[go: up one dir, main page]

 
 

Epithelial Cell Mechanics: From Physiology to Pathology

A special issue of Cells (ISSN 2073-4409). This special issue belongs to the section "Cell Motility and Adhesion".

Deadline for manuscript submissions: closed (30 June 2021) | Viewed by 74206

Special Issue Editor


E-Mail Website
Guest Editor
Institute of Molecular and Cellular Anatomy, RWTH Aachen University, 52074 Aachen, Germany
Interests: epithelia; cytoskeleton; junctions; mechanobiology
Special Issues, Collections and Topics in MDPI journals

Special Issue Information

Dear Colleagues,

Epithelia protect the body against external mechanical forces while maintaining a tight barrier to the underlying connective tissue. They have developed a particularly resilient cytoskeleton consisting of interconnected filament networks that are anchored to adhesion sites connecting cells to each other and the extracellular matrix. The resulting transcellular scaffold supports epithelial barrier functions, withstands extreme mechanical deformation, and responds to different mechanical environments through mechanosensing and re-organization. The precise contribution of the different cellular and extracellular matrix components to the overall mechanical properties of epithelial tissues is slowly emerging with the advent of novel tools, software, and image processing routines to study mechanical properties in a vital 3D tissue context.

This current volume aims to present new ideas and novel findings on how mechanical factors determine:

  • Epithelial tissue differentiation (e.g., stem cell differentiation, luminogenesis, stratification);
  • Epithelial physiology (e.g., ciliary function, single cell extrusion);
  • Epithelial wound healing;
  • Intraepithelial inflammatory responses;
  • Invasion of epithelial cells into the connective tissue compartment;

and how these insights can be exploited to improve epithelial tissue engineering (e.g., the production of functional organoids).

Dr. Rudolf Leube
Guest Editor

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. Cells is an international peer-reviewed open access semimonthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. The Article Processing Charge (APC) for publication in this open access journal is 2700 CHF (Swiss Francs). Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.

Keywords

  • Epithelial mechanobiology
  • Cytoskeleton
  • Adhesion
  • Extracellular matrix
  • Tissue engineering
  • Mechanical probing

Benefits of Publishing in a Special Issue

  • Ease of navigation: Grouping papers by topic helps scholars navigate broad scope journals more efficiently.
  • Greater discoverability: Special Issues support the reach and impact of scientific research. Articles in Special Issues are more discoverable and cited more frequently.
  • Expansion of research network: Special Issues facilitate connections among authors, fostering scientific collaborations.
  • External promotion: Articles in Special Issues are often promoted through the journal's social media, increasing their visibility.
  • e-Book format: Special Issues with more than 10 articles can be published as dedicated e-books, ensuring wide and rapid dissemination.

Further information on MDPI's Special Issue polices can be found here.

Published Papers (8 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Research

Jump to: Review

22 pages, 5534 KiB  
Article
From Microspikes to Stress Fibers: Actin Remodeling in Breast Acini Drives Myosin II-Mediated Basement Membrane Invasion
by Julian Eschenbruch, Georg Dreissen, Ronald Springer, Jens Konrad, Rudolf Merkel, Bernd Hoffmann and Erik Noetzel
Cells 2021, 10(8), 1979; https://doi.org/10.3390/cells10081979 - 4 Aug 2021
Cited by 8 | Viewed by 3323
Abstract
The cellular mechanisms of basement membrane (BM) invasion remain poorly understood. We investigated the invasion-promoting mechanisms of actin cytoskeleton reorganization in BM-covered MCF10A breast acini. High-resolution confocal microscopy has characterized actin cell protrusion formation and function in response to tumor-resembling ECM stiffness and [...] Read more.
The cellular mechanisms of basement membrane (BM) invasion remain poorly understood. We investigated the invasion-promoting mechanisms of actin cytoskeleton reorganization in BM-covered MCF10A breast acini. High-resolution confocal microscopy has characterized actin cell protrusion formation and function in response to tumor-resembling ECM stiffness and soluble EGF stimulation. Traction force microscopy quantified the mechanical BM stresses that invasion-triggered acini exerted on the BM–ECM interface. We demonstrate that acini use non-proteolytic actin microspikes as functional precursors of elongated protrusions to initiate BM penetration and ECM probing. Further, these microspikes mechanically widened the collagen IV pores to anchor within the BM scaffold via force-transmitting focal adhesions. Pre-invasive basal cells located at the BM–ECM interface exhibited predominantly cortical actin networks and actin microspikes. In response to pro-invasive conditions, these microspikes accumulated and converted subsequently into highly contractile stress fibers. The phenotypical switch to stress fiber cells matched spatiotemporally with emerging high BM stresses that were driven by actomyosin II contractility. The activation of proteolytic invadopodia with MT1-MMP occurred at later BM invasion stages and only in cells already disseminating into the ECM. Our study demonstrates that BM pore-widening filopodia bridge mechanical ECM probing function and contractility-driven BM weakening. Finally, these EMT-related cytoskeletal adaptations are critical mechanisms inducing the invasive transition of benign breast acini. Full article
(This article belongs to the Special Issue Epithelial Cell Mechanics: From Physiology to Pathology)
Show Figures

Figure 1

Figure 1
<p>Breast acini form dynamic BM-breaching MS. The spatial localization of the analyzed confocal image is indicated as a black horizontal bar within the schematically shown red acinus (<b>A</b>,<b>C</b>,<b>F</b>). (<b>A</b>–<b>E</b>): MCF10A acini were fixed and IF-stained. Confocal images of breast acini treated with pro-invasive cues (24 h EGF treatment, glass). In all micrographs, collagen IV signal is shown in red and actin in cyan. (<b>A</b>) Overview of an EHS matrix-embedded hd-BM acinus with BM piercing CPs. (<b>B</b>) Magnification of an intact BM scaffold highlights the collagen IV network pores (white arrows) interleaved by short actin MS. (<b>C</b>) Heterogeneous arrangement and distribution of laterally elongated MS breaching the BM barrier. (<b>D</b>) A local hot spot of high MS density and (<b>E</b>) the formation of parallel MS bundles. (<b>F</b>) Image series of a living MCF10A/RFP-LifeAct acinus (EGF-treated). The image series depicts the lifetime of MS formation (white arrows) (min:sec). For complete image series, see <a href="#app1-cells-10-01979" class="html-app">Supplementary Movie S1</a>. Scale bars: (<b>A</b>,<b>C</b>) = 20 µm; (<b>B</b>,<b>D</b>–<b>F</b>) = 5 µm.</p>
Full article ">Figure 2
<p>MS formation is fueled by tumor-like stiffness and oncogenic EGF. For analyses, MCF10A acini were fixed and IF-stained. (<b>A</b>) The schematic illustrates the acinar area analyzed for MS quantification (yellow area: start at 10 µm above the glass up to the equatorial plane, defined as the cross-section with the largest diameter). MS were counted for several confocal image planes in this area (see material and methods for a detailed description). (<b>B</b>) Scatter plot summarizes MS numbers in non-invasive acini depending on the BM state and oncogenic EGF treatment over time. Counts of protrusions (x) were divided by individual BM perimeters (in µm) of each analyzed confocal image. For a summary of observed acinar radii for BM perimeter calculation, see <a href="#app1-cells-10-01979" class="html-app">Supplementary Figure S1</a>. (<b>C</b>) The experimental workflow of the designed invasion assay illustrates the invasive transition of benign acini transferred on a stiff substrate and conditional EGF stimulation (black box: area of invasion at the cell–BM–substrate interface) (<b>D</b>) A representative image shows local BM disruption (white box) and invasive cell transmigration (white arrow indicates the direction of cell migration) in an EGF-treated acinus. (<b>E</b>) Collectively outgrowing cell clusters showed frequent filopodial CP at the migration front. For additional images, see <a href="#app1-cells-10-01979" class="html-app">Supplementary Movie S4</a>. (<b>F</b>) Scatter plot compares the lateral MS formation in invasive and non-invasive acini fractions depending on pro-invasive cues (24 h on glass). Sample size <span class="html-italic">n</span>: number of analyzed images of at least three independent experiments. Scatter plots: bars: median with 95% confidence interval. For statistical tests mean values per acinus were used: Mann–Whitney U-test with: **** = <span class="html-italic">p</span> &lt; 0.0001, *** = <span class="html-italic">p</span> &lt; 0.001, * = <span class="html-italic">p</span> &lt; 0.05. Scale bars: (<b>D</b>,<b>E</b>) = 20 µm.</p>
Full article ">Figure 3
<p>Subcellular localization of invadopodia markers in acinar and invasive cells. MCF10A acini were fixed and IF-stained. (<b>A</b>–<b>F</b>): Confocal images of pre-invasive breast acini with intact BM treated with pro-invasive cues (ld-BM, 24 h EGF treatment, glass). (<b>A</b>,<b>B</b>): The actin cortex (cyan) of the basal cell layer of pre-invasive acini co-localizes with the invadopodia marker proteins MT1-MMP (magenta), Tks5 (yellow), and cortactin (green) (for single MT1-MMP, TKS5, and cortactin signal see <a href="#app1-cells-10-01979" class="html-app">Supplementary Figure S2A</a>. (<b>C</b>–<b>E</b>): Visualization of the actin cytoskeleton with small single actin-based lateral MS, co-localized with MT1-MMP (<b>C</b>), Tks5 (<b>D</b>) and cortactin (<b>E</b>). (<b>F</b>) Area of high MS density without co-localization of MT1-MMP and actin MS. (<b>G</b>,<b>H</b>) MT1-MMP protein localization at cortical actin structures in cells at the BM–substrate interface (cf. <b>A</b>) (white arrows). (<b>I</b>,<b>J</b>): Micrographs showing MT1-MMP at the migration front of invasive cells after BM breakdown. Antibody control staining is provided in <a href="#app1-cells-10-01979" class="html-app">Supplementary Figure S2D</a>. Scale bars: (<b>A</b>–<b>C</b>,<b>G</b>–<b>J</b>) = 20 µm, magnifications in (<b>C</b>–<b>F</b>) = 5 µm.</p>
Full article ">Figure 4
<p>Tumor-like substrate stiffness and EGF trigger actin remodeling at the acinus-BM–ECM interface. Ld-BM acini were treated with tumor-like conditions (+/− EGF for 1 and 24 h on stiff substrate) and subsequently fixed and IF-stained for analyses. (<b>A</b>) The schematic illustrates the analysis area at the cell–ECM interface (yellow). (<b>B</b>) Confocal image series shows the basal acinar cell layer (1 h on glass, −EGF). (<b>C</b>) A representative image shows the actin networks in MS and SF cells (1 h on glass, +EGF). (<b>D</b>) The schematic illustrates the observed spatial orientation of actin dots, spikes and fibers. (<b>E</b>) A scatter plot for MS and SF cell formation at cell-ECM contact area (basal cell layer of an acinus as shown in (<b>C</b>)) depending on substrate stiffness and duration of EGF stimulation. Cell fractions (%) were calculated by dividing the MS and SF cell numbers by the total cell number at the contact plane of individual acini. (<b>F</b>) A scatter plot for the size of MS and SF cells at the contact area depending on the duration of contact to a stiff substrate and EGF stimulation (<span class="html-italic">n</span> = all analyzed single cells/cell fractions of the observed acini in (<b>E</b>)). The contact area (%) of cell fractions was calculated by dividing the cell area (pixel) of MS and SF cells by the total contact area of individual acini (pixel). (<b>G</b>,<b>H</b>) The graphs show microspike coverage in MS cells (<b>G</b>) and stress fiber coverage in SF cells (<b>H</b>) depending on tumor-like ECM. <span class="html-italic">n</span> = total number of analyzed acini (<b>E</b>) or cells (<b>F</b>–<b>H</b>). Scatter plots and bar charts (<b>E</b>–<b>H</b>): bars: median with 95% confidence interval. For statistical tests mean values per acinus were used: Mann–Whitney U-test with: n.s.: <span class="html-italic">p</span> ≥ 0.05; *: <span class="html-italic">p</span> &lt; 0.05; **: <span class="html-italic">p</span> &lt; 0.01; ***: <span class="html-italic">p</span> &lt; 0.001; ****: <span class="html-italic">p</span> &lt; 0.0001. Scale bars: (<b>B</b>,<b>C</b>) = 20 µm.</p>
Full article ">Figure 5
<p>The actin cytoskeleton intercalates with the BM scaffold. Pre-invasive MCF10A acini were fixed and IF-stained. (<b>A</b>) The orthogonal projection of the basal cell layer that is in direct contact with the underlying substrate (not visible) highlights the connections between actin (cyan) and the BM (collagen IV, red). Left box: F-actin located on top of collagen IV network. Right box: F-actin penetrating the collagen IV network. (<b>B</b>) The micrograph shows BM traversing microspikes (white arrows) in an MS cell of an EGF triggered ld-BM acinus (7 h on glass). (<b>C</b>) Overview image of the acinus–BM–substrate interface shows different BM porosity beneath MS cells (upper box) and SF cells (lower box) (ld-BM, 1 h on glass, –EGF) (same sample analyzed as in <a href="#cells-10-01979-f004" class="html-fig">Figure 4</a>C) (<b>D</b>) Acinus with MS cells (ld-BM, 1 h on glass, –EGF). (<b>E</b>) The orthogonal projection of MS cells shows actin-rich microspikes penetrating the BM collagen IV scaffold. (<b>F</b>) Acinus with SF cells (ld-BM, 24 h on glass, +EGF). (<b>G</b>) The orthogonal projection of SF cells shows actin stress fibers ends at the top of the collagen IV network without detectable penetration. (<b>H</b>) The orthogonal projection of an SF cell highlights a BM-penetrating stress fiber-tip (cf. box in (<b>F</b>)). Scale bars: (<b>C</b>,<b>D</b>,<b>F</b>) = 20 µm, others = 5 µm.</p>
Full article ">Figure 6
<p>Focal adhesion-mediated actomyosin activation in MS and SF cells. MCF10A acini were fixed and IF-stained. (<b>A</b>–<b>D</b>): Representative confocal images of the acinus–BM–substrate interface. (<b>A</b>,<b>C</b>) MS cells (1h on glass, −EGF) and (<b>B</b>,<b>D</b>) SF cells (24 h on glass, +EGF) show the co-localization of actin microspikes and stress fiber ends with vinculin and talin (arrowheads) ((<b>A</b>,<b>B</b>): same sample analyzed as in <a href="#cells-10-01979-f005" class="html-fig">Figure 5</a>D,F). (<b>E</b>) The orthogonal cross sections highlight the co-localization of vinculin with microspikes in MS cells. (<b>F</b>) SF cell with co-localization of vinculin with stress fiber ends within the BM scaffold (yellow). White box: position as indicated in ((<b>B</b>), right arrow) (collagen IV and actin signal confocal images in (<b>E</b>,<b>F</b>) are reused from <a href="#cells-10-01979-f005" class="html-fig">Figure 5</a>E,G). (<b>G</b>) Overview image of the acinus–BM–substrate interface demonstrates pMLCK-actin co-localization. (<b>H</b>) Zoom-in on an invasive cell shows actin stress fiber and pMLCK co-localization after BM transmigration. (<b>I</b>) Detailed view on pMLCK bound to actin microspikes in MS cells and (<b>J</b>) to stress fibers in SF cells. A pMLCK staining of cells on a planar substrate and a secondary antibody control staining is provided in <a href="#app1-cells-10-01979" class="html-app">Supplementary Figure S3B,C</a>. Scale bars in (<b>A</b>,<b>B</b>,<b>G</b>–<b>J</b>) = 20 µm and (<b>C</b>–<b>F</b>) = 5 µm.</p>
Full article ">Figure 7
<p>SF cells of invasion-triggered breast acini exert force hot spots on the acinus–BM–ECM interface. TFM image sequences show mechanical stress at the BM–substrate interface exerted by MCF10A/RFP-LifeAct acini (ld-BM, +EGF) on 12 kPa substrates over time. (<b>A</b>,<b>B</b>) Confocal images of the acinus–ECM interface representing the actin cytoskeleton. (<b>C</b>,<b>D</b>) Matching stress maps (linear lookup table for pseudo-color display) used to calculate total strain energy within the field of view (indicated by white numbers in (<b>C</b>,<b>D</b>,<b>H</b>)). SF cells were marked manually (magenta). The stress map at one hour for Example 1 is provided in <a href="#app1-cells-10-01979" class="html-app">Supplementary Figure S4A</a>. (<b>E</b>,<b>F</b>) Hot spots (hs) in stress maps (light blue patches) were segmented automatically and counted for the spatial correlation with SF cells. (<b>G</b>) The representative image illustrates the counting procedure and the quantitative result of hot spots and stress coherence (in total 292 hot spots of four independent analyzed acini (30 images)). Hs segmentation (light blue patches) and counting (green and red numbers) was carried out every two hours for each acinus within a measuring period of 18–20 h. The localization of 237 hot spots matched with SF cells (green numbers), and 55 did not (red numbers). For a further example of matching hot spots with SF, see <a href="#app1-cells-10-01979" class="html-app">Supplementary Figure S4C</a>. (<b>H</b>) Representative acinus with MS cell treated with blebbistatin (25 µM). Left: actin cytoskeleton, right: TFM stress map (for additional time points, see <a href="#app1-cells-10-01979" class="html-app">Supplementary Figure S4B</a>). (<b>I</b>) Calculated strain energies (SE) of acini treated with blebbistatin compared to untreated samples. The graph shows the SE values measured over a period of 18–20 h (images were taken every seven minutes). Bars show median with 95% confidence interval; Untreated acini: median SE = 32 fJ (<span class="html-italic">n</span> = 4, in total 554 images); Blebbistatin treated acini: median SE = 1.4 fJ (<span class="html-italic">n</span> = 4, in total 429 images); DMSO control (not shown): median SE = 45 fJ (<span class="html-italic">n</span> = 3, in total 438 images). Mann–Whitney U-test with: **** = <span class="html-italic">p</span> &lt; 0.0001. Scale bars: 20 µm. For complete image series of the actin dynamics, see <a href="#app1-cells-10-01979" class="html-app">Supplementary Movies S6 and S7</a>.</p>
Full article ">Figure 8
<p>Graphical summary. A mechanistic model for the dynamic switch from non-proteolytic BM-traversing actin protrusions (CP) to highly contractile ventral stress fibers during the invasive transition of benign breast acini. Under non-invasive conditions and at the early stages of BM invasion, CPs act as mechanical probing units at the acinus–BM–ECM interface. In the course of the invasive transition, actin cytoskeleton remodeling mediates the switch from MS cells (rich in microspikes) to SF cells (rich in stress fibers). Oncogenic EGF and tumor-like ECM stiffness fuel this process. Progressively reinforced SF cells generate local hot spots of high contractile forces at the BM–ECM interface. This stress weakens the BM barrier chronically. The invasive transition of initially benign breast cells converges into BM disruption and finally cell invasion into the microenvironment.</p>
Full article ">
18 pages, 5355 KiB  
Article
Skin under Strain: From Epithelial Model Tissues to Adult Epithelia
by Robin Püllen, Jens Konrad, Rudolf Merkel and Bernd Hoffmann
Cells 2021, 10(7), 1834; https://doi.org/10.3390/cells10071834 - 20 Jul 2021
Cited by 9 | Viewed by 3069
Abstract
Formation of a barrier capable of protecting tissue from external damage, chemical factors, and pathogens is one of the main functions of the epidermis. Furthermore, upon development and during aging, mechanoprotective epidermal functions change dramatically. However, comparative studies between embryonic and adult skin [...] Read more.
Formation of a barrier capable of protecting tissue from external damage, chemical factors, and pathogens is one of the main functions of the epidermis. Furthermore, upon development and during aging, mechanoprotective epidermal functions change dramatically. However, comparative studies between embryonic and adult skin in comparison to skin equivalents are still scarce which is especially due to the lack of appropriate measurement systems with sufficient accuracy and long-term tissue compatibility. Our studies fill this gap by developing a combined bioreactor and tensile testing machine for biomechanical analysis of living epithelia. Based on this tissue stretcher, our data clearly show that viscoelastic and plastic deformation behavior of embryonic and adult skin differ significantly. Tissue responses to static strain compared to cyclic strain also show a clear dependence on differentiation stage. Multilayered unkeratinized epidermis equivalents, on the other hand, respond very similar to mechanical stretch as adult tissue. This mechanical similarity is even more evident after a single cycle of mechanical preconditioning. Our studies therefore suggest that skin equivalents are well suited model systems to analyze cellular interactions of epidermal cells in natural tissues. Full article
(This article belongs to the Special Issue Epithelial Cell Mechanics: From Physiology to Pathology)
Show Figures

Figure 1

Figure 1
<p>Tissue stretcher setup for tensile measurements. The test specimen is mounted between a stationary fixture and a moving grip that can be locked in position until the holding frame is placed in the chamber unit and the needle is hooked into a titanium wire loop. Afterwards, it is precisely moved by a linear step motor and resulting forces are measured by a load cell. Samples can be kept fully viable in a liquid and temperature-controlled environment during the whole experiment. A front glass plate of the incubation chamber allows low-magnification imaging during experiments.</p>
Full article ">Figure 2
<p>Setup calibration: (<b>A</b>) Two different elastomer samples of defined geometry and known Young’s modulus were strained stepwise from 0–0.04, 0.04–0.08 and 0.08–0.12. The resulting forces were measured with a rate of 10 Hz; (<b>B</b>) Taking the sample’s dimensions into consideration, the linear correlation of stress and strain enables the calculation of expected Young’s moduli for every step (indicated in red, green and blue, respectively, in (<b>A</b>) and (<b>B</b>)). Stated here is the mean Young’s modulus of all steps; (<b>C</b>) The sensitivity of the setup was shown by measuring the very small buoyancy correction to the force. To this end, the instrument’s needle was raised in liquid without a sample mounted to the amplitudes used in our experiments.</p>
Full article ">Figure 3
<p>Skin model systems to analyze dermal and epidermal mechanical properties. Three models of different maturation state and composition were comparably prepared: (<b>A</b>) Adult rat skin explants were excised from back skin and the test dimension used for analysis previously marked in situ. The mean dermo-epidermal thickness (exemplarily labeled with red line) was manually determined from tissue slices (<span class="html-italic">n</span> = 16). Underlying hypodermal tissue was omitted here; (<b>B</b>) Back skin from embryonic day 18 rats was cut into indicated test dimensions after isolation. Analysis length was set to 10 mm initial length using a C-shaped PVDF frame (gray). Mean tissue thickness was measured from 14 independent sections (<span class="html-italic">n</span> = 14); (<b>C</b>) Similar PVDF frames were used for transfer of multilayered, simplified epidermis equivalents (SEEs). Therefore, the 20 × 20 mm cell sheets were folded 2 times along one axis to obtain the 5 × 20 mm test dimension. The mean cell sheet thickness is indicated (<span class="html-italic">n</span> = 17); (<b>A</b>–<b>C</b>) Right images show mounted samples previous to start of the test protocol. The lateral white PVDF transfer frame in fetal skin and SEE samples were cut after mounting and had therefore no influence on subsequent force measurements. Scale: 2 mm. Red label illustrates the outline of SEE; (<b>D</b>) Viability assay of epidermal cells via immunofluorescence was performed with fetal rat skin after transfer and stretching for at least 4 h. Non-transferred but identically prepared samples served as controls. Fetal skin in positive controls were deliberately killed with azide. Scale: 50 µm.</p>
Full article ">Figure 4
<p>Repetitive static stretch protocol uncovers viscoelastic properties and tissue strain recovery: (<b>A</b>) The test protocol provides four cycles of 1 h at strain position followed by 1 h at zero position; (<b>B</b>–<b>D</b>) The stress curves of one representative experiment for each model system are shown. Tissue stress was calculated from the estimated cross-sectional area after isolation and fixation (see <a href="#cells-10-01834-f003" class="html-fig">Figure 3</a>). Data was recorded every 0.1 s; (<b>E</b>) Images on the right were taken at indicated time points after moving back to zero position. The tissue dimension before and during stretch are indicated on the left. The visible bending and length recovery (indicated by red arrowheads) was reproducibly observed over several samples and all repetitive cycles.</p>
Full article ">Figure 5
<p>Tissue stress relaxation during repetitive static stretch and rest cycles: (<b>A</b>) To compare tissue behavior for repetitive cycles, force values of every stretch cycle were normalized to its initial peak force; (<b>B</b>–<b>D</b>) The initial stress relaxation curves for each tissue system were plotted as mean in dark color and SD in bright color over time (fetal: <span class="html-italic">n</span> = 3, adult and SEE: <span class="html-italic">n</span> = 4).</p>
Full article ">Figure 6
<p>Tissue force recovery upon repetitive static stretch and rest cycles: (<b>A</b>–<b>C</b>) The initial force decay for each tissue system was plotted as mean in dark color and SD in bright color at relative time scale (fetal: <span class="html-italic">n</span> = 3, adult and SEE: <span class="html-italic">n</span> = 4); (<b>D</b>) The tissue force recovery was calculated as mean and SD of the cycle’s peak force divided by the highest, first peak force value.</p>
Full article ">Figure 7
<p>Static and cyclic strain affects tissues differently: (<b>A</b>) The different tissue models were subjected to either static stretch with a constant load or trapezoid-cyclic stretch with 0.3 Hz but equal strain velocity and amplitude; (<b>B</b>–<b>D</b>) Normalized force relaxation curves of individual experiments were plotted. Note that both stretch protocols were performed at two different amplitudes. For better visualization, only peak forces of cyclic stretch experiments are shown.</p>
Full article ">
15 pages, 4111 KiB  
Article
Mechanical Adaptations of Epithelial Cells on Various Protruded Convex Geometries
by Sun-Min Yu, Bo Li, Steve Granick and Yoon-Kyoung Cho
Cells 2020, 9(6), 1434; https://doi.org/10.3390/cells9061434 - 9 Jun 2020
Cited by 6 | Viewed by 4929
Abstract
The shape of epithelial tissue supports physiological functions of organs such as intestinal villi and corneal epithelium. Despite the mounting evidence showing the importance of geometry in tissue microenvironments, the current understanding on how it affects biophysical behaviors of cells is still elusive. [...] Read more.
The shape of epithelial tissue supports physiological functions of organs such as intestinal villi and corneal epithelium. Despite the mounting evidence showing the importance of geometry in tissue microenvironments, the current understanding on how it affects biophysical behaviors of cells is still elusive. Here, we cultured cells on various protruded convex structure such as triangle, square, and circle shape fabricated using two-photon laser lithography and quantitatively analyzed individual cells. Morphological data indicates that epithelial cells can sense the sharpness of the corner by showing the characteristic cell alignments, which was caused by actin contractility. Cell area was mainly influenced by surface convexity, and Rho-activation increased cell area on circle shape. Moreover, we found that intermediate filaments, vimentin, and cytokeratin 8/18, play important roles in growth and adaptation of epithelial cells by enhancing expression level on convex structure depending on the shape. In addition, microtubule building blocks, α-tubulin, was also responded on geometric structure, which indicates that intermediate filaments and microtubule can cooperatively secure mechanical stability of epithelial cells on convex surface. Altogether, the current study will expand our understanding of mechanical adaptations of cells on out-of-plane geometry. Full article
(This article belongs to the Special Issue Epithelial Cell Mechanics: From Physiology to Pathology)
Show Figures

Figure 1

Figure 1
<p>The shape of epithelial tissue and fabricated convex geometric structure. (<b>A</b>) A schematic of flat and curved epithelium on the designed protruded geometry; triangle, square and circle shape, respectively. The height and diameter of triangle, square, and circle ring structure is fixed as 60 μm. For the cell culture, fibronectin was fully covered on both protruded structure and flat surface. (<b>B</b>) The schematics showing cell orientation (∆<span class="html-italic">θ</span>) on the structure. ∆<span class="html-italic">θ</span> were measured from main axis of cells from the center of each structure (<span class="html-italic">P<sub>a</sub></span>); triangle, square, circle geometry. When the cells are oriented from reference axis from <span class="html-italic">P<sub>a</sub></span> in longitudinal (L) or perpendicular (P) direction, ∆<span class="html-italic">θ</span> equals 0° or 90°, respectively. The scanning electron microscope (SEM) images of the fabricated structure are shown in the bottom panel.</p>
Full article ">Figure 2
<p>Adaptive morphology of epithelial cells on convex geometric structure. (<b>A</b>) Representative fluorescence images of MDCK cell morphology on flat and protruded structures. F-actin (green) was stained with phalloidin-FITC. (<b>B</b>) Averaged cell alignments (∆<span class="html-italic">θ</span>) of the cells on corner of triangle structure (navy), square (orange), and circle (red) shape. (<b>C</b>) Averaged ∆<span class="html-italic">θ</span> of MDCK cells on corner and arm of triangle (navy) and square (orange) structure. (<b>D</b>) Averaged area of the cells on corner of triangle structure, square and circle shape. (<b>E</b>) Averaged area of MDCK cells on corner and arm of triangle and square structure. Number of analyzed cells (<math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>c</mi> <mi>e</mi> <mi>l</mi> <mi>l</mi> </mrow> </msub> </mrow> </semantics></math>) for flat surface and corner of triangle, square, circle structure was 1014 and 212, 471, 749, respectively (<span class="html-italic">N</span> = 3). <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>c</mi> <mi>e</mi> <mi>l</mi> <mi>l</mi> </mrow> </msub> </mrow> </semantics></math> for arm of triangle and square structure were 412 and 580 (<span class="html-italic">N</span> = 3). Error bars in graph indicate the standard error of mean (S.E.M.). One-way ANOVA test with post-hoc Fisher’s least significant difference (LSD) was used. **** <span class="html-italic">p</span> &lt; 0.0001; *** <span class="html-italic">p</span> &lt; 0.001; * <span class="html-italic">p</span> &lt; 0.05; NS, not significant.</p>
Full article ">Figure 3
<p>Actin contractility controls the morphological adaptation of MDCK cells on convex geometric structure. (<b>A</b>) Representative fluorescence images of MDCK cells on the 3D geometry treated with CN03 and Lat-A. F-actin (green) and nucleus (blue) were stained with phalloidin-FITC and 4′,6-diamidino-2-phenylindole (DAPI), respectively. Averaged (<b>B</b>) ∆<span class="html-italic">θ</span>, and (<b>C</b>) area of the cells on corner of triangle (navy), square (orange), and circle (red) structure with CN03 and Lat-A. <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>c</mi> <mi>e</mi> <mi>l</mi> <mi>l</mi> </mrow> </msub> </mrow> </semantics></math> for CN03 treatment on the triangle, square and circle structure were 198, 364, and 430, respectively (<span class="html-italic">N</span> = 3). <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>c</mi> <mi>e</mi> <mi>l</mi> <mi>l</mi> </mrow> </msub> </mrow> </semantics></math> for Lat-A treatment on the triangle, square and circle structure were 287, 343, and 435, respectively (<span class="html-italic">N</span> = 3). Error bars in graph indicate S.E.M. One-way ANOVA test with post-hoc Fisher’s LSD was used. **** <span class="html-italic">p</span> &lt; 0.0001; *** <span class="html-italic">p</span> &lt; 0.001; ** <span class="html-italic">p</span> &lt; 0.01; * <span class="html-italic">p</span> &lt; 0.05; NS, not significant.</p>
Full article ">Figure 4
<p>Enhanced vimentin expression on convex geometric structure. (<b>A</b>) Representative fluorescence images of vimentin (red) in MDCK cells on the flat and triangle, square, circle structure. (<b>B</b>) The box plots indicate the total intensity of vimentin per individual cells on corner of triangle (navy), square (orange) and circle (red) structure. The result from flat surface is plotted in gray color as a control (<span class="html-italic">N</span> = 3). (<b>C</b>) Mean relative vimentin intensity of the cells on the geometric structure compared to that on the flat surface. <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>c</mi> <mi>e</mi> <mi>l</mi> <mi>l</mi> </mrow> </msub> </mrow> </semantics></math> for flat surface and triangle, square, circle structure was 457 and 306, 473, 919, respectively (<span class="html-italic">N</span> = 3). Error bars in graph indicate S.E.M. One-way ANOVA test with post-hoc Fisher’s LSD was used. **** <span class="html-italic">p</span> &lt; 0.0001; ** <span class="html-italic">p</span> &lt; 0.01; NS, not significant.</p>
Full article ">Figure 5
<p>Enhanced cytokeratin 8/18 expression on convex geometric structure. (<b>A</b>) Representative fluorescence images of cytokeratin 8/18 (CK8/18; red) in MDCK cells on the flat and triangle, square, circle structure. (<b>B</b>) The box plots indicate the total intensity of CK8/18 per individual cells on corner of triangle (navy), square (orange), and circle (red) structure. The result from flat surface is plotted in gray color as a control (<span class="html-italic">N</span> = 5). (<b>C</b>) Mean relative CK8/18 intensity of the cells on the geometric structure compared to that on the flat surface. <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>c</mi> <mi>e</mi> <mi>l</mi> <mi>l</mi> </mrow> </msub> </mrow> </semantics></math> for flat surface and triangle, square, circle structure was 4761 and 571, 799, 1324, respectively (<span class="html-italic">N</span> = 5). Error bars in graph indicate S.E.M. One-way ANOVA with post-hoc Fisher’s LSD was used. **** <span class="html-italic">p</span> &lt; 0.0001; ** <span class="html-italic">p</span> &lt; 0.01; NS, non-significant.</p>
Full article ">Figure 6
<p>Disturbing vimentin intermediate filaments (IFs) networks influence cell growth and adaptation on convex structure. Representative fluorescence images of MDCK cells after 3 days of incubation without- and with 2mM acrylamide, causing collapse of vimentin filaments organization. F-actin (green) and nucleus (blue) were stained with phalloidin-FITC and DAPI, respectively.</p>
Full article ">Figure 7
<p>Enhanced α-tubulin expression on convex geometric structure. (<b>A</b>) Representative fluorescence images of α-tubulin (red) in MDCK cells on the flat, triangle, square, and circle structure. (<b>B</b>) The box plots indicate the total intensity of α-tubulin per individual cells on corner of triangle (navy), square (orange), and circle (red) structure. The result from flat surface is plotted in gray color as a control (<span class="html-italic">N</span> = 6). (<b>C</b>) Mean relative α-tubulin intensity of the cells on the geometric structure compared to that on the flat surface. <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>c</mi> <mi>e</mi> <mi>l</mi> <mi>l</mi> </mrow> </msub> </mrow> </semantics></math> for flat surface and triangle, square, circle structure was 4893 and 603, 704, 1487, respectively (<span class="html-italic">N</span> = 6). Error bars in graph indicate S.E.M. One-way ANOVA with post-hoc Fisher’s LSD was used. **** <span class="html-italic">p</span> &lt; 0.0001; NS, non-significant.</p>
Full article ">
12 pages, 3591 KiB  
Article
Transcriptional and Ultrastructural Analyses Suggest Novel Insights into Epithelial Barrier Impairment in Celiac Disease
by Agnieszka Sowińska, Yasser Morsy, Elżbieta Czarnowska, Beata Oralewska, Ewa Konopka, Marek Woynarowski, Sylwia Szymańska, Maria Ejmont, Michael Scharl, Joanna B. Bierła, Marcin Wawrzyniak and Bożena Cukrowska
Cells 2020, 9(2), 516; https://doi.org/10.3390/cells9020516 - 24 Feb 2020
Cited by 11 | Viewed by 3547
Abstract
Disruption of epithelial junctional complex (EJC), especially tight junctions (TJ), resulting in increased intestinal permeability, is supposed to activate the enhanced immune response to gluten and to induce the development of celiac disease (CD). This study is aimed to present the role of [...] Read more.
Disruption of epithelial junctional complex (EJC), especially tight junctions (TJ), resulting in increased intestinal permeability, is supposed to activate the enhanced immune response to gluten and to induce the development of celiac disease (CD). This study is aimed to present the role of EJC in CD pathogenesis. To analyze differentially expressed genes the next-generation mRNA sequencing data from CD326+ epithelial cells isolated from non-celiac and celiac patients were involved. Ultrastructural studies with morphometry of EJC were done in potential CD, newly recognized active CD, and non-celiac controls. The transcriptional analysis suggested disturbances of epithelium and the most significant gene ontology enriched terms in epithelial cells from CD patients related to the plasma membrane, extracellular exome, extracellular region, and extracellular space. Ultrastructural analyses showed significantly tighter TJ, anomalies in desmosomes, dilatations of intercellular space, and shorter microvilli in potential and active CD compared to controls. Enterocytes of fetal-like type and significantly wider adherence junctions were observed only in active CD. In conclusion, the results do not support the hypothesis that an increased passage of gluten peptides by unsealing TJ precedes CD development. However, increased intestinal permeability due to abnormality of epithelium might play a role in CD onset. Full article
(This article belongs to the Special Issue Epithelial Cell Mechanics: From Physiology to Pathology)
Show Figures

Figure 1

Figure 1
<p>Multivariate visualization of the analyzed genes. (<b>a</b>) PCA scatter plot PCA plot showing variance between CD patients’ samples and non-celiac control samples (PC1), and heterogeneity between the five biological replicates in each group (PC2). (<b>b</b>) Volcano plot representing the results of the analysis. Each dot representing one gene, and the blue highlighted genes were significantly differentially expressed (<span class="html-italic">p</span>-value &lt; 0.05) and log fold change cut-off 2. (<b>c</b>) Heatmap shows hierarchical clustering of genes on the left side and the clustering of the samples on the top. The histogram represents the expression data of the significant differentially expressed genes (green is the down-regulated and red is the up-regulated).</p>
Full article ">Figure 2
<p>Results of functional enrichment analysis. The outer circle shows a scatter plot of each term, including their corresponding genes (blue dots down-regulated and red dot up-regulated). Z-score indicates the tendency to increase or decrease each of the gene ontology (GO) terms based on the ratio of the differentially expressed genes. Only the significant terms are displayed in the three main categories: (<b>a</b>) biological process, (<b>b</b>) cellular components, and (<b>c</b>) molecular function.</p>
Full article ">Figure 3
<p>Ultrastructure of small intestine enterocytes from the non-CD control group, patients with potential and active CD (<b>a</b>). Cross-sections of non-CD enterocytes with anchoring filaments (AF), endosomes (e) and no intracellular dilatations between cells present in the control group. Enterocytes in potential and active CD exhibit numerous endosomes (e) and tubules of an apical canicular system (ACS), and dilated intercellular spaces (*). The length and width of the brush border microvilli (<b>b</b>). Measurements were done with the use of the morphometric iTEM program (Olympus) in 10 selected epithelial areas at a magnification of ×60,000, and at least 3 values/patient were obtained. All measurements are presented. Statistical analysis was performed with the use of one-way ANOVA with Tukey correction for multiple comparisons. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> ≤ 0.01, *** <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 4
<p>The proximal region of the enterocytes with tight junctions (TJ), adherence junctions (AJ) and desmosomes (D) from the non-CD control group, patients with potential and active CD (<b>a</b>), and ultrastructural features of intercellular junctions (<b>b</b>,<b>c</b>). Desmosomes with an incorrect asymmetrical structure (D*) present in a patient with active CD. The widths (<b>b</b>) and lengths (<b>c</b>) of EJC were measured using the morphometric iTEM program (Olympus) at a magnification of ×60,000. Measurements were done in 10 selected epithelial areas through longitudinally sectioned intercellular junctions, and at least 5 values of each type of junction/patient were obtained. All measurements are presented. Statistical analysis was performed with the use of one-way ANOVA with Tukey correction for multiple comparisons. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> ≤ 0.01, **** <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">

Review

Jump to: Research

19 pages, 3758 KiB  
Review
How Mechanical Forces Change the Human Endometrium during the Menstrual Cycle in Preparation for Embryo Implantation
by Anna K. Sternberg, Volker U. Buck, Irmgard Classen-Linke and Rudolf E. Leube
Cells 2021, 10(8), 2008; https://doi.org/10.3390/cells10082008 - 6 Aug 2021
Cited by 19 | Viewed by 12729
Abstract
The human endometrium is characterized by exceptional plasticity, as evidenced by rapid growth and differentiation during the menstrual cycle and fast tissue remodeling during early pregnancy. Past work has rarely addressed the role of cellular mechanics in these processes. It is becoming increasingly [...] Read more.
The human endometrium is characterized by exceptional plasticity, as evidenced by rapid growth and differentiation during the menstrual cycle and fast tissue remodeling during early pregnancy. Past work has rarely addressed the role of cellular mechanics in these processes. It is becoming increasingly clear that sensing and responding to mechanical forces are as significant for cell behavior as biochemical signaling. Here, we provide an overview of experimental evidence and concepts that illustrate how mechanical forces influence endometrial cell behavior during the hormone-driven menstrual cycle and prepare the endometrium for embryo implantation. Given the fundamental species differences during implantation, we restrict the review to the human situation. Novel technologies and devices such as 3D multifrequency magnetic resonance elastography, atomic force microscopy, organ-on-a-chip microfluidic systems, stem-cell-derived organoid formation, and complex 3D co-culture systems have propelled the understanding how endometrial receptivity and blastocyst implantation are regulated in the human uterus. Accumulating evidence has shown that junctional adhesion, cytoskeletal rearrangement, and extracellular matrix stiffness affect the local force balance that regulates endometrial differentiation and blastocyst invasion. A focus of this review is on the hormonal regulation of endometrial epithelial cell mechanics. We discuss potential implications for embryo implantation. Full article
(This article belongs to the Special Issue Epithelial Cell Mechanics: From Physiology to Pathology)
Show Figures

Figure 1

Figure 1
<p>The plasticity and contractility of the human uterus require the constant adjustment of the endometrium to prepare and support embryo implantation and enable subsequent embryo development. The scheme highlights the multilayered wall structure of the uterus and menstrual cycle-dependent endometrial changes. The uterus is positioned in the pelvis below the peritoneal cavity. The inverted cone-shaped organ is connected through the Fallopian tube to the ovaries and opens into the vagina. It consists of three major layers: the outer connective tissue-rich perimetrium, the smooth muscle-containing contractile myometrium, and the inner endometrium. The endometrium can be further subdivided into the permanent stratum basale and the transitory stratum functionale, which is shed during desquamation and rebuilt during the proliferative phase, with changes in differentiation during the secretory phase to prepare for blastocyst implantation during the window of implantation. For further details, see Info Boxes 1 and 2 (<a href="#app1-cells-10-02008" class="html-app">Appendix A</a>).</p>
Full article ">Figure 2
<p>Schematic representation of blastocyst adhesion to the endometrial epithelium and subsequent trophoblast invasion into the endometrium. The developing blastocyst—consisting of the fluid-filled blastocoel, the embryoblast encompassing the inner cell mass, and the outer trophoblast cell layer—adheres to the uterine surface epithelium. Adhesion induces processes in both the embryonic blastocyst and the maternal epithelial and stromal cell layers, which result in the transmigration of the trophoblast through the epithelial layer and the invasion of the endometrial stromal compartment. Subsequently, specialized single trophoblast cells, the extravillous trophoblasts, invade the decidua and erode endothelia and glandular epithelia from their respective basal sides. For further details, see Info Box 3 (<a href="#app1-cells-10-02008" class="html-app">Appendix A</a>).</p>
Full article ">Figure 3
<p>The schemes highlight endometrial changes affecting endometrial biomechanics during the menstrual cycle in preparation for blastocyst adhesion and invasion. Key features of the endometrium during the proliferative phase are depicted at the top, and key features during the window of implantation are depicted at the bottom. Note the differences in the endometrial epithelial cell layer including cell height, apical surface specializations (e.g., microvilli and pinopodes), cytoskeletal organization, nuclear morphology and position, distribution of lateral cell–cell junctions, and basolateral plasma membrane invaginations. Additionally note the changes in the connective tissue compartment with a reduced fiber content and altered fibroblast morphology during the window of implantation, thus providing a different biomechanical environment for the epithelial cell layer. The adhering blastocyst impacts the epithelium from the apical side. The red arrows indicate different types of forces that act in all spatial directions within and on the endometrial epithelium. They include shear stress from the uterine lumen, hydrostatic pressure from the adhering blastocyst, tensile forces within the epithelium, traction forces resulting from the epithelial–stromal interphase, and intracellular contractile forces. Epithelial viscoelasticity and extracellular matrix stiffness have impacts on force distribution.</p>
Full article ">Figure 4
<p>The lateralization of desmosomal cell–cell adhesions coincides with the window of implantation and polarization of endometrial epithelial cell lines. The confocal fluorescence micrographs (inverse presentation; single focal planes are shown in (<b>a</b>–<b>c</b>); the projections of 8 consecutive focal planes shown in (<b>d</b>–<b>f</b>)) reveal anti-desmoplakin (Dsp) reactivity that detects punctate desmosomes in the endometrial epithelial cell layer of the human endometrium obtained at different days of the menstrual cycle (<b>a</b>–<b>c</b>) and in gland-like spheroids (<b>d</b>–<b>f</b>) derived from endometrial adenocarcinoma cell lines with high polarity (HEC-1-A), intermediate polarity (Ishikawa), and low polarity (RL95-2). Note the different distributions of desmoplakin-positive desmosomes along the basolateral plasma membrane (arrows). Lu, lumen. Scale bars: 20 μm (<b>a</b>–<b>c</b>), 10 μm (<b>d</b>–<b>f</b>). The images were modified from [<a href="#B27-cells-10-02008" class="html-bibr">27</a>,<a href="#B31-cells-10-02008" class="html-bibr">31</a>].</p>
Full article ">
20 pages, 1453 KiB  
Review
Cell Mechanics in Embryoid Bodies
by Kira Zeevaert, Mohamed H. Elsafi Mabrouk, Wolfgang Wagner and Roman Goetzke
Cells 2020, 9(10), 2270; https://doi.org/10.3390/cells9102270 - 11 Oct 2020
Cited by 34 | Viewed by 11508
Abstract
Embryoid bodies (EBs) resemble self-organizing aggregates of pluripotent stem cells that recapitulate some aspects of early embryogenesis. Within few days, the cells undergo a transition from rather homogeneous epithelial-like pluripotent stem cell colonies into a three-dimensional organization of various cell types with multifaceted [...] Read more.
Embryoid bodies (EBs) resemble self-organizing aggregates of pluripotent stem cells that recapitulate some aspects of early embryogenesis. Within few days, the cells undergo a transition from rather homogeneous epithelial-like pluripotent stem cell colonies into a three-dimensional organization of various cell types with multifaceted cell–cell interactions and lumen formation—a process associated with repetitive epithelial-mesenchymal transitions. In the last few years, culture methods have further evolved to better control EB size, growth, cellular composition, and organization—e.g., by the addition of morphogens or different extracellular matrix molecules. There is a growing perception that the mechanical properties, cell mechanics, and cell signaling during EB development are also influenced by physical cues to better guide lineage specification; substrate elasticity and topography are relevant, as well as shear stress and mechanical strain. Epithelial structures outside and inside EBs support the integrity of the cell aggregates and counteract mechanical stress. Furthermore, hydrogels can be used to better control the organization and lineage-specific differentiation of EBs. In this review, we summarize how EB formation is accompanied by a variety of biomechanical parameters that need to be considered for the directed and reproducible self-organization of early cell fate decisions. Full article
(This article belongs to the Special Issue Epithelial Cell Mechanics: From Physiology to Pathology)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of commonly used methods for the formation of embryoid bodies. (<b>A</b>) EBs are commonly produced from monolayer cultures of pluripotent stem cells using enzymatic treatment to fragment and detach cell colonies from the culture well. (<b>B</b>) Seeding single pluripotent stem cells in hanging drops facilitates the generation of homogeneous EBs. Alternatively, single cells can be spun down in (<b>C</b>) U-shaped bottom wells or (<b>D</b>) AggreWells<sup>TM</sup> to facilitate homogeneous cell aggregation. (<b>E</b>) Bioprinting using cell-laden biological ink allows better control over the composition of EBs. Finally, (<b>F</b>) hydrogels provide an elastic 3D environment which allows the spontaneous formation and differentiation of EBs. (<b>G</b>) Exemplary phase contrast images of EBs produced by enzymatic treatment and (<b>H</b>) EBs produced by a Spin-EB assay starting with a specific cell number (D0) to generate aggregates that are more uniform in size (D1).</p>
Full article ">Figure 2
<p>Gene expression and morphological changes during the course of EB development. (<b>A</b>) Gene expression patterns of embryoid bodies show simultaneous upregulation of germ layer-specific markers (Brachyury, PAX6, GATA6) and downregulation of pluripotency-related genes (OCT4, NANOG, SOX2) over the course of their differentiation. (<b>B</b>) Morphologically, EBs transition from a dense mass of cells into fluid-filled cavitated structures that can later develop additional appendages.</p>
Full article ">Figure 3
<p>Cellular patterning during embryoid body formation. Around day 2, cells on the surface of the EB form a layer of primitive endoderm which exhibits an epithelial morphology with an increased expression of E-cadherin and a sensitive response to FGF signaling. These cells deposit a basement membrane which is rich in laminin and collagen IV. Cells adjacent to the basement membrane receive survival signals, whereas cells without contact undergo caspase-dependent apoptosis. Cavitation inside the EB forms a yolk sac-like structure and results in the formation of a columnar epithelium. Cells between the primitive endoderm and the columnar epithelium are highly responsive to different signaling pathways, including Wnt, Nodal, and BMP signaling.</p>
Full article ">
31 pages, 1435 KiB  
Review
The Intestinal Barrier and Current Techniques for the Assessment of Gut Permeability
by Ida Schoultz and Åsa V. Keita
Cells 2020, 9(8), 1909; https://doi.org/10.3390/cells9081909 - 17 Aug 2020
Cited by 259 | Viewed by 24124
Abstract
The intestinal barrier is essential in human health and constitutes the interface between the outside and the internal milieu of the body. A functional intestinal barrier allows absorption of nutrients and fluids but simultaneously prevents harmful substances like toxins and bacteria from crossing [...] Read more.
The intestinal barrier is essential in human health and constitutes the interface between the outside and the internal milieu of the body. A functional intestinal barrier allows absorption of nutrients and fluids but simultaneously prevents harmful substances like toxins and bacteria from crossing the intestinal epithelium and reaching the body. An altered intestinal permeability, a sign of a perturbed barrier function, has during the last decade been associated with several chronic conditions, including diseases originating in the gastrointestinal tract but also diseases such as Alzheimer and Parkinson disease. This has led to an intensified interest from researchers with diverse backgrounds to perform functional studies of the intestinal barrier in different conditions. Intestinal permeability is defined as the passage of a solute through a simple membrane and can be measured by recording the passage of permeability markers over the epithelium via the paracellular or the transcellular route. The methodological tools to investigate the gut barrier function are rapidly expanding and new methodological approaches are being developed. Here we outline and discuss, in vivo, in vitro and ex vivo techniques and how these methods can be utilized for thorough investigation of the intestinal barrier. Full article
(This article belongs to the Special Issue Epithelial Cell Mechanics: From Physiology to Pathology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A schematic drawing of the intestinal barrier and passage routes across the epithelium. Solutes can pass the intestinal epithelium via either the (<b>A</b>) paracellular route (larger hydrophilic solutes); (<b>B</b>) transcellular route (small hydrophilic and lipophilic solutes) (<b>C</b>); transcellular route via aqueous pores (small hydrophilic solutes) or active carrier-mediated absorption (nutrients); or (<b>D</b>) endocytosis, followed by transcytosis and exocytosis (larger particles, peptides and proteins). The barrier constitutes of (1) the lumen, bacteria and antigens are degraded by biliary juices, gastric and pancreatic acids and the colonization of pathogens is inhibited by commensal bacteria producing antimicrobial substances; (2) the microclimate; unstirred water layer, glycocalyx, bacterial adhesion is prevented by mucus and IgA secretion; (3) the epithelial cells; luminal content is transported while noxious stimuli is impeded by chloride secretion and production of antimicrobial peptides (AMP), junctional complexes between the cells regulate permeability, for details see right panel; (4) the lamina propria; immunoglobulins and cytokines are secreted from cells of both the innate and acquired immunity with direct or indirect effects on permeability, interactions with the endocrine and enteric nervous system. TAMP: tight junction-associated-MARVEL proteins including occludin, tricellulin and Marvel D3; JAM: junctional adhesion molecule; MLC: myosin light chain; MLCK: MLC of myosin II kinase.</p>
Full article ">Figure 2
<p>Overview of the different techniques used to measure intestinal barrier function. It is important to consider the aim of the study and the resources in the laboratory prior choosing the methodology. For a thorough assessment of intestinal barrier function the techniques are preferably combined, for example in vivo and ex vivo techniques can be combined with in vitro studies for a more mechanistic approach. TER: transepithelial resistance.</p>
Full article ">
27 pages, 1685 KiB  
Review
Ciliary Genes in Renal Cystic Diseases
by Anna Adamiok-Ostrowska and Agnieszka Piekiełko-Witkowska
Cells 2020, 9(4), 907; https://doi.org/10.3390/cells9040907 - 8 Apr 2020
Cited by 22 | Viewed by 8762
Abstract
Cilia are microtubule-based organelles, protruding from the apical cell surface and anchoring to the cytoskeleton. Primary (nonmotile) cilia of the kidney act as mechanosensors of nephron cells, responding to fluid movements by triggering signal transduction. The impaired functioning of primary cilia leads to [...] Read more.
Cilia are microtubule-based organelles, protruding from the apical cell surface and anchoring to the cytoskeleton. Primary (nonmotile) cilia of the kidney act as mechanosensors of nephron cells, responding to fluid movements by triggering signal transduction. The impaired functioning of primary cilia leads to formation of cysts which in turn contribute to development of diverse renal diseases, including kidney ciliopathies and renal cancer. Here, we review current knowledge on the role of ciliary genes in kidney ciliopathies and renal cell carcinoma (RCC). Special focus is given on the impact of mutations and altered expression of ciliary genes (e.g., encoding polycystins, nephrocystins, Bardet-Biedl syndrome (BBS) proteins, ALS1, Oral-facial-digital syndrome 1 (OFD1) and others) in polycystic kidney disease and nephronophthisis, as well as rare genetic disorders, including syndromes of Joubert, Meckel-Gruber, Bardet-Biedl, Senior-Loken, Alström, Orofaciodigital syndrome type I and cranioectodermal dysplasia. We also show that RCC and classic kidney ciliopathies share commonly disturbed genes affecting cilia function, including VHL (von Hippel-Lindau tumor suppressor), PKD1 (polycystin 1, transient receptor potential channel interacting) and PKD2 (polycystin 2, transient receptor potential cation channel). Finally, we discuss the significance of ciliary genes as diagnostic and prognostic markers, as well as therapeutic targets in ciliopathies and cancer. Full article
(This article belongs to the Special Issue Epithelial Cell Mechanics: From Physiology to Pathology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The structure of primary cilium. The axoneme is composed of nine pairs of microtubules, anchored in the cell by the basal body. The latter is a modified centriole, consisting of nine triplets of microtubules. The mother centriole plays a key role in ciliogenesis, recruiting the molecules required for axoneme elongation. The daughter centriole results from duplication of mother centriole during S phase [<a href="#B4-cells-09-00907" class="html-bibr">4</a>]). The arrows indicate the key structural cilium elements (the axoneme, transition zone and basal body) as well as proteins involved in kidney ciliopathies and renal cell carcinoma (RCC) [<a href="#B4-cells-09-00907" class="html-bibr">4</a>,<a href="#B5-cells-09-00907" class="html-bibr">5</a>,<a href="#B17-cells-09-00907" class="html-bibr">17</a>,<a href="#B18-cells-09-00907" class="html-bibr">18</a>].</p>
Full article ">Figure 2
<p>The structure of polycystins. LRR: leucine-rich repeats; WSC: cell wall integrity and stress response component domain; Lectin C: lectin C type-3 domain; LDL-A: low-density lipoprotein-A domain, PKD: polycystic kidney disease repeats; GAIN: G-protein-coupled receptor (GPCR) autoproteolysis-inducing domain, PLAT: Polycystin-1, Lipoxygenase, Alpha-Toxin domain. The Figure reprinted and modified with permission from [<a href="#B45-cells-09-00907" class="html-bibr">45</a>] under Creative Commons Attribution 4.0 International (CC BY 4.0).</p>
Full article ">
Back to TopTop