G Model

ARTICLE IN PRESS

YSCDB 1719 1–8

Seminars in Cell & Developmental Biology xxx (2015) xxx–xxx

Contents lists available at ScienceDirect

Seminars in Cell & Developmental Biology journal homepage: www.elsevier.com/locate/semcdb

1

Review

2

Mechanobiology of lymphatic contractions Lance L. Munn ∗

3

Q1

4

Q2 Department of Radiation Oncology, Massachusetts General Hospital, Boston, MA, United States

5

a r t i c l e

6 19

i n f o

a b s t r a c t

7

Article history: Available online xxx

8 9 10

Keywords: Lymphatic 13 Contraction 14 Mechanosensor 15 Mechanobiology 16 Stretch-activated channel Shear stress 17 18 Q4 Nitric oxide 11 12

The lymphatic system is responsible for controlling tissue fluid pressure by facilitating flow of lymph (i.e. the plasma and cells that enter the lymphatic system). Because lymph contains cells of the immune system, its transport is not only important for fluid homeostasis, but also immune function. Lymph drainage can occur via passive flow or active pumping, and much research has identified the key biochemical and mechanical factors that affect output. Although many studies and reviews have addressed how tissue properties and fluid mechanics (i.e. pressure gradients) affect lymph transport [1–3] there is less known about lymphatic mechanobiology. As opposed to passive mechanical properties, mechanobiology describes the active coupling of mechanical signals and biochemical pathways. Lymphatic vasomotion is the result of a fascinating system affected by mechanical forces exerted by the flowing lymph, including pressure-induced vessel stretch and flow-induced shear stresses. These forces can trigger or modulate biochemical pathways important for controlling the lymphatic contractions. Here, I review the current understanding of lymphatic vessel function, focusing on vessel mechanobiology, and summarize the prospects for a comprehensive understanding that integrates the mechanical and biomechanical control mechanisms in the lymphatic system. © 2015 Published by Elsevier Ltd.

Contents

20

1. 2.

21 22 23

3. 4. 5.

24 25 26 27 28

Lymphatic anatomy and network topology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Lymphatic contractions: calcium dynamics, pacemaker potentials and mechanobiology of stretching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1. Mechanical activation of contractions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Lymphatic relaxation: endothelial derived relaxing factors and fluid shear stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Integrating lymphatic physiology using mathematical models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

00 00 00 00 00 00 00 00

29

1. Lymphatic anatomy and network topology

30

Blood microvasculature is somewhat permeable to water and 32 proteins, and a fraction of the circulating plasma extravasates 33 into tissue wherever there is a transmural pressure driving force. This can establish pressure gradients that extend from the blood 34 vessels to the lymphatic system, allowing for flow of plasma 35 36Q6 through the tissue, into the lymphatics. The process of flushing the extravascular space with plasma is likely useful for conditioning the 37 extracellular matrix and providing biomechanical signals to cells, 38 31Q5

Q3

∗ Tel.: +1 6177264085. E-mail address: [email protected]

but it relies on proper pressure distribution in the tissue: if the pressure in the lymphatic bed is not lower than that in the surrounding tissue – at least transiently – then there is accumulation of pressure and fluid, resulting in edema. The lymphatic system is responsible for maintaining proper tissue-fluid balance [4] and organizing the immune response. Fluid from the tissue first enters blind-ending lymphatic capillaries, termed initial lymphatics (Fig. 1). The initial lymphatic walls are formed by overlapping endothelial cells that act as one-way valves to allow fluid into the vessels, but not out [5,6]. Initial lymphatics pass the fluid to collecting lymphatics, which constitute a system of variable-demand, distributed pumps [7]. Distinct compartments (lymphangions) within each collecting lymphatic vessel are defined by intraluminal luminal one-way valves [8–10]; lymphangions can

http://dx.doi.org/10.1016/j.semcdb.2015.01.010 1084-9521/© 2015 Published by Elsevier Ltd.

Please cite this article in press as: Munn LL. Mechanobiology of lymphatic contractions. Semin Cell Dev Biol (2015), http://dx.doi.org/10.1016/j.semcdb.2015.01.010

39 40 41 42 43 44 45 46 47 48 49 50 51 52

G Model YSCDB 1719 1–8 2

ARTICLE IN PRESS L.L. Munn / Seminars in Cell & Developmental Biology xxx (2015) xxx–xxx

Fig. 1. Lymphatic network topology and anatomy.

53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78

drive flow via contractions of their muscle-invested walls. The collecting lymphatic vessels are arranged in a converging tree-like network, and lymph nodes are distributed throughout, so that lymph passes through them on the way back to the blood circulation [5]. However, active lymphatic pumping is not always operational, and deficient lymphatic pumping is involved in multiple clinical problems, including metabolic disorders [11], local immunocompromise [12,13], and lymphedema (fluid accumulation in tissue). The lymphatic system is also involved in cancer progression, as entry of metastatic cancer cells into the lymphatic system can result in lymph node metastases. Thus, the lymphatic system is central to a variety of pathological processes. A key question is how a connected network of pumps (lymphangions) can coordinate the contractions to move fluid efficiently back to the large veins. In the diverging arterial network of the cardiovascular system, flow is achieved by a single, central pump (the heart), and only relatively small (but important) local adjustments of diameters are able to achieve correct flow distribution to the capillaries. The inverse problem, which exists in the converging lymphatic network, is more difficult. One-way valves must be arranged appropriately, and individual contractions coordinated along vessels and at branch points so that the system does not “fight against itself”. Therefore, an overarching communication system might exist to provide the necessary coordination of the contractions. Although

many studies have delineated the various mechanical and chemical perturbations that affect phasic contractions, there are still questions about their role in organizing the contractions. 2. Lymphatic contractions: calcium dynamics, pacemaker potentials and mechanobiology of stretching To actively move fluid, the lymphangions have to contract, increasing local fluid pressure, which closes upstream – and opens downstream – valve. The contractions are driven by Ca2+ fluxes from extracellular and intracellular stores, which operate through myosin light chain kinase (MLCK) to allow binding of actin with the myosin light chain crossbridges, generating force [14–17]. To control lymphatic pumping on the larger scale, there needs to be synchronization of the calcium dynamics [18]. Coordination of the pumping should lead to a decrease in pressure pulses over the network, and indeed this is observed, especially with high lymph flow states. This implies that the active pumping is organized so that pressure pulses are smoothed as the lymph moves up the lymphatic tree [19]. Conceptually, there are many possible modes in which the contractions might synchronize. For example, in a series of lymphangions, all the odd numbered segments might contract at the same time that the even numbered segments are relaxing (Fig. 2A). This would serve to move the fluid effectively in “bucket brigade” style. Alternatively, two lymphangions might contract together, moving fluid downstream to the next pair of lymphangions,

Please cite this article in press as: Munn LL. Mechanobiology of lymphatic contractions. Semin Cell Dev Biol (2015), http://dx.doi.org/10.1016/j.semcdb.2015.01.010

79 80 81

82 83

84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102

G Model YSCDB 1719 1–8

ARTICLE IN PRESS L.L. Munn / Seminars in Cell & Developmental Biology xxx (2015) xxx–xxx

3

Fig. 2. Examples of lymphangion coordination. (A) Contractions might alternate in adjacent lymphangions. (B) It is also possible that two or more adjacent lymphangions contract and relax together. (C) If there is asymmetric contraction/relaxation in terms of the number of in-phase adjacent lymphangions, then the stroke volumes must not be the same (e.g., lymphangions 1 and 2 will have lower stroke volume than lymphangion 3).

103 104 105 106 107 108 109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147

which are relaxing to accept the fluid (Fig. 2B). This scheme can be extended to even larger numbers of adjacent lymphangions in synchrony, with each ejecting and accepting equivalent volumes of fluid in a given cycle. But other modes are also conceivable. For example, two lymphangions might contract together, while only one of the downstream lymphangions dilates (Fig. 2C). This is possible only if the stroke volume of the relaxing compartment is larger than that of the individual contracting compartments, since it has to accept more fluid than if only a single upstream compartment were contracting. Although highly simplified, these examples illustrate the need for overarching control of the contractions, especially considering the additional complications at converging bifurcations, where the downstream branch accepts fluid from two upstream lymphangions. Thus, it is likely that some mechanism involving local feedback helps control the contractions and relaxations. Smooth muscle cell contractions are dependent on intracellular calcium, and the calcium concentrations are subject to multiple control mechanisms, summarized in Fig. 3. Ca2+ levels are normally low in the cytosol, but can undergo rapid changes due to influx through channels in the plasma membrane and sarcoplasmic reticulum. The contraction process starts as calcium fluctuations due to influx through channels are amplified to become spontaneous transient depolarizations (STDs). Similar to the Na+ /K+ system in nerve action potentials, it requires only a threshold level of Ca2+ in the cytosol to initiate the depolarization and cause a subsequent, larger calcium spike [20–23]. In the lymphatic muscle cells (LMCs), the Ca2+ influx appears to be through IP3 receptors [23], and STDs can be blocked by chelation of cytosolic Ca2+ [22]. These STDs can be modulated by activators or blockers of inositol 1,4,5-trisphosphate receptors (IP3Rs) or calcium-activated chloride channel (CaCC) blockers, so it is thought that Ca2+ release through IP3 receptors opens CaCC channels to generate STDs [23]. This process does not have to be present in all the muscle cells in the vessel wall, but can be driven by a subset of cells that act as pacemakers to initiate the contractions [24,25]. Firing of a pacemaker cell leads to propagation of the Ca2+ wave to adjacent cells, inducing a series of action potential-like spikes of calcium that cause the contractions. This mechanism of STDs is distinct from cardiac-like electrical pacemaker activity, which would require electrical control through nerve action potentials [22,25,26]. Certain lymphatic vessels have nerves in their walls that, when stimulated, induce an increase in pumping contraction frequency: experiments in bovine mesenteric vessels show that contraction is preceded by a single action potential [27], stimulated by the release of both ATP and noradrenaline

[28]. However, it remains to be seen how prevalent lymphatic innervation is or how such action potentials would be coordinated throughout the network. Although it is not clear how the calcium fluctuations progress to become STDs in an individual pacemaker cell, many biochemical pathways may be involved. Neurotransmitters such as noradrenaline, isoproterenol [28,29] and substance P [30,31] and inflammatory mediators such as histamine [13,17,20] have been shown to affect lymphatic contractions by affecting the calcium fluxes. Endothelial derived agents such as endothelin-1 (ET-1) can also enhance lymphatic vasomotion. ET-1 acts through IP3, directly affecting the IP3R-mediated release of Ca2+ [32]. 2.1. Mechanical activation of contractions Lymphatic vessel contractions are also influenced by mechanical signals. In the blood vasculature, resistance vessels respond to increased luminal pressure by constricting, a protective mechanism that prevents damage to the microvessels. The process is controlled by stretch-activation of calcium channels [33,34]. Lymphatic vessels have similar responses to increased pressure [35]. The effects are triggered by physical distension of the vessel wall, as they can be induced either by pressurizing the vessel or by stretching it with a mechanical device. For example, plasma dilution can be used to increase lymph production and the load on the lymphatic system. In response to plasma dilution, rat mesenteric collecting lymphatics respond with increased diastolic diameter, contraction frequency and amplitude, likely caused by intrinsic stretch-dependent mechanisms [19]. Because of the importance of wall stretch in lymphatic physiology, many groups have developed methodologies for studying vessel contractions under controlled conditions of preload and afterload [36–40]. In isolated mesenteric lymphatics, increasing luminal pressure without changing the axial pressure driving force increases both the frequency and force of the individual contractions [37,38] which are preceded by calcium transients [40]. This causes an increase in vessel output. However, at higher pressures, output decreases, likely due to pressure-limited stroke volume [39,41]. This autoregulation of contractions based on transmural pressure can increase lymph drainage in response to pressures caused by tissue edema or positional/gravitational changes. Interestingly, vessels isolated from various locations have different tension vs. length relationships. Applying isometric force to various rat lymphatic vessels (thoracic duct, cervical, and femoral lymph vessels), Gashev and coworkers [42] found similar tension

Please cite this article in press as: Munn LL. Mechanobiology of lymphatic contractions. Semin Cell Dev Biol (2015), http://dx.doi.org/10.1016/j.semcdb.2015.01.010

148 149 150 151 152 153 154 155 156 157 158 159

160

161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190

G Model YSCDB 1719 1–8 4

ARTICLE IN PRESS L.L. Munn / Seminars in Cell & Developmental Biology xxx (2015) xxx–xxx

Fig. 3. Control of calcium kinetics. Calcium enters the cytosol through ion channels (L-type, T-type, voltage gated, stretch-activated) in the plasma membrane or smooth endoplasmic reticulum (SER; SERCA, IP3 R) and acts through myosin light chain kinase (MLCK) to phosphorylate MLC, allowing formation of the myosin–actin crossbridges and cell contraction. Stretch-activated channels can also allow calcium to enter the cell. Calcium-activated chloride channels (CaCCs) can enhance depolarization during STD generation. Endothelial cells produce endothelial derived relaxing factors (EDRFs) such as histamine and NO in response to fluid shear. The EDRFs act on adjacent muscle cells through soluble guanylyl cyclase (sGC), cyclic GMP, protein kinase G (PKG) and myosin light chain phosphatase (MLCP) to reduce intracellular Ca2+ and can dephosphorylate MLC, causing relaxation.

191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225

vs. length curves; all vessels generated near-maximal tensions over a relatively wide range of preloads, but there were differences in the highest pressure each vessel could withstand. This suggests that lymph vessels adapt their contractile force depending on their location in the lymphatic system. More evidence of distension-induced contractions has been documented using servo-controlled wire-myograph devices [43]. These devices allow isometric stretching of the vessel wall and measurement of the resulting contractions, independent of fluid pressure. Similar to increased fluid pressure, application of isometric stretch increases the amplitude and the frequency of phasic activity before declining at higher preloads [44], although the range of outputs differs significantly between the two methods [45]. In addition, the behavior is dependent on the rate of stress application: fast stretching induces bursts of higher frequency contractions [44]. In lymphatic smooth muscle cells, there are two types of voltage-dependent calcium channels (VDCCs) that appear to be important in regulating the phasic contractions: L-type (“longlasting”) and T-type (“transient) channels. It has recently been shown that these channels differentially regulate the strength and frequency of stretch-induced contractions. Nifedipine and diltiazem (which specifically block L-type channels) decreased the strength of contractions, while mibefradil and nickel (which block T-type channels) decreased contraction frequency [46]. Experiments such as described above suggest that the sensitivity of the calcium release mechanism is modulated by mechanical stresses. Our knowledge of mechanically activated channels lags behind that of voltage or ion-gated channels, but a number of candidates have been identified that could be involved in the lymphatic response to mechanical distension. Obvious candidates for mediating the stretch-induced changes in contractions are the stretch-activated ion channels (Fig. 4). Originally identified in nerve cells where they mediate processes such as sensation of touch, pain and hearing, stretch-activated channels

have received more attention recently in vasculature, where they are postulated to transduce fluid mechanical signals. It is known that endothelial [47–49] and smooth muscle cells [50,51] have stretch-activated ion channels that can initiate Ca2+ mobilization. These channels undergo conformational changes in response to membrane deformations. They have been implicated in the control of cell volume, mediating ion exchange as the channels are opened due to membrane stretch [52]. They are also involved in migration of epithelial cells, modulating local calcium fluxes at the cell trailing edge [53]. Stretch-activated channels can be identified and characterized using patch clamp studies in which the membrane is placed under

Fig. 4. Stretch activated calcium channels in the vessel wall change conformation in response to membrane tension, increasing calcium entry into the cell.

Please cite this article in press as: Munn LL. Mechanobiology of lymphatic contractions. Semin Cell Dev Biol (2015), http://dx.doi.org/10.1016/j.semcdb.2015.01.010

226 227 228 229 230 231 232 233 234 235 236 237

G Model YSCDB 1719 1–8

ARTICLE IN PRESS L.L. Munn / Seminars in Cell & Developmental Biology xxx (2015) xxx–xxx

238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262

263 264

265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300

tension by suction and the changes in ion flux measured. Much work on blood vessel endothelial cells [47,49] and smooth muscle cells [50,51] has concluded that calcium channels open more frequently under conditions of membrane stretch. Similar results are seen when endothelial cells are stretched on elastic substrates. Stretch allows Ca2+ entry from outside the cell, which then stimulates release of calcium from internal stores [48]. In endothelial cells, the increase in intracellular Ca2+ with stretch is dependent on Ca2+ in the external media and it can be blocked with gadolinium (Gd3+), which is specific for cation-selective stretch-activated channels [54]. In contrast, it is not affected by nifedipine, which blocks L-type voltage-gated calcium channels [48]. Although the identity of the mechanically responsive calcium channel in lymphatic vessels is not known, recent work has identified Piezo1 as an important stretch-sensitive channel in the endothelial cells of blood vessels. Piezo proteins transduce mechanical forces to cationic currents, and are evolutionarily conserved [55]. They can be blocked with the peptide GsMTx4 as well as gadolinium [56,57]. Also prominent in sensory neurons, Piezo1 proteins exist as subunits of Ca2+ -permeable channels, and are modulated by fluid forces [56]. They play an essential role in vascular development, as Piezo1-deficient endothelial cells have abnormal stress fibers and do not align with shear stress [58]. It remains to be seen whether Piezo proteins play a role in lymphatic contractions in response to stretch.

3. Lymphatic relaxation: endothelial derived relaxing factors and fluid shear stress In addition to the vasoactive agents that enhance lymphatic muscle contractions, there are mechanisms for damping the Ca2+ dynamics. The most well-studied is nitric oxide (NO), a vasodilator that acts at multiple points in the Ca2+ -contraction pathway. NO produced by the endothelium can quickly diffuse to the muscle cells to affect pumping [13,59–67]. It operates by modulating Ca2+ release and uptake [68], as well as the enzymes responsible for force production [69,70]. NO activates cytoplasmic guanylate cyclase in vascular smooth muscle to decrease vascular tone through cGMP-dependent mechanisms. NO decreases intracellular Ca2+ by inhibiting its entry from internal stores. This effects relaxation by decreasing the activity of myosin light chain kinase (Fig. 3). Specifically, NO can act through sGC and PKG to inhibit IP3 receptor channels, limiting Ca2+ influx. At the same time, it can activate SERCA and BKCa channels to increase Ca2+ outflux [59,67]. NO also directly induces relaxation by dismantling the myosin light chain crossbridges through PKG and myosin light chain phosphatase [71–73]. In general, elevated NO leads to dilated lymphatic vessels with decreased contraction frequency, consistent with the role of NO as a vasodilator able to blunt the Ca2+ response. Under normal conditions, the primary source of NO is the endothelium. Blood and lymphatic endothelial cells produce NO in response to fluid flow [63,67,74–77], and removal of the endothelium mimics reagents that block NO production [75]. Flow-induced NO is produced by eNOS in the endothelial cells [78], but during inflammation, NO can be produced by stromal cells via iNOS, independent of the endothelium. The resulting excess of NO can inhibit pumping [13], and iNOS-induced NO is also implicated the dysfunction of aging lymphatics [79]. Interestingly, it appears that NO is not the only endothelialderived relaxing factor (EDRF) important in lymphatic physiology. In isolated vessels, exposure to exogenous NO can reproduce the changes in pumping produced by flow, but blocking NO synthase does not completely abolish the effects of flow [80]. It has been shown that histamine may be another EDRF that compensates

5

when NO is blocked [20]. In rat mesenteric vessels, histamine induces vessel relaxation through the H1 and H2 histamine receptors, and works in conjunction with NO through sGC to decrease tone and contraction frequency [81]. Only by blocking both NO and histamine can the relaxation be blocked [82]. It is appropriate that EDRFs are produced in response to fluid shear forces. They represent another mechanobiological feedback system that, together with stretch-activated calcium channels, can help control the phasic contractions. They provide a counterbalance to the calcium-induced contractions, maintaining vessel dilation in circumstances when low resistance is needed to allow pressuredriven flow [80]. Unfortunately, despite much research activity, relatively little is known about endothelial mechanobiology and the structures in the cells that transduce shear stress signals. Although channels such as Piezo1 might be involved, other structures such as adhesion molecules [83] and glycocalyx components [84,85] have been implicated in vessel mechanobiology. Furthermore, it is likely that cells in the lymphatic wall use distinct pathways to discriminate fluid shear stress and pressure-induced stretch, given the distinct responses to these stimuli. Thus, more research is needed to identify and characterize the mechanosensors involved in lymphatic physiology.

4. Integrating lymphatic physiology using mathematical models The complex behavior of the lymphatic system has led many groups to develop mathematical models to characterize and study its performance analytically. Some models have increased our understanding of the trade-off between active pumping and flow resistance and have explored various putative modes of pumping. Others have focused on basic properties of interstitial transport and lymphatic uptake. Lymph flow is not only driven by pumping collecting lymphatics, but also by pressure imbalances in the tissue. Swartz et al. used data from mouse tail lymphatics to numerically model the pressures and hydraulic conductivities that drive fluid into the lymphatics. They noted that the pressure driving forces are sensitive to tissue elasticity, which can decrease as tissue remodels during chronic edema [86]. More recent analyses have addressed the issue of initial lymphatic network topology, and how the shape of the network affects lymph drainage. Using homogenization theory, it is possible to show that most of the resistance to lymph flow occurs at the entrance to the initial lymphatics, and that the observed hexagonal arrangement of lymphatics in the mouse tail provides the optimum drainage in this tissue [87]. Some modeling efforts have addressed lymphatic pumping efficiency, valve function or vessel mechanics to quantify lymphatic performance. One category of models uses measured physical and functional properties of lymphatic vessels to simulate pumping. These models characterize the resulting performance of a single lymphatic vessel, represented by a number of contracting lymphangions in series. In general, input functions are imposed to drive flow or to cycle the contractions of individual lymphangions, and the output of the simulated vessel is calculated [88–90]. Such models can also be used to predict the conditions under which the system fails or to determine which parameters most strongly affect flow. For example, excessive back-pressure can prevent flow by causing valve leakage [91], and there is an optimum transmural pressure and number of lymphangion segments for a given length of vessel [92]. These models highlight the dual role of lymphatic vessels as conduits as well as pumps, and have shown that, in addition to contraction amplitude, the overall flow resistance through the vessel is a critical determinant of output [92]. Thus, the observed inhibition of contractions (which tend to increase flow resistance) by fluid flow is advantageous under conditions where

Please cite this article in press as: Munn LL. Mechanobiology of lymphatic contractions. Semin Cell Dev Biol (2015), http://dx.doi.org/10.1016/j.semcdb.2015.01.010

301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322

323 324

325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363

G Model YSCDB 1719 1–8

ARTICLE IN PRESS L.L. Munn / Seminars in Cell & Developmental Biology xxx (2015) xxx–xxx

6

390

pressure-driven flow is possible [93]. Key parameters in these models are the pressure-diameter relationship [94], the calcium force production (often a function of diameter) and the measured period of the contractions. When input into a fluid dynamics model, these can predict the lymph flow produced in a collecting lymphatic, and compare well with experimental results [95]. A more recent focus of experimental and modeling efforts has been the performance of the intraluminal valves during a contraction cycle. Rather than perfect one-way check valves, these structures allow some back flow, which appears to be dependent on many factors, including vessel diameter and axial pressure. Mathematical models that incorporate these subtleties of valve efficiency have been able to more accurately predict in vivo results [89]. As we have seen, lymphatic vessels can adjust their vasomotion to maintain fluid homeostasis. It is thought that they minimize contractions if existing fluid pressure gradients can drive flow, but actively pump to drive fluid otherwise [96]. This variable behavior has made it difficult to converge on a comprehensive, conceptual model of lymphatic function. Given that fluid shear stress and mechanical stretch play central roles in lymphatic function, it is likely that these vessels integrate physical signals to control lymph drainage. Indeed, in any control system, direct feedback between the quantity being controlled (fluid pressure and flow rate in this case) and the operation (lymphatic pumping) is needed. Furthermore, this control has to operate locally, but coordinate pumping over larger lengths of vessels. The nature of the feedback signals and the intramural communication are still areas of active investigation.

391

5. Conclusions

364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389

392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427

Mechanobiological control mechanisms can potentially explain the robust transport functions observed in the lymphatic system. They allow lymphatics to adapt to changes in the fluid environment to optimize lymph transport, automatically reacting to changes in tissue fluid pressures. It is interesting that many of the observed mechanism have analogs in the blood vasculature, but serve slightly different purposes in the lymphatic system. For example, EDRFs can relax both blood and lymphatic vessels, and they also react similarly to changes in pressure. Excess pressure in blood vessels can induce rapid constriction to limit flow to the microvasculature, and the same stretch-induced constriction appears to be one of the mechanical drivers of lymphatic phasic contractions. The periodic vasomotion of blood vessels is perhaps most closely related to lymphatic phasic contractions, but its purpose is not well-understood. Although not necessary for pumping blood, it may be involved in shuttling flow between capillary beds either randomly or in response to metabolic or thermal stimuli. The magnitudes of the fluid pressures and shear stresses vary significantly in the blood and lymphatic systems. For vessel stretch, the relevant quantity is the transwall pressure, which ranges from ∼150 mmHg in the aorta to ∼5–10 mmHg in the blood microvessels and collecting lymphatics [97,98]. However, mechanobiological mechanisms underlying the stretch response are sensitive to wall strain rather than fluid pressure. Furthermore, vessels with higher transwall pressures have more muscular walls, limiting the pressure-induced strain. Thus, it is possible that smooth muscle strains are similar throughout the blood and lymphatic systems. Shear stresses also vary somewhat throughout the blood vasculature, ranging from ∼10 to 20 dyn/cm2 in the large arteries and capillaries but increasing to as high as 100 dyn/cm2 in arterioles [99,100]. Lymphatic vessels lie at the low end of this range. Average shear rates in collecting lymphatic vessels have been reported as 0.64 ± 0.14 dyn/cm2 , with peaks of 4–12 dyn/cm2 [101], but can increase to 40 dyn/cm2 during edemagenic stress [102]. The large range of shear stresses in the blood circulation and the cyclic shear stress experienced by lymphatic endothelium during lymphatic

pumping raise the question of whether there is a floating threshold tuned to different levels in different vessels to determine EDRF production. It has also been proposed that it is not the absolute level of shear stress that determines endothelial mechanobiology, but the temporal changes in shear stress [103]. Thus, lymphatic endothelium may be more responsive by virtue of the dynamic shear stress during phasic contractions. The reliance of lymphatic pumping on mechanosensors provides additional opportunities for pharmacological intervention in lymphatic disorders. More research is needed to identify the mechanosensors that respond to stretch and shear stress in lymphatics so that drugs can be developed that specifically enhance or block these pathways. The hope would be to find targets that are specific to lymphatic mechanobiology with minimal toxicity to blood vessels and other tissues. Acknowledgment This work was funded in part by National Institutes of Health Q7 Grant NIH R01CA149285. References [1] Breslin JW. Mechanical forces and lymphatic transport. Microvasc Res 2014;96C:46–54. [2] Nipper ME, Dixon JB. Engineering the lymphatic system. Cardiovasc Eng Technol 2011;2:296–308. [3] Negrini D, Moriondo A. Lymphatic anatomy and biomechanics. J Physiol 2011;589:2927–34. [4] Thomas SN, Rutkowski JM, Pasquier M, Kuan EL, Alitalo K, Randolph GJ, et al. Impaired humoral immunity and tolerance in K14-VEGFR-3-Ig mice that lack dermal lymphatic drainage. J Immunol 2012;189:2181–90. [5] Mendoza E, Schmid-Schonbein GW. A model for mechanics of primary lymphatic valves. J Biomech Eng 2003;125:407–14. [6] Schmid-Schönbein GW. Microlymphatics and lymph flow. Physiol Rev 1990;70:987. [7] Roddie IC, Mawhinney HJ, McHale NG, Kirkpatrick CT, Thornbury K. Lymphatic motility. Lymphology 1980;13:166–72. [8] Gashev AA, Zawieja DC. Physiology of human lymphatic contractility: a historical perspective. Lymphology 2001;34:124–34. [9] Bazigou E, Wilson JT, Moore Jr JE. Primary and secondary lymphatic valve development: molecular, functional and mechanical insights. Microvasc Res Q8 2014. [10] Vittet D. Lymphatic collecting vessel maturation and valve morphogenesis. Microvasc Res 2014. [11] Zawieja SD, Wang W, Wu X, Nepiyushchikh ZV, Zawieja DC, Muthuchamy M. Impairments in the intrinsic contractility of mesenteric collecting lymphatics in a rat model of metabolic syndrome. Am J Physiol Heart Circ Physiol 2012;302:H643–53. [12] Elias RM, Johnston MG, Hayashi A, Nelson W. Decreased lymphatic pumping after intravenous endotoxin administration in sheep. Am J Physiol 1987;253:H1349–57. [13] Liao S, Cheng G, Conner DA, Huanga Y, Kucherlapati RS, Munn LL, et al. Impaired lymphatic contraction associated with immunosuppression. PNAS 2011;108:18784. [14] Ledvora RF, Barany M, Barany K. Myosin light chain phosphorylation and tension development in stretch-activated arterial smooth muscle. Clin Chem 1984;30:2063–8. [15] Nepiyushchikh ZV, Chakraborty S, Wang W, Davis MJ, Zawieja DC, Muthuchamy M. Differential effects of myosin light chain kinase inhibition on contractility, force development and myosin light chain 20 phosphorylation of rat cervical and thoracic duct lymphatics. J Physiol 2011;589:5415–29. [16] Wang W, Nepiyushchikh Z, Zawieja DC, Chakraborty S, Zawieja SD, Gashev AA, et al. Inhibition of myosin light chain phosphorylation decreases rat mesenteric lymphatic contractile activity. Am J Physiol Heart Circ Physiol 2009;297:H726–34. [17] Von Der Weid PY. Review article: lymphatic vessel pumping and inflammation – the role of spontaneous constrictions and underlying electrical pacemaker potentials. Aliment Pharmacol Ther 2001;15:1115–29. [18] McHale NG, Meharg MK. Co-ordination of pumping in isolated bovine lymphatic vessels. J Physiol 1992;450:503–12. [19] Benoit JN, Zawieja DC, Goodman AH, Granger HJ. Characterization of intact mesenteric lymphatic pump and its responsiveness to acute edemagenic stress. Am J Physiol 1989;257:H2059–69. [20] Fox JL, von der Weid PY. Effects of histamine on the contractile and electrical activity in isolated lymphatic vessels of the guinea-pig mesentery. Br J Pharmacol 2002;136:1210–8. [21] McHale NG, Allen JM, Iggulden HL. Mechanism of alpha-adrenergic excitation in bovine lymphatic smooth muscle. Am J Physiol 1987;252:H873–8.

Please cite this article in press as: Munn LL. Mechanobiology of lymphatic contractions. Semin Cell Dev Biol (2015), http://dx.doi.org/10.1016/j.semcdb.2015.01.010

428 429 430 431 432 433 434 435 436 437 438 439 440 441 442

443

444 445

446

447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502

G Model YSCDB 1719 1–8

ARTICLE IN PRESS L.L. Munn / Seminars in Cell & Developmental Biology xxx (2015) xxx–xxx

503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588

[22] Van Helden DF. Pacemaker potentials in lymphatic smooth muscle of the guinea-pig mesentery. J Physiol 1993;471:465–79. [23] von der Weid PY, Rahman M, Imtiaz MS, van Helden DF. Spontaneous transient depolarizations in lymphatic vessels of the guinea pig mesentery: pharmacology and implication for spontaneous contractility. Am J Physiol Heart Circ Physiol 2008;295:H1989–2000. [24] Akl TJ, Nepiyushchikh ZV, Gashev AA, Zawieja DC, Cot GL. Measuring contraction propagation and localizing pacemaker cells using high speed video microscopy. J Biomed Opt 2011;16:026016. [25] McCloskey KD, Hollywood MA, Thornbury KD, Ward SM, McHale NG. Kit-like immunopositive cells in sheep mesenteric lymphatic vessels. Cell Tissue Res 2002;310:77–84. [26] Van Helden DF, Zhao J. Lymphatic vasomotion. Clin Exp Pharmacol Physiol 2000;27:1014–8. [27] McHale NG. Lymphatic innervation. Blood Vessels 1990;27:127–36. [28] Hollywood MA, McHale NG. Mediation of excitatory neurotransmission by the release of ATP and noradrenaline in sheep mesenteric lymphatic vessels. J Physiol 1994;481(Pt 2):415–23. [29] Von der Weid PY, Van Helden DF. Beta-adrenoceptor-mediated hyperpolarization in lymphatic smooth muscle of guinea pig mesentery. Am J Physiol 1996;270:H1687–95. [30] Davis MJ, Lane MM, Davis AM, Durtschi D, Zawieja DC, Muthuchamy M, et al. Modulation of lymphatic muscle contractility by the neuropeptide substance P. Am J Physiol Heart Circ Physiol 2008;295:H587–97. [31] Chakraborty S, Nepiyushchikh Z, Davis MJ, Zawieja DC, Muthuchamy M. Substance P activates both contractile and inflammatory pathways in lymphatics through the neurokinin receptors NK1R and NK3R. Microcirculation 2011;18:24–35. [32] Zhao J, van Helden DF. ET-1-associated vasomotion and vasospasm in lymphatic vessels of the guinea-pig mesentery. Br J Pharmacol 2003;140:1399–413. [33] Setoguchi M, Ohya Y, Abe I, Fujishima M. Stretch-activated whole-cell currents in smooth muscle cells from mesenteric resistance artery of guineapig. J Physiol 1997;501(Pt 2):343–53. [34] Gustafsson D, Grande PO, Borgstrom P, Lindberg L. Effects of calcium antagonists on myogenic and neurogenic control of resistance and capacitance vessels in cat skeletal muscle. J Cardiovasc Pharmacol 1988;12:413–22. [35] Scallan JP, Wolpers JH, Muthuchamy M, Zawieja DC, Gashev AA, Davis MJ. Independent and interactive effects of preload and afterload on the pump function of the isolated lymphangion. Am J Physiol Heart Circ Physiol 2012;303:H809–24. [36] Kornuta JA, Brandon Dixon J. Ex vivo lymphatic perfusion system for independently controlling pressure gradient and transmural pressure in isolated vessels. Ann Biomed Eng 2014;42:1691–704. [37] McHale NG, Roddie IC. The effect of transmural pressure on pumping activity in isolated bovine lymphatic vessels. J Physiol 1976;261:255–69. [38] Davis MJ, Scallan JP, Wolpers JH, Muthuchamy M, Gashev AA, Zawieja DC. Intrinsic increase in lymphangion muscle contractility in response to elevated afterload. Am J Physiol Heart Circ Physiol 2012;303:H795–808. [39] Johnston MG, Elias R. The regulation of lymphatic pumping. Lymphology 1987;20:215–8. [40] Shirasawa Y, Benoit JN. Stretch-induced calcium sensitization of rat lymphatic smooth muscle. Am J Physiol Heart Circ Physiol 2003;285:H2573–7. [41] McGeown JG, McHale NG, Roddie IC, Thornbury K. Peripheral lymphatic responses to outflow pressure in anaesthetized sheep. J Physiol 1987;383:527–36. [42] Gashev AA, Zhang RZ, Muthuchamy M, Zawieja DC, Davis MJ. Regional heterogeneity of length–tension relationships in rat lymph vessels. Lymphat Res Biol 2012;10:14–9. [43] Davis MJ, Lane MM, Scallan JP, Gashev AA, Zawieja DC. An automated method to control preload by compensation for stress relaxation in spontaneously contracting, isometric rat mesenteric lymphatics. Microcirculation 2007;14:603–12. [44] Davis MJ, Davis AM, Lane MM, Ku CW, Gashev AA. Rate-sensitive contractile responses of lymphatic vessels to circumferential stretch. J Physiol 2009;587:165–82. [45] Zhang R, Gashev AA, Zawieja DC, Lane MM, Davis MJ. Length-dependence of lymphatic phasic contractile activity under isometric and isobaric conditions. Microcirculation 2007;14:613–25. [46] Lee S, Roizes S, von der Weid PY. Distinct roles of L- and T-type voltagedependent Ca2+ channels in regulation of lymphatic vessel contractile activity. J Physiol 2014;592:5409–27. [47] Lansman JB, Hallam TJ, Rink TJ. Single stretch-activated ion channels in vascular endothelial cells as mechanotransducers. Nature 1987;325:811–3. [48] Naruse K, Sokabe M. Involvement of stretch-activated ion channels in Ca2+ mobilization to mechanical stretch in endothelial cells. Am J Physiol 1993;264:C1037–44. [49] Ito S, Suki B, Kume H, Numaguchi Y, Ishii M, Iwaki M, et al. Actin cytoskeleton regulates stretch-activated Ca2+ influx in human pulmonary microvascular endothelial cells. Am J Respir Cell Mol Biol 2010;43:26–34. [50] Davis MJ, Donovitz JA, Hood JD. Stretch-activated single-channel and whole cell currents in vascular smooth muscle cells. Am J Physiol 1992;262:C1083–8. [51] Ducret T, El Arrouchi J, Courtois A, Quignard JF, Marthan R, Savineau JP. Stretch-activated channels in pulmonary arterial smooth muscle cells from normoxic and chronically hypoxic rats. Cell Calcium 2010;48:251–9.

7

[52] Christensen O. Mediation of cell volume regulation by Ca2+ influx through stretch-activated channels. Nature 1987;330:66–8. [53] Lee J, Ishihara A, Oxford G, Johnson B, Jacobson K. Regulation of cell movement is mediated by stretch-activated calcium channels. Nature 1999;400:382–6. [54] Yang XC, Sachs F. Block of stretch-activated ion channels in Xenopus oocytes by gadolinium and calcium ions. Science 1989;243:1068–71. [55] Coste B, Xiao B, Santos JS, Syeda R, Grandl J, Spencer KS, et al. Piezo proteins are pore-forming subunits of mechanically activated channels. Nature 2012;483:176–81. [56] Li J, Hou B, Tumova S, Muraki K, Bruns A, Ludlow MJ, et al. Piezo1 integration of vascular architecture with physiological force. Nature 2014. [57] Bae C, Sachs F, Gottlieb PA. The mechanosensitive ion channel Piezo1 is inhibited by the peptide GsMTx4. Biochemistry 2011;50:6295–300. [58] Ranade SS, Qiu Z, Woo SH, Hur SS, Murthy SE, Cahalan SM, et al. Piezo1, a mechanically activated ion channel, is required for vascular development in mice. Proc Natl Acad Sci U S A 2014;111:10347–52. [59] Mizuno R, Koller A, Kaley G. Regulation of the vasomotor activity of lymph microvessels by nitric oxide and prostaglandins. AJP Regul Integr Comp Physiol 1998;274:R790–6. [60] Shirasawa Y, Ikomi F, Ohhashi T. Physiological roles of endogenous nitric oxide in lymphatic pump activity of rat mesentery in vivo. AJP Gastrointest Liver Physiol 2000;278:G551–6. [61] von der Weid PY, Zhao J, Van Helden DF. Nitric oxide decreases pacemaker activity in lymphatic vessels of guinea pig mesentery. AJP Heart Circ Physiol 2001;280:H2707–16. [62] Bohlen HG, Wang W, Gashev A, Gasheva O, Zawieja D. Phasic contractions of rat mesenteric lymphatics increase basal and phasic nitric oxide generation in vivo. AJP Heart Circ Physiol 2009;297:H1319–28. [63] Bohlen HG, Gasheva OY, Zawieja DC. Nitric oxide formation by lymphatic bulb and valves is a major regulatory component of lymphatic pumping. Am J Physiol Heart Circ Physiol 2011;301:H1897–906. [64] Scallan JP, Davis MJ. Genetic removal of basal nitric oxide enhances contractile activity in isolated murine collecting lymphatic vessels. J Physiol 2013;591:2139–56. [65] Gasheva OY, Zawieja DC, Gashev AA. Contraction-initiated NO-dependent lymphatic relaxation: a self-regulatory mechanism in rat thoracic duct. J Physiol 2006;575:821. [66] Koller A, Mizuno R, Kaley G. Flow reduces the amplitude and increases the frequency of lymphatic vasomotion: role of endothelial prostanoids. AJP Regul Integr Comp Physiol 1999;277:R1683–9. [67] von der Weid PY, Crowe MJ, Van Helden DF. Endothelium-dependent modulation of pacemaking in lymphatic vessels of the guinea-pig mesentery. J Physiol 1996;493(Pt 2):563–75. [68] Cohen RA, Weisbrod RM, Gericke M, Yaghoubi M, Bierl C, Bolotina VM. Mechanism of nitric oxide-induced vasodilatation: refilling of intracellular stores by sarcoplasmic reticulum Ca2+ ATPase and inhibition of store-operated Ca2+ influx. Circ Res 1999;84:210–9. [69] Barouch LA, Harrison RW, Skaf MW, Rosas GO, Cappola TP, Kobeissi ZA, et al. Nitric oxide regulates the heart by spatial confinement of nitric oxide synthase isoforms. Nature 2002;416:337–9. [70] Dimmeler S, Fleming I, Fisslthaler B, Hermann C, Busse R, Zeiher AM. Activation of nitric oxide synthase in endothelial cells by Akt-dependent phosphorylation. Nature 1999;399:601–5. [71] Bolz SS, Vogel L, Sollinger D, Derwand R, de Wit C, Loirand G, et al. Nitric oxideinduced decrease in calcium sensitivity of resistance arteries is attributable to activation of the myosin light chain phosphatase and antagonized by the RhoA/Rho kinase pathway. Circulation 2003;107:3081–7. [72] Murphy RA, Walker JS. Inhibitory mechanisms for cross-bridge cycling: the nitric oxide–cGMP signal transduction pathway in smooth muscle relaxation. Acta Physiol Scand 1998;164:373–80. [73] Blatter LA, Wier WG. Nitric oxide decreases [Ca2+]i in vascular smooth muscle by inhibition of the calcium current. Cell Calcium 1994;15:122–31. [74] Shyy JYJ, Chien S. Role of integrins in endothelial mechanosensing of shear stress. Circ Res 2002;91:769–75. [75] Ferrusi I, Zhao J, van Helden D, von der Weid PY. Cyclopiazonic acid decreases spontaneous transient depolarizations in guinea pig mesenteric lymphatic vessels in endothelium-dependent and -independent manners. Am J Physiol Heart Circ Physiol 2004;286:H2287–95. [76] Kuchan MJ, Jo H, Frangos JA. Role of G proteins in shear stress-mediated nitric oxide production by endothelial cells. AJP Cell Physiol 1994;267:C753–8. [77] Kuchan M, Frangos J. Role of calcium and calmodulin in flow-induced nitric oxide production in endothelial cells. AJP Cell Physiol 1994;266:C628–36. [78] Gasheva OY, Knippa K, Nepiushchikh ZV, Muthuchamy M, Gashev AA. Agerelated alterations of active pumping mechanisms in rat thoracic duct. Microcirculation 2007;14:827–39. [79] Gashev AA, Zawieja DC. Hydrodynamic regulation of lymphatic transport and the impact of aging. Pathophysiology 2010;17:277–87. [80] Gashev AA, Davis MJ, Zawieja DC. Inhibition of the active lymph pump by flow in rat mesenteric lymphatics and thoracic duct. J Physiol 2002;540: 1023–37. [81] Kurtz KH, Moor AN, Souza-Smith FM, Breslin JW. Involvement of h1 and h2 receptors and soluble guanylate cyclase in histamine-induced relaxation of rat mesenteric collecting lymphatics. Microcirculation 2014;21:593–605. [82] Nizamutdinova IT, Maejima D, Nagai T, Bridenbaugh E, Thangaswamy S, Chatterjee V, et al. Involvement of histamine in endothelium-dependent relaxation of mesenteric lymphatic vessels. Microcirculation 2014;21:640–8.

Please cite this article in press as: Munn LL. Mechanobiology of lymphatic contractions. Semin Cell Dev Biol (2015), http://dx.doi.org/10.1016/j.semcdb.2015.01.010

589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674

G Model YSCDB 1719 1–8 8 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702

ARTICLE IN PRESS L.L. Munn / Seminars in Cell & Developmental Biology xxx (2015) xxx–xxx

[83] Ross TD, Coon BG, Yun S, Baeyens N, Tanaka K, Ouyang M, et al. Integrins in mechanotransduction. Curr Opin Cell Biol 2013;25:613–8. [84] Qazi H, Palomino R, Shi ZD, Munn LL, Tarbell JM. Cancer cell glycocalyx mediates mechanotransduction and flow-regulated invasion. Integr Biol 2013;5:1334–43. [85] Tarbell JM, Simon SI, Curry FR. Mechanosensing at the vascular interface. Annu Rev Biomed Eng 2014;16:505–32. [86] Swartz MA, Kaipainen A, Netti PA, Brekken C, Boucher Y, Grodzinsky AJ, et al. Mechanics of interstitial-lymphatic fluid transport: theoretical foundation and experimental validation. J Biomech 1999;32:1297–307. [87] Roose T, Swartz MA. Multiscale modeling of lymphatic drainage from tissues using homogenization theory. J Biomech 2012;45:107–15. [88] Macdonald AJ, Arkill KP, Tabor GR, McHale NG, Winlove CP. Modeling flow in collecting lymphatic vessels: one-dimensional flow through a series of contractile elements. Am J Physiol Heart Circ Physiol 2008;295:H305–13. [89] Bertram CD, Macaskill C, Moore Jr JE. Incorporating measured valve properties into a numerical model of a lymphatic vessel. Comput Methods Biomech Biomed Eng 2013. [90] Rahbar E, Moore Jr JE. A model of a radially expanding and contracting lymphangion. J Biomech 2011;44:1001–7. [91] Bertram CD, Macaskill C, Moore Jr JE. Simulation of a chain of collapsible contracting lymphangions with progressive valve closure. J Biomech Eng 2011;133:011008. [92] Jamalian S, Bertram CD, Richardson WJ, Moore Jr JE. Parameter sensitivity analysis of a lumped-parameter model of a chain of lymphangions in series. Am J Physiol Heart Circ Physiol 2013;305:H1709–17. [93] Quick CM, Venugopal AM, Gashev AA, Zawieja DC, Stewart RH. Intrinsic pump-conduit behavior of lymphangions. Am J Physiol Regul Integr Comp Physiol 2007;292:R1510–8.

[94] Rahbar E, Weimer J, Gibbs H, Yeh AT, Bertram CD, Davis MJ, et al. Passive pressure-diameter relationship and structural composition of rat mesenteric lymphangions. Lymphat Res Biol 2012;10:152–63. [95] Bertram CD, Macaskill C, Davis MJ, Moore Jr JE. Development of a model of a multi-lymphangion lymphatic vessel incorporating realistic and measured parameter values. Biomech Model Mechanobiol 2013. [96] Gashev AA. Physiologic aspects of lymphatic contractile function: current perspectives. Ann N Y Acad Sci 2002;979:178–87, discussion 88–96. [97] Boucher Y, Jain RK. Microvascular pressure is the principal driving force for interstitial hypertension in solid tumors: implications for vascular collapse. Cancer Res 1992;52:5110–4. [98] Bouta EM, Wood RW, Brown EB, Rahimi H, Ritchlin CT, Schwarz EM. In vivo quantification of lymph viscosity and pressure in lymphatic vessels and draining lymph nodes of arthritic joints in mice. J Physiol 2014;592: 1213–23. [99] Pries AR, Secomb TW, Gaehtgens P. Design principles of vascular beds. Circ Res 1995;77:1017–23. [100] Cheng C, Helderman F, Tempel D, Segers D, Hierck B, Poelmann R, et al. Large variations in absolute wall shear stress levels within one species and between species. Atherosclerosis 2007;195:225–35. [101] Dixon JB, Greiner ST, Gashev AA, Cote GL, Moore JE, Zawieja DC. Lymph flow, shear stress, and lymphocyte velocity in rat mesenteric prenodal lymphatics. Microcirculation 2006;13:597–610. [102] Rahbar E, Akl T, Cote GL, Moore Jr JE, Zawieja DC. Lymph transport in rat mesenteric lymphatics experiencing edemagenic stress. Microcirculation 2014. [103] Bao X, Lu C, Frangos JA. Temporal gradient in shear but not steady shear stress induces PDGF-A and MCP-1 expression in endothelial cells: role of NO NF kappa B, and egr-1. Arterioscler Thromb Vasc Biol 1999;19:996–1003.

Please cite this article in press as: Munn LL. Mechanobiology of lymphatic contractions. Semin Cell Dev Biol (2015), http://dx.doi.org/10.1016/j.semcdb.2015.01.010

703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731

Mechanobiology of lymphatic contractions.

The lymphatic system is responsible for controlling tissue fluid pressure by facilitating flow of lymph (i.e. the plasma and cells that enter the lymp...
1MB Sizes 0 Downloads 6 Views