Dajac Labs, Trevose, PA) with a 1% (w/v) solution of photoinitiator (2,2-dimethoxy-2-phenyl acetophenone, n-vinyl-2-pyrrolidinone [300 mg/mL]) was incorporated into the twist-braid scaff olds in two diff erent ways. In the case of the hydrogel scaff olds (HS), the scaff olds aft er braiding were soaked in the hydrogel and subsequently cross-linked under UV light (365 nm). With the hydrogel yarns (HY), each yarn was soaked in hydrogel, cross-linked, and braided.

Pore size and distributionScaff old pore size was determined using the LEP-1100A Liquid Extrusion Porosimeter (PMI, Ithaca, NY). Scaff olds, approximately 1 cm in length were fi rst vacuum-soaked in wetting fl uid (Galwick, surface tension 15.9 dynes/cm, PMI, Ithaca, NY) for 30 minutes and then left to soak over-night. A small rectangular section was cut in the center of a non-porous circular piece of plastic and the scaff old to be tested was glued over it. Th e entire rectangular portion was completely covered by the scaff old, leaving no gaps. Th is was placed in the testing chamber, and then the sample chamber was placed over it and fi lled with Galwick. Th e scaff old was allowed to soak in it for seven to ten minutes before the test was started. Th e median pore size and pore size distributions were obtained for the braided-only and the twist-braid scaff olds (n = 3).

Mechanical characterizationTh e Instron (Model 5869, Norwood, MA) with a 50-kN load cell and MTS (Tytron 250, Eden Prairie, MN) mechanical testing systems were used to determine the scaff old mechanical properties. Scaff old samples of length 3.5 cm were used for the tests. Th e ends of each scaff old were coated with a castable epoxy (Smooth-on, Inc. Macungie, PA) to prevent them from slipping out of the grips during a test. A gauge length of 15 mm and n = 5 was chosen for all the tests. All the scaff olds were vacuum soaked in phosphate-buff ered saline (PBS) for 30 minutes prior to testing. During the tests, PBS was sprayed on the scaff olds to keep them moist.

Stress relaxation tests were performed using the Instron to determine the scaff old viscoelastic properties. Th e scaff olds were loaded to 50 N at 0.3 mm/sec and then allowed to relax for 1,800 seconds. Th e normalized stresses at intervals of 200 seconds were calculated for all the groups, i.e. the BO, TB, HS and HY scaff olds by dividing the stresses at each time point by the maximum stress.

To test for tensile properties, the scaff olds were fi rst preloaded to 0.2 N and then extended to failure at a strain rate of 2%/s (0.3 mm/sec)25. Th e ultimate tensile strength (UTS), elastic modulus of the linear region, the length of the toe-region and the strain at failure were characterized.

Scaff olds were subjected to cyclic strain using the MTS Tytron 250 in order to study their creep properties. Th ey were fi rst extended to 10% their gauge length, i.e. to 16.5 mm. Th e 10% was chosen based on the data that was obtained from the tensile tests, which showed that the toe region of the scaff olds was about 8–10%. Following this, cyclic strain of

2%, i.e. 0.3 mm was applied at a frequency of 1 Hz for 2,000 cycles. Th e samples were then stretched out to failure at 0.3 mm/sec using the Instron. Th e resultant tensile properties, i.e. the maximum load, UTS, strain at failure and modulus were compared to those obtained from the tensile tests conducted prior to fatigue.

Sterilization of scaffoldsTh e four diff erent types of scaff olds (BO, TB, HS, HY and n = 6 for each), each of length 9 mm were fi rst vacuum soaked in 70% ethanol for 30 minutes to allow ethanol to penetrate into the scaff olds and reduce the risk of possible contamination. Th e scaff olds were then placed in separate wells in a 48-well plate and sterilized by exposing each side to UV light for 30 minutes.

Cell seedingRat patellar tendon (rPT) fi broblasts were used in the cell studies. Th ese cells had been previously isolated in the lab using the explant+outgrowth method from the patellar tendons of a skeletally mature male Sprague–Dawley rat. Cryopreserved cells were thawed and expanded in α-MEM (Gibco® by Life Technologies, Carlsbad, CA) supplemented with 10% fetal bovine serum (FBS) and 1% penicillin and streptomycin (P/S) (Gibco® by Life Technologies, Carlsbad, CA). Th e cells were grown to about 80% confl uence in the growth medium at 37°C and 5% CO2. Th e scaff olds, aft er sterilization, were soaked in media (α-MEM, 10% FBS, 1% P/S) overnight. Th is was done to promote cell adhesion to the scaff olds when cells were seeded the next morning. Th ey were seeded at a density of 100,000 cells/scaff old. Cells were also seeded on the TCP to serve as an additional control. Th e cell seeding effi ciency was calculated at day 1 by counting the number of cells that were attached to the well plate. Media was changed three times a week.

Cell viability: PrestoBlue assayTh e PrestoBlue assay measures mitochondrial activity which was used an indirect measure of cell viability. Th e PrestoBlue assay (n = 6) was run on day 1 aft er transferring the scaff olds into a new 48-well plate. Th e PrestoBlue solution was prepared using media and the PrestoBlue® reagent in a ratio of 10:1. Th e PrestoBlue solution was added to each well aft er removing media and washing the cells with PBS. Th e well plate was incubated for 1 hour. Following this, triplicates of each sample were added to a 96-well plate and then placed in a plate reader. Since the PrestoBlue reagent is photosensitive, care was taken to avoid exposure to light during all procedures involving this reagent. Th is assay was run on days 1, 3, 7, 14, 21 and 28 and then at 6 weeks and 8 weeks.

Total protein estimation: BCA assayTh e amount of total protein that was released into the media was estimated at days 1, 3, 7, 14, 21 and 28 and at 6 weeks and 8 weeks (n = 6) using the Bicinchoninic acid (BCA) assay. For this assay, media was collected at every time point that it had to be changed, i.e. three times a week. A 10× dilution of samples for each time point was used. Th e standards for the assay were prepared according to the Pierce BCA Protein Assay protocol. Twenty-fi ve microliters of the standards and samples (duplicates) were pipetted into a 96-well plate. Two hundred microliters of the BCA working reagent was then added to each well. Aft er incubating for 30 minutes, the plate was placed in a plate reader and the values of absorbance were noted. A standard curve was plotted using the values obtained for the standards. Th e values of the samples were then compared to the standard curve to get the amount of total protein present in each of the samples. Th e total protein concentration obtained for the media was subtracted from the concentration obtained for each individual sample to normalize the results.

Picrosirius red staining and quantifi cationTh e scaff old samples, aft er 8 weeks, were stained for collagen using the Picrosirius Red Stain Kit (Polyscience, Inc., Warrington, PA). Images were

1750006.indd 1001750006.indd 100 5/31/2017 10:52:56 AM5/31/2017 10:52:56 AM

Page 4: Design and characterization of three-dimensional twist-braid … · 2019-04-30 · Design and characterization of three-dimensional twist-braid scaffolds for anterior cruciate ligament

ARTICLE

101TECHNOLOGY l VOLUME 5 • NUMBER 2 • JUNE 2017© World Scientific Publishing Co.

obtained using a LG Google Nexus 5 8-megapixel camera and analyzed using ImageJ. Aft er converting them to 8-bit grayscale images, the mean intensity of the pixels was calculated and compared.

Statistical analysis of dataAll results were expressed as mean ± standard deviation. Diff erences between the groups were analyzed using one-way ANOVA with Fisher’s LSD, with p < 0.05 considered to be signifi cant.

RESULTS

Cross-sectional area and weight of scaffoldsFigure 1 shows that the TB scaffolds displayed a significantly larger cross-sectional area (8.23 ± 0.29 mm2), than the BO scaff olds (6.84 ± 0.32 mm2). Th e two scaff olds with the hydrogel also exhibited signifi cantly greater cross-sectional areas (HS 8.75 ± 0.62 mm2 and HY 9.56 ± 0.35 mm2) when compared to the TB and the BO scaff olds. Th e cross-sectional area of the HY scaff olds was also signifi cantly higher than that of the HS scaff olds. Th e TB, BO, HS and HY sam-ples were weighed to determine the weight per unit length (i.e. 1 cm). Th e TB scaff olds (50.18 ± 2.97 mg), HS scaff olds (52.39 ± 1.89 mg) and the HY scaff olds (57.78 ± 4.37 mg) all weighed signifi cantly more than the BO scaff olds (47.83 ± 2.17 mg). Th e weight of the HY scaff olds was also signifi cantly greater than that of the TB and the HS scaff olds.

Pore size and distributionTh e BO scaff olds displayed a median pore diameter of 73.51 ± 10.58 μm. Th e median pore diameter of the TB scaff olds was signifi cantly lower at 39.04 ± 7.78 μm. Th e median pore diameter indicates that 50% of the pores have pore diameters smaller than this value and the other 50% have larger diameters. For the BO scaff olds, it was observed that approximately 50% of pores were in the range of 70 to 100 μm. For the TB scaff olds, more than 80% of the pore sizes were below 70 μm.

Viscoelastic propertiesTh e diff erences between the TB, HS and HY were statistically insignifi cant at all the time points. Normalized stresses were calculated by dividing the stresses at each time point by the

maximum stress. Figure 2 shows that the BO scaff olds, when com-pared to the other three, displayed significantly higher normalized stresses at all the time points. Normalized stresses for the TB scaff olds (with and without hydro-gel) at 1,000 seconds ranged from 61.3 ± 4.8% for the HS, 64.7 ± 3.3% for the HY, to 64.9 ± 2.8% for the TB scaff olds. Th e normalized stress for BO scaff olds was signifi cantly higher at 70.5 ± 2.1%.

Tensile propertiesThe data from the tensile tests (Fig. 3) showed that the BO and TB scaffolds were able to withstand maximum tensi le loads of 614.76 ± 27.17 N and 593.35 ± 18.28 N, respectively. Th e HS (691.34 ± 27.19 N) and HY (734.42 ± 30.18 N) scaffolds displayed significantly higher maximum loads than both the BO

and TB scaffolds. The BO scaffolds (86.38 ± 5.46 MPa) exhibited the highest UTS. The UTS for the TB (70.83 ± 3.40 MPa) and HY (77.41 ± 4.37 MPa) were significantly lower than that of the BO scaffolds. The TB scaffolds showed a significantly lower UTS than the HS (81.59 ± 5.96 MPa) and the HY scaffolds. The end of the toe region was noted as the point where the slope increased by more than 200%. The scaffolds showed toe regions ranging from about 8–10%. The BO and HS scaffolds displayed toe region of 8.42 ± 0.75% and 8.53 ± 1.66%. The HY scaffolds displayed a significantly larger toe region (9.74 ± 1.34%) than the TB scaffolds (8.01 ± 0.9%). The two scaf-folds with the hydrogel-HS (64.64 ± 7.20%) and HY (72.46 ± 4.59%) exhibited significantly higher stains at failure than the TB (50.46 ± 5.30%) and BO (49.00 ± 5.40%) scaffolds. The HY also displayed a significantly higher strain at failure than the HS scaffolds. The BO scaffolds exhibited the highest elastic modulus at 425.79 ± 38.68 MPa. This is significantly higher than that of the TB, HS and HY scaffolds which displayed moduli of 315.28 ± 18.81 MPa, 293.01 ± 43.43 MPa,

Figure 1 Left to right: cross-sectional areas of four scaffolds, weight per unit length (1 cm) of the scaffolds. An asterisk (*) indicates a signifi cant difference (p < 0.05). BO: braided only; TB: twist-braid; HS: hydrogel scaf-fold; HY: hydrogel yarns.

0.0

17.5

35.0

52.5

70.0

BO

Wei

ght (

mg)

TB HS HY

*

0.0

2.5

5.0

7.5

10.0

12.5

BO TB HS HYC

ross

-sec

tiona

l Are

a (m

m2 )

*

Figure 2 Stress-relaxation data obtained for all the scaffolds after a load of 50 N (n = 5). BO: braided only; TB: twist-braid; HS: hydrogel scaffold; HY: hydrogel yarns.

0.500

0.550

0.600

0.650

0.700

0.750

0.800

0 200 400 600 800 1000 1200 1400 1600 1800 2000

Nor

mal

ized

Ste

ss

BO

Time (Sec)

TB HS HY

1750006.indd 1011750006.indd 101 5/31/2017 10:52:57 AM5/31/2017 10:52:57 AM

Page 5: Design and characterization of three-dimensional twist-braid … · 2019-04-30 · Design and characterization of three-dimensional twist-braid scaffolds for anterior cruciate ligament

ARTICLE

102 TECHNOLOGY l VOLUME 5 • NUMBER 2 • JUNE 2017© World Scientific Publishing Co.

and 248.90 ± 15.47 MPa, respectively. The TB and HS scaffolds also showed significantly higher moduli than the HY scaffolds.

Fatigue analysisA significant decrease in the maximum load after 2,000 cycles was observed for all groups except the TB scaff olds. In case of maximum extension before failure, a signifi cant diff erence was only observed in case of HY scaff olds. Th e ultimate tensile strength signifi cantly decreased for all but the TB scaff olds. A signifi cant drop in the strain at failure is observed only with HY scaff olds. It was also observed that elastic moduli of all four types of scaff olds decreased aft er subjecting them to 2,000 cycles of strain but this drop was only signifi cant in the BO and HS scaff olds.

Cell viabilityTh e seeding effi ciency calculated at day 1 was approximately 95% for the BO scaff olds, 82% for the TB scaff olds, 80% for the HS scaff olds and 91% for the HY scaff olds. Th e data obtained from the PrestoBlue assays showed that the cell viability increased steadily until day 7 following which a drop was observed at day 14. At day 21, a signifi cant decrease in the viability was seen following which there were very small changes in the cell viability until 8 weeks.

Total protein estimationTh e data obtained from the BCA assay showed that there was a signifi cant increase in the amount of protein released into the media at day 14 for all of the scaff olds.

Staining for collagen As shown in Fig. 6a, the blank scaff old (one with no cells seeded on it) showed an absence of any color indicating that the scaff old itself did not get stained and that the color observed in case of the other scaff olds was due to the collagen deposited by the cells. A reddish-pink color was observed in case of all four types of scaff olds (BO, TB, HS, HY) indicating the presence of collagen (Figs. 6c–f). Figure 6b shows a blank scaff old with PEGDA (with no cells seeded on it but containing hydrogel) which was stained a yellow-orange color indicating that the hydrogel retained some stain. Table 1 shows the color intensity values obtained using ImageJ for the diff erent types of scaff olds.

DISCUSSIONThe objective of this study was to develop a novel scaffold for ACL regeneration. In order for such a scaffold to be successful, it has to display mechanical properties similar to or superior to that of the native ligament. It must also enable cell adhesion and proliferation and promote tissue integration. To achieve this, 3D braiding technology was used in combination with fi ber twisting to develop 3D TB scaff olds. PLLA fi bers used in this study were each composed of 30 microfi laments. Th e sequential arrangement of fi bers, bundles, yarns and braids was used to achieve the hierarchal structure of native ligaments while the element of twisting was incorporated to replicate their crimp pattern.

Th e TB scaff olds were compared to BO scaff olds to assess the impact that fi ber twisting has on the scaff old mechanical properties. Further, hydrogel PEGDA was incorporated into these scaffolds to study its impact on their mechanical behavior. Data showed that TB, HS and

Figure 3 Top (left to right): Stress–strain data obtained for the different scaffolds; maximum tensile load; ultimate tensile strength (UTS). Bottom (left to right): Toe region; strain at failure (%); elastic modulus (MPa). BO: braided only; TB: twist-braid; HS: hydrogel scaffold; HY: hydrogel yarns.

0

-25

25

50

75

100

0 30 60 90 120

Stre

ss (

MPa

)

Strain %

BO TB HS HY

0

25

50

75

100

BO TB

Stre

ss (

MPa

)

HS HY

*

0

200

400

600

800

1000

BO

Max

Loa

d (N

)

TB HS HY*

0

3

6

9

12

BO

% S

trai

n

TB HS HY

*

0

20

40

60

80

100

BO TB HS HY

Stra

in a

t fai

lure

(%

)

*

0

125

250

375

500

BO TB

Mod

ulus

(M

Pa)

HS HY

*

1750006.indd 1021750006.indd 102 5/31/2017 10:52:57 AM5/31/2017 10:52:57 AM

Page 6: Design and characterization of three-dimensional twist-braid … · 2019-04-30 · Design and characterization of three-dimensional twist-braid scaffolds for anterior cruciate ligament

ARTICLE

103TECHNOLOGY l VOLUME 5 • NUMBER 2 • JUNE 2017© World Scientific Publishing Co.

HY scaff olds all had signifi cantly larger cross-sectional areas than the BO scaff olds. It was also observed that the incorporation of hydrogel resulted in a signifi cant increase in the cross-sectional area, with the HY scaff olds exhibiting the largest cross-sectional area. An increasing trend was also observed in terms of the weight per unit length (BO < TB < HS < HY). Th is observation that the HY scaff olds weighed more than the HS scaff olds, shows that signifi cantly more hydrogel was incorporated into the HY scaff olds. It also indicated that despite braiding aft er the incorporation of the hydrogel onto the individual yarns, there was not a signifi cant loss of hydrogel during the braiding process. Th e data obtained from liquid extrusion porosimetry showed that the median pore diameter was signifi cantly higher for the BO scaff olds. Th e pore size distribution data showed that BO scaff olds had a higher percentage of larger pores when compared to the TB scaff olds.

Th e results from the mechanical tests show that both the BO and TB scaff olds (both with and without hydrogel) exhibit favorable mechanical properties. Th e data from the stress-relaxation tests showed that the normalized stress values obtained for all the scaff olds at 1,000 seconds were comparable to the 70% obtained for the native ACL37. Th is data also showed that the incorporation of twisting lead to the TB scaff olds being signifi cantly more viscoelastic than the BO scaff olds. Th e addition of the twisting allows the stress at the beginning of the test to be used for stretching out the twisting pattern. Th is leads to less of the stress being applied to direct fi ber stretching in these scaff olds than in the BO scaff olds and a larger degree of viscoelastic (or rate dependent) behavior. Results from the tensile tests showed that the scaff olds with hydrogel,

i.e. the HS and HY scaff olds, can withstand signifi cantly higher tensile loads. Hydrogels absorb water and release it when under stress. Th is places less stress on the fi bers at the beginning of the mechanical tests. While the addition of twisting did make the scaff olds viscoelastic, hy-drogel PEGDA was incorporated to more closely mimic the viscoelastic feature of ligaments. Incorporation of PEGDA lead to a larger toe region, increased load-bearing capabilities and overall mechanical properties that were comparable to the native ACL. The TB and HS scaffolds showed signifi cantly higher moduli than the HY scaff olds. Th is again shows that the incorporation of the hydrogel signifi cantly reduced the stiff ness of the scaff olds which would help prevent them from failure under high load conditions. Following the application of cyclic strain, a signifi cant decrease in maximum load was observed for the BO, HS and HY scaff olds. Th e UTS also showed a signifi cant decrease for all but the TB scaff olds, indicating that these scaff olds are able to withstand cyclic strain and are not as susceptible to fatigue as the rest of the scaff olds. Aft er cyclic loading, the changes observed in mechanical properties of TB scaff olds were not signifi cant, which could be a result of fi ber twisting. Once again, incorporation of fi ber twisting allows the scaff olds to be strained to a greater extent without directly stretching out the individual PLLA fi bers, similar to the stretching out of the crimp pattern observed in native ligaments11. Fiber-twisting replicates the crimp pattern, improving strength and fatigue resistance of the scaff olds. Aft er the fatigue tests, a significant difference in maximum extension before failure was only observed in the HY scaff olds. Th is signifi cant decrease can be attributed to the loss of hydrogel or the scraping away of hydrogel

Figure 4 The trends observed before and after applying cyclic strain (2,000 cycles). Top (left to right): Maximum load; maximum extension; UTS. Bottom (left to right): Strain at failure; elastic modulus. BO: braided only; TB: twist-braid; HS: hydrogel scaffold; HY: hydrogel yarns.

0

200

400

600

800

1000

Before After

)

BO

Max

Loa

d (N

TB HS HY*

0

3

6

9

12

15

Before After

BO TB

Max

Ext

ensi

on (

mm

)

HS HY

*

0

25

50

75

100

Before After

UT

S

BO TB HS HY

*

0

20

40

60

80

100

Before After

BO TB HS HY

Stra

in a

t fai

lure

(%

)

*

0

125

250

375

500

Before After

BO

Mod

ulus

TB HS HY

*

1750006.indd 1031750006.indd 103 5/31/2017 10:52:57 AM5/31/2017 10:52:57 AM

Page 7: Design and characterization of three-dimensional twist-braid … · 2019-04-30 · Design and characterization of three-dimensional twist-braid scaffolds for anterior cruciate ligament

ARTICLE

104 TECHNOLOGY l VOLUME 5 • NUMBER 2 • JUNE 2017© World Scientific Publishing Co.

due to the application of cyclic strain for an extended period of time. A signifi cant drop in the strain at failure was also observed only with HY scaff olds. Th is can again be due to the loss of hydrogel during the process of cyclic strain. Table 2 shows that results obtained for all the scaff olds — BO, TB, HS and HY — and a comparison to the mechanical properties of the native human ACL. It can be observed that all scaff olds have mechanical properties that are superior to the native human ACL. Th is is essential to maintaining the mechanical properties throughout the polymer degrada-tion process. Th e results obtained from the PrestoBlue assay show that the cell viability increased steadily until day 7 for all the scaf-folds, following which a drop was observed at day 14. Th is indicates that the cells were growing at a steady pace for the fi rst 7 days. Th e drop at day 14 could be a result of the cells reaching confl uence and starting to lay down extracellular matrix. Th e data obtained from the BCA assay showed a signifi cant increase in the amount of protein released at day 14. Th is shows that the fi broblasts did indeed lay down extracellular matrix between days 7 and 14. At day 21, a signifi cant decrease in the cell viability was seen following which there were very small changes in the cell viability until the end of 8 weeks. Th is decrease in viability shows that the cells may have gone into a state of quiescence aft er reaching confl uence. There may have been some amount of cell death due to the cells not receiving suffi cient nutrients. As the cell debris was washed away (through changing media), the remaining cells received nutrients bringing them out of their state of quiescence and leading to cell growth. Th is may be the cause of the small increase in cell viability that is observed from

day 28 to 8 weeks. Th e images obtained aft er staining the scaff old with picrosirius red (aft er 8 weeks) showed that the HS and HY scaff olds exhibited lower color intensities. Th is was due to the retention of stain by the hydrogel PEGDA as shown with the blank scaff old containing PEDGA that was stained a yellow-orange color. Th e lower intensity of color does not indicate that there was more collagen deposition with these scaff olds. From the values of color intensity values shown in Table 1 and the images in Fig. 6, it can be concluded that collagen was deposited on all four types of scaff olds.

Figure 6 Top: Scaffolds stained with picrosirius red. Bottom: Sections (zoomed in) used for ImageJ analysis. (a) Blank scaffold (no cells), (b) Blank Scaffold with PEGDA, (c) BO: braided only, (d) TB: twist-braid, (e) HS: hydrogel scaf-fold and (f) HY: hydrogel yarns.

Figure 5 Data obtained from the (left) PrestoBlue Assay and (right) BCA Assay over the course of the 8-week study (n = 6). BO: braided only; TB: twist-braid; HS: hydrogel scaffold; HY: hydrogel yarns.

-1125

0

1125

2250

3375

1 7 14 21 28 42 56

BO

Time (Days)

TB HS HY

Tot

al P

rote

in R

elea

sed

by th

e ce

lls(µ

g/m

l)

-9375

0

9375

18750

28125

1 3 7 14 21 28 42 56

BO

Time (Days)

TB

Rel

ativ

e Fl

uore

scen

ce U

nits

(R

FU)

HS HY

Table 1 Intensity of color (Mean; SD: standard deviation; Min: minimum; Max: maximum) obtained for diff erent scaff olds using ImageJ. Maxi-mum intensity is denoted by 0 and the minimum is 255 for 8-bit images. BO:  braided only; TB: twist-braid; HS: hydrogel scaff old; HY: hydrogel yarns.

Scaff old Area Mean SD Min Max

Blank 41154 219.019 5.687 190 241

Blank with PEGDA 41154 110.897 13.769 79 158

BO 41154 143.426 17.722 98 211

TB 41154 164.971 16.017 122 220

HS 41154 108.713 15.638 72 197

HY 41154 102.556 14.255 78 194

1750006.indd 1041750006.indd 104 5/31/2017 10:52:58 AM5/31/2017 10:52:58 AM

Page 8: Design and characterization of three-dimensional twist-braid … · 2019-04-30 · Design and characterization of three-dimensional twist-braid scaffolds for anterior cruciate ligament

ARTICLE

105TECHNOLOGY l VOLUME 5 • NUMBER 2 • JUNE 2017© World Scientific Publishing Co.

CONCLUSIONAn ideal tissue-engineered graft for ligament replacement has to be biocompatible, biodegradable and mimic the biomechanical proper-ties of the native human ligament. PLLA was the material of choice because it is biodegradable, approved by the FDA for various clinical applications, and does not induce a permanent foreign body response. PLLA also degrades over a longer period of time (>24 months) when compared to other polyhydroxyesters such as PGA (6 to 12 months) and PLGA (1 to 6 months)22. Th is is essential as the graft has to allow regeneration of new tissue while maintaining its mechanical proper-ties. In this study, the four diff erent types of scaff olds, i.e. the BO, the TB and the two scaff olds with the hydrogel PEGDA (HS, HY) were all found to mimic the mechanical properties of the native human ACL. Aft er subjecting all the scaff olds to 2,000 cycles of strain, only the TB scaff olds showed changes in mechanical properties that were not signifi cant. Th e TB scaff olds displayed resistance to fatigue, indicating that the incorporation of fi ber-twisting can provide better long-term mechanical stability. Cell adhesion, proliferation and ECM deposition were observed with all the scaff olds, proving their biocompatibility. Future work would involve creating a steady supply of nutrients to the cells and the application of diff erentiation cues in order to study tissue integration and vascularization. Th e concentration of PEGDA could be varied and diff erent hydrogel phases could be evaluated to optimize the viscoelastic behavior of the TB scaff olds. Air could also be bubbled into the hydrogel phase before cross-linking to increase porosity. Th e hydrogel phase could also be made more bioactive by cross-linking PEGDA, which is relatively biologically inert, with another polymer/monomer or using a diff erent hydrogel entirely. Scanning electron microscopy can be used to obtain images throughout the course of the cell study to study cell adhesion on diff erent types of scaff olds as well as to see if there are any diff erences in how the cells proliferate on them. Further, in-vivo studies will have to be performed to show if the implanted scaff olds can promote tissue regeneration and maintain mechanical stability.

REFERENCES 1. Chung, E.J., Sugimoto, M.J., Koh, J.L. & Ameer, G.A. A biodegradable tri-component

graft for anterior cruciate ligament reconstruction. J. Tissue Eng. Regen. Med. 11(3), 704–712 (2014).

2. Yates, E.W., Rupani, A., Foley, G.T., Khan, W.S., Cartmell, S. & Anand, S.J. Ligament tissue engineering and its potential role in anterior cruciate ligament reconstruction. Stem Cells Int. 2012, 438125 (2012).

3. Bach, J.S., Cherkaoui, M., Corté, L., Cantournet, S. & Ku, D.N. Design considerations for a prosthetic anterior cruciate ligament. J. Med. Devices 6(4), 045004 (2012).

4. Laurencin, C.T. & Freeman, J.W. Ligament tissue engineering: An evolutionary materials science approach. Biomaterials 26(36), 7530–7536 (2005).

5. Wang, J.H.C. Mechanobiology of tendon. J. Biomech. 39(9), 1563–1582 (2006). 6. Frank, C.B. Ligament structure, physiology and function. J. Musculoskel. Neuronal

Interact. 4(2), 199 (2004). 7. Walsh, W.R. Repair and Regeneration of Ligaments, Tendons, and Joint Capsule (Humana

Press, 2006). 8. Freeman, J.W., Woods, M.D., Cromer, D.A., Ekwueme, E.C., Andric, T., Atiemo, E.A.,

Bijoux, C.H. & Laurencin, C.T. Evaluation of a hydrogel–fi ber composite for ACL tissue engineering. J. Biomech. 44(4), 694–699 (2011).

9. Freeman, J.W. Tissue engineering options for ligament healing. Bone Tissue Regen. Insights 2, 13 (2009).

10. Kwansa, A.L., Empson, Y.M., Ekwueme, E.C., Walters, V.I., Freeman, J.W. & Laurencin, C.T. Novel matrix based anterior cruciate ligament (ACL) regeneration. Soft Matter 6(20), 5016–5025 (2010).

11. Freeman, J.W., Woods, M.D. & Laurencin, C.T. Tissue engineering of the anterior cruciate ligament using a braid–twist scaffold design. J. Biomech. 40(9), 2029–2036 (2007).

12. Shirazi, A.N., Chrzanowski, W., Khademhosseini, A. & Dehghani, F. Anterior cruciate ligament: Structure, injuries and regenerative treatments. In: Engineering Mineralized and Load Bearing Tissues (ed.) Bertassoni, L.E. & Coelho, P.G. (Springer, 2015), pp. 161–186.

13. Ma, J., Smietana, M.J., Kostrominova, T.Y., Wojtys, E.M., Larkin, L.M. & Arruda, E.M. Three-dimensional engineered bone–ligament–bone constructs for anterior cruciate ligament replacement. Tissue Eng. Part A 18(1–2), 103–116 (2011).

14. Agel, J., Arendt, E.A. & Bershadsky, B. Anterior cruciate ligament injury in National Collegiate Athletic Association basketball and soccer a 13-year review. Am. J. Sports Med. 33(4), 524–531 (2005).

15. Yu, B. & Garrett, W.E. Mechanisms of non-contact ACL injuries. Br. J. Sports Med. 41(Suppl. 1), i47–i51 (2007).

16. Barber, J.G., Handorf, A.M., Allee, T.J. & Li, W.J. Braided nanofi brous scaffold for tendon and ligament tissue engineering. Tissue Eng. Part A 19(11–12), 1265–1274 (2011).

17. Liu, Y., Ramanath, H.S. & Wang, D.A. Tendon tissue engineering using scaffold enhancing strategies. Trends Biotechnol. 26(4), 201–209 (2008).

18. Shelton, W.R. & Fagan, B.C. Autografts commonly used in anterior cruciate ligament reconstruction. J. Am. Acad. Orthop. Surg. 19(5), 259–264 (2011).

19. Chen, C.-H., Chen, W.-J. & Shih, C.-H. Arthroscopic anterior cruciate ligament reconstruction with quadriceps tendon-patellar bone autograft. J. Trauma Acute Care Surg. 46(4), 678–682 (1999).

20. Fu, F.H., Bennett, C.H., Lattermann, C. & Ma, C.B. Current trends in anterior cruciate ligament reconstruction part 1: Biology and biomechanics of reconstruction. Am. J. Sports Med. 27(6), 821–830 (1999).

21. Mengsteab, P.Y., Nair, L.S. & Laurencin, C.T. The past, present and future of ligament regenerative engineering. Regen. Med. 11(8), 871–881 (2016).

22. Petrigliano, F.A., McAllister, D.R. & Wu, B.M. Tissue engineering for anterior cruciate ligament reconstruction: A review of current strategies. Arthroscopy 22(4), 441–451 (2006).

23. Cooper, J.A., Jr., Bailey, L.O., Carter, J.N., Castiglioni, C.E., Kofron, M.D., Ko, F.K. & Laurencin, C.T. Evaluation of the anterior cruciate ligament, medial collateral ligament, Achilles tendon and patellar tendon as cell sources for tissue-engineered ligament. Biomaterials 27(13), 2747–2754 (2006).

24. Legnani, C., Ventura, A., Terzaghi, C., Borgo, E. & Albisetti, W. Anterior cruciate liga-ment reconstruction with synthetic grafts. A review of literature. Int. Orthop. 34(4), 465–471 (2010).

25. Altman, G.H., Horan, R.L., Lu, H.H., Moreau, J., Martin, I., Richmond, J.C. & Kaplan, D.L. Silk matrix for tissue engineered anterior cruciate ligaments. Biomaterials 23(20), 4131–4141 (2002).

26. Cooper, J.A., Jr., Lu, H.H., Ko, F.K., Freeman, J.W. & Laurencin, C.T. Fiber-based tissue-engineered scaffold for ligament replacement: Design considerations and in-vitro evaluation. Biomaterials 26(13), 1523–1532 (2005).

27. Beasley, L.S., Weiland, D.E., Vidal, A.F., Chhabra, A., Herzka, A.S., Feng, M.T. & West, R.V. Anterior cruciate ligament reconstruction: A literature review of the anatomy, biomechanics, surgical considerations, and clinical outcomes. Oper. Tech. Orthop. 15(1), 5–19 (2005).

28. Dunn, M.G., Liesch, J.B., Tiku, M.L. & Zawadsky, J.P. Development of fi broblast-seeded ligament analogs for ACL reconstruction. J. Biomed. Mater. Res. 29(11), 1363–1371 (1995).

29. Ge, Z., Yang, F., Goh, J.C., Ramakrishna, S. & Lee, E.H. Biomaterials and scaffolds for ligament tissue engineering. Biomed. Mater. Res. Part A 77(3), 639–652 (2006).

30. Silver, F.H., Tria, A.J., Zawadsky, J.P. & Dunn, M.G. Anterior cruciate ligament replace-ment: A review. J. Long Term Eff. Med. Implants 1(2), 135–154 (1990).

31. Lu, H.H., Cooper, J.A., Jr., Manuel, S., Freeman, J.W., Attawia, M.A., Ko, F.K. & Laurencin, C.T. Anterior cruciate ligament regeneration using braided biodegradable scaffolds: In-vitro optimization studies. Biomaterials 26(23), 4805–4816 (2005).

32. Leong, N.L., Kabir, N., Arshi, A., Nazemi, A., Wu, B., Petrigliano, F.A. & McAllister, D.R. Evaluation of polycaprolactone scaffold with basic fi broblast growth factor and fi broblasts in an athymic rat model for anterior cruciate ligament reconstruction. Tissue Eng. Part A 21(11–12), 1859–1868 (2015).

Table 2 Comparison of scaff old mechanical properties and the native human ACL.

Property Braided only Twist-braid Hydrogel scaff old Hydrogel yarns Native human ACL37

UTS (MPa) 86.379 ± 5.460 70.826 ± 3.402 81.592 ± 5.956 77.411 ± 4.366 37.8 ± 9.3

Toe region (% Strain) 8.417 ± 0.753 8.007 ± 0.909 8.535 ± 1.662 9.741 ± 1.344 2 - 4.8

Strain at failure (% Strain) 49.004 ± 5.402 50.460 ± 5.302 64.641 ± 7.204 72.460 ± 4.586 44.3 ± 8.5

Modulus (MPa) 425.790 ± 38.675 315.280 ± 18.808 293.014 ± 43.429 248.902 ± 15.472 111 ± 26

1750006.indd 1051750006.indd 105 5/31/2017 10:52:59 AM5/31/2017 10:52:59 AM

Page 9: Design and characterization of three-dimensional twist-braid … · 2019-04-30 · Design and characterization of three-dimensional twist-braid scaffolds for anterior cruciate ligament

ARTICLE

106 TECHNOLOGY l VOLUME 5 • NUMBER 2 • JUNE 2017© World Scientific Publishing Co.

33. Machotka, Z., Scarborough, I., Duncan, W., Kumar, S. & Perraton, L. Anterior cruciate ligament repair with LARS (ligament advanced reinforcement system): A systematic review. Sports Med. Arthrosc. Rehabil. Ther. Technol. 2(1), 29 (2010).

34. Mooney, D.J. & Vacanti, J.P. Tissue engineering using cells and synthetic polymers. Transplant. Rev. 7(3), 153–162 (1993).

35. Vacanti, J.P. Tissue engineering and the road to whole organs. Br. J. Surg. 99(4), 451–453 (2012).

36. Kimura, Y., Hokugo, A., Takamoto, T., Tabata, Y. & Kurosawa, H. Regeneration of anterior cruciate ligament by biodegradable scaffold combined with local controlled release of basic fi broblast growth factor and collagen wrapping. Tissue Eng. Part C Methods 14(1), 47–57 (2008).

37. Noyes, F.R. & Grood, E.S. The strength of the anterior cruciate ligament in humans and Rhesus monkeys. J. Bone Joint Surg. Am. 58(8), 1074–1082 (1976).

1750006.indd 1061750006.indd 106 5/31/2017 10:52:59 AM5/31/2017 10:52:59 AM