Two-photon polymerization has been used to create micro- and nanoscale structures. Complex 3D structures can be fabricated by selectively polymerizing photoresists at the focus of a laser with the ability to produce small features. 3D extracellular environments can be synthesized to have tailored properties including stiffness, porosity, roughness, adhesion propensity, etc. Measurement of the mechanical properties of the samples herein are tested by using two TPP-printed movable plates for actuation and force sensing. These moveable plates are mechanically coupled by some sample and supported by vertical, flexible beams. Understanding the stress, strain, and recovery of different samples can used to predict the long-term industrial application of materials.
Studying cellular behavior in 3D by environment requires knowledge of the mechanics of the scaffold in liquid under physiological conditions. Mechanical studies of on TPP-printed structures and narrow wires have been previously studied, but studies like this had limitations. Herein, the method utilizes two TPP printed-moveable plates for actuation and force-sensing which are mechanically couples by microfibers in a liquid environment.
Adhesive organelles between neighboring epithelial cells form an integrated network to withstand external and internal forces. As part of normal physiology, this integrated network is constantly exposed to mechanical stress and strain, which is essential to normal cellular activities, such as proliferation, migration, differentiation, and gene regulation in the process of a diverse portfolio of functions in tissue morphogenesis and wound healing. A host of developmental defects or clinical pathologies in the form of compromised cell-cell associations will arise when cells fail to withstand external mechanical stress due to genetic mutations or pathological perturbations. Indeed, since the mechanical stress is mainly sustained by the intercellular junctions, mutations or disease-induced changes injunction molecules and components in adherens junctions and desmosomes lead to cell layer fracture and tissue fragility, which exacerbate the pathological conditions. This clinical relevance gives rise to the importance of the understanding of biophysical transformations when cells are subjected to load.
Cell-cell adhesions are often subjected to mechanical strains of different rates and magnitudes in normal tissue function. This is particularly true under large strain conditions which may potentially lead to cell-cell adhesion dissociation and ultimately tissue fracture. However, the rate-dependent mechanical behavior of individual cell-cell adhesion complexes has not been fully characterized due to the lack of proper experimental techniques and therefore remains elusive.
Cells often experience strains of tens to a few hundred percent at strain rates of 10 to 100% s−1 in normal physiological conditions. They have many mechanisms to dissipate internal stress produced by external strain to avoid fracture, often via cytoskeleton remodeling and cell-cell adhesion enhancement. The coping mechanisms are different in time scale. Cytoskeleton remodeling can dissipate mechanical stress promptly due to its viscoelastic nature and actomyosin-mediated cell contractility. One study in cell monolayers showed the actomyosin regulated stress relaxation when cells are connected with robust adherens junctions in a biphasic response, with an initial viscoelastic phase within a few seconds and a prolonged response of a few minutes.
Adhesion enhancement at the cell-cell contact is more complex in terms of time scale. Load-induced cell-cell adhesion strengthening has been shown by the increase in the number of adhesion complexes or clustering of adhesion complexes, which occurs on the order of a few minutes to a few hours after cells experience an initial load. Studies also showed that increased load on the cell-cell contact results in a prolonged cell-cell dissociation time, suggesting cadherin bonds may transition to catch bonds in certain loading conditions, which can occur within seconds.
It is generally accepted that stress accumulation in the cytoskeleton network and potentially in the cytoplasm is strain-rate dependent. With the increase in cellular tension, failure to dissipate the stress within the cell layer at a rate faster than the accumulation will inevitably lead to the fracture of the cell layer. To date, there is a lack of understanding about the rate-dependent behavior of the cell-cell adhesion complex, particularly about which of the aforementioned coping mechanisms are at play across the spectrum of strain rates, and about how the stress relaxation by the cytoskeleton coordinates with the enhancement of the cell-cell adhesion under large strains leading to fracture.
To characterize the intricacy of the biophysical and biochemical response of individual cell-cell adhesion complex under large strain, a functional technique needs to fulfill the following requirements. First, it should have a highly sensitive force sensing component that allows easy quantification of pico- or nano-newton forces. Second, it should have the capability to apply mechanical strain or stress in a controlled manner. Third, the testing can be conducted under physiologically relevant conditions, especially allowing the formation of mature cell-cell junctions and cell-ECM adhesions. Several widely used techniques in the quantitative assessment of cell-generated forces include traction force microscopy and elastomer-based micropillar arrays. In addition, micro-scaffolds fabricated by 3D printing have been used to measure cell forces in a 3D microenvironment.
Although these quantification methods provide great insight into the actin-based ECM adhesion networks, they are restricted to static observations and fail to apply desirable mechanical stimuli. Techniques exist to apply mechanical strain to a monolayer of cells, but the stress within an individual cell-cell adhesion cannot be determined. Further, when a load is applied to individual cell-cell junctions, the majority of the studies are carried out on isolated suspended cells where mature intercellular junctions are yet to form, and the focus can only be placed on the separation of the cadherin bonds while the effect of stress relaxation of the cytoskeleton and the cell-ECM interactions are ignored.
The present disclosure provides systems and methods for the design and fabrication of a polymeric microstructure using two-photon polymerization and systems and methods for performing a displacement-controlled tensile test on a sample.
In an embodiment, the disclosure provides a method of measuring a stress-strain curve in a sample. The method includes providing a structure including a first movable island supported by a first beam, a second movable island supported by a second beam, and a gap therebetween connected by the sample wherein the sample has an initial length. The second movable island may be moved with a defined displacement. The displacement of the first movable island may be determined based on moving the second moveable island. The difference between the displacement of the first movable island and the defined displacement of the second movable island is calculated based on moving the second movable island. An applied strain in the sample is determined based on the difference dividing by the initial length of the sample. A force on the sample is calculated based on the displacement of the first moveable island. A stress on the sample is calculated based on the force. The stress-strain curve of the sample is determined by plotting the calculated stress against the applied strain.
In some embodiments, moving the second moveable island includes using atomic force microscopy (AFM). In some embodiments, moving the second moveable island includes using a nanopositioner.
In some embodiments, the structure is developed based on a nanofabricated polymeric structure using two-photon polymerization (TPP).
In some embodiments the first beam has a defined stiffness and the second beam has a defined stiffness.
The present disclosure provides an apparatus for preforming displacement-controlled tensile test of a sample. The apparatus includes a first moveable island supported by a first supporting beam having a first defined stiffness, and a second moveable island supported by a second supporting beam having a second defined stiffness. The first moveable island and the second moveable island defining a junction therebetween having an initial length.
In some embodiments, the apparatus further includes a first sample anchoring structure attached to the first moveable island and a second sample anchoring structure attached to the second moveable island. In some embodiments, the anchor is a confinement for cells. In other embodiments, the anchor is a resin or printing material. In certain embodiments, a sample coupled to the first sample anchoring structure and the second sample anchoring structure is a printed microfiber.
In some embodiments, the first moveable island and the second moveable island are attached to an optically transparent substrate. The optically transparent substrate is optically coupled to an inverted microscope configured to monitor movement of the first moveable island and the second moveable island using digital image correlation (DIC).
In some embodiments, the apparatus is configured to stretch the junction at a controlled strain rate by applying force to the second moveable island using atomic force microscopy (AFM). In some embodiments, the apparatus is configured to stretch the junction at a controlled strain rate by applying force to the second moveable island using a nanopositioner.
In some embodiments, at least a portion of the first movable island or the second moveable island is made using a low autofluorescence resin. In some embodiments, the sample anchoring structure includes a low autofluorescence resin and the method further includes performing fluorescence imaging of the sample attached to the sample anchoring structure.
The present disclosure provides systems and methods for the design and fabrication of a polymeric microstructure using two-photon polymerization and systems and methods for performing a displacement-controlled tensile test of various samples including fibers, epithelial cells, polymers, etc.
The present disclosure provides systems and methods for the design and fabrication of a polymeric microstructure using two-photon polymerization and systems and methods for performing a displacement-controlled tensile test of a pair of adherent epithelial cells. Straining the cytoskeleton-cell adhesion complex system reveals a shear-thinning viscoelastic behavior and a rate-dependent stress accumulation phenomenon that agrees with a linear cytoskeleton growth model. Further, under considerably large strain (>150%), cadherin bond dissociation exhibits rate-dependent strengthening, in which increased strain rate results in elevated stress levels at which cadherin bonds fail. The remarkable tensile strength of a single cell adhesion complex under large strains facilitated by cytoskeleton stress relaxation and cadherin bond strengthening are discussed.
A single cell-cell adhesion interrogation and stimulation platform is developed based on nanofabricated polymeric structures using two-photon polymerization (TPP). This is a platform that allows in situ investigation of stress-strain characteristics of a cell-cell junction through defined strain and strain rate. Two movable islands, supported with beams of known or defined stiffness, are mechanically coupled through the formation of a mature junction between epithelial cells on each island. Integrating the polymeric microstructure with atomic force microscopy (AFM) enables the cell pair to stretch with precisely controlled strain rates, while the deformation of the supporting beams informs the resultant stress accumulated at the cell-cell junction.
The resolution of the sensing beams, capable of resolving the discrete breakage of a few bonds, enables the study of the adhesion failure as a collection of the bond rupture events. With this technique, biophysical phenomena can be revealed at the single cell-cell adhesion interface that was previously not possible to be observed using existing techniques, promoting a paradigm shift in the mechanical characterization of cell-cell adhesions. A single cell pair system behaves like a shear-thinning viscoelastic material under tensile stress, following an active mechanosensing constitutive model. The single cell adhesion complex between an adherent cell pair fails at remarkably large strains in a symmetrical failure pattern with discrete bond ruptures at the edge of the cell-cell contact. Further, the rate-dependent dissociation of cell-cell adhesion complexes is described.
Thus, in one aspect, the disclosure provides a method of measuring a stress-strain curve in a cell-cell adhesion interface, including: providing a structure including a first movable island supported by a first beam, a second movable island supported by a second beam, and a gap therebetween connected by a pair of cells forming a junction, and the pair of cells including a cell-cell adhesion interface having an initial length defined by a distance between nuclei of the pair of cells; moving the second movable island with a defined displacement; determining a displacement of the first movable island based on moving the second movable island; calculating a difference between the displacement of the first movable island and the defined displacement of the second movable island based on moving the second movable island; determining an applied strain in the cell-cell adhesion interface between the pair of cells based on the difference divided by the initial length of the cell-cell adhesion interface; calculating a force between the cell-cell adhesion interface of the pair of cells based on the displacement of the first movable island; calculating a stress in the cell-cell adhesion interface between the pair of cells based on the force; and determining the stress-strain curve of the cell-cell adhesion interface between the pair of cells by plotting the calculated stress against the applied strain.
In some embodiments of the method, moving the second movable island may include moving the second movable island using atomic force microscopy (AFM). In other embodiments of the method, moving the second movable island may include moving the second movable island using a nanopositioner. In various embodiments of the method, the pair of cells may form the junction after culturing of the cells for a period of time. In certain embodiments of the method, calculating a stress in the cell-cell adhesion interface may include: calculating the stress in the cell-cell adhesion interface based on dividing the applied force at the cell-cell adhesion interface by a cross-sectional area of the cell-cell adhesion interface. In some embodiments of the method, the structure may be developed based on a nanofabricated polymeric structure using two-photon polymerization. In particular embodiments of the method, each of the first beam may have a first defined stiffness and the second beam may have a second defined stiffness. In certain embodiments of the method, at least one of the first defined stiffness or the second defined stiffness may be measured by deforming the first beam or the second beam using an AFM probe having a known stiffness. Various embodiments of the method may further include applying a stain to the pair of cells to visualize the cell-cell adhesion between the pair of cells and the focal adhesion points between each of the pair of cells and the structure. In some embodiments of the method, the structure may further include a cell confinement structure, wherein a first portion of the cell confinement structure may be attached to the first movable island and a second portion of the cell confinement structure may be attached to the second movable island, and each of the pair of cells may be disposed within the first portion or the second portion of the cell confinement structure such that the pair of cells forms the junction between them to connect the two movable islands. In particular embodiments of the method, moving the second movable island may include: moving the second movable island in a direction away from the first movable island. In some embodiments of the method, determining a displacement of the first movable island may include: determining a displacement of the first movable island using digital image correction (DIC). In various embodiments of the method, moving the second movable island with a defined displacement may further include: measuring the defined displacement using digital image correction (DIC).
In another aspect, the disclosure provides an apparatus for performing a displacement-controlled tensile test of a pair of cells, including: a first movable island supported by a first supporting beam having a first defined stiffness; and a second movable island supported by a second supporting beam having a second defined stiffness, the first moveable island and the second moveable island defining a junction therebetween having an initial length.
Some embodiments of the apparatus may further include a first cell confinement structure attached to the first moveable island and a second cell confinement structure attached to the second movable island. In various embodiments of the apparatus, a pair of cells may be disposed within the first and second cell confinement structures. In certain embodiments of the apparatus, the first moveable island and the second moveable island may be attached to an optically transparent substrate. In particular embodiments of the apparatus, the optically transparent substrate may be optically coupled to an inverted microscope configured to monitor movement of the first moveable island and the second moveable island using digital image correlation (DIC). In certain embodiments of the apparatus, the apparatus may be configured to stretch the junction at a controlled strain rate by applying force to the second moveable island using atomic force microscopy (AFM). In some embodiments, the apparatus may be configured to stretch the junction at a controlled strain rate by applying force to the second moveable island using a nanopositioner.
The patent or application file contains at least one drawing executed in color. Copies of this patent or patent application publication with color drawing(s) will be provided by the Office upon request and payment of the necessary fee.
The following drawings are provided to help illustrate various features of example embodiments of the disclosure and are not intended to limit the scope of the disclosure or exclude alternative implementations.
The following discussion is presented to enable a person skilled in the art to make and use embodiments of the invention. Various modifications to the illustrated embodiments will be readily apparent to those skilled in the art, and the generic principles herein can be applied to other embodiments and applications without departing from embodiments of the invention. Thus, embodiments of the invention are not intended to be limited to embodiments shown but are to be accorded the widest scope consistent with the principles and features disclosed herein. The following detailed description is to be read with reference to the figures, in which like elements in different figures have like reference numerals. The figures, which are not necessarily to scale, depict selected embodiments and are not intended to limit the scope of embodiments of the invention. Skilled artisans will recognize the examples provided herein have many useful alternatives and fall within the scope of embodiments of the invention.
Before any embodiments of the invention are explained in detail, it is to be understood that the invention is not limited in its application to the details of construction and the arrangement of components set forth in the following description or illustrated in the attached drawings. The invention is capable of other embodiments and of being practiced or of being carried out in various ways. Also, it is to be understood that the phraseology and terminology used herein is for the purpose of description and should not be regarded as limiting. For example, the use of “including,” “comprising,” or “having” and variations thereof herein is meant to encompass the items listed thereafter and equivalents thereof as well as additional items.
As used herein, unless otherwise specified or limited, the terms “mounted,” “connected,” “supported,” and “coupled” and variations thereof are used broadly and encompass both direct and indirect mountings, connections, supports, and couplings. Further, unless otherwise specified or limited, “connected” and “coupled” are not restricted to physical or mechanical connections or couplings.
The term “about,” as used herein, refers to variation in the numerical quantity that may occur, for example, through typical measuring and manufacturing procedures used for articles of footwear or other articles of manufacture that may include embodiments of the disclosure herein; through an inadvertent error in these procedures; through differences in the manufacture, source, or purity of the ingredients used to make the compositions or mixtures or carry out the methods; and the like. Throughout the disclosure, the terms “about” and “approximately” refer to a range of values ±5% of the numeric value that the term precedes.
A microstructure (i.e., a structure with micrometer-scale features) has been designed and fabricated to interrogate the mechanical behavior of the cell-cell adhesion complex under large strain (
The stiffness of the beams was designed to be as close to the stiffness of the cell junction (0.01 N/m-0.5 N/m) as possible to acquire the best balance between force-sensing resolution and applied strain to the cell-cell junction, with the ability to measure a stress range of 0-12 kPa and force range of 0-50 nN at the junction. Compared with horizontal beams, vertical beams offer greater control of their length which allows for easy adaptation to this desired stiffness and offers better structural stability during the TPP fabrication process (as discussed below with reference to
To measure the stiffness of the “A-shaped” beam structure, a tipless cantilever probe with a known and thermally tuned stiffness was used to apply force on an isolated sensing structure with beam thickness of 6 μm (
Cells are deposited into the bowtie structure using an Eppendorf single cell isolation setup which includes a pressure controller, a 3D manipulator, and microcapillary (as discussed below with reference to
To apply strain to the cell-cell junction, the test platform was placed with deposited cells under the AFM integrated with an inverted microscope. An AFM probe with a through-hole drilled at the front end using a focused ion beam (FIB) is positioned above the micropillar on Island 2 and then lowered to capture it within the through-hole. With this, displacement was applied and displacement rates were tried to investigate the mechanical behavior of the cell-cell junction with obtained stress-strain curves (as discussed below with reference to
Four strain rates were examined ranging from 0.5% s−1 to 50% s−1 and different modes of stress relaxation and cell-cell adhesion failure were observed that are strongly strain-rate dependent. A 0.5% s−1 strain rate (100 nm/s in displacement rate) was applied and substantially none of the junctions failed at the end of the 50 μm displacement. The stress-strain curve exhibits a typical viscoelastic behavior wherein the stress increases nonlinearly with a decreasing rate as the strain increases. A typical set of time series images shows that there is no sign of rupture in the cell-cell junction, which was elongated to 221.8±8.0% strain and tolerated maximum stress of 3.8±1.6 kPa (
Tensile tests demonstrate that the cell pair can withstand a remarkably large strain level before it fails through cell adhesion rupture. At low rates, the cell-cell junction remains largely intact even when the strain is higher than 200%. Comparing with the 50% s−1 strain rate, the lower maximum stress under the strain rate of 0.5% s−1, where cell-cell adhesion complexes remain largely intact, indicates the existence of another effective stress dissipation scheme inside cells. Considering the dynamic nature of cytoskeletons among all the intracellular structures, the mechanical stress can be dissipated via remodeling and reorganization of their cytoskeletons. However, under high strain rates, the cell pair dissipates stress primarily through the dissociation of cell-cell adhesion complexes and complete breakage occurs at a strain level of ˜200%. In addition, all failures occur at the cell-cell contact symmetrically through the rupture of the cell-cell adhesion complex. The image series of the tensile test (
The stress-strain relationship from the four types of tensile tests of varied strain rates can be well fitted with an empirical exponential growth function plus a linear function:
σ=−Ae−BE+Cε,
supporting an overall viscoelastic behavior. To delineate the viscoelastic behavior of the cell pair before cell-cell adhesion rupture under the mechanical stretch of different strain rates, a phenomenological constitutive model was developed that effectively incorporates a mechanosensing component to account for the stress dissipation mediated by cytoskeleton remodeling. Briefly, when a pair of cells are stretched, the cell membrane deforms along with their intracellular components. The viscoelastic response of the cell can be modeled using a modified standard linear solid (MSLS) model as shown in
σS2=E2(εS2−ε0) (1)
where εS2 and ε0 are the total strain of the second spring and the strain resulting from the continuous growth of the cytoskeleton, respectively. The cytoskeleton growth rate is related to the strain rate of the second spring through a model parameter α:
{dot over (ε)}0=α{dot over (ε)}s2 (2)
where 0≤α≤1. When α=0, {dot over (ε)}0=0, suggesting that the cytoskeleton does not grow at all, which corresponds to the condition of a very high strain rate stretch. When α=1, Eqn. (2) reduces to {dot over (ε)}0={dot over (ε)}S2, indicating that the growth of the cytoskeleton is able to completely release the passive stress, which could occur under an extremely low strain rate stretching. Therefore, the value of a is an effective parameter to indicate the growth level of the cytoskeleton during the stretching test and thus the stress dissipation efficiency. The model predicts the following time-dependent relationship between stress (σtot) and strain (εtot):
Under a constant strain rate condition, Eqn. (3) yields:
As shown in
The cell pair was treated with cellular contractility modulators, RhoA Activator I. CN01, and myosin II inhibitor: blebbistatin (Bleb), to examine the impact of actomyosin activity on the mechanical behavior of the cell pair under mechanical stress. Stress-strain curves collected at 0.5% s−1 strain rate show a clear contrast between samples treated with CN01, Bleb, and DMSO control. Specifically, CN01 raises the overall stress level compared with controls at the same strain, while Bleb reduces the stress accumulation (
The necking process can be attributed to the rupture of cell-cell adhesion bonds, which is most apparent under the intermediate strain rate. A few cadherin bonds are ruptured in discrete steps at the edge of the cell-cell junction, which corresponds to a small drop in the measured forces in the force-displacement curve (
The bond dissociation events also exhibit strong strain-rate dependency. First, at a very low strain rate (0.5% s−1), the absence of bond rupture may be attributed to cadherin strengthening. It has been observed that cadherin bond clustering in epithelial cells under tensile load occurs in a time scale of minutes, right in line with the time span of a low-strain-rate tensile test (about 10 minutes). Second, the stress level at which cadherin bonds show initial signs of dissociation, or critical stress, increases significantly with increasing strain rate. As shown in
The mechanical stretch at different strain rates reveals three different rate-dependent modes of stress dissipation and failure phenomenon at the cell-cell adhesion complex. First, the viscoelastic behavior of the cell pair at different strain rates depends on a robust intercellular adhesion. At low strain rate levels (such as {dot over (ε)}=0.5% s−1), cell-cell adhesion through cadherin bonds remains intact, allowing continuous remodeling of the cytoskeleton through the alignment of the cytoskeleton to the tensile load direction. More importantly, it leads to the growth of actin filaments (α is high or close to 1), and thus the continuous stress relaxation in the network of the cytoskeleton and the cell-cell adhesion complex (as illustrated in
The platform developed has distinct advantages over AFM-based single-cell force spectroscopy (SCFS) and dual micropipette aspiration (DPA) techniques, which have been previously used to study adhesion mechanics in isolated cell pairs. A major limitation of SCFS is an inability to interrogate mature cell-cell junctions because the system is limited by the adhesive strength between the cell and AFM tip, which is lower than the strength of a mature cell-cell junction. In addition, in SCFS, it is impossible to image the cell-cell junction as the junction moves vertically as it is stretched, leaving the focus plane. A major drawback of DPA is a lack of a mature cell-ECM junction. As the cells are held to the micropipette tip through negative pressure, they do not form a junction, and the sometimes extreme deformation of the cell at the micropipette tip may induce internal biochemical changes which may impact the physiology of the cell-cell junction. In addition, a constant strain rate cannot be achieved because the strain is applied in incremental steps. A common drawback between each of these methods is throughput for interrogating mature cell-cell junctions, as cells would need to be held in place by these devices for a long period of time before a single test could be performed. The design of the device according to an embodiment combines the advantages of each system while eliminating or mitigating these drawbacks. The arrangement of the cells allows for imaging of the cell-cell junction, cells can form strong and mature cell-ECM junctions with the device and cell-cell junctions with each other, and continuous strain can be applied. In addition, throughput for mature cell-cell junction interrogation is increased due to parallel sample preparation and testing, as the equipment for manipulating or stretching cells do not need to be used to hold cells in place during junction maturation. The presence of the mature cell-ECM junction allows for application of large strains as in DPA, whereas the force sensitivity of the beams achieves stress and strain resolution comparable to SCFS. Finally, another technique that has been used to interrogate adhesion molecules, such as cadherins, is single-molecule force spectroscopy. While this technique can accurately measure forces within bonds at a single-molecule level, the internal response from cells to stretching, which is crucial in understanding cell-cell adhesion mechanics, is lost in this experimental setup and fully captured in the design.
Integrated within a microscopy imaging system, the mechanical characterization studies can be combined with fluorescent imaging of cytoskeleton deformation and localization of cadherins and linker molecules when the single cell adhesion complex is subject to a tensile load of varying amplitudes and strain rates. Further, the tensile strength within the cytoskeleton-cell adhesion-cytoskeleton system can correlate with tensional fluorescence resonance energy transfer (FRET) sensors within the cadherin or linker molecules, and this correlation may ultimately delineate the force contribution of each component in maintaining the mechanical integrity of the complex and reveal mechanisms of mechanotransduction in a concerted effort with other cellular elements, such as the cytoskeleton and the cell-ECM adhesion. Despite the promising propositions, a limitation still exists in performing real-time fluorescence imaging with cells on the microstructures due to the strong auto-fluorescence of the polymer materials used for the TPP fabrication. Research efforts are ongoing to address this critical issue.
In summary, a polymeric microstructure was fabricated using TPP for displacement application and force sensing to examine the rate-dependent mechanical behavior of a single cell-cell adhesion complex. This platform can target the cell-cell contact of a single cell pair and strain their mutual junction, enabling the quantitative assessment of its mechanics at controlled strain rates and the examination of its failure at large strains. The fine resolution of the force sensing beams also enables capturing the dissociation of cell-cell adhesion bonds to reveal its failure mechanism. Displacement-controlled tensile tests reveal that the single cell-cell adhesion complex composed of the cytoskeleton structures from the cell pair and the cadherin adhesion molecules fails at a remarkably large strain level, and the failure process exhibits strain rate-dependent phenomena. This is predominantly facilitated by the relaxation of the actin networks and rate-dependent strengthening of cadherin molecules.
Embodiments of the invention described herein can be incorporated into a variety of applications and disciplines. For example, embodiments of the invention can be incorporated into medical devices to facilitate the study of drug penetration through barriers, diagnostics to study skin and heart diseases and cancer metastasis, and in biomedical engineering applications for predicting deformation and failure in artificial tissues. Accordingly, non-limiting example materials, methods, designs, fabrication, and calculations are discussed below.
A431 E-cadherin GFP-tagged cells were cultured in a growth medium composed of Dulbecco's modified Eagle's medium (DMEM) and supplemented with 10% fetal bovine serum (Chemie Brunschwig AG) and 1% penicillin-streptomycin (Invitrogen). The medium included C02-independent growth medium (Gipco) supplemented with 2 mM L-glutamine (Gipco), 10% fetal bovine serum, and 1% penicillin-streptomycin. All solutions were filtered through 0.22 μm pore-size filters before use. Shortly before each experiment, PBS was replaced with 2 ml of the experimental medium. All experiments were performed in a temperature-controlled enclosed chamber at 37° C. Transfection of E-cadherin siRNA (Santa Cruz Biotechnology; SC35242) and control siRNA (Santa Cruz Biotechnology; SC37007) were performed using Lipofectamine RNAiMAX Transfection Reagent (Invitrogen), according to the manufacturer's protocol. The expression of GFP was analyzed by fluorescence microscopy after 48 hours.
Full-length human E-cadherin fused at its C-terminus to GFP was constructed by first inserting an E-cadherin cDNA into pEGFP-N2 (Clontech, Mountain View, CA) and then inserting the tagged construct into a derivative of the LZRS retroviral expression vector. The final cDNA construct was fully sequenced to ensure no errors were introduced during subcloning.
3D models of the micromechanical structures for biological cell mechanical interrogation were compiled in COMSOL using the built-in CAD module. The compiled models were evaluated using the Solid Mechanics module (linear elastic materials approximation). Finite element analysis (FEA) in COMSOL allowed estimation of the spring constant of the flexible beams supporting the microscale plates for cell attachment. Various preliminary designs, including planar structures, were fabricated using TPP stereolithography and tested for stability during fabrication and susceptibility to damage by capillary forces after fabrication. The rationale behind this design is as follows. First, compared to doubly clamped (bridge) structures, singly clamped (cantilever) beams provide a more linear elastic response with significantly lower sensitivity to intrinsic stresses. Second, the parallelogram arrangement of the twin-beam leaf springs improves the leveling of the cell-bearing platforms and the overall mechanical stability of the devices. Furthermore, vertical beams separated by larger distances from the substrate are preferable over horizontal beams closer to the substrate due to the better ability of the former to withstand capillary forces after fabrication. Finally, the thinnest beams that could be reliably fabricated with high accuracy and yield were approximately 2.5 mm thick. This minimum thickness, combined with the targeted stiffness, dictated the width and the length of the beams in the implemented structures.
To fabricate the structures shown in
Glass coverslips with diameters ranging from 11 to 25 mm and thicknesses of approximately 160 μm were used as substrates in the present study. Prior to 3D printing, the glass substrates were coated with indium tin oxide (ITO) to achieve optical reflectivity of the IP-S/substrate interface sufficient for autofocusing. The ITO layer had a thickness of approximately 50 nm and was deposited using direct current sputtering of an ITO target in an Ar plasma. It was found that mechanical 3D structures printed directly on ITO-coated glass had insufficient adhesion and would detach from the substrate after prolonged soaking or incubation in aqueous solutions. To address this commonly encountered issue of insufficient adhesion between smooth substrates and 3D structures fabricated using TPP, an in-house developed protocol was used in which an additional layer of porous silicon oxide (PSO) was deposited on top of ITO-coated coverslips. PSO with a thickness of approximately 2 μm and a high density of nanopores was found to act as an excellent anchoring layer, eliminating detachment of the 3D printed structures from the substrate during soaking and subsequent experiments in aqueous solutions. For all experiments, arrays of structures (varying from 5×4 up to 6×6) were fabricated on each coverslip, allowing for increased throughput in testing.
The structures were placed inside of a glass-bottom petri dish, washed with 70% ethanol, and immediately soaked with PBS for 10 minutes until all the ethanol dissolved. The substrate was then submerged in 0.3% volume ratio Sudan Black B (Sigma-Aldrich) in 70% ethanol for one hour to eliminate the autofluorescence of the polymer. To dissolve excessive Sudan Black, the substrate was submerged in 70% ethanol for 1 hour and then soaked with PBS for 10 minutes. The substrate was then coated with fibronectin to enhance the adhesion and growth of the cells on the structures. Fibronectin solution with a concentration of 50 μg/ml in PBS was placed on the substrate and left in the incubator for 2 hours. Finally, the fibronectin solution was removed, and the substrate was washed with PBS two times.
The structures were placed inside of a glass-bottom petri dish, washed with 70% ethanol, and immediately soaked with PBS for 10 minutes until all the ethanol dissolved. The substrate was then coated with fibronectin (50 μg/ml in PBS) to enhance the adhesion and growth of the cells on the structures. The fibronectin solution was placed on the substrate and left in the incubator for 2 hours. The solution was removed, and the substrate was washed with PBS.
An Eppendorf single-cell isolation setup was used to pick up and position cells on the stretching structure. This setup has a microcapillary (Piezo Drill Tip ICSI, Eppendorf) with a tip inner diameter of 6 μm. The microcapillary is connected to a pressure controller (CellTram® 4r Air/Oil, Eppendorf) which can control the inside pressure of the pipette. The micropipette position is controlled with a 3D manipulator (TransferMan® 4r, Eppendorf) on an inverted microscope. First, the microcapillary is positioned just above a cell on the substrate and brought into contact with the cell membrane. Then, a negative pressure is applied, suctioning the cell onto the pipette tip. Finally, the cell is retracted from the surface and positioned on the structure and detached from the pipette tip by applying positive pressure. The same procedure is performed to pick up and position the second cell (
To apply displacement to the structure, AFM probes were used. For this purpose, the AFM probe was drilled using FIB etching to make a circular hole with a diameter of 15 μm so that it could capture the pillar (10 μm diameter) on the structure.
The A431 cells were E-cadherin GFP-tagged to visualize the cell-cell junctions. Alexa Fluor™ 657 Phalloidin (Invitrogen) was used to stain the actin filaments and the nuclei were stained with DAPI (Invitrogen). The structures with deposited cells were placed in a glass-bottom petri dish. The cells were washed twice with PBS, pH 7.4, and fixed using 4% formaldehyde solution in PBS for 15 minutes at room temperature, and then washed two times with PBS. Subsequently, they were permeabilized with a solution of 0.1% Triton X-100 in PBS for 15 minutes and then washed twice with PBS. To enhance the quality of the actin fluorescent intensity, 4 drops of Image-iT™ FX Signal Enhancer (Thermofisher) were added and incubated at room temperature with a humid environment for 30 minutes. After removing the solution and washing with PBS, the Phalloidin staining solution with a ratio of 1:100 in PBS was placed on the substrate for 30 minutes at room temperature and then washed with PBS. Next, the DAPI solution with a ratio of 1:1000 with PBS was placed on the substrate and incubated for 10 minutes at room temperature. The solution was removed, and the substrate was washed with PBS. Finally, 3 ml of pure water was added to the petri dish for imaging.
Zyxin staining was performed to visualize the focal adhesion points between cells and the structure. After fixing the cells (see above), the anti-zyxin antibody (Sigma) with a ratio of 1:250 with PBS was added to the sample and refrigerated for 24 hours. The solution was then removed, and the sample was washed with PBS. PBS was replaced by Goat anti-Rabbit IgG (H+L), Superclonal™ Recombinant Secondary Antibody, Alexa Fluor 647 (Thermofisher) and incubated for 1 hour at 37° C. Finally, the sample was washed with PBS and the actin and nuclei staining protocol were performed. Pharmacological treatments modulating cell contractility included 3 μM blebbistatin (Bleb) (Sigma-Aldrich) for 2 h and 1 unit/ml Rho Activator I (CN01; Cytoskeleton, Inc., Denver, CO) for 30 min.
A Nikon Al-NiE upright confocal system (60× water immersion objective) driven by NIS-Elements Confocal image acquisition and analysis program (Nikon software) was used for immunofluorescent imaging of cells on the structures. All image reconstructions and channel alignments were performed within the Nikon software. Zeiss Axio 7 was used for the stretch test. An AFM setup (Nanosurf AG, Switzerland) was installed on the microscope to apply the displacement to the structures.
A431 GFP-tagged E-cadherin cells were lysed with RIPA buffer (50 mM Tris-HCl, pH 7.4, 150 mM NaCl, 5 mM EDTA, 2 mM dithiothreitol, 1 mM PMSF and 1% Triton X-100) containing a protease inhibitor cocktail (58830; Sigma). Whole-cell lysates were incubated on ice for 30 min and then centrifuged at 14000 g for 20 min at 4° C. Proteins were separated by SDS-PAGE using 8% gels and blotted onto PVDF (polyvinylidene fluoride) membranes. The blots were incubated overnight at 4° C. with anti-E-cadherin (BD Biosciences; 610181), or anti-3-Actin (Santa Cruz Biotechnology; SC-47778). Blots were then washed and incubated with HRP-conjugated anti-mouse (Jackson Immunoresearch), followed by washing and detection of immunoreactivity with enhanced chemiluminescence (Santa Cruz Biotechnology).
A modified version of MATLAB digital image correlation (DIC) was used to analyze the frames from the stretch test. The first frame was considered as the reference and the rest of the frames were compared to the reference frame to calculate the displacement of each island. A region of interest with markers within the region was defined for both islands. Then, the MATLAB code calculated the markers' new coordinates with respect to the first frame, from which the displacement of the islands was calculated. The force is defined by the Island 1 displacement multiplied by its stiffness, and the stress is acquired by dividing the force by the junction cross-section (approximately 120 μm2). The strain is then calculated as the difference between the islands' displacements divided by the cell-cell junction's initial length (approximately 20 μm).
Several generations of the sensing beam structure have been designed, fabricated, and tested, and their stiffness was calculated using COMSOL Multiphysics simulation software. The first generation was a group of parallel horizontal beams. A design with 5 sets of beams was proposed as the first design. After the simulation, the calculated stiffness was K=1e5 N/m, which, compared to biological samples, was too large to measure the stress in the cell-cell junction (FIG. 7A). By reducing the number of beams, decreasing the beam width from 5 μm to 2.5 μm, and increasing the beam length from 80 μm to 150 μm, the stiffness was decreased to 4.6 N/m (
All of the horizontal beam designs have a stiffness higher than desired values (0.01 N/m-0.5 N/m). So, a vertical beam design was proposed (
The modulus of elasticity of the printed material varies with laser power and print speed during TPP fabrication. The modulus of elasticity is very important since it affects the stiffness of the structure which is further used to calculate the force and stress. First, a deflection equation was derived for the actuating beam using beam theory for a fixed and guided beam to find the relation between the applied force and the displacement. Then, data from AFM force spectroscopy experiments on the structure were averaged and used to find the actuating beam stiffness. Finally, from the AFM data and the theoretical model, the sensing beam stiffness is obtained.
The actuating side of the microstructure is composed of four main beams, two cross beams, and a top plate. The force is assumed to be evenly distributed to every beam. Further, building on this assumption, it was assumed that each beam would deflect the same. Next, because of the crossbar and the coupling it provides on each beam, the torque that could be attributed to the applied force and the horizontal distance from the base to the top of the beam was neglected. Lastly, the top plate maintained that the end of each beam remained parallel, therefore the system was treated as a fixed and guided beam. For the structure, the beam thickness is consistent, butits cross-section varies (
b(x)=bo(x)−bi(x) (5)
where bo(x) and bi(x) are the length of the outer and inner construction triangles, respectively. The moment of inertia of the beam, I(x), can be expressed as:
I(x)=4*( 1/12b(x)t3)=⅓b(x)t3
Here, four is the number of beams. Using the basic differential equations of the deflection curve of the beam, the deflection of the beam, δact, is derived:
Here, P is the applied force, MB is the moment produced by the top plate on the end of the beam, E is the modulus of elasticity, w is a constant equal to
z is a structural constant equal to z=h−L, and C1 and C2 are the integration constants that come from the boundary conditions for a fixed guided beam:
C
1
=Pz·ln(h)+MB ln(h)−Ph (8)
C
2
=Pz·h(ln(h)+1)+MB·h(ln(h)+1)−½Ph2−C1h (9)
Finally, the stiffness of the structure can be predicted for a given applied force, and the structural constants by the following equation:
As mentioned in the paper, a tipless cantilever probe with a known and thermally tuned stiffness, kp, was used to press on the actuating structure (
P
AFM
=Δx
p
·k
p
=Δx
act.
·k
act.
As a result:
Equation (7) shows that the stiffness is proportional to the cubic thickness, that is:
k
act.
∝t
3 (14)
For the higher resolution experiments, the sensing structure with 2 μm beam thickness was used. The stiffness of this structure will be:
Since the material used for the structure is a polymer, it is possible that viscoelastic effects during the deformation of the structure may result in a nonlinear, rate-dependent relationship between beam deflection and junction stress. To examine the elasticity of the structure, two experiments were performed with a controlled displacement and release. The first one was a 25 μm displacement and sudden release of the structure and the second one was a 50 μm displacement and sudden release. Since lower strain rates have more impact on the viscoelastic properties of the material, 100 nm/s (0.5% s−1) was used for both experiments.
Cell manipulation was performed using the Eppendorf cell isolation system. This setup consists of a microcapillary (Piezo Drill Tip ICSI, Eppendorf) integrated with a pressure controller (CellTram® 4r Air/Oil, Eppendorf) and a 3D manipulator (TransferMan® 4r, Eppendorf), allowing for precise 3D cell manipulation. The inner diameter of the microcapillary was chosen based on the cell diameter (approximately 15 μm). To aspirate and hold a cell on the needle tip, the inner diameter should be less than the cell diameter. Based on available needle sizes from Eppendorf, Piezo Drill Tip ICSI with 6 μm inner diameter was selected. The needle is connected to the capillary and through a tube to the pressure controller. The tube is filled with mineral oil, and a small displacement of the pressure controller cylinder creates positive or negative pressure at the needle tip.
The needle approaches the cell using the 3D manipulator (
Each stretch test was recorded with a screen recorder software (Camtasia) and the movie was divided into frames that were analyzed with a customized DIC-based program to calculate each island's movement. In this method, a region of interest is defined and, within this region, a few markers are placed. The higher the number of markers, the better the resolution of the calculated displacement is. Then, when the coordinates of these markers change, a corresponding red marker appears (
It is assumed that a pair of cells that have a junction in between are attached fully to the substrate and when the force is applied, deformation occurs to half of each cell where cell-ECM adhesion, Therefore, the initial length, L0, is approximately the distance between the two cells' nuclei, which is measured to be approximately 20 μm. D is the forcing island displacement, δ is the sensing island displacement, L0 is the initial length (
Investigation of cellular contractility was performed using CN01, control DMSO, and Bleb with a 0.5% s−1 (100 nm/s) strain rate, and representative frames are shown in
To calculate the junction length, ImageJ (NIH funded software) was used. The scale is assigned to the frames of interest and a freehand line was drawn on the junction (
E-Cadherin siRNA Knockdown
To determine the E-cadherin bond effect on the stress-strain curve and bond rupture initiation, E-cadherin siRNA was transfected into A431 GFP-tagged E-cadherin cells. Cells were incubated with E-cadherin siRNA and Lipofectamine RNAiMAX, and with control siRNA and Lipofectamine RNAiMAX as the control sample for 48 h, 72 h, and 96 h. The inhibition in the expression of the E-cadherin protein was confirmed by fluorescence microscopy and immunoblotting. After 48 h, a specific knockdown of E-cadherin expression was visualized by fluorescence microscopy. Immunoblotting shows that the protein levels of E-cadherin were dramatically decreased in A431 GFP-tagged E-cadherin cells compared to control siRNA. These results demonstrate that E-cadherin siRNA can downregulate the E-cadherin expression effectively.
Each of the following references is incorporated by reference in its entirety for all purposes:
In recent years, two-photon polymerization (TPP), or direct laser writing (DLW), has been widely used to create micro- and nanoscale structures such as microfluidic chambers, three-dimensional (3D) tissue scaffolds, cellular mechanical interrogators, and for in vivo implantation, providing a suite of toolboxes for studying cell mechanobiology in 3D. Complex 3D structures can be fabricated by selectively polymerizing photoresists at the focus of a femtosecond-laser, with the ability to produce features of less than 100 nm. Besides the high resolution, these 3D extracellular environments can be synthesized with tailored mechanical properties, such as stiffness, and topographic characteristics, such as porosity, roughness, and adhesion propensity. Coupling this with excellent biocompatibility of the TPP resins and precision surface functionalization, TPP-printed scaffolds enable the study of a host of cellular behavior, such as 3D cell migration, cancer cell invasion, 3D cell extracellular matrix (ECM) adhesion, and stem cell differentiation. They have also been increasingly used to measure forces generated by single cells, affording a new frontier in mechanobiology.
Studying cellular behavior in 3D environments requires knowledge of the mechanics of the scaffolds in liquid and under physiological conditions. This knowledge plays a crucial role in the quantification of cell mechanical behaviors during cellular interaction with the scaffolds and in guiding the selection of TPP writing parameters to generate scaffolds with desired mechanical behavior. However, previous mechanical studies on TPP-printed structures were carried out in air, due to the limitations of conventional micro- and nanoscale mechanical testing methods. For instance, tensile testing of TPP-printed beam-like structures and nanowires has been performed in air using a conventional micro-electro-mechanical systems (MEMS)-based tensile tester, yielding a Young's modulus of a few GPa for an IP-DIP (a proprietary photoresist from Nanoscribe GmbH) nanowire with a diameter of a few hundred nanometers. These values are an order of magnitude higher than reported from a lone mechanical characterization of Ormocomp performed in liquid with AFM indentation. This large discrepancy in measured Young's moduli lends support to the development of a liquid-based mechanical characterization method. In addition, mechanical properties of the TPP-printed structures vary significantly among a large selection of TPP materials, such as acrylates, SU-8 resins, and hydrogels, and their mechanical properties can be significantly modified in the presence of liquid. Moreover, the choice of writing parameters for polymerization, i.e., laser intensity and writing speed, introduces additional variabilities to the mechanical properties of the resulting scaffolds. Finally, a liquid based mechanical characterization method enables quantitative examination of 4D printed microstructures during their shape-morphing process. 4D printing refers to shape-morphing 3D printed structures, and it has been realized in cellular scaffolds. One of the major environmental stimuli that induce shape-morphing is water content, along with temperature, light, and force interactions. Therefore, a precisely defined mechanical testing method in liquid for TPP-printed structures is strongly warranted.
In this contribution, we report a new method for the mechanical characterization of the mechanical properties of TPP-printed microfibers in liquid environment. The method utilizes two TPP-printed movable plates for actuation and force sensing, which are mechanically coupled by microfibers printed in between and supported by vertical flexible beams. The stress-strain relationship of the microfiber is obtained by applying a displacement with a controlled rate to the actuation plate and evaluating the resulting displacement of the sensing plate. With these stress-strain characterizations, we show that the Young's moduli of the TPP-printed microfibers are significantly reduced in liquid as compared to values obtained in air. Moreover, we show that the mechanical behavior of the microfibers can be tailored by controlling the TPP writing parameters, i.e., laser intensity and writing speed. Further, a size-dependent shape memory effect was observed on the microfibers. Plastically deformed microfibers can recover their pre-deformed shapes in water but not in air. Lastly, the viscoelastic behavior of the microfiber is characterized using tensile tests at different strain rates. The reported methods use of a two-step TPP process can be adopted to perform mechanical characterization of a wide spectrum of biomaterials in liquid conditions, paving the way for reaching the full potential of TPP fabricated 3D scaffolds for mechanobiological studies.
a microscale tensile testing (μTT) device fabricated by TPP
As shown in
The final structures, consisting of the testing device and microfiber specimens, were examined in a scanning electron microscope (SEM) (
The stiffness of the sensing structure, k, was calibrated using AFM. Briefly, a tipless cantilever probe with a calibrated stiffness, kp, was used to apply a force, FAFM, to a sensing stage (
The entire tensile testing setup, including the testing device and the sample along with the AFM probe (attached to the actuation stage with a pin-hole connection), can be completely immersed in liquid to perform tensile tests. The displacement applied to the sample is controlled by the AFM which has a sub-nanometer resolution. The displacements of both stages and the deformation of the microfiber during tensile loading, as shown in a typical sequence in
The working principle of the testing device can be further explained by using a representative experiment described in
where σ0 is the amplitude of relaxation and B is the residual stress in microfibers.
Mechanical Properties of TPP-Printed Microfibers from Tensile Testing
The TPP fabrication process is highly complex and dynamic. The mechanisms and processes have been explored both theoretically and experimentally. As illustrated in
With fixed printing parameters, we first compared the mechanical properties tested in air to those tested in water. As shown in
We then conducted the tensile tests on microfibers printed with IP-Visio resin at different laser powers, writing speeds, and designed cross section dimensions. The experimental data shows that the Young's moduli can be tuned over a range of a few hundred MPa by controlling the printing parameters, indicating a high degree of tunability. A higher laser writing power results in increased Young's modulus and yield strength (
The viscoelastic behavior of the microfibers has been illustrated from the stress relaxation shown in
The observed trend in mechanical properties of microfibers consistently hold for other different combinations of writing parameters and strain rates (
There have been reports showing mechanical properties of TPP materials tested in air correlate with printing parameters. For example, a previous study showed a linear relationship between the measured Young's modulus and the laser writing power. By contrast, our tests conducted in water shows that such correlation is stronger than a linear relationship (
The microfiber tested in water can recover its original shape from a buckled state. This shape memory effect is further explored in this section. First, we show that the microfiber tested in air does not exhibit shape recovery. Like the previous experiment, the microfibers are first stretched in air by pulling the actuation stage to a set displacement of 20 m, followed by an immediate retraction to its initial position (
To further demonstrate the smaller fiber's ability to rapidly recover, the fiber was stretched, allowed to recover, and stretched again.
Shape memory in polymers has been extensively studied in literature, and most recently in TPP printed polymers. In general, the cross-linked shape memory polymer network consists of two segregated components: netpoints and molecular switches. The netpoints are the hard segments such as the covalent bonds between polymer chains, which hold the shape of the polymer, while the molecular switches are the switchable segments. At ground state, the switchable segments are polymer chains relaxed around the netpoints. When external load is applied, the switchable segments are elongated and can be settled at a different morphology, resulting in a shape change. Beyond a critical transition temperature, the switchable segments become flexible and provide entropic elastic behavior, which can drive the polymer to its original relaxed state. This critical transition temperature is usually higher than room temperature, such as recently demonstrated in TPP printed microstructure. However, it is possible to lower the temperature to room temperature by diffusion of low molecular weight molecules into the polymer, such as water in our case, which works as a plasticizer. In our experiments, water molecules diffuse into the microfiber, forming hydrogen bonds with the polymer chains, which increases the chain dynamics and results in the shape recovery. Understanding these recovery dynamics is crucial for the design of 4D printed materials for probing cellular behaviors.
TPP-printed polymeric structures offer unique properties as cellular scaffolds that enable a wide spectrum of biological studies. In this paper, a new experimental method is developed for the characterization of the mechanical behavior of TPP-printed structures in a liquid, a more physiologically relevant environment. The method is demonstrated on TPP printed microfibers as a model study. It is found that the Young's moduli and yield strength of the microfibers are significantly reduced in liquid as compared to values obtained in air, and the mechanical behavior can be tailored over a wide range by controlling the TPP printing parameters. In addition, a size-dependent shape memory effect on the microfiber is found in liquid but not in air. It is envisioned that both the tunable mechanical behavior and shape memory effect could have great potential in mechanobiology applications. Further, the experimental method presented here represents a significant advancement in mechanical testing of TPP fabricated structures and can be used to determine mechanical properties of 3D scaffolds in mechanobiology studies.
TPP fabrication. For this study TT devices are 3D printed using a Photonic Professional GT tool (Nanoscribe GmbH). The TT device is produced using commercially available photo resin IP-S monomer: (7,7,9 (or 7,9,9)-trimethyl-4,13-dioxo-3,14-dioxa-5,12-diazahexadecane-1,16-diyl bismethacrylate; photo initiator: 4,4′-bis(diethylamino) benzophenone) (Nanoscribe GmbH). Once an array of structures is produced, the substrate is removed from the tool, developed in SU-8 developer (1-Methoxy-2-propyl acetate), rinsed with isopropanol and blow dried with a gentle stream of filtered nitrogen. Next, the material resin of interest (IP-Visio, monomer: 7,7,9 (or 7,9,9)-trimethyl-4,13-dioxo-3,14-dioxa-5,12-diazahexadecane-1,16-diyl bismethacrylate; photo initiator: phenyl bis(2,4,6-trimethylbenzoyl)-phosphine oxide) is applied to the substrate and the microscope objective of the TPP tool is again dipped into and focused within the resin. The previously printed structures are found and aligned, then the dumb-bell shaped fiber structures are printed on top of the existing structures. Both resins are acrylate based and bonding the IP-Visio dumbbell to the existing IP-S structure worked well. In the fabrication process, the laser power, the laser writing speed, and the CAD dimension of the fiber were varied.
AFM stiffness calibration. The TPP TT device stiffness was calibrated using AFM. To calibrate the TT device, the structures were printed onto a silicon chip mounted vertically in the sample holder of the TPP tool. The silicon chip was then flipped on its side and secured to another substrate, which produced horizontal structures compatible with AFM probes for force analysis. AFM force-displacement curves were then collected on the horizontal structures in air and in water. The detailed calibration process is described in the SI, Section 2.
SEM imaging to measure fiber dimension. The fiber structures for tensile testing were designed with a square shape and specified width and height parameters. Additionally, the laser writing speed and laser writing power were varied parameters in the writing of the tensile fiber structures. All the mentioned parameters proved to influence the cross-section size of the fiber. To measure the beam cross section, control sets of prints were fabricated to evaluate the effect of the printing parameters. Furthermore, the final row of each testing column of the structure arrays were left un-stretched and used as control measurements for the tensile beams. SEM images were taken of the top view, and a side view with 450 angle of the beams. From here, the corrected height of the beam was then calculated, and the cross section of the beam was found.
Tensile testing experiments. Experiments were conducted under an optical microscope and AFM set-up. The view was in the plane of the structure platform, allowing for precise measurement of the structure displacements. An AFM cantilever probe was milled to have a hole using focused ion beam (FEI Helios FIB/SEM 660). The micro-structure substrate was place onto a piezo-actuated nano-manipulator stage, and then the AFM is mounted on top of it. The AFM probe was then moved into place and hooked onto the peg of the platform. A horizontal displacement d and a displacement rate d were set, and the experiment was then initiated and recorded. Analysis was performed using DIC. It is worth mentioning that the Young's modulus data is calculated by taking the first-order derivative of the linear portions of the stress-strain curves and yield strength is calculated as the 2% yield offset of the corresponding stress-strain curve (
Digital image correlation for data processing. From the DIC, the set AFM displacement was used to calibrate the measurement through a pixel to micron conversion constant ratio. This was a found and averaged value used for all experiments (Table 4). After finding the micron displacement, the force could then be extracted. Further, using the SEM fiber measurements, the associated stress could then be found. For strain, by strategically setting the grids, the difference between the bottom line of the top grid, and the top line of the bottom grid was used to find the length and the strain of the fiber. The DIC tracks three main features: 1) the dumbbell pattern, represented by the cyan box, 2) the edge of the structure, represented by the red line, and 3) the fiber, represented by the green line (
Each of the following references is incorporated by reference in its entirety for all purposes:
The study of mechanical properties of cell-cell junctions has seen increased interest in recent years due to advanced technologies that allow for precise cell manipulation and complex microstructures with fine resolution to interact with cells. For studying cell-cell junction mechanics, two common approaches are dual micropipette aspiration (DPA) and single cell force spectroscopy (SCFS). In DPA, two cells are suctioned onto the tip of two micropipettes, are brought into contact and subsequently pulled apart, and the force within the junction is determined based on the deflection of the micropipette tips. A major limitation of this experimental setup is that cells are not given time to form mature focal adhesions, which play a critical role in adhesion-mediated mechanotransduction. In addition, this platform has difficulty in combining mechanical measurements with fluorescence microscopy. In SCFS, one cell is attached to an atomic force microscopy (AFM) probe and brought into contact with another cell on a petri dish, and then the cells are pulled apart while force is read with the AFM. While this technique allows for a high force sensing resolution, a major limitation is the difficulty in continuously observing the cell-cell junction during stretching due to the orientation of the stretching direction with respect to the imaging plane. In addition, like with DPA, mature cell-cell junctions cannot be given time to form without sacrificing throughput.
To address these limitations, our group designed a single-cell adhesion micro tensile tester (SCAμTT) platform, which allows for stretching a single pair of cells connected by a junction while imaging it and recording stress and strain. In addition, cells are allowed for prolonged growth on the platform to form mature cell-cell junctions and focal adhesions, and since an array of these platforms are fabricated on one substrate, parallel operation and high throughput is achieved. However, a major limitation of this platform for studying mechanotransduction is the optical properties of IP-S, the photoresin it was originally designed to be fabricated with, which produces high autofluorescence during fluorescent imaging. Fluorescent imaging is critical in investigating mechanotransduction, as it allows for real-time visualization of the expression and organization of proteins tagged with fluorescent markers as well as investigation of force-induced protein unfolding with fluorescence resonant energy transfer (FRET). The high background signal produced by IP-S makes imaging these fluorescent tags difficult and makes FRET studies nearly impossible due to its inherently weak signal.
In this report we detail the design and fabrication of a new multi-material based SCAμTT platform compatible with fluorescent imaging. The platform takes advantage of IP-Visio, a new photoresin developed by Nanoscribe with reduced autofluorescence, and IP-S, which was found to have superior mechanical properties and produce more stable platforms than IP-Visio alone. In addition, we incorporated integrated apertures made by evaporating gold on to the substrate to prevent the illumination of IP-S, further reducing background noise and improving signal-to-noise ratio during imaging. With this design, we demonstrated the ability to image F-actin and the nucleus in a pair of keratinocytes as they are stretched up to 250% strain, allowing us to observe junction rupture and F-actin retraction while simultaneously recording the accumulation of up to 80 kPa of stress in the junction. The platform presented here will enable potential studies on mechanotransduction at the cell-cell junction through monitoring the expression and organization of fluorescently tagged proteins and tension levels in mechanosensitive networks using FRET, and the fabrication techniques can be integrated into other TPP-printed platforms for use in cell mechanics studies.
TPP processing was carried out using a Photonic Professional GT (Nanoscribe GmbH) instrument and two proprietary photoresins, IP-S and IP-visio supplied by the vendor. The 3D computer assisted designs (CAD) of SCAμTT platforms were compiled using the 3D graphic editor built into COMSOL Multiphysics software. Subsequently, STL files of 3D CAD designs were exported and converted into job files using DeScribe software. When printing IP-S parts of SCAμTT platforms, the laser beam scanning velocity and laser power were set to 55 mm/s and 65%, respectively. The interlayer and raster distances (commonly referred to as “slicing” and “hatching” distances, respectively) were set to be 0.5 μm and 0.4 μm. For printing IP-Visio parts, scanning velocity, laser power, slicing distance and hatching distance were 30 mm/s, 90%, 0.4 μm and 0.3 μm, respectively. These small slicing and hatching distances provide sufficiently high resolution and surface smoothness the while printing time was reasonably short. After printing arrays of the main (bottom) parts of SCAμTT platforms, the substrates were successively soaked in SU-8 developer for 30 minutes, rinsed with isopropanol, dried and loaded into the TPP tool. A similar development and drying procedures were used after printing the top IP-Visio platforms. Before printing the top IP-Visio platforms alignment of the substrate with already printed IP-S printed structures was done manually by centering each IP-S structure in the field of view of the tool's optical camera.
A431 cells were cultured in Dulbecco's modified Eagle's medium (DMEM) (Thermo Fisher) supplemented with 10% fetal bovine serum (FBS) (Thermo Fisher) and 1% penicillin-streptomycin (P/S) (Thermo Fisher). HaCaT cells were cultured in DMEM with low calcium concentration (Thermo Fisher) and supplemented with 10% FBS, 1% P/S, and 1% GlutaMAX (Thermo Fisher). For experiments, cells were grown in C02 independent DMEM (Thermo Fisher) supplemented with 10% FBS and 1% P/S.
A431 cells were first fixed with 4% paraformaldehyde diluted from 16% paraformaldehyde (Thermo Fisher) in PBS for 10 minutes. Then, cells were permeabilized with 0.1% Triton X-100 (Sigma-Aldrich) for 5 minutes. To stain zyxin, anti-zyxin antibody (Sigma) was diluted to a ratio of 1:250 in PBS, added to the sample, refrigerated for 24 hours, and finally washed twice with PBS. Then, goat anti-rabbit IgG (H+L), Superclonal Recombinant Secondary Antibody, Alexa Fluor 647 (Thermo Fisher) was incubated with the sample for 1 hour at 37° C. To stain F-actin, Alexa Fluor 488 Phalloidin (Invitrogen) diluted in to a 1× concentration in PBS and incubated in the sample at room temperature for 30. Finally, to stain the nucleus, DAPI (Thermo Fisher) was diluted 1:1000 in PBS and incubated with the sample for 10 minutes at room temperature and then filled with PBS for imaging. Between each of the above steps, the sample was washed twice with PBS. HaCaT cells were fixed, permeabilized, and stained for F-actin and the nucleus with the same parameters as A431 cells. Between each step, cells were washed two times by incubating in PBS at room temperature for 4 minutes.
Prior to cell deposition, SCAμTT platform arrays were placed in a glass bottom petri dish and sterilized with 70% ethanol for 1 minute and then soaked in phosphate-buffered saline (PBS) (Thermo Fisher) for 1 hour to dissolve any remaining ethanol. Then, the platforms were coated with Geltrex (Thermo Fisher) for 1 hour to promote adhesion of cells onto the platform. In preparation for cell deposition, 2 mL of C02-independent DMEM was placed in the petri dish. After cells were passaged, 100 μL of cells suspended in DMEM after passaging was dropped on top of the platforms. For depositing cells on the platforms, an Eppendorf single-cell isolation setup was used. The setup consists of a pressure controller (CellTram 4r Air/Oil, Eppendorf) which controls the pressure inside a microcapillary (Piezo Drill Tip ICSI, Eppendorf) with a tip inner diameter of 6 μm. The pressure controller is positioned with a 3D manipulator (TransferMan 4r, Eppendorf), which is on Nikon Eclipse Ti-S microscope within a temperature-controlled chamber at 37° C. To deposit cells, the tip of the microcapillary is brought into contact with a cell on the substrate. Subsequently, a negative pressure is applied to suction the cell to the tip. The cell is brought into one side of the bowtie confinement on top of the SCAμTT platform and the pressure is released to place the cell. The same procedure is then repeated to place another cell on the other side of the confinement, and this process is repeated for each platform in the array.
The cell stretching setup consists of an AFM (Nanosurf AG, Switzerland) on a Zeiss Axio Observer 7 microscope. A hole is drilled in the tip of the AFM probe with FIB etching in a FEI Helios NanoLab 660. For stretching cells on the hybrid design, first the islands are separated by breaking the tethers using the microneedle tip used for cell deposition. The sample is placed under the AFM and the AFM tip with a through hole is lowered to the device to capture the pin on the forcing island. The cell pair was stretched in increments of 5 μm by moving the AFM with the AFM software and imaged at each point from 0 μm to 50 μm.
To quantify the signal intensity from images of disks in the initial comparison of autofluorescence between IP-S and IP-Visio, a circular region of interest was drawn on each disk and the intensity for each channel was calculated with Zeiss software. An average of these values for all disks was then calculated. For quantifying the intensity of stains of zyxin, F-actin, and DAPI, small regions of interest were defined in areas with high signal with well-defined features of the proteins, and the average intensity of these regions of interest was calculated.
To quantify the signal intensity from images of the original SCAμTT platform as well as the hybrid SCAμTT platform without and with integrated apertures, a small rectangular region of interest (
To quantify the intensity of stained nuclei on the platforms, a custom CellProfiler (Broad Institute) pipeline was created to identify the nuclei and calculate their average intensity. To calculate the background signal for each nucleus, images of the platforms without stained cells were taken with the same image acquisition parameters. Then, based on the border of each nuclei identified by CellProfiler, the intensity within the regions that the nuclei contain were calculated. The intensity of an individual nucleus is then calculated by subtracting this value from the original intensity, which would include signal from both the nucleus and background. To calculate signal to noise ratio, this value was divided by the background value for each nucleus, and then these values were averaged. To calculate the adjusted nucleus intensity, which takes into account the potential for variability of staining between samples, the average nucleus intensity in cells growing on the petri dish within which the platforms are placed was calculated, and the nucleus intensity of cells on the platform was divided by this value. This allows us to gauge accurately the effect SCAμTT platforms have on inhibiting signal excitation and collection.
Top plate displacement assumed to match input, used to find relationship between pixels and distance in um. Bottom plate displacement found by tracking bright spots on edge of bowtie with a MATLAB code. To convert bottom plate displacement to force, the stiffness of the vertical beams of the bottom island was calculated based on its dimensions and the dimensions and stiffness of the vertical beams of the original platform. The effective stiffness of the vertical beams supporting the sensing island in the original design was 0.11 N/m. Based on beam bending theory, the stiffness of the beams of the hybrid design with a height of 125 μm and thickness of 4.5 μm was calculated to be 4.85 N/m. From here, stress and strain were calculated as described previously.
Imaging was done on a Zeiss LSM800 Confocal laser scanning microscope and a Zeiss Axio Observer 7 microscope. Images for the initial comparison of autofluorescence between IP-S and IP-Visio were done on the confocal microscope, and all other images were taken on the Observer 7 unless specified otherwise. Imaging parameters such as laser power, master gain, and objective pinhole diameter were kept consistent between experiments to ensure consistency, except for experiments to determine the effect of changing these parameters in the supplemental information. SEM images were taken on a FEI Helios NanoLab 660 SEM.
To calculate the Young's modulus of IP-S and IP-Visio, a Hysitron TI 950 Triboindenter was used to indent disks of IP-S and IP-Visio. First, the reduced modulus Er was calculated through the indention experiment for each material. The Young's modulus was then calculated based on the following equation: E=Er(1−v2), assuming v=0.4.
As a first test to gauge the fluorescent properties of IP-Visio, disks were fabricated from IP-S and IP-Visio and imaged with laser wavelengths of 405 nm, 488 nm, and 640 nm to measure their fluorescence across a range of possible imaging wavelengths. Compared with IP-S, IP-Visio produces far less fluorescence in all channels (
Fabrication of SCAμTT Platform with IP-Visio
After demonstrating the potential of IP-Visio for use in fluorescent imaging of cells, we next sought to fabricate our SCAμTT platform with this material. The platform features two moveable islands supported by vertical beams of known stiffness, on which two cells are placed within the bowtie confinement area on each island. After cells adhere to the islands and form a mutual junction, one island is displaced with an AFM probe, stretching the cells and deflecting the second island, the displacement of which is tracked to determine junctional stress based on the stiffness of the vertical beams (
To produce a stable platform that can be used in experiments, a new design of our SCAμTT platform that incorporates both IP-S and IP-Visio was made. The design uses IP-S for the vertical beams due to its superior strength and crosslink stability, and IP-Visio for the top island where cells are cultured and grow to allow for imaging with reduced background from autofluorescence. In this two-stage fabrication process, IP-S resin is cast on a glass slide coated in a porous silicon oxide layer and is crosslinked to yield the vertical beams. Next, IP-Visio resin is cast on the slide and crosslinked to yield the islands with bowtie confinements (
After preliminary images showed that the illumination of IP-S beams still caused increased background signal projected onto the cell confinement area, we developed a fabrication procedure to integrate apertures made with a gold coating into the porous silicon oxide layer, which can block illumination of the IP-S component. To fabricate the apertures, a disk of IP-S with a diameter matching the desired aperture diameter is crosslinked onto the glass slide before adding the porous silicon oxide layer. Next, a layer of gold is evaporated on the glass slide. As the porous silicon oxide layer helps promote adhesion between TPP printed parts and the glass slide, in its absence the disks can be easily washed off, leaving behind a pinhole in the gold coating layer through which light can pass. After adding the porous silicon oxide layer, the hybrid device can be fabricated as normal after carefully aligning the apertures on the substrate with the TPP printer (
To compare the suppression of background signal generated by autofluorescence with these two new designs and the original design fabricated from IP-S, each platform was imaged and the background signal inside of the bowtie confinement, where cells would be during an experiment, was measured with input laser wavelengths of 353 nm, 488 nm, 545 nm, and 650 nm. For instance, in the 488 nm channel, the original design fabricated with IP-S has a much higher background signal compared to the hybrid design both without and with an optical blocking aperture (
While the optical blocking apertures can suppress background signal produced by the platform, it is important to consider their effect on exciting and collecting signal from cells on the device. Based on the objective and immersion liquid used in imaging, there is a cone of light defined by an angle α within which signal can be collected by the objective. Based on the height of the stretcher and aperture diameter, another cone of light (β) is defined that determines the maximum cone of light produced by signal from the cells that can be seen from underneath. If β is larger than α, signal collection is not impacted (
Representative images of cells on a device with a small angle (14.5 degrees), an intermediate angle (36.9 degrees), and with no aperture (90 degrees) are shown in
We finally demonstrated the utility of the device by stretching a pair of HaCaT cells on the hybrid platform with an optical blocking aperture. After the cells were stained, the IP-Visio links were broken using the deposition microneedle to disconnect the islands and allow for stretching the stained pair of cells. The cells were stretched in increments of 5 μm and imaged for the nucleus and F-actin (
We report the design, fabrication, and testing of a novel TPP-printed platform for investigating mechanotransduction in a pair of cells connected by a mutual junction. First, the fluorescent properties of IP-Visio, a new proprietary photoresist from Nanoscribe, were investigated and compared with IP-S, a photoresist from Nanoscribe that has been used extensively in research that utilizes TPP. Compared with IP-S, IP-Visio was found to be significantly less autoflourescent across a spectrum of tested excitation wavelengths and staining of cells on top of IP-S and IP-Visio showed the potential of stimulating and collecting signal from cells on IP-Visio. However, as revealed with a nanoindentation test, IP-Visio has a lower Young's modulus compared to IP-S, and therefore may not be able to be used in place of IP-S in TPP platforms with flexible components, such as our SCAμTT platform.
To address this shortcoming, we designed a new hybrid multi-material based SCAμTT platform that uses IP-S for the thin vertical beams and IP-Visio for part of the islands on which cells are imaged. To further combat increased background that results in illumination of IP-S during imaging, we developed a fabrication approach for integrating optical apertures with a gold coating on the glass slide on which SCAμTT platforms are printed. Our results showed that, compared to the original SCAμTT platform fabricated with IP-S, the hybrid design produces significantly lower background signal within the cell confinement region, and the integrated apertures further reduce background. Next, considering the potential for the integrated apertures to interfere with the stimulation and collection and signal from cells, we quantified the intensity of nuclei of cells stained on platforms with varying combinations of aperture diameter and height. We found that combinations of these parameters that result in a small characteristic angle α result in reduced signal stimulation and collection with a low signal-to-noise ratio, and platforms with no coating have a low signal-to-noise ratio due to illumination of IP-S increasing background signal. However, platforms with intermediate angles allow for stimulation and collection of almost all signal from cells on the platform while increasing signal-to-noise ratio by blocking the illumination of IP-S. Finally, we demonstrated the potential of the platform for imaging a pair of cells as they are stretched while simultaneously recording stress and strain in the junction. After staining the cells and disconnecting the islands by breaking the connecting tethers, the cells were stretched to almost 250% strain with a peak stress of around 80 kPa. Imaging of F-actin allowed for direct visualization of the cytoskeleton being deformed and the junction rupturing, which was initiated around a strain of 150% just before the peak stress was observed.
In other studies that use TPP platforms for studying cells, a variety of approaches have been used to either limit fluorescence of the TPP materials or to mitigate the effects of illuminating TPP materials on signal-to-noise ratio. To limit fluorescence of TPP materials, materials such as Sudan black have been used to quench autofluorescence in all channels, and dyes have been incorporated the TPP materials to induce high fluorescence in a specific channel and reduce autofluorescence in other channels. While this approach can reduce autofluorescence, these additives can be toxic and can influence the health of cells growing on the platforms. In addition, as the photoinitiator used for crosslinking in TPP is a significant contributor to autofluorescence, another approach is using photoinitiators with lower autofluorescence. However, this can result in a variety of changes to physical or chemical properties of the resulting device, and therefore is not always an option. Also, prolonged illumination of printed materials can reduce their autofluorescence, and can be used as a pretreatment step. However, this method cannot completely eliminate autofluorescence, and some autofluorescence recovers after time and is therefore not compatible for applications where cells are cultured for many hours. To avoid autofluorescence, infrared fluorophores can be used, as materials generally have lower autofluorescence at the excitation wavelength for these fluorophores. However, microscopes equipped with the appropriate filters and detectors for imaging these fluorophores are not common. In addition, image processing techniques can be used estimate autofluorescence by imaging at a slightly different excitation wavelength and subtracting it from the image. However, this assumes that autofluorescence will be the same at different wavelengths, which is not always true, and additionally requires extra images to be taken which can slow the image acquisition rate. Another approach is using microscopes with different methods of image acquisition that mitigate the autofluorescence of TPP materials. For example, in two-photon imaging, infra-red light is focused on the sample, where fluorescence is induced by the simultaneous absorption of two photons in a fluorophore. Infra-red light can pass through materials more easily compared to visible light, which limits exposure of these materials and mitigates autofluorescence. In addition, as we showed here (
The platform designed in this report, as well as the ideas and techniques used in its design, can be used for fluorescent imaging of cells on TPP platforms on standard fluorescent microscopes. The hybrid SCAμTT platform can be used to study mechanotransduction processes in relation to stress and strain in the junction. Signal from FRET is inherently weak, and the high signal-to-noise ratio of our SCAμTT platform with integrated apertures has potential to allow for the stimulation and capturing of this signal while stretching the cell-cell junction. For example, mechanosensitive proteins within cell-cell junctions can be probed with a pair of FRET fluorophores to study the transmission of forces in cell-cell junctions. Forces has been shown to be transmitted through desmosomes and adherens junctions using FRET sensors in desmoplakin and E-cadherin, respectively, under externally applied stretch, and the hybrid SCAμTT platform can be used to further investigate how forces within these individual junctions relate to stress within the entire cell-cell contact. In addition, FRET sensors have been integrated into mechanosensors such as α-catenin to study conformation changes that expose a cryptic binding cite for vinculin, and our platform can be used to investigate the levels of stress in the cell-cell junction that are required for its activation. In addition, the platform can be used for studying phenomena such as force-induced clustering of E-cadherin in the cell-cell junction under applied stress using cells expressing fluorescently tagged E-cadherin, which has been theorized to occur based on simulation studies of E-cadherin dynamics. Further, remodeling of the cytoskeleton has been observed in response to applied forces, such as force across VE-cadherin inducing F-actin polymerization, and can be investigated further on our platform using cells expressing fluorescent tags on cytoskeleton proteins such as F-actin or intermediate filaments. Finally, as the platform allows time for the formation of mature focal adhesions, the transmission of force from the cell-cell junction to cell-ECM junctions can be studied with a variety of techniques, including FRET, fluorescent tags, and cell-ECM force probes such as tension gauge tethers and DNA hairpin sensors. The approach detailed in this report of integrated apertures to block illumination of autofluorescent materials can also be used in other studies. For example, integrated apertures could be used in microfluidic devices fabricated with TPP to block illumination of the materials that make the microchannel, allowing for focusing on cells cultured within the channel.
Despite these promising results, there remain some limitations to our approach. First, we have observed that the hybrid fabrication approach with IP-S and IP-Visio results in residual stress accumulation and dimensional inaccuracies of IP-S components. We hypothesize that these observations are the result of interaction between crosslinked IP-S and uncrosslinked IP-Visio resin during the two-stage fabrication. Residual stress could lead to inaccuracies in calculating the stress in the junction due to the small length scale at which the sensing island deforms under forces tolerated by the cell-cell junction and will be investigated further.
Each of the following references is incorporated by reference in its entirety for all purposes:
Thus, while the invention has been described above in connection with particular embodiments and examples, the invention is not necessarily so limited, and that numerous other embodiments, examples, uses, modifications and departures from the embodiments, examples and uses are intended to be encompassed by the claims attached hereto.
This application is based on, claims priority to, and incorporates herein by reference in their entirety U.S. Provisional Application Ser. No. 63/077,264, filed Sep. 11, 2020 and entitled, “In Situ Mechanical Characterization of a Single Cell-Cell Adhesion Interface Under Large Strain”; U.S. application Ser. No. 17/473,090 filed Sep. 13, 2021 and entitled “In Situ Mechanical Characterization of a Single Cell-Cell Adhesion Interface Under Large Strain”; U.S. Provisional Application Ser. No. 63/398,121 filed Aug. 15, 2022 and entitled “A Mechanical Characterization Method for Two Photon Polymerized Microfibers in Liquid”; and U.S. Provisional Application Ser. No. 63/398,128, filed Aug. 15, 2022 and entitled “A Multi-Material Platform for Imaging of Single Cell-Cell Junctions under Tensile Load Fabricated with Two Photon Polymerization.”
This invention was made with government support under P20 GM113126 and P30 GM127200 awarded by the National Institutes of Health, under 1826135 and 1936065 awarded by the National Science Foundation, and under DE-AC05-000R22725 awarded by the U.S. Department of Energy. The government has certain rights in the invention.
Number | Date | Country | |
---|---|---|---|
63398128 | Aug 2022 | US | |
63398121 | Aug 2022 | US | |
63077264 | Sep 2020 | US |
Number | Date | Country | |
---|---|---|---|
Parent | 17473090 | Sep 2021 | US |
Child | 18449597 | US |