Skip to main content

Xenon preconditioning: molecular mechanisms and biological effects

Abstract

Xenon is one of noble gases and has been recognized as an anesthetic for more than 50 years. Xenon possesses many of the characteristics of an ideal anesthetic, but it is not widely applied in clinical practice mainly because of its high cost. In recent years, numerous studies have demonstrated that xenon as an anesthetic can exert neuroprotective and cardioprotective effects in different models. Moreover, xenon has been applied in the preconditioning, and the neuroprotective and cardioprotective effects of xenon preconditioning have been investigated in a lot of studies in which some mechanisms related to these protections are proposed. In this review, we summarized these mechanisms and the biological effects of xenon preconditioning.

The noble gases are a group of chemical elements with very similar properties: they are all odorless, colorless, monatomic gases with very low chemical reactivity under standard conditions. The six noble gases occurring naturally are helium (He), neon (Ne), argon (Ar), krypton (Kr), xenon (Xe), and the radioactive radon (Rn) [1]. Among these noble gases, xenon is the most frequently investigated and widely applied in medicine. Xenon has been used as an anesthetic [2], to treat brain and heart injuries due to its neuroprotection and cardioprotection [3, 4] and in single photon emission computed tomography (SPECT) [5].

Preconditioning is a process by which an organism’s exposure to a stress/stimulus permits it to decrease cellular damage or death when exposed to a subsequent greater or more sustained stress. To date, a large number of strategies have been developed for preconditioning such as lipopolysaccharide, heat and seizure, hypoxia and hyperoxia [6]. In this brief review, we introduce the protective effects of xenon preconditioning, not the xenon treatment.

Introduction of xenon

Xenon is a colorless, heavy, odorless noble gas and was discovered by William Ramsay and Morris Travers in 1898. Xenon derives its name from the Greek word for “stranger” [7]. Xenon is a trace gas in Earth’s atmosphere and much more expensive than the lighter noble gases due to its very low concentration in air (0.5 ppm) [2]. It is inert to most common chemical reactions because the outer valence shell contains eight electrons.

To date, xenon has been commercially used for lasers, high intensity lamps, flash bulbs, jet propellant in the aerospace industry, X-ray tubes, and in medicine [7].

In medicine, xenon has been used experimentally in clinical anesthetic practice for more than 50 years. As a general anesthetic, xenon possesses many advantages. For example, its blood-gas partition coefficient is extremely small (0.115), which results in a rapid onset and offset of its action. It lacks teratogenicity and can produce profound analgesia which thereby inhibits the surgery induced hemodynamic and catecholamine responses. It is also a potent hypnotic and does not produce hemodynamic depression because it at least in part has no influence on some important ion channels [2, 7]. In addition, it has been confirmed that xenon can confer neuroprotective [3] and cardioprotective [4, 8] effects. Unlike nitrous oxide (N2O), xenon is not a greenhouse gas and so it is also viewed as environmentally friendly [9]. Xenon that is vented into the atmosphere is being returned to its original source, and thus environmental pollution is unlikely. However, its relatively high cost has precluded its more widespread clinical use.

Mechanisms underlying bioeffects of xenon

Numerous studies have been conducted to investigate the mechanisms of xenon’s bioeffects. Xenon can potently inhibit the N-methyl-D-aspartate (NMDA) receptors non-competitively, with little effect on the γ-aminobutyric acid A (GABAA) receptor and non-NMDA glutamatergic receptor [10]. In the study of Yamakura and Harris, their results showed N2O (0.58 atmosphere [atm]) and xenon (0.46 atm) exhibited similar effects on various receptors, and the NMDA receptors, and nACh receptors composed of β2 subunits are likely targets for NO and xenon [11]. In addition, xenon (and NO) has been reported to competitively inhibit the 5-hydroxytryptamine receptor 3A (5HT3A receptor) expressed in the xenopus oocytes [12] but this has yet to be confirmed in mammalian cells.

In addition, the xenon induced anesthesia is related to the inhibition of the calcium ATPase pump on the cell membrane of synapses [13], which results from a conformational change when xenon binds to nonpolar sites inside the protein [14], and the non-specific interactions between the xenon and the lipid membrane [15].

Mechanisms of protective effects of xenon preconditioning

Cardioprotection, neuroprotection, renoprotective and endothelial protection

In 2005, a German group found that exposure to 70% xenon 45 min before myocardial ischemia could confer cardioprotection in a rat model, in which the activation of ε isoform of protein kinase C and its downstream target p38 mitogen-activated protein kinase (MAPK) is a central molecular mechanism [16]. Eight months later, the same group preconditioned rats with same methods, but hearts without experiencing ischemia/reperfusion were collected for detection. Their results demonstrated that xenon preconditioning enhanced the translocation of heat shock protein 27 (HSP27) to the particulate fraction and increased F-actin polymerization and activated MAPKAPK-2 and HSP27 downstream of PKC and p38 MAPK. Moreover, F-actin and pHSP27 were colocalized after xenon preconditioning. Their findings link xenon preconditioning to the cytoskeleton, revealing new insights into the mechanisms of xenon preconditioning in vivo[17]. In another study of Weber et al., they found xenon preconditioning increased the translocation of PKC-ε to membrane regions, and the mitochondrial adenosine triphosphate (ATP) dependent potassium (KATP) channels and 3 phosphatidylinositol-dependent kinase-1 (PDK-1) located upstream of PKC-ε and were crucial for the activation of PKC-ε [18]. In addition, they noted ERK 1/2, but not JNK, was also a mediator of xenon preconditioning [19]. A group in the USA found the cardioprotection of xenon preconditioning was attributed to the phosphorylation of Akt, glycogen synthase kinase 3 β (GSK-3β), preservation of mitochondrial function, and inhibition of Ca2+-induced mitochondrial permeability transition (MPT) pore opening (Table 1) [20].

Table 1 Mechanisms of protective effects of xenon preconditioning

The neuroprotection of xenon preconditioning was attributed to the transcription of activity-dependent neuroprotective protein (ADNP) in neonatal rats [21], the opening of plasmalemmal KATP channels in neuronal-glial cocultures [22] and a reduction in the plasma IL-1β and an up-regulation of hippocampal HSP72 in the surgery and/or isoflurane induced postoperative cognitive decline (POCD) model [23]. Moreover, the neuroprotection of xenon preconditioning was independent of gender in mouse transient middle cerebral artery occlusion model [24].

In the study of Ma et al., the xenon preconditioning was found to be a natural inducer of hypoxia-inducible factor (HIF-1α) [24, 25]. Xenon preconditioning can activate HIF-1α and its downstream effectors erythropoietin (EPO) and vascular endothelial growth factor (VEGF) in a time-dependent manner in the kidneys in vivo and in vitro[25, 26].

In the human umbilical vein endothelial cells, xenon preconditioning was found to prevent tumor necrosis factor-α (TNF-α) induced mRNA and protein expression of intracellular cell adhesion molecule 1 (ICAM-1) and vascular cell adhesion molecule 1 (VCAM-1) and decreased the TNF-α induced transcriptional activity of nuclear factor κB (NF-κB), but had no effect on the TNF-α induced E-selectin expression [27].

Contribution of anesthetic property of xenon to protective effects

Studies also revealed that preconditioning with other noble gases without anesthetic properties could exert protective effect [28, 29]. These findings supported the contention that the protection of preconditioning with noble gases was independent of the anesthetic properties per se. However, the protective effect of preconditioning with other noble gases was not found in the human tubular kidney cells, and even helium by comparison significantly enhanced the cell injury [26]. Additionally, in a study of Baumert et al., they did not confirm the cardioprotective effect of brief, intermittent xenon preconditioning, but the xenon anesthesia (xenon 70%, continued before and after myocardial ischemia) exert protective effect on myocardial ischemia [30]. Moreover, the xenon exposure induced gene expression profile was also found to be different from that following N2O (another anesthetic) exposure [31].

Late protective effect of xenon preconditioning

In previous studies, pharmacological preconditioning not only produces early protection but induces late protective effect. This was true in the xenon preconditioning [24, 3234]. The late myocardial protective effect of xenon preconditioning was found to be closely related to the cyclooxygenase-2 (COX-2) activity because inhibition of COX-2 abolished this cardioprotective effect and the mRNA and protein expression of COX-2 remained unchanged following xenon preconditioning [32], the enhanced phosphorylated cyclic adenosine monophosphate response element binding protein signaling [33] and the phosphorylated cAMP-response element binding protein (pCREB)-regulated synthesis of proteins that promote survival against neuronal injury (Table 1) [33, 34].

Comparisons of xenon preconditioning with other strategies for preconditioning

Investigators also compared the protective effects of xenon preconditioning with those of ischemia, anesthetic(s) and hypoxia preconditioning. In the experimental myocardial infarction model, results showed combined isoflurane/xenon preconditioning reduced infarct size but not more than isoflurane alone. Ischemic preconditioning was more effective than the anesthetics [35]. In the rat N2O- and isoflurane-induced neuroapoptosis model, xenon preconditioning was found to prevent N2O oxide- and isoflurane-induced neuroapoptosis (in vivo and in vitro) and cognitive deterioration (in vivo). However, N2O - and isoflurane-induced neuroapoptosis was exacerbated by hypoxic pretreatment. N2O pretreatment had no effect [36]. In the study of Baumert et al., myocardial infarct size was reduced by ischemic preconditioning but less so by xenon anesthesia, and brief, intermittent exposure to xenon before myocardial ischemia did not reduce myocardial infarct size [30]. However, in the study of Weber et al., ischemic preconditioning induced by 3 × 5 min coronary artery occlusion reduced infarct size to a similar extent like anesthetic induced preconditioning [16].

Conclusion and perspectives

Although the anesthetic properties of xenon have been known for more than 50 years and the neuroprotection and cardioprotection of xenon demonstrated for more than 10 years, xenon preconditioning is still in its infant stage. The neuroprotective and cardioprotective effects of xenon preconditioning have been confirmed in the majority of studies, but clinical studies have not been reported. In addition, numerous signaling pathways and a large amount of molecules have been shown to involve in the protective effects of xenon preconditioning, but the relative contribution of each pathway or molecule is unclear to the protective effects of xenon preconditioning.

As shown above, the cost of xenon for anesthesia and treatment is high due to the large amount of xenon used, which is the major factor limiting the wide application of xenon. However, in the available experiments on xenon preconditioning, xenon was used in combination with oxygen at a ratio of 7:3 (v/v), and preconditioning was done with 3 cycles of xenon/oxygen administered for 5 min periods [20] or for up to 20 min [23]. If the protection of this strategy is confirmed, the cost of xenon may be significantly reduced as compared to that in the anesthesia and treatment. With the development of technology, closed-circuit xenon delivery is introduced. In the study of Chakkarapani et al., they reported that the xenon consumption was minimal (<$2/h at $10/L) when the gas exchange was maintained [37]. Thus, xenon preconditioning is very cost-effective. Furthermore, xenon is not flammable and can be easily applied at bedside. Thus, we postulate that xenon preconditioning may become an alternative strategy for the prevention of diseases or injuries.

References

  1. Noble gas: http://en.wikipedia.org/wiki/Noble_gas,

  2. Jordan BD, Wright EL: Xenon as an anesthetic agent. AANA J. 2010, 78: 387-392.

    PubMed  Google Scholar 

  3. Banks P, Franks NP, Dickinson R: Competitive inhibition at the glycine site of the N-methyl-D-aspartate receptor mediates xenon neuroprotection against hypoxia-ischemia. Anesthesiology. 2010, 112: 614-622. 10.1097/ALN.0b013e3181cea398.

    Article  CAS  PubMed  Google Scholar 

  4. Pagel PS: Cardioprotection by noble gases. J Cardiothorac Vasc Anesth. 2010, 24: 143-163. 10.1053/j.jvca.2009.03.016.

    Article  CAS  PubMed  Google Scholar 

  5. Suga K, Kawakami Y, Yamashita T, Zaki M, Matsunaga N: Characterization of 133Xe gas washout in pulmonary emphysema with dynamic 133Xe SPECT functional images. Nucl Med Commun. 2006, 27: 71-80. 10.1097/01.mnm.0000188222.07204.62.

    Article  PubMed  Google Scholar 

  6. Liu W, Khatibi N, Sridharan A, Zhang JH: Application of medical gases in the field of neurobiology. Med Gas Res. 2011, 1: 13-10.1186/2045-9912-1-13.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  7. Sanders R, Franks N, Maze M: Xenon: no stranger to anaesthesia. Br J Anaesth. 2003, 91: 709-717. 10.1093/bja/aeg232.

    Article  CAS  PubMed  Google Scholar 

  8. Preckel B, Müllenheim J, Moloschavij A, Thämer V, Schlack W: Xenon administration during early reperfusion reduces infarct size after regional ischemia in the rabbit heart in vivo. Anesth Analg. 2000, 91: 1327-1332. 10.1097/00000539-200012000-00003.

    Article  CAS  PubMed  Google Scholar 

  9. Goto T, Nakata Y, Morita S: Will xenon be a stranger or a friend? the cost, benefit, and future of xenon anesthesia. Anesthesiology. 2003, 98: 1-2. 10.1097/00000542-200301000-00002.

    Article  PubMed  Google Scholar 

  10. Franks NP, Dickinson R, de Sousa SL, Hall AC, Lieb WR: How does xenon produce anesthesia?. Nature. 1998, 396: 324-10.1038/24525.

    Article  CAS  PubMed  Google Scholar 

  11. Yamakura T, Harris RA: Effects of gaseous anesthetics nitrous oxide and xenon on ligand-gated ion channels. Comparison with isoflurane and ethanol. Anesthesiology. 2000, 93: 1095-1101. 10.1097/00000542-200010000-00034.

    Article  CAS  PubMed  Google Scholar 

  12. Suzuki T, Koyama H, Sugimoto M, Uchida I, Mashimo T: The diverse actions of volatile and gaseous anesthetics on human-cloned 5-hydroxytryptamine 3 receptors expressed in Xenopus oocytes. Anesthesiology. 2002, 96: 699-704. 10.1097/00000542-200203000-00028.

    Article  CAS  PubMed  Google Scholar 

  13. Franks JJ, Horn JL, Janicki PK, Singh G: Halothane, isoflurane, xenon, and nitrous oxide inhibit calcium ATPase pump activity in rat brain synaptic plasma membranes. Anesthesiology. 1995, 82: 108-117. 10.1097/00000542-199501000-00015.

    Article  CAS  PubMed  Google Scholar 

  14. Lopez MM, Kosk-Kosicka D: How do volatile anesthetics inhibit Ca2+-ATPases?. J Biol Chem. 1995, 270 (47): 28239-28245. 10.1074/jbc.270.47.28239.

    Article  CAS  PubMed  Google Scholar 

  15. Xu Y, Tang P: Amphiphilic sites for general anesthetic action? Evidence from 129Xe-[1H] intermolecular nuclear Overhauser effects. Biochim Biophys Acta. 1997, 1323 (1): 154-162. 10.1016/S0005-2736(96)00184-8.

    Article  CAS  PubMed  Google Scholar 

  16. Weber NC, Toma O, Wolter JI, Obal D, Müllenheim J, Preckel B, Schlack W: The noble gas xenon induces pharmacological preconditioning in the rat heart in vivo via induction of PKC-epsilon and p38 MAPK. Br J Pharmacol. 2005, 144: 123-132.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  17. Weber NC, Toma O, Wolter JI, Wirthle NM, Schlack W, Preckel B: Mechanisms of xenon- and isoflurane-induced preconditioning - a potential link to the cytoskeleton via the MAPKAPK-2/HSP27 pathway. Br J Pharmacol. 2005, 146: 445-455. 10.1038/sj.bjp.0706324.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  18. Weber NC, Toma O, Damla H, Wolter JI, Schlack W, Preckel B: Upstream signaling of protein kinase C-epsilon in xenon-induced pharmacological preconditioning. Implication of mitochondrial adenosine triphosphate dependent potassium channels and phosphatidylinositol dependent kinase-1. Eur J Pharmacol. 2006, 539: 1-9. 10.1016/j.ejphar.2006.03.054.

    Article  CAS  PubMed  Google Scholar 

  19. Weber NC, Stursberg J, Wirthle NM, Toma O, Schlack W, Preckel B: Xenon preconditioning differently regulates p44/42 MAPK (ERK 1/2) and p46/54 MAPK (JNK 1/2 and 3) in vivo. Br J Anaesth. 2006, 97: 298-306. 10.1093/bja/ael153.

    Article  CAS  PubMed  Google Scholar 

  20. Mio Y, Shim YH, Richards E, Bosnjak ZJ, Pagel PS, Bienengraeber M: Xenon preconditioning: the role of prosurvival signaling, mitochondrial permeability transition and bioenergetics in rats. Anesth Analg. 2009, 108: 858-866. 10.1213/ane.0b013e318192a520.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  21. Cattano D, Valleggi S, Ma D, Kastsiuchenka O, Abramo A, Sun P, Cavazzana AO, Natale G, Maze M, Giunta F: Xenon induces transcription of ADNP in neonatal rat brain. Neurosci Lett. 2008, 440: 217-221. 10.1016/j.neulet.2008.05.086.

    Article  CAS  PubMed  Google Scholar 

  22. Bantel C, Maze M, Trapp S: Neuronal preconditioning by inhalational anesthetics: evidence for the role of plasmalemmal adenosine triphosphate-sensitive potassium channels. Anesthesiology. 2009, 110: 986-995. 10.1097/ALN.0b013e31819dadc7.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  23. Vizcaychipi MP, Lloyd DG, Wan Y, Palazzo MG, Maze M, Ma D: Xenon pretreatment may prevent early memory decline after isoflurane anesthesia and surgery in mice. PLoS One. 2011, 6: e26394-10.1371/journal.pone.0026394.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  24. Limatola V, Ward P, Cattano D, Gu J, Giunta F, Maze M, Ma D: Xenon preconditioning confers neuroprotection regardless of gender in a mouse model of transient middle cerebral artery occlusion. Neuroscience. 2010, 165: 874-881. 10.1016/j.neuroscience.2009.10.063.

    Article  CAS  PubMed  Google Scholar 

  25. Ma D, Lim T, Xu J, Tang H, Wan Y, Zhao H, Hossain M, Maxwell PH, Maze M: Xenon preconditioning protects against renal ischemic-reperfusion injury via HIF-1alpha activation. J Am Soc Nephrol. 2009, 20: 713-720. 10.1681/ASN.2008070712.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  26. Rizvi M, Jawad N, Li Y, Vizcaychipi MP, Maze M, Ma D: Effect of noble gases on oxygen and glucose deprived injury in human tubular kidney cells. Exp Biol Med (Maywood). 2010, 235: 886-891. 10.1258/ebm.2010.009366.

    Article  CAS  Google Scholar 

  27. Weber NC, Kandler J, Schlack W, Grueber Y, Frädorf J, Preckel B: Intermitted pharmacologic pretreatment by xenon, isoflurane, nitrous oxide, and the opioid morphine prevents tumor necrosis factor alpha-induced adhesion molecule expression in human umbilical vein endothelial cells. Anesthesiology. 2008, 108: 199-207. 10.1097/01.anes.0000299441.32091.ed.

    Article  CAS  PubMed  Google Scholar 

  28. Pagel PS, Krolikowski JG, Shim YH, Venkatapuram S, Kersten JR, Weihrauch D, Warltier DC, Pratt PF: Noble gases without anesthetic properties protect myocardium against infarction by activating prosurvival signaling kinases and inhibiting mitochondrial permeability transition in vivo. Anesth Analg. 2007, 105: 562-569. 10.1213/01.ane.0000278083.31991.36.

    Article  CAS  PubMed  Google Scholar 

  29. Pagel PS, Krolikowski JG, Pratt PF, Shim YH, Amour J, Warltier DC, Weihrauch D: The mechanism of helium-induced preconditioning: a direct role for nitric oxide in rabbits. Anesth Analg. 2008, 107: 762-768. 10.1213/ane.0b013e3181815995.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  30. Baumert JH, Hein M, Gerets C, Baltus T, Hecker KE, Rossaint R: The effect of xenon anesthesia on the size of experimental myocardial infarction. Anesth Analg. 2007, 105: 1200-1206. 10.1213/01.ane.0000284697.73471.9c.

    Article  CAS  PubMed  Google Scholar 

  31. Valleggi S, Cavazzana AO, Bernardi R, Ma D, Natale G, Maze M, Cattano D, Giunta F: Xenon up-regulates several genes that are not up-regulated by nitrous oxide. J Neurosurg Anesthesiol. 2008, 20: 226-232. 10.1097/ANA.0b013e31817da878.

    Article  PubMed  Google Scholar 

  32. Weber NC, Frässdorf J, Ratajczak C, Grueber Y, Schlack W, Hollmann MW, Preckel B: Xenon induces late cardiac preconditioning in vivo: a role for cyclooxygenase 2?. Anesth Analg. 2008, 107: 1807-1813. 10.1213/ane.Ob013e31818874bf.

    Article  CAS  PubMed  Google Scholar 

  33. Luo Y, Ma D, Ieong E, Sanders RD, Yu B, Hossain M, Maze M: Xenon and sevoflurane protect against brain injury in a neonatal asphyxia model. Anesthesiology. 2008, 109: 782-789. 10.1097/ALN.0b013e3181895f88.

    Article  CAS  PubMed  Google Scholar 

  34. Ma D, Hossain M, Pettet GK, Luo Y, Lim T, Akimov S, Sanders RD, Franks NP, Maze M: Xenon preconditioning reduces brain damage from neonatal asphyxia in rats. J Cereb Blood Flow Metab. 2006, 26: 199-208. 10.1038/sj.jcbfm.9600184.

    Article  CAS  PubMed  Google Scholar 

  35. Baumert JH, Hein M, Gerets C, Baltus T, Hecker KE, Rossaint R: The effect of xenon on isoflurane protection against experimental myocardial infarction. J Cardiothorac Vasc Anesth. 2009, 23: 614-618. 10.1053/j.jvca.2009.01.028.

    Article  CAS  PubMed  Google Scholar 

  36. Shu Y, Patel SM, Pac-Soo C, Fidalgo AR, Wan Y, Maze M, Ma D: Xenon pretreatment attenuates anesthetic-induced apoptosis in the developing brain in comparison with nitrous oxide and hypoxia. Anesthesiology. 2010, 113: 360-368. 10.1097/ALN.0b013e3181d960d7.

    Article  CAS  PubMed  Google Scholar 

  37. Chakkarapani E, Thoresen M, Hobbs CE, Aquilina K, Liu X, Dingley J: A closed-circuit neonatal xenon delivery system: a technical and practical neuroprotection feasibility study in newborn pigs. Anesth Analg. 2009, 109: 451-460. 10.1213/ane.0b013e3181aa9550.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgement

This study was partially supported by the National Natural Science Foundation of China (No. 81000493/H0906).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Xuejun Sun.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

Tao HY and Sun XJ outlined this review; Liu WW, Liu Y, Chen H and Liu K searched the database and summarized the findings; Liu WW and Liu Y drafted this manuscript; Tao HY and Sun XJ revised this review. All authors read and approved the final manuscript

Wenwu Liu, Ying Liu contributed equally to this work.

Rights and permissions

Open Access This article is published under license to BioMed Central Ltd. This is an Open Access article is distributed under the terms of the Creative Commons Attribution License ( https://creativecommons.org/licenses/by/2.0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Liu, W., Liu, Y., Chen, H. et al. Xenon preconditioning: molecular mechanisms and biological effects. Med Gas Res 3, 3 (2013). https://doi.org/10.1186/2045-9912-3-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/2045-9912-3-3

Keywords