Open access peer-reviewed chapter

Early Cell Response to Mechanical Stimuli during TBI

Written By

Mohammad Mehdi Maneshi and Susan Z. Hua

Reviewed: 30 June 2020 Published: 09 September 2020

DOI: 10.5772/intechopen.93295

From the Edited Volume

Recent Advances in Biomechanics

Edited by Redha Taiar

Chapter metrics overview

1,006 Chapter Downloads

View Full Metrics

Abstract

Traumatic brain injury (TBI) refers to brain damage resulting from external mechanical forces such as a blast or crash. The tissue and cell deformations caused by shear forces are the most common pathological features in TBI and lead to long-term symptoms. Our current understanding of TBI derives mainly from in vivo studies of poststimulus pathology and the effects on brain function. Little is known about the early responses of brain cells during mechanical stimuli. In this chapter, we evaluate the early cell response to the rapid shear forces in vitro. We introduce advanced technologies capable of generating fast shear stimuli mimicking forces occurring in TBI and reporting internal forces in specific proteins at the time of injury. We define the threshold shear forces for calcium influx using an astrocyte model. We describe the spatiotemporal distribution of cytoskeletal forces and correlate them with variations in cell membrane tension. This chapter makes a strong argument that cells’ response to external forces is nonlinear. The heterogeneous distribution of cytoskeletal tension and the gradient of protein forces play a key role in the cells’ response to mechanical stimuli.

Keywords

  • traumatic brain injury (TBI)
  • cell mechanics
  • astrocytes
  • shear stress
  • cytoskeletal forces
  • cell membrane tension

1. Introduction

Traumatic brain injury (TBI) is a form of acquired brain injury that is caused by abrupt external mechanical forces. TBI occurs in both civilians and military veterans from events such as an explosive blast, a blunt impact, and uncontrolled acceleration, deceleration, or rotation of the head. The injury occurs in two phases: a primary injury caused by mechanical forces that occur at the incipient moment of injury and a secondary injury that involves subsequent biochemical and metabolic dysfunction [1, 2, 3]. Depending on the type of forces, the injury is normally classified as either a focal injury that results in cerebral contusions in a specific location or a diffuse brain injury that occurs over a widespread area due to shear forces generated by rapid acceleration or deceleration [4, 5, 6]. While the focal injuries are common in moderate to severe brain injuries, diffuse injuries are the main cause of mild TBI [7]. The deformations caused by shear forces are more difficult to define because they occur throughout the brain and are often invisible to commonly available imaging techniques at the early stages of injury [8, 9].

Most TBI patients exhibit mild or minimally observable damage upon the initial shock, but serious pathology can develop within hours and days [8, 10, 11]. Cell death may proceed from multiple locations within the brain. Many of these sites contain cells with no initially observable injury or proximity to the damage site [12]. Thus, brain cells can respond to low mechanical forces that do not cause an immediate structural damage. Understanding early cell response to mechanical stimuli will help provide valuable insight into the origin and evolution of long-term pathological changes.

Our current understanding of TBI is derived from both in vitro and in vivo studies. In vivo studies provide data on the state of cells well after (hours to days) the injury and probably reflect injuries from downstream processing [13, 14, 15]. In vitro models, using cultured cells, examine cells’ response at the expense of normal physiological interactions [2, 16]. The applied forces for in vitro models are often not on the same timescale as TBI. Recent research has begun using the so-called next-generation injury assessment tools to study specific cell responses in real time.

This chapter examines the earliest cell responses during TBI as seen in astrocytes via changes in Ca2+ levels in vitro. In this chapter, we will introduce advanced technologies that can generate fast shear stimuli mimicking forces that cause TBI and can measure internal forces in specific proteins at the very point of time of injury. We will determine the features of force stimuli that are most susceptible to cells and that lead to long-term alterations. The mechanical properties of cells will be evaluated by real-time measurements of forces in specific cytoskeletal proteins and in cell membranes.

Advertisement

2. Brain structure and traumatic brain injury

2.1 Head structure and brain cells

The brain by its very nature is made of soft substances that are submerged in a thin layer of fluid inside the skull. Underneath the skull, there are multiple layers of tissues separated by intracranial spaces containing water-like liquid, called the cerebrospinal fluid (CSF), as shown in Figure 1. It is believed that this fluid plays an important role in the shock-absorbing capacity of the brain. Further inwards, it is the brain itself that consists of a network of functional brain cells. These cells are arranged into several specialized areas, each performing distinct physiological functions.

Figure 1.

Schematic illustration of brain structure and brain cells.

During TBI, the transient external forces cause linear and angular accelerations, resulting in a range of injuries to the brain. Two main types of injuries may occur: the impact injury due to the brain directly hitting the skull at the point of impact that also generates a whipsaw effect at the opposite side of the brain [17]; and the shear injury due to relative movements between the brain and surrounding tissues and between tissues of different densities [18]. For example, with a sudden acceleration, movement of the brain lags behind that of the skull, producing shear stresses at various interfaces between the brain and the cortical tissues. Similarly, in a deceleration injury, the brain continues its inertial path after the skull has been abruptly halted. Shear injuries commonly occur at gray/white matter junctions, but they are also found in the deeper white matter of the corpus callosum, brain stem, and cerebral cortex [19, 20]. The shear strains and stresses are responsible for producing the loss of consciousness during diffuse axonal injury that accounts for ~60% of hospitalized TBI cases [9, 21].

2.2 Brain cells

Two major cell types in the brain are the neurons and glia. Astrocytes are the most abundant glial cells. While the neurons play the role of processing and transmitting information, astrocytes provide the critical link between the circulatory system and the structural support and the maintenance of neurons (tripartite synapses, neurotransmitter processing, etc.) [22]. Figure 1 (zoom-in panel) illustrates the configuration of the brain cells. During TBI, astrocytes transmit mechanical forces to neurons and (by Newton’s third law) accept forces from them. They can also propagate damage signals via Ca2+ waves, signal to neurons through neurotransmitters, and alter ion concentrations in the extracellular space [22, 23]. By applying a blast-like overpressure to brain slices, it has been shown that astrocyte injury precedes the neuronal injury. This raises the possibility that astrocyte Ca2+ may be one of the earliest responses modulating the brain function [24].

2.3 Earliest cellular events in injury models

Two early events that are caused by external forces are known: a Ca2+ influx and an associated neurotransmitter release [25, 26, 27]. An increase in intracellular Ca2+ is universally seen in brain cells that have been subjected to mechanical perturbations regardless of the nature of the forces [28, 29, 30]. There are many Ca2+-dependent processes that can be activated during TBI, including cell swelling [31, 32, 33], cytoskeletal disruption [34, 35, 36], membrane poration [37, 38], failure of ATP-dependent membrane pumps [39, 40], and mitochondrial dysfunction and apoptosis [41, 42, 43, 44]. The Ca2+-loaded or injured neurons release neurotransmitters including glutamate [45, 46]. The release of glutamate and other excitatory amino acids, in turn, can lead to cytotoxic injury and cell death [47, 48, 49]. These occurrences clearly show that the early responses of brain cells to mechanical stimuli are responsible for the subsequent pathology.

Advertisement

3. Shear forces and their assays

3.1 TBI-induced internal forces

The external force results in the development of internal forces within the skull, causing brain injury. Several types of internal forces can occur depending on the type of incidents. For instance, a blast-generated shockwave initially changes the pressure inside the skull. The abrupt motion of the head causes the acceleration or deceleration to generate compression, stretching, and shear forces at multiple places inside the skull [2].

In vitro studies using cultured cells show that excessive pressure (above 10 atm) is required to stimulate a cellular response. This level of pressure is much higher than the overall pressure measured in vivo within the brain [50]. Using a blast chamber to apply rapid pressures and shear forces to cell cultures, Raven et al. showed that cells are more responsive to shear forces than to hydrostatic pressures [51].

Tensile forces are primarily localized at impact points and have been commonly studied using stretching methods. In stretching models, cell cultures or tissues are placed on a flexible substrate that can be stretched with a vacuum pulse [52] or by piezoelectric actuators [53]. In these studies, cells produced a measurable Ca2+ response only when they experienced very large strains (up to 40%) [54, 55, 56]. In contrast, an overall cell deformation of ~4% due to shear pulses was sufficient to cause a similar Ca2+ response [57]. These studies show that brain cells are more susceptible to shear forces than other types of internal force.

Most of the studies that have investigated diffusive brain injury due to shear forces have focused on the injury of neurons. A sliding between two tissue layers can break the long thin nerve fibers (called axons) of neurons by their extension across the layers [58, 59]. However, shear stresses have been found to induce drastic responses in cells (neurons and astrocytes) even without any noticeable damage to the axons, leading to cell damage and cell death [60]. In other words, subcellular signaling is likely to be a more causal pathway than obvious physical damage in these cells.

3.2 Microfluidic assays for shear forces

Earlier studies variously used a rotating disk, air blow, or pulsed media to apply shear stresses to the cells [61, 62, 63]. Following the application of force stimuli, it was found that transient shear deformations tended to increase the membrane permeability to Ca2+ and small dye molecules and decrease the cell viability [64]. A blast chamber was also used to apply rapid pressure and shear forces to cell cultures. It was found that Ca2+ response was more sensitive to shear forces than to hydrostatic pressures [51]. These methods provide controlled shear stimuli, followed by live cell measurements. In real scenarios of TBI, the forces increase in milliseconds, and small elastic deformations in cells occur in real time [16]. Thus, the challenge for the shear stress assays is the ability to rapidly ramp the shear forces and an ability to reliably measure cells’ response in real time.

In this chapter, we introduce a high-speed programmable microfluidic assay tool to apply precise shear forces to cultured cells [30]. The fluid shear is controlled by a fast pressure servo [65] that is capable of generating pressure pulses with a time resolution of ~1 ms in the microfluidic chamber [30, 65]. The millisecond resolution allows the ability to program arbitrary waveforms emulating the waveforms recorded in TBI [66]. The microfluidic chamber allows high-resolution optical microscopy, which enables in situ measurements of changes in intracellular Ca2+ and protein forces in single cells during the very application of forces. Figure 2 displays the schematic of the experimental setup.

Figure 2.

Schematic of the pressure servo-driven microfluidic chamber. High-speed pressure servo is connected to a microfluidic chamber to generate controlled fluid shear stimuli on cultured cells.

Advertisement

4. The response of cells to transient shear forces

The severity of TBI is conventionally classified as mild, moderate, or severe, depending on the extent of the damage to the brain tissue. In the 1950s, the head injury tolerance to external forces began to be evaluated by dropping embalmed corpse heads on a rigid surface and examining the resulting profile of the skull fracture. Linear skull fracture came to be used as the criterion of injury. These studies led to the first quantitative tolerance curve to be established, now known as the Wayne State Tolerance Curve (WSTC) [67, 68]. This curve describes the influence of magnitude and duration of external loads on the severity of injury to the head.

However, the internal forces are not directly proportional to the external load, which comprise a more complex load profile depending on the type of forces. It is now also known that most cells show a physiological response at much lower internal force levels that are below the injury criteria. Therefore, the thresholds for specific internal forces need to be defined.

4.1 Shear stress thresholds

Using the previously described high-speed servo to generate well-defined shear stresses in a microfluidic chamber, the threshold shear stress to cell response in cultured astrocytes was analyzed. Ca2+ rise being the earliest measurable signal in cells, the Ca2+ rise in the cells was measured in real time. The thresholds were defined with multiple input parameters including magnitude, duration, and load rise time.

The cells responded reliably to a shear stimulus of 23 dyn/cm2 with a transient Ca2+ increase, but they did not respond up to a pulse of 11.5 dyn/cm2, thereby setting an effective threshold [30]. However, the amplitude of the force stimuli is not the only parameter, and a variety of relaxation times has been observed. Changes in pulse duration and rise time also affect the subthreshold responses. For example, a low amplitude of short duration (11.5 dyn/cm2, 10 ms) did not trigger a measurable Ca2+ elevation, whereas a pulse width of 1000 ms of the same amplitude did. Together, the two parameters establish a multidimensional threshold matrix for pulses with a square profile, as shown in Figure 3. The relative significance of each stimulus parameter was evaluated by fitting the profile of Ca2+ response with a two-state Boltzmann equation, which shows the amplitude is the primary determinant [30].

Figure 3.

(a) Profile of shear stimulus thresholds. (b) The damage threshold is defined by a 5% maximum response.

4.2 Cell response is nonlinear

The cell response to shear stimuli is nonlinear. The nonlinear characteristic of Ca2+ response is demonstrated in Figure 4. For example, an abrupt increase in shear stress caused the Ca2+ rise, but a slow increase to the same amplitude failed to activate the cells, as shown in Figure 4(a). The cell sensitivity decays rapidly with an increase in force rise time. Moreover, a brief mechanical shock is more important than a sustained shear force. As shown in Figure 4(b), at short intervals of 10 ms, the pulse train caused a response that was three times higher than a single pulse of the same amplitude and a total pulse duration of 100 ms [30]. Thus, the total energy input of the stimulus is not the only determinant in Ca2+ response; the kinetics of the stimulus (force loading rate) also plays an essential role. This emphasizes the viscoelastic/plastic nature of the force transduction processes in cells [64, 69]. By using sine wave stimuli of different frequencies, it has been shown that cells’ response is also frequency dependent; the peaks in response occur at a higher frequency (>80 Hz) or a lower frequency (<20 Hz) [30].

Figure 4.

Nonlinear Ca2+ response of cells. (a) Ca2+ response to shear pulses of different rise times. (b) Ca2+ response to a train of 10 narrow consecutive pulses (10 ms wide) is three times higher than a single broad pulse (100 ms wide) of the same amplitude. (c) and (d) Ca2+ response to repeated pulses (11.5 dyn/cm2, 10 ms) with intervals of 1 s and 10s, respectively, showing that the response is nonlinear.

The strain rate dependence of cell injury has been observed in 3D matrix cultured neurons and astrocytes, showing an increase in membrane permeability to small molecules and an increase in post-insult cell death [70, 71]. However, this effect was not observed in hippocampal tissues under biaxial stretch at strain rates ranging from 0.1 to 50 s−1 [72]; this rate of ramping is likely outside the sensitivity range of the cells [73]. In addition, different cell types have been found to exhibit different properties, which is not surprising.

4.3 Response to repeated challenges

It has been known for a long time that repeated low-amplitude shocks that never had any clear damage to the cells could be lethal in animal models of TBI [74, 75, 76]. This suggests that low subthreshold mechanical forces can be registered and accumulated by the cells. In other words, cells have a long-term pernicious “memory” to repeated stimuli. To study this memory effect, cells were exposed to repetitive low-amplitude stimuli that alone are not able to produce a Ca2+ rise. The response to a pulse train with 1 s intervals of 10 ms duration revealed that individual cells were activated at different times. About 20% of the cells responded to the second pulse, while 60% of the cells responded to the sixth and subsequent pulses (Figure 4(c)). Reducing the frequency sufficiently can eliminate the response (Figure 4(d)). There are two possibilities for this behavior: the plastic mechanical deformation due to cytoskeletal bonds break and reform, or the accumulation of second messenger (Ca2+) pools. Recent studies on cytoskeletal protein forces show that rapid shear stress can generate prolonged cytoskeletal tension, implying that the cytoskeletal deformation plays a key role.

As discussed in detail in the following sections, cells withstand mechanical loading through support from the cytoskeleton consisting of dynamically cross-linked structural proteins [77, 78, 79, 80]. The deformation of a cell depends on the intrinsic elastic deformation of fibrous cytoskeletal proteins and plastic deformation involving reversible cross-linking and cytoskeleton reorganization [81, 82]. Slow shear loading engages the plastic processes that significantly modify the local forces around the Ca2+ transducers.

The nonlinear response of cells to external forces provides evidence that cells use complex force transduction mechanisms to register the forces. Understanding the force transduction pathways is an important step toward understanding the effectors that lead to TBI pathology.

Advertisement

5. Force transduction mechanisms

While fluid shear stresses act on the apical surface of the cells, this force can be transmitted to force-sensitive molecules that trigger ionic trafficking such as Ca2+ influx and other biochemical reactions via a variety of mechanosensitive mechanisms, a process called mechanotransduction. Recent studies suggest that shear stress transduction is mediated by the cytoskeleton, a filamentous network that connects different regions of the cells and holds the mechanical structure of cells [83, 84, 85].

5.1 Role of actin cytoskeleton in force transduction

The cytoskeleton is the universal mechanical structure of animal cells and responsible for cell shape, topography, and facilitating cell locomotion [86]. Cells resist the external mechanical forces via the cytoskeleton. The cytoskeleton consists of filamentous proteins that are interlinked to form filaments. Actin filaments are the smallest type of filaments. Single filaments of actin bind together via cross-linking proteins, such as α-actinin, forming a network that connects different regions of the cells. Actin bundles are found across the cell body; actin mesh is also found beneath the cell membrane, supporting the plasma membrane.

Like many other subcellular organelles, the actin cytoskeleton can be affected during TBI. Cytoskeleton damage has been observed in various moderate to severe TBI conditions [87, 88]. A number of studies now suggest that trauma-induced neuronal cell death may be preceded by the disruption of the neuronal cytoskeleton [89, 90]. Through its connections, the cytoskeleton can also transmit forces to different functional elements of the cell to alter their functions [86, 91, 92, 93]. Among them are the mechanosensitive channels (MSCs) that permeate cation ions such as Ca2+.

5.2 Mechanosensitive channels and Ca2+ signaling

Mechanosensitive ionic channels are transmembrane proteins that form a pore structure across the cell membrane. These channels are linked to the cytoskeletal proteins; thus, changes in cytoskeletal stresses may open the mechanosensitive channels [94, 95]. Similarly, changes in membrane tension could also alter the channel configuration.

Several MSCs have been identified in astrocytes that are members of the transient receptor potential (TRP) family including TRPV4, TRPC1, TRPC5, and TRPA1 [96, 97, 98, 99]. Studies of sensory neurons have suggested that Piezo channels play an important role in brain cells [100, 101]. Piezo-type MSCs are also present in astrocytes that can be inhibited with specific Piezo channel inhibitor [102]. Using an astrocyte model, a recent study has shown that N-Methyl-D-aspartic acid receptors (NMDARs) are the primary Ca2+ source in astrocytes and fluid shear stimuli can activate NMDARs in the absence of agonists [57].

The fluid shear forces can alter the channel activities via several mechanisms. They can modify the cytoskeletal stresses, and the cytoskeleton under high tension can pull the channel proteins via their links. Many MSCs are known to link with the actin cytoskeleton with cross-linking proteins [103, 104, 105]. In astrocytes and neurons, α-actinin binds to NMDA receptors, providing a mechanical link between NMDA receptors with the underlying cytoskeleton [106]. Shear stress can also cause transient deformation and bending of the lipid bilayer, altering the MSCs directly [107, 108, 109]. Most likely, both mechanisms are correlated.

Advertisement

6. Heterogeneous force distribution in single cells subjected to shear pulses

The shear-induced strains and stresses in the cytoskeleton could be the core of the mechanical response of a cell. Since the cytoskeleton is an anisotropic collection of dynamically cross-linked structural proteins, the cytoskeletal stresses are intrinsically nonuniform. This heterogeneity could be responsible for the nonlinear behavior of cells in response to shear stresses.

6.1 Cytoskeletal force measurements

The recent development of force probes based on fluorescence resonance energy transfer (FRET) technique [110, 111, 112, 113] provides a powerful tool to observe cytoskeletal protein forces in live cells. The probe consists of two mutant fluorophores, a donor and an acceptor, linked with an elastic structure. Mechanical forces change the distance and/or the angle between the donor and the acceptor, resulting in a change in the energy transfer efficiency from donor to the acceptor [112, 113]. These probes were genetically encoded in cross-linking proteins, such as α-actinin and actin, and they are being used to report the strains in cytoskeletal linking proteins when cells are subjected to shear pulses. Using FRET probes, the distribution of protein forces induced by fluid shear stress has been mapped in real time [114].

6.2 Heterogeneous force distribution in single cells

By inserting the FRET-based force probes in actin-binding protein (α-actinin) in astrocytes, the cytoskeletal protein response to shear stress was measured for the first time. Results show that rapid shear stimuli generate nonuniform protein forces in single cells. Figure 5 demonstrates that a narrow square shear pulse produced an immediate increase in force in α-actinin at the upstream end of the cell and a small negative force at the downstream end, resulting in a force gradient along the flow direction. In contrast, a slow ramp to the same force amplitude caused a minimal and more uniform change in actinin force. Moreover, the cytoskeleton structure and its prestress states also influence the cell response. The shear pulse produced significant internal force gradients in softer cells that have fewer bundled actin filaments [114]. These experiments demonstrate that uniform shear stress can generate heterogeneous cytoskeletal forces in single cells. Importantly, the time dependence of the stimulus plays a critical role.

Figure 5.

Rapid fluid shear pulse generates nonuniform force distribution in astrocytes. Panel (a) shows fluorescent images (YFP, CFP) and inverse FRET ratio representing actinin forces in an astrocyte cell. (b) Changes in actinin force in two ROIs in response to a square shear pulse (23 dyn/cm2, 15 ms), showing that the narrow shear pulse produce tension in actinin at the upstream edge and compression at downstream edge. (c) Finite element analysis model of a viscoelastic cells under fluid shear stress.

While fluid velocity in the flow chamber was precisely controlled, the body stresses produced in the cells depend on cell geometry and the distribution of stresses within the cytoskeleton (we now know that stress distribution is heterogeneous). This heterogeneity likely accounts for the observed variability between cells. Thus, an averaged impact force cannot be used as a parameter to specify the activation of cell activities; the stress at local points can be an order of magnitude higher than the average.

If the local cytoskeletal stresses are the direct cause of Ca2+ rise, then they should be correlated in time and space. This correlation cannot be resolved by simple statistical comparisons. The distribution of forces in the structural proteins and the changes in Ca2+ was measured simultaneously using a Quad-View optical imaging system. It was found that the activation of Ca2+ began in regions of higher strain, which normally occurred at the upstream edge of the cells and propagated from the edge to the somata of the cell as a wave [114]. In animal models, the NMDAR subunits are found primarily in the astrocytic processes [115]. This finding supports the earlier speculations that applied force may primarily affect the processes where the cytoskeletal structure is most pronounced.

We have noticed that shear stresses above the Ca2+ threshold did not disrupt cell adhesions nor produced any visible changes in cytoskeletal structures, indicating that the activation of Ca2+-permeable ion channels require much lower forces than breaking of the bonds or irreversible deformations of the cytoskeleton.

Advertisement

7. Shear-induced cell membrane tension

Since the mechanosensitive ion channels are membrane bound, changes in cell membrane tension inevitably affect the ion channel configurations that mediate Ca2+ influx. Indeed, it is known that mechanosensitive channels can be opened by membrane tension [116, 117].

7.1 The cell membrane

The cell membrane consists of a lipid bilayer incorporating the membrane proteins including integral proteins such as transmembrane ion channels and receptors and peripheral proteins that loosely attach to the outer side of the cell membrane. Through the functional proteins, the cell membrane selectively controls the transport of ions, water, and macromolecules between the intracellular and extracellular compartments. Inside the cell, the lipid bilayer intimately adheres to the cortical cytoskeleton that provides the support for membrane topography and integrity [118]. On the outside is a hair-like structure called the glycocalyx. Depending on the cell type and local environments, the cell membrane may have tension at its resting state, called pretension. Several factors contribute to the pretension, including internal forces exerted by the cytoskeleton, osmotic pressure from the cytosol, and the forces resulting from cell-substrate interactions at adhesions that can be passed by cytoskeleton [119].

7.2 Effect of shear stress on cell membrane tension

The bilayer tension can be measured using lipid-soluble molecular rotor probe FCVJ [120] whose mobility is commonly used to extract the lipid bilayer fluidity or viscosity. An increase in bilayer tension increases the fluidity in cell membrane, causing a decrease fluorescent intensity of probes [121]. We measured the bilayer tension in astrocyte membranes using the molecular rotor probe FCVJ with the above-described microfluidic chip. As demonstrated in Figure 6(a), a square pulse of fluid shear (23 dyn/cm2, 400 ms) generates a gradient in the membrane tension, with higher tension at the upstream edge of the cell and a lower tension (compression) at the distal edge. Both tension and compression recover back to the initial state within ~30 ms. In comparison, the same shear pulse generated a much longer-lasting tension in actinin at the upstream edge of the cell. Figure 6(b) and (c) illustrates these different characteristics. Interestingly, the membrane tension at the front edge increased much slower than compression at the downstream edge, suggesting that there exists a pretension in the membrane probably via the cortical actin cytoskeleton. The pretension resists the effect of shear force at the upstream edge. In addition, buckling (rapid compression) can occur at the downstream edge. It has been shown that buckling of the lipid membrane can occur at a similar timescale that takes ~150 ms to saturate [122].

Figure 6.

Change in membrane tension measured using FCVJ molecular rotor incorporated into the astrocyte membrane. (a) Left panel: Membrane tension in response to a shear pulse (23 dyn/cm2, 400 ms), where downward inflection indicates tension and upward inflection compression. Labels 1–4 correspond to the regions shown in the image in the right panel. (b) and (c) comparison of membrane tension and actinin force, respectively, in response to a same shear pulse. They show membrane tension at upstream edge increases much slower than actinin force.

A rapid shear pulse generates a gradient in the membrane tension. The spatiotemporal distribution of tension is dependent on the rise time of the shear force. When subjected to a slow ramp-up shear stress, the tension gradient was reduced significantly.

7.3 Membrane tension gradients are coupled to cytoskeletal forces

It has been shown that mechanosensitive channels can be activated by bilayer tension in the lipid vesicles without the cytoskeleton [123]. However, membrane tension measurements show that the membrane tension at the upstream edge increases rather slowly compared with the Ca2+ rise. This suggests that the Ca2+ channel could not have been activated by the bilayer tension alone and additional mechanisms are likely involved. Slowly ramping the shear stress was not able to change the membrane tension at the upstream edge until it reaches a shear stress threshold. This threshold exactly matches the threshold observed for the rise in the cytoskeletal tension. Therefore, both cytoskeleton and membrane tensions are involved.

Time-dependent analysis shows that shear pulse generates faster and longer-lasting tension in actinin at the front edge of the cell compared to membrane tension, suggesting the cytoskeleton imposes the limiting force at the upstream edge of the cells. In the presence of cytochalasin D that disrupts F-actin, this limit is lifted, and a more predominant membrane tension is seen at the upstream edge compared with controls. While disruption of F-actin diminished tension gradients in the cytoskeleton, it also eliminated the fast influx of Ca2+ at the upstream edge of the cell [114]. Thus, cell membrane tension is modulated by the cytoskeleton stresses since the bilayer alone has a minimal ability to bear a large tension gradient.

Moreover, fluid shear by itself is not likely to affect the cell membrane directly since most of the velocity gradient is absorbed by the glycocalyx, leaving little friction at the bilayer [124]. Fluid drag applied to the cell body will pull on the cytoskeleton, thereby increasing its stress [116, 125].

7.4 Effect of shear stress on cell membrane

The effect of shear stress on the lipid membrane response has been variously modeled. The generally proposed model for lipid bilayers is the fluid mosaic model [126]. This model describes the structure of the plasma membrane as a mosaic of components including phospholipids, cholesterol, proteins, and carbohydrates that give the membrane a 2D viscous fluid characteristic. Since the lipid membrane is expected to be intrinsically incompressible, the membrane tension changes could not be explained by solely considering the pure fluid characteristics. Hence, lipid bilayer and MSC interaction via a fluid, or instability of curved stress states of a non-lamellar lipid bilayer, is not easily justifiable within the scope of the fluid mosaic model [127]. Subsequent studies suggest that the lipid membrane is more mosaic rather than fluid [128]. Other models such as the surface model [129] could predict the interaction of lipids and proteins better by allowing both compression and tension changes in the membrane. Based on such models, interactions of lipid and MSCs could be explained by an increase in distance of lipid hydrophobic head groups in the vicinity of proteins and a decrease in outer layer viscosity of membrane [130]. However, without an input from underneath cytoskeleton, the tension gradient in bilayer is minimal.

This leads to the conclusion that shear stress generates tension gradients in the cell cortex and that membrane tension gradients are coupled to cytoskeletal forces to mediate Ca2+ influx. The time-dependent membrane tension gradient is modulated by the intact cytoskeleton forces.

Advertisement

8. Evaluation of astrocytic Ca2+ response to shear in brain slices

In the brain, neurons and astrocytes are intimately connected and function through a three-dimensional circuit that passes information waves. The interplay between them is evident in bidirectional glutamatergic astrocyte-neuron signaling in a Ca2+-dependent fashion. A common consequence of TBI is the alternations of this information flow.

While in vitro experiments described above permit high-resolution measurements, the environment differs from in vivo. To better approximate the in situ conditions, and to see how much of the in vitro results are applicable in vivo, similar sets of experiments can be performed to brain slides, since slides would contain the native cell types and their local environments as in vivo. To a closer approximation, a modified shear chamber was used to apply fluid shear stress to mechanically stimulate the slices.

The Ca2+ response in acute slices from rats is demonstrated in Figure 7, which shows how shear stimuli modulate Ca2+ response in cells under physiological conditions (with the caveat that these may be treated as samples of extreme TBI). To discriminate the astrocytes from neurons, the slices were loaded with SR101 that serves as a marker for astrocytes, as shown in Figure 7(a). Shear stimulated slices showed an acute Ca2+ increase in selected cells that peaked in 1 to 4 s and returned to baseline levels within 20 s, consistent with observations in cell cultures. Most of the cells showed one dominant peak, but some (~20% of cells) responded with multiple peaks (trace 4, Figure 7(c)). The average peak Ca2+ was much higher than the spontaneous Ca2+ transients that were 10–20% of the shear-induced peak (traces 5 and 6, Figure 7(c)). These Ca2+ peaks were eliminated with 10 μM Gd3+, which is a nonspecific MSC blocker. This confirms the observations that shear stress-induced transient Ca2+ peaks are via MSCs.

Figure 7.

Astrocyte Ca2+ response to a shear pulse in a hippocampal slice. (a) The slice is co-loaded with Flou-4 (green) and SR101 (red). (b) Time sequence of Ca2+ images showing Ca2+ peaks at different times in selected cells. (c) Typical traces of astrocyte Ca2+ response of individual cells. (d) Statistics of peak amplitudes and frequency.

It is worthwhile pointing out that the fluid shear stress generates well-controlled forces on the apical surface of cell cultures. In the slide experiments, once the deformation reaches a deeper layer of the cells, the poroelastic nature of the tissue will also modify the forces, so the stimulus profile itself may change as it propagates.

Advertisement

9. Conclusion

Using advanced technologies that can generate fast shear stimuli mimicking forces that cause TBI, we have demonstrated that cell response to mechanical stimuli is nonlinear and the features of the stimuli play a critical role. Using FRET-based force probes, the nonlinearity is shown to be a direct result of nonuniform force distribution within the cytoskeleton. Rapid shear pulses generate a heterogeneous distribution of cytoskeletal forces in cells, in both time and space. The cytoskeletal forces and their modulation on cell membrane tension open MSCs that mediate Ca2+ response. These early response signals can be small and transient. However, the integration of these signals leads to pathology and the progression of TBI.

Advertisement

Acknowledgments

This work was supported by the National Institutes of Health grant NS085517 and National Science Foundation grants CMMI-1537239 and CMMI-2015964.

References

  1. 1. North SH, Shriver-Lake LC, Taitt CR, Ligler FS. Rapid analytical methods for on-site triage for traumatic brain injury. Annual Review of Analytical Chemistry (Palo Alto, California). 2012;5:35-56
  2. 2. Chen YC, Smith DH, Meaney DF. In-vitro approaches for studying blast-induced traumatic brain injury. Journal of Neurotrauma. 2009;26(6):861-876
  3. 3. Meaney DF, Smith DH. Biomechanics of concussion. Clinics in Sports Medicine. 2011;30(1):19-31
  4. 4. Garman RH, Jenkins LW, Switzer RC 3rd, Bauman RA, Tong LC, Swauger PV, et al. Blast exposure in rats with body shielding is characterized primarily by diffuse axonal injury. Journal of Neurotrauma. 2011;28(6):947-959
  5. 5. DeKosky ST, Blennow K, Ikonomovic MD, Gandy S. Acute and chronic traumatic encephalopathies: Pathogenesis and biomarkers. Nature Reviews. Neurology. 2013;9(4):192-200
  6. 6. Sanborn B, Nie X, Chen W, Weerasooriya T. Inertia effects on characterization of dynamic response of brain tissue. Journal of Biomechanics. 2012;45(3):434-439
  7. 7. Graham DI, Adams JH, Nicoll JA, Maxwell WL, Gennarelli TA. The nature, distribution and causes of traumatic brain injury. Brain Pathology. 1995;5(4):397-406
  8. 8. Kan EM, Ling EA, Lu J. Microenvironment changes in mild traumatic brain injury. Brain Research Bulletin. 2012;87(4-5):359-372
  9. 9. Johnson VE, Stewart W, Smith DH. Axonal pathology in traumatic brain injury. Experimental Neurology. 2013;246:35-43
  10. 10. Koliatsos VE, Cernak I, Xu L, Song Y, Savonenko A, Crain BJ, et al. A mouse model of blast injury to brain: Initial pathological, neuropathological, and behavioral characterization. Journal of Neuropathology and Experimental Neurology. 2011;70(5):399-416
  11. 11. Saljo A, Svensson B, Mayorga M, Hamberger A, Bolouri H. Low-level blasts raise intracranial pressure and impair cognitive function in rats. Journal of Neurotrauma. 2009;26(8):1345-1352
  12. 12. Ellis EF, McKinney JS, Willoughby KA, Liang S, Povlishock JT. A new model for rapid stretch-induced injury of cells in culture: Characterization of the model using astrocytes. Journal of Neurotrauma. 1995;12(3):325-339
  13. 13. Kamnaksh A, Kovesdi E, Kwon SK, Wingo D, Ahmed F, Grunberg NE, et al. Factors affecting blast traumatic brain injury. Journal of Neurotrauma. 2011;28(10):2145-2153
  14. 14. Bramlett HM, Dietrich WD. Progressive damage after brain and spinal cord injury: Pathomechanisms and treatment strategies. Progress in Brain Research. 2007;161:125-141
  15. 15. Rubovitch V, Ten-Bosch M, Zohar O, Harrison CR, Tempel-Brami C, Stein E, et al. A mouse model of blast-induced mild traumatic brain injury. Experimental Neurology. 2011;232(2):280-289
  16. 16. Morrison B 3rd, Elkin BS, Dolle JP, Yarmush ML. In vitro models of traumatic brain injury. Annual Review of Biomedical Engineering. 2011;13:91-126
  17. 17. Ommaya AK, Hirsch AE. Tolerances for cerebral concussion from head impact and whiplash in primates. Journal of Biomechanics. 1971;4(1):13-21
  18. 18. Parker RS. Concussive Brain Trauma: Neurobehavioral Impairment & Maladaptation. Boca Raton, FL, USA: CRC Press; 2011
  19. 19. Ivancevic VG. New mechanics of traumatic brain injury. Cognitive Neurodynamics. 2009;3(3):281-293
  20. 20. Taber KH, Warden DL, Hurley RA. Blast-related traumatic brain injury: What is known? The Journal of Neuropsychiatry and Clinical Neurosciences. 2006;18(2):141-145
  21. 21. van Eijck MM, Schoonman GG, van der Naalt J, de Vries J, Roks G. Diffuse axonal injury after traumatic brain injury is a prognostic factor for functional outcome: A systematic review and meta-analysis. Brain Injury. 2018;32(4):395-402
  22. 22. Chen Y, Swanson RA. Astrocytes and brain injury. Journal of Cerebral Blood Flow and Metabolism. 2003;23(2):137-149
  23. 23. Niggel J, Sigurdson W, Sachs F. Mechanically induced calcium movements in astrocytes, bovine aortic endothelial cells and C6 glioma cells. The Journal of Membrane Biology. 2000;174(2):121-134
  24. 24. Sarntinoranont M, Lee SJ, Hong Y, King MA, Subhash G, Kwon J, et al. High-strain-rate brain injury model using submerged acute rat brain tissue slices. Journal of Neurotrauma. 2012;29(2):418-429
  25. 25. Fineman I, Hovda DA, Smith M, Yoshino A, Becker DP. Concussive brain injury is associated with a prolonged accumulation of calcium: A 45Ca autoradiographic study. Brain Research. 1993;624(1-2):94-102
  26. 26. Shapira Y, Yadid G, Cotev S, Shohami E. Accumulation of calcium in the brain following head trauma. Neurological Research. 1989;11(3):169-172
  27. 27. Mills LR, Velumian AA, Agrawal SK, Theriault E, Fehlings MG. Confocal imaging of changes in glial calcium dynamics and homeostasis after mechanical injury in rat spinal cord white matter. NeuroImage. 2004;21(3):1069-1082
  28. 28. LaPlaca MC, Thibault LE. Dynamic mechanical deformation of neurons triggers an acute calcium response and cell injury involving the N-methyl-D-aspartate glutamate receptor. Journal of Neuroscience Research. 1998;52(2):220-229
  29. 29. Weber JT, Rzigalinski BA, Ellis EF. Traumatic injury of cortical neurons causes changes in intracellular calcium stores and capacitative calcium influx. The Journal of Biological Chemistry. 2001;276(3):1800-1807
  30. 30. Maneshi MM, Sachs F, Hua SZ. A threshold shear force for calcium influx in an astrocyte model of traumatic brain injury. Journal of Neurotrauma. 2015;32(13):1020-1029
  31. 31. Bullock R, Maxwell WL, Graham DI, Teasdale GM, Adams JH. Glial swelling following human cerebral contusion: An ultrastructural study. Journal of Neurology, Neurosurgery, and Psychiatry. 1991;54(5):427-434
  32. 32. Kimelberg HK, Rutledge E, Goderie S, Charniga C. Astrocytic swelling due to hypotonic or high K+ medium causes inhibition of glutamate and aspartate uptake and increases their release. Journal of Cerebral Blood Flow and Metabolism. 1995;15(3):409-416
  33. 33. Povlishock JT, Becker DP, Cheng CL, Vaughan GW. Axonal change in minor head injury. Journal of Neuropathology and Experimental Neurology. 1983;42(3):225-242
  34. 34. Monnerie H, Tang-Schomer MD, Iwata A, Smith DH, Kim HA, Le Roux PD. Dendritic alterations after dynamic axonal stretch injury in vitro. Experimental Neurology. 2010;224(2):415-423
  35. 35. Tang-Schomer MD, Patel AR, Baas PW, Smith DH. Mechanical breaking of microtubules in axons during dynamic stretch injury underlies delayed elasticity, microtubule disassembly, and axon degeneration. The FASEB Journal. 2010;24(5):1401-1410
  36. 36. Siedler DG, Chuah MI, Kirkcaldie MT, Vickers JC, King AE. Diffuse axonal injury in brain trauma: Insights from alterations in neurofilaments. Frontiers in Cellular Neuroscience. 2014;8:429
  37. 37. Farkas O, Lifshitz J, Povlishock JT. Mechanoporation induced by diffuse traumatic brain injury: An irreversible or reversible response to injury? The Journal of Neuroscience. 2006;26(12):3130-3140
  38. 38. Lafrenaye AD, McGinn MJ, Povlishock JT. Increased intracranial pressure after diffuse traumatic brain injury exacerbates neuronal somatic membrane poration but not axonal injury: Evidence for primary intracranial pressure-induced neuronal perturbation. Journal of Cerebral Blood Flow and Metabolism. 2012;32(10):1919-1932
  39. 39. Unterberg AW, Stover J, Kress B, Kiening KL. Edema and brain trauma. Neuroscience. 2004;129(4):1021-1029
  40. 40. Davalos D, Grutzendler J, Yang G, Kim JV, Zuo Y, Jung S, et al. ATP mediates rapid microglial response to local brain injury in vivo. Nature Neuroscience. 2005;8(6):752-758
  41. 41. Pivovarova NB, Andrews SB. Calcium-dependent mitochondrial function and dysfunction in neurons. The FEBS Journal. 2010;277(18):3622-3636
  42. 42. Orrenius S, Zhivotovsky B, Nicotera P. Regulation of cell death: The calcium-apoptosis link. Nature Reviews. 2003;4(7):552-565
  43. 43. Singh IN, Sullivan PG, Deng Y, Mbye LH, Hall ED. Time course of post-traumatic mitochondrial oxidative damage and dysfunction in a mouse model of focal traumatic brain injury: Implications for neuroprotective therapy. Journal of Cerebral Blood Flow and Metabolism. 2006;26(11):1407-1418
  44. 44. Sauerbeck A, Hunter R, Bing G, Sullivan PG. Traumatic brain injury and trichloroethylene exposure interact and produce functional, histological, and mitochondrial deficits. Experimental Neurology. 2012;234(1):85-94
  45. 45. Yi JH, Hazell AS. Excitotoxic mechanisms and the role of astrocytic glutamate transporters in traumatic brain injury. Neurochemistry International. 2006;48(5):394-403
  46. 46. Persson L, Hillered L. Chemical monitoring of neurosurgical intensive care patients using intracerebral microdialysis. Journal of Neurosurgery. 1992;76(1):72-80
  47. 47. Lee JM, Zipfel GJ, Choi DW. The changing landscape of ischaemic brain injury mechanisms. Nature. 1999;399(6738 Suppl):A7-A14
  48. 48. Ankarcrona M, Dypbukt JM, Bonfoco E, Zhivotovsky B, Orrenius S, Lipton SA, et al. Glutamate-induced neuronal death: A succession of necrosis or apoptosis depending on mitochondrial function. Neuron. 1995;15(4):961-973
  49. 49. McAllister TW. Neurobiological consequences of traumatic brain injury. Dialogues in Clinical Neuroscience. 2011;13(3):287-300
  50. 50. Murphy EJ, Horrocks LA. A model for compression trauma: Pressure-induced injury in cell cultures. Journal of Neurotrauma. 1993;10(4):431-444
  51. 51. Ravin R, Blank PS, Steinkamp A, Rappaport SM, Ravin N, Bezrukov L, et al. Shear forces during blast, not abrupt changes in pressure alone, generate calcium activity in human brain cells. PLoS One. 2012;7(6):e39421
  52. 52. Morrison B III, Meaney DF, McIntosh TK. Mechanical characterization of an in vitro device designed to quantitatively injure living brain tissue. Annals of Biomedical Engineering. 1998;26(3):381-390
  53. 53. King AI, Ruan J, Zhou C, Hardy WN, Khalil T. Recent advances in biomechanics of brain injury research: A review. Journal of Neurotrauma. 1995;12(4):651-658
  54. 54. Rzigalinski BA, Weber JT, Willoughby KA, Ellis EF. Intracellular free calcium dynamics in stretch-injured astrocytes. Journal of Neurochemistry. 1998;70(6):2377-2385
  55. 55. Alford PW, Dabiri BE, Goss JA, Hemphill MA, Brigham MD, Parker KK. Blast-induced phenotypic switching in cerebral vasospasm. Proceedings of the National Academy of Sciences of the United States of America. 2011;108(31):12705-12710
  56. 56. Singh P, Doshi S, Spaethling JM, Hockenberry AJ, Patel TP, Geddes-Klein DM, et al. N-methyl-D-aspartate receptor mechanosensitivity is governed by C terminus of NR2B subunit. The Journal of Biological Chemistry. 2012;287(6):4348-4359
  57. 57. Maneshi MM, Maki B, Gnanasambandam R, Belin S, Popescu GK, Sachs F, et al. Mechanical stress activates NMDA receptors in the absence of agonists. Scientific Reports. 2017;7:39610
  58. 58. Pena A, Pickard J, Stiller D, Harris N, Schuhmann M. Brain tissue biomechanics in cortical contusion injury: A finite element analysis. In: Intracranial Pressure and Brain Monitoring XII. Vienna, Austria: Springer; 2005. pp. 333-336
  59. 59. Zhang L, Yang KH, King AI. A proposed injury threshold for mild traumatic brain injury. Journal of Biomechanical Engineering. 2004;126(2):226-236
  60. 60. Ravin R, Blank PS, Steinkamp A, Rappaport SM, Ravin N, Bezrukov L, et al. Shear forces during blast, not abrupt changes in pressure alone, generate calcium activity in human brain cells. PLoS One. 2012;7(6):e39421
  61. 61. LaPlaca MC, Thibault LE. Dynamic mechanical deformation of neurons triggers an acute calcium response and cell injury involving the N-methyl-D-aspartate glutamate receptor. Journal of Neuroscience Research. 1998;52(2):220-229
  62. 62. Chung RS, Staal JA, McCormack GH, Dickson TC, Cozens MA, Chuckowree JA, et al. Mild axonal stretch injury in vitro induces a progressive series of neurofilament alterations ultimately leading to delayed axotomy. Journal of Neurotrauma. 2005;22(10):1081-1091
  63. 63. Chung RS, Staal JA, McCormack GH, Dickson TC, Cozens MA, Chuckowree JA, et al. Mild axonal stretch injury in vitro induces a progressive series of neurofilament alterations ultimately leading to delayed axotomy. Journal of Neurotrauma. 2005;22(10):1081-1091
  64. 64. Cullen DK, Vernekar VN, LaPlaca MC. Trauma-induced plasmalemma disruptions in three-dimensional neural cultures are dependent on strain modality and rate. Journal of Neurotrauma. 2011;28(11):2219-2233
  65. 65. Besch SR, Suchyna T, Sachs F. High-speed pressure clamp. Pflügers Archiv / European Journal of Physiology. 2002;445(1):161-166
  66. 66. Taylor PA, Ludwigsen JS, Ford CC. Investigation of blast-induced traumatic brain injury. Brain Injury. 2014;28(7):879-895
  67. 67. Gurdjian ES, Webster JE. Surgical considerations in cerebro-vascular accidents. Transactions of the American Neurological Association. 1950;51:121-123. Discussion 3-5
  68. 68. Lissner HR, Lebow M, Evans FG. Experimental studies on the relation between acceleration and intracranial pressure changes in man. Surgery, Gynecology & Obstetrics. 1960;111:329-338
  69. 69. LaPlaca MC, Cullen DK, McLoughlin JJ, Cargill RS 2nd. High rate shear strain of three-dimensional neural cell cultures: A new in vitro traumatic brain injury model. Journal of Biomechanics. 2005;38(5):1093-1105
  70. 70. Cullen DK, Vernekar VN, LaPlaca MC. Trauma-induced plasmalemma disruptions in three-dimensional neural cultures are dependent on strain modality and rate. Journal of Neurotrauma. 2011;28(11):2219-2233
  71. 71. Tamayol A, Bahrami M, editors. Analytical solutions for laminar fully-developed flow in microchannels with non-circular cross-section. In: ASME 2009 Fluids Engineering Division Summer Meeting. New York, USA: American Society of Mechanical Engineers; 2009
  72. 72. Cater HL, Sundstrom LE, Morrison B. Temporal development of hippocampal cell death is dependent on tissue strain but not strain rate. Journal of Biomechanics. 2006;39(15):2810-2818
  73. 73. Yeung EW, Whitehead NP, Suchyna TM, Gottlieb PA, Sachs F, Allen DG. Effects of stretch-activated channel blockers on [Ca2+] i and muscle damage in the mdx mouse. The Journal of Physiology. 2005;562(2):367-380
  74. 74. Koch W, Filene W. Beiträge zur experimentellen Chirurgie. Über die commotio cerebri. Arch Klin Chir. 1874;17:P190
  75. 75. Shultz SR, Bao F, Omana V, Chiu C, Brown A, Cain DP. Repeated mild lateral fluid percussion brain injury in the rat causes cumulative long-term behavioral impairments, neuroinflammation, and cortical loss in an animal model of repeated concussion. Journal of Neurotrauma. 2012;29(2):281-294
  76. 76. Smith DH, Johnson VE, Stewart W. Chronic neuropathologies of single and repetitive TBI: Substrates of dementia? Nature Reviews. Neurology. 2013;9(4):211-221
  77. 77. Verma D, Ye N, Meng F, Sachs F, Rahimzadeh J, Hua SZ. Interplay between cytoskeletal stresses and cell adaptation under chronic flow. PLoS One. 2012;7(9):e44167
  78. 78. Meng F, Sachs F. Orientation-based FRET sensor for real-time imaging of cellular forces. Journal of Cell Science. 2012;125(3):743-750
  79. 79. Meng F, Sachs F. Measuring strain of structural proteins in vivo in real time. In: Kohl P, Sachs F, Franz MR, editors. Cardiac Mechano-Electric Coupling and Arrhythmia: From Pipette to Patient. Oxford, UK: Oxford University Press; 2011. pp. 431-434
  80. 80. Rahimzadeh J, Meng F, Sachs F, Wang J, Verma D, Hua SZ. Real-time observation of flow-induced cytoskeletal stress in living cells. American Journal of Physiology. Cell Physiology. 2011;301(3):C646-CC52
  81. 81. Gardel M, Shin J, MacKintosh F, Mahadevan L, Matsudaira P, Weitz D. Elastic behavior of cross-linked and bundled actin networks. Science. 2004;304(5675):1301-1305
  82. 82. Bursac P, Lenormand G, Fabry B, Oliver M, Weitz DA, Viasnoff V, et al. Cytoskeletal remodelling and slow dynamics in the living cell. Nature Materials. 2005;4(7):557-561
  83. 83. Ito S, Majumdar A, Kume H, Shimokata K, Naruse K, Lutchen KR, et al. Viscoelastic and dynamic nonlinear properties of airway smooth muscle tissue: Roles of mechanical force and the cytoskeleton. American Journal of Physiology. Lung Cellular and Molecular Physiology. 2006;290(6):L1227-L1237
  84. 84. Alenghat FJ, Nauli SM, Kolb R, Zhou J, Ingber DE. Global cytoskeletal control of mechanotransduction in kidney epithelial cells. Experimental Cell Research. 2004;301(1):23-30
  85. 85. Vogel V, Sheetz M. Local force and geometry sensing regulate cell functions. Nature Reviews. Molecular Cell Biology. 2006;7:265-275
  86. 86. Stricker J, Falzone T, Gardel ML. Mechanics of the F-actin cytoskeleton. Journal of Biomechanics. 2010;43(1):9-14
  87. 87. Deng Y, Thompson BM, Gao X, Hall ED. Temporal relationship of peroxynitrite-induced oxidative damage, calpain-mediated cytoskeletal degradation and neurodegeneration after traumatic brain injury. Experimental Neurology. 2007;205(1):154-165
  88. 88. Graham DI, Saatman KE, McIntosh TK. The neuronal cytoskeleton is at risk after mild and moderate brain injury. Journal of Neurotrauma. 1998;15(12):1047-1058
  89. 89. Newcomb JK, Kampfl A, Posmantur RM, Zhao X, Pike BR, Liu S-J, et al. Immunohistochemical study of calpain-mediated breakdown products to α-spectrin following controlled cortical impact injury in the rat. Journal of Neurotrauma. 1997;14(6):369-383
  90. 90. Posmantur R, Newcomb J, Kampfl A, Hayes R. Light and confocal microscopic studies of evolutionary changes in neurofilament proteins following cortical impact injury in the rat. Experimental Neurology. 2000;161(1):15-26
  91. 91. Orr AW, Helmke BP, Blackman BR, Schwartz MA. Mechanisms of Mechanotransduction. Developmental Cell. 2006;10(1):11-20
  92. 92. Gardel ML, Shin JH, MacKintosh FC, Mahadevan L, Matsudaira P, Weitz DA. Elastic behavior of cross-linked and bundled actin networks. Science. 2004;304(5675):1301-1305
  93. 93. Gardel ML, Nakamura F, Hartwig JH, Crocker JC, Stossel TP, Weitz DA. Prestressed F-actin networks cross-linked by hinged filamins replicate mechanical properties of cells. Proceedings of the National Academy of Sciences of the United States of America. 2006;103(6):1762-1767
  94. 94. Mills J, Mandel L. Cytoskeletal regulation of membrane transport events. The FASEB Journal. 1994;8(14):1161-1165
  95. 95. Hayakawa K, Tatsumi H, Sokabe M. Actin stress fibers transmit and focus force to activate mechanosensitive channels. Journal of Cell Science. 2008;121:496-503
  96. 96. Ostrow LW, Sachs F. Mechanosensation and endothelin in astrocytes--hypothetical roles in CNS pathophysiology. Brain Research. Brain Research Reviews. 2005;48(3):488-508
  97. 97. Suchyna TM, Tape SE, Koeppe RE, Andersen OS, Sachs F, Gottlieb PA. Bilayer-dependent inhibition of mechanosensitive channels by neuroactive peptide enantiomers. Nature. 2004;430(6996):235-240
  98. 98. Ostrow LW, Suchyna TM, Sachs F. Stretch induced endothelin-1 secretion by adult rat astrocytes involves calcium influx via stretch-activated ion channels (SACs). Biochemical and Biophysical Research Communications. 2011;410(1):81-86
  99. 99. Shigetomi E, Tong X, Kwan KY, Corey DP, Khakh BS. TRPA1 channels regulate astrocyte resting calcium and inhibitory synapse efficacy through GAT-3. Nature Neuroscience. 2012;15(1):70-80
  100. 100. Coste B, Mathur J, Schmidt M, Earley TJ, Ranade S, Petrus MJ, et al. Piezo1 and Piezo2 are essential components of distinct mechanically activated cation channels. Science. 2010;330(6000):55-60
  101. 101. Blumenthal NR, Hermanson O, Heimrich B, Shastri VP. Stochastic nanoroughness modulates neuron–astrocyte interactions and function via mechanosensing cation channels. Proceedings of the National Academy of Sciences. 2014;111(45):16124-16129
  102. 102. Suchyna TM, Tape SE, Koeppe RE 2nd, Andersen OS, Sachs F, Gottlieb PA. Bilayer-dependent inhibition of mechanosensitive channels by neuroactive peptide enantiomers. Nature. 2004;430(6996):235-240
  103. 103. Müller P, Langenbach A, Kaminski A, Rychly J. Modulating the actin cytoskeleton affects mechanically induced signal transduction and differentiation in mesenchymal stem cells. PLoS One. 2013;8(7):e71283
  104. 104. Wechsler A, Teichberg VI. Brain spectrin binding to the NMDA receptor is regulated by phosphorylation, calcium and calmodulin. The EMBO Journal. 1998;17(14):3931-3939
  105. 105. Martinac B. The ion channels to cytoskeleton connection as potential mechanism of mechanosensitivity. Biochimica et Biophysica Acta (BBA) - Biomembranes. 2014;1838(2):682-691
  106. 106. Sekino Y, Kojima N, Shirao T. Role of actin cytoskeleton in dendritic spine morphogenesis. Neurochemistry International. 2007;51(2-4):92-104
  107. 107. Haidekker MA, L’Heureux N, Frangos JA. Fluid shear stress increases membrane fluidity in endothelial cells: A study with DCVJ fluorescence. American Journal of Physiology. Heart and Circulatory Physiology. 2000;278(4):H1401-H14H6
  108. 108. White CR, Frangos JA. The shear stress of it all: The cell membrane and mechanochemical transduction. Philosophical Transactions of the Royal Society, B: Biological Sciences. 2007;362(1484):1459-1467
  109. 109. Bloom M, Evans E, Mouritsen OG. Physical properties of the fluid lipid-bilayer component of cell membranes: A perspective. Quarterly Reviews of Biophysics. 1991;24(03):293-397
  110. 110. Meng F, Suchyna TM, Sachs F. A fluorescence energy transfer-based mechanical stress sensor for specific proteins in situ. The FEBS Journal. 2008;275(12):3072-3087
  111. 111. Grashoff C, Hoffman BD, Brenner MD, Zhou R, Parsons M, Yang MT, et al. Measuring mechanical tension across vinculin reveals regulation of focal adhesion dynamics. Nature. 2010;466(7303):263-266
  112. 112. Meng F, Sachs F. Orientation-based FRET sensor for real-time imaging of cellular forces. Journal of Cell Science. 2012;125(Pt 3):743-750
  113. 113. Guo J, Wang Y, Sachs F, Meng F. Actin stress in cell reprogramming. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(49):E5252-E5261
  114. 114. Maneshi MM, Sachs F, Hua SZ. Heterogeneous cytoskeletal force distribution delineates the onset Ca(2+) influx under fluid shear stress in astrocytes. Frontiers in Cellular Neuroscience. 2018;12:69
  115. 115. Conti F, DeBiasi S, Minelli A, Melone M. Expression of NR1 and NR2A/B subunits of the NMDA receptor in cortical astrocytes. Glia. 1996;17(3):254-258
  116. 116. Cox CD, Bae C, Ziegler L, Hartley S, Nikolova-Krstevski V, Rohde PR, et al. Removal of the mechanoprotective influence of the cytoskeleton reveals PIEZO1 is gated by bilayer tension. Nature Communications. 2016;7:10366
  117. 117. Sachs F. Mechanical transduction by ion channels; a cautionary tale. World Journal of Neurology. 2015;5(3):74-87
  118. 118. Kapus A, Janmey P. Plasma membrane--cortical cytoskeleton interactions: A cell biology approach with biophysical considerations. Comprehensive Physiology. 2013;3(3):1231-1281
  119. 119. Kozlov MM, Chernomordik LV. Membrane tension and membrane fusion. Current Opinion in Structural Biology. 2015;33:61-67
  120. 120. Haidekker MA, Ling T, Anglo M, Stevens HY, Frangos JA, Theodorakis EA. New fluorescent probes for the measurement of cell membrane viscosity. Chemistry & Biology. 2001;8(2):123-131
  121. 121. Markin VS, Sachs F. Free volume in membranes: Viscosity or tension? Open Journal of Biophysics. 2015;05(03):80-83
  122. 122. Helfer E, Harlepp S, Bourdieu L, Robert J, MacKintosh F, Chatenay D. Buckling of actin-coated membranes under application of a local force. Physical Review Letters. 2001;87(8):088103
  123. 123. Kloda A, Lua L, Hall R, Adams DJ, Martinac B. Liposome reconstitution and modulation of recombinant N-methyl-D-aspartate receptor channels by membrane stretch. Proceedings of the National Academy of Sciences. 2007;104(5):1540-1545
  124. 124. Florian JA, Kosky JR, Ainslie K, Pang Z, Dull RO, Tarbell JM. Heparan sulfate proteoglycan is a mechanosensor on endothelial cells. Circulation Research. 2003;93(10):e136-ee42
  125. 125. Akinlaja J, Sachs F. The breakdown of cell membranes by electrical and mechanical stress. Biophysical Journal. 1998;75(1):247-254
  126. 126. Singer S, Nicolson GL. The fluid mosaic model of the structure of cell membranes. Science. 1972;175(4023):720-731
  127. 127. Bagatolli LA, Mouritsen OG. Is the fluid mosaic (and the accompanying raft hypothesis) a suitable model to describe fundamental features of biological membranes? What may be missing? Frontiers in Plant Science. 2013;4:457
  128. 128. Engelman DM. Membranes are more mosaic than fluid. Nature. 2005;438(7068):578-580
  129. 129. Brown MF. Curvature forces in membrane lipid–protein interactions. Biochemistry. 2012;51(49):9782-9795
  130. 130. Haswell ES, Phillips R, Rees DC. Mechanosensitive channels: What can they do and how do they do it? Structure. 2011;19(10):1356-1369

Written By

Mohammad Mehdi Maneshi and Susan Z. Hua

Reviewed: 30 June 2020 Published: 09 September 2020