请输入您要查询的百科知识:

 

词条 Nuclear clock
释义

  1. Principle of Operation

  2. Different nuclear clock concepts

      The single-ion nuclear clock    The solid-state nuclear clock  

  3. Transition requirements

  4. History

      The history of the nuclear clock    The history of 229mTh  

  5. Applications

  6. References

A nuclear clock or nuclear optical clock is a notional clock that would use the frequency of a nuclear transition as its reference frequency

[1], in the same manner as an atomic clock uses the frequency of an electronic transition in an atom's shell. Such a clock is expected to be by a factor of about 10 more accurate than the best current atomic clocks, with an achievable accuracy approaching the 10−19 level.

[2]

The only nuclear state suitable for the development of a nuclear clock using existing technology is thorium-229m, a nuclear isomer of thorium-229 and the lowest-energy nuclear isomer known. With an energy of 7.8±0.5 eV

[3]

, the corresponding ground-state transition is expected to be in the vacuum ultraviolet wavelength region around 160 nm, which would make it accessible to laser excitation.

[4]

Principle of Operation

Modern optical atomic clocks are by today the most accurate time-keeping devices. Their underlying principle of operation is based on the fact that the energy of an atomic transition (the energy difference between two atomic states) is independent on space and time. The atomic transition energy corresponds to a particular frequency of a light wave, which is required to drive the transition. Therefore, an atomic transtion can be excited with the help of laser light, if the laser frequency is exactly matching the frequency corresponding to the energy of the atomic transition. Thus, in turn, the laser frequency can be stabilized to match the corresponding atomic transition energy by continuous verification of a successful laser excitation of the atomic transition. In case of successful stabilization to an atomic transition, the frequency of the laser light will always be the same (independent on space and time).

It is technologically possible to measure the frequency of laser light to extraordinary high accuracy by counting the oscillations of the light wave with the help of a frequency comb. This allows to measure time, simply by counting the number of ocillations of the laser light, that has been stabilized to a particular atomic transition. Such a device is known as optical atomic clock.

[5]

One prominent example for an optical atomic clock is the Ytterbium lattice clock, where a particular transition in the Ytterbium(Yb)-171 isotope is used for laser stabilization.

[6]

In this case, one second has elapsed after exactly 518,295,836,590,864 oscillations of the laser light stabilized to the corresponding transition. Other examples for optical atomic clocks of highest accuracies are the Ytterbium(Yb)-171 single-ion clock

[7]

, the Strontium(Sr)-87 optical lattice clock

[8]

and the Aluminum(Al)-27 single-ion clock.[9]

The achieved accuarcies of these clocks vary around 10−18, corresponding to about 1 second of inaccuracy in 30 billion years, significantly longer than the age of the universe.

For a nuclear optical clock the principle of operation remains unchanged, however, with the important difference that a nuclear transition instead of an atomic shell transition is used for laser stabilization.[1] The expected advantage of a nuclear clock compared to an atomic clock is that, figuratively speaking, the atomic nucleus is by up to five orders of magnitude smaller than the atomic shell and therefore (due to small magnetic dipole and electric quadrupole moments) significantly less affected by external influences like, e.g., electric and magnetic fields. Such external perturbations are the limiting factor for the achieved accuracies of atomic-shell based clocks. Due to this conceptual advantage, a nuclear optical clock is expected to achieve a time accuracy approaching 10−19.[2]

Different nuclear clock concepts

The single-ion nuclear clock

Two differenct concepts for a nuclear optical clock have been discussed in literature. The first concept, known as the single-ion nuclear clock[1][2], uses the nucleus of an individual trapped and laser-cooled ion for laser stabilization. Such a clock is expected to achieve the highest time accuracy, as the ion is to a large extent isolated from its environment. There is no conceptual hindrance to also use an ion cloud for the purpose of laser stabilization.

The solid-state nuclear clock

As the nucleus is largely unaffected by the atomic shell, it is also intruiging to embed many nuclei into a crystal lattice environment. This concept is known as the solid-state nuclear clock, which poses the second nuclear clock concept.[1] Due to the high density of embedded nuclei of up to 1018 per cm3, this concept would allow to irradiate a huge amount of nuclei in parallel, thereby drastically increasing the achievable signal-to-noise ratio[10]

, however, on the cost of potentially higher external perturbances.[11]

Transition requirements

From the principle of operation of a nuclear optical clock it is evident, that direct laser excitation of a nuclear state is a central requirement for the development of a nuclear clock. Until today no direct nuclear laser excitation has been achieved. The central reason is that the typical energy range of nuclear transitions (keV to MeV) is orders of magnitude above the maximum energy which is accessible with significant intensity by today's narrowbandwith laser technology (a few eV). There are only two nuclear excited states known, which possess an extraordinary low excitation energy (below 100 eV). These are 229mTh, a metastable nuclear excited state of the isotope Thorium-229 with an excitation energy of only 7.8 eV [3] and 235mU, a metastable excited state of Uranium-235 with an energy of 76.7 eV.

[12] For nuclear structure reasons, only 229mTh offers a realistic chance for direct nuclear laser excitation.

Further requirements for the development of a nuclear clock are, that the lifetime of the nuclear excited state is relatively long, thereby leading to a resonance of narrow bandwith (a high quality factor) and that the ground-state nucleus is easily available and sufficiently long-lived to allow to work with moderate quantities of the material. Fortunately, with a radiative lifetime of 103 to 104 seconds of 229mTh [13][14] and a lifetime of about 7917 years of a 229Th nucleus in its ground state[15]., both conditions are fulfilled for 229mTh, making it an ideal candidate for the development of a nuclear clock.

History

The history of the nuclear clock

A nuclear optical clock based on 229mTh was first proposed in 2003 by E. Peik and C. Tamm, who developed an idea of U. Sterr.[1] The paper contains both concepts, the single-ion nuclear clock, as well as the solid-state nuclear clock.

In their pioneering work, Peik and Tamm proposed to use individual laser-cooled 229Th3+ ions in a Paul trap to perform nuclear laser spectroscopy.[1] Here the 3+ charge state is advantageous, as it possesses a shell structure suitable for direct laser cooling. It was further proposed to excite an electronic shell state, to achieve 'good' quantum numbers of the total system of the shell plus nucleus that will lead to a reduction of the influence induced by external perturbing fields. A central idea is to probe the successful laser excitation of the nuclear state via the hyperfine-structure shift induced into the electronic shell due to the different nuclear spins of ground- and excited state. This method is known as the double-resonance method.

The expected performance of a single-ion nuclear clock was further investigated in 2012 by C. Campbell et al. with the result that a systematic frequency uncertainty (accuracy) of the clock of 1.5·10-19 could be achieved, which would be by about an order of magnitude better than the accuracy achieved by the best optical atomic clocks today.[2] The nuclear clock approach proposed by Campbell et al. slightly differs from the original one proposed by Peik and Tamm. Instead of exciting an electronic shell state in order to obtain the highest insensitivity against external perturbing fields, the nuclear clock propsed by Campbell et al. uses a stretched pair of nuclear hyperfine states in the electronic ground-state configuration, which appears to be advantageous in terms of the achievable quality factor and an improved suppression of the quadratic Zeeman shift.

The solid-state nuclear clock approach was further developed in 2010 by W.G. Rellergert et al. [11] with the result of an expected long-term accuracy of about 2·10-16. Although expected to be less accurate than the single-ion nuclear clock approach due to line-broadening effects and temperature shifts in the crystal lattice environement, this approach may have advantages in terms of compactness, robustness and power consumption. The expected stability performance was investigated by G. Kazakov et al. in 2012.[10]

Important steps on the road towards a nuclear clock were the successful direct laser cooling of 229Th3+ ions in a Paul trap achieved in 2011[16], the direct detection of the 229mTh decay in 2016 [17] and a first detection of the isomer-induced hyperfine-structure shift, enabling the double-resonance method to probe a successful nuclear excitation in 2018.[18]

The history of 229mTh

{{main|Thorium-229m}}

Since 1976, the 229Th nucleus has been known to possess a low energy excited state[19], which was constrained to be of below 10 eV excitation energy in 1990[20] and for which an energy value of 3.5±1.0 eV was determined in 1994.[21] As early as 1996 it was proposed to use the nuclear excitation as a highly stable source of light for metrology by E.V. Tkalya.[22]

At the time of the nuclear clock proposal in 2003 the parameters of 229mTh, in particular its energy, were not known to sufficient precision to allow for nuclear laser spectroscopy of individual thorium ions and thus the development of a nuclear clock. This fact triggered a multitude of experimental efforts to pin down the excited state's parameters like energy and half-life. The detection of light emitted in the direct decay of 229mTh would significantly help to determine its energy to higher precision, however until today all efforts failed to observe a secure signal of light emitted in the decay of 229mTh.[4] The failure of early experiments to observe any direct 229mTh decay signal can partly be explained by a correction of the energy value to 7.6±0.5 eV in 2007[23] (slightly shifted to 7.8±05 eV in 2009[3]). However, also all recent experiments failed to observe any signal of light emitted in the direct decay, potentially pointing towards a strong non-radiative decay channel.[24][25][26][27] In 2012[28] and again in 2018[29] the detection of light emitted in the decay of 229mTh was reported, but the observed signals are subject to controversial discussions within the community.[30]

A direct detection of electrons as being emitted in the isomer's internal conversion decay channel was achieved in 2016.[17] This detection laid the foundation for the determination of the 229mTh half-life in neutral, surface-bound atoms in 2017[31] and a first laser-spectroscopic characterization in 2018.[18] However, the most important question, namely a precise determination of the 229mTh energy value, remained unanswered. An experiment having the potential to directly determine the isomeric energy value via spectroscopy of the internal conversion electrons emitted in the 229mTh decay is currently conducted.[32]

Applications

When operational, a nuclear optical clock is expected to be applicable in various fields. Potential applications may arise in the field where already today's atomic clocks are in operation, like e.g., satellite-based navigation or data transfer. However, also potentially new applications may arise in the fields of relativistic geodesy, the search for topological dark matter or the determination of time-variations of fundamental constants.[33]

Especially a high sensitivity of a nuclear clock for potential time variations of fundamental constants, e.g., the fine-structure constant, has been highlighted.[34] The central idea is that a nuclear transition couples differently to the fine-structure constant than an atomic shell transition does. For this reason a comparison of the frequency of a nuclear clock with an atomic clock could lead to an extraordinary high sensitivity for potential time variations of the fine structure constant. The achievable factor of sensitivity, however, remains subject to speculation. A recent measurement is consistent with enhancement factors between 1 (no enhancement) and 104.[18]

References

1. ^{{cite journal| first1=E. |last1=Peik| first2=Chr. |last2=Tamm| title = Nuclear laser spectroscopy of the 3.5 eV transition in 229Th| url = http://www.ptb.de/cms/fileadmin/internet/fachabteilungen/abteilung_4/4.4_zeit_und_frequenz/pdf/th001.pdf| journal = Europhysics Letters| volume = 61| issue = 2| pages = 181–186| date = 2003-01-15| doi = 10.1209/epl/i2003-00210-x| bibcode = 2003EL.....61..181P}}
2. ^{{cite journal| first1=C. |last1=Campbell| first2=A.G. |last2=Radnaev| first3=A. |last3=Kuzmich| first4=V.A. |last4=Dzuba| first5=V.V. |last5=Flambaum| first6=A. |last6=Derevianko| title = A single ion nuclear clock for metrology at the 19th decimal place| journal = Phys. Rev. Lett.| volume = 108| pages = 120802| date = 2012| doi = 10.1103/PhysRevLett.108.120802| arxiv= 1110.2490}}
3. ^{{cite conference| title = Improved value for the energy splitting of the ground-state doublet in the nucleus 229Th|vauthors=Beck BR, Wu CY, Beiersdorfer P, Brown GV, Becker JA, Moody KJ, Wilhelmy JB, Porter FS, Kilbourne CA, Kelley RL | date = 2009-07-30| conference = 12th Int. Conf. on Nuclear Reaction Mechanisms| location = Varenna, Italy| id = LLNL-PROC-415170| url = https://e-reports-ext.llnl.gov/pdf/375773.pdf}}
4. ^{{cite journal| title=Towards a 229Th-based nuclear clock| first1=Lars |last1=von der Wense| first2=Benedict |last2=Seiferle| first3=Peter G. |last3=Thirolf| journal=Measurement Techniques |volume=60 |issue=12 |pages=1178–1192 |date=March 2018| doi=10.1007/s11018-018-1337-1| arxiv=1811.03889}}
5. ^{{cite journal| title=Optical atomic clocks| first1=A.D. |last1=Ludlow| journal=Rev. Mod. Phys. |volume=87 |pages=637-699 |date=2015| doi=10.1103/RevModPhys.87.637| arxiv=1407.3493|display-authors=etal}}
6. ^{{cite journal| title=Atomic clock performance enabling geodesy below the centimetre level| first1=W.F. |last1=McGrew| first2=X. |last2=Zhang| first3=R.J. |last3=Fasano| first4=S.A. |last4=Schäffer| first5=K. |last5=Beloy| first6=D. |last6=Nicolody| first7=R.C. |last7=Brown| first8=N. |last8=Hinkley| first9=G. |last9=Milani| first10=M. |last10=Schioppo| first11=T.H. |last11=Yoon| first12=A.D. |last12=Ludlow| journal=Nature |volume=564 |pages=87-90 |date=2018| doi=10.1038/s41586-018-0738-2| arxiv=1807.11282}}
7. ^{{cite journal| title=Single-ion atomic clock with 3·10-18 systematic uncertainty| first1=N. |last1=Huntemann| journal=Phys. Rev. Lett. |volume=116 |pages=063001 |date=2016| doi=10.1103/PhysRevLett.116.063001| arxiv=1602.03908|display-authors=etal}}
8. ^{{cite journal| title= Systematic evaluation of an atomic clock at 2·10-18 total uncertainty| first1=T.L. |last1=Nicholson| journal=Nature Commun. |volume=6 |pages=6896 |date=2015| doi=10.1038/ncomms7896| arxiv=1412.8261|display-authors=etal}}
9. ^{{cite journal| title= An 27Al+ quantum-logic clock with systematic uncertainty below 10-18 | first1=S.M. |last1=Brewer| arxiv=1902.07694|display-authors=etal}}
10. ^{{cite journal |arxiv=1204.3268 |title=Performance of a 229 Thorium solid-state nuclear clock |author1=G. A. Kazakov |author2=A. N. Litvinov |author3=V. I. Romanenko |author4=L. P. Yatsenko |author5=A. V. Romanenko |author6=M. Schreitl |author7=G. Winkler |author8=T. Schumm |date=2 Oct 2012 |doi=10.1088/1367-2630/14/8/083019 |volume=14 |issue=8 |journal=New Journal of Physics |page=083019 |bibcode=2012NJPh...14h3019K}}
11. ^{{cite journal| title= Constraining the evolution of the fundamental constants with a solid-state optical frequency reference based on the 229Th nucleus| first1=W.G. |last1=Rellergert| journal=Phys. Rev. Lett. |volume=104 |pages=200802 |date=2010| doi=10.1103/PhysRevLett.104.200802|url=http://www.nat.vu.nl/~wimu/Varying-Constants-Papers/Th-229-Rellergert-2010.pdf|display-authors=etal}}
12. ^{{cite journal| title= Accurate measurement of the first excited nuclear state in 235U| first1=F. |last1=Ponce| journal=Phys. Rev. C |volume=97 |pages=054310 |date=2018| doi=10.1103/PhysRevC.97.054310|display-authors=etal}}
13. ^{{cite journal| last1=Tkalya |first1=E.V. |last2=Schneider |first2=C. |last3=Jeet |first3=J. |last4=Hudson |first4=E.R.| title = Radiative lifetime and energy of the low-energy isomeric level in 229Th| journal = Phys. Rev. C| volume = 92| year = 2015| pages = 054324| doi = 10.1103/PhysRevC.92.054324| arxiv=1509.09101}}
14. ^{{cite journal| last1=Minkov |first1=N. |last2=Pálffy |first2=A.| title = Reduced transition probabilities for the gamma decay of the 7.8 eV isomer in 229mTh | journal = Phys. Rev. Lett.| volume = 118| year = 2017| pages = 212501| doi = 10.1103/PhysRevLett.118.212501| arxiv=1704.07919}}
15. ^Varga et al.: Determination of the 229Th half-life In: Phys. Rev. C. 89, 064310 (2014); 10.1103/PhysRevC.89.064310.
16. ^{{cite journal| first1=C.J. |last1=Campbell| first2=A.G. |last2=Radnaev| first3=A. |last3=Kuzmich| title = Wigner crystals of 229Th for optical excitation of the nuclear isomer| journal = Phys. Rev. Lett.| volume = 106| pages = 223001| date = 2011| doi = 10.1103/PhysRevLett.106.223001| arxiv= 1110.2339 }}
17. ^{{cite journal | journal=Nature| volume=533 | issue=7601 | pages=47–51| date=5 May 2016| title = Direct detection of the 229Th nuclear clock transition| first1=Lars | last1=von der Wense| first2=Benedict | last2=Seiferle| first3=Mustapha | last3=Laatiaoui| first4=Jürgen B. | last4=Neumayr| first5=Hans-Jörg | last5=Maier| first6=Hans-Friedrich | last6=Wirth| first7=Christoph | last7=Mokry| first8=Jörg | last8=Runke| first9=Klaus | last9=Eberhardt| first10=Christoph E. | last10=Düllmann| first11=Norbert G. | last11=Trautmann| first12=Peter G. | last12=Thirolf| doi=10.1038/nature17669| pmid=27147026 | arxiv=1710.11398| bibcode=2016Natur.533...47V}}
18. ^{{cite journal| last1=Thielking |first1=J. |last2=Okhapkin |first2=M.V. |last3=Przemyslaw |first3=G.|last4=Meier |first4=D.M. |last5=von der Wense |first5=L. |last6=Seiferle |first6=B.|last7=Düllmann |first7=C.E. |last8=Thirolf |first8=P.G. |last9=Peik |first9=E.| title = Laser spectroscopic characterization of the nuclear-clock isomer 229mTh| journal = Nature| volume = 556| year = 2018| pages = 321| doi = 10.1038/s41586-018-0011-8| arxiv=1709.05325}}
19. ^{{cite journal| first1=L.A. |last1=Kroger| first2=C.W. |last2=Reich| title = Features of the low energy level scheme of 229Th as observed in the α decay of 233U| journal = Nucl. Phys. A| volume = 259| pages = 29-60| date = 1976| doi = 10.1016/0375-9474(76)90494-2}}
20. ^{{cite journal| title = Energy separation of the doublet of intrinsic states at the ground state of 229Th|author1=Reich, C. W. |author2=Helmer, R. G.| journal = Phys. Rev. Lett.| volume = 64| issue = 3| pages = 271–273|date=Jan 1990| doi = 10.1103/PhysRevLett.64.271|pmid=10041937 | url = http://link.aps.org/doi/10.1103/PhysRevLett.64.271| publisher = American Physical Society|bibcode = 1990PhRvL..64..271R }}
21. ^{{cite journal | journal=Physical Review C| volume=49 | issue=4| pages=1845–1858 | date=April 1994 | last1=Helmer |first1=R. G. |last2=Reich |first2=C. W.| title = An Excited State of 229Th at 3.5 eV | doi=10.1103/PhysRevC.49.1845 | bibcode = 1994PhRvC..49.1845H}}
22. ^{{cite journal| first1=E.V. |last1=Tkalya| first2=V.O. |last2=Varlamov| first3=V.V. |last3=Lomonosov| first4=S.A. |last4=Nikulin| title = Processes of the nuclear isomer 229mTh(3/2+, 3.5±1.0 eV) Resonant excitation by optical photons| journal = Physica Scripta| volume = 53| pages = 296-299| date = 1996}}
23. ^{{cite journal|author=B. R. Beck|title=Energy splitting in the ground state doublet in the nucleus 229Th |journal=Physical Review Letters|volume= 98|pages=142501|date=2007-04-06 |doi=10.1103/PhysRevLett.98.142501 |pmid=17501268 |bibcode=2007PhRvL..98n2501B|issue=14|display-authors=etal}}
24. ^{{Cite journal|last=Jeet|first=Justin|date=2015|title=Results of a Direct Search Using Synchrotron Radiation for the Low-Energy|journal=Physical Review Letters|volume=114|issue=25|doi=10.1103/physrevlett.114.253001|arxiv=1502.02189|bibcode=2015PhRvL.114y3001J}}
25. ^{{Cite journal|last=Yamaguchi|first=A.|last2=Kolbe|first2=M.|last3=Kaser|first3=H.|last4=Reichel|first4=T.|last5=Gottwald|first5=A.|last6=Peik|first6=E.|date=2015|title=Experimental search for the low-energy nuclear transition in 229 Th with undulator radiation|url=http://stacks.iop.org/1367-2630/17/i=5/a=053053|journal=New Journal of Physics|language=en|volume=17|issue=5|pages=053053|doi=10.1088/1367-2630/17/5/053053|issn=1367-2630|bibcode=2015NJPh...17e3053Y}}
26. ^{{cite book| first1=L. |last1= von der Wense| title = On the direct detection of 229mTh| journal = Springer Theses, Berlin| date = 2018| ISBN=978-3-319-70461-6| url = https://edoc.ub.uni-muenchen.de/20492/7/Wense_Lars_von_der.pdf}}
27. ^{{cite journal| title=Attempt to optically excite excite the nuclear isomer in 229Th| first1=S. |last1=Stellmer| first2=G. |last2=Kazakov| first3=M. |last3=Schreitl| first4=H. |last4=Kaser| first5=M. |last5=Kolbe| first6=T. |last6=Schumm| journal=Phys. Rev. A |volume=97 |pages=062506 |date=2018| doi= 10.1103/PhysRevA.97.062506| arxiv=1803.09294}}
28. ^{{cite journal|last=Zhao|first=Xinxin|year=2012|title=Observation of the Deexcitation of the ^{229m}Th Nuclear Isomer|journal=Physical Review Letters|volume=109|issue=16|doi=10.1103/PhysRevLett.109.160801|issn=0031-9007|author2=Yenny Natali Martinez de Escobar|author3=Robert Rundberg,|author4=Evelyn M. Bond,|author5=Allen Moody|author6=David J. Vieira|bibcode=2012PhRvL.109p0801Z}}
29. ^{{Cite arxiv|last=Borisyuk|first=P. V.|last2=Chubunova|first2=E. V.|last3=Kolachevsky|first3=N. N.|last4=Lebedinskii|first4=Yu Yu|last5=Vasiliev|first5=O. S.|last6=Tkalya|first6=E. V.|date=2018-04-01|title=Excitation of 229Th nuclei in laser plasma: the energy and half-life of the low-lying isomeric state|eprint=1804.00299|class=nucl-th}}
30. ^{{Cite journal|last=Peik|first=Ekkehard|last2=Zimmermann|first2=Kai|date=2013-07-03|title=Comment on "Observation of the Deexcitation of the 229mTh Nuclear Isomer"|journal=Physical Review Letters|volume=111|issue=1|pages=018901|doi=10.1103/PhysRevLett.111.018901|pmid=23863029|bibcode=2013PhRvL.111a8901P}}
31. ^{{cite journal| last1=Seiferle |first1=B. |last2=von der Wense |first2=L. |last3=Thirolf |first3=P.G.| title = Lifetime measurement of the 229Th nuclear isomer | journal = Phys. Rev. Lett.| volume = 118| year = 2017| pages = 042501| doi = 10.1103/PhysRevLett.118.042501| arxiv=1801.05205}}
32. ^{{Cite arxiv|last=Seiferle|first=B.|last2=von der Wense|first2=L.|last3=Amersdorffer|first3=I.|last4=Arlt|first4=N.|last5=Kotulski|first5=B.|last6=Thirolf|first6=P.G.|date=2018|title=Towards a precise determination of the excitation energy of the thorium nuclear isomer using a magnetic bottle spectrometer|eprint=1812.04621|class=physics.ins-det]}}
33. ^{{cite journal| last1=Thirolf |first1=P.G. |last2=Seiferle |first2=B. |last3=von der Wense |first3=L.| title = Improving Our Knowledge on the 229mThorium Isomer: Toward a Test Bench for Time Variations of Fundamental Constants | journal = Annalen der Physik| year = 2019| pages = 1800381| doi = 10.1002/andp.201800381}}
34. ^{{cite journal| last1=Flambaum |first1=V.V.| title = Enhanced Effect of Temporal Variation of the Fine Structure Constant and the Strong Interaction in 229mTh| journal = Phys. Rev. Lett.| volume = 97| year = 2006| pages = 092502| doi = 10.1103/PhysRevLett.97.092502| arxiv=physics/0604188}}

1 : Nuclear physics

随便看

 

开放百科全书收录14589846条英语、德语、日语等多语种百科知识,基本涵盖了大多数领域的百科知识,是一部内容自由、开放的电子版国际百科全书。

 

Copyright © 2023 OENC.NET All Rights Reserved
京ICP备2021023879号 更新时间:2024/9/25 22:17:17