Structural and Biochemical Consequences of Disease-Causing Mutations in the Ankyrin Repeat Domain of the Human TRPV4 Channel
Abstract

The TRPV4 calcium-permeable cation channel plays important physiological roles in osmosensation, mechanosensation, cell barrier formation, and bone homeostasis. Recent studies reported that mutations in TRPV4, including some in its ankyrin repeat domain (ARD), are associated with human inherited diseases, including neuropathies and skeletal dysplasias, probably because of the increased constitutive activity of the channel. TRPV4 activity is regulated by the binding of calmodulin and small molecules such as ATP to the ARD at its cytoplasmic N-terminus. We determined structures of ATP-free and -bound forms of human TRPV4-ARD and compared them with available TRPV-ARD structures. The third inter-repeat loop region (Finger 3 loop) is flexible and may act as a switch to regulate channel activity. Comparisons of TRPV-ARD structures also suggest an evolutionary link between ARD structure and ATP binding ability. Thermal stability analyses and molecular dynamics simulations suggest that ATP increases stability in TRPV-ARDs that can bind ATP. Biochemical analyses of a large panel of TRPV4-ARD mutations associated with human inherited diseases showed that some impaired thermal stability while others weakened ATP binding ability, suggesting molecular mechanisms for the diseases.
Funding Statement
This work was funded by National Institutes of Health Grant R01GM081340 to R.G. This work is based upon research conducted at the Advanced Photon Source on the Northeastern Collaborative Access Team beamlines supported by Grant RR-15301 from the National Center for Research Resources. The Advanced Photon Source is supported by the U.S. Department of Energy, Office of Basic Energy Sciences, under Contract DE-AC02-06CH11357. M.S. was a Howard Hughes Medical Institute Fellow of the Helen Hay Whitney Foundation.
‡ Author Present Address
Howard Hughes Medical Institute and Department of Biochemistry, University of Washington, Seattle, WA 98195.
Experimental Procedures
Expression Constructs
Protein Production and Purification
Crystallization of Human TRPV4-ARD
Data Collection, Structure Determination, and Analysis
crystal form I (ATP-free) | crystal form II (with ATP) | |
---|---|---|
Data Collection | ||
space group | P3121 | P3221 |
wavelength (Å) | 0.97917 | 0.97917 |
cell dimensions a, b, c (Å) | 53.30, 53.37, 440.71 | 147.89, 147.89, 93.90 |
resolution (Å)a | 40.0–2.85(2.90–2.85) | 40.0–2.95(3.00–2.95) |
Rsyma | 0.115 (0.614) | 0.099 (0.714) |
I/σ(I)a | 12.3 (1.8) | 17.6 (2.4) |
completeness (%)a | 99.6 (99.3) | 100.0 (100.0) |
redundancya | 5.3. (5.4) | 5.5 (5.6) |
Refinement | ||
resolution (Å)a | 39.2–2.85(2.92–2.85) | 38.86–2.95(3.03–2.95) |
no. of reflections | 17306 | 23958 |
Rwork/Rfree | 0.215/0.278 | 0.177/0.220 |
no. of molecules per asymmetric unit | 2 | 2 |
no. of residues in model | 148–397(chain A) | 148–394 (ATP-bound) |
148–396(chain B) | 148–392 (ATP-unbound) | |
no. of atoms | ||
protein | 1995 (chain A) | 1987 (ATP-bound) |
1986 (chain B) | 1963 (ATP-unbound) | |
ligand | 25 (PO4) | 31 (ATP) |
6 (glycerol) | 6 (glycerol) | |
12 (glucose) | ||
water | 18 | 72 |
B factor (Å) | ||
protein | 80.91 | 62.12 (ATP-bound) |
72.76 (ATP-unbound) | ||
ligand | 113.98 (PO4) | 91.00 (ATP) |
103.44 (glycerol) | 72.51 (glycerol) | |
90.86 (glucose) | ||
water | 76.06 | 54.84 |
rmsd | ||
bond lengths (Å) | 0.013 | 0.013 |
bond angles (deg) | 1.51 | 1.80 |
Values from the highest-resolution shell are in parentheses.
Thermal Stability Assay by Circular Dichroism Spectroscopy
Molecular Dynamics Simulation of rTRPV1-ARD
ATP-Agarose Pull-Down Assays
TRPV1 Cysteine Modification Assay
Statistical Analyses
Results
Structures of Human TRPV4-ARD
Figure 1

Figure 1. Structural comparison of human and chicken TRPV4-ARDs. (A) Superimposed ribbon diagrams of ATP-bound (magenta) and ATP-free (blue) hTRPV4-ARD. ATP is shown as sticks. (B) Superimposed Cα traces of human and chicken TRPV4-ARD. Finger 3 is twisted and shrunken in the ATP-bound (magenta) and ATP-unbound (green) hTRPV4-ARD structures, while the finger is extended in ATP-free hTRPV4-ARD (blue) and cTRPV4-ARD (gray). Several Finger 3 residues are disordered in three of six TRPV4-ARD structures. The structure of Finger 2 in the ATP-bound and -unbound forms differs from that in ATP-free forms.
Structural Consequences of ATP Binding
Figure 2

Figure 2. Aromatic residues on Fingers 2 and 3 have varied positions in hTRPV4-ARD structures. (A) The hTRPV4-ARD structures of ATP-bound (magenta), ATP-unbound (green), and ATP-free (blue) forms are superimposed. Aromatic residues are shown as sticks. (B) Detail of the Finger 2 and 3 loops. F272 and F273 on Finger 3 (black rectangles) are embedded in the aromatic cluster in the ATP-bound and -unbound forms but exposed in the ATP-free form. Y235 and Y236 on Finger 2 and Y281 and F282 on Finger 3 are located in similar positions but show variable orientations. F231, F282, Y283, and F284 show less variation.
Structural Comparison of TRPV-ARDs
Figure 3

Figure 3. Structural comparison of TRPV-ARDs. (A) Superimposed main chain structure of TRPV-ARDs (rTRPV1-ARD, gray; rat and human TRPV2-ARD, cyan; ATP-bound hTRPV4-ARD, magenta; ATP-unbound hTRPV4-ARD, green; ATP-free hTRPV4-ARD, blue; and mouse TRPV6-ARD, black). Finger 3 and a part of Finger 2 are highly flexible. Several residues on Finger 3 are missing in one of two TRPV1-ARD structures and four of seven TRPV2-ARD structures. (B) ATP-binding site of hTRPV4-ARD and rTRPV1-ARD. Residues (sticks) within 4 Å of the ATP molecule and a surface map of the ATP-binding site in hTRPV4-ARD (left) and the corresponding residues in the rTRPV1-ARD ATP-binding site (right). The bBound ATP molecule is shown as sticks (orange and yellow). (C) Finger 2 (2) and Finger 3 (3) structures of ATP-bound rat TRPV1-ARD (gray), rat and human TRPV2-ARD (cyan), human ATP-bound TRPV4-ARD (magenta), and mouse TRPV6-ARD (black). (D) Aromatic residue positioned behind the adenine base of ATP in Finger 2 (F231 in human TRPV4-ARD). (E) This aromatic residue is conserved in TRPV-ARDs that bind ATP (red rectangle). (F and G) ATP-agarose pull-down assays for wild-type and mutant rTRPV1-ARD (F) or hTRPV4-ARD (G). Coomassie-stained gels (top) of wild-type and mutant proteins loaded (left) and bound to ATP-agarose in the absence (middle) or presence (right) of competing free ATP. The normalized intensity of protein recovered (mean ± SD; n = 3) is plotted below. The statistical significance of the change in binding to ATP-agarose with respect to the wild type (WT) was determined by a multiple-comparison test using Dunnett’s method, with p < 0.01 indicated by an asterisk.
ATP Binding Stabilizes TRPV4-ARD and TRPV1-ARD
Figure 4

Figure 4. Effect of ATP on hTRPV4-ARD thermal stability. (A) Representative circular dichroism spectra of the purified TRPV4-ARD protein (3.4 μM) in the presence of ATP, AMP, or phosphate (1 mM each) at 10 °C. The wavelength (λ) of 222 nm used for thermostability assays is indicated by a vertical red line. (B) Representative traces of the thermostability assay. The molar ellipticity at 222 nm was measured as the protein solutions were heated at a rate of 1 °C/min. (C) Tm of TRPV4-ARD in the presence of 1 mM ATP, AMP, or phosphate. The statistical significance of the change in Tm was determined by a multiple-comparison test using the Tukey–Kramer method, with p < 0.05 and p < 0.01 indicated by one asterisk and two asterisks, respectively.
Figure 5

Figure 5. Effect of ATP binding on protein stability in rTRPV1-ARD. (A) Structure of TRPV1-ARD (gray) bound to ATP (green, sticks), with buried Cys157 highlighted (spheres). (B) TRPV1-ARD is modified at cysteine residues by PEG-maleimide (mPEG), causing an electrophoretic mobility shift on a Coomassie-stained SDS gel. Abbreviations: WT, wild type; CL, a cysteine-less variant; C157, CL-TRPV1-ARD C157 single-cysteine variant. Shown is a representative Coomassie-stained gel from one of three experiments. (C) Time course for modification of single-cysteine TRPV1-ARD variants C157 and C362 with 0.5 mM mPEG at room temperature. (D) Data from four experiments like that depicted in panel C were quantified, and the mean ± standard deviation was plotted. (E and F) Molecular dynamics simulation in which the termini of the ATP-bound TRPV1-ARD (E) or TRPV1-ARD structure with ATP removed prior to equilibrating the system (F) are pulled apart at a rate of 20 nm/ns. Superimposed are the structures at the start (gold) and end (blue) of the simulations. (G and H) Root-mean-square deviation of each Cα atom over the course of the simulation mapped onto the starting models with (G) or without (H) ATP. The change in color from blue to red indicates changes in rmsd from 0 to 80 Å. Simulations in which the termini were pulled apart at a rate of 2 nm/ns gave similar results. See Table S3 of the Supporting Information for experimental details.
Structural Analysis of TRPV4 Mutations Associated with Human Diseases
Figure 6

Figure 6. Mutations associated with human diseases in hTRPV4-ARD. (A) Positions of mutations associated with human inherited diseases that lie within hTRPV4-ARD. Abbreviations: SEDM, spondyloepiphyseal dysplasia, type Maroteaux; SMDK, spondylometaphyseal dysplasia, type Kozolowski; MD, metatropic dysplasia; SMA, spinal muscular atrophy; SPMA, scapuloperoneal spinal muscular atrophy; CMTC2, Charcot-Marie-Tooth disease type 2C; HMSN2C, hereditary motor and sensory neuropathy 2C. This figure was inspired by ref 51. (B) Location of the disease-causing mutations within TRPV4-ARD. Shown as spheres are 12 residue positions at which a total of 15 mutations causing human inherited diseases have been identified. The ATP molecule is shown as sticks. Skeletal dysplasia and neurophathy mutations are depicted as green and blue spheres, respectively. (C) Leu199 is located at the hydrophobic interface between ANK2 and ANK3. (D) Glu183 and Arg232 form a salt bridge on the convex face of TRPV4-ARD.
Figure 7

Figure 7. Thermal stability and ATP binding of hTRPV4-ARD mutants associated with inherited diseases. (A) The Tm determined by CD spectrometry in a phosphate-based buffer is plotted for wild-type and mutant hTRPV4-ARDs. The statistical significance is shown in Table S4 of the Supporting Information. (B) Coomassie-stained gels show wild-type and mutant TRPV4-ARDs loaded (top) and bound to ATP-agarose (bottom). (C) Normalized intensity of recovered protein (mean ± SD; n = 3). The statistical significance of the change in binding to ATP-agarose with respect to wild type (WT) was determined by a multiple-comparison test using Dunnett’s method, with p < 0.05 and p < 0.01 indicated by one asterisk and two asterisks, respectively.
ATP Binding by hTRPV4-ARD Mutants
Discussion
Supporting Information
TRPV-ARD structures used in this study (Table S1), structural similarity between TRPV4-ARD and other TRPV-ARDs (Table S2), molecular dynamics simulations of rat TRPV1-ARD (Table S3), Tm values of wild-type and mutant TRPV4-ARD proteins (Table S4), hTRPV4-ARD cysteine modification assay (Figure S1), stabilities of TRPV1-ARD in equilibrium and SMD simulations (Figure S2), and thermal stabilities of wild-type and mutant TRPV4-ARD proteins (Figure S3). This material is available free of charge via the Internet at http://pubs.acs.org.
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgment
We thank David Neau for assistance with data collection, Dr. Wilhelm A. Weihofen for help with data processing, Dr. Charlotte J. Sumner for providing human TRPV4 cDNA, Dr. Ute Hellmich for comments, and current and former lab members for technical help and discussions.
ANK | ankyrin repeat |
ARD | ankyrin repeat domain |
CaM | calmodulin |
CD | circular dichroism |
DTT | dithiothreitol |
EDTA | ethylenediaminetetraacetic acid |
HEPES | 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid |
IPTG | isopropyl β-d-thiogalactopyranoside |
MD | molecular dynamics |
PAGE | polyacrylamide gel electrophoresis |
PDB | Protein Data Bank |
PEG | polyethylene glycol |
rmsd | root-mean-square deviation |
SDS | sodium dodecyl sulfate |
SMD | steered molecular dynamics |
TRP | transient receptor potential |
TRPV | TRP vanilloid. |
References
This article references 56 other publications.
- 1Clapham, D. E. (2003) TRP channels as cellular sensors Nature 426, 517– 524Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXpsVejtrc%253D&md5=f45fe3448869a0e1c8623290df1cd218TRP channels as cellular sensorsClapham, David E.Nature (London, United Kingdom) (2003), 426 (6966), 517-524CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)A review. TRP channels are the vanguard of our sensory systems, responding to temp., touch, pain, osmolarity, pheromones, taste and other stimuli. But their role is much broader than classical sensory transduction. They are an ancient sensory app. for the cell, not just the multicellular organism, and they have been adapted to respond to all manner of stimuli, from both within and outside the cell.
- 2Clapham, D. E., Montell, C., Schultz, G., and Julius, D. (2003) International Union of Pharmacology. XLIII. Compendium of voltage-gated ion channels: Transient receptor potential channels Pharmacol. Rev. 55, 591– 596Google ScholarThere is no corresponding record for this reference.
- 3Wu, L. J., Sweet, T. B., and Clapham, D. E. (2010) International Union of Basic and Clinical Pharmacology. LXXVI. Current progress in the mammalian TRP ion channel family Pharmacol. Rev. 62, 381– 404Google ScholarThere is no corresponding record for this reference.
- 4Tominaga, M. and Tominaga, T. (2005) Structure and function of TRPV1 Pfluegers Arch. 451, 143– 150Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXhtVKqu7rE&md5=35d28025b866fe7e955bd2861dd5a9aaStructure and function of TRPV1Tominaga, Makoto; Tominaga, TomokoPfluegers Archiv (2005), 451 (1), 143-150CODEN: PFLABK; ISSN:0031-6768. (Springer GmbH)A review. Capsaicin, the main ingredient in hot chili peppers, elicits a sensation of burning pain by selectively activating sensory neurons that convey information about noxious stimuli to the central nervous system. The capsaicin receptor, transient receptor potential vanilloid 1 (TRPV1), is predicted to have 6 transmembrane (TM) domains and a short, pore-forming hydrophobic stretch between the 5th and 6th TM domains, and is activated not only by capsaicin but also by heat (>43°), acid, and various lipids. Within the TPRV1 protein, many regions and amino acids involved in specific functions (multimerization, capsaicin action, proton action, heat activation, desensitization, permeability, phosphorylation and modulation by lipids) have been identified since the cloning in 1997. Given the fact that TRPV1 is a key mol. in peripheral nociception, these regions and amino acids could prove useful for the development of novel anti-nociceptive or anti-inflammatory agents.
- 5van de Graaf, S. F., Hoenderop, J. G., and Bindels, R. J. (2006) Regulation of TRPV5 and TRPV6 by associated proteins Am. J. Physiol. 290, F1295– F1302Google ScholarThere is no corresponding record for this reference.
- 6Rosenbaum, T., Gordon-Shaag, A., Munari, M., and Gordon, S. E. (2004) Ca2+/calmodulin modulates TRPV1 activation by capsaicin J. Gen. Physiol. 123, 53– 62Google ScholarThere is no corresponding record for this reference.
- 7Lishko, P. V., Procko, E., Jin, X., Phelps, C. B., and Gaudet, R. (2007) The ankyrin repeats of TRPV1 bind multiple ligands and modulate channel sensitivity Neuron 54, 905– 918Google ScholarThere is no corresponding record for this reference.
- 8Phelps, C. B., Wang, R. R., Choo, S. S., and Gaudet, R. (2010) Differential regulation of TRPV1, TRPV3, and TRPV4 sensitivity through a conserved binding site on the ankyrin repeat domain J. Biol. Chem. 285, 731– 740Google ScholarThere is no corresponding record for this reference.
- 9Prescott, E. D. and Julius, D. (2003) A modular PIP2 binding site as a determinant of capsaicin receptor sensitivity Science 300, 1284– 1288Google ScholarThere is no corresponding record for this reference.
- 10Ufret-Vincenty, C. A., Klein, R. M., Hua, L., Angueyra, J., and Gordon, S. E. (2011) Localization of the PIP2 sensor of TRPV1 ion channels J. Biol. Chem. 286, 9688– 9698Google ScholarThere is no corresponding record for this reference.
- 11Numazaki, M., Tominaga, T., Takeuchi, K., Murayama, N., Toyooka, H., and Tominaga, M. (2003) Structural determinant of TRPV1 desensitization interacts with calmodulin Proc. Natl. Acad. Sci. U.S.A. 100, 8002– 8006Google ScholarThere is no corresponding record for this reference.
- 12Liedtke, W., Choe, Y., Marti-Renom, M. A., Bell, A. M., Denis, C. S., Sali, A., Hudspeth, A. J., Friedman, J. M., and Heller, S. (2000) Vanilloid receptor-related osmotically activated channel (VR-OAC), a candidate vertebrate osmoreceptor Cell 103, 525– 535Google ScholarThere is no corresponding record for this reference.
- 13Guler, A. D., Lee, H., Iida, T., Shimizu, I., Tominaga, M., and Caterina, M. (2002) Heat-evoked activation of the ion channel, TRPV4 J. Neurosci. 22, 6408– 6414Google ScholarThere is no corresponding record for this reference.
- 14Watanabe, H., Vriens, J., Suh, S. H., Benham, C. D., Droogmans, G., and Nilius, B. (2002) Heat-evoked activation of TRPV4 channels in a HEK293 cell expression system and in native mouse aorta endothelial cells J. Biol. Chem. 277, 47044– 47051Google ScholarThere is no corresponding record for this reference.
- 15Watanabe, H., Vriens, J., Prenen, J., Droogmans, G., Voets, T., and Nilius, B. (2003) Anandamide and arachidonic acid use epoxyeicosatrienoic acids to activate TRPV4 channels Nature 424, 434– 438Google ScholarThere is no corresponding record for this reference.
- 16Liedtke, W. (2005) TRPV4 as osmosensor: A transgenic approach Pfluegers Arch. 451, 176– 180Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXhtVKqu73E&md5=ea165f301f59cf2468ae28e73ae08ae9TRPV4 as osmosensor: a transgenic approachLiedtke, WolfgangPfluegers Archiv (2005), 451 (1), 176-180CODEN: PFLABK; ISSN:0031-6768. (Springer GmbH)A review. The transient receptor potential vanilloid 4 (TRPV4) ion channel was named initially vanilloid-receptor-related osmotically activated channel (VR-OAC). Preliminary answers to the question, "What is the function of the trpv4 gene in live animals " are highlighted briefly in this review. In trpv4 null mice, TRPV4 is necessary for the maintenance of osmotic equil., and in Caenorhabditis elegans transgenic for mammalian TRPV4, TRPV4 directs the osmotic avoidance response in the context of the ASH "nociceptive" neuron. The mol. mechanisms of gating of TRPV4 in vivo need to be detd.; in particular, whether TRPV4 in live animals is gated via phosphorylation of defined amino-acid residues or more directly through the osmotic stimulus itself.
- 17Tominaga, M. and Caterina, M. J. (2004) Thermosensation and pain J. Neurobiol. 61, 3– 12Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BD2cvlt12hsQ%253D%253D&md5=ac03f5366520d918056a6c30efee8850Thermosensation and painTominaga Makoto; Caterina Michael JJournal of neurobiology (2004), 61 (1), 3-12 ISSN:0022-3034.We feel a wide range of temperatures spanning from cold to heat. Within this range, temperatures over about 43 degrees C and below about 15 degrees C evoke not only a thermal sensation, but also a feeling of pain. In mammals, six thermosensitive ion channels have been reported, all of which belong to the TRP (transient receptor potential) superfamily. These include TRPV1 (VR1), TRPV2 (VRL-1), TRPV3, TRPV4, TRPM8 (CMR1), and TRPA1 (ANKTM1). These channels exhibit distinct thermal activation thresholds (>43 degrees C for TRPV1, >52 degrees C for TRPV2, > approximately 34-38 degrees C for TRPV3, > approximately 27-35 degrees C for TRPV4, < approximately 25-28 degrees C for TRPM8 and <17 degrees C for TRPA1), and are expressed in primary sensory neurons as well as other tissues. The involvement of TRPV1 in thermal nociception has been demonstrated by multiple methods, including the analysis of TRPV1-deficient mice. TRPV2, TRPM8, and TRPA1 are also very likely to be involved in thermal nociception, because their activation thresholds are within the noxious range of temperatures.
- 18Mochizuki, T., Sokabe, T., Araki, I., Fujishita, K., Shibasaki, K., Uchida, K., Naruse, K., Koizumi, S., Takeda, M., and Tominaga, M. (2009) The TRPV4 cation channel mediates stretch-evoked Ca2+ influx and ATP release in primary urothelial cell cultures J. Biol. Chem. 284, 21257– 21264Google ScholarThere is no corresponding record for this reference.
- 19Sokabe, T. and Tominaga, M. (2010) The TRPV4 cation channel: A molecule linking skin temperature and barrier function Commun. Integr. Biol. 3, 619– 621Google Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC3M3gsVGmug%253D%253D&md5=090f2a24c4f290f61a7f74a22fa4f301The TRPV4 cation channel: A molecule linking skin temperature and barrier functionSokabe Takaaki; Tominaga MakotoCommunicative & integrative biology (2010), 3 (6), 619-21 ISSN:.The skin barrier function is indispensable for terrestrial animals to avoid dehydration. The function is achieved by a hydrophobic cornified layer consisting of dead keratinocytes and lipids, and by an intercellular junction barrier formed among differentiated keratinocytes. A recent report demonstrated that TRPV4, one of the temperature-sensitive cation channels, contributes to the formation and maintenance of the intercellular junction-dependent barrier in the skin. TRPV4 associates with the E-cadherin complex via β-catenin, and thereby participates in the promotion of cell-cell junction development. TRPV4 allows influx of Ca(2+) ions from the extracellular space at physiological skin temperatures. The Ca(2+) influx induces Rho activation and promotes actin fiber organization and junction formation, thereby augmenting barrier integrity. Indeed, the intercellular junction structures and the skin barrier function were impaired in TRPV4-deficeint mice. This novel role of TRPV4 in keratinocytes may explain the significant correlation between temperature and the condition of skin.>
- 20Shibasaki, K., Suzuki, M., Mizuno, A., and Tominaga, M. (2007) Effects of body temperature on neural activity in the hippocampus: Regulation of resting membrane potentials by transient receptor potential vanilloid 4 J. Neurosci. 27, 1566– 1575Google Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXitl2gu78%253D&md5=f2d68caf39b30e807957c8cca4f9f76cEffects of body temperature on neural activity in the hippocampus: regulation of resting membrane potentials by transient receptor potential vanilloid 4Shibasaki, Koji; Suzuki, Makoto; Mizuno, Atsuko; Tominaga, MakotoJournal of Neuroscience (2007), 27 (7), 1566-1575CODEN: JNRSDS; ISSN:0270-6474. (Society for Neuroscience)Physiol. body temp. is an important determinant for neural functions, and it is well established that changes in temp. have dynamic influences on hippocampal neural activities. However, the detailed mol. mechanisms have never been clarified. Here, we show that hippocampal neurons express functional transient receptor potential vanilloid 4 (TRPV4), one of the thermosensitive TRP (transient receptor potential) channels, and that TRPV4 is constitutively active at physiol. temp. Activation of TRPV4 at 37°C depolarized the resting membrane potential in hippocampal neurons by allowing cation influx, which was obsd. in wild-type (WT) neurons, but not in TRPV4-deficient (TRPV4KO) cells, although dendritic morphol., synaptic marker clustering, and synaptic currents were indistinguishable between the two genotypes. Furthermore, current injection studies revealed that TRPV4KO neurons required larger depolarization to evoke firing, equiv. to WT neurons, indicating that TRPV4 is a key regulator for hippocampal neural excitabilities. We conclude that TRPV4 is activated by physiol. temp. in hippocampal neurons and thereby controls their excitability.
- 21Alessandri-Haber, N., Dina, O. A., Joseph, E. K., Reichling, D., and Levine, J. D. (2006) A transient receptor potential vanilloid 4-dependent mechanism of hyperalgesia is engaged by concerted action of inflammatory mediators J. Neurosci. 26, 3864– 3874Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xjs1yksbo%253D&md5=30220a30e99775236cd7876c1b121d7bA transient receptor potential vanilloid 4-dependent mechanism of hyperalgesia is engaged by concerted action of inflammatory mediatorsAlessandri-Haber, Nicole; Dina, Olayinka A.; Joseph, Elizabeth K.; Reichling, David; Levine, Jon D.Journal of Neuroscience (2006), 26 (14), 3864-3874CODEN: JNRSDS; ISSN:0270-6474. (Society for Neuroscience)The transient receptor potential vanilloid 4 (TRPV4) is a primary afferent transducer that plays a crucial role in neuropathic hyperalgesia for osmotic and mech. stimuli, as well as in inflammatory mediator-induced hyperalgesia for osmotic stimuli. In view of the clin. importance of mech. hyperalgesia in inflammatory states, the present study investigated the role of TRPV4 in mech. hyperalgesia induced by inflammatory mediators and the second-messenger pathways involved. Intradermal injection of either the inflammogen carrageenan or a soup of inflammatory mediators enhanced the nocifensive paw-withdrawal reflex elicited by hypotonic or mech. stimuli in rat. Spinal administration of TRPV4 antisense oligodeoxynucleotide blocked the enhancement without altering baseline nociceptive threshold. Similarly, in TRPV4-/- knock-out mice, inflammatory soup failed to induce any significant mech. or osmotic hyperalgesia. In vitro investigation showed that inflammatory mediators engage the TRPV4-mediated mechanism of sensitization by direct action on dissocd. primary afferent neurons. Addnl. behavioral observations suggested that multiple mediators are necessary to achieve sufficient activation of the cAMP pathway to engage the TRPV4-dependent mechanism of hyperalgesia. In addn., direct activation of protein kinase A or protein kinase C ε, two pathways that mediate inflammation-induced mech. hyperalgesia, also induced hyperalgesia for both hypotonic and mech. stimuli that was decreased by TRPV4 antisense and absent in TRPV4-/- mice. We conclude that TRPV4 plays a crucial role in the mech. hyperalgesia that is generated by the concerted action of inflammatory mediators present in inflamed tissues.
- 22Masuyama, R., Vriens, J., Voets, T., Karashima, Y., Owsianik, G., Vennekens, R., Lieben, L., Torrekens, S., Moermans, K., Vanden Bosch, A., Bouillon, R., Nilius, B., and Carmeliet, G. (2008) TRPV4-mediated calcium influx regulates terminal differentiation of osteoclasts Cell Metab. 8, 257– 265Google Scholar22https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXhtFeju77E&md5=a6987e7efdaad7228132430519f13e7aTRPV4-mediated calcium influx regulates terminal differentiation of osteoclastsMasuyama, Ritsuko; Vriens, Joris; Voets, Thomas; Karashima, Yuji; Owsianik, Grzegorz; Vennekens, Rudi; Lieben, Liesbet; Torrekens, Sophie; Moermans, Karen; Vanden Bosch, An; Bouillon, Roger; Nilius, Bernd; Carmeliet, GeertCell Metabolism (2008), 8 (3), 257-265CODEN: CMEEB5; ISSN:1550-4131. (Cell Press)Calcium signaling controls multiple cellular functions and is regulated by the release from internal stores and entry from extracellular fluid. In bone, osteoclast differentiation is induced by RANKL (receptor activator of NF-κB ligand)-evoked intracellular Ca2+ oscillations, which trigger nuclear factor-activated T cells (NFAT) c1-responsive gene transcription. However, the Ca2+ channels involved remain largely unidentified. Here we show that genetic ablation in mice of Trpv4, a Ca2+-permeable channel of the transient receptor potential (TRP) family, increases bone mass by impairing bone resorption. TRPV4 mediates basolateral Ca2+ influx specifically in large osteoclasts when Ca2+ oscillations decline. TRPV4-mediated Ca2+ influx hereby secures intracellular Ca2+ concns., ensures NFATc1-regulated gene transcription, and regulates the terminal differentiation and activity of osteoclasts. In conclusion, our data indicate that Ca2+ oscillations and TRPV4-mediated Ca2+ influx are sequentially required to sustain NFATc1-dependent gene expression throughout osteoclast differentiation, and we propose TRPV4 as a therapeutic target for bone diseases.
- 23Nilius, B. and Owsianik, G. (2010) Transient receptor potential channelopathies Pfluegers Arch. 460, 437– 450Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXntlGns70%253D&md5=e1f538ce79e09a627cec9424d2c45fdbTransient receptor potential channelopathiesNilius, Bernd; Owsianik, GrzegorzPfluegers Archiv (2010), 460 (2), 437-450CODEN: PFLABK; ISSN:0031-6768. (Springer GmbH)A review. In the past years, several hereditary diseases caused by defects in transient receptor potential channels (TRP) genes have been described. This review summarizes our current knowledge about TRP channelopathies and their possible pathomechanisms. Based on available genetic indications, we will also describe several putative pathol. conditions in which (mal)function of TRP channels could be anticipated.
- 24Verma, P., Kumar, A., and Goswami, C. (2010) TRPV4-mediated channelopathies Channels 4, 319– 328Google ScholarThere is no corresponding record for this reference.
- 25Rock, M. J., Prenen, J., Funari, V. A., Funari, T. L., Merriman, B., Nelson, S. F., Lachman, R. S., Wilcox, W. R., Reyno, S., Quadrelli, R., Vaglio, A., Owsianik, G., Janssens, A., Voets, T., Ikegawa, S., Nagai, T., Rimoin, D. L., Nilius, B., and Cohn, D. H. (2008) Gain-of-function mutations in TRPV4 cause autosomal dominant brachyolmia Nat. Genet. 40, 999– 1003Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXptVKku7g%253D&md5=33da69ac6e23dbe409b9c1807c9f504dGain-of-function mutations in TRPV4 cause autosomal dominant brachyolmiaRock, Matthew J.; Prenen, Jean; Funari, Vincent A.; Funari, Tara L.; Merriman, Barry; Nelson, Stanley F.; Lachman, Ralph S.; Wilcox, William R.; Reyno, Soraya; Quadrelli, Roberto; Vaglio, Alicia; Owsianik, Grzegorz; Janssens, Annelies; Voets, Thomas; Ikegawa, Shiro; Nagai, Toshiro; Rimoin, David L.; Nilius, Bernd; Cohn, Daniel H.Nature Genetics (2008), 40 (8), 999-1003CODEN: NGENEC; ISSN:1061-4036. (Nature Publishing Group)Daniel Cohn and colleagues identify mutations in the gene encoding the calcium-permeable cation channel TRPV4 in families with autosomal dominant brachyolmia. Functional studies show that the mutations result in gain-of-function of channel activation. Daniel Cohn and colleagues identify mutations in the gene encoding the calcium-permeable cation channel TRPV4 in families with autosomal dominant brachyolmia. Functional studies show that the mutations result in gain-of-function of channel activation. The brachyolmias constitute a clin. and genetically heterogeneous group of skeletal dysplasias characterized by a short trunk, scoliosis and mild short stature. Here, we identify a locus for an autosomal dominant form of brachyolmia on chromosome 12q24.1-12q24.2. Among the genes in the genetic interval, we selected TRPV4, which encodes a calcium permeable cation channel of the transient receptor potential (TRP) vanilloid family, as a candidate gene because of its cartilage-selective gene expression pattern. In two families with the phenotype, we identified point mutations in TRPV4 that encoded R616Q and V620I substitutions, resp. Patch clamp studies of transfected HEK cells showed that both mutations resulted in a dramatic gain of function characterized by increased constitutive activity and elevated channel activation by either mechano-stimulation or agonist stimulation by arachidonic acid or the TRPV4-specific agonist 4α-phorbol 12,13-didecanoate (4αPDD). This study thus defines a previously unknown mechanism, activation of a calcium-permeable TRP ion channel, in skeletal dysplasia pathogenesis.
- 26Krakow, D., Vriens, J., Camacho, N., Luong, P., Deixler, H., Funari, T. L., Bacino, C. A., Irons, M. B., Holm, I. A., Sadler, L., Okenfuss, E. B., Janssens, A., Voets, T., Rimoin, D. L., Lachman, R. S., Nilius, B., and Cohn, D. H. (2009) Mutations in the gene encoding the calcium-permeable ion channel TRPV4 produce spondylometaphyseal dysplasia, Kozlowski type and metatropic dysplasia Am. J. Hum. Genet. 84, 307– 315Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXmvFCmu7c%253D&md5=f0e29ebe74f51d9bd0539e0c2ae805d4Mutations in the gene encoding the calcium-permeable ion channel TRPV4 produce spondylometaphyseal dysplasia, Kozlowski type and metatropic dysplasiaKrakow, Deborah; Vriens, Joris; Camacho, Natalia; Luong, Phi; Deixler, Hannah; Funari, Tara L.; Bacino, Carlos A.; Irons, Mira B.; Holm, Ingrid A.; Sadler, Laurie; Okenfuss, Ericka B.; Janssens, Annelies; Voets, Thomas; Rimoin, David L.; Lachman, Ralph S.; Nilius, Bernd; Cohn, Daniel H.American Journal of Human Genetics (2009), 84 (3), 307-315CODEN: AJHGAG; ISSN:0002-9297. (Cell Press)The spondylometaphyseal dysplasias (SMDs) are a group of short-stature disorders distinguished by abnormalities in the vertebrae and the metaphyses of the tubular bones. SMD Kozlowski type (SMDK) is a well-defined autosomal-dominant SMD characterized by significant scoliosis and mild metaphyseal abnormalities in the pelvis. The vertebrae exhibit platyspondyly and overfaced pedicles similar to autosomal-dominant brachyolmia, which can result from heterozygosity for activating mutations in the gene encoding TRPV4, a calcium-permeable ion channel. Mutation anal. in six out of six patients with SMDK demonstrated heterozygosity for missense mutations in TRPV4, and one mutation, predicting a R594H substitution, was recurrent in four patients. Similar to autosomal-dominant brachyolmia, the mutations altered basal calcium channel activity in vitro. Metatropic dysplasia is another SMD that has been proposed to have both clin. and genetic heterogeneity. Patients with the nonlethal form of metatropic dysplasia present with a progressive scoliosis, widespread metaphyseal involvement of the appendicular skeleton, and carpal ossification delay. Because of some similar radiog. features between SMDK and metatropic dysplasia, TRPV4 was tested as a disease gene for nonlethal metatropic dysplasia. In two sporadic cases, heterozygosity for de novo missense mutations in TRPV4 was found. The findings demonstrate that mutations in TRPV4 produce a phenotypic spectrum of skeletal dysplasias from the mild autosomal-dominant brachyolmia to SMDK to autosomal-dominant metatropic dysplasia, suggesting that these disorders should be grouped into a new bone dysplasia family.
- 27Deng, H. X., Klein, C. J., Yan, J., Shi, Y., Wu, Y., Fecto, F., Yau, H. J., Yang, Y., Zhai, H., Siddique, N., Hedley-Whyte, E. T., Delong, R., Martina, M., Dyck, P. J., and Siddique, T. (2010) Scapuloperoneal spinal muscular atrophy and CMT2C are allelic disorders caused by alterations in TRPV4 Nat. Genet. 42, 165– 169Google Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhs1Skur7F&md5=ebf469c8e33fcf491b99b00619b7e9adScapuloperoneal spinal muscular atrophy and CMT2C are allelic disorders caused by alterations in TRPV4Deng, Han-Xiang; Klein, Christopher J.; Yan, Jian-Hua; Shi, Yong; Wu, Yan-Hong; Fecto, Faisal; Yau, Hau-Jie; Yang, Yi; Zhai, Hong; Siddique, Nailah; Hedley-Whyte, E. Tessa; De Long, Robert; Martina, Marco; Dyck, Peter J.; Siddique, TeepuNature Genetics (2010), 42 (2), 165-169CODEN: NGENEC; ISSN:1061-4036. (Nature Publishing Group)Scapuloperoneal spinal muscular atrophy (SPSMA) and hereditary motor and sensory neuropathy type IIC (HMSN IIC, also known as HMSN2C or Charcot-Marie-Tooth disease type 2C (CMT2C)) are phenotypically heterogeneous disorders involving topog. distinct nerves and muscles. We originally described a large New England family of French-Canadian origin with SPSMA and an American family of English and Scottish descent with CMT2C. We mapped SPSMA and CMT2C risk loci to 12q24.1-q24.31 with an overlapping region between the two diseases. Further anal. reduced the CMT2C risk locus to a 4-Mb region. Here we report that SPSMA and CMT2C are allelic disorders caused by mutations in the gene encoding the transient receptor potential cation channel, subfamily V, member 4 (TRPV4). Functional anal. revealed that increased calcium channel activity is a distinct property of both SPSMA- and CMT2C-causing mutant proteins. Our findings link mutations in TRPV4 to altered calcium homeostasis and peripheral neuropathies, implying a pathogenic mechanism and possible options for therapy for these disorders.
- 28Landoure, G., Zdebik, A. A., Martinez, T. L., Burnett, B. G., Stanescu, H. C., Inada, H., Shi, Y., Taye, A. A., Kong, L., Munns, C. H., Choo, S. S., Phelps, C. B., Paudel, R., Houlden, H., Ludlow, C. L., Caterina, M. J., Gaudet, R., Kleta, R., Fischbeck, K. H., and Sumner, C. J. (2010) Mutations in TRPV4 cause Charcot-Marie-Tooth disease type 2C Nat. Genet. 42, 170– 174Google Scholar28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhs1Skur%252FM&md5=91ca81c1ac5b02d2f7429e50e99e6a86Mutations in TRPV4 cause Charcot-Marie-Tooth disease type 2CLandoure, Guida; Zdebik, Anselm A.; Martinez, Tara L.; Burnett, Barrington G.; Stanescu, Horia C.; Inada, Hitoshi; Shi, Yi-Jun; Taye, Addis A.; Kong, Ling-Ling; Munns, Clare H.; Choo, Shelly S.; Phelps, Christopher B.; Paudel, Reema; Houlden, Henry; Ludlow, Christy L.; Caterina, Michael J.; Gaudet, Rachelle; Kleta, Robert; Fischbeck, Kenneth H.; Sumner, Charlotte J.Nature Genetics (2010), 42 (2), 170-174CODEN: NGENEC; ISSN:1061-4036. (Nature Publishing Group)Charcot-Marie-Tooth disease type 2C (CMT2C) is an autosomal dominant neuropathy characterized by limb, diaphragm and laryngeal muscle weakness. Two unrelated families with CMT2C showed significant linkage to chromosome 12q24.11. We sequenced all genes in this region and identified two heterozygous missense mutations in the TRPV4 gene, C805T and G806A, resulting in the amino acid substitutions R269C and R269H. TRPV4 is a well-known member of the TRP superfamily of cation channels. In TRPV4-transfected cells, the CMT2C mutations caused marked cellular toxicity and increased constitutive and activated channel currents. Mutations in TRPV4 were previously assocd. with skeletal dysplasias. Our findings indicate that TRPV4 mutations can also cause a degenerative disorder of the peripheral nerves. The CMT2C-assocd. mutations lie in a distinct region of the TRPV4 ankyrin repeats, suggesting that this phenotypic variability may be due to differential effects on regulatory protein-protein interactions.
- 29Jin, X., Touhey, J., and Gaudet, R. (2006) Structure of the N-terminal ankyrin repeat domain of the TRPV2 ion channel J. Biol. Chem. 281, 25006– 25010Google ScholarThere is no corresponding record for this reference.
- 30Otwinowski, Z. and Minor, W. (1997) Processing of X-ray Diffraction Data Collected in Oscillation Mode Methods Enzymol. 276, 307– 326Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXivFehsbw%253D&md5=c9536971d4e32cc35352c40fb9368131Processing of x-ray diffraction data collected in oscillation modeOtwinowski, Zbyszek; Minor, WladekMethods in Enzymology (1997), 276 (Macromolecular Crystallography, Part A), 307-326CODEN: MENZAU; ISSN:0076-6879. (Academic)Macromol. crystallog. is an iterative process. Rarely do the first crystals provide all the necessary data to solve the biol. problem being studied. Each step benefits from experience learned in previous steps. To monitor the progress, the HKL package provides 2 tools: (1) statistics, both weighted (χ2) and unweighted (R-merge), are provided, and the Bayesian reasoning and multicomponent error model facilitates obtaining the proper error ests. and (2) visualization of the process plays a double role by helping the operator to confirm that the process of data redn., including the resulting statistics, is correct, and allowing one to evaluate problems for which there are no good statistical criteria. Visualization also provides confidence that the point of diminishing returns in data collection and redn. has been reached. At that point, the effort should be directed to solving the structure. The methods presented here have been applied to solve a large variety of problems, from inorg. mols. with 5 Å unit cell to rotavirus of 700 Å diam. crystd. in 700 × 1000 × 1400 Å cell. Overall quality of the method was tested by many researchers by successful application of the programs to MAD structure detns.
- 31Vagin, A. and Teplyakov, A. (2000) An approach to multi-copy search in molecular replacement Acta Crystallogr. D56, 1622– 1624Google ScholarThere is no corresponding record for this reference.
- 32McCoy, A. J., Grosse-Kunstleve, R. W., Adams, P. D., Winn, M. D., Storoni, L. C., and Read, R. J. (2007) Phaser crystallographic software J. Appl. Crystallogr. 40, 658– 674Google Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXnslWqsLk%253D&md5=c63b722ae97e0a74e6a5a079d388f09fPhaser crystallographic softwareMcCoy, Airlie J.; Grosse-Kunstleve, Ralf W.; Adams, Paul D.; Winn, Martyn D.; Storoni, Laurent C.; Read, Randy J.Journal of Applied Crystallography (2007), 40 (4), 658-674CODEN: JACGAR; ISSN:0021-8898. (International Union of Crystallography)Phaser is a program for phasing macromol. crystal structures by both mol. replacement and exptl. phasing methods. The novel phasing algorithms implemented in Phaser have been developed using max. likelihood and multivariate statistics. For mol. replacement, the new algorithms have proved to be significantly better than traditional methods in discriminating correct solns. from noise, and for single-wavelength anomalous dispersion exptl. phasing, the new algorithms, which account for correlations between F+ and F-, give better phases (lower mean phase error with respect to the phases given by the refined structure) than those that use mean F and anomalous differences ΔF. One of the design concepts of Phaser was that it be capable of a high degree of automation. To this end, Phaser (written in C++) can be called directly from Python, although it can also be called using traditional CCP4 keyword-style input. Phaser is a platform for future development of improved phasing methods and their release, including source code, to the crystallog. community.
- 33Emsley, P. and Cowtan, K. (2004) Coot: Model-building tools for molecular graphics Acta Crystallogr. D60, 2126– 2132Google Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXhtVars73P&md5=1be390f3bb6fd584468499ad0921161eCoot: model-building tools for molecular graphicsEmsley, Paul; Cowtan, KevinActa Crystallographica, Section D: Biological Crystallography (2004), D60 (12, Pt. 1), 2126-2132CODEN: ABCRE6; ISSN:0907-4449. (Blackwell Publishing Ltd.)CCP4mg is a project that aims to provide a general-purpose tool for structural biologists, providing tools for x-ray structure soln., structure comparison and anal., and publication-quality graphics. The map-fitting tools are available as a stand-alone package, distributed as 'Coot'.
- 34Murshudov, G. N., Vagin, A. A., and Dodson, E. J. (1997) Refinement of macromolecular structures by the maximum-likelihood method Acta Crystallogr. D53, 240– 255Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXjs1Gnsb4%253D&md5=ec7f141ce1542f7ff458b98ecfe3f8afRefinement of macromolecular structures by the maximum-likelihood methodMurshudov, Garib N.; Vagin, Alexei A.; Dodson, Eleanor J.Acta Crystallographica, Section D: Biological Crystallography (1997), D53 (3), 240-255CODEN: ABCRE6; ISSN:0907-4449. (Munksgaard)A review with many refs. on the math. basis of max. likelihood. The likelihood function for macromol. structures is extended to include prior phase information and exptl. std. uncertainties. The assumption that different parts of a structure might have different errors is considered. A method for estg. σA using "free" reflections is described and its effects analyzed. The derived equations have been implemented in the program REFMAC. This has been tested on several proteins at different stages of refinement (bacterial α-amylase, cytochrome c', cross-linked insulin and oligopeptide binding protein). The results derived using the max.-likelihood residual are consistently better than those obtained from least-squares refinement.
- 35Humphrey, W., Dalke, A., and Schulten, K. (1996) VMD: Visual molecular dynamics J. Mol. Graphics 14, 27– 38Google ScholarThere is no corresponding record for this reference.
- 36Phillips, J. C., Braun, R., Wang, W., Gumbart, J., Tajkhorshid, E., Villa, E., Chipot, C., Skeel, R. D., Kale, L., and Schulten, K. (2005) Scalable molecular dynamics with NAM J. Comput. Chem. 26, 1781– 1802Google ScholarThere is no corresponding record for this reference.
- 37MacKerell, A. D., Jr., Bashford, D., Bellott, M., Dunbrack, R. L., Jr., Evanseck, J. D., Field, M. J., Fischer, S., Gao, J., Guo, H., Ha, S., Joseph-McCarthy, D., Kuchnir, L., Kuczera, K., Lau, F. T. K., Mattos, C., Michnick, S., Ngo, T., Nguyen, D. T., Prodhom, B., Reiher, W. E., III, Roux, B., Schlenkrich, M., Smith, J. C., Stote, R., Straub, J., Watanabe, M., Wiórkiewicz-Kuczera, J., Yin, D., and Karplus, M. (1998) All-Atom Empirical Potential for Molecular Modeling and Dynamics Studies of Proteins J. Phys. Chem. B 102, 3586– 3616Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXivVOlsb4%253D&md5=ebb5100dafd0daeee60ca2fa66c1324aAll-Atom Empirical Potential for Molecular Modeling and Dynamics Studies of ProteinsMacKerell, A. D., Jr.; Bashford, D.; Bellott, M.; Dunbrack, R. L.; Evanseck, J. D.; Field, M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.; Joseph-McCarthy, D.; Kuchnir, L.; Kuczera, K.; Lau, F. T. K.; Mattos, C.; Michnick, S.; Ngo, T.; Nguyen, D. T.; Prodhom, B.; Reiher, W. E., III; Roux, B.; Schlenkrich, M.; Smith, J. C.; Stote, R.; Straub, J.; Watanabe, M.; Wiorkiewicz-Kuczera, J.; Yin, D.; Karplus, M.Journal of Physical Chemistry B (1998), 102 (18), 3586-3616CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)New protein parameters are reported for the all-atom empirical energy function in the CHARMM program. The parameter evaluation was based on a self-consistent approach designed to achieve a balance between the internal (bonding) and interaction (nonbonding) terms of the force field and among the solvent-solvent, solvent-solute, and solute-solute interactions. Optimization of the internal parameters used exptl. gas-phase geometries, vibrational spectra, and torsional energy surfaces supplemented with ab initio results. The peptide backbone bonding parameters were optimized with respect to data for N-methylacetamide and the alanine dipeptide. The interaction parameters, particularly the at. charges, were detd. by fitting ab initio interaction energies and geometries of complexes between water and model compds. that represented the backbone and the various side chains. In addn., dipole moments, exptl. heats and free energies of vaporization, solvation and sublimation, mol. vols., and crystal pressures and structures were used in the optimization. The resulting protein parameters were tested by applying them to noncyclic tripeptide crystals, cyclic peptide crystals, and the proteins crambin, bovine pancreatic trypsin inhibitor, and carbonmonoxy myoglobin in vacuo and in a crystal. A detailed anal. of the relationship between the alanine dipeptide potential energy surface and calcd. protein φ, χ angles was made and used in optimizing the peptide group torsional parameters. The results demonstrate that use of ab initio structural and energetic data by themselves are not sufficient to obtain an adequate backbone representation for peptides and proteins in soln. and in crystals. Extensive comparisons between mol. dynamics simulation and exptl. data for polypeptides and proteins were performed for both structural and dynamic properties. Calcd. data from energy minimization and dynamics simulations for crystals demonstrate that the latter are needed to obtain meaningful comparisons with exptl. crystal structures. The presented parameters, in combination with the previously published CHARMM all-atom parameters for nucleic acids and lipids, provide a consistent set for condensed-phase simulations of a wide variety of mols. of biol. interest.
- 38Mackerell, A. D., Jr., Feig, M., and Brooks, C. L., III (2004) Extending the treatment of backbone energetics in protein force fields: Limitations of gas-phase quantum mechanics in reproducing protein conformational distributions in molecular dynamics simulations J. Comput. Chem. 25, 1400– 1415Google Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXlsVOgt7c%253D&md5=b2451bb5df548447f8b172a211bc1848Extending the treatment of backbone energetics in protein force fields: Limitations of gas-phase quantum mechanics in reproducing protein conformational distributions in molecular dynamics simulationsMacKerell, Alexander D., Jr.; Feig, Michael; Brooks, Charles L., IIIJournal of Computational Chemistry (2004), 25 (11), 1400-1415CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Computational studies of proteins based on empirical force fields represent a powerful tool to obtain structure-function relationships at an at. level, and are central in current efforts to solve the protein folding problem. The results from studies applying these tools are, however, dependent on the quality of the force fields used. In particular, accurate treatment of the peptide backbone is crucial to achieve representative conformational distributions in simulation studies. To improve the treatment of the peptide backbone, quantum mech. (QM) and mol. mech. (MM) calcns. were undertaken on the alanine, glycine, and proline dipeptides, and the results from these calcns. were combined with mol. dynamics (MD) simulations of proteins in crystal and aq. environments. QM potential energy maps of the alanine and glycine dipeptides at the LMP2/cc-pVxZ/MP2/6-31G* levels, where x = D, T, and Q, were detd., and are compared to available QM studies on these mols. The LMP2/cc pVQZ//MP2/6-31G* energy surfaces for all three dipeptides were then used to improve the MM treatment of the dipeptides. These improvements included addnl. parameter optimization via Monte Carlo simulated annealing and extension of the potential energy function to contain peptide backbone .vphi., ψ dihedral crossterms or a .vphi., ψ grid-based energy correction term. Simultaneously, MD simulations of up to seven proteins in their cryst. environments were used to validate the force field enhancements. Comparison with QM and crystallog. data showed that an addnl. optimization of the .vphi., ψ dihedral parameters along with the grid-based energy correction were required to yield significant improvements over the CHARMM22 force field. However, systematic deviations in the treatment of .vphi. and ψ in the helical and sheet regions were evident. Accordingly, empirical adjustments were made to the grid-based energy correction for alanine and glycine to account for these systematic differences. These adjustments lead to greater deviations from QM data for the two dipeptides but also yielded improved agreement with exptl. crystallog. data. These improvements enhance the quality of the CHARMM force field in treating proteins. This extension of the potential energy function is anticipated to facilitate improved treatment of biol. macromols. via MM approaches in general.
- 39McCleverty, C. J., Koesema, E., Patapoutian, A., Lesley, S. A., and Kreusch, A. (2006) Crystal structure of the human TRPV2 channel ankyrin repeat domain Protein Sci. 15, 2201– 2206Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xpt1ejsbc%253D&md5=c31904fbd84d2334c3a2b276a603344aCrystal structure of the human TRPV2 channel ankyrin repeat domainMcCleverty, Clare J.; Koesema, Eric; Patapoutian, Ardem; Lesley, Scott A.; Kreusch, AndreasProtein Science (2006), 15 (9), 2201-2206CODEN: PRCIEI; ISSN:0961-8368. (Cold Spring Harbor Laboratory Press)TRPV channels are important polymodal integrators of noxious stimuli mediating thermosensation and nociception. An ankyrin repeat domain (ARD), which is a common protein-protein recognition domain, is conserved in the N-terminal intracellular domain of all TRPV channels and predicted to contain 3-4 ankyrin repeats. Here, the authors report the 1st structure from the TRPV channel subfamily, a 1.7-Å resoln. crystal structure of human TRPV2 ARD. The crystal structure revealed a 6-ankyrin repeat stack with multiple insertions in each repeat generating several unique features compared with a canonical ARD. The surface typically used for ligand recognition, the ankyrin groove, contained extended loops with an exposed hydrophobic patch and a prominent kink resulting from a large rotational shift of the last 2 repeats. The TRPV2 ARD provided the 1st structural insight into a domain that coordinates nociceptive sensory transduction and is likely to be a prototype for other TRPV channel ARDs.
- 40Phelps, C. B., Huang, R. J., Lishko, P. V., Wang, R. R., and Gaudet, R. (2008) Structural analyses of the ankyrin repeat domain of TRPV6 and related TRPV ion channels Biochemistry 47, 2476– 2484Google ScholarThere is no corresponding record for this reference.
- 41Croy, C. H., Bergqvist, S., Huxford, T., Ghosh, G., and Komives, E. A. (2004) Biophysical characterization of the free IκBα ankyrin repeat domain in solution Protein Sci. 13, 1767– 1777Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXlsFWktLw%253D&md5=de0d4921c20309c8f99199b5233a252fBiophysical characterization of the free IκBα ankyrin repeat domain in solutionCroy, Carrie Hughes; Bergqvist, Simon; Huxford, Tom; Ghosh, Gourisankar; Komives, Elizabeth A.Protein Science (2004), 13 (7), 1767-1777CODEN: PRCIEI; ISSN:0961-8368. (Cold Spring Harbor Laboratory Press)The crystal structure of IκBα in complex with the transcription factor, nuclear factor κ-B (NF-κB) shows six ankyrin repeats, which are all ordered. Electron d. was not obsd. for most of the residues within the PEST sequence, although it is required for high-affinity binding. To characterize the folded state of IκBα (67-317) when it is not in complex with NF-κB, we have carried out CD spectroscopy, 8-anilino-1-naphthalenesulfonic acid (ANS) binding, differential scanning calorimetry, and amide hydrogen/deuterium exchange expts. The CD spectrum shows the presence of helical structure, consistent with other ankyrin repeat proteins. The large amt. of ANS-binding and amide exchange suggest that the protein may have molten globule character. The amide exchange expts. show that the third ankyrin repeat is the most compact, the second and fourth repeats are somewhat less compact, and the first and sixth repeats are solvent exposed. The PEST extension is also highly solvent accessible. IκBα unfolds with a Tm of 42°, and forms a sol. aggregate that sequesters helical and variable loop parts of the first, fourth, and sixth repeats and the PEST extension. The second and third repeats, which conform most closely to a consensus for stable ankyrin repeats, appear to remain outside of the aggregate. The ramifications of these observations for the biol. function of IκBα are discussed.
- 42Truhlar, S. M., Torpey, J. W., and Komives, E. A. (2006) Regions of IκBα that are critical for its inhibition of NF-κB·DNA interaction fold upon binding to NF-κB Proc. Natl. Acad. Sci. U.S.A. 103, 18951– 18956Google ScholarThere is no corresponding record for this reference.
- 43Barrick, D., Ferreiro, D. U., and Komives, E. A. (2008) Folding landscapes of ankyrin repeat proteins: Experiments meet theory Curr. Opin. Struct. Biol. 18, 27– 34Google Scholar43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXitVGitrY%253D&md5=f9e5a9b6eb66598ba45fe81cddedb69bFolding landscapes of ankyrin repeat proteins: Experiments meet theoryBarrick, Doug; Ferreiro, Diego U.; Komives, Elizabeth A.Current Opinion in Structural Biology (2008), 18 (1), 27-34CODEN: COSBEF; ISSN:0959-440X. (Elsevier B.V.)A review. Nearly 6% of eukaryotic protein sequences contain ankyrin repeat (AR) domains, which consist of several repeats and often function in binding. AR proteins show highly cooperative folding despite a lack of long-range contacts. Both theory and expt. converge to explain that formation of the interface between elements is more favorable than formation of any individual repeat unit. IκBα and Notch both undergo partial folding upon binding perhaps influencing the binding free energy. The simple architecture, combined with identification of consensus residues that are important for stability, has enabled systematic perturbation of the energy landscape by single point mutations that affect stability or by addn. of consensus repeats. The folding energy landscapes appear highly plastic, with small perturbations re-routing folding pathways.
- 44Salazar, H., Llorente, I., Jara-Oseguera, A., Garcia-Villegas, R., Munari, M., Gordon, S. E., Islas, L. D., and Rosenbaum, T. (2008) A single N-terminal cysteine in TRPV1 determines activation by pungent compounds from onion and garlic Nat. Neurosci. 11, 255– 261Google Scholar44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXisFOru74%253D&md5=b7b57fe9b43ab99748bb95084f193161A single N-terminal cysteine in TRPV1 determines activation by pungent compounds from onion and garlicSalazar, Hector; Llorente, Itzel; Jara-Oseguera, Andres; Garcia-Villegas, Refugio; Munari, Mika; Gordon, Sharona E.; Islas, Leon D.; Rosenbaum, TamaraNature Neuroscience (2008), 11 (3), 255-261CODEN: NANEFN; ISSN:1097-6256. (Nature Publishing Group)Some members of the transient receptor potential (TRP) family of cation channels mediate sensory responses to irritant substances. Although it is well known that TRPA1 channels are activated by pungent compds. found in garlic, onion, mustard and cinnamon exts., activation of TRPV1 by these exts. remains controversial. Here the authors establish that TRPV1 is activated by pungent exts. from onion and garlic, as well as by allicin, the active compd. in these prepns., and participates together with TRPA1 in the pain-related behavior induced by this compd. The authors found that in TRPV1 these agents act by covalent modification of cysteine residues. In contrast to TRPA1 channels, modification of a single cysteine located in the N-terminal region of TRPV1 was necessary and sufficient for all the effects the authors obsd. The authors findings point to a conserved mechanism of activation in TRP channels, which provides new insights into the mol. basis of noxious stimuli detection.
- 45Isralewitz, B., Gao, M., and Schulten, K. (2001) Steered molecular dynamics and mechanical functions of proteins Curr. Opin. Struct. Biol. 11, 224– 230Google Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXjsVyisro%253D&md5=09b7a191f8ffa19d316ec0f4421c3731Steered molecular dynamics and mechanical functions of proteinsIsralewitz, Barry; Gao, Mu; Schulten, KlausCurrent Opinion in Structural Biology (2001), 11 (2), 224-230CODEN: COSBEF; ISSN:0959-440X. (Elsevier Science Ltd.)A review with 50 refs. At. force microscopy of single mols., steered mol. dynamics and the theory of stochastic processes have established a new field that investigates mech. functions of proteins, such as ligand-receptor binding/unbinding and elasticity of muscle proteins during stretching. The combination of these methods yields information on the energy landscape that controls mech. function and on the force-bearing components of proteins, as well as on the underlying phys. mechanisms.
- 46Camacho, N., Krakow, D., Johnykutty, S., Katzman, P. J., Pepkowitz, S., Vriens, J., Nilius, B., Boyce, B. F., and Cohn, D. H. (2010) Dominant TRPV4 mutations in nonlethal and lethal metatropic dysplasia Am. J. Med. Genet., Part A 152A, 1169– 1177Google Scholar46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXnsFylurg%253D&md5=00b52e12ff7d4602a43cf7270dc2384bDominant TRPV4 mutations in nonlethal and lethal metatropic dysplasiaCamacho, Natalia; Krakow, Deborah; Johnykutty, Sharlin; Katzman, Philip J.; Pepkowitz, Samuel; Vriens, Joris; Nilius, Bernd; Boyce, Brendan F.; Cohn, Daniel H.American Journal of Medical Genetics, Part A (2010), 152A (5), 1169-1177CODEN: AJMGB8; ISSN:1552-4825. (Wiley-Liss, Inc.)Metatropic dysplasia is a clin. heterogeneous skeletal dysplasia characterized by short extremities, a short trunk with progressive kyphoscoliosis, and craniofacial abnormalities that include a prominent forehead, midface hypoplasia, and a squared-off jaw. Dominant mutations in the gene encoding TRPV4, a calcium permeable ion channel, were identified all 10 of a series of metatropic dysplasia cases, ranging in severity from mild to perinatal lethal. These data demonstrate that the lethal form of the disorder is dominantly inherited and suggest locus homogeneity in the disease. Electrophysiol. studies demonstrated that the mutations activate the channel, indicating that the mechanism of disease may result from increased calcium in chondrocytes. Histol. studies in two cases of lethal metatropic dysplasia revealed markedly disrupted endochondral ossification, with reduced nos. of hypertrophic chondrocytes and presence of islands of cartilage within the zone of primary mineralization. These data suggest that altered chondrocyte differentiation in the growth plate leads to the clin. findings in metatropic dysplasia.
- 47Dai, J., Kim, O. H., Cho, T. J., Schmidt-Rimpler, M., Tonoki, H., Takikawa, K., Haga, N., Miyoshi, K., Kitoh, H., Yoo, W. J., Choi, I. H., Song, H. R., Jin, D. K., Kim, H. T., Kamasaki, H., Bianchi, P., Grigelioniene, G., Nampoothiri, S., Minagawa, M., Miyagawa, S. I., Fukao, T., Marcelis, C., Jansweijer, M. C., Hennekam, R. C., Bedeschi, F., Mustonen, A., Jiang, Q., Ohashi, H., Furuichi, T., Unger, S., Zabel, B., Lausch, E., Superti-Furga, A., Nishimura, G., and Ikegawa, S. (2010) Novel and recurrent TRPV4 mutations and their association with distinct phenotypes within the TRPV4 dysplasia family J. Med. Genet. 47, 704– 709Google Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhsVags7nM&md5=565af1fbc841fd2fa39201dccd9c26caNovel and recurrent TRPV4 mutations and their association with distinct phenotypes within the TRPV4 dysplasia familyDai, J.; Kim, O.-H.; Cho, T.-J.; Schmidt-Rimpler, M.; Tonoki, H.; Takikawa, K.; Haga, N.; Miyoshi, K.; Kitoh, H.; Yoo, W.-J.; Choi, I.-H.; Song, H.-R.; Jin, D.-K.; Kim, H.-T.; Kamasaki, H.; Bianchi, P.; Grigelioniene, G.; Nampoothiri, S.; Minagawa, M.; Miyagawa, S.-i.; Fukao, T.; Marcelis, C.; Jansweijer, M. C. E.; Hennekam, R. C. M.; Bedeschi, F.; Mustonen, A.; Jiang, Q.; Ohashi, H.; Furuichi, T.; Unger, S.; Zabel, B.; Lausch, E.; Superti-Furga, A.; Nishimura, G.; Ikegawa, S.Journal of Medical Genetics (2010), 47 (10), 704-709CODEN: JMDGAE; ISSN:0022-2593. (BMJ Publishing Group)Mutations in TRPV4, a gene that encodes a Ca2+ permeable non-selective cation channel, have recently been found in a spectrum of skeletal dysplasias that includes brachyolmia, spondylometaphyseal dysplasia, Kozlowski type (SMDK) and metatropic dysplasia (MD). Only a total of seven missense mutations were detected, however. The full spectrum of TRPV4 mutations and their phenotypes remained unclear. To examine TRPV4 mutation spectrum and phenotype-genotype assocn., we searched for TRPV4 mutations by PCR-direct sequencing from genomic DNA in 22 MD and 20 SMDK probands. TRPV4 mutations were found in all but one MD subject. In total, 19 different heterozygous mutations were identified in 41 subjects; two were recurrent and 17 were novel. In MD, a recurrent P799L mutation was identified in nine subjects, as well as 10 novel mutations including F471 del, the first deletion mutation of TRPV4. In SMDK, a recurrent R594H mutation was identified in 12 subjects and seven novel mutations. An assocn. between the position of mutations and the disease phenotype was also obsd. Thus, P799 in exon 15 is a hot codon for MD mutations, as four different amino acid substitutions have been obsd. at this codon; while R594 in exon 11 is a hotspot for SMDK mutations. The TRPV4 mutation spectrum in MD and SMDK, which showed genotype-phenotype correlation and potential functional significance of mutations that are non-randomly distributed over the gene, was presented in this study. The results would help diagnostic labs. establish efficient screening strategies for genetic diagnosis of the TRPV4 dysplasia family diseases.
- 48Nishimura, G., Dai, J., Lausch, E., Unger, S., Megarbane, A., Kitoh, H., Kim, O. H., Cho, T. J., Bedeschi, F., Benedicenti, F., Mendoza-Londono, R., Silengo, M., Schmidt-Rimpler, M., Spranger, J., Zabel, B., Ikegawa, S., and Superti-Furga, A. (2010) Spondylo-epiphyseal dysplasia, Maroteaux type (pseudo-Morquio syndrome type 2), and parastremmatic dysplasia are caused by TRPV4 mutations Am. J. Med. Genet., Part A 152A, 1443– 1449Google Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC3czmsVKrsg%253D%253D&md5=fc3a83e2d0d91659a246956bbb8a42daSpondylo-epiphyseal dysplasia, Maroteaux type (pseudo-Morquio syndrome type 2), and parastremmatic dysplasia are caused by TRPV4 mutationsNishimura Gen; Dai Jin; Lausch Ekkehart; Unger Sheila; Megarbane Andre; Kitoh Hiroshi; Kim Ok Hwa; Cho Tae-Joon; Bedeschi Francesca; Benedicenti Francesco; Mendoza-Londono Roberto; Silengo Margherita; Schmidt-Rimpler Maren; Spranger Jurgen; Zabel Bernhard; Ikegawa Shiro; Superti-Furga AndreaAmerican journal of medical genetics. Part A (2010), 152A (6), 1443-9 ISSN:.Recent discoveries have established the existence of a family of skeletal dysplasias caused by dominant mutations in TRPV4. This family comprises, in order of increasing severity, dominant brachyolmia, spondylo-metaphyseal dysplasia Kozlowski type, and metatropic dysplasia. We tested the hypothesis that a further condition, Spondylo-epiphyseal dysplasia (SED), Maroteaux type (MIM 184095; also known as pseudo-Morquio syndrome type 2), could be caused by TRPV4 mutations. We analyzed six individuals with Maroteaux type SED, including three who had previously been reported. All six patients were found to have heterozygous TRPV4 mutations; three patients had unreported mutations, while three patients had mutations previously described in association with metatropic dysplasia. In addition, we tested one individual with a distinct rare disorder, parastremmatic dysplasia (MIM 168400). This patient had a common, recurrent mutation seen in several patients with Kozlowski type spondylo-metaphyseal dysplasia. We conclude that SED Maroteaux type and parastremmatic dysplasia are part of the TRPV4 dysplasia family and that TRPV4 mutations show considerable variability in phenotypic expression resulting in distinct clinical-radiographic phenotypes.
- 49Zimon, M., Baets, J., Auer-Grumbach, M., Berciano, J., Garcia, A., Lopez-Laso, E., Merlini, L., Hilton-Jones, D., McEntagart, M., Crosby, A. H., Barisic, N., Boltshauser, E., Shaw, C. E., Landoure, G., Ludlow, C. L., Gaudet, R., Houlden, H., Reilly, M. M., Fischbeck, K. H., Sumner, C. J., Timmerman, V., Jordanova, A., and Jonghe, P. D. (2010) Dominant mutations in the cation channel gene transient receptor potential vanilloid 4 cause an unusual spectrum of neuropathies Brain 133, 1798– 1809Google ScholarThere is no corresponding record for this reference.
- 50Clapham, D. E. and Miller, C. (2011) A thermodynamic framework for understanding temperature sensing by transient receptor potential (TRP) channels Proc. Natl. Acad. Sci. U.S.A. 108, 19492– 19497Google ScholarThere is no corresponding record for this reference.
- 51Dai, J., Cho, T. J., Unger, S., Lausch, E., Nishimura, G., Kim, O. H., Superti-Furga, A., and Ikegawa, S. (2010) TRPV4-pathy, a novel channelopathy affecting diverse systems J. Hum. Genet. 55, 400– 402Google Scholar51https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC3cnps1ajtw%253D%253D&md5=37b381dc5fe3485dc27ef8fde886164eTRPV4-pathy, a novel channelopathy affecting diverse systemsDai Jin; Cho Tae-Joon; Unger Sheila; Lausch Ekkehart; Nishimura Gen; Kim Ok-Hwa; Superti-Furga Andrea; Ikegawa ShiroJournal of human genetics (2010), 55 (7), 400-2 ISSN:.Transient receptor potential cation channel, subfamily V, member 4 (TRPV4) is a calcium-permeable nonselective cation channel of unknown biological function. TRPV4 mutation was first identified in brachyolmia, and then in a spectrum of autosomal-dominant skeletal dysplasias, which includes Kozlowski type of spondylometaphyseal dysplasia, metatropic dysplasia, Maroteaux type of spondyloepiphyseal dysplasia and parastremmatic dysplasia. Recently, TRPV4 mutation has also been identified in a spectrum of neuromuscular diseases that includes congenital distal spinal muscular atrophy (SMA), scapuloperoneal SMA, and hereditary motor and sensory neuropathy type IIC. These diverse spectrums of diseases compose a novel channelopathy, TRPV4-pathy, which could further include polygenic traits such as serum sodium concentration and a chronic obstructive pulmonary disease. In this review, we clarified the TRPV4 mutation spectrum, and discussed the phenotypic complexity of TRPV4-pathy and its pathogenic mechanisms. TRPV4-pathy may extend further to other monogenic and polygenic diseases.
- 52Andreucci, E., Aftimos, S., Alcausin, M., Haan, E., Hunter, W., Kannu, P., Kerr, B., McGillivray, G., Gardner, R. M., Patricelli, M. G., Sillence, D., Thompson, E., Zacharin, M., Zankl, A., Lamande, S. R., and Savarirayan, R. (2011) TRPV4 related skeletal dysplasias: A phenotypic spectrum highlighted byclinical, radiographic, and molecular studies in 21 new families Orphanet. J. Rare Dis. 6, 37Google Scholar52https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC3MnnvVSmuw%253D%253D&md5=b06a8af5d8f9d02931e0300f3bb90cd8TRPV4 related skeletal dysplasias: a phenotypic spectrum highlighted byclinical, radiographic, and molecular studies in 21 new familiesAndreucci Elena; Aftimos Salim; Alcausin Melanie; Haan Eric; Hunter Warwick; Kannu Peter; Kerr Bronwyn; McGillivray George; McKinlay Gardner R J; Patricelli Maria G; Sillence David; Thompson Elizabeth; Zacharin Margaret; Zankl Andreas; Lamande Shireen R; Savarirayan RaviOrphanet journal of rare diseases (2011), 6 (), 37 ISSN:.BACKGROUND: The TRPV4 gene encodes a calcium-permeable ion-channel that is widely expressed, responds to many different stimuli and participates in an extraordinarily wide range of physiologic processes. Autosomal dominant brachyolmia, spondylometaphyseal dysplasia Kozlowski type (SMDK) and metatropic dysplasia (MD) are currently considered three distinct skeletal dysplasias with some shared clinical features, including short stature, platyspondyly, and progressive scoliosis. Recently, TRPV4 mutations have been found in patients diagnosed with these skeletal phenotypes. METHODS AND RESULTS: We critically analysed the clinical and radiographic data on 26 subjects from 21 families, all of whom had a clinical diagnosis of one of the conditions described above: 15 with MD; 9 with SMDK; and 2 with brachyolmia. We sequenced TRPV4 and identified 9 different mutations in 22 patients, 4 previously described, and 5 novel. There were 4 mutation-negative cases: one with MD and one with SMDK, both displaying atypical clinical and radiographic features for these diagnoses; and two with brachyolmia, who had isolated spine changes and no metaphyseal involvement. CONCLUSIONS: Our data suggest the TRPV4 skeletal dysplasias represent a continuum of severity with areas of phenotypic overlap, even within the same family. We propose that AD brachyolmia lies at the mildest end of this spectrum and, since all cases described with this diagnosis and TRPV4 mutations display metaphyseal changes, we suggest that it is not a distinct entity but represents the mildest phenotypic expression of SMDK.
- 53Klein, C. J., Shi, Y., Fecto, F., Donaghy, M., Nicholson, G., McEntagart, M. E., Crosby, A. H., Wu, Y., Lou, H., McEvoy, K. M., Siddique, T., Deng, H. X., and Dyck, P. J. (2011) TRPV4 mutations and cytotoxic hypercalcemia in axonal Charcot-Marie-Tooth neuropathies Neurology 76, 887– 894Google ScholarThere is no corresponding record for this reference.
- 54Fecto, F., Shi, Y., Huda, R., Martina, M., Siddique, T., and Deng, H. X. (2011) Mutant TRPV4-mediated toxicity is linked to increased constitutive function in axonal neuropathies J. Biol. Chem. 286, 17281– 17291Google ScholarThere is no corresponding record for this reference.
- 55Loukin, S., Su, Z., and Kung, C. (2011) Increased basal activity is a key determinant in the severity of human skeletal dysplasia caused by TRPV4 mutations PLoS One 6, e19533Google ScholarThere is no corresponding record for this reference.
- 56Schaefer, M. (2005) Homo- and heteromeric assembly of TRP channel subunits Pfluegers Arch. 451, 35– 42Google Scholar56https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXhtVKqu7rF&md5=86c1dc77b33044bb46a302f596fbe758Homo- and heteromeric assembly of TRP channel subunitsSchaefer, MichaelPfluegers Archiv (2005), 451 (1), 35-42CODEN: PFLABK; ISSN:0031-6768. (Springer GmbH)A review. Mammalian homologs of the Drosophila melanogaster transient receptor potential (TRP) channels are the 2nd largest cation channel family within the superfamily of hexahelical cation channels. Most mammalian TRP channels function as homo-oligomers and mediate monovalent or divalent cation entry upon activation by a variety of stimuli. Because native TRP channels may be multimeric proteins of possibly complex compn., it has been difficult to compare cation conductances in native tissues to those of clearly defined homomeric TRP channel complexes in living cells. Therefore, the possibility of heteromeric TRP channel assembly has been investigated in recent years by several groups. As a major conclusion of these studies, most heteromeric TRP channel complexes appear to consist of subunit combinations only within relatively narrow confines of phylogenetic subfamilies. Although the general capability of heteromer formation between closely related TRP channel subunits is now clearly established, investigators are only beginning to understand whether these heteromeric complexes are of physiol. significance. Here, the author summarizes current knowledge on the promiscuity and specificity of the assembly of channel complexes composed of TRPC-, TRPV-, and TRPM-subunits of mammalian TRP channels.
Cited By
This article is cited by 69 publications.
- Xiaopeng Fan, Renjian Xie, Wenjing Song, Kunfu Ouyang, Li Ren. Biomimetic Hyaluronic Acid-Based Brush Polymers Modulate Chondrocyte Homeostasis via ROS/Ca2+/TRPV4. Biomacromolecules 2023, 24
(9)
, 4240-4252. https://doi.org/10.1021/acs.biomac.3c00547
- Jacob K. Hilton, Parthasarathi Rath, Cole V. M. Helsell, Oliver Beckstein, and Wade D. Van Horn . Understanding Thermosensitive Transient Receptor Potential Channels as Versatile Polymodal Cellular Sensors. Biochemistry 2015, 54
(15)
, 2401-2413. https://doi.org/10.1021/acs.biochem.5b00071
- Miao Zhang, Yueming Ma, Xianglu Ye, Ning Zhang, Lei Pan, Bing Wang. TRP (transient receptor potential) ion channel family: structures, biological functions and therapeutic interventions for diseases. Signal Transduction and Targeted Therapy 2023, 8
(1)
https://doi.org/10.1038/s41392-023-01464-x
- Haofeng Chen, Chong Sun, Yongsheng Zheng, Junxiong Yin, Mingshi Gao, Chongbo Zhao, Jie Lin. A TRPV4 mutation caused Charcot-Marie-Tooth disease type 2C with scapuloperoneal muscular atrophy overlap syndrome and scapuloperoneal spinal muscular atrophy in one family: a case report and literature review. BMC Neurology 2023, 23
(1)
https://doi.org/10.1186/s12883-023-03260-0
- Benedikt Goretzki, Christoph Wiedemann, Brett A. McCray, Stefan L. Schäfer, Jasmin Jansen, Frederike Tebbe, Sarah-Ana Mitrovic, Julia Nöth, Ainara Claveras Cabezudo, Jack K. Donohue, Cy M. Jeffries, Wieland Steinchen, Florian Stengel, Charlotte J. Sumner, Gerhard Hummer, Ute A. Hellmich. Crosstalk between regulatory elements in disordered TRPV4 N-terminus modulates lipid-dependent channel activity. Nature Communications 2023, 14
(1)
https://doi.org/10.1038/s41467-023-39808-4
- Lin Niu, Yu‐jie Lu, Xing‐wang Zu, Wen Yang, Fu‐kui Shen, Yan‐yan Xu, Min Jiang, Yang Xie, Su‐yun Li, Jie Gao, Gang Bai. Magnolol alleviates pulmonary fibrosis inchronic obstructive pulmonary disease by targeting transient receptor potential vanilloid 4‐ankyrin repeat domain. Phytotherapy Research 2023, 37
(9)
, 4282-4297. https://doi.org/10.1002/ptr.7907
- Aiqin Mao, Dongxu He, Ka Zhang, Mengru Gao, Hao Kan, Zhiwei Wang, Chunlei Tang, Xin Ma. Functional role of coupling the endothelial TRPV4 and K
Ca
3.1 channels in regulating coronary vascular tone. British Journal of Pharmacology 2023, 180
(17)
, 2266-2279. https://doi.org/10.1111/bph.16082
- A. F. Murtazina, P. N. Tsabay, G. E. Rudenskaya, L. A. Bessonova, F. M. Bostanova, D. M. Guseva, I. V. Sharkova, O. A. Shchagina, A. A. Orlova, O. P. Ryzhkova, T. V. Markova, A. S. Kuchina, S. S. Nikitin, E. L. Dadali. Phenotypic variability in TRPV4-associated neuropathies and neuronopathies: a case series. Neuromuscular Diseases 2023, 13
(2)
, 42-55. https://doi.org/10.17650/2222-8721-2023-13-2-42-55
- Sébastien Chaigne, Solène Barbeau, Thomas Ducret, Romain Guinamard, David Benoist. Pathophysiological Roles of the TRPV4 Channel in the Heart. Cells 2023, 12
(12)
, 1654. https://doi.org/10.3390/cells12121654
- Xin Wen, Yidi Peng, Bohao Zheng, Shaying Yang, Jing Han, Fan Yu, Tingting Zhou, Li Geng, Zhiming Yu, Lei Feng. Silybin induces endothelium-dependent vasodilation via TRPV4 channels in mouse mesenteric arteries. Hypertension Research 2022, 45
(12)
, 1954-1963. https://doi.org/10.1038/s41440-022-01000-4
- Benedikt Goretzki, Frederike Tebbe, Sarah-Ana Mitrovic, Ute A. Hellmich. Backbone NMR assignments of the extensive human and chicken TRPV4 N-terminal intrinsically disordered regions as important players in ion channel regulation. Biomolecular NMR Assignments 2022, 16
(2)
, 205-212. https://doi.org/10.1007/s12104-022-10080-9
- Ezequiel A. Galpern, Jacopo Marchi, Thierry Mora, Aleksandra M. Walczak, Diego U. Ferreiro. Evolution and folding of repeat proteins. Proceedings of the National Academy of Sciences 2022, 119
(31)
https://doi.org/10.1073/pnas.2204131119
- Anna M. Bagnell, Charlotte J. Sumner, Brett A. McCray. TRPV4: A trigger of pathological RhoA activation in neurological disease. BioEssays 2022, 44
(6)
https://doi.org/10.1002/bies.202100288
- Di Zhang, Bei Jing, Xin Li, Huimei Shi, Zhenni Chen, Shiquan Chang, Yachun Zheng, Yi Lin, Yuwei Pan, Jianxin Sun, Guoping Zhao. Antihyperalgesic Effect of Paeniflorin Based on Chronic Constriction Injury in Rats. Revista Brasileira de Farmacognosia 2022, 32
(3)
, 375-385. https://doi.org/10.1007/s43450-022-00251-z
- Maria V. Yelshanskaya, Alexander I. Sobolevsky. Ligand-Binding Sites in Vanilloid-Subtype TRP Channels. Frontiers in Pharmacology 2022, 13 https://doi.org/10.3389/fphar.2022.900623
- William H. Aisenberg, Brett A. McCray, Jeremy M. Sullivan, Erika Diehl, Lauren R. DeVine, Jonathan Alevy, Anna M. Bagnell, Patrice Carr, Jack K. Donohue, Benedikt Goretzki, Robert N. Cole, Ute A. Hellmich, Charlotte J. Sumner. Multiubiquitination of TRPV4 reduces channel activity independent of surface localization. Journal of Biological Chemistry 2022, 298
(4)
, 101826. https://doi.org/10.1016/j.jbc.2022.101826
- Na Liu, Weiwei Lu, Xiaolin Dai, Xiaowen Qu, Chongtao Zhu. The role of TRPV channels in osteoporosis. Molecular Biology Reports 2022, 49
(1)
, 577-585. https://doi.org/10.1007/s11033-021-06794-z
- Swapnil K. Sonkusare, Victor E. Laubach. Endothelial TRPV4 channels in lung edema and injury. 2022, 43-62. https://doi.org/10.1016/bs.ctm.2022.07.001
- Brett A. McCray, Erika Diehl, Jeremy M. Sullivan, William H. Aisenberg, Nicholas W. Zaccor, Alexander R. Lau, Dominick J. Rich, Benedikt Goretzki, Ute A. Hellmich, Thomas E. Lloyd, Charlotte J. Sumner. Neuropathy-causing TRPV4 mutations disrupt TRPV4-RhoA interactions and impair neurite extension. Nature Communications 2021, 12
(1)
https://doi.org/10.1038/s41467-021-21699-y
- Chenfan Ji, Christopher A. McCulloch. TRPV4 integrates matrix mechanosensing with Ca
2+
signaling to regulate extracellular matrix remodeling. The FEBS Journal 2021, 288
(20)
, 5867-5887. https://doi.org/10.1111/febs.15665
- Miao Chen, Xiucun Li. Role of TRPV4 channel in vasodilation and neovascularization. Microcirculation 2021, 28
(6)
https://doi.org/10.1111/micc.12703
- Trine L. Toft-Bertelsen, Nanna MacAulay. TRPing to the Point of Clarity: Understanding the Function of the Complex TRPV4 Ion Channel. Cells 2021, 10
(1)
, 165. https://doi.org/10.3390/cells10010165
- Tingting Zhou, Zhiwei Wang, Mengting Guo, Ka Zhang, Li Geng, Aiqin Mao, Yujiao Yang, Fan Yu. Puerarin induces mouse mesenteric vasodilation and ameliorates hypertension involving endothelial TRPV4 channels. Food & Function 2020, 11
(11)
, 10137-10148. https://doi.org/10.1039/D0FO02356F
- Osama F. Harraz, David Hill-Eubanks, Mark T. Nelson. PIP
2
: A critical regulator of vascular ion channels hiding in plain sight. Proceedings of the National Academy of Sciences 2020, 117
(34)
, 20378-20389. https://doi.org/10.1073/pnas.2006737117
- Ruth A. Pumroy, Edwin C. Fluck, Tofayel Ahmed, Vera Y. Moiseenkova-Bell. Structural insights into the gating mechanisms of TRPV channels. Cell Calcium 2020, 87 , 102168. https://doi.org/10.1016/j.ceca.2020.102168
- Ernesto Ladrón-de-Guevara, Laura Dominguez, Gisela E. Rangel-Yescas, Daniel A. Fernández-Velasco, Alfredo Torres-Larios, Tamara Rosenbaum, Leon D. Islas. The Contribution of the Ankyrin Repeat Domain of TRPV1 as a Thermal Module. Biophysical Journal 2020, 118
(4)
, 836-845. https://doi.org/10.1016/j.bpj.2019.10.041
- Alexandra E. Hochstetler, Makenna M. Reed, Bonnie L. Blazer-Yost. TRPV4, a Regulatory Channel in the Production of Cerebrospinal Fluid by the Choroid Plexus. 2020, 173-191. https://doi.org/10.1007/978-1-0716-0536-3_7
- Deng Li, Tsu-Hsin Kao, Shu-Wei Chang. The structural changes of the mutated ankyrin repeat domain of the human TRPV4 channel alter its ATP binding ability. Journal of the Mechanical Behavior of Biomedical Materials 2020, 101 , 103407. https://doi.org/10.1016/j.jmbbm.2019.103407
- Peng Gao, Li Li, Xiao Wei, Miao Wang, Yangning Hong, Hao Wu, Yanjia Shen, Tianyi Ma, Xing Wei, Qin Zhang, Xia Fang, Lijuan Wang, Zhencheng Yan, Guan-Hua Du, Hongting Zheng, Gangyi Yang, Daoyan Liu, Zhiming Zhu. Activation of Transient Receptor Potential Channel Vanilloid 4 by DPP-4 (Dipeptidyl Peptidase-4) Inhibitor Vildagliptin Protects Against Diabetic Endothelial Dysfunction. Hypertension 2020, 75
(1)
, 150-162. https://doi.org/10.1161/HYPERTENSIONAHA.119.13778
- Jing Shao, Jing Han, Yifei Zhu, Aiqin Mao, Zhiwei Wang, Ka Zhang, Xiaodong Zhang, Yufeng Zhang, Chunlei Tang, Xin Ma. Curcumin Induces Endothelium-Dependent Relaxation by Activating Endothelial TRPV4 Channels. Journal of Cardiovascular Translational Research 2019, 12
(6)
, 600-607. https://doi.org/10.1007/s12265-019-09928-8
- Masakazu Atobe. Activation of Transient Receptor Potential Vanilloid (TRPV) 4 as a Therapeutic Strategy in Osteoarthritis. Current Topics in Medicinal Chemistry 2019, 19
(24)
, 2254-2267. https://doi.org/10.2174/1568026619666191010162850
- Dawn Hayward, Valentina L. Kouznetsova, Hannah E. Pierson, Nesrin M. Hasan, Estefany R. Guzman, Igor F. Tsigelny, Svetlana Lutsenko. ANKRD9 is a metabolically-controlled regulator of IMPDH2 abundance and macro-assembly. Journal of Biological Chemistry 2019, 294
(39)
, 14454-14466. https://doi.org/10.1074/jbc.RA119.008231
- Wenjun Zheng, Han Wen. Heat activation mechanism of TRPV1: New insights from molecular dynamics simulation. Temperature 2019, 6
(2)
, 120-131. https://doi.org/10.1080/23328940.2019.1578634
- Jose Velilla, Michael Mario Marchetti, Agnes Toth-Petroczy, Claire Grosgogeat, Alexis H. Bennett, Nikkola Carmichael, Elicia Estrella, Basil T. Darras, Natasha Y. Frank, Joel Krier, Rachelle Gaudet, Vandana A. Gupta. Homozygous
TRPV4
mutation causes congenital distal spinal muscular atrophy and arthrogryposis. Neurology Genetics 2019, 5
(2)
, e312. https://doi.org/10.1212/NXG.0000000000000312
- Limeng Pu, Rajiv Gandhi Govindaraj, Jeffrey Mitchell Lemoine, Hsiao-Chun Wu, Michal Brylinski, . DeepDrug3D: Classification of ligand-binding pockets in proteins with a convolutional neural network. PLOS Computational Biology 2019, 15
(2)
, e1006718. https://doi.org/10.1371/journal.pcbi.1006718
- Benedikt Goretzki, Nina A. Glogowski, Erika Diehl, Elke Duchardt-Ferner, Carolin Hacker, Rachelle Gaudet, Ute A. Hellmich. Structural Basis of TRPV4 N Terminus Interaction with Syndapin/PACSIN1-3 and PIP2. Structure 2018, 26
(12)
, 1583-1593.e5. https://doi.org/10.1016/j.str.2018.08.002
- Zeyaul Islam, Raghavendra Sashi Krishna Nagampalli, Munazza Tamkeen Fatima, Ghulam Md Ashraf. New paradigm in ankyrin repeats: Beyond protein-protein interaction module. International Journal of Biological Macromolecules 2018, 109 , 1164-1173. https://doi.org/10.1016/j.ijbiomac.2017.11.101
- Zengqin Deng, Navid Paknejad, Grigory Maksaev, Monica Sala-Rabanal, Colin G. Nichols, Richard K. Hite, Peng Yuan. Cryo-EM and X-ray structures of TRPV4 reveal insight into ion permeation and gating mechanisms. Nature Structural & Molecular Biology 2018, 25
(3)
, 252-260. https://doi.org/10.1038/s41594-018-0037-5
- Han Wen, Wenjun Zheng. Decrypting the Heat Activation Mechanism of TRPV1 Channel by Molecular Dynamics Simulation. Biophysical Journal 2018, 114
(1)
, 40-52. https://doi.org/10.1016/j.bpj.2017.10.034
- Gongxin Wang, KeWei Wang. The Ca
2+
-Permeable Cation Transient Receptor Potential TRPV3 Channel: An Emerging Pivotal Target for Itch and Skin Diseases. Molecular Pharmacology 2017, 92
(3)
, 193-200. https://doi.org/10.1124/mol.116.107946
- . Molecular Mechanisms of Temperature Gating in TRP Channels. 2017, 11-26. https://doi.org/10.1201/9781315152837-3
- Fan Yang, Jie Zheng. Understand spiciness: mechanism of TRPV1 channel activation by capsaicin. Protein & Cell 2017, 8
(3)
, 169-177. https://doi.org/10.1007/s13238-016-0353-7
- Han Wen, Feng Qin, Wenjun Zheng. Toward elucidating the heat activation mechanism of the TRPV1 channel gating by molecular dynamics simulation. Proteins: Structure, Function, and Bioinformatics 2016, 84
(12)
, 1938-1949. https://doi.org/10.1002/prot.25177
- Radim Mazanec, Jana Neupauerová, Daniel Baumgartner, Veronika Potočková, Petra Laššuthová, Dana Šafka-Brožková, Pavel Seeman. Hereditary motor neuropathies. Neurologie pro praxi 2016, 17
(6)
, 354-358. https://doi.org/10.36290/neu.2016.074
- B. Gandolfi, S. Alamri, W.G. Darby, B. Adhikari, J.C. Lattimer, R. Malik, C.M. Wade, L.A. Lyons, J. Cheng, J.F. Bateman, P. McIntyre, S.R. Lamandé, B. Haase. A dominant TRPV4 variant underlies osteochondrodysplasia in Scottish fold cats. Osteoarthritis and Cartilage 2016, 24
(8)
, 1441-1450. https://doi.org/10.1016/j.joca.2016.03.019
- John P. M. White, Mario Cibelli, Laszlo Urban, Bernd Nilius, J. Graham McGeown, Istvan Nagy. TRPV4: Molecular Conductor of a Diverse Orchestra. Physiological Reviews 2016, 96
(3)
, 911-973. https://doi.org/10.1152/physrev.00016.2015
- Jeremy M. Sullivan, Christina M. Zimanyi, William Aisenberg, Breanne Bears, Dong-Hui Chen, John W. Day, Thomas D. Bird, Carly E. Siskind, Rachelle Gaudet, Charlotte J. Sumner. Novel mutations highlight the key role of the ankyrin repeat domain in
TRPV4
-mediated neuropathy. Neurology Genetics 2015, 1
(4)
, e29. https://doi.org/10.1212/NXG.0000000000000029
- Francisco J. Taberner, Gregorio Fernández-Ballester, Asia Fernández-Carvajal, Antonio Ferrer-Montiel. TRP channels interaction with lipids and its implications in disease. Biochimica et Biophysica Acta (BBA) - Biomembranes 2015, 1848
(9)
, 1818-1827. https://doi.org/10.1016/j.bbamem.2015.03.022
- Wenjun Zheng, Feng Qin. A combined coarse-grained and all-atom simulation of TRPV1 channel gating and heat activation. Journal of General Physiology 2015, 145
(5)
, 443-456. https://doi.org/10.1085/jgp.201411335
- Balázs István Tóth, Bernd Nilius. Transient Receptor Potential Dysfunctions in Hereditary Diseases. 2015, 13-33. https://doi.org/10.1016/B978-0-12-420024-1.00002-3
- Nicholas A. Veldhuis, Daniel P. Poole, Megan Grace, Peter McIntyre, Nigel W. Bunnett, . The G Protein–Coupled Receptor–Transient Receptor Potential Channel Axis: Molecular Insights for Targeting Disorders of Sensation and Inflammation. Pharmacological Reviews 2015, 67
(1)
, 36-73. https://doi.org/10.1124/pr.114.009555
- Shiro SUETSUGU, Nobuaki TAKAHASHI, Yuzuru ITOH, Kazuhiro TAKEMURA, Atsushi SHIMADA, Akio KITAO, Yasuo MORI. TRPV4 Channel Activity Is Modulated by Direct Interaction of the Ankyrin Domain to PI(4,5)P2. Seibutsu Butsuri 2015, 55
(5)
, 262-265. https://doi.org/10.2142/biophys.55.262
- Nobuaki Takahashi, Sayaka Hamada-Nakahara, Yuzuru Itoh, Kazuhiro Takemura, Atsushi Shimada, Yoshifumi Ueda, Manabu Kitamata, Rei Matsuoka, Kyoko Hanawa-Suetsugu, Yosuke Senju, Masayuki X. Mori, Shigeki Kiyonaka, Daisuke Kohda, Akio Kitao, Yasuo Mori, Shiro Suetsugu. TRPV4 channel activity is modulated by direct interaction of the ankyrin domain to PI(4,5)P2. Nature Communications 2014, 5
(1)
https://doi.org/10.1038/ncomms5994
- Pau Doñate-Macián, Alex Perálvarez-Marín, . Dissecting Domain-Specific Evolutionary Pressure Profiles of Transient Receptor Potential Vanilloid Subfamily Members 1 to 4. PLoS ONE 2014, 9
(10)
, e110715. https://doi.org/10.1371/journal.pone.0110715
- Kevin W. Huynh, Matthew R. Cohen, Vera Y. Moiseenkova-Bell. Application of Amphipols for Structure–Functional Analysis of TRP Channels. The Journal of Membrane Biology 2014, 247
(9-10)
, 843-851. https://doi.org/10.1007/s00232-014-9684-6
- S. Duchatelet, L. Guibbal, S. Veer, S. Fraitag, P. Nitschké, M. Zarhrate, C. Bodemer, A. Hovnanian. Olmsted syndrome with erythromelalgia caused by recessive transient receptor potential vanilloid 3 mutations. British Journal of Dermatology 2014, 171
(3)
, 675-678. https://doi.org/10.1111/bjd.12951
- Bernd Nilius, Arpad Szallasi, . Transient Receptor Potential Channels as Drug Targets: From the Science of Basic Research to the Art of Medicine. Pharmacological Reviews 2014, 66
(3)
, 676-814. https://doi.org/10.1124/pr.113.008268
- Kevin W. Huynh, Matthew R. Cohen, Sudha Chakrapani, Heather A. Holdaway, Phoebe L. Stewart, Vera Y. Moiseenkova-Bell. Structural Insight into the Assembly of TRPV Channels. Structure 2014, 22
(2)
, 260-268. https://doi.org/10.1016/j.str.2013.11.008
- Ute A. Hellmich, Rachelle Gaudet. Structural Biology of TRP Channels. 2014, 963-990. https://doi.org/10.1007/978-3-319-05161-1_10
- Bernd Nilius, Veit Flockerzi. What Do We Really Know and What Do We Need to Know: Some Controversies, Perspectives, and Surprises. 2014, 1239-1280. https://doi.org/10.1007/978-3-319-05161-1_20
- Jeremy M. Sullivan, Thomas E. Lloyd, Charlotte J. Sumner. Hereditary Channelopathies Caused by TRPV4 Mutations. 2014, 413-440. https://doi.org/10.1007/978-3-642-40282-1_21
- Pu Yang, Michael X. Zhu. TRPV3. 2014, 273-291. https://doi.org/10.1007/978-3-642-54215-2_11
- Matthew R. Cohen, Vera Y. Moiseenkova-Bell. Structure of Thermally Activated TRP Channels. 2014, 181-211. https://doi.org/10.1016/B978-0-12-800181-3.00007-5
- Kristin K. Jernigan, Seth R. Bordenstein. Ankyrin domains across the Tree of Life. PeerJ 2014, 2 , e264. https://doi.org/10.7717/peerj.264
- Maofu Liao, Erhu Cao, David Julius, Yifan Cheng. Structure of the TRPV1 ion channel determined by electron cryo-microscopy. Nature 2013, 504
(7478)
, 107-112. https://doi.org/10.1038/nature12822
- Di-Jing Shi, Sheng Ye, Xu Cao, Rongguang Zhang, KeWei Wang. Crystal structure of the N-terminal ankyrin repeat domain of TRPV3 reveals unique conformation of finger 3 loop critical for channel function. Protein & Cell 2013, 4
(12)
, 942-950. https://doi.org/10.1007/s13238-013-3091-0
- Anna Garcia-Elias, Sanela Mrkonjić, Carlos Pardo-Pastor, Hitoshi Inada, Ute A. Hellmich, Fanny Rubio-Moscardó, Cristina Plata, Rachelle Gaudet, Rubén Vicente, Miguel A. Valverde. Phosphatidylinositol-4,5-biphosphate-dependent rearrangement of TRPV4 cytosolic tails enables channel activation by physiological stimuli. Proceedings of the National Academy of Sciences 2013, 110
(23)
, 9553-9558. https://doi.org/10.1073/pnas.1220231110
- Bernd Nilius, Thomas Voets. The puzzle of TRPV4 channelopathies. EMBO reports 2013, 14
(2)
, 152-163. https://doi.org/10.1038/embor.2012.219
- Sang Sun Kang, Sung Hwa Shin, Chung-Kyoon Auh, Jaesun Chun. Human skeletal dysplasia caused by a constitutive activated transient receptor potential vanilloid 4 (TRPV4) cation channel mutation. Experimental & Molecular Medicine 2012, 44
(12)
, 707. https://doi.org/10.3858/emm.2012.44.12.080
Abstract
Figure 1
Figure 1. Structural comparison of human and chicken TRPV4-ARDs. (A) Superimposed ribbon diagrams of ATP-bound (magenta) and ATP-free (blue) hTRPV4-ARD. ATP is shown as sticks. (B) Superimposed Cα traces of human and chicken TRPV4-ARD. Finger 3 is twisted and shrunken in the ATP-bound (magenta) and ATP-unbound (green) hTRPV4-ARD structures, while the finger is extended in ATP-free hTRPV4-ARD (blue) and cTRPV4-ARD (gray). Several Finger 3 residues are disordered in three of six TRPV4-ARD structures. The structure of Finger 2 in the ATP-bound and -unbound forms differs from that in ATP-free forms.
Figure 2
Figure 2. Aromatic residues on Fingers 2 and 3 have varied positions in hTRPV4-ARD structures. (A) The hTRPV4-ARD structures of ATP-bound (magenta), ATP-unbound (green), and ATP-free (blue) forms are superimposed. Aromatic residues are shown as sticks. (B) Detail of the Finger 2 and 3 loops. F272 and F273 on Finger 3 (black rectangles) are embedded in the aromatic cluster in the ATP-bound and -unbound forms but exposed in the ATP-free form. Y235 and Y236 on Finger 2 and Y281 and F282 on Finger 3 are located in similar positions but show variable orientations. F231, F282, Y283, and F284 show less variation.
Figure 3
Figure 3. Structural comparison of TRPV-ARDs. (A) Superimposed main chain structure of TRPV-ARDs (rTRPV1-ARD, gray; rat and human TRPV2-ARD, cyan; ATP-bound hTRPV4-ARD, magenta; ATP-unbound hTRPV4-ARD, green; ATP-free hTRPV4-ARD, blue; and mouse TRPV6-ARD, black). Finger 3 and a part of Finger 2 are highly flexible. Several residues on Finger 3 are missing in one of two TRPV1-ARD structures and four of seven TRPV2-ARD structures. (B) ATP-binding site of hTRPV4-ARD and rTRPV1-ARD. Residues (sticks) within 4 Å of the ATP molecule and a surface map of the ATP-binding site in hTRPV4-ARD (left) and the corresponding residues in the rTRPV1-ARD ATP-binding site (right). The bBound ATP molecule is shown as sticks (orange and yellow). (C) Finger 2 (2) and Finger 3 (3) structures of ATP-bound rat TRPV1-ARD (gray), rat and human TRPV2-ARD (cyan), human ATP-bound TRPV4-ARD (magenta), and mouse TRPV6-ARD (black). (D) Aromatic residue positioned behind the adenine base of ATP in Finger 2 (F231 in human TRPV4-ARD). (E) This aromatic residue is conserved in TRPV-ARDs that bind ATP (red rectangle). (F and G) ATP-agarose pull-down assays for wild-type and mutant rTRPV1-ARD (F) or hTRPV4-ARD (G). Coomassie-stained gels (top) of wild-type and mutant proteins loaded (left) and bound to ATP-agarose in the absence (middle) or presence (right) of competing free ATP. The normalized intensity of protein recovered (mean ± SD; n = 3) is plotted below. The statistical significance of the change in binding to ATP-agarose with respect to the wild type (WT) was determined by a multiple-comparison test using Dunnett’s method, with p < 0.01 indicated by an asterisk.
Figure 4
Figure 4. Effect of ATP on hTRPV4-ARD thermal stability. (A) Representative circular dichroism spectra of the purified TRPV4-ARD protein (3.4 μM) in the presence of ATP, AMP, or phosphate (1 mM each) at 10 °C. The wavelength (λ) of 222 nm used for thermostability assays is indicated by a vertical red line. (B) Representative traces of the thermostability assay. The molar ellipticity at 222 nm was measured as the protein solutions were heated at a rate of 1 °C/min. (C) Tm of TRPV4-ARD in the presence of 1 mM ATP, AMP, or phosphate. The statistical significance of the change in Tm was determined by a multiple-comparison test using the Tukey–Kramer method, with p < 0.05 and p < 0.01 indicated by one asterisk and two asterisks, respectively.
Figure 5
Figure 5. Effect of ATP binding on protein stability in rTRPV1-ARD. (A) Structure of TRPV1-ARD (gray) bound to ATP (green, sticks), with buried Cys157 highlighted (spheres). (B) TRPV1-ARD is modified at cysteine residues by PEG-maleimide (mPEG), causing an electrophoretic mobility shift on a Coomassie-stained SDS gel. Abbreviations: WT, wild type; CL, a cysteine-less variant; C157, CL-TRPV1-ARD C157 single-cysteine variant. Shown is a representative Coomassie-stained gel from one of three experiments. (C) Time course for modification of single-cysteine TRPV1-ARD variants C157 and C362 with 0.5 mM mPEG at room temperature. (D) Data from four experiments like that depicted in panel C were quantified, and the mean ± standard deviation was plotted. (E and F) Molecular dynamics simulation in which the termini of the ATP-bound TRPV1-ARD (E) or TRPV1-ARD structure with ATP removed prior to equilibrating the system (F) are pulled apart at a rate of 20 nm/ns. Superimposed are the structures at the start (gold) and end (blue) of the simulations. (G and H) Root-mean-square deviation of each Cα atom over the course of the simulation mapped onto the starting models with (G) or without (H) ATP. The change in color from blue to red indicates changes in rmsd from 0 to 80 Å. Simulations in which the termini were pulled apart at a rate of 2 nm/ns gave similar results. See Table S3 of the Supporting Information for experimental details.
Figure 6
Figure 6. Mutations associated with human diseases in hTRPV4-ARD. (A) Positions of mutations associated with human inherited diseases that lie within hTRPV4-ARD. Abbreviations: SEDM, spondyloepiphyseal dysplasia, type Maroteaux; SMDK, spondylometaphyseal dysplasia, type Kozolowski; MD, metatropic dysplasia; SMA, spinal muscular atrophy; SPMA, scapuloperoneal spinal muscular atrophy; CMTC2, Charcot-Marie-Tooth disease type 2C; HMSN2C, hereditary motor and sensory neuropathy 2C. This figure was inspired by ref 51. (B) Location of the disease-causing mutations within TRPV4-ARD. Shown as spheres are 12 residue positions at which a total of 15 mutations causing human inherited diseases have been identified. The ATP molecule is shown as sticks. Skeletal dysplasia and neurophathy mutations are depicted as green and blue spheres, respectively. (C) Leu199 is located at the hydrophobic interface between ANK2 and ANK3. (D) Glu183 and Arg232 form a salt bridge on the convex face of TRPV4-ARD.
Figure 7
Figure 7. Thermal stability and ATP binding of hTRPV4-ARD mutants associated with inherited diseases. (A) The Tm determined by CD spectrometry in a phosphate-based buffer is plotted for wild-type and mutant hTRPV4-ARDs. The statistical significance is shown in Table S4 of the Supporting Information. (B) Coomassie-stained gels show wild-type and mutant TRPV4-ARDs loaded (top) and bound to ATP-agarose (bottom). (C) Normalized intensity of recovered protein (mean ± SD; n = 3). The statistical significance of the change in binding to ATP-agarose with respect to wild type (WT) was determined by a multiple-comparison test using Dunnett’s method, with p < 0.05 and p < 0.01 indicated by one asterisk and two asterisks, respectively.
References
ARTICLE SECTIONSThis article references 56 other publications.
- 1Clapham, D. E. (2003) TRP channels as cellular sensors Nature 426, 517– 524Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXpsVejtrc%253D&md5=f45fe3448869a0e1c8623290df1cd218TRP channels as cellular sensorsClapham, David E.Nature (London, United Kingdom) (2003), 426 (6966), 517-524CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)A review. TRP channels are the vanguard of our sensory systems, responding to temp., touch, pain, osmolarity, pheromones, taste and other stimuli. But their role is much broader than classical sensory transduction. They are an ancient sensory app. for the cell, not just the multicellular organism, and they have been adapted to respond to all manner of stimuli, from both within and outside the cell.
- 2Clapham, D. E., Montell, C., Schultz, G., and Julius, D. (2003) International Union of Pharmacology. XLIII. Compendium of voltage-gated ion channels: Transient receptor potential channels Pharmacol. Rev. 55, 591– 596Google ScholarThere is no corresponding record for this reference.
- 3Wu, L. J., Sweet, T. B., and Clapham, D. E. (2010) International Union of Basic and Clinical Pharmacology. LXXVI. Current progress in the mammalian TRP ion channel family Pharmacol. Rev. 62, 381– 404Google ScholarThere is no corresponding record for this reference.
- 4Tominaga, M. and Tominaga, T. (2005) Structure and function of TRPV1 Pfluegers Arch. 451, 143– 150Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXhtVKqu7rE&md5=35d28025b866fe7e955bd2861dd5a9aaStructure and function of TRPV1Tominaga, Makoto; Tominaga, TomokoPfluegers Archiv (2005), 451 (1), 143-150CODEN: PFLABK; ISSN:0031-6768. (Springer GmbH)A review. Capsaicin, the main ingredient in hot chili peppers, elicits a sensation of burning pain by selectively activating sensory neurons that convey information about noxious stimuli to the central nervous system. The capsaicin receptor, transient receptor potential vanilloid 1 (TRPV1), is predicted to have 6 transmembrane (TM) domains and a short, pore-forming hydrophobic stretch between the 5th and 6th TM domains, and is activated not only by capsaicin but also by heat (>43°), acid, and various lipids. Within the TPRV1 protein, many regions and amino acids involved in specific functions (multimerization, capsaicin action, proton action, heat activation, desensitization, permeability, phosphorylation and modulation by lipids) have been identified since the cloning in 1997. Given the fact that TRPV1 is a key mol. in peripheral nociception, these regions and amino acids could prove useful for the development of novel anti-nociceptive or anti-inflammatory agents.
- 5van de Graaf, S. F., Hoenderop, J. G., and Bindels, R. J. (2006) Regulation of TRPV5 and TRPV6 by associated proteins Am. J. Physiol. 290, F1295– F1302Google ScholarThere is no corresponding record for this reference.
- 6Rosenbaum, T., Gordon-Shaag, A., Munari, M., and Gordon, S. E. (2004) Ca2+/calmodulin modulates TRPV1 activation by capsaicin J. Gen. Physiol. 123, 53– 62Google ScholarThere is no corresponding record for this reference.
- 7Lishko, P. V., Procko, E., Jin, X., Phelps, C. B., and Gaudet, R. (2007) The ankyrin repeats of TRPV1 bind multiple ligands and modulate channel sensitivity Neuron 54, 905– 918Google ScholarThere is no corresponding record for this reference.
- 8Phelps, C. B., Wang, R. R., Choo, S. S., and Gaudet, R. (2010) Differential regulation of TRPV1, TRPV3, and TRPV4 sensitivity through a conserved binding site on the ankyrin repeat domain J. Biol. Chem. 285, 731– 740Google ScholarThere is no corresponding record for this reference.
- 9Prescott, E. D. and Julius, D. (2003) A modular PIP2 binding site as a determinant of capsaicin receptor sensitivity Science 300, 1284– 1288Google ScholarThere is no corresponding record for this reference.
- 10Ufret-Vincenty, C. A., Klein, R. M., Hua, L., Angueyra, J., and Gordon, S. E. (2011) Localization of the PIP2 sensor of TRPV1 ion channels J. Biol. Chem. 286, 9688– 9698Google ScholarThere is no corresponding record for this reference.
- 11Numazaki, M., Tominaga, T., Takeuchi, K., Murayama, N., Toyooka, H., and Tominaga, M. (2003) Structural determinant of TRPV1 desensitization interacts with calmodulin Proc. Natl. Acad. Sci. U.S.A. 100, 8002– 8006Google ScholarThere is no corresponding record for this reference.
- 12Liedtke, W., Choe, Y., Marti-Renom, M. A., Bell, A. M., Denis, C. S., Sali, A., Hudspeth, A. J., Friedman, J. M., and Heller, S. (2000) Vanilloid receptor-related osmotically activated channel (VR-OAC), a candidate vertebrate osmoreceptor Cell 103, 525– 535Google ScholarThere is no corresponding record for this reference.
- 13Guler, A. D., Lee, H., Iida, T., Shimizu, I., Tominaga, M., and Caterina, M. (2002) Heat-evoked activation of the ion channel, TRPV4 J. Neurosci. 22, 6408– 6414Google ScholarThere is no corresponding record for this reference.
- 14Watanabe, H., Vriens, J., Suh, S. H., Benham, C. D., Droogmans, G., and Nilius, B. (2002) Heat-evoked activation of TRPV4 channels in a HEK293 cell expression system and in native mouse aorta endothelial cells J. Biol. Chem. 277, 47044– 47051Google ScholarThere is no corresponding record for this reference.
- 15Watanabe, H., Vriens, J., Prenen, J., Droogmans, G., Voets, T., and Nilius, B. (2003) Anandamide and arachidonic acid use epoxyeicosatrienoic acids to activate TRPV4 channels Nature 424, 434– 438Google ScholarThere is no corresponding record for this reference.
- 16Liedtke, W. (2005) TRPV4 as osmosensor: A transgenic approach Pfluegers Arch. 451, 176– 180Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXhtVKqu73E&md5=ea165f301f59cf2468ae28e73ae08ae9TRPV4 as osmosensor: a transgenic approachLiedtke, WolfgangPfluegers Archiv (2005), 451 (1), 176-180CODEN: PFLABK; ISSN:0031-6768. (Springer GmbH)A review. The transient receptor potential vanilloid 4 (TRPV4) ion channel was named initially vanilloid-receptor-related osmotically activated channel (VR-OAC). Preliminary answers to the question, "What is the function of the trpv4 gene in live animals " are highlighted briefly in this review. In trpv4 null mice, TRPV4 is necessary for the maintenance of osmotic equil., and in Caenorhabditis elegans transgenic for mammalian TRPV4, TRPV4 directs the osmotic avoidance response in the context of the ASH "nociceptive" neuron. The mol. mechanisms of gating of TRPV4 in vivo need to be detd.; in particular, whether TRPV4 in live animals is gated via phosphorylation of defined amino-acid residues or more directly through the osmotic stimulus itself.
- 17Tominaga, M. and Caterina, M. J. (2004) Thermosensation and pain J. Neurobiol. 61, 3– 12Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BD2cvlt12hsQ%253D%253D&md5=ac03f5366520d918056a6c30efee8850Thermosensation and painTominaga Makoto; Caterina Michael JJournal of neurobiology (2004), 61 (1), 3-12 ISSN:0022-3034.We feel a wide range of temperatures spanning from cold to heat. Within this range, temperatures over about 43 degrees C and below about 15 degrees C evoke not only a thermal sensation, but also a feeling of pain. In mammals, six thermosensitive ion channels have been reported, all of which belong to the TRP (transient receptor potential) superfamily. These include TRPV1 (VR1), TRPV2 (VRL-1), TRPV3, TRPV4, TRPM8 (CMR1), and TRPA1 (ANKTM1). These channels exhibit distinct thermal activation thresholds (>43 degrees C for TRPV1, >52 degrees C for TRPV2, > approximately 34-38 degrees C for TRPV3, > approximately 27-35 degrees C for TRPV4, < approximately 25-28 degrees C for TRPM8 and <17 degrees C for TRPA1), and are expressed in primary sensory neurons as well as other tissues. The involvement of TRPV1 in thermal nociception has been demonstrated by multiple methods, including the analysis of TRPV1-deficient mice. TRPV2, TRPM8, and TRPA1 are also very likely to be involved in thermal nociception, because their activation thresholds are within the noxious range of temperatures.
- 18Mochizuki, T., Sokabe, T., Araki, I., Fujishita, K., Shibasaki, K., Uchida, K., Naruse, K., Koizumi, S., Takeda, M., and Tominaga, M. (2009) The TRPV4 cation channel mediates stretch-evoked Ca2+ influx and ATP release in primary urothelial cell cultures J. Biol. Chem. 284, 21257– 21264Google ScholarThere is no corresponding record for this reference.
- 19Sokabe, T. and Tominaga, M. (2010) The TRPV4 cation channel: A molecule linking skin temperature and barrier function Commun. Integr. Biol. 3, 619– 621Google Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC3M3gsVGmug%253D%253D&md5=090f2a24c4f290f61a7f74a22fa4f301The TRPV4 cation channel: A molecule linking skin temperature and barrier functionSokabe Takaaki; Tominaga MakotoCommunicative & integrative biology (2010), 3 (6), 619-21 ISSN:.The skin barrier function is indispensable for terrestrial animals to avoid dehydration. The function is achieved by a hydrophobic cornified layer consisting of dead keratinocytes and lipids, and by an intercellular junction barrier formed among differentiated keratinocytes. A recent report demonstrated that TRPV4, one of the temperature-sensitive cation channels, contributes to the formation and maintenance of the intercellular junction-dependent barrier in the skin. TRPV4 associates with the E-cadherin complex via β-catenin, and thereby participates in the promotion of cell-cell junction development. TRPV4 allows influx of Ca(2+) ions from the extracellular space at physiological skin temperatures. The Ca(2+) influx induces Rho activation and promotes actin fiber organization and junction formation, thereby augmenting barrier integrity. Indeed, the intercellular junction structures and the skin barrier function were impaired in TRPV4-deficeint mice. This novel role of TRPV4 in keratinocytes may explain the significant correlation between temperature and the condition of skin.>
- 20Shibasaki, K., Suzuki, M., Mizuno, A., and Tominaga, M. (2007) Effects of body temperature on neural activity in the hippocampus: Regulation of resting membrane potentials by transient receptor potential vanilloid 4 J. Neurosci. 27, 1566– 1575Google Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXitl2gu78%253D&md5=f2d68caf39b30e807957c8cca4f9f76cEffects of body temperature on neural activity in the hippocampus: regulation of resting membrane potentials by transient receptor potential vanilloid 4Shibasaki, Koji; Suzuki, Makoto; Mizuno, Atsuko; Tominaga, MakotoJournal of Neuroscience (2007), 27 (7), 1566-1575CODEN: JNRSDS; ISSN:0270-6474. (Society for Neuroscience)Physiol. body temp. is an important determinant for neural functions, and it is well established that changes in temp. have dynamic influences on hippocampal neural activities. However, the detailed mol. mechanisms have never been clarified. Here, we show that hippocampal neurons express functional transient receptor potential vanilloid 4 (TRPV4), one of the thermosensitive TRP (transient receptor potential) channels, and that TRPV4 is constitutively active at physiol. temp. Activation of TRPV4 at 37°C depolarized the resting membrane potential in hippocampal neurons by allowing cation influx, which was obsd. in wild-type (WT) neurons, but not in TRPV4-deficient (TRPV4KO) cells, although dendritic morphol., synaptic marker clustering, and synaptic currents were indistinguishable between the two genotypes. Furthermore, current injection studies revealed that TRPV4KO neurons required larger depolarization to evoke firing, equiv. to WT neurons, indicating that TRPV4 is a key regulator for hippocampal neural excitabilities. We conclude that TRPV4 is activated by physiol. temp. in hippocampal neurons and thereby controls their excitability.
- 21Alessandri-Haber, N., Dina, O. A., Joseph, E. K., Reichling, D., and Levine, J. D. (2006) A transient receptor potential vanilloid 4-dependent mechanism of hyperalgesia is engaged by concerted action of inflammatory mediators J. Neurosci. 26, 3864– 3874Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xjs1yksbo%253D&md5=30220a30e99775236cd7876c1b121d7bA transient receptor potential vanilloid 4-dependent mechanism of hyperalgesia is engaged by concerted action of inflammatory mediatorsAlessandri-Haber, Nicole; Dina, Olayinka A.; Joseph, Elizabeth K.; Reichling, David; Levine, Jon D.Journal of Neuroscience (2006), 26 (14), 3864-3874CODEN: JNRSDS; ISSN:0270-6474. (Society for Neuroscience)The transient receptor potential vanilloid 4 (TRPV4) is a primary afferent transducer that plays a crucial role in neuropathic hyperalgesia for osmotic and mech. stimuli, as well as in inflammatory mediator-induced hyperalgesia for osmotic stimuli. In view of the clin. importance of mech. hyperalgesia in inflammatory states, the present study investigated the role of TRPV4 in mech. hyperalgesia induced by inflammatory mediators and the second-messenger pathways involved. Intradermal injection of either the inflammogen carrageenan or a soup of inflammatory mediators enhanced the nocifensive paw-withdrawal reflex elicited by hypotonic or mech. stimuli in rat. Spinal administration of TRPV4 antisense oligodeoxynucleotide blocked the enhancement without altering baseline nociceptive threshold. Similarly, in TRPV4-/- knock-out mice, inflammatory soup failed to induce any significant mech. or osmotic hyperalgesia. In vitro investigation showed that inflammatory mediators engage the TRPV4-mediated mechanism of sensitization by direct action on dissocd. primary afferent neurons. Addnl. behavioral observations suggested that multiple mediators are necessary to achieve sufficient activation of the cAMP pathway to engage the TRPV4-dependent mechanism of hyperalgesia. In addn., direct activation of protein kinase A or protein kinase C ε, two pathways that mediate inflammation-induced mech. hyperalgesia, also induced hyperalgesia for both hypotonic and mech. stimuli that was decreased by TRPV4 antisense and absent in TRPV4-/- mice. We conclude that TRPV4 plays a crucial role in the mech. hyperalgesia that is generated by the concerted action of inflammatory mediators present in inflamed tissues.
- 22Masuyama, R., Vriens, J., Voets, T., Karashima, Y., Owsianik, G., Vennekens, R., Lieben, L., Torrekens, S., Moermans, K., Vanden Bosch, A., Bouillon, R., Nilius, B., and Carmeliet, G. (2008) TRPV4-mediated calcium influx regulates terminal differentiation of osteoclasts Cell Metab. 8, 257– 265Google Scholar22https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXhtFeju77E&md5=a6987e7efdaad7228132430519f13e7aTRPV4-mediated calcium influx regulates terminal differentiation of osteoclastsMasuyama, Ritsuko; Vriens, Joris; Voets, Thomas; Karashima, Yuji; Owsianik, Grzegorz; Vennekens, Rudi; Lieben, Liesbet; Torrekens, Sophie; Moermans, Karen; Vanden Bosch, An; Bouillon, Roger; Nilius, Bernd; Carmeliet, GeertCell Metabolism (2008), 8 (3), 257-265CODEN: CMEEB5; ISSN:1550-4131. (Cell Press)Calcium signaling controls multiple cellular functions and is regulated by the release from internal stores and entry from extracellular fluid. In bone, osteoclast differentiation is induced by RANKL (receptor activator of NF-κB ligand)-evoked intracellular Ca2+ oscillations, which trigger nuclear factor-activated T cells (NFAT) c1-responsive gene transcription. However, the Ca2+ channels involved remain largely unidentified. Here we show that genetic ablation in mice of Trpv4, a Ca2+-permeable channel of the transient receptor potential (TRP) family, increases bone mass by impairing bone resorption. TRPV4 mediates basolateral Ca2+ influx specifically in large osteoclasts when Ca2+ oscillations decline. TRPV4-mediated Ca2+ influx hereby secures intracellular Ca2+ concns., ensures NFATc1-regulated gene transcription, and regulates the terminal differentiation and activity of osteoclasts. In conclusion, our data indicate that Ca2+ oscillations and TRPV4-mediated Ca2+ influx are sequentially required to sustain NFATc1-dependent gene expression throughout osteoclast differentiation, and we propose TRPV4 as a therapeutic target for bone diseases.
- 23Nilius, B. and Owsianik, G. (2010) Transient receptor potential channelopathies Pfluegers Arch. 460, 437– 450Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXntlGns70%253D&md5=e1f538ce79e09a627cec9424d2c45fdbTransient receptor potential channelopathiesNilius, Bernd; Owsianik, GrzegorzPfluegers Archiv (2010), 460 (2), 437-450CODEN: PFLABK; ISSN:0031-6768. (Springer GmbH)A review. In the past years, several hereditary diseases caused by defects in transient receptor potential channels (TRP) genes have been described. This review summarizes our current knowledge about TRP channelopathies and their possible pathomechanisms. Based on available genetic indications, we will also describe several putative pathol. conditions in which (mal)function of TRP channels could be anticipated.
- 24Verma, P., Kumar, A., and Goswami, C. (2010) TRPV4-mediated channelopathies Channels 4, 319– 328Google ScholarThere is no corresponding record for this reference.
- 25Rock, M. J., Prenen, J., Funari, V. A., Funari, T. L., Merriman, B., Nelson, S. F., Lachman, R. S., Wilcox, W. R., Reyno, S., Quadrelli, R., Vaglio, A., Owsianik, G., Janssens, A., Voets, T., Ikegawa, S., Nagai, T., Rimoin, D. L., Nilius, B., and Cohn, D. H. (2008) Gain-of-function mutations in TRPV4 cause autosomal dominant brachyolmia Nat. Genet. 40, 999– 1003Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXptVKku7g%253D&md5=33da69ac6e23dbe409b9c1807c9f504dGain-of-function mutations in TRPV4 cause autosomal dominant brachyolmiaRock, Matthew J.; Prenen, Jean; Funari, Vincent A.; Funari, Tara L.; Merriman, Barry; Nelson, Stanley F.; Lachman, Ralph S.; Wilcox, William R.; Reyno, Soraya; Quadrelli, Roberto; Vaglio, Alicia; Owsianik, Grzegorz; Janssens, Annelies; Voets, Thomas; Ikegawa, Shiro; Nagai, Toshiro; Rimoin, David L.; Nilius, Bernd; Cohn, Daniel H.Nature Genetics (2008), 40 (8), 999-1003CODEN: NGENEC; ISSN:1061-4036. (Nature Publishing Group)Daniel Cohn and colleagues identify mutations in the gene encoding the calcium-permeable cation channel TRPV4 in families with autosomal dominant brachyolmia. Functional studies show that the mutations result in gain-of-function of channel activation. Daniel Cohn and colleagues identify mutations in the gene encoding the calcium-permeable cation channel TRPV4 in families with autosomal dominant brachyolmia. Functional studies show that the mutations result in gain-of-function of channel activation. The brachyolmias constitute a clin. and genetically heterogeneous group of skeletal dysplasias characterized by a short trunk, scoliosis and mild short stature. Here, we identify a locus for an autosomal dominant form of brachyolmia on chromosome 12q24.1-12q24.2. Among the genes in the genetic interval, we selected TRPV4, which encodes a calcium permeable cation channel of the transient receptor potential (TRP) vanilloid family, as a candidate gene because of its cartilage-selective gene expression pattern. In two families with the phenotype, we identified point mutations in TRPV4 that encoded R616Q and V620I substitutions, resp. Patch clamp studies of transfected HEK cells showed that both mutations resulted in a dramatic gain of function characterized by increased constitutive activity and elevated channel activation by either mechano-stimulation or agonist stimulation by arachidonic acid or the TRPV4-specific agonist 4α-phorbol 12,13-didecanoate (4αPDD). This study thus defines a previously unknown mechanism, activation of a calcium-permeable TRP ion channel, in skeletal dysplasia pathogenesis.
- 26Krakow, D., Vriens, J., Camacho, N., Luong, P., Deixler, H., Funari, T. L., Bacino, C. A., Irons, M. B., Holm, I. A., Sadler, L., Okenfuss, E. B., Janssens, A., Voets, T., Rimoin, D. L., Lachman, R. S., Nilius, B., and Cohn, D. H. (2009) Mutations in the gene encoding the calcium-permeable ion channel TRPV4 produce spondylometaphyseal dysplasia, Kozlowski type and metatropic dysplasia Am. J. Hum. Genet. 84, 307– 315Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXmvFCmu7c%253D&md5=f0e29ebe74f51d9bd0539e0c2ae805d4Mutations in the gene encoding the calcium-permeable ion channel TRPV4 produce spondylometaphyseal dysplasia, Kozlowski type and metatropic dysplasiaKrakow, Deborah; Vriens, Joris; Camacho, Natalia; Luong, Phi; Deixler, Hannah; Funari, Tara L.; Bacino, Carlos A.; Irons, Mira B.; Holm, Ingrid A.; Sadler, Laurie; Okenfuss, Ericka B.; Janssens, Annelies; Voets, Thomas; Rimoin, David L.; Lachman, Ralph S.; Nilius, Bernd; Cohn, Daniel H.American Journal of Human Genetics (2009), 84 (3), 307-315CODEN: AJHGAG; ISSN:0002-9297. (Cell Press)The spondylometaphyseal dysplasias (SMDs) are a group of short-stature disorders distinguished by abnormalities in the vertebrae and the metaphyses of the tubular bones. SMD Kozlowski type (SMDK) is a well-defined autosomal-dominant SMD characterized by significant scoliosis and mild metaphyseal abnormalities in the pelvis. The vertebrae exhibit platyspondyly and overfaced pedicles similar to autosomal-dominant brachyolmia, which can result from heterozygosity for activating mutations in the gene encoding TRPV4, a calcium-permeable ion channel. Mutation anal. in six out of six patients with SMDK demonstrated heterozygosity for missense mutations in TRPV4, and one mutation, predicting a R594H substitution, was recurrent in four patients. Similar to autosomal-dominant brachyolmia, the mutations altered basal calcium channel activity in vitro. Metatropic dysplasia is another SMD that has been proposed to have both clin. and genetic heterogeneity. Patients with the nonlethal form of metatropic dysplasia present with a progressive scoliosis, widespread metaphyseal involvement of the appendicular skeleton, and carpal ossification delay. Because of some similar radiog. features between SMDK and metatropic dysplasia, TRPV4 was tested as a disease gene for nonlethal metatropic dysplasia. In two sporadic cases, heterozygosity for de novo missense mutations in TRPV4 was found. The findings demonstrate that mutations in TRPV4 produce a phenotypic spectrum of skeletal dysplasias from the mild autosomal-dominant brachyolmia to SMDK to autosomal-dominant metatropic dysplasia, suggesting that these disorders should be grouped into a new bone dysplasia family.
- 27Deng, H. X., Klein, C. J., Yan, J., Shi, Y., Wu, Y., Fecto, F., Yau, H. J., Yang, Y., Zhai, H., Siddique, N., Hedley-Whyte, E. T., Delong, R., Martina, M., Dyck, P. J., and Siddique, T. (2010) Scapuloperoneal spinal muscular atrophy and CMT2C are allelic disorders caused by alterations in TRPV4 Nat. Genet. 42, 165– 169Google Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhs1Skur7F&md5=ebf469c8e33fcf491b99b00619b7e9adScapuloperoneal spinal muscular atrophy and CMT2C are allelic disorders caused by alterations in TRPV4Deng, Han-Xiang; Klein, Christopher J.; Yan, Jian-Hua; Shi, Yong; Wu, Yan-Hong; Fecto, Faisal; Yau, Hau-Jie; Yang, Yi; Zhai, Hong; Siddique, Nailah; Hedley-Whyte, E. Tessa; De Long, Robert; Martina, Marco; Dyck, Peter J.; Siddique, TeepuNature Genetics (2010), 42 (2), 165-169CODEN: NGENEC; ISSN:1061-4036. (Nature Publishing Group)Scapuloperoneal spinal muscular atrophy (SPSMA) and hereditary motor and sensory neuropathy type IIC (HMSN IIC, also known as HMSN2C or Charcot-Marie-Tooth disease type 2C (CMT2C)) are phenotypically heterogeneous disorders involving topog. distinct nerves and muscles. We originally described a large New England family of French-Canadian origin with SPSMA and an American family of English and Scottish descent with CMT2C. We mapped SPSMA and CMT2C risk loci to 12q24.1-q24.31 with an overlapping region between the two diseases. Further anal. reduced the CMT2C risk locus to a 4-Mb region. Here we report that SPSMA and CMT2C are allelic disorders caused by mutations in the gene encoding the transient receptor potential cation channel, subfamily V, member 4 (TRPV4). Functional anal. revealed that increased calcium channel activity is a distinct property of both SPSMA- and CMT2C-causing mutant proteins. Our findings link mutations in TRPV4 to altered calcium homeostasis and peripheral neuropathies, implying a pathogenic mechanism and possible options for therapy for these disorders.
- 28Landoure, G., Zdebik, A. A., Martinez, T. L., Burnett, B. G., Stanescu, H. C., Inada, H., Shi, Y., Taye, A. A., Kong, L., Munns, C. H., Choo, S. S., Phelps, C. B., Paudel, R., Houlden, H., Ludlow, C. L., Caterina, M. J., Gaudet, R., Kleta, R., Fischbeck, K. H., and Sumner, C. J. (2010) Mutations in TRPV4 cause Charcot-Marie-Tooth disease type 2C Nat. Genet. 42, 170– 174Google Scholar28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhs1Skur%252FM&md5=91ca81c1ac5b02d2f7429e50e99e6a86Mutations in TRPV4 cause Charcot-Marie-Tooth disease type 2CLandoure, Guida; Zdebik, Anselm A.; Martinez, Tara L.; Burnett, Barrington G.; Stanescu, Horia C.; Inada, Hitoshi; Shi, Yi-Jun; Taye, Addis A.; Kong, Ling-Ling; Munns, Clare H.; Choo, Shelly S.; Phelps, Christopher B.; Paudel, Reema; Houlden, Henry; Ludlow, Christy L.; Caterina, Michael J.; Gaudet, Rachelle; Kleta, Robert; Fischbeck, Kenneth H.; Sumner, Charlotte J.Nature Genetics (2010), 42 (2), 170-174CODEN: NGENEC; ISSN:1061-4036. (Nature Publishing Group)Charcot-Marie-Tooth disease type 2C (CMT2C) is an autosomal dominant neuropathy characterized by limb, diaphragm and laryngeal muscle weakness. Two unrelated families with CMT2C showed significant linkage to chromosome 12q24.11. We sequenced all genes in this region and identified two heterozygous missense mutations in the TRPV4 gene, C805T and G806A, resulting in the amino acid substitutions R269C and R269H. TRPV4 is a well-known member of the TRP superfamily of cation channels. In TRPV4-transfected cells, the CMT2C mutations caused marked cellular toxicity and increased constitutive and activated channel currents. Mutations in TRPV4 were previously assocd. with skeletal dysplasias. Our findings indicate that TRPV4 mutations can also cause a degenerative disorder of the peripheral nerves. The CMT2C-assocd. mutations lie in a distinct region of the TRPV4 ankyrin repeats, suggesting that this phenotypic variability may be due to differential effects on regulatory protein-protein interactions.
- 29Jin, X., Touhey, J., and Gaudet, R. (2006) Structure of the N-terminal ankyrin repeat domain of the TRPV2 ion channel J. Biol. Chem. 281, 25006– 25010Google ScholarThere is no corresponding record for this reference.
- 30Otwinowski, Z. and Minor, W. (1997) Processing of X-ray Diffraction Data Collected in Oscillation Mode Methods Enzymol. 276, 307– 326Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXivFehsbw%253D&md5=c9536971d4e32cc35352c40fb9368131Processing of x-ray diffraction data collected in oscillation modeOtwinowski, Zbyszek; Minor, WladekMethods in Enzymology (1997), 276 (Macromolecular Crystallography, Part A), 307-326CODEN: MENZAU; ISSN:0076-6879. (Academic)Macromol. crystallog. is an iterative process. Rarely do the first crystals provide all the necessary data to solve the biol. problem being studied. Each step benefits from experience learned in previous steps. To monitor the progress, the HKL package provides 2 tools: (1) statistics, both weighted (χ2) and unweighted (R-merge), are provided, and the Bayesian reasoning and multicomponent error model facilitates obtaining the proper error ests. and (2) visualization of the process plays a double role by helping the operator to confirm that the process of data redn., including the resulting statistics, is correct, and allowing one to evaluate problems for which there are no good statistical criteria. Visualization also provides confidence that the point of diminishing returns in data collection and redn. has been reached. At that point, the effort should be directed to solving the structure. The methods presented here have been applied to solve a large variety of problems, from inorg. mols. with 5 Å unit cell to rotavirus of 700 Å diam. crystd. in 700 × 1000 × 1400 Å cell. Overall quality of the method was tested by many researchers by successful application of the programs to MAD structure detns.
- 31Vagin, A. and Teplyakov, A. (2000) An approach to multi-copy search in molecular replacement Acta Crystallogr. D56, 1622– 1624Google ScholarThere is no corresponding record for this reference.
- 32McCoy, A. J., Grosse-Kunstleve, R. W., Adams, P. D., Winn, M. D., Storoni, L. C., and Read, R. J. (2007) Phaser crystallographic software J. Appl. Crystallogr. 40, 658– 674Google Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXnslWqsLk%253D&md5=c63b722ae97e0a74e6a5a079d388f09fPhaser crystallographic softwareMcCoy, Airlie J.; Grosse-Kunstleve, Ralf W.; Adams, Paul D.; Winn, Martyn D.; Storoni, Laurent C.; Read, Randy J.Journal of Applied Crystallography (2007), 40 (4), 658-674CODEN: JACGAR; ISSN:0021-8898. (International Union of Crystallography)Phaser is a program for phasing macromol. crystal structures by both mol. replacement and exptl. phasing methods. The novel phasing algorithms implemented in Phaser have been developed using max. likelihood and multivariate statistics. For mol. replacement, the new algorithms have proved to be significantly better than traditional methods in discriminating correct solns. from noise, and for single-wavelength anomalous dispersion exptl. phasing, the new algorithms, which account for correlations between F+ and F-, give better phases (lower mean phase error with respect to the phases given by the refined structure) than those that use mean F and anomalous differences ΔF. One of the design concepts of Phaser was that it be capable of a high degree of automation. To this end, Phaser (written in C++) can be called directly from Python, although it can also be called using traditional CCP4 keyword-style input. Phaser is a platform for future development of improved phasing methods and their release, including source code, to the crystallog. community.
- 33Emsley, P. and Cowtan, K. (2004) Coot: Model-building tools for molecular graphics Acta Crystallogr. D60, 2126– 2132Google Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXhtVars73P&md5=1be390f3bb6fd584468499ad0921161eCoot: model-building tools for molecular graphicsEmsley, Paul; Cowtan, KevinActa Crystallographica, Section D: Biological Crystallography (2004), D60 (12, Pt. 1), 2126-2132CODEN: ABCRE6; ISSN:0907-4449. (Blackwell Publishing Ltd.)CCP4mg is a project that aims to provide a general-purpose tool for structural biologists, providing tools for x-ray structure soln., structure comparison and anal., and publication-quality graphics. The map-fitting tools are available as a stand-alone package, distributed as 'Coot'.
- 34Murshudov, G. N., Vagin, A. A., and Dodson, E. J. (1997) Refinement of macromolecular structures by the maximum-likelihood method Acta Crystallogr. D53, 240– 255Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2sXjs1Gnsb4%253D&md5=ec7f141ce1542f7ff458b98ecfe3f8afRefinement of macromolecular structures by the maximum-likelihood methodMurshudov, Garib N.; Vagin, Alexei A.; Dodson, Eleanor J.Acta Crystallographica, Section D: Biological Crystallography (1997), D53 (3), 240-255CODEN: ABCRE6; ISSN:0907-4449. (Munksgaard)A review with many refs. on the math. basis of max. likelihood. The likelihood function for macromol. structures is extended to include prior phase information and exptl. std. uncertainties. The assumption that different parts of a structure might have different errors is considered. A method for estg. σA using "free" reflections is described and its effects analyzed. The derived equations have been implemented in the program REFMAC. This has been tested on several proteins at different stages of refinement (bacterial α-amylase, cytochrome c', cross-linked insulin and oligopeptide binding protein). The results derived using the max.-likelihood residual are consistently better than those obtained from least-squares refinement.
- 35Humphrey, W., Dalke, A., and Schulten, K. (1996) VMD: Visual molecular dynamics J. Mol. Graphics 14, 27– 38Google ScholarThere is no corresponding record for this reference.
- 36Phillips, J. C., Braun, R., Wang, W., Gumbart, J., Tajkhorshid, E., Villa, E., Chipot, C., Skeel, R. D., Kale, L., and Schulten, K. (2005) Scalable molecular dynamics with NAM J. Comput. Chem. 26, 1781– 1802Google ScholarThere is no corresponding record for this reference.
- 37MacKerell, A. D., Jr., Bashford, D., Bellott, M., Dunbrack, R. L., Jr., Evanseck, J. D., Field, M. J., Fischer, S., Gao, J., Guo, H., Ha, S., Joseph-McCarthy, D., Kuchnir, L., Kuczera, K., Lau, F. T. K., Mattos, C., Michnick, S., Ngo, T., Nguyen, D. T., Prodhom, B., Reiher, W. E., III, Roux, B., Schlenkrich, M., Smith, J. C., Stote, R., Straub, J., Watanabe, M., Wiórkiewicz-Kuczera, J., Yin, D., and Karplus, M. (1998) All-Atom Empirical Potential for Molecular Modeling and Dynamics Studies of Proteins J. Phys. Chem. B 102, 3586– 3616Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXivVOlsb4%253D&md5=ebb5100dafd0daeee60ca2fa66c1324aAll-Atom Empirical Potential for Molecular Modeling and Dynamics Studies of ProteinsMacKerell, A. D., Jr.; Bashford, D.; Bellott, M.; Dunbrack, R. L.; Evanseck, J. D.; Field, M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.; Joseph-McCarthy, D.; Kuchnir, L.; Kuczera, K.; Lau, F. T. K.; Mattos, C.; Michnick, S.; Ngo, T.; Nguyen, D. T.; Prodhom, B.; Reiher, W. E., III; Roux, B.; Schlenkrich, M.; Smith, J. C.; Stote, R.; Straub, J.; Watanabe, M.; Wiorkiewicz-Kuczera, J.; Yin, D.; Karplus, M.Journal of Physical Chemistry B (1998), 102 (18), 3586-3616CODEN: JPCBFK; ISSN:1089-5647. (American Chemical Society)New protein parameters are reported for the all-atom empirical energy function in the CHARMM program. The parameter evaluation was based on a self-consistent approach designed to achieve a balance between the internal (bonding) and interaction (nonbonding) terms of the force field and among the solvent-solvent, solvent-solute, and solute-solute interactions. Optimization of the internal parameters used exptl. gas-phase geometries, vibrational spectra, and torsional energy surfaces supplemented with ab initio results. The peptide backbone bonding parameters were optimized with respect to data for N-methylacetamide and the alanine dipeptide. The interaction parameters, particularly the at. charges, were detd. by fitting ab initio interaction energies and geometries of complexes between water and model compds. that represented the backbone and the various side chains. In addn., dipole moments, exptl. heats and free energies of vaporization, solvation and sublimation, mol. vols., and crystal pressures and structures were used in the optimization. The resulting protein parameters were tested by applying them to noncyclic tripeptide crystals, cyclic peptide crystals, and the proteins crambin, bovine pancreatic trypsin inhibitor, and carbonmonoxy myoglobin in vacuo and in a crystal. A detailed anal. of the relationship between the alanine dipeptide potential energy surface and calcd. protein φ, χ angles was made and used in optimizing the peptide group torsional parameters. The results demonstrate that use of ab initio structural and energetic data by themselves are not sufficient to obtain an adequate backbone representation for peptides and proteins in soln. and in crystals. Extensive comparisons between mol. dynamics simulation and exptl. data for polypeptides and proteins were performed for both structural and dynamic properties. Calcd. data from energy minimization and dynamics simulations for crystals demonstrate that the latter are needed to obtain meaningful comparisons with exptl. crystal structures. The presented parameters, in combination with the previously published CHARMM all-atom parameters for nucleic acids and lipids, provide a consistent set for condensed-phase simulations of a wide variety of mols. of biol. interest.
- 38Mackerell, A. D., Jr., Feig, M., and Brooks, C. L., III (2004) Extending the treatment of backbone energetics in protein force fields: Limitations of gas-phase quantum mechanics in reproducing protein conformational distributions in molecular dynamics simulations J. Comput. Chem. 25, 1400– 1415Google Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXlsVOgt7c%253D&md5=b2451bb5df548447f8b172a211bc1848Extending the treatment of backbone energetics in protein force fields: Limitations of gas-phase quantum mechanics in reproducing protein conformational distributions in molecular dynamics simulationsMacKerell, Alexander D., Jr.; Feig, Michael; Brooks, Charles L., IIIJournal of Computational Chemistry (2004), 25 (11), 1400-1415CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Computational studies of proteins based on empirical force fields represent a powerful tool to obtain structure-function relationships at an at. level, and are central in current efforts to solve the protein folding problem. The results from studies applying these tools are, however, dependent on the quality of the force fields used. In particular, accurate treatment of the peptide backbone is crucial to achieve representative conformational distributions in simulation studies. To improve the treatment of the peptide backbone, quantum mech. (QM) and mol. mech. (MM) calcns. were undertaken on the alanine, glycine, and proline dipeptides, and the results from these calcns. were combined with mol. dynamics (MD) simulations of proteins in crystal and aq. environments. QM potential energy maps of the alanine and glycine dipeptides at the LMP2/cc-pVxZ/MP2/6-31G* levels, where x = D, T, and Q, were detd., and are compared to available QM studies on these mols. The LMP2/cc pVQZ//MP2/6-31G* energy surfaces for all three dipeptides were then used to improve the MM treatment of the dipeptides. These improvements included addnl. parameter optimization via Monte Carlo simulated annealing and extension of the potential energy function to contain peptide backbone .vphi., ψ dihedral crossterms or a .vphi., ψ grid-based energy correction term. Simultaneously, MD simulations of up to seven proteins in their cryst. environments were used to validate the force field enhancements. Comparison with QM and crystallog. data showed that an addnl. optimization of the .vphi., ψ dihedral parameters along with the grid-based energy correction were required to yield significant improvements over the CHARMM22 force field. However, systematic deviations in the treatment of .vphi. and ψ in the helical and sheet regions were evident. Accordingly, empirical adjustments were made to the grid-based energy correction for alanine and glycine to account for these systematic differences. These adjustments lead to greater deviations from QM data for the two dipeptides but also yielded improved agreement with exptl. crystallog. data. These improvements enhance the quality of the CHARMM force field in treating proteins. This extension of the potential energy function is anticipated to facilitate improved treatment of biol. macromols. via MM approaches in general.
- 39McCleverty, C. J., Koesema, E., Patapoutian, A., Lesley, S. A., and Kreusch, A. (2006) Crystal structure of the human TRPV2 channel ankyrin repeat domain Protein Sci. 15, 2201– 2206Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xpt1ejsbc%253D&md5=c31904fbd84d2334c3a2b276a603344aCrystal structure of the human TRPV2 channel ankyrin repeat domainMcCleverty, Clare J.; Koesema, Eric; Patapoutian, Ardem; Lesley, Scott A.; Kreusch, AndreasProtein Science (2006), 15 (9), 2201-2206CODEN: PRCIEI; ISSN:0961-8368. (Cold Spring Harbor Laboratory Press)TRPV channels are important polymodal integrators of noxious stimuli mediating thermosensation and nociception. An ankyrin repeat domain (ARD), which is a common protein-protein recognition domain, is conserved in the N-terminal intracellular domain of all TRPV channels and predicted to contain 3-4 ankyrin repeats. Here, the authors report the 1st structure from the TRPV channel subfamily, a 1.7-Å resoln. crystal structure of human TRPV2 ARD. The crystal structure revealed a 6-ankyrin repeat stack with multiple insertions in each repeat generating several unique features compared with a canonical ARD. The surface typically used for ligand recognition, the ankyrin groove, contained extended loops with an exposed hydrophobic patch and a prominent kink resulting from a large rotational shift of the last 2 repeats. The TRPV2 ARD provided the 1st structural insight into a domain that coordinates nociceptive sensory transduction and is likely to be a prototype for other TRPV channel ARDs.
- 40Phelps, C. B., Huang, R. J., Lishko, P. V., Wang, R. R., and Gaudet, R. (2008) Structural analyses of the ankyrin repeat domain of TRPV6 and related TRPV ion channels Biochemistry 47, 2476– 2484Google ScholarThere is no corresponding record for this reference.
- 41Croy, C. H., Bergqvist, S., Huxford, T., Ghosh, G., and Komives, E. A. (2004) Biophysical characterization of the free IκBα ankyrin repeat domain in solution Protein Sci. 13, 1767– 1777Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXlsFWktLw%253D&md5=de0d4921c20309c8f99199b5233a252fBiophysical characterization of the free IκBα ankyrin repeat domain in solutionCroy, Carrie Hughes; Bergqvist, Simon; Huxford, Tom; Ghosh, Gourisankar; Komives, Elizabeth A.Protein Science (2004), 13 (7), 1767-1777CODEN: PRCIEI; ISSN:0961-8368. (Cold Spring Harbor Laboratory Press)The crystal structure of IκBα in complex with the transcription factor, nuclear factor κ-B (NF-κB) shows six ankyrin repeats, which are all ordered. Electron d. was not obsd. for most of the residues within the PEST sequence, although it is required for high-affinity binding. To characterize the folded state of IκBα (67-317) when it is not in complex with NF-κB, we have carried out CD spectroscopy, 8-anilino-1-naphthalenesulfonic acid (ANS) binding, differential scanning calorimetry, and amide hydrogen/deuterium exchange expts. The CD spectrum shows the presence of helical structure, consistent with other ankyrin repeat proteins. The large amt. of ANS-binding and amide exchange suggest that the protein may have molten globule character. The amide exchange expts. show that the third ankyrin repeat is the most compact, the second and fourth repeats are somewhat less compact, and the first and sixth repeats are solvent exposed. The PEST extension is also highly solvent accessible. IκBα unfolds with a Tm of 42°, and forms a sol. aggregate that sequesters helical and variable loop parts of the first, fourth, and sixth repeats and the PEST extension. The second and third repeats, which conform most closely to a consensus for stable ankyrin repeats, appear to remain outside of the aggregate. The ramifications of these observations for the biol. function of IκBα are discussed.
- 42Truhlar, S. M., Torpey, J. W., and Komives, E. A. (2006) Regions of IκBα that are critical for its inhibition of NF-κB·DNA interaction fold upon binding to NF-κB Proc. Natl. Acad. Sci. U.S.A. 103, 18951– 18956Google ScholarThere is no corresponding record for this reference.
- 43Barrick, D., Ferreiro, D. U., and Komives, E. A. (2008) Folding landscapes of ankyrin repeat proteins: Experiments meet theory Curr. Opin. Struct. Biol. 18, 27– 34Google Scholar43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXitVGitrY%253D&md5=f9e5a9b6eb66598ba45fe81cddedb69bFolding landscapes of ankyrin repeat proteins: Experiments meet theoryBarrick, Doug; Ferreiro, Diego U.; Komives, Elizabeth A.Current Opinion in Structural Biology (2008), 18 (1), 27-34CODEN: COSBEF; ISSN:0959-440X. (Elsevier B.V.)A review. Nearly 6% of eukaryotic protein sequences contain ankyrin repeat (AR) domains, which consist of several repeats and often function in binding. AR proteins show highly cooperative folding despite a lack of long-range contacts. Both theory and expt. converge to explain that formation of the interface between elements is more favorable than formation of any individual repeat unit. IκBα and Notch both undergo partial folding upon binding perhaps influencing the binding free energy. The simple architecture, combined with identification of consensus residues that are important for stability, has enabled systematic perturbation of the energy landscape by single point mutations that affect stability or by addn. of consensus repeats. The folding energy landscapes appear highly plastic, with small perturbations re-routing folding pathways.
- 44Salazar, H., Llorente, I., Jara-Oseguera, A., Garcia-Villegas, R., Munari, M., Gordon, S. E., Islas, L. D., and Rosenbaum, T. (2008) A single N-terminal cysteine in TRPV1 determines activation by pungent compounds from onion and garlic Nat. Neurosci. 11, 255– 261Google Scholar44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1cXisFOru74%253D&md5=b7b57fe9b43ab99748bb95084f193161A single N-terminal cysteine in TRPV1 determines activation by pungent compounds from onion and garlicSalazar, Hector; Llorente, Itzel; Jara-Oseguera, Andres; Garcia-Villegas, Refugio; Munari, Mika; Gordon, Sharona E.; Islas, Leon D.; Rosenbaum, TamaraNature Neuroscience (2008), 11 (3), 255-261CODEN: NANEFN; ISSN:1097-6256. (Nature Publishing Group)Some members of the transient receptor potential (TRP) family of cation channels mediate sensory responses to irritant substances. Although it is well known that TRPA1 channels are activated by pungent compds. found in garlic, onion, mustard and cinnamon exts., activation of TRPV1 by these exts. remains controversial. Here the authors establish that TRPV1 is activated by pungent exts. from onion and garlic, as well as by allicin, the active compd. in these prepns., and participates together with TRPA1 in the pain-related behavior induced by this compd. The authors found that in TRPV1 these agents act by covalent modification of cysteine residues. In contrast to TRPA1 channels, modification of a single cysteine located in the N-terminal region of TRPV1 was necessary and sufficient for all the effects the authors obsd. The authors findings point to a conserved mechanism of activation in TRP channels, which provides new insights into the mol. basis of noxious stimuli detection.
- 45Isralewitz, B., Gao, M., and Schulten, K. (2001) Steered molecular dynamics and mechanical functions of proteins Curr. Opin. Struct. Biol. 11, 224– 230Google Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXjsVyisro%253D&md5=09b7a191f8ffa19d316ec0f4421c3731Steered molecular dynamics and mechanical functions of proteinsIsralewitz, Barry; Gao, Mu; Schulten, KlausCurrent Opinion in Structural Biology (2001), 11 (2), 224-230CODEN: COSBEF; ISSN:0959-440X. (Elsevier Science Ltd.)A review with 50 refs. At. force microscopy of single mols., steered mol. dynamics and the theory of stochastic processes have established a new field that investigates mech. functions of proteins, such as ligand-receptor binding/unbinding and elasticity of muscle proteins during stretching. The combination of these methods yields information on the energy landscape that controls mech. function and on the force-bearing components of proteins, as well as on the underlying phys. mechanisms.
- 46Camacho, N., Krakow, D., Johnykutty, S., Katzman, P. J., Pepkowitz, S., Vriens, J., Nilius, B., Boyce, B. F., and Cohn, D. H. (2010) Dominant TRPV4 mutations in nonlethal and lethal metatropic dysplasia Am. J. Med. Genet., Part A 152A, 1169– 1177Google Scholar46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXnsFylurg%253D&md5=00b52e12ff7d4602a43cf7270dc2384bDominant TRPV4 mutations in nonlethal and lethal metatropic dysplasiaCamacho, Natalia; Krakow, Deborah; Johnykutty, Sharlin; Katzman, Philip J.; Pepkowitz, Samuel; Vriens, Joris; Nilius, Bernd; Boyce, Brendan F.; Cohn, Daniel H.American Journal of Medical Genetics, Part A (2010), 152A (5), 1169-1177CODEN: AJMGB8; ISSN:1552-4825. (Wiley-Liss, Inc.)Metatropic dysplasia is a clin. heterogeneous skeletal dysplasia characterized by short extremities, a short trunk with progressive kyphoscoliosis, and craniofacial abnormalities that include a prominent forehead, midface hypoplasia, and a squared-off jaw. Dominant mutations in the gene encoding TRPV4, a calcium permeable ion channel, were identified all 10 of a series of metatropic dysplasia cases, ranging in severity from mild to perinatal lethal. These data demonstrate that the lethal form of the disorder is dominantly inherited and suggest locus homogeneity in the disease. Electrophysiol. studies demonstrated that the mutations activate the channel, indicating that the mechanism of disease may result from increased calcium in chondrocytes. Histol. studies in two cases of lethal metatropic dysplasia revealed markedly disrupted endochondral ossification, with reduced nos. of hypertrophic chondrocytes and presence of islands of cartilage within the zone of primary mineralization. These data suggest that altered chondrocyte differentiation in the growth plate leads to the clin. findings in metatropic dysplasia.
- 47Dai, J., Kim, O. H., Cho, T. J., Schmidt-Rimpler, M., Tonoki, H., Takikawa, K., Haga, N., Miyoshi, K., Kitoh, H., Yoo, W. J., Choi, I. H., Song, H. R., Jin, D. K., Kim, H. T., Kamasaki, H., Bianchi, P., Grigelioniene, G., Nampoothiri, S., Minagawa, M., Miyagawa, S. I., Fukao, T., Marcelis, C., Jansweijer, M. C., Hennekam, R. C., Bedeschi, F., Mustonen, A., Jiang, Q., Ohashi, H., Furuichi, T., Unger, S., Zabel, B., Lausch, E., Superti-Furga, A., Nishimura, G., and Ikegawa, S. (2010) Novel and recurrent TRPV4 mutations and their association with distinct phenotypes within the TRPV4 dysplasia family J. Med. Genet. 47, 704– 709Google Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXhsVags7nM&md5=565af1fbc841fd2fa39201dccd9c26caNovel and recurrent TRPV4 mutations and their association with distinct phenotypes within the TRPV4 dysplasia familyDai, J.; Kim, O.-H.; Cho, T.-J.; Schmidt-Rimpler, M.; Tonoki, H.; Takikawa, K.; Haga, N.; Miyoshi, K.; Kitoh, H.; Yoo, W.-J.; Choi, I.-H.; Song, H.-R.; Jin, D.-K.; Kim, H.-T.; Kamasaki, H.; Bianchi, P.; Grigelioniene, G.; Nampoothiri, S.; Minagawa, M.; Miyagawa, S.-i.; Fukao, T.; Marcelis, C.; Jansweijer, M. C. E.; Hennekam, R. C. M.; Bedeschi, F.; Mustonen, A.; Jiang, Q.; Ohashi, H.; Furuichi, T.; Unger, S.; Zabel, B.; Lausch, E.; Superti-Furga, A.; Nishimura, G.; Ikegawa, S.Journal of Medical Genetics (2010), 47 (10), 704-709CODEN: JMDGAE; ISSN:0022-2593. (BMJ Publishing Group)Mutations in TRPV4, a gene that encodes a Ca2+ permeable non-selective cation channel, have recently been found in a spectrum of skeletal dysplasias that includes brachyolmia, spondylometaphyseal dysplasia, Kozlowski type (SMDK) and metatropic dysplasia (MD). Only a total of seven missense mutations were detected, however. The full spectrum of TRPV4 mutations and their phenotypes remained unclear. To examine TRPV4 mutation spectrum and phenotype-genotype assocn., we searched for TRPV4 mutations by PCR-direct sequencing from genomic DNA in 22 MD and 20 SMDK probands. TRPV4 mutations were found in all but one MD subject. In total, 19 different heterozygous mutations were identified in 41 subjects; two were recurrent and 17 were novel. In MD, a recurrent P799L mutation was identified in nine subjects, as well as 10 novel mutations including F471 del, the first deletion mutation of TRPV4. In SMDK, a recurrent R594H mutation was identified in 12 subjects and seven novel mutations. An assocn. between the position of mutations and the disease phenotype was also obsd. Thus, P799 in exon 15 is a hot codon for MD mutations, as four different amino acid substitutions have been obsd. at this codon; while R594 in exon 11 is a hotspot for SMDK mutations. The TRPV4 mutation spectrum in MD and SMDK, which showed genotype-phenotype correlation and potential functional significance of mutations that are non-randomly distributed over the gene, was presented in this study. The results would help diagnostic labs. establish efficient screening strategies for genetic diagnosis of the TRPV4 dysplasia family diseases.
- 48Nishimura, G., Dai, J., Lausch, E., Unger, S., Megarbane, A., Kitoh, H., Kim, O. H., Cho, T. J., Bedeschi, F., Benedicenti, F., Mendoza-Londono, R., Silengo, M., Schmidt-Rimpler, M., Spranger, J., Zabel, B., Ikegawa, S., and Superti-Furga, A. (2010) Spondylo-epiphyseal dysplasia, Maroteaux type (pseudo-Morquio syndrome type 2), and parastremmatic dysplasia are caused by TRPV4 mutations Am. J. Med. Genet., Part A 152A, 1443– 1449Google Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC3czmsVKrsg%253D%253D&md5=fc3a83e2d0d91659a246956bbb8a42daSpondylo-epiphyseal dysplasia, Maroteaux type (pseudo-Morquio syndrome type 2), and parastremmatic dysplasia are caused by TRPV4 mutationsNishimura Gen; Dai Jin; Lausch Ekkehart; Unger Sheila; Megarbane Andre; Kitoh Hiroshi; Kim Ok Hwa; Cho Tae-Joon; Bedeschi Francesca; Benedicenti Francesco; Mendoza-Londono Roberto; Silengo Margherita; Schmidt-Rimpler Maren; Spranger Jurgen; Zabel Bernhard; Ikegawa Shiro; Superti-Furga AndreaAmerican journal of medical genetics. Part A (2010), 152A (6), 1443-9 ISSN:.Recent discoveries have established the existence of a family of skeletal dysplasias caused by dominant mutations in TRPV4. This family comprises, in order of increasing severity, dominant brachyolmia, spondylo-metaphyseal dysplasia Kozlowski type, and metatropic dysplasia. We tested the hypothesis that a further condition, Spondylo-epiphyseal dysplasia (SED), Maroteaux type (MIM 184095; also known as pseudo-Morquio syndrome type 2), could be caused by TRPV4 mutations. We analyzed six individuals with Maroteaux type SED, including three who had previously been reported. All six patients were found to have heterozygous TRPV4 mutations; three patients had unreported mutations, while three patients had mutations previously described in association with metatropic dysplasia. In addition, we tested one individual with a distinct rare disorder, parastremmatic dysplasia (MIM 168400). This patient had a common, recurrent mutation seen in several patients with Kozlowski type spondylo-metaphyseal dysplasia. We conclude that SED Maroteaux type and parastremmatic dysplasia are part of the TRPV4 dysplasia family and that TRPV4 mutations show considerable variability in phenotypic expression resulting in distinct clinical-radiographic phenotypes.
- 49Zimon, M., Baets, J., Auer-Grumbach, M., Berciano, J., Garcia, A., Lopez-Laso, E., Merlini, L., Hilton-Jones, D., McEntagart, M., Crosby, A. H., Barisic, N., Boltshauser, E., Shaw, C. E., Landoure, G., Ludlow, C. L., Gaudet, R., Houlden, H., Reilly, M. M., Fischbeck, K. H., Sumner, C. J., Timmerman, V., Jordanova, A., and Jonghe, P. D. (2010) Dominant mutations in the cation channel gene transient receptor potential vanilloid 4 cause an unusual spectrum of neuropathies Brain 133, 1798– 1809Google ScholarThere is no corresponding record for this reference.
- 50Clapham, D. E. and Miller, C. (2011) A thermodynamic framework for understanding temperature sensing by transient receptor potential (TRP) channels Proc. Natl. Acad. Sci. U.S.A. 108, 19492– 19497Google ScholarThere is no corresponding record for this reference.
- 51Dai, J., Cho, T. J., Unger, S., Lausch, E., Nishimura, G., Kim, O. H., Superti-Furga, A., and Ikegawa, S. (2010) TRPV4-pathy, a novel channelopathy affecting diverse systems J. Hum. Genet. 55, 400– 402Google Scholar51https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC3cnps1ajtw%253D%253D&md5=37b381dc5fe3485dc27ef8fde886164eTRPV4-pathy, a novel channelopathy affecting diverse systemsDai Jin; Cho Tae-Joon; Unger Sheila; Lausch Ekkehart; Nishimura Gen; Kim Ok-Hwa; Superti-Furga Andrea; Ikegawa ShiroJournal of human genetics (2010), 55 (7), 400-2 ISSN:.Transient receptor potential cation channel, subfamily V, member 4 (TRPV4) is a calcium-permeable nonselective cation channel of unknown biological function. TRPV4 mutation was first identified in brachyolmia, and then in a spectrum of autosomal-dominant skeletal dysplasias, which includes Kozlowski type of spondylometaphyseal dysplasia, metatropic dysplasia, Maroteaux type of spondyloepiphyseal dysplasia and parastremmatic dysplasia. Recently, TRPV4 mutation has also been identified in a spectrum of neuromuscular diseases that includes congenital distal spinal muscular atrophy (SMA), scapuloperoneal SMA, and hereditary motor and sensory neuropathy type IIC. These diverse spectrums of diseases compose a novel channelopathy, TRPV4-pathy, which could further include polygenic traits such as serum sodium concentration and a chronic obstructive pulmonary disease. In this review, we clarified the TRPV4 mutation spectrum, and discussed the phenotypic complexity of TRPV4-pathy and its pathogenic mechanisms. TRPV4-pathy may extend further to other monogenic and polygenic diseases.
- 52Andreucci, E., Aftimos, S., Alcausin, M., Haan, E., Hunter, W., Kannu, P., Kerr, B., McGillivray, G., Gardner, R. M., Patricelli, M. G., Sillence, D., Thompson, E., Zacharin, M., Zankl, A., Lamande, S. R., and Savarirayan, R. (2011) TRPV4 related skeletal dysplasias: A phenotypic spectrum highlighted byclinical, radiographic, and molecular studies in 21 new families Orphanet. J. Rare Dis. 6, 37Google Scholar52https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC3MnnvVSmuw%253D%253D&md5=b06a8af5d8f9d02931e0300f3bb90cd8TRPV4 related skeletal dysplasias: a phenotypic spectrum highlighted byclinical, radiographic, and molecular studies in 21 new familiesAndreucci Elena; Aftimos Salim; Alcausin Melanie; Haan Eric; Hunter Warwick; Kannu Peter; Kerr Bronwyn; McGillivray George; McKinlay Gardner R J; Patricelli Maria G; Sillence David; Thompson Elizabeth; Zacharin Margaret; Zankl Andreas; Lamande Shireen R; Savarirayan RaviOrphanet journal of rare diseases (2011), 6 (), 37 ISSN:.BACKGROUND: The TRPV4 gene encodes a calcium-permeable ion-channel that is widely expressed, responds to many different stimuli and participates in an extraordinarily wide range of physiologic processes. Autosomal dominant brachyolmia, spondylometaphyseal dysplasia Kozlowski type (SMDK) and metatropic dysplasia (MD) are currently considered three distinct skeletal dysplasias with some shared clinical features, including short stature, platyspondyly, and progressive scoliosis. Recently, TRPV4 mutations have been found in patients diagnosed with these skeletal phenotypes. METHODS AND RESULTS: We critically analysed the clinical and radiographic data on 26 subjects from 21 families, all of whom had a clinical diagnosis of one of the conditions described above: 15 with MD; 9 with SMDK; and 2 with brachyolmia. We sequenced TRPV4 and identified 9 different mutations in 22 patients, 4 previously described, and 5 novel. There were 4 mutation-negative cases: one with MD and one with SMDK, both displaying atypical clinical and radiographic features for these diagnoses; and two with brachyolmia, who had isolated spine changes and no metaphyseal involvement. CONCLUSIONS: Our data suggest the TRPV4 skeletal dysplasias represent a continuum of severity with areas of phenotypic overlap, even within the same family. We propose that AD brachyolmia lies at the mildest end of this spectrum and, since all cases described with this diagnosis and TRPV4 mutations display metaphyseal changes, we suggest that it is not a distinct entity but represents the mildest phenotypic expression of SMDK.
- 53Klein, C. J., Shi, Y., Fecto, F., Donaghy, M., Nicholson, G., McEntagart, M. E., Crosby, A. H., Wu, Y., Lou, H., McEvoy, K. M., Siddique, T., Deng, H. X., and Dyck, P. J. (2011) TRPV4 mutations and cytotoxic hypercalcemia in axonal Charcot-Marie-Tooth neuropathies Neurology 76, 887– 894Google ScholarThere is no corresponding record for this reference.
- 54Fecto, F., Shi, Y., Huda, R., Martina, M., Siddique, T., and Deng, H. X. (2011) Mutant TRPV4-mediated toxicity is linked to increased constitutive function in axonal neuropathies J. Biol. Chem. 286, 17281– 17291Google ScholarThere is no corresponding record for this reference.
- 55Loukin, S., Su, Z., and Kung, C. (2011) Increased basal activity is a key determinant in the severity of human skeletal dysplasia caused by TRPV4 mutations PLoS One 6, e19533Google ScholarThere is no corresponding record for this reference.
- 56Schaefer, M. (2005) Homo- and heteromeric assembly of TRP channel subunits Pfluegers Arch. 451, 35– 42Google Scholar56https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXhtVKqu7rF&md5=86c1dc77b33044bb46a302f596fbe758Homo- and heteromeric assembly of TRP channel subunitsSchaefer, MichaelPfluegers Archiv (2005), 451 (1), 35-42CODEN: PFLABK; ISSN:0031-6768. (Springer GmbH)A review. Mammalian homologs of the Drosophila melanogaster transient receptor potential (TRP) channels are the 2nd largest cation channel family within the superfamily of hexahelical cation channels. Most mammalian TRP channels function as homo-oligomers and mediate monovalent or divalent cation entry upon activation by a variety of stimuli. Because native TRP channels may be multimeric proteins of possibly complex compn., it has been difficult to compare cation conductances in native tissues to those of clearly defined homomeric TRP channel complexes in living cells. Therefore, the possibility of heteromeric TRP channel assembly has been investigated in recent years by several groups. As a major conclusion of these studies, most heteromeric TRP channel complexes appear to consist of subunit combinations only within relatively narrow confines of phylogenetic subfamilies. Although the general capability of heteromer formation between closely related TRP channel subunits is now clearly established, investigators are only beginning to understand whether these heteromeric complexes are of physiol. significance. Here, the author summarizes current knowledge on the promiscuity and specificity of the assembly of channel complexes composed of TRPC-, TRPV-, and TRPM-subunits of mammalian TRP channels.
Supporting Information
Supporting Information
ARTICLE SECTIONSTRPV-ARD structures used in this study (Table S1), structural similarity between TRPV4-ARD and other TRPV-ARDs (Table S2), molecular dynamics simulations of rat TRPV1-ARD (Table S3), Tm values of wild-type and mutant TRPV4-ARD proteins (Table S4), hTRPV4-ARD cysteine modification assay (Figure S1), stabilities of TRPV1-ARD in equilibrium and SMD simulations (Figure S2), and thermal stabilities of wild-type and mutant TRPV4-ARD proteins (Figure S3). This material is available free of charge via the Internet at http://pubs.acs.org.
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.