Amin
Orash Mahmoudsalehi†
a,
Maryam
Soleimani†
b,
Kevin
Stalin Catzim Rios
a,
Wendy
Ortega-Lara
*a and
Narsimha
Mamidi
*c
aTecnologico de Monterrey, Escuela de Ingeniería y Ciencias, Av. Eugenio Garza Sada 2501 Sur, Monterrey 64849, Mexico. E-mail: wlortega@tec.mx
bSilesian University of Technology, Faculty of Mechanical Engineering, Department of Didactic Laboratory of Nanotechnology and Material Technologies, 18a Konareskiego Str, 44-100 Gliwice, Poland
cSchool of Pharmacy, Wisconsin Center for NanoBioSystems, University of Wisconsin-Madison, Madison, Wisconsin, USA. E-mail: nmamidi@wisc.edu; narsimhachem06@gmail.com
First published on 12th March 2025
Corneal stromal defects represent a significant global cause of blindness, necessitating innovative therapeutic strategies to address the limitations of conventional treatments, such as corneal transplantation. Tissue engineering, a cornerstone of regenerative medicine, offers a transformative approach by leveraging biomaterial-based solutions to restore damaged tissues. Among these, three-dimensional (3D) scaffolds fabricated using advanced techniques like 3D printing have emerged as a promising platform for corneal regeneration. These scaffolds replicate the native extracellular matrix (ECM) architecture, providing a biomimetic microenvironment that supports cell proliferation, differentiation, and tissue integration. This review highlights recent advances in the design and fabrication of 3D scaffolds for corneal stroma engineering (CSE), emphasizing the critical interplay between scaffold architecture, mechanical properties, and bioactive signaling in directing cellular behavior and tissue regeneration. Likewise, we emphasize the diverse range of biomaterials utilized in scaffold fabrication, highlighting their influence on cellular interactions and tissue reconstruction. By elucidating the complex relationship between scaffold design and biologics, this review aims to illuminate the evolution of next-generation strategies for engineering functional corneal tissue. Eventually, this review will provide a comprehensive synthesis of the current state-of-the-art in 3D scaffold-based corneal tissue engineering (CTE), offering insights that could advance progress toward effective vision restoration therapies.
To address these challenges, regenerative medicine approaches, including stem cell therapy and tissue engineering, have emerged as promising alternatives for generating bioengineered corneas or individual corneal layers.20–22 Tissue engineering, which integrates cells, bioactive molecules, and scaffolds to create functional tissues, has shown particular potential in this regard.23–25 Notable progress in corneal surgery, human corneal cell keratoplasty (HCCK), employs transparent carriers to enhance the behavior of human corneal cells (HCCs), demonstrating success in both lamellar and full-thickness corneal transplants.26–28 However, these approaches still depend on donor corneas, which remain limited in supply and are susceptible to immune rejection.29–31 Recent research has highlighted the potential of maintaining the proliferative capacity of corneal epithelial cells (CEpCs), offering a promising strategy for CTE and the repair of corneal damage (Fig. 1).32–37
Significant progress has been made in developing tissue-engineered scaffolds that replicate the structural and functional properties of the cornea.38 This review focuses on the anatomy and physiology of the corneal stroma, common ocular disorders affecting this layer, and the latest strategies for engineering stromal layer scaffolds. We emphasize the structural and biological requirements for designing effective stromal scaffolds and provide a comprehensive analysis of recent advances in scaffold design, characterization, and evaluation for CSE. Our discussion offers a novel perspective, highlighting the engineering principles underlying scaffold fabrication and the incorporation of functional additives. While fabrication methods alone may not fully dictate scaffold properties, the interplay between fabrication processes and additive selection critically influences the mechanical and biological performance of CSE scaffolds. We begin by emphasizing the structural properties essential for effective 3D scaffolds, drawing on recent studies that have explored various fabrication techniques and their impact on scaffold characteristics. Subsequently, we explore the biological engineering aspects of 3D scaffolds, reviewing the components that can enhance stromal tissue regeneration. By integrating structural and biological considerations, this review aims to advance the field of corneal stroma engineering and contribute to the development of innovative solutions for restoring vision in patients with corneal disorders.
![]() | ||
Fig. 2 (A) Illustration of the eye showing its anatomical structures, including the cornea. Reprinted with permission.14 Copyright © [2024], Elsevier. (B) Diagram of the human corneal layers: the outermost stratified squamous epithelium, followed by Bowman's layer, the stroma with keratocytes, Descemet's membrane, and the single-layer endothelium. Reprinted with permission.14 Copyright © [2024], Elsevier. (C) Overview of corneal tissue transplantation techniques: PK involves full-thickness corneal replacement of the central part; ALK replaces only the stroma; DSAEK and DMEK are partial-thickness transplants targeting the endothelium, with DMEK providing a pure endothelial replacement without stromal cells. Reprinted with permission.14 Copyright © [2024], Elsevier. (D) Progression of in vitro HCCs culture techniques. The shift from conventional 2D CEC culture to an advanced 3D corneal structure requires accurately mimicking the natural in vivo environment. Critical factors include scaffold-cell interactions, cell–cell communication—such as endothelial-stromal signaling—and endothelial-Descemet's membrane connections, as well as cell-surface interactions, all of which contribute to developing a 3D corneal graft. The concept of 4D CEC culture introduces a dynamic aspect, where the corneal structure evolves. This self-adapting transformation depends on carefully engineered bioactuators, and dynamic cell-scaffold interactions. Reprinted with permission.42 Copyright © 2020, Springer Nature. (E) Summary of bioengineering strategies for corneal tissue replacement. Reprinted with permission.14 Copyright © 2024, Elsevier. (F) Depiction of an injured human cornea undergoing either uncontrolled or controlled CSE. Uncontrolled regeneration (top) may result from untreated severe stromal injuries or unsuitable implants, leading to irreversible keratocyte-to-(myo)fibroblast transitions and disorganized ECM production. Controlled regeneration (bottom) is supported by a well-integrated bioengineered implant that promotes keratocyte infiltration, reversible keratocyte-to-fibroblast transition, and organized ECM deposition. Reproduced with permission.43 Copyright © 2024, Wiley. (G) Common materials, modifications, and fabrication techniques used for designing stromal constructs with optimal mechanical, optical, and biocompatible properties. Reproduced with permission.43 Copyright © 2024, Wiley. |
The stromal layer, at approximately 500 μm, is the thickest, while the epithelium and endothelium are less than 50 μm thick. Occupying over 90% of the cornea, the stromal layer is crucial for corneal main functions due to its transparency and mechanical strength, influenced by COL alignment and fibril cross-linking. Its mechanical properties vary across the peripheral and central sections, affecting COL orientation and corneal cell behavior.9 The stroma contains almost 8% CKCs and is an acellular but dense connective layer derived from neural crest cells.44 Consisting of over 200 noncellular collagenous lamellae with uniformly aligned small COL fibers, the stroma's flattened fibroblasts are activated during injuries to produce COL, stabilize collagenous lamellae, and secrete stromal components.45 A healthy stroma exhibits optical transparency and suitable mechanical strength, which is crucial for maintaining light transmittance. Challenges for tissue engineers include achieving equal mechanical stability and high optical transparency.46
The corneal stroma comprises both extracellular and cellular elements. Cellular components within the mature corneal stroma are known as CKCs.40 These cells, with a dendritic morphology, play a crucial role in maintaining the ECM of the stroma.47 CKCs actively produce keratocan and lumican, essential factors in preserving the shape and transparency of the stroma. Keratocan and lumican, belonging to the small leucine-rich protein family, serve as vital keratan sulfate proteoglycans in the corneal stroma.48 While keratocan is uniquely found as a proteoglycan in the cornea, lumican may also exist as a glycosylated protein in various tissues. Both keratocan and lumican interact with COL fibrils, regulating the tissue's structure within defined limits for specific properties. Notably, keratocan is crucial in preserving the corneal structure, as previous studies have shown.48–50
During wound healing processes, the dendritic morphology of CKCs transforms into a fibroblastic appearance. This transformation is associated with a decrease in two critical functions of keratocytes—expression of keratocan and synthesis of keratan sulfate-highlighting changes during fibroblast/myofibroblast transformation.46 Both isolated keratocytes from the corneal stroma and cultured keratocytes exhibit fibroblastic/myofibroblast phenotypes, demonstrating decreased keratocan expression and keratan sulfate synthesis, mirroring in vivo wound healing processes. This underscores that keratocan can serve as an indicator of the native keratocyte phenotype.43,51,52
Viewing it from a material perspective (Fig. 2(E)), studies have demonstrated that natural biomaterials and dC tissues enhance critical aspects like cell adhesion, viability, and differentiation.42,79–81 However, limitations such as poor mechanical properties and a high degradation rate are associated with these natural biomaterials.82,83 Conversely, synthetic biomaterials exhibit favorable attributes like acceptable mechanical strength, a low degradation rate, and tunable geometry, making them appealing to researchers.80,83 Nonetheless, challenges such as the absence of cell binding sites and potential inflammatory responses leading to graft rejection are considerations with synthetic biomaterials.83 Consequently, many studies explore the synergistic use of various biomaterials to regenerate damaged corneas, with some even opting for approaches that eliminate the necessity for any biomaterial in the CSE process.84 Characterizing scaffolds, encompassing factors such as stiffness, surface topology, degradation rate, and cytocompatibility, plays a pivotal role in influencing cell differentiation and growth in scaffold-based cell delivery methods.79 Optimal mechanical properties are essential to emulate the microenvironment of the host tissue and promote the differentiation of cells into the desired cell type.80 Considering the tensile strength of the cornea, adjusting the scaffold's stiffness through variations in polymer/crosslinker concentration or crosslinking time significantly impacts cell differentiation and the reconstruction of damaged tissue.75 The corneal stroma's mechanical properties, such as strength, elasticity, and viscoelasticity, are primarily governed by its unique architecture and composition at microscopic and macroscopic levels. COL fibers provide strength and elasticity, while proteoglycans and cellular components contribute to viscoelastic properties. However, most bioengineered stromal tissues lack cellular elements, making viscoelasticity less relevant and highlighting the importance of strength and elasticity. The Young's modulus and ultimate tensile strength are commonly used metrics to assess these properties, where the Young's modulus represents resistance to elastic deformation and varies widely (0.1 to 57 MPa) depending on the stromal region, donor age, and testing methods.81 The tensile strength, indicating the force required to break the tissue, is typically around 3.8 MPa. These mechanical parameters are crucial for maintaining corneal structure and supporting keratocyte functions such as adhesion, proliferation, and differentiation. Changes in stiffness can affect keratocyte behavior. Consequently, designing bioengineered stromal implants with suitable mechanical characteristics is essential to ensure long-term functional integration, provide biomechanical cues for keratocyte activity, and withstand clinical handling and surgical procedures without compromising structural integrity.81
The structural features of scaffolds significantly influence cellular behavior, as CSCs in vivo are situated between layers of well-organized COL fibrils (Fig. 2(F)). To simulate these topographical cues in vitro, researchers have employed channels at both micro- and nanoscale dimensions. Such aligned substrates have been shown to enhance stromal cell alignment and direct migration, as well as support the maintenance of a keratocyte phenotype. Similarly, nanoscale channels designed to mimic the architecture of the corneal basement membrane have demonstrated the ability to regulate corneal cell elongation, adhesion, proliferation, and gene expression. To create a more physiologically relevant environment, 3D scaffolds incorporating topographical features have been developed. For example, Wilson et al. incorporated aligned nanofibers into hydrogels, resulting in an upregulation of keratocyte-specific genes and a reduction in myofibroblastic gene expression.94 Beyond surface topography, recent research has shown that surface curvature can also influence the orientation and phenotypic behavior of both corneal stromal and epithelial cells, further underscoring the importance of scaffold design in modulating cellular responses.
When designing scaffolds for corneal regeneration, it is essential to consider their light transmittance properties, ensuring they mimic the natural cornea's ability to transmit visible light (400–780 nm) while limiting UV light (<400 nm). UV transmission varies between the central and peripheral regions of the cornea, and excessive UV exposure can damage the retina.76 Additionally, the cornea's shape is critical for focusing light onto the retina, with irregularities like astigmatism causing blurred vision. The transparency of the cornea largely depends on the highly organized COL fibril arrangement in the stroma. Small, evenly spaced fibrils allow light to pass through; any disruption from injury or disease reduces transparency.62 Moreover, keratocytes in the stroma contain crystalline proteins, such as ALDH1A1 and ALDH3A1, which minimize light scattering. When these cells are activated, the reduction in these proteins leads to increased light scattering, affecting corneal clarity.65–67
Biodegradability stands out as a critical aspect of scaffold-based CSE.95 An ideal cell carrier should undergo degradation as transported cells secrete ECM, allowing the produced matrix to replace the biomaterial.96 The interplay between mechanical properties and degradation is noteworthy, and this relationship can be carefully controlled to ensure proper healing of damaged tissue.50 Additionally, the swelling ratio is influenced by the stiffness and hydrophilicity of the scaffold, directly impacting the material's water content.79 This water content, crucial for biocompatibility and cell growth, is directly related to the scaffold's hydrophilicity and inversely related to its mechanical strength.97 Notably, the equilibrium water content of the cornea is reported to be 80%.82 Scaffold design not only influences cellular behavior but is also subject to remodeling by the cells through the application of mechanical forces, enzymatic activity, and the secretion of ECM components. Therefore, scaffolds must be engineered to support cellular remodeling while maintaining their structural integrity, ultimately resulting in tissue formation that closely resembles the native cornea.24 As the scaffold degrades, it should ideally be replaced by newly synthesized ECM, ensuring that the tissue retains its functional properties and transparency. Proper alignment of COL fibrils is crucial for maintaining corneal clarity. It can be enhanced by initially structuring the scaffold in a manner that mimics the organization of natural COL fibers.36 It should be noted that COL hydrogels with low polymer concentrations tend to contract into opaque masses unless stabilized by external supports such as rings. Parameters like cell seeding density, culture medium composition, scaffold degradation rates, material concentration, and porosity must be meticulously tuned to guide the remodeling process. While certain scaffolds are designed to be stable and resistant to degradation, they may lack the ability to integrate with host tissue and facilitate long-term regeneration.34 Although such designs may offer advantages, such as minimizing the risk of premature degradation and reducing variability in clinical outcomes, they are likely to fail due to the absence of dynamic tissue regeneration. Therefore, scaffolds engineered to promote controlled remodeling and support tissue regeneration present a more viable approach for CSE.
The cytocompatibility of biomaterials emerges as another critical parameter influencing not only cell viability but also the material's ability to support cell growth, migration, and ECM deposition during tissue regeneration factor, significantly influencing cell–cell and cell–tissue interactions. The careful consideration and optimization of these scaffold characteristics are imperative to enhance the overall effectiveness of CSE. Therefore, ensuring appropriate surface topology and properties becomes crucial for promoting cell growth and differentiation in the context of CSE.50,80,81 Thus, the careful selection of materials with suitable properties, combined with advanced scaffold fabrication techniques, is essential to developing constructs that closely mimic the unique characteristics of corneal stromal tissue. Achieving the appropriate balance of transparency, mechanical strength, and biocompatibility is critical to creating structures that meet the functional and structural requirements of the corneal stromal tissue. By addressing these considerations, researchers can advance the field of CSE and contribute to the development of innovative solutions for treating corneal disorders (Fig. 2(G)).
Biomaterials | Pros | Cons | Ref. |
---|---|---|---|
COL | • Biodegradable | • Unstable | 101–106 |
• Biocompatible | • Degrades rapidly | ||
• Encourages cell adhesion | • Expensive | ||
• Easy to produce | • Shrinks at cell introduction | ||
• Component of stroma, therefore compatible with corneal cells | • Poor mechanical toughness and elasticity | ||
• Can be crosslinked to improve properties | |||
• Can be organized to match native tissue | |||
GEL | • Cost-effective compared to COL | • Degrades at a faster rate than COL due to the removal of cross-linkages. | 102 and 107 |
• Derived from COL | |||
• Possesses the same amino acid structure as COL | |||
• Exhibits superior transparency compared to COL | |||
• Demonstrates a good elastic modulus | |||
CHI | • Found naturally in insects, algae, fungi, and crustacean shells | • Can swell in aqueous conditions if not crosslinked or combined with a stabilizer | 108–110 |
• Non-toxic | • Degrades in less than 8 weeks | ||
• Biodegradable | • Compact internal structure leading to reduced cell proliferation | ||
• Encourages wound healing | |||
• Transparent | |||
• Exhibits good mechanical strength | |||
HA | • Naturally occurring in the vitreous humor of the eye | • Occasionally, it may trigger allergic reactions. Additionally, when injected into the eye, it has the potential to elevate intraocular pressure. | 111 |
• Demonstrates effective electrospinning capability | |||
SF | • Highly versatile material that can be molded into various scaffold types, and cost-effectivy | • Vulnerable to UV radiation, lacking durability | 112–114 |
• Biodegradable with good cell viability | • Exhibits opacity as a material characteristic | ||
• Robust in water | • Limited biocompatibility | ||
• Exhibits good mechanical properties | |||
• Provides porous scaffolds | |||
• Can be combined with COL to enhance cell attachment and proliferation | |||
CS | • Resembles the structure of HA | • During storage, there has been an observed increase in corneal thickness. No reported side effects or disadvantages have been noted. | 115 |
• Naturally occurring in the ECM | |||
• Employed for therapeutic treatment of disorders | |||
• Transparency progressively enhances over time | |||
dC | • Exhibits properties comparable to the native cornea | • Dependent on donations from human corneas | 116–118 |
• Demonstrates low rates of immune rejection | • Mandates screening for diseases | ||
• Offers improved transparency | |||
PLA | • Biodegradable | • Takes 10 months to fully degrade | 119 and 120 |
• FDA approved | • Exhibits poor cell adhesion without surface modification | ||
• Naturally passes from the body during degradation | |||
• Flexible | |||
• Can be produced from plant extracts | |||
• Can be either semi-crystalline or amorphous | |||
PLGA | • Lower stiffness in comparison to PGA | • Shrinkage may occur upon cell introduction | 121 |
• Degradation rate can be adjusted by altering the PLA to PGA ratio | • Exhibits poor cell adhesion without surface modification | ||
• Non-harmful to the body | • Displays inadequate mechanical strength | ||
PCL | • Biocompatible | • Extremely slow degradation speed | 122 and 123 |
• Biodegradable | • Low-melting temperature | ||
• Exhibits good cell viability | • Highly hydrophobic | ||
• Approved as a cell carrier for the retina and conjunctiva due to non-toxicity | • Poor transparency | ||
• Flexible | • Exhibits poor cell adhesion without surface modification | ||
• Cost-effective | |||
• Capable of producing fine fibers when electrospun | |||
PVA | • Biocompatible | • Water-soluble | 124 |
• Exhibits good transparency | • Rapid degradation | ||
• Capable of producing nanofibers when electrospun | • Limited mechanical strength | ||
• Poor cell adhesion without surface modification | |||
PGS | • Elastic | • Challenging to obtain | 125 |
• Transparent polymer | • Requires crosslinking | ||
• Biocompatible | • Rapid degradation | ||
• Biodegradable | • Poor cell adhesion without surface modification | ||
• Adaptable for controlling degradation rates and mechanical strength | |||
• Utilizable as a shape memory polymer | |||
PLLA | • Isomer form of PLA | • Less transparent than PLA | 126 |
• Capable of producing fibrous scaffolds for wound healing | • Degrades at a slower rate than PLA | ||
• Transparency increases over time | • Poor cell adhesion without surface modification | ||
• Exhibits good biocompatibility | |||
pNIPAM | • Promotes cell adhesion and growth | • Monomer is toxic to neural tissue | 127–129 |
• Non-cytotoxic and biocompatible | • Considered expensive | ||
• Easy to manufacture | |||
• Relatively simple to adjust mechanical properties | |||
• Attachment properties can be altered by changing tissue temperature | |||
• Exhibits thermosensitive hydrophobic/hydrophilic changes | |||
• Can be combined with other materials to enhance properties | |||
PGA | • Biodegradable | • Excessively rigid | 109 and 110 |
• FDA approved | • Mechanical properties deteriorate after 6 weeks | ||
• Naturally passes from the body during degradation | • Exhibits poor cell adhesion without surface modification | ||
• Exhibits greater strength compared to PLA | |||
• Requires 4 months for complete degradation | |||
• Achieves transparency approximately 8 weeks into the degradation process |
Type of material | Material source | In vivo test | Ref. |
---|---|---|---|
COL | Porcine Type I | Rabbit | 130 |
Rat Type I | Rabbit | 131 | |
Bovine Type I | Rabbit | 132 | |
Bovine Type I | Dog | 133 | |
Rat Type I | Rabbit | 134 | |
Bovine Type I | Pig | 135 | |
Porcine Type I | Pig | 136 | |
Recombinant human Type I and III | Mini-Pig | 137 | |
Recombinant human Type III | Human | 138 | |
SF | Antheraea mylitta | Rabbit | 139 |
Bombyx mori and CHI | Rabbit | 140 | |
Bombyx mori and COL | Rabbit | 141 | |
Bombyx mori | Rabbit | 142 | |
Bombyx mori and CHI | Rabbit | 143 | |
Bombyx mori and RGD peptides | Rabbit | 144 | |
GEL | GelMA | Rabbit | 145 |
GEL | Rabbit | 146 | |
GEL + GAG | Rabbit | 147 | |
dC | PdCT | Rabbit | 148 |
PdCT | Dog | 149 | |
HdCT | Rabbit | 150 | |
HdCT | Human | 151 | |
HdCT | Rabbit | 152 | |
PdCT | Rabbit | 153 | |
PdCT | Human | 154 | |
PdCT | Nonhuman primate | 155 | |
PdCT | Rabbit | 156 | |
HdCT | Rabbit | 157 | |
PdCT | Rabbit | 158 |
Though COL's mechanical strength and elasticity may be insufficient for certain applications, crosslinking techniques have been employed to enhance its properties without compromising its biological benefits.89,92 Chemical crosslinkers such as glutaraldehyde and genipin are widely used; however, their application in corneal bioengineering is limited due to the potential for toxic residues.94,98 In contrast, physical crosslinking methods, such as UV treatment and enzymatic crosslinking using transglutaminase, offer alternatives with fewer toxic by-products (98). For example, Duan et al. demonstrated that different crosslinking methods yield varying optical and mechanical properties of COL membranes. Among these, EDC crosslinking has been noted for producing superior results, closely resembling the mechanical properties of the native human cornea.90 UV-induced glucose crosslinking has also been shown to produce COL membranes with favorable mechanical and optical characteristics while supporting corneal stromal cell adhesion, proliferation, and ECM secretion.88 Recent advances in COL-based scaffolds have focused on improving their mechanical resilience and transparency. For instance, COL composites incorporating substances like chondroitin sulfate and tobramycin have been developed to enhance mechanical properties and transparency. Additionally, COL-vitrigel composites offer improved strength and clarity, making them suitable for corneal tissue engineering applications. These modifications not only enhance the mechanical properties of COL scaffolds but also improve their ability to support cellular interactions, ECM deposition, and tissue remodeling.89,100,159 Despite these improvements, challenges remain in accurately replicating the complex structure of the cornea. Researchers are actively exploring methods to mimic the intricate organization of COL fibrils in the corneal stroma, which is critical for achieving transparency and functionality. For example, techniques such as electrospinning and 3D bioprinting are being investigated to create COL-based scaffolds with aligned fibrillar structures that closely resemble the native corneal ECM.160–170
![]() | ||
Fig. 4 (A) Pathway for creating patient-specific devices using 3D printing technology. Reprinted with permission.192 Copyright © [2024], Elsevier. (B) Schematic of the basic techniques of 3D printing: Droplet printing (thermal (a), piezoelectric (b)), Extrusion printing (piston (c), pneumatic (d)), and Laser printing (stereolithographic (e)).193 |
On the other hand, bioprinting, which utilizes cellular encapsulated biological materials as bioinks is considered a suitable platform for the simulation of various tissues.194 Scaffolds printed with cells are produced in situ, necessitating a sterile environment compatible with the cells. Maintaining structural integrity and desired mechanical properties in the printed structure limits the selection of cell-compatible materials.195 Appropriate rheological parameters must be chosen to minimize shear pressure during the printing process. However, the resolution of the printed substrate is compromised in cell-loaded bioprinting. Additionally, achieving a critical cell density relative to the surface area is paramount.11,196 The standard healthy threshold for cell density in solid organs is approximately 109 to 1010 cells per cell culture well. Bioprinted hydrogel scaffolds, however, typically exhibit a cell density ranging between 105 and 107 cells per cell culture well, falling short of this threshold.197
Bioprinting techniques have been successfully applied by researchers, demonstrating reliable properties for generating ex vivo constructs and membranes.198 Various approaches, including microextrusion, laser-based methods, and droplet-based techniques, have shown promising results in creating constructs for neural studies, 3D models of interacting HECs, cancer studies, and replication of the native ECM of cartilage.199–201 Bioprinted scaffolds, considered a top-down approach, is recognized as a biofabrication technology for artificially fabricating various ex vivo membranes and tissues through consecutive deposition of cell-loaded layers.202–205 Different bioprinting methods, such as laser-based, droplet-based, and extrusion-based techniques, are compatible with various bioinks, each with unique challenges in optimization according to specific bioprinting techniques (Fig. 4(B)).193 Extrusion-based methods are particularly popular due to their compatibility with most injectable hydrogel platforms for biomedical engineering and regenerative medicine applications.206–209 In brief, this method involves extruding pre-polymerized bioink through a single nozzle under pressure. Pressurized air can be applied to the printer head, producing a 3D construction layer by layer.
Corneal bioprinting presents a promising avenue for addressing contemporary challenges and requirements in CSE, encompassing a controllable structure, properties akin to natural mechanical strength, and the ability to fabricate fully organized corneal constructs.193 As delineated in prior sections, the cornea. Comprises three transparent layers, with the stromal layer being the thickest (∼500 μm), and the delicate epithelium and endothelium layers measuring less than 50 μm.4 The stromal layer, constituting over 90% of the corneal structure, is of paramount importance in CTE due to its transparency, resulting from aligned COL lamellae and proteoglycan expressions, and mechanical performance, facilitated by cross-linked COL fibrils.46 It should be noted that the mechanical properties vary between the peripheral and central sections of the stromal layer, influencing COL content orientation, corneal cell differentiation, and alignment.8,210 Hence, accurate replication of the micro and macrostructure is crucial for enhancing the functionality of the stromal part, as the physical, mechanical, and chemical properties significantly impact biological factors.211 Recent studies (Table 3) have explored various bioprinting techniques for CSE, considering factors such as structure, transparency, and mechanical properties.1,44,87,212–222 Sorkio et al. investigated the potential of laser-assisted bioprinting for CSE using COL type-I and LSCs. Their study demonstrated the feasibility of constructing 3D cornea-like structures by combining hESC-LESCs and hASCs. By employing optimized bioinks based on recombinant laminin and human COL type-I, the researchers successfully bioprinted layered structures, including stratified epithelium, lamellar stroma, and combined epithelial-stromal tissues. The constructs showed promising cellular viability, proliferation, and key protein expressions, with CKCs organizing horizontally, mimicking native stromal structures. However, the sensitivity of CKCs impacted the mechanical properties and structural integrity of the bioprinted constructs, highlighting challenges in achieving fully functional tissue substitutes. Although the stromal layers adhered to porcine corneal organ cultures and showed initial signs of integration, the mechanical performance and stability of the bioprinted tissues require further refinement. This study underscores the promise of laser-assisted bioprinting in CSE while emphasizing the need for continued optimization to achieve structures with enhanced mechanical and biological properties (Fig. 5).212
Bioprinting method | Material | Cell source | Results | Ref. |
---|---|---|---|---|
Extrusion | SOALG, COL bioink, | CKCs | • Exhibits a structural resemblance to the native corneal architecture, with stromal cells encapsulated utilizing a COL-based bio-ink | 192 |
FRESH support | • Demonstrates notable viability of stromal cells within the construct | |||
Laser | Matrigel, COL bioink | LECs | • The printed membranes exhibited commendable cell viability and positive COL labeling | 212 |
• Encountered challenges associated with insufficient transparency | ||||
Extrusion | COL, dC | CKCs | • The differentiation potential of hTMSCs was exclusively evident when using the dC-COL membrane. | 220 |
• The COL-dC scaffold demonstrated both adequate mechanical flexibility and enhanced transparency properties, surpassing those of the conventional COL scaffold | ||||
Droplet | COL, AG | CKCs | • Maintaining the native keratocyte phenotype along with appropriate cellular elongation was achieved | 213 |
• Comparable transparency was observed in relation to the stromal layer | ||||
Extrusion | GelMA, reinforced with PEG, PCL | LSSCs | • Creating an optimal environment conducive to the preservation of the keratocyte phenotype | 219 |
Fibers | ||||
Extrusion | GelMA | CKCs | • Keratocytes exhibited sustained preservation of their phenotype | 44 |
• Demonstrated transparency akin to the native cornea | ||||
• Possessed satisfactory mechanical stability | ||||
Extrusion | dC | CKCs | • Determining the ideal nozzle diameter for the bioprinting of cornea-like aligned COL fibrils | 223 |
• Identifying the optimal nozzle diameter essential for maintaining the morphology and phenotype of keratocytes | ||||
• Achieving outstanding transparency | ||||
• Preserving the keratocyte phenotype throughout the process | ||||
Extrusion | HA | hASCs and hASC-dCSKCs | • The bioink exhibited remarkable characteristics, including superior shear-thinning properties, viscosity, printability, shape fidelity, and self-healing capabilities, all while maintaining high cytocompatibility. | 224 |
• The cells within the printed structures demonstrated robust tissue formation, and the 3D bioprinted cornea structures showcased exceptional ex vivo integration with host tissue, along with notable in vitro innervation. | ||||
Extrusion | GelMA and MC | CKCs | • GelMA7/MC8 hydrogel exhibited superior mechanical strength, with a maximum compressive strength of 69 kPa, compared to 54 kPa for GelMA5/MC8, attributed to higher cross-linking density. | 225 |
• Light transmission improved to ∼80% by day 14, while biocompatibility assessments showed excellent cell adhesion and proliferation, with GelMA5/MC8 supporting higher viability due to its larger pore size (∼674 nm) compared to GelMA7/MC8 (∼412 nm). | ||||
Laser | dC and GelMA | hCFs | • In vitro experiments underscored the hydrogel's ability to sustain elevated cell viability and express core proteins. | 226 |
• In vivo assessments suggested that the hydrogel holds potential for fostering epithelial regeneration, maintaining matrix alignment, and restoring clarity. | ||||
• dC plays pivotal role in creating a conducive environment that promotes the transformation of cellular function. | ||||
Laser | GelMA | CSCs | • The cytocompatibility and elongation of CSCs within 12.5% GelMA scaffolds denote robust cell attachment, growth, and integration within the scaffold. | 227 |
• Over time, there was a noteworthy increase in the gene expression levels of COL type I, lumican, and keratan sulfate in cells cultured within 12.5% GelMA scaffolds compared to those cultured on conventional plastic tissue culture plates. | ||||
Laser | HA-aldehyde HA-ALD | CKCs | • The storage modulus of cell-laden constructs increased from 0.069 ± 0.005 kPa to 0.141 ± 0.184 kPa after 14 days, demonstrating that cell proliferation and tissue formation significantly enhanced the mechanical properties of the bioprinted structures. | 228 |
• Despite high cell proliferation, bioprinted constructs retained transparency within a 67–84% transmittance range, aligning with corneal transparency classification and supporting their potential application in corneal tissue engineering. | ||||
Extrusion | dC | CKCs | • Developed a corneal stroma-specific bioink using hASC–CKCs-derived ECM, eliminating donor corneas and enabling scalable, cost-effective production. | 229 |
• Ensured low immunogenicity with chemical and enzymatic methods, preserving ECM composition for clinical use. |
![]() | ||
Fig. 5 (A) Illustration of the laser-assisted bioprinting system to create 3D structures that mimic the corneal stroma. These structures consist of ten alternating layers of hASCs and a cell-free bioink. Each cell-containing layer has two layers of hASCs, while the acellular layer comprises four printed layers. (B) 3D bioprinted stroma within a porcine corneal organ culture model after seven days (a)–(c), compared with blank Matriderm® sheets (d)–(f) and acellular bioink used as controls (g)–(i). Immunohistochemical staining for the human cell marker TRA-1-85 (shown in brown) indicates the successful integration of hASCs into the porcine corneal stroma in the bioprinted structures. In contrast, Matriderm® and acellular bioink showed limited stromal interaction. HE staining of the acellular bioink in the organ culture revealed porcine epithelial overgrowth—scale bars: 200 μm. (C) 3D cornea bioprinted from hESC-LESCs and hASCs using laser-assisted bioprinting. The 3D cornea printed on a PET substrate (a) demonstrates moderate transparency, whereas printing on the non-transparent Matriderm® substrate (b) was necessary to prevent structural shrinkage during culture. Panel (D) compares the bioprinted corneal tissue with a native human cornea. Immunofluorescence staining on cryosections shows multilayered corneal progenitor marker p40 (red)-positive hESC-LESCs on the 3D scaffold two days post-printing. HE staining highlights the tissue structure in the bioprinted sample, while cryosections of the human cornea serve as a control. Scale bars 200 μm. Reprinted with permission.212 Copyright © [2024], Elsevier. |
Isaacson et al. explored the application of 3D bioprinting in CSE, employing a bioink composed of ALG-COL type-I with encapsulated CKCs. Using the FRESH method, the bioink was injected into a 3D mold to fabricate corneal structures mimicking the native human corneal stroma. The constructs achieved promising transparency and maintained high CKC viability, with over 90% viability observed on day 1 and 83% on day 7 post-printing. These results highlight the feasibility of 3D bioprinting as a novel method for creating synthetic corneal prostheses. However, despite the initial success in transparency and cell viability, the constructs were unable to adequately support the CKCs in adopting their native dendritic morphology, a critical characteristic for functional corneal stroma. This limitation underscores the challenges in replicating the intricate cellular architecture and biomechanical properties of the native tissue. While the study demonstrated significant progress in using 3D bioprinting to engineer artificial corneal structures, further optimization of the bioink composition and printing parameters is required to achieve fully functional and clinically translatable corneal tissue substitutes (Fig. 6A and B).192
![]() | ||
Fig. 6 (A) Process of support structure creation: (a) Human cornea, showing natural size and curvature; (b) Initial corneal model converted to a solid for Boolean operations; (c) Cornea sealed with a planar circle for volume subtraction; (d) Wireframe view of cornea centered within a cuboid before subtraction; (e) Digital support structure after volume subtraction; (f) 3D-printed plastic support structure. Reprinted with permission.192 Copyright © [2024], Elsevier. (B) Utilizing the support structure for 3D printing a corneal model with 3% alginate (200 μm nozzle diameter) and refining bio-inks for corneal printing. (a) Digital cornea loaded into 3D printer software slicer, showing concentric print direction preview; (b) support structure coated with FRESH hydrogel to aid in 3D bioprinting of corneal structures; (c) image of 3D printing in progress, using 3% AG bio-ink with trypan blue for visibility; (d) 3D-printed corneal structure before incubation; (e) FRESH hydrogel aspirated after 8 minutes of incubation, with corneal structure carefully removed, though structure begins to unravel one day post-printing with keratocytes in alginate bio-ink; (f) images of corneal structures printed with composite bio-inks; (g) Correlation between nozzle diameter and printed thickness of corneal structures (left) and transparency of structures printed with COL Type I bio-ink; (h) brightfield image of 3D-printed corneal structure with cells on day 1 and cell viability over 7 days (right); (i) live/dead staining fluorescence images at days 1 and 7 in COL Type I bio-ink. Reprinted with permission.192 Copyright © [2024], Elsevier. (C) CAD design and slicing for corneal bioprinting: (a) dome-shaped artificial cornea designed in CAD; (b) model sliced using custom software compatible with DoD bioprinting, where alternating blue and red drops are printed with spaces calculated for green and beige filling drops; (c) and (d) 3D structures bioprinted with human corneal stromal keratocytes showing transparency and optical qualities. Transparent dome-shaped artificial corneas consist of 0.5% AG and 0.2% COL Type I. (e) Live/dead staining of human CKCs one day after printing. Immunohistochemical and immunocytochemical staining on human corneal tissue and CKCs-loaded bioprinted samples: (f) keratocan (Kera), (g) lumican (Lum), and (h) SMA staining on human corneal tissue section (5 μm); (i) Kera, (k) Lum, and (m) SMA staining on CSK-loaded AG-COL blends 7 days post-printing. Reproduced with permission.213 Copyright © 2024, Wiley. |
Campos et al. tackled the challenges of maintaining CKCs vitality and proliferation by employing a droplet-based bioprinting technique to fabricate a COL type-I-AG bioink construct. Using a layer-by-layer printing strategy, they successfully created a dome-shaped structure miming the native corneal stromal architecture. The printed constructs displayed cell vitality and proliferation rates comparable to control samples, maintaining high levels of CKCs viability throughout the culture period. Notably, the bioprinted keratocytes expressed lumican and keratocan, essential markers for keratocyte functionality and stromal tissue integrity, underscoring the suitability of the approach for producing biomimetic corneal tissue (Fig. 6(C)). This innovative drop-on-demand technique enabled the fabrication of translucent corneal stromal equivalents with optical properties comparable to native tissue, as demonstrated by optical coherence tomography. The bioprinting method preserved CKCs phenotypes, maintaining their native characteristics for up to 7 days in vitro. These results highlight the potential of droplet-based bioprinting to generate 3D corneal models as viable alternatives to donor tissue for patients with corneal stromal diseases. While the study demonstrated substantial progress in replicating the structural and functional aspects of the human corneal stroma, further work is required to optimize the constructs for long-term functionality and integration in clinical settings.213
Regenerating the stromal layer poses a considerable challenge due to its intricate microstructure comprising randomly oriented COL lamellae.8 Kong et al. developed a novel approach for CSE by fabricating an aligned PCL-PEG microfibrous scaffold infused with 15% GelMA hydrogel. This 3D fiber-hydrogel construct was engineered to mimic the intricate architecture of the native corneal stroma, with fiber spacing optimized for structural and functional resemblance. The scaffold demonstrated enhanced mechanical strength, suture-ability, and high transparency, making it a promising candidate for CSE. LSCs seeded on the construct exhibited improved expression of keratocyte-specific markers, suggesting that the scaffold provided an optimal environment for maintaining the keratocyte phenotype while preventing fibroblast transformation (Fig. 7(A)). The study also explored the synergistic effects of scaffold topology and serum-free media on keratocyte phenotype maintenance and stromal regeneration. The fiber-hydrogel scaffold, combined with serum-free culture conditions and chemical factors like ascorbic acid, insulin, and β-FGF, created a microenvironment conducive to LSC differentiation into keratocytes. Both in vitro and in vivo evaluations revealed that the construct facilitated the regeneration of damaged corneal stroma with properties closely resembling native tissue. These findings highlight the potential of aligned PCL-PEG fibers infused with GelMA as a biomimetic platform for corneal repair, addressing key challenges in CSE through mechanical robustness and biological compatibility.219
![]() | ||
Fig. 7 (A) The custom-designed direct writing device is shown along with a schematic of the direct writing process, illustrating a polymer jet ejected from a Taylor cone. (b) Images of the produced curly lines (scale bar: 500 μm), grids (scale bar: 500 μm; enlarged image: 150 μm), straight lines, and a specially designed pattern (scale bar: 10 mm). (c) Images demonstrate stable jet deposition at different substrate moving speeds under specific conditions (scale bar: 100 μm). (d) SEM images of grid scaffolds with fiber spacings ranging from 50–500 μm at varying magnifications (scale bar: 50 μm). (e) LSSCs inoculated on the grid scaffolds align along the fiber direction, stained with phalloidin (red) and DAPI (blue) (scale bar: 100 μm). (f) Macroscopic and microscopic morphology of GelMA hydrogels at 5%, 10%, and 15% concentrations (scale bar for macroscopic images: 10 mm; SEM images: 80 μm). (g) Viability of LSSCs within the 5% GelMA hydrogel, with live cells in green and dead cells in red (scale bar: 500 μm). (h) Cytoskeleton (red) and DAPI (blue) staining of LSSCs within the 5% GelMA hydrogel (scale bar: 100 μm). (i)–(l) Fabricated 100 μm-spaced grid fiber-reinforced GelMA hydrogel, with scale bars in (i)–(k) at 10 mm, the enlarged image in (k) at 1 mm, and (l) at 100 μm. Staining in (e), (g), and (h) was repeated three times with consistent results. Reprinted with permission.219 Copyright © 2024, Springer Nature. (B) Schematic illustration of the preparation process for human keratocyte (HK)-loaded hydrogels, showing (a) GelMA slabs and (b) 3D bioprinted GelMA hydrogels Reproduced ref. 44 with permission from the Royal Society of Chemistry. (C) (a) Schematic of dC ECM gel preparation and validation, (b) H&E-stained images using a rabbit model, optical micrographs, OCT images with H&E staining on day 28, and (c) quantification of immune cells on days 14 and 28 (scale bar: 50 μm). Reproduced with permission.221 Copyright © 2024, SAGE Publications. |
Recent advancements in bioprinting have enabled the creation of corneal stroma equivalents with enhanced biological and mechanical properties. A notable study developed 3D bioprinted constructs using GelMA hydrogels optimized for CKCs printing by fine-tuning parameters such as nozzle speed and spindle speed. These constructs exhibited exceptional stability in physiological conditions, with only an 8% weight loss over three weeks and a 98% cell viability on day 21. Mechanical properties of the cell-loaded hydrogels increased twofold during incubation, reaching a tensile strength close to that of the native cornea. The constructs also maintained transparency above 80% at 700 nm, comparable to the native cornea. They supported the expression of key corneal markers such as COL types I and V and decorin, demonstrating their potential to replicate the biological and optical properties of the corneal stroma (Fig. 7(B)).44
Another innovative approach employed a cornea-derived dC ECM bioink for CSE, leveraging its composition rich in COL and glycosaminoglycans to mimic the native cornea closely. The dC ECM bioink supported the differentiation of hTMSCs into CKCs, showcasing its cornea-specific capabilities. In vivo evaluations in mice and rabbits demonstrated excellent biocompatibility and safety comparable to clinical-grade COL, with no cytotoxic effects observed. Additionally, the dC ECM constructs preserved CKCs-specific characteristics, maintained proper transparency for vision, and offered design flexibility through 3D cell printing. These findings position dC ECM as a promising bioink for treating various corneal diseases, addressing the challenges of donor shortages and material deterioration in artificial corneas (Fig. 7(C)).221
Recently, Vijayaraghavan et al. developed a 3D-printed GelMA-MC hydrogel for CSE, demonstrating stable optical activity, biocompatibility, and printability at room temperature. Rheological analysis confirmed a steady decrease in complex viscosity with increasing shear rate, ensuring shape fidelity, while optimal printing conditions for GelMA 5 wt%/MC 8 wt% and GelMA 7 wt%/MC 8 wt% were determined as 120 kPa and 145 kPa at 2 mm s−1. FE-SEM analysis revealed that GelMA7/MC8 had smaller pores (∼412 nm) than GelMA5/MC8 (∼674 nm), leading to improved mechanical properties, with GelMA7/MC8 achieving a maximum compressive strength of 69 kPa compared to 54 kPa for GelMA5/MC8. The increased GelMA concentration likely contributed to higher cross-linking densities, enhancing stress–strain behavior, which is crucial for implantation and surgical applications. Additionally, light transmission improved to ∼80% by day 14. Biocompatibility assessments with goat stromal keratocytes confirmed excellent adhesion and proliferation, with MTT assays indicating superior viability in GelMA5/MC8, attributed to its larger pore size. Keratan sulfate and vimentin expression confirmed sustained keratocyte phenotype up to day 14. While the study presents promising advancements, limitations include the need for in vivo validation, the exclusive use of goat stromal keratocytes, and the necessity for further exploration of hydrogel-cell interactions, such as gene expression analysis, to enhance translational potential for corneal stromal disorders.225
Building upon previous studies, Puistola et al. reported the development of a novel multi-material 3D bioprinting strategy to fabricate biomimetic corneal stromal structures. Utilizing hASCs and hyaluronic acid-based bioinks of varying stiffness, they bioprinted alternating soft and stiff filaments in perpendicular layers to replicate the native stromal microarchitecture. The soft bioink supported cell proliferation and tissue formation, while the stiff bioink provided mechanical integrity and guided cellular organization. Notably, the storage modulus of cell-laden constructs increased from 0.069 ± 0.005 kPa to 0.141 ± 0.184 kPa after 14 days, demonstrating cell-driven mechanical enhancement. Additionally, transparency remained within an acceptable range (67–84%) despite high cell proliferation. Ex vivo cornea organ culture confirmed good integration of the constructs, with host porcine stromal cells migrating towards the bioprinted composite. These findings highlight the potential of multi-material 3D bioprinting in corneal tissue engineering, though further in vivo studies are required to validate long-term performance.228
![]() | ||
Fig. 8 Schematic illustration of the electrospinning setup highlighting different collector configurations and their corresponding fiber orientations: (a) rotating mandrel collector. (b) parallel plate collector. (c) four-post collector. (d) pin-and-cup collector.211 |
Over polymers have been successfully electrospun into nano or micro-scale fibers, marking significant progress in tissue engineering. However, the widespread application of electrospun scaffolds hinges on exceptional biocompatibility and specific properties tailored for CSE.231–234Table 4 serves as a comprehensive compilation of polymers investigated for this purpose. Among the commonly employed materials are both single polymers and blended combinations. The properties of the polymer solution, including concentration and molecular weight, play pivotal roles in shaping the electrospun fiber morphology. Precise control over electrospinning parameters allows the creation of scaffolds with varied fiber arrangements, mirroring the corneal stromal layer's structure. Material components, fiber morphology, surface chemical modifications, and incorporated information molecules collectively impact scaffold functionality. The strategic manipulation of these factors facilitates the engineering of scaffolds with superior biological and mechanical properties and high transparency, thus propelling advancements in CSE.5
Scaffold type | Cell Source | Application | Ref. |
---|---|---|---|
COL/PCL membranes | HCECs | • Biodegradable COL/PCL membranes could facilitate cell attachment and proliferation | 235 |
• Substrate showed promise as corneal grafts for CTE | |||
GEL and PLLA fibers | CSCs | • Uniaxially aligned fibers can induce mechanical and transparency properties and enhance polarized ingrowth of keratocytes | 126 |
PCL and PCL/dC ECM electrospun scaffold | CSCs | • Promising cell-free implant with aligned electrospun PCL fibers with incorporated ECM that encourage endogenous repopulation by neighboring cells | 49 |
PCL/GEL electrospun scaffold | HCSCs | • PCL-GEL composition can be considered as an option in the selection of CTE application | 236 |
Aligned electrospun PCL/PGS | HCSCs | • Promising as corneal's replacement and promote cell organization and immunocompatibility; consequently, this PGS-PCL composition could be applied to support corneal tissue repair | 106 and 237 |
Aligned and non-aligned electrospun PCL/SF/AV | HCKCs | • Aligned scaffold act as an alternative scaffold for corneal stromal layer regeneration | 234, 238 and 239 |
• Membrane has a sufficient property to plays like native stromal structure | |||
PLDLA | HCSCs | • Biocompatible and facilitates the transformation of corneal fibroblasts back to a keratocyte phenotype | 240 |
• Utilizes orthogonal multilayers with aligned fibers for each layer. | |||
COL | CKCs and Fibroblast Cells | • Facilitates cell attachment, growth, and increased ECM deposition | 241–244 |
• Demonstrates biocompatibility, leading to a decrease in myofibroblast phenotype expression on the aligned scaffold | |||
• Exhibits high transparency, attributed to the presence of aligned fibers | |||
PEUU | HCSCs | • Facilitates the differentiation of stem cells into keratocytes and stimulates the production of a COL matrix | 28 |
PHBV | CKCs | • Enhances both cell attachment and proliferation | 245 |
Natural materials like COL, SF, and GEL, known for biocompatibility, biodegradability, and low immunogenicity, have found extensive use in CSE. Notable examples include COL-based scaffolds that replicate the corneal microenvironment. However, natural materials often exhibit limitations in mechanical strength. Synthetic polymers like PLGA, PCL, and PLA offer excellent mechanical properties, allowing for various cell types of inoculation on electrospun surfaces.245
Blended electrospun scaffolds, crafted from polymer blends, aim to balance mechanical, chemical, and biological properties for optimal CSE. These scaffolds offer a versatile solution by combining the biocompatibility of natural polymers like GEL with the mechanical prowess of synthetic polymers like PHBV or PCL. Adjusting polymer ratios within blends allows fine-tuning mechanical properties and degradation rates to align with the native ECM. COL/HA/PEO and SF/P (LLA-CL) exemplify successful blended scaffolds, exhibiting excellent biocompatibility, mechanical properties, and the ability to stimulate corneal regeneration. These advancements bridge the gap between the biological and mechanical demands of CTE, bringing us closer to clinically viable solutions.246,247
Fabricating electrospun scaffolds tailored for CSE involves meticulous control over structural properties. Key considerations include fiber morphology, diameter, and orientation, each impacting the scaffold's performance.8 Notably, the use of special collectors (Fig. 8) allows for creating either isotropic or anisotropic structures, influencing cell behavior in response to micro- and nano-topography cues. Traditional collectors like cylinders or metal mesh yield random fibers suitable for non-woven scaffolds.49 Scaffolds with aligned fibers offer unique properties that are beneficial for guiding cell growth in desired anisotropic directions. Various collection methods align fibers on the collector surface, including high rotational speed, wire drum collectors, auxiliary electrodes, and parallel double-thin plate collectors. Aligned fibers, mimicking the natural orientation of tissues, promote favorable cell adhesion, migration, and proliferation. In CSE, where alignment is crucial for mechanical and transparency properties, studies show that aligned COL nanofibrous scaffolds lead to down-regulated expression of smooth muscle actin in corneal fibroblasts.241–244 Other investigations with aligned PEU and poly (L, D lactic acid) scaffolds demonstrate their potential for reverting corneal fibroblasts to a keratocyte phenotype. These aligned electrospun scaffolds serve as ideal matrices for guiding corneal cells into organized tissues that closely resemble the native corneal environment.28
Constructing corneal substitutes in vitro requires meticulous attention to both the microstructure and crucial characteristics like transparency and mechanical properties in electrospun scaffolds.238 Optical transparency is a pivotal consideration when developing bioengineered corneal constructs. The cornea's role in clear vision and its contribution of two-thirds of the eye's refractive power underscore the significance of achieving transparency in corneal equivalents. A successful tissue-engineered cornea should efficiently transmit visible light, mirroring the natural behavior of the native cornea.230 Despite the versatility of polymers for electrospinning into nanofibrous scaffolds, not all materials are suitable for CSE due to opacity or low transparency.229 Three primary methods address this challenge. Firstly, post-modification via plasma discharge treatment proves effective. Plasma treatment modifies surface chemistry in an eco-friendly manner and enhances cell adhesion properties.8 For instance, He/O2 plasma treatment of electrospun PCL nanofibrous scaffolds significantly increased light transmission compared to untreated equivalents.122 Secondly, blending different materials, particularly natural and synthetic polymers, proves effective.35 Understanding the importance of keratocytes in maintaining corneal transparency, a third method involves keeping keratocytes in a quiescent state. Recent studies highlight the significance of intracellular protein expression, such as TKT and ALDH1A1, in achieving and preserving corneal transparency. Maintaining the quiescent phenotype is accomplished by providing the right topographical and chemical cues.238 For example, L. Samantha Wilson et al. demonstrated the critical role of topographical and chemical cues in reverting corneal fibroblasts to a keratocyte phenotype within 3D multi-layered constructs. Their study utilized aligned electrospun nanofiber meshes arranged orthogonally to mimic the stromal organization of the native cornea. These nanofibers aligned individual cells and facilitated interlayer migration, while also providing a mechanically robust structure with a higher initial modulus. The combination of serum-free media and insulin supplementation created an optimal chemical environment, reducing contraction and fibroblast activation, as evidenced by changes in elastic modulus, dimensional stability, and gene expression profiles. Constructs cultured under these conditions maintained a quiescent keratocyte phenotype, contrasting with those cultured in serum-containing media, which exhibited increased contraction and a fibroblast-like phenotype due to growth factor stimulation. The synergistic effect of nanofibers, serum-free media, and insulin supplementation offers a promising pathway for fabricating electrospun scaffolds tailored for CSE. This approach enhances mechanical strength, transparency, and cellular organization, addressing the challenges associated with developing in vitro tissue-engineered strategies for corneal repair. Optical coherence tomography and qPCR analyses further validated the constructs' capacity to revert fibroblasts to keratocytes, confirming their phenotypical and genotypical fidelity to native stromal tissue. These advances significantly contribute to the development of transparent, mechanically robust scaffolds that closely mimic the native cornea, opening new avenues for CSE applications (Fig. 9).248
![]() | ||
Fig. 9 (A) Diagram illustrating the process of creating aligned, portable nanofiber meshes and structured multi-nanofiber mesh cell constructs. (a) SEM image showing aligned electrospun nanofibers (scale bar = 15 μm). (b) Cellulose acetate frames are affixed to the aligned nanofibers, and a scalpel is used to cut the frames from the collecting apparatus. (c) Portable fiber meshes are seeded with cells and arranged in an orthogonal pattern using a layer-by-layer technique, with acetate and filter paper as spacers. (d) After cells adhere for 2–3 hours, COL solution is introduced to infiltrate the fibers and secure them within the construct. (e) Excess fibers and acetate frames are removed, resulting in finalized nanofiber-cell constructs. Reproduced with permission.248 Copyright © 2024, Wiley. (B) (a) Aspect ratio for AHDCS cells in hydrogel constructs with and without nanofiber meshes after 7 and 14 days in F, K, and K* media. *p ≤ 0.05, **p ≤ 0.01, ***p ≤ 0.001. (b)–(g) Representative actin-stained cells in COL hydrogel constructs without (b)–(d) and with (e)–(g) nanofiber meshes after 14 days in F, K, and K* media, respectively. In fiber-free constructs, cells in serum-containing F media exhibit a shorter, fusiform shape (b) compared to the elongated shape in cells cultured in serum-free K and K* media (c) and (d). The addition of nanofibers enhanced cellular alignment and promoted a more elongated cell morphology (e)–(g). Yellow arrows indicate fiber direction (scale bar = 50 μm). Reproduced with permission.248 Copyright © 2024, Wiley. |
Beyond the imperative for high transparency and optimal refractive power, electrospun scaffolds must bear the mechanical stresses induced by intraocular pressure and eye movements.3 Achieving mechanical properties akin to natural corneal tissue is paramount. For example, Haleh Bakhshandeh et al. explored the development of a two-part artificial cornea designed to substitute for penetrating HCCK in patients with corneal blindness. The peripheral part, composed of plasma-treated electrospun PCL nanofibers, was engineered to enhance mechanical integrity and facilitate tissue biointegration. The Young's modulus of the PCL skirt was measured at 7.5 MPa, aligning well with the natural human cornea's elasticity range (0.3–7 MPa). This skirt was paired with a PVA hydrogel core as the optical center, exhibiting light transmittance greater than 85% across the 400–800 nm range, ensuring excellent optical properties. Morphological assessments revealed a highly porous PCL scaffold structure conducive to cell integration, while biocompatibility tests demonstrated robust adhesion and proliferation of rabbit limbal stem cells within 10 days of culture. These findings highlight the artificial cornea's potential for clinical application, with suitable mechanical properties, transparency, and epithelialization capabilities crucial for corneal regeneration.249 In another study, Bhattacharjee et al. reported that incorporating SF into PCL nanofibrous scaffolds, combined with plasma treatment, improved transparency, mechanical strength, and hydrophilicity, making them more suitable for corneal stromal reconstruction. The study demonstrated that PCL/SF blends remained stable for up to 6–8 hours without phase separation, allowing for uniform scaffold fabrication. Plasma treatment further enhanced cell attachment and proliferation, with significantly higher keratocyte adhesion observed after 5 hours and increased proliferation over 21 days (p < 0.05). The presence of SF not only improved cytocompatibility but also modified the nanoscale properties of the scaffolds, promoting the adsorption of adhesive molecules and optimizing cellular interactions. In addition, SF accelerated scaffold degradation (∼15 wt% weight loss in 12 hours), facilitating corneal tissue remodeling compared to the slow degradation of pure PCL. The study also confirmed that PCL/SF scaffolds supported higher expression of keratocyte-associated markers and corneal ECM-specific proteins (p < 0.01), indicating a more favorable environment for stromal regeneration. The increased fibroin content contributed to controlled degradation, enhanced oxygen permeability, and optimized repair processes. Overall, the findings highlight the potential of PCL/SF scaffolds as a promising biomaterial for CTE by balancing structural integrity, transparency, and biodegradation.250 Khaow Tonsomboon et al. presented an innovative strategy to address the mechanical limitations of hydrogel scaffolds in CSE by reinforcing ALG hydrogels with electrospun GEL nanofibers. This integration resulted in transparent, mechanically robust hydrogels that mimicked the structural and functional properties of native corneal tissue. The incorporation of non-crosslinked GEL nanofibers significantly enhanced the tensile elastic modulus of the hydrogels from 78 ± 19 kPa to 450 ± 100 kPa, demonstrating a marked improvement in mechanical strength while retaining optical clarity. Additionally, pre-crosslinking the nanofibers increased the modulus to 820 ± 210 kPa but slightly reduced transparency. This approach addresses the critical challenge of developing load-bearing, transparent scaffolds that meet the mechanical demands of CSE, offering a promising alternative to overcome the global shortage of donor corneas. The study underscores the potential of fiber-reinforced hydrogels as durable and functional materials for clinical applications in corneal transplantation (Fig. 10(A)).251 Salehi et al. explored the mechanical properties of electrospun-aligned nanofibers made from varying blends of PGS and PCL for use in CSE. They fabricated fibers using different weight ratios of PGS and PCL, ranging from 1:
1 to 4
:
1, and found that the elastic modulus of the fibers decreased as the PGS/PCL blend ratio increased. This change was attributed to the increasing amount of amorphous PGS in the composition, as confirmed by DSC and XRD measurements, which showed a reduction in overall crystallinity. Interestingly, when the surface modulus was measured through nano-indentation, it was found to be two orders of magnitude higher than the bulk elastic modulus, with a notable increase corresponding to higher PGS content. This phenomenon is believed to be due to the increased PGS content, which forces the PCL into more confined, cross-linked domains near the surface of the fibers, contributing to enhanced surface stiffness. This study highlights the complex interplay between material composition and mechanical properties, providing insights into designing nanofiber scaffolds for CSE.252
![]() | ||
Fig. 10 (A) Fabrication process of a gelatin fiber-alginate hydrogel: (a)–(d) A gelatin mat is electrospun and infiltrated with alginate solution, (e) resulting in a gelatin fiber-alginate film after rinsing in distilled water, and (f) forming a gelatin fiber-alginate hydrogel after crosslinking in CaCl2. Reprinted with permission.251 Copyright © [2024], Elsevier. (B) Schematic procedure for creating chitosan-modified biomimetic nanofiber membranes (CBNMs) designed to replicate the microstructure and architecture of devitalized tissue membranes (e.g., HAM) for corneal repair. (a)–(c) Macroscopic morphology and nanoscale structure of HAM's base layer, with randomly arranged COL fibers. (d) Schematic overview of the CBNM fabrication process. Reproduced with permission.253 Copyright © The Royal Society of Chemistry. (C) SEM images of (a) aligned and (b) random electrospun fibrous scaffolds, with inset images showing lower magnifications. Scale bars indicate 2 μm for higher magnifications and 10 μm for lower magnifications. (c) Tensile properties of randomly oriented versus aligned scaffolds. Reproduced with permission.254 Copyright © 2024, Wiley. |
Silva et al. reported that their C-MSC-laden GelNF-HA scaffold effectively mimicked corneal stromal tissue, supporting cellular adhesion, proliferation, and ECM remodeling. Their ex vivo organ culture studies demonstrated that corneas treated with the C-MSC-laden GelNF-HA scaffold exhibited a 68.3% reduction in stromal haze and a 72.5% decrease in α-SMA expression compared to the untreated keratectomy-only group. The scaffold also improved cell retention compared to the direct topical application of C-MSCs. Furthermore, the GelNF-HA scaffold exhibited a refractive index of 1.322, slightly lower than the human cornea (1.376), and a Young's modulus of 1.66 ± 0.59 MPa, aligning with the physiological range (1.3–5.9 MPa). These findings suggest that the GelNF-HA scaffold enhances corneal wound healing by providing structural support and maintaining an organized stromal architecture, reducing fibrosis and promoting transparency.255
Juan Ye et al. developed a highly effective method combining electrospinning and surface modification techniques to create biomimetic nanofibrous membranes of COL, HA, and PEO. These membranes demonstrated exceptional mechanical properties, with a tensile strength exceeding 20 MPa even in the wet state, indicating their robust structural stability. The electrospun membranes, designed to mimic the natural ECM, were optimized for their mechanical and biological performance. They showed superior results compared to traditional HAM, particularly in their ability to selectively promote the adhesion of epithelial cells and fibroblasts, which are crucial for CSE. In an in vivo rat model of alkali-burned corneal injury, the COL/HA/PEO membranes significantly enhanced re-epithelialization within just one week. This versatile approach holds great potential for biomedical applications, offering a promising alternative to HAM for corneal repair and other wound healing treatments while mitigating concerns related to infectious disease transmission (Fig. 10(B)).253
Jing Yan et al. investigated the impact of fiber alignment on the mechanical properties and cellular behaviors in a GEL/PLLA nanofibrous scaffold. Their study demonstrated that scaffolds with aligned fibers exhibited significantly improved mechanical properties compared to those with randomly oriented fibers. Specifically, aligned scaffolds showed a higher tensile modulus, greater break strength, and lower elongation at break, indicating the importance of fiber orientation in enhancing the mechanical robustness of the scaffold. Moreover, the research highlighted how fiber alignment influences cellular responses, with CKCs favoring alignment in the scaffolds for better adhesion and proliferation. In contrast, CECs showed enhanced interaction with randomly oriented scaffolds. This study underscores the potential of fiber alignment as a crucial factor in scaffold design, suggesting that scaffolds incorporating both aligned and randomly oriented fibers could be optimized to support the growth of different corneal cell types, a key consideration for the successful development of CSE applications (Fig. 10(C)).254
Fabrication method | Material | Cell source | Results | Ref. |
---|---|---|---|---|
Cell sheet engineering | Silk/PolyNIPA GEL/PolyNIPA | GKCs | • Patterned substrates directed corneal cell alignment along the direction of substrate | 256 |
• Gene expression was higher over GEL-PolyNIPA matrix | ||||
Fibroblast cells | HCCs | • Endothelial and epithelial cells attached to the reconstructed stroma | 257 | |
• Epithelial cells formed epithelium layer, and endothelial cells formed endothelium layer | ||||
Self-assembly | Cyclodextrins/COL | HCCs | • The implants demonstrated tissue integration and supported re-epithelialization | 258 |
• Mimetic substitutes had progressive functional and structural properties | ||||
Microfabrication | Silk protein films | CFCs | • Improvement in alignment, proliferation and ECM expression profile on these films | 114 |
• Silk protein to produce a substrate with a well-defined topography developed the epithelial cells behavior toward directional epithelialization | ||||
COL | CKCs | • Composite COL-based hydrogels with a transparent core and degradable peripheral skirt were developed to restore corneal transparency and act as reservoirs for cells and drugs. | 130 | |
• The hydrogels supported cell populations in vitro and maintained corneal integrity in vivo, with controlled skirt degradation promoting host cell integration and nerve regeneration. | ||||
COL | CKCs | • Orthogonally aligned multilayer COL fibril scaffolds, prepared using a high magnetic field, successfully recreated the organized architecture of the corneal stroma and supported the in vitro reconstruction of human hemi-corneas with differentiated epithelium and aligned keratocytes. | 131 | |
• In vivo experiments in a rabbit model demonstrated the scaffolds' potential for anterior stromal repair, enabling re-epithelialization, restored transparency, and maintained ultrastructural organization. | ||||
COL | CKCs | • COL film exhibits excellent mechanical properties, optical transparency, nutrient permeability, and stable morphology, making it suitable for diverse dimensions and easy fabrication. | 132 | |
• In vivo corneal lamellar HCCK in rabbits demonstrated that the COL. film supports rapid re-epithelialization, restored transparency, and shows no rejection or neovascularization within two months, highlighting its potential for cost-effective corneal repair. |
In the context of corneal structure, cell sheet engineering in a bottom-up strategy proves promising for mimicking the ECM of the cornea and specific corneal functions. Ultrathin membranes, often porous layers of polymeric resources, serve as the physical carriers for cell sheets. This method is mainly helpful for selectively replacing damaged tissue in endothelium and epithelium layers due to their dimensions, especially for the endothelium layer (Table 6). To replace the corneal endothelium, the Descemet membrane of the endothelial cell carrier must also be replaced or repaired. Free-standing functionality on corneal endothelial cell sheets is achieved under specific conditions, often involving thermo-sensitive P(NIPAM-co-GMA) polymer membranes.74,259,260 Several studies have explored the use of cell sheet engineering for keratocyte replacement and CSE. For example, Nara et al. utilized DWA to create microperiodic parallel patterns using silk–PolyNIPA and gelatin–PolyNIPA composites. These semi-interpenetrating networks demonstrated temperature-responsive properties suitable for cell sheet applications. At 37 °C, both hybrids displayed hydrophobic surfaces, with silk–PolyNIPA showing higher surface roughness compared to gelatin–PolyNIPA. When the temperature was lowered to 20 °C, surface roughness and contact angle values decreased, promoting better cell interactions. The parallel-patterned substrates aligned corneal cells along the patterns, enhancing COL-I and aggrecan gene expression, particularly in the gelatin–PolyNIPA matrix. Furthermore, gelatin–PolyNIPA promoted higher metabolic activity and vinculin expression, indicating improved biocompatibility. Although cell sheets detached quickly from planar films (within 10–30 minutes), intact recovery from patterned substrates remains challenging, highlighting the need for further optimization to develop adequate corneal stromal substitutes (Fig. 11(A)).256 Proulx et al. developed a tissue-engineered cornea by culturing fibroblasts in serum and ascorbic acid-enriched medium, leading to the formation of ECM sheets that were layered to reconstruct the corneal stroma. Endothelial and epithelial cells were then seeded on opposite sides of the stroma. After 10 days of air–liquid interface culture, the epithelial cells stratified into 4–5 layers, forming distinct basal and wing cells expressing Na +/K + -ATPase α1 and keratins. Endothelial cells also formed a tightly packed monolayer, expressing similar functional proteins. This fully biological corneal model closely resembled native corneal tissue, making it a promising tool for studying corneal pathologies and potentially for clinical applications (Fig. 11(B)).257
Therapy technique | Pros | Cons | Ref. |
---|---|---|---|
Natural scaffolds | • Bioactivity | • Limited mechanical strength | 261 |
• Compatibility with living tissues | • Suboptimal optical transparency | ||
• Degradation through natural processes | • Invasive nature | ||
• Absence of cytotoxic effects | • Variability across diverse samples | ||
• Induction of biological recognition responses | • Potential for disease transmission | ||
• Promotion of cell adhesion and functionality | • Susceptibility to rejection | ||
Synthetic scaffolds | • Abundant supply | • Lack of cellular recognition signals | 262 |
• Outstanding mechanical characteristics | • Minimal impact on proliferation, adhesion, and differentiation | ||
• Adjustable properties | • Potential for inflammatory reactions | ||
• Cost-effective | • Bio-toxicity concerns | ||
• Manageable degradability | • Limited biocompatibility | ||
• Invasive nature | |||
Cell therapy | • Straightforward | • Uneven cell distribution | 263–266 |
• Minimally invasive | • Limited persistence | ||
• Inefficiency in repair | |||
• Requirement for prolonged patient stability in an eye-down position | |||
• Reduced adhesion at the injection site | |||
• Inappropriate cell migration | |||
• Restricted proliferation of primary HCECs | |||
Xenograft transplantation | • Endless availability | • Potential transmission of zoonotic diseases | 267 |
• Comparable anatomical composition | • Risk of hyperacute rejection | ||
• Invasive nature | |||
Autograft transplantation | • Presently considered the gold standard in treatment | • Limited donor availability | 268 |
• Potential for infection transmission | |||
• Invasive nature | |||
• Susceptibility to rejection | |||
Cell sheet | • Preservation of ECM, microenvironment, and cell recognition signals | • Invasive nature | 264 |
• Impact on cellular proliferation, adhesion, and differentiation | • Time-intensive synthesis of cell sheets | ||
• Elimination of the need for sutures | • Suboptimal mechanical strength | ||
• Optical clarity |
![]() | ||
Fig. 11 (A) Demonstration of DWA for fabricating microperiodic parallel patterns using silk–PolyNIPA and gelatin–PolyNIPA composites. Reproduced with permission.256 Copyright the Royal Society of Chemistry. (B) Macroscopic appearance of the bioengineered cornea: a. Arial 12 pt font dots visible through the bioengineered cornea. (b). The bioengineered cornea exhibits a slight haze. Tissue-engineered human cornea: c. Masson's Trichrome staining highlights a well-differentiated epithelial layer on the surface, a monolayer of endothelial cells at the base, and a self-assembled stromal matrix in between. (d). Transmission electron microscopy (TEM) of the epithelial basal membrane reveals numerous hemidesmosomes (arrows). (e). TEM of the corneal endothelium shows a monolayer of flattened cells. Scale bars: 50 μm (A), 1 μm (d), 0.5 μm (e) Reproduced with permission.257 Copyright © 2010 Molecular Vision. |
Self-assembly is a fundamental technique in CSE, drawing inspiration from natural tissue formation processes. Mimicking the anabolism observed in in vitro organogenesis, self-assembly relies on structural motive forces, short-range interactions, and equilibrium conditions. This method facilitates the modular assembly of engineered tissues, replicating the interactions between the ECM and cells crucial for tissue formation. Although the precise assembly protocol for corneal fibers remains incompletely understood, it is believed to involve layers of COL aligned with the corneal tissue's anatomy and physiology, which are vital for optical transparency. In CSE, self-assembly offers adaptability in scaffold production by allowing adjustments in solvent, concentration, pH, temperature, and ionic strength to modify structural integrity. This reversible process, driven by non-covalent bonding mechanisms, occurs within controlled environments to yield diverse structures, reminiscent of natural processes like lipid organization, bone formation, and shell construction. Scaffold formation via self-assembly involves the aggregation of constituent fragments, initially maintaining separation between solvent and solute until reaching a critical solute concentration, leading to the formation of various micellar shapes.
Synthetic corneas have been successfully created using self-assembly, integrating three distinct cell types. Fibroblasts contribute to the fabrication of a stromal matrix resembling the ECM environment. At the same time, endothelial and epithelial cells are seeded on opposing sides of the self-assembled stroma to mimic the native corneal structure. Over 10 days, stratified epithelial and endothelial cells form a cohesive monolayer resembling native corneal tissue features. Beyond CTE, self-assembled scaffolds have widespread applications in drug delivery systems. They offer advantages such as mitigating drug side effects and enabling precise control over timing, solubility, and dosage at specific sites through modulation of the carrier's degradation rate. However, ongoing efforts are focused on enhancing self-assembled structures to minimize cytotoxicity and optimize functionality under specific environmental conditions.
Majumdar et al. explored the use of CDs to control COL fibrillogenesis and structure, creating transparent and mechanically robust COL vitrigel implants that replicate native corneal architecture. CDs interacted with COL during gelation and vitrification, guiding fibril formation into aligned lamellae stacks. Custom molds shaped these materials into implants that matched corneal curvature. In a rabbit keratectomy model, the implants showed good biocompatibility, and tissue integration and supported re-epithelialization, making them promising candidates for corneal reconstruction (Fig. 12).258
![]() | ||
Fig. 12 (A) Role of cyclodextrins in modulating fibril formation and alignment in vitrified COL corneal substitutes through COL interactions: (a) illustration depicting how cyclodextrins interact with COL, influencing fibril arrangement and spacing. (b) Visual comparison of cyclodextrin-COL composites with vitrified COL layers of varying thicknesses. (c) Image showing 500 μm-thick implants fabricated using βCD-Suc and COL, highlighting their optical clarity. (d) Lamellar structures observed in the implants, were demonstrated by manually peeling individual layers. (e) Set up for suture pullout testing to evaluate the suture retention capacity of biomaterials. (f) Peak load data measuring suture retention strength of vitrified gels (n = 6). Results are presented as mean ± SD; statistical analysis by unpaired Student's t-test, ***P < 0.0001. (B) Functional performance and surgical outcomes of biomimetic COL-CD implants: a. Diagram illustrating the implantation of CD/COL materials in ex vivo rabbit corneas. (b) Time-lapse of corneal epithelial wound healing over 72 hours, visualized using fluorescein staining (yellow-green areas indicate non-epithelialized regions under blue light). (c) Epithelial proliferation across implants at 72 hours, assessed via H&E staining. (d) Immunohistochemical staining of ex vivo cornea sections using DAPI (blue) and K-14 (red) to assess epithelial cell maturation and functionality. (e) Custom molds for fabricating implants with the desired curvature. (f) Lenticular-shaped βCD/Col implants mimicked the curvature of healthy corneas. (g) Schematic of in vivo implantation of corneal implants, secured with interrupted sutures. (h) Gross image of ocular surgery in a rabbit model on the day of implantation. (i) Gross appearance of the ocular surface 31 days post-surgery. (j) Fluorescein staining (under blue light) on day 31 to evaluate reepithelialization of the βCD/Col implant. (k) Histological analysis using Masson's Trichrome staining to examine the implant at 14 days. (l) Immunostaining for laminin (red) and DAPI (blue) to assess epithelial cell activity in peripheral and central wound regions, showing laminin expression post-epithelial maturation. Reproduced with permission.258 Copyright © 2024, Wiley. |
Inspired by geometrical patterns from ECM surface topography, microfabrication techniques aim to evaluate physiological functions and cell behavior on desired structures. The goal is to achieve the natural corneal architecture by utilizing information from surface effects on cells. This focus on scaffold construction aims to create the most effective topography for cell-scaffold interactions, leveraging insights gained from the influence of surface topography on cell behaviors such as adhesion, migration, and differentiation. In the context of CSE, mimicking topographies of various corneal layers is crucial for the cells in different layers. According to corneal anatomy, epithelial cells exhibit enhanced interaction on 2D surfaces, while keratocytes show better alignment in 3D scaffolds.217 Lawrence et al. designed silk protein films to replicate the architecture of corneal stromal tissue. The films, measuring 2 μm in thickness, were surface patterned to guide cell alignment and featured pores (0.5–5.0 μm) to enhance nutrient diffusion and cell interactions. In both 2D and 3D cultures, the films supported human and rabbit corneal fibroblast proliferation, alignment, and ECM expression. Their favorable mechanical properties, optical clarity, and ability to support corneal cell functions make these silk films a promising biomaterial for CSE (Fig. 13(A and B)). Xiong et al. (265) developed micro-grooved COL films to study the impact of surface topography on corneal cell behavior and antifibrosis. The micro-grooved patterns significantly influenced cell orientation, proliferation, and migration, promoting epithelial cell movement and wound healing while reducing keratocyte fibrosis. The films demonstrated similar swelling, optical clarity, and biodegradability to native cornea, making them a promising option for CTE applications (Fig. 13(C and D)).114
![]() | ||
Fig. 13 (A) (a) Grooved patterned silk films and (c). flat silk films with corresponding SEM images displaying their surface topography (b) and (d). (e). Silk films fabricated with pore sizes ranging from 500 nm to 5 mm, as shown from surface (f) and cross-sectional (g) perspectives. The films were 2 mm thick with pores extending throughout the cross-sectional area (b). Reprinted with permission.114 Copyright © [2024], Elsevier. (B) (a) Phase contrast image showing a GFP-rCF cellular process aligned along the groove axis of the patterned silk film (arrow) after 4 days of culture. (b) Fluorescent image confirming the cellular structure of the same process. (g) and (h) Actin filament staining on patterned and flat silk film surfaces, respectively, after 10 days of culture. False-color images depict actin filaments (red) and GFP fluorescence (green). (i) Statistical analysis revealed a significant difference in actin alignment between patterned and flat silk film surfaces (p < 0.05; error bars = SD; n = 3). (b), (c), (e), and (f) Fluorescent images illustrating actin filament organization at varying time points and cell densities for GFP-rCFs cultured on patterned (b) and (c) and flat (e) and (f) silk films. At lower cell densities, cells exhibited greater spreading, while higher densities led to more compact and aligned organization along the groove axis (c), compared to flat surfaces (f). (i) Increased cell density on patterned silk films significantly enhanced actin filament alignment (p < 0.05; error bars = SD; n = 3). Reprinted with permission.114 Copyright © [2024], Elsevier. (C) Microgrooved COL films were developed to evaluate their potential application in CTE. Reproduced with permission.265 Copyright the Royal Society of Chemistry. (D) Orientation of CECs and keratocytes cultured on COL films: a. Fluorescence image showing CECs and keratocytes after 48 hours of culture on COL films, stained for F-actin and nuclei (scale bar = 250 μm). (b) Alignment index indicating the percentage of aligned CECs among total cells. (c) Alignment index showing the percentage of aligned keratocytes among total cells (*p < 0.05, **p < 0.01). Reproduced with permission.265 Copyright the Royal Society of Chemistry. (E) Schematic illustration of the fabrication process for composite core-and-skirt hydrogels. (a) The setup for creating a mechanically compressed COL skirt involves COL without a cross-linker, placed between a nylon mesh template and compressed under a weight to facilitate fluid removal (indicated by white arrows). (b) A photograph of the COL skirt, with transparency adjustable based on COL concentration and thickness. (c) Central holes are created in the skirt to enable unobstructed vision. (d) The skirt (with holes) is embedded into a transparent COL core solution containing a cross-linker and allowed to cure. Scanning electron micrographs display the nanostructures of COL fibrils in the core, the core-and-skirt composite, and the skirt-only material. Reprinted with permission.130 Copyright © [2024], Elsevier. |
Rafat et al. addressed the critical challenges of corneal blindness and graft survival by developing innovative composite COL-based hydrogels designed to restore corneal transparency and serve as reservoirs for cells and drugs. These hydrogels feature a transparent core and a degradable peripheral skirt (Fig. 13(E)), enabling adjustable transparency and degradability. The skirt exhibited faster degradation in vitro and integrated seamlessly with the core to distribute mechanical loads. In vitro studies confirmed the hydrogels' support for HCCs populations, while in vivo transplantation in rabbit corneas demonstrated maintained corneal shape and integrity over three months. Controlled skirt degradation facilitated host cell migration, stromal integration, and nerve regeneration, highlighting the potential of these composites for clinical applications in corneal repair.130
In another study, Shen et al. reported that a dual-crosslinked hybrid hydrogel composed of HAMA and pDCSM-G exhibited promising mechanical and biological properties for corneal defect repair. The hydrogel displayed over 80% light transmittance across the visible spectrum (380–800 nm), maintaining optical clarity. Mechanical assessments revealed that the hybrid hydrogel had a tensile strength of ∼43 kPa, demonstrating superior deformation resistance due to its high crosslink density. Biocompatibility evaluations showed that, unlike HAMA alone, it effectively supported corneal stromal cell adhesion, viability, and proliferation. Gene expression analysis indicated upregulation of keratocan and lumican, crucial for maintaining keratocyte phenotype and corneal transparency, while myofibroblast markers were downregulated, minimizing the risk of corneal scarring. In vivo results further confirmed rapid and complete re-epithelialization on hydrogel-treated corneas, with intact endothelial layers observed even eight weeks post-operation. Compared to HAMA and pDCSM-G alone, the hybrid hydrogel reduced corneal fibrosis and promoted long-term stromal regeneration, making it a potential candidate for CTE applications.266
Huang et al. reported that they fabricated GelMA hydrogels using gelatin derived from COL to mimic the ECM, providing a supportive scaffold for hAESC-CKCs. Their hydrogel demonstrated tunable mechanical properties, reinforcing the structural integrity of the corneal stroma in vitro. When applied in situ, it functioned as a bioactive platform promoting epithelial regeneration without the need for fibrin glue or sutures, significantly reducing the risk of postoperative complications. AS-OCT measurements confirmed seamless sealing of corneal incisions, ensuring proper integration between the hydrogel and host tissue for improved wound repair. Furthermore, their results aligned with previous studies showing that cell-laden COL matrices support epithelial stratification and corneal regeneration. In addition, their study demonstrated that the hydrogel effectively modulated the inflammatory response by inhibiting early immune activation, leading to the repression of TGF-β-mediated fibrosis and VEGF-induced neovascularization. The hAESC-CKCs exhibited potent immunomodulatory effects by secreting anti-inflammatory factors and inhibiting M1 macrophage polarization, fostering a favorable tissue microenvironment. Notably, the hydrogel exhibited superior anti-scarring effects, highlighting the critical role of keratocyte-like hAESC-CKCs in scarless corneal repair. These findings suggest that combining keratocytes with CSSCs could enhance therapeutic outcomes, though challenges remain in optimizing cell ratios and procedural consistency. Despite promising results, the study was limited by its short observation period, necessitating further long-term evaluations to assess its full therapeutic potential in preventing and treating corneal scarring.267
Clinical trials evaluate the safety and efficacy of CSE. Key parameters include graft integration, optical clarity, and supporting epithelial and endothelial regeneration. A significant challenge in CSE is the development of biomaterials that possess the necessary mechanical properties and biocompatibility to promote tissue regeneration. For example, González-Andrades et al. initiated a clinical trial to evaluate the safety and efficacy of a bioengineered nanostructured fibrin-agarose corneal substitute. This substitute, incorporating allogeneic cells, aims to replicate the optical, mechanical, and biological properties of the native cornea.268 In another study, Xeroudaki et al. explored a collagen-based hydrogel for CSE. This transparent, mechanically stable hydrogel supports cell migration and maintains corneal thickness, offering a promising approach for corneal tissue replacement and regeneration (Fig. 14).271
![]() | ||
Fig. 14 (A) Properties of the BPC: SEM images show the surface morphology of the BPC, revealing long, parallel-aligned collagen fibers at both low (a) and high (b) magnifications. (c) Optical transmittance curve of the BPC hydrogel compared to the human cornea. (d) HCECs cultured in vitro without BPC and (e) cultured on the surface of BPC hydrogel, with live cells highlighted by green fluorescence. (f) In an ex vivo test, the BPC was sutured into a porcine eye using eight interrupted sutures. (g) SEM images show the interface between the sutured BPC hydrogel and the host cornea, highlighting microfractures caused by sutures cutting through both the BPC and corneal tissue (arrows). (B) Surgical Technique for Femtosecond Laser-Assisted ALK: (a) In the autograft procedure, a femtosecond laser is used to cut a stromal tissue disc, which is then reinserted into the bed and secured with eight interrupted sutures. (b) In the BPC hydrogel implantation, the native stromal tissue disc is removed and replaced with the BPC hydrogel, with mattress sutures placed to avoid suturing through the hydrogel. (c) Time-lapse images showing the laser cutting the stromal disc. (d) Time-lapse photos document the stromal disc's extraction, and its replacement with either the autograft tissue or BPC hydrogel, followed by the placement of interrupted or mattress sutures (e). Reproduced with permission.271 Copyright © 2024, Springer Nature. |
Building upon previous research, Rafat et al. introduce BPCDX, a novel cell-free engineered solution for CSE targeting visual impairment from corneal stromal diseases (Fig. 15). This minimally invasive approach, combined with a pilot feasibility study in India and Iran (NCT04653922), evaluated BPCDX implantation in 20 patients with advanced keratoconus. The study aimed to reshape the corneal stroma without tissue removal or sutures. Results demonstrated safety over a 24-month follow-up, with positive outcomes including increased corneal thickness (India: 209 ±
18
μm, Iran: 285
±
99
μm), reduced keratometry (India: 13.9
±
7.9
D, Iran: 11.2
±
8.9
D), and improved visual acuity. Notably, contact lens-corrected acuity averaged 20/26 in India, while spectacle-corrected acuity reached 20/58 in Iran. Remarkably, all initially blind subjects regained vision (20/36 best corrected) and the ability to wear contact lenses. These findings suggest that BPCDX is a potentially safer, facile, and more accessible alternative to traditional corneal transplantation for vision restoration.270
![]() | ||
Fig. 15 (A) Biomaterial properties of BPCDX. (a) Appearance of BPCDX demonstrating its transparency and refractive properties. (b) Light transmission through 550-μm-thick samples of BPCDX, single-crosslinked BPC, and human cornea. The human cornea, containing epithelial cells that absorb UV light, contrasts with the cell-free bioengineered materials. Data represent means and standard deviations from three independent samples. (c) Mechanical properties of BPCDX compared to single-crosslinked BPC and previously published data of bioengineered constructs made from porcine COL, with reference values from the human cornea for comparison. Measurements for BPCDX represent means and standard deviations from 22 independent samples (550-μm-thick ‘dog-bone’ specimens from different production batches). (d) SEM images showing the surface and bulk (cross-section) structures of BPCDX and porcine cornea, illustrating densely packed COL fibrils in BPCDX, with diameters slightly larger than those in the native porcine cornea. Representative images from three samples per cornea type show consistent results. (e) Degradation of BPCDX, single-crosslinked BPC, and human donor cornea in 1 mg mL−1 collagenase, with data representing means and standard deviations from three independent samples for bioengineered materials and two independent samples of human donor cornea. (f) Attachment and growth of HCE-2 human corneal epithelial cells on BPCDX compared to control culture plates after 16 days. Cells attached to BPCDX, with NucBlue staining, indicating live, viable cell morphology. BPCDX demonstrated greater cell density than culture plasticware. Data from three control and six BPCDX samples, with error bars representing mean and standard deviation (P = 0.003, two-sided independent t-test). Scale bar, 100 μm. (B) Clinical data from a subject in India receiving BPCDX. (a) Pre-operative (left) and one-day post-operative slit-lamp photographs show immediate corneal thickness and curvature changes. (b) OCT scans illustrate sustained thickening and regularization of corneal curvature after implantation of 280-μm-thick BPCDX. Anterior and posterior surfaces of BPCDX are indicated by white arrows. (c) Topographic maps (left, keratometric power in diopters), anterior surface elevation maps (center, μm displacement from a best-fit sphere), and OCT pachymetric maps (right, thickness in μm) from the same subject, showing significant flattening of the steepest pre-operative region (black arrow) and a notable increase in corneal thickness post-operatively. The subject's initial BCLVA of 20/600 improved to 20/30 at 24 months. (C) Clinical data from subjects in Iran receiving BPCDX. (a) Keratometric and corneal thickness maps of the same subject show a thin, steep pre-operative cornea, which was thickened and flattened following intrastromal implantation of a 440-μm-thick BPCDX. OCT cross-sectional scans show corneal thickness and shape before and after implantation, with white arrows indicating the anterior and posterior borders of BPCDX. The subject's initial best spectacle-corrected visual acuity (BSCVA) of 20/200 improved to 20/50 at 24 months. (b) and (c) Photographs of eyes from two subjects showing maintenance of corneal transparency four months post-operatively. (d) and (e) In vivo confocal microscopy images from a single subject showing sub-basal nerve presence (d) and endothelial cell mosaic (e) at 6 months postoperatively. The endothelial cell density was 2222 ± 62 cells/mm2. These images, obtained from one subject, may not be representative. Images in (d) and (e) are 400 × 400 μm2. Reproduced with permission.270 Copyright © 2024, Springer Nature. |
CSE | Corneal stroma engineering |
3D | Three-dimensional |
WHO | World health organization |
HCCs | Human corneal cells |
LSCs | Limbus (Limbal) stem cells |
CSCs | Corneal stromal cells |
OH | Hydroxyl |
rCEN | rabbit corneal endothelial |
HCLE | Human corneal limbal epithelial |
HCECs | Human corneal epithelial cells |
PEG/PAA | Poly(ethylene glycol)/Poly(acrylic acid) |
AM | Amniotic membrane |
PGS | Poly glycerol sebacate |
CAD | Computer-aided design |
PEGDA | Poly ethylene glycol diacrylate |
GEL | Gelatin |
ALP | Alkaline phosphatase |
FRESH | Freeform reversible embedding of suspended hydrogels |
HA | Hyaluronic acid |
dC | Decellularized cornea |
PGA | Poly glycolic acid |
PVAc | Poly vinyl acetate |
PLLA | Poly-L-lactic acid |
PHEMA | Poly(2-hydroxyethyl methacrylate) |
TACs | Transient amplifying cells |
LECs | Limbal epithelial cells |
ALG | Alginate |
CECs | Corneal endothelial cells |
hTMSC | Human turbinate-derived mesenchymal stem cells |
hASC-dCSKCs | hASC-derived corneal stromal keratocytes |
ALDH1A1 | Aldehyde dehydrogenase class 1A1 |
PNIPAAm | Poly(N-isopropylacrylamide) |
GlMA | glycidyl methacrylate |
He/O2 | Helium–oxygen |
CFCs | Corneal fibroblast cells |
GKCs | Goat keratocyte cells |
SS | Silk sericin |
CHI | Chitosan |
HPCHI | Hydroxypropyl chitosan |
FIBn | Fibronectin |
PdCT | Porcine decellularized corneal tissue |
PEU | Poly(ester urethane) |
RPE | Retinal pigment epithelium |
hESC-LESCs | Human embryonic stem cell-derived limbal epithelial stem cells |
FEM | Finite element method |
HE | Hematoxylin and eosin |
DWA | Direct-write assembly |
ECM | Extracellular matrix |
BPCDX | Bioengineered porcine construct double-crosslinked |
BPC | Bioengineered porcine construct |
MAPK | Mitogen-activated protein kinase |
TGF- β | Transforming growth factor–β |
miRNAs | microRNAs |
IL-1 | Interleukin-1 |
EBM | Epithelial basement membrane |
BCLVA | Best contact lens-corrected visual acuity |
ALD | Aldehyde |
pDCSM | Porcine decellularized corneal stroma matrix |
EB | Embryoid body |
GelNF-HA | Gelatin nanofibers coated with hyaluronic acid |
hAESCs | Human amniotic epithelial stem cells |
2D | Two-dimensional |
4D | Four-dimensional |
HCCK | Human corneal cell keratoplasty |
CHCCs | Cultured human corneal cells |
CKCs | Corneal keratocyte cells |
UV | Ultraviolet |
BCECs | Bovine corneal epithelial cells |
PDL | Poly-D-lysine |
HECs | Human endothelial cells |
FMN | Flavin-mononucleotide |
HLSCs | Human limbal stem cells |
LECs | Limbal epithelial cells |
PCL | Polycaprolactone |
MRI | Magnetic resonance imaging |
GelMA | Gelatin methacrylate |
LAP | Lithium phenyl (2,4,6-trimethyl benzoyl) phosphinate |
OCT | Optical coherence tomography |
SOALG | Sodium alginate |
CS | Chondroitin sulphate |
PLA | Poly lactic acid |
PLGA | Poly lactide-co-glycolide |
PVA | Poly vinyl alcohol |
pNIPAM | Poly-N-isopropylacrylamide |
PEGDA | Polyethylene (glycol) diacrylate |
COL | Collagen |
AG | Agarose |
LSSCs | Limbal stromal stem cells |
CEpCs | Corneal epithelial cells |
hASCs | human adipose stem cells |
TKT | Transketolase |
hCFs | Human corneal fibroblasts |
PIPAAm | Poly(N-isopropylacrylamide) |
LCST | Lower critical solution temperature |
AuNPs | Gold nanoparticles |
PCECs | Primary corneal epithelial cells |
SF | Silk fibroin |
AVE | Aloe vera |
HECTS | Hydroxyethyl chitosan |
FIB | Fibrin |
HdCT | Human decellularized corneal tissue |
BdCT | Bovine decellularized corneal tissue |
P(NIPAM-co-GMA) | Poly (N-Isopropylacrylamide-co-Glycidylmethacrylate) |
ALK | Anterior lamellar keratoplasty |
PET | Polyethylene terephthalate |
DoD | Drop-on-demand |
hTMSCs | Human turbinate-derived mesenchymal stem cells |
CDs | Cyclodextrins |
LASIK | Laser-assisted in situ keratomileusis |
SEM | Scanning electron microscope |
FAK | Focal adhesion kinase |
CSSCs | Corneal stromal stem cells |
Smad | Signal transduction by mothers against decapentaplegic |
CTE | Corneal tissue engineering |
PDGF- BB | Platelet-derived growth factor BB |
αSMA | α-Smooth muscle actin |
MC | Methylcellulose |
HAMA | Methacrylated hyaluronic acid |
hiPSCs | human-induced pluripotent stem cells |
C-MSCs | Corneal Mesenchymal stromal cells |
AS-OCT | Anterior segment optical coherence tomography |
Footnote |
† These authors have contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |