• No results found

University of Groningen Klotho in vascular biology Mencke, Rik

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Klotho in vascular biology Mencke, Rik"

Copied!
19
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

University of Groningen

Klotho in vascular biology

Mencke, Rik

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date: 2018

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Mencke, R. (2018). Klotho in vascular biology. Rijksuniversiteit Groningen.

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

309

Chapter 10

Characterization of vascular function in

Klotho deficiency

R. Mencke A. van Buiten N. Stege O.W. Moe H. Buikema J.L. Hillebrands Manuscript in preparation

(3)

310

Abstract

Endothelial dysfunction is an important risk factor for cardiovascular disease. It is generally found in conditions that also feature a deficiency in Klotho, a renal anti-ageing protein. Conversely, full Klotho knockout mice have long been known to develop severe vascular abnormalities and heterozygous mice have been found to have endothelial dysfunction. The goal of this study was to characterize vascular function in Klotho-/- mice.

We used aortic rings from Klotho-/-, Klotho+/-, and WT mice to determine the relaxation

response to acetylcholine after pre-contraction with phenylephrine and thromboxane A2 receptor agonist U46619. We determined phenylephrine and U46619 dose-response relationships. Finally, we incubated Klotho-/- and Klotho+/- rings with recombinant Klotho to

determine whether exogenous Klotho affected subsequent acetylcholine-induced vasodilation.

Klotho-/- mice displayed marked endothelial dysfunction compared to Klotho+/- and WT mice.

Phenylephrine-induced pre-contraction, however, was also significantly higher in Klotho

-/-rings. This was not the case with U46619 and Klotho-/- rings still displayed a markedly impaired

dilatory response. Neither in young (8-12-week-old), nor in old (52-55-week-old) Klotho

+/-mice could endothelial dysfunction be observed. Incubation with recombinant Klotho improved acetylcholine-induced relaxation of Klotho-/- and Klotho+/- rings, albeit not

significantly.

In conclusion, this is the first study to characterize vascular function in Klotho-/- aortic rings

and we confirm the presence of endothelial dysfunction in Klotho deficiency. There was also an increase in sensitivity to phenylephrine, as may occur in ageing. We could not confirm endothelial dysfunction in Klotho+/- mice and we provide indications of direct Klotho effects

(4)

311

Introduction

Endothelial dysfunction is a common finding in ageing, hypertension, and chronic kidney disease (CKD) and it is thought to be involved in the early pathogenesis of atherosclerotic cardiovascular disease (1-5). These conditions are all known experimentally to be accompanied by various degrees of deficiency in Klotho (6-8), which is considered a renal anti-ageing protein (9, 10). Klotho is a type I transmembrane protein expressed primarily in distal convoluted epithelial cells (11, 12), from which it is also cleaved, giving rise to soluble Klotho in blood and urine (13), which then exerts its effects on target tissues.

Similar to CKD patients, which are progressively Klotho-deficient, with endothelial dysfunction and related comorbidities, it has been reported that partial deficiency of Klotho in mice results in endothelial dysfunction (14-16), which comports with the coincident decline in renal function and development of hypertension in these mice (17, 18). This renders Klotho a potentially causal factor in the development of both endothelial dysfunction and associated vascular abnormalities and end-organ damage. Profound effects on the vasculature are attributed to Klotho, since its absence leads to a phenotype of vascular calcification (9, 19, 20), arteriolar hyalinosis, and endothelial hyperpermeability (21) in full Klotho deficiency, and to a phenotype of hypertension (17, 18), and arterial stiffening (22-24) in partial Klotho deficiency. Endothelial dysfunction could play a role in the pathogenesis of many of these vascular conditions that also occur in humans during ageing, and especially in CKD. Promisingly, overexpression of Klotho has been shown to prevent endothelial dysfunction (15), but it is unknown whether this effect is directly mediated by soluble Klotho. As there have so far been no follow-up reports to initial indications of endothelial dysfunction in partial Klotho deficiency and vascular function in conduit arteries from Klotho-/- mice have so far not been

examined, the aim of this study was therefore to further characterize endothelial dysfunction in the complete absence of Klotho and to explore the direct effects of exogenous Klotho on endothelial dysfunction.

Materials and Methods

Animals

Klotho+/- mice were a generous gift from dr. Joost Hoenderop (Department of Physiology,

Radboud University Medical Center, Nijmegen, The Netherlands) and Klotho-/-, Klotho+/-, and

WT mice were bred at the University Medical Center Groningen. Animal experiments were performed in accordance with the NIH Guide for the Care and Use of Laboratory Animals.

(5)

312

Klotho-/- mice were 8-11 weeks old, Klotho+/- mice were 12 weeks old, and WT mice were

8-13 weeks old at the time of the baseline experiments. Vascular function in old Klotho+/- and

WT mice was also measured at 52-55 weeks of age and rings from 90-week-old Klotho+/- mice

were used for stimulation with recombinant Klotho.

Vascular function measurements

The thoracic aorta was removed after termination by cardiac puncture under isoflurane anaesthesia. Vascular ring preparations of approximately 2.5 mm were prepared and mounted on a small vessel wire myograph (Multi Myograph 610 M, Danish Myo Technology, Aarhus, Denmark) connected to a force displacement transducer, in an organ bath with Krebs solution bubbled with carbogen at 37 °C. A computer-assisted normalization protocol was used to pre-stretch rings to 90% of the diameter mimicking a transmural pressure of 100 mmHg, as described previously (25). Rings were allowed to equilibrate for 1 h after which they were primed and checked for viability by repeated stimulation with 60 mM KCl (three times with intermediate washings and re-stabilization). Subsequently, rings were pre-constricted with phenylephrine (PE, 1×10-6 M) or U46619 (3.0×10-8 M) and investigated for endothelial

dysfunction by cumulative stimulation with acetylcholine (ACh, 0.1 nM - 10 µM), followed by the administration of a single high dose of sodium nitroprusside (SNP, 0.1 mM) to induce maximal endothelium-independent relaxation. Phenylephrine-induced contraction curves (0.1 nM - 30 µM) and U46619-induced contraction curves (0.1 nM – 1 µM) were additionally generated in a separate or in the same set of rings after washouts, followed by post-constriction with 60 mM KCl to determine maximal receptor-independent contraction in the same setting. Recombinant mouse Klotho (kindly provided by dr. Orson Moe, UT Southwestern, Dallas, USA) was used at a concentration of 0.4 nM for pre-incubation for 2 hours at 37 °C, in addition to during acetylcholine stimulation. PE, U46619, SNP, and ACh were from Sigma-Aldrich (USA).

Statistical analysis

Ring preparations producing less than 0.5 mN force in response to constriction stimuli were excluded from analysis. ACh- and SNP-induced relaxation responses were expressed as a percentage of pre-constriction. PE- and U46619-induced contraction responses were expressed as a percentage of maximal contraction to 60 mM KCl given at the end of the protocol. Distribution of data was assessed using the Kolmogorov-Smirnov test. Normally distributed data were compared using ANOVA (followed by Bonferroni’s post-hoc correction) or Student’s t test and non-normally distributed data were compared using the Kruskal-Wallis (followed by Dunn’s post-hoc correction) or Mann-Whitney U test, depending on the number of groups. Paired data (rings from the same mouse incubated with PBS or with Klotho) were compared using a paired t test or Wilcoxon signed rank test, depending on the distribution.

(6)

313

GraphPad Prism version 5.0 (GraphPad Software, USA) was used for statistical analysis and a

p value < 0.05 was considered significant.

Results

To assess whether Klotho deficiency induces endothelial dysfunction, we used aortic rings from Klotho-/-, Klotho+/-, and WT mice for ex vivo vascular function measurements. After

pre-constriction with phenylephrine (PE), incubation with increasing concentrations of acetylcholine (ACh) revealed that Klotho-/- rings displayed impaired relaxation compared to

Klotho+/- and WT rings (Figure 1A). Notably, there was no difference between Klotho+/- and

WT rings. However, even though force exerted by all rings was similar when repeatedly

Figure 1. Klotho-/- aortic rings display impaired endothelium-mediated acetylcholine-induced dilation after

pre-constriction with phenylephrine (PE). (A) KCl responses (mean ± SD) from rings depicted in (C). (B) Phenylephrine responses (mean ± SD) from rings depicted in (C). * p < 0.05, ** p < 0.01. (C) Responses from Klotho-/- (N = 6 rings

from 4 mice), Klotho+/- (N = 7 rings from 4 mice), and WT (N = 9 rings from 5 mice) aortic rings to acetylcholine

after pre-constriction with phenylephrine (mean ± SEM). * p < 0.05 (Klotho-/- vs. Klotho+/-). (D) Sodium

(7)

314

incubated with KCl (Figure 1B), the subsequent PE-induced pre-contraction was significantly stronger in Klotho-/- rings (Figure 1C). Incubation with sodium nitroprusside (SNP) resulted in

complete relaxation of WT, Klotho+/-, and Klotho-/- rings (Figure 1D).

To assess whether the relative difference between rings in relaxation could be partially attributed to the difference in PE-induced pre-contraction levels, we performed another series of experiments employing thromboxane A2 receptor agonist U46619 to induce pre-constriction. Comparing Klotho+/- and Klotho-/- rings, we again found that Klotho-/- rings

displayed a marked dilatation impairment (Figure 2A). KCl incubations again produced similar contractions (Figure 2B) and, unlike with PE, there was no significant difference between Klotho-/- and Klotho+/- rings in U46619-induced pre-contraction (Figure 2C). SNP responses

after U41669 pre-contraction were similar across genotypes, albeit incomplete (Figure 2D). These latter data using U41669 as a means to induce comparable pre-constriction seems to preclude the possibility that the attenuated relaxation response to

Figure 2. Klotho-/- aortic rings display impaired endothelium-mediated acetylcholine-induced dilation after

pre-constriction with U46619. (A) KCl responses (mean ± SD) from rings depicted in (C). (B) U46619 responses (mean ± SD) from rings depicted in (C). (C) Responses from Klotho-/- (N = 6 rings from 3 mice) and Klotho+/- (N = 8 rings

from 4 mice) aortic rings to acetylcholine after pre-constriction with U46619 (mean ± SEM). * p < 0.05, ** p < 0.01, *** p < 0.001. (D) Sodium nitroprusside (SNP) responses (mean ± SD) from rings depicted in (C).

(8)

315

ACh in Klotho-/- rings might have been the result of functional antagonism rather than

indicative of impaired endothelial dysfunction. Collectively, the data indicate that Klotho

-/-mice, but not Klotho+/- mice, display endothelial dysfunction.

To confirm the above described differences in PE-induced pre-constriction levels, we additionally assessed per se contractility to PE and U46619 by generating full concentration-response curves to PE and U46619. While concentration-responses to KCl, albeit after a single high dose, were similar between the groups, again contraction responses to PE displayed a with a higher maximal contraction (Figure 3A), suggesting that PE in Klotho-/- aortic rings exhibits a higher

efficacy (rather than potency), as compared to Klotho+/- and WT rings (Figure 3A), although

this difference did not reach statistical significance.U46619 dose-response curves were similar across genotypes (Figure 3C) and KCl post-contractions were also similar after both PE and U46619 stimulations (Figure 3B, D).

Figure 3. Phenylephrine (PE) and U46619 dose-response curves. (A) Responses from Klotho-/- (N = 6 rings from

5 mice), Klotho+/- (N = 8 rings from 5 mice), and WT (N = 11 rings from 5 mice) aortic rings to phenylephrine

relative to the post-constriction KCl response (mean ± SEM). (B) KCl responses from rings depicted in (A). (C) Responses from Klotho-/- (N = 3 from 3 mice), Klotho+/- (N = 3 from 3 mice), and WT (N = 3 from 3 mice) aortic

rings to U46619 relative to the post-constriction KCl response (mean ± SEM). (D) KCl responses from rings depicted in (C).

(9)

316

To assess whether Klotho+/- mice develop endothelial dysfunction at a later age, we repeated

the vascular function measurements in mice that were 1 year old. However, neither with PE as pre-constrictor (Figure 4A), nor with U46619 (Figure 4C) did we observe a difference between Klotho+/- and WT rings. PE- and U46619-induced pre-contractions were similar also

at this age (Figure 4B, D).

Finally, to investigate whether soluble Klotho is capable of directly and acutely rescuing endothelial dysfunction, we incubated Klotho-/- aortic rings with 0.4 nM recombinant mouse

Klotho (a kind gift from Dr. Orson Moe, UT Southwestern University, Dallas, USA) or PBS for 2 hours at 37 °C and during incubation with ACh. There was an increase in ACh-induced relaxation in rings incubated with Klotho (Figure 5A), although this difference was not significant. Note that this set of experiments was performed separately from the experiments in Figure 2A and that the acetylcholine response after U46619 pre-contraction

Figure 4. Klotho+/- aortic rings from 1-year-old mice do not display impaired endothelium-mediated

acetylcholine-induced dilation after pre-constriction with phenylephrine (PE) or U46619. (A) Responses from Klotho+/- (N = 12 rings from 6 mice), and WT (N = 11 rings from 6 mice) aortic rings to acetylcholine after

pre-constriction with phenylephrine (mean ± SEM). One spasmic ring was excluded. (B) Phenylephrine responses (mean ± SD) from rings depicted in (A). (C) Responses from Klotho+/- (N = 12 rings from 6 mice) and WT(N = 10

rings from 5 mice) aortic rings to acetylcholine after pre-constriction with U46619 (mean ± SEM). Two spasmic rings were excluded. (D) U46619 responses for rings depicted in (C).

(10)

317

in this set of experiments was much weaker. Similarly, using rings from old (90-week-old) Klotho+/- mice, incubation with Klotho induced an increase in ACh-induced relaxation (Figure

5B), which was significant only at 3.0*10-6 M ACh.

Figure 5. Recombinant mouse Klotho effects on acetylcholine-induced vasodilation. (A) Responses from rings from Klotho-/- mice (N = 3) treated with 0.4 nM Klotho (N = 5 rings) or PBS (N = 4 rings) to acetylcholine (mean ±

SEM). (B) Responses from rings from 90-week-old Klotho+/- mice (N = 4) treated with 0.4 nM Klotho (N = 8 rings)

(11)

318

Discussion

The goal of this study was to re-visit the relationship between Klotho and vascular function and we found that Klotho-/- mice have marked endothelial dysfunction. Surprisingly, these

mice were more sensitive to phenylephrine. Also surprisingly, we could not observe endothelial dysfunction in Klotho+/- mice. Finally, this study provides indications that soluble

Klotho may rapidly improve endothelial function in Klotho-/- and Klotho+/- aortic rings.

After the serendipitous discovery of Klotho was first described in 1997, the first study on vascular function in Klotho-deficient mice followed suit in 1998 and was performed by Saito

et al. They describe that kl/+ mice (heterozygotes of the original kl/kl mice in which the Klotho

promoter is disrupted) display endothelial dysfunction (14). Also, Saito et al. investigated the response of cremaster arterioles (33 ± 3 µm in diameter) in vivo in kl/kl mice and found that relaxation of these arterioles in response to ACh was impaired. However, the aortic rings from

kl/kl mice were too calcified to dilate or contract in response to ACh. The Klotho-/- mice we

used do not develop severe vascular calcification, likely due to a combination of factors, including a genetic background more resistant to the development of vascular calcification, diet composition, and housing conditions. The lack of calcification allowed us to evaluate the vascular function in aortic rings of Klotho-/- mice for the first time. It was not necessarily a

given that large conduit arteries from Klotho-/- mice would have endothelial dysfunction

because it had been detected before in heterozygote kl/+ mice. After all, kl/+ mice develop hypertension (albeit at a later age) (17, 18), whereas kl/kl mice develop hypotension (26), likely due to the additional presence of hypovolemia (27). Our results in Klotho-/- mice further

cement that endothelial dysfunction in Klotho deficiency is likely a primary phenomenon, rather than secondary to, for instance, hypertension.

An interesting finding in this study was the increased sensitivity to PE that we observed in Klotho-/- mice. This may be in line with the increased response to norepinephrine in kl/+ rings

compared to WT rings as previously reported (16). An increased sensitivity to PE has been observed before in conduit arteries from hypertensive rats (28) (which likely also had decreased Klotho levels (29)). In this study, however, there was an increased potency rather than efficacy of PE. An increased efficacy of PE has also been observed, for instance, due to increased oxidative stress (30). Of note, PE responses in aged Klotho+/- and WT animals were

similar to the PE response in young Klotho-/- mice, indicating that perhaps sensitivity to PE

increases with age and the premature ageing syndrome seen in Klotho-/- mice is accompanied

by a prematurely increased response to PE. Evidence from rat studies appears to be in line with this hypothesis (31, 32), as well as evidence from the SAM-P8 mouse (33), another mouse model of premature ageing.

Our results disagree with the literature in an important aspect, which is that we could not identify any differences in vascular function between Klotho+/- and WT animals. Since a

(12)

319

number of phenotypic aspects in this model are less severe than in other Klotho-deficient mice (like the development of vascular calcification), it is possible that Klotho+/- mice are protected

from developing endothelial dysfunction as well. Since Klotho+/- mice are generally

indistinguishable from WT littermates, save for their increased vulnerability to renal, cardiac, and other damage, it is clear that Klotho levels of approximately 50% of WT levels are typically enough to prevent the spontaneous development of ageing-related phenotypes. It is possible that the mice used in this study remained above the threshold below which endothelial dysfunction develops and that a different diet, different housing conditions or a different genetic background would have potentiated the development of endothelial dysfunction also in Klotho+/- mice.

Saito et al. also observed that vascular function in kl/+ mice is restored by parabiosis with WT littermates (14), indicating that (a) soluble factor(s) more abundant in WT mice, or WT-assisted excretion of (a) soluble factor(s) more abundant in Klotho-deficient mice is responsible for improving vascular function. Given the beneficial effect of Klotho overexpression on endothelial dysfunction in OLETF rats (15), and the occurrence of Klotho as a soluble factor in the circulation, it is tempting to speculate that soluble Klotho itself from WT mice was responsible for improving vascular function in kl/+ mice after parabiosis. However, as Klotho deficiency entails profound disturbances in a great many pathways, there are certainly other soluble factors that are plausibly of relevance. For instance, hyperphosphatemia can cause endothelial dysfunction (34, 35), as can high levels of uremic toxins (36, 37). Both phosphate and indoxyl sulfate are elevated in Klotho-deficient mice (38). The function of membrane-bound Klotho as a co-receptor to fibroblast growth factor (FGF) receptor 1c for FGF23 (39, 40)

results in high FGF23 levels in the absence of Klotho and high FGF23 levels can also cause endothelial dysfunction (41, 42). Despite these alternative reasons for endothelial dysfunction in Klotho deficiency, our data support that there could also be a direct effect of soluble Klotho, although our experiments lack the power for conclusive statements. Because aortic rings were pre-incubated with recombinant Klotho and Klotho was present during ACh stimulation, we cannot currently discriminate between an effect initiated by recombinant Klotho incubation and potential changes in protein expression, or an acute effect of the presence of Klotho during relaxation, as described by Six et al. (43). However, the divergence of the curves occurred only during stimulation with higher concentrations of ACh (at a constant Klotho concentration), which is suggestive of an effect mediated by Klotho pre-incubation. Further study of this phenomenon will be required to provide clear answers.

The nature of the interaction between Klotho and the endothelium is an important piece in many puzzles in Klotho biology. There are indications that Klotho binds directly to the endothelium (44) and it has been shown that Klotho prevents excessive calcium influx into endothelial cells by promoting internalization of calcium channel TRPC1 via a complex with Klotho and VEGFR2 (21). Furthermore, while it recently became clear that Klotho binds to lipid rafts in the cell membrane (45), it is unclear whether this also occurs in endothelial cells and there also appears to be a yet to be elucidated way for Klotho to cross the endothelium via

(13)

320

transcytosis (46-48), through which it may reach target cell types in the body. Further study of the interaction between Klotho and the endothelium therefore holds great potential for vascular disease and ageing-related diseases in general.

In conclusion, we can confirm that Klotho deficiency is accompanied by endothelial dysfunction. Furthermore, Klotho-/- mice display an exaggerated response to phenylephrine,

which is potentially ageing-related, and there may be direct protective effects of soluble Klotho on endothelial dysfunction. A therapeutic approach of preventing endothelial dysfunction using Klotho-based treatment merits further investigation.

(14)

321

Acknowledgments

(15)

322

References

1. P. O. Bonetti, L. O. Lerman, A. Lerman, Endothelial dysfunction: a marker of atherosclerotic risk. Arterioscler. Thromb. Vasc. Biol. 23, 168-175 (2003).

2. J. P. Halcox, W. H. Schenke, G. Zalos, R. Mincemoyer, A. Prasad, M. A. Waclawiw, K. R. Nour, A. A. Quyyumi, Prognostic value of coronary vascular endothelial dysfunction. Circulation. 106, 653-658 (2002).

3. D. R. Seals, K. L. Jablonski, A. J. Donato, Aging and vascular endothelial function in humans. Clin. Sci. (Lond). 120, 357-375 (2011).

4. J. Malyszko, Mechanism of endothelial dysfunction in chronic kidney disease. Clin. Chim. Acta. 411, 1412-1420 (2010).

5. C. H. Bolton, L. G. Downs, J. G. Victory, J. F. Dwight, C. R. Tomson, M. I. Mackness, J. H. Pinkney, Endothelial dysfunction in chronic renal failure: roles of lipoprotein oxidation and pro-inflammatory cytokines. Nephrol. Dial. Transplant. 16, 1189-1197 (2001).

6. Y. Wang, Z. Sun, Klotho gene delivery prevents the progression of spontaneous hypertension and renal damage. Hypertension. 54, 810-817 (2009).

7. M. C. Hu, M. Shi, J. Zhang, H. Quinones, C. Griffith, M. Kuro-o, O. W. Moe, Klotho deficiency causes vascular calcification in chronic kidney disease. J. Am. Soc. Nephrol. 22, 124-136 (2011).

8. Z. Zuo, H. Lei, X. Wang, Y. Wang, W. Sonntag, Z. Sun, Aging-related kidney damage is associated with a decrease in klotho expression and an increase in superoxide production. Age (Dordr). 33, 261-274 (2011).

9. M. Kuro-o, Y. Matsumura, H. Aizawa, H. Kawaguchi, T. Suga, T. Utsugi, Y. Ohyama, M. Kurabayashi, T. Kaname, E. Kume, H. Iwasaki, A. Iida, T. Shiraki-Iida, S. Nishikawa, R. Nagai, Y. I. Nabeshima, Mutation of the mouse klotho gene leads to a syndrome resembling ageing. Nature. 390, 45-51 (1997).

10. H. Kurosu, M. Yamamoto, J. D. Clark, J. V. Pastor, A. Nandi, P. Gurnani, O. P. McGuinness, H. Chikuda, M. Yamaguchi, H. Kawaguchi, I. Shimomura, Y. Takayama, J. Herz, C. R. Kahn, K. P. Rosenblatt, M. Kuro-o, Suppression of aging in mice by the hormone Klotho. Science. 309, 1829-1833 (2005).

11. Y. Kato, E. Arakawa, S. Kinoshita, A. Shirai, A. Furuya, K. Yamano, K. Nakamura, A. Iida, H. Anazawa, N. Koh, A. Iwano, A. Imura, T. Fujimori, M. Kuro-o, N. Hanai, K. Takeshige, Y. Nabeshima, Establishment of the anti-Klotho monoclonal antibodies and detection of Klotho protein in kidneys. Biochem. Biophys. Res. Commun. 267, 597-602 (2000).

12. H. Olauson, R. Mencke, J. L. Hillebrands, T. E. Larsson, Tissue expression and source of circulating alphaKlotho. Bone. (2017).

13. A. Imura, A. Iwano, O. Tohyama, Y. Tsuji, K. Nozaki, N. Hashimoto, T. Fujimori, Y. Nabeshima, Secreted Klotho protein in sera and CSF: implication for post-translational cleavage in release of Klotho protein from cell membrane. FEBS Lett. 565, 143-147 (2004).

(16)

323

14. Y. Saito, T. Yamagishi, T. Nakamura, Y. Ohyama, H. Aizawa, T. Suga, Y. Matsumura, H. Masuda, M. Kurabayashi, M. Kuro-o, Y. Nabeshima, R. Nagai, Klotho protein protects against endothelial dysfunction. Biochem. Biophys. Res. Commun. 248, 324-329 (1998).

15. Y. Saito, T. Nakamura, Y. Ohyama, T. Suzuki, A. Iida, T. Shiraki-Iida, M. Kuro-o, Y. Nabeshima, M. Kurabayashi, R. Nagai, In vivo klotho gene delivery protects against endothelial dysfunction in multiple risk factor syndrome. Biochem. Biophys. Res. Commun. 276, 767-772 (2000).

16. T. Nakamura, Y. Saito, Y. Ohyama, H. Masuda, H. Sumino, M. Kuro-o, Y. Nabeshima, R. Nagai, M. Kurabayashi, Production of nitric oxide, but not prostacyclin, is reduced in klotho mice. Jpn. J. Pharmacol. 89, 149-156 (2002). 17. X. Zhou, K. Chen, Y. Wang, M. Schuman, H. Lei, Z. Sun, Antiaging Gene Klotho Regulates Adrenal CYP11B2 Expression and Aldosterone Synthesis. J. Am. Soc. Nephrol. (2015).

18. X. Zhou, K. Chen, H. Lei, Z. Sun, Klotho gene deficiency causes salt-sensitive hypertension via monocyte chemotactic protein-1/CC chemokine receptor 2-mediated inflammation. J. Am. Soc. Nephrol. 26, 121-132 (2015).

19. M. Ohnishi, T. Nakatani, B. Lanske, M. S. Razzaque, In vivo genetic evidence for suppressing vascular and soft-tissue calcification through the reduction of serum phosphate levels, even in the presence of high serum calcium and 1,25-dihydroxyvitamin d levels. Circ. Cardiovasc. Genet. 2, 583-590 (2009).

20. M. Ohnishi, T. Nakatani, B. Lanske, M. S. Razzaque, Reversal of mineral ion homeostasis and soft-tissue calcification of klotho knockout mice by deletion of vitamin D 1alpha-hydroxylase. Kidney Int. 75, 1166-1172 (2009).

21. T. Kusaba, M. Okigaki, A. Matui, M. Murakami, K. Ishikawa, T. Kimura, K. Sonomura, Y. Adachi, M. Shibuya, T. Shirayama, S. Tanda, T. Hatta, S. Sasaki, Y. Mori, H. Matsubara, Klotho is associated with VEGF receptor-2 and the transient receptor potential canonical-1 Ca2+ channel to maintain endothelial integrity. Proc. Natl. Acad. Sci. U. S. A. 107, 19308-19313 (2010).

22. K. Chen, X. Zhou, Z. Sun, Haplodeficiency of Klotho Gene Causes Arterial Stiffening via Upregulation of Scleraxis Expression and Induction of Autophagy. Hypertension. 66, 1006-1013 (2015).

23. D. Gao, Z. Zuo, J. Tian, Q. Ali, Y. Lin, H. Lei, Z. Sun, Activation of SIRT1 Attenuates Klotho Deficiency-Induced Arterial Stiffness and Hypertension by Enhancing AMP-Activated Protein Kinase Activity. Hypertension. 68, 1191-1199 (2016).

24. Y. Lin, J. Chen, Z. Sun, Antiaging Gene Klotho Deficiency Promoted High-Fat Diet-Induced Arterial Stiffening via Inactivation of AMP-Activated Protein Kinase. Hypertension. 67, 564-573 (2016).

25. I. V. Samarska, H. R. Bouma, H. Buikema, H. E. Mungroop, M. C. Houwertjes, A. R. Absalom, A. H. Epema, R. H. Henning, S1P1 receptor modulation preserves vascular function in mesenteric and coronary arteries after CPB in the rat independent of depletion of lymphocytes. PLoS One. 9, e97196 (2014).

26. J. Voelkl, I. Alesutan, C. B. Leibrock, L. Quintanilla-Martinez, V. Kuhn, M. Feger, S. Mia, M. S. Ahmed, K. P. Rosenblatt, M. Kuro-O, F. Lang, Spironolactone ameliorates PIT1-dependent vascular osteoinduction in klotho-hypomorphic mice. J. Clin. Invest. 123, 812-822 (2013).

(17)

324

27. C. B. Leibrock, I. Alesutan, J. Voelkl, T. Pakladok, D. Michael, E. Schleicher, Z. Kamyabi-Moghaddam, L. Quintanilla-Martinez, M. Kuro-O, F. Lang, NH4Cl Treatment Prevents Tissue Calcification in Klotho Deficiency. J. Am. Soc. Nephrol. (2015).

28. R. M. White, C. O. Rivera, C. B. Davison, Differential contribution of endothelial function to vascular reactivity in conduit and resistance arteries from deoxycorticosterone-salt hypertensive rats. Hypertension. 27, 1245-1253 (1996).

29. H. Aizawa, Y. Saito, T. Nakamura, M. Inoue, T. Imanari, Y. Ohyama, Y. Matsumura, H. Masuda, S. Oba, N. Mise, K. Kimura, A. Hasegawa, M. Kurabayashi, M. Kuro-o, Y. Nabeshima, R. Nagai, Downregulation of the Klotho gene in the kidney under sustained circulatory stress in rats. Biochem. Biophys. Res. Commun. 249, 865-871 (1998). 30. C. R. Tirapelli, C. R. De Andrade, M. Lieberman, F. R. Laurindo, H. P. De Souza, A. M. de Oliveira, Vitamin K1 (phylloquinone) induces vascular endothelial dysfunction: role of oxidative stress. Toxicol. Appl. Pharmacol. 213, 10-17 (2006).

31. C. F. Reyes-Toso, D. Obaya-Naredo, C. R. Ricci, F. M. Planells, J. E. Pinto, L. M. Linares, D. P. Cardinali, Effect of melatonin on vascular responses in aortic rings of aging rats. Exp. Gerontol. 42, 337-342 (2007).

32. A. Harvey, A. C. Montezano, R. M. Touyz, Vascular biology of ageing-Implications in hypertension. J. Mol. Cell. Cardiol. 83, 112-121 (2015).

33. S. Llorens, R. M. de Mera, A. Pascual, A. Prieto-Martin, Y. Mendizabal, C. de Cabo, E. Nava, J. Jordan, The senescence-accelerated mouse (SAM-P8) as a model for the study of vascular functional alterations during aging. Biogerontology. 8, 663-672 (2007).

34. K. K. Stevens, L. Denby, R. K. Patel, P. B. Mark, S. Kettlewell, G. L. Smith, M. J. Clancy, C. Delles, A. G. Jardine, Deleterious effects of phosphate on vascular and endothelial function via disruption to the nitric oxide pathway. Nephrol. Dial. Transplant. 32, 1617-1627 (2017).

35. G. S. Di Marco, M. Konig, C. Stock, A. Wiesinger, U. Hillebrand, S. Reiermann, S. Reuter, S. Amler, G. Kohler, F. Buck, M. Fobker, P. Kumpers, H. Oberleithner, M. Hausberg, D. Lang, H. Pavenstadt, M. Brand, High phosphate directly affects endothelial function by downregulating annexin II. Kidney Int. 83, 213-222 (2013).

36. M. Yu, Y. J. Kim, D. H. Kang, Indoxyl sulfate-induced endothelial dysfunction in patients with chronic kidney disease via an induction of oxidative stress. Clin. J. Am. Soc. Nephrol. 6, 30-39 (2011).

37. S. Chu, X. Mao, H. Guo, L. Wang, Z. Li, Y. Zhang, Y. Wang, H. Wang, X. Zhang, W. Peng, Indoxyl sulfate potentiates endothelial dysfunction via reciprocal role for reactive oxygen species and RhoA/ROCK signaling in 5/6 nephrectomized rats. Free Radic. Res. 51, 237-252 (2017).

38. K. Yang, C. Du, X. Wang, F. Li, Y. Xu, S. Wang, S. Chen, F. Chen, M. Shen, M. Chen, M. Hu, T. He, Y. Su, J. Wang, J. Zhao, Uremic solute indoxyl sulfate-induced platelet hyperactivity contributes to CKD-associated thrombosis in mice. Blood. (2017).

39. I. Urakawa, Y. Yamazaki, T. Shimada, K. Iijima, H. Hasegawa, K. Okawa, T. Fujita, S. Fukumoto, T. Yamashita, Klotho converts canonical FGF receptor into a specific receptor for FGF23. Nature. 444, 770-774 (2006).

(18)

325

40. H. Kurosu, Y. Ogawa, M. Miyoshi, M. Yamamoto, A. Nandi, K. P. Rosenblatt, M. G. Baum, S. Schiavi, M. C. Hu, O. W. Moe, M. Kuro-o, Regulation of fibroblast growth factor-23 signaling by klotho. J. Biol. Chem. 281, 6120-6123 (2006).

41. N. Silswal, C. D. Touchberry, D. R. Daniel, D. L. McCarthy, S. Zhang, J. Andresen, J. R. Stubbs, M. J. Wacker, FGF23 directly impairs endothelium-dependent vasorelaxation by increasing superoxide levels and reducing nitric oxide bioavailability. Am. J. Physiol. Endocrinol. Metab. 307, E426-36 (2014).

42. M. I. Yilmaz, A. Sonmez, M. Saglam, H. Yaman, S. Kilic, E. Demirkaya, T. Eyileten, K. Caglar, Y. Oguz, A. Vural, M. Yenicesu, C. Zoccali, FGF-23 and vascular dysfunction in patients with stage 3 and 4 chronic kidney disease. Kidney Int. 78, 679-685 (2010).

43. I. Six, H. Okazaki, P. Gross, J. Cagnard, C. Boudot, J. Maizel, T. B. Drueke, Z. A. Massy, Direct, acute effects of Klotho and FGF23 on vascular smooth muscle and endothelium. PLoS One. 9, e93423 (2014).

44. T. Takenaka, T. Inoue, Y. Ohno, T. Miyazaki, A. Nishiyama, N. Ishii, H. Suzuki, Calcitriol supplementation improves endothelium-dependent vasodilation in rat hypertensive renal injury. Kidney Blood Press. Res. 39, 17-27 (2014).

45. G. Dalton, S. W. An, S. I. Al-Juboori, N. Nischan, J. Yoon, E. Dobrinskikh, D. W. Hilgemann, J. Xie, K. Luby-Phelps, J. J. Kohler, L. Birnbaumer, C. L. Huang, Soluble klotho binds monosialoganglioside to regulate membrane microdomains and growth factor signaling. Proc. Natl. Acad. Sci. U. S. A. 114, 752-757 (2017).

46. M. C. Hu, M. Shi, J. Zhang, J. Pastor, T. Nakatani, B. Lanske, M. S. Razzaque, K. P. Rosenblatt, M. G. Baum, M. Kuro-o, O. W. Moe, Klotho: a novel phosphaturic substance acting as an autocrine enzyme in the renal proximal tubule. FASEB J. 24, 3438-3450 (2010).

47. P. Ravikumar, L. Li, J. Ye, M. Shi, M. Taniguchi, J. Zhang, M. Kuro-O, M. C. Hu, O. W. Moe, C. C. Hsia, Alpha-Klotho deficiency in Acute Kidney Injury Contributes to Lung Damage. J. Appl. Physiol. (1985)., jap.00792.2015 (2015).

48. M. C. Hu, M. Shi, J. Zhang, T. Addo, H. J. Cho, S. L. Barker, P. Ravikumar, N. Gillings, A. Bian, S. S. Sidhu, M. Kuro-O, O. W. Moe, Renal Production, Uptake, and Handling of Circulating alphaKlotho. J. Am. Soc. Nephrol. 27, 79-90 (2016).

(19)

Referenties

GERELATEERDE DOCUMENTEN

We confirmed the expression of both membrane-bound and alternative Klotho mRNA transcripts in human kidney, primary human renal tubular epithelial cells (RTECs), and a human

(C) Autoradiography on aortas from Klotho -/- mice at 7 weeks of age reveals that the pattern develops in the aortic arch and in the abdominal aorta. Patchy involvement of the

Given the profound protective effects of Klotho on endothelial function, smooth muscle cell differentiation, and the kidney, we explored whether a relationship exists

To address whether there is a relationship between Klotho and intimal hyperplasia, we investigated various Klotho-deficient mouse strains and we studied the development of

GBM cell lines (A172, U251, U87, and GSC23) were stimulated with recombinant human Klotho (aa 34-981; R&amp;D Systems, USA) at a concentration of 40 pM, equivalent to 5.2 ng/ml) for

To summarise, reports indicate that soluble Klotho, either directly supplemented as recombinant protein or derived from induced or constitutive overexpression, is capable of

We have assessed studies on Klotho gene variants, Klotho promoter methylation, Klotho mRNA expression, Klotho protein expression, and soluble Klotho levels, with regard to

In short, we found that certain Klotho SNPs are more frequent in ESRD patients and that rs577912 and rs553791 in recipients are associated with an increased risk of graft loss, which