Network


Latest external collaboration on country level. Dive into details by clicking on the dots.

Hotspot


Dive into the research topics where J. V. Kratz is active.

Publication


Featured researches published by J. V. Kratz.


Nuclear Physics | 1997

Invariant-mass spectroscopy of 10Li and 11Li

M. Zinser; F. Humbert; T. Nilsson; W. Schwab; Horst Simon; T. Aumann; M. J. G. Borge; L. V. Chulkov; J. Cub; Th. W. Elze; H. Emling; H. Geissel; D. Guillemaud-Mueller; P. G. Hansen; R. Holzmann; H. Irnich; B. Jonson; J. V. Kratz; R. Kulessa; Y. Leifels; H. Lenske; A. Magel; A. C. Mueller; G. Münzenberg; F. Nickel; G. Nyman; A. Richter; K. Riisager; C. Scheidenberger; G. Schrieder

Break-up of secondary Li-11 ion beams (280 MeV/nucleon) on C and Pb targets into Li-9 and neutrons is studied experimentally. Cross sections and neutron multiplicity distributions are obtained, characterizing different reaction mechanisms. Invariant-mass spectroscopy for Li-11 and Li-10 is performed. The E1 strength distribution, deduced from electromagnetic excitation of Li-11 up to an excitation energy of 4 MeV comprises similar to 8% of the Thomas-Reiche-Kuhn energy-weighted sumrule strength. Two low-lying resonance-like structures are observed for Li-10 at decay energies of 0.21(5) and 0.62(10) MeV, the former one carrying 26(10)% of the strength and likely to be associated with an s-wave neutron decay. A strong di-neutron correlation in Li-11 can be discarded. Calculations in a quasi-particle RPA approach are compared with the experimental results for Li-10 and Li-11


Nuclear Instruments & Methods in Physics Research Section A-accelerators Spectrometers Detectors and Associated Equipment | 1992

A Large area detector for high-energy neutrons

Th. Blaich; Th. W. Elze; H. Emling; H. Freiesleben; K. Grimm; W. Henning; R. Holzmann; G. Ickert; J. G. Keller; H. Klingler; W. Kneissl; R. König; R. Kulessa; J. V. Kratz; D. Lambrecht; J.S. Lange; Y. Leifels; E. Lubkiewicz; M. Proft; W. Prokopowicz; C. Schütter; R. Schmidt; H. Spies; K. Stelzer; J. Stroth; W. Walus; E. Wajda; H. J. Wollersheim; M. Zinser; E. Zude

Abstract We present design studies, results of test measurements, and Monte Carlo simulations which served as a basis for the realization of a large area neutron detector (LAND). It has a front area of 2m×2m and a depth of 1 m, and features a multilayer structure of passive converter and active scintillator material. The detector is subdivided in independently operating paddles which allow time-of-flight and position measurement. An energy resolution of ΔT n / T n =5.3% for a flight path of 15 m and an overall detection efficiency of ϵ ≈ 1 is anticipated for neutrons with T n ≈ 1 GeV. The operation of LAND at the SIS facility of GSI is described.


Nature | 1997

Chemical properties of element 106 (seaborgium)

M. Schädel; W. Brüchle; R. Dressler; B. Eichler; H. W. Gäggeler; R. Günther; Kenneth E. Gregorich; Darleane C. Hoffman; S. Hübener; D.T. Jost; J. V. Kratz; W. Paulus; D. Schumann; S. N. Timokhin; N. Trautmann; A. Türler; G. Wirth; A. Yakuschev

The synthesis, via nuclear fusion reactions, of elements heavier than the actinides, allows one to probe the limits of the periodic table as a means of classifying the elements. In particular, deviations in the periodicity of chemical properties for the heaviest elements are predicted as a consequence of increasingly strong relativistic effects on the electronic shell structure. The transactinide elements have now been extended up to element 112 (ref. 8), but the chemical properties have been investigated only for the first two of the transactinide elements, 104 and 105 (refs 9,10,11,12,13,14,15,16,17,18,19). Those studies showed that relativistic effect render these two elements chemically different from their lighter homologues in the same columns of the periodic table (Fig. 1). Here we report the chemical separation of element 106 (seaborgium, Sg) and investigations of its chemical behaviour in the gas phase and in aqueous solution. The methods that we use are able to probe the reactivity of individual atoms, and based on the detection of just seven atoms of seaborgium we find that it exhibits properties characteristic of the group 6 homologues molybdenum and tungsten. Thus seaborgium appears to restore the trends of the periodic table disrupted by relativistic effects in elements 104 and 105.


Nuclear Physics | 1985

Fusion near the threshold: A comparative study of the systems 40Ar + 112, 116, 122sn and 40Ar + 144, 148, 154Sm

W. Reisdorf; F.P. Hessberger; K.D. Hildenbrand; S. Hofmann; G. Münzenberg; K. H. Schmidt; J.H.R. Schneider; W.F.W. Schneider; K. Sümmerer; G. Wirth; J. V. Kratz; K. Schlitt

Abstract Fusion excitation functions for the systems 40 Ar + 112, 116, 122 Sn and 40 Ar + 144, 148, 154 Sm have been determined, covering cross sections ranging from several hundred mb down to the μb level. The data show a pronounced correlation of the subbarrier behaviour with low-energy collective properties of the nuclei involved and are well reproduced by simplified coupled-channel calculations coupling fusion to inelastic channels. The possibilities of parameterizing the data in terms of a simple dynamic barrier-fluctuation phenomenon are discussed and result in the prediction of remarkably diffuse partial-wave distributions above the barrier. This is shown to be important for the analysis of deexcitation phenomena following fusion reactions.


Spectrochimica Acta Part B: Atomic Spectroscopy | 2000

Production of monodisperse uranium oxide particles and their characterization by scanning electron microscopy and secondary ion mass spectrometry

Nicole Erdmann; Maria Betti; O Stetzer; Gabriele Tamborini; J. V. Kratz; N. Trautmann; J. Van Geel

Abstract Secondary ion mass spectrometry (SIMS) can be confidently used to measure uranium isotopic ratios in single particles. Dense particles of known isotopic composition and size allow the precision and the accuracy of the applied procedure to be estimated. These particles can be obtained by dissolving standard reference uranium materials, nebulizing the solution in droplets of proper diameter and collecting the particles after the desolvation and calcination of the droplets. A new instrumental set up, based on a commercial vibrating orifice aerosol generator to generate monodisperse droplets of the solutions from four uranium oxide reference materials, is described. The droplets were dried and calcined in a sequence of three furnaces. The morphology of the monodisperse uranium oxide particles was studied by scanning electron microscopy. It was observed that the particles were nearly spherical and consisted of dense material. Their diameter distribution evidenced the presence of two populations mainly, the first showing a narrow distribution with a maximum centered at approximately 1 μm. The first statistical moment ratios between the two populations remained practically constant at 1.24±0.01. This demonstrated that the second population was due to the formation of one particle from two droplets of solution (theoretical double mass≡diameter ratio of 2 3 =1.26). Secondary ion mass spectrometry was used to verify the isotopic composition of the produced particles. Typical accuracies of better than 0.4% for 235U/238U and a few percent for the minor isotopes have been achieved. For the determination of the 236U content, the signal at mass M=239 (due to 238UH+) was used to correct the 235UH+ contribution to 236U at mass M=236, greatly improving the accuracy of the 236/238 ratio with increasing enrichment of the 235U isotope.


Journal of Alloys and Compounds | 1998

Determination of the first ionization potential of nine actinide elements by resonance ionization mass spectroscopy (RIMS)

Nicole Erdmann; M. Nunnemann; K. Eberhardt; G. Herrmann; G. Huber; S. Köhler; J. V. Kratz; G. Passler; J.R. Peterson; N. Trautmann; A. Waldek

The high sensitivity of RIMS enables the precise determination of the first ionization potential of actinide elements with a sample size of ≤1012 atoms. By multiple resonant laser excitation, the actinide atoms under investigation are ionized in the presence of an electric field, and the ions are mass-selectively detected in a time-of-flight spectrometer. The first ionization potential is obtained by scanning the wavelength of the laser used for the last excitation step across the ionization threshold Wth—indicated by a sudden increase of the ion count rate—at various electric field strengths. Extrapolation of Wth to electric field strength zero leads directly to the first ionization potential. The first ionization potentials (IP) of Am, Cm, Bk, Cf and Es were determined for the first time as IPAm=5.9736(3) eV, IPCm=5.9914(2) eV, IPBk=6.1979(2) eV, IPCf=6.2817(2) eV, IPEs=6.3676(5) eV with samples of 1012 atoms. Furthermore, the ionization potentials of Th, U, Np and Pu were remeasured.


Radiochimica Acta | 1989

ARCA II – A New Apparatus for Fast, Repetitive HPLC Separations

M. Schädel; W. Brüchle; Egon Jäger; E. Schimpf; J. V. Kratz; U. W. Scherer; H. P. Zimmermann

The microcomputer controlled Automated Rapid Chemistry Apparatus, ARCA, is described in its newly designed version for the study of chemical properties of element 105 in aqueous solutions. This improved version, ARCA II, is adapted to the needs of fast and repetitive separations to be carried out in a chemically inert automated micro high performance liquid chromatography system. As an example, the separation of several group HIB, IVB, and VB elements in the system triisooctylamine/hydrochloric acid within 30 s is demonstrated. Furthermore, a new method for the fast preparation of samples for α-particle spectroscopy by evaporation of the aqueous effluent with an intense light source is presented.


Nature | 2015

Measurement of the first ionization potential of lawrencium, element 103

T. K. Sato; M. Asai; A. Borschevsky; T. Stora; N. Sato; Y. Kaneya; K. Tsukada; Ch. E. Düllmann; K. Eberhardt; E. Eliav; S. Ichikawa; U. Kaldor; J. V. Kratz; Sunao Miyashita; Y. Nagame; K. Ooe; A. Osa; D. Renisch; J. Runke; M. Schädel; P. Thörle-Pospiech; A. Toyoshima; N. Trautmann

The chemical properties of an element are primarily governed by the configuration of electrons in the valence shell. Relativistic effects influence the electronic structure of heavy elements in the sixth row of the periodic table, and these effects increase dramatically in the seventh row—including the actinides—even affecting ground-state configurations. Atomic s and p1/2 orbitals are stabilized by relativistic effects, whereas p3/2, d and f orbitals are destabilized, so that ground-state configurations of heavy elements may differ from those of lighter elements in the same group. The first ionization potential (IP1) is a measure of the energy required to remove one valence electron from a neutral atom, and is an atomic property that reflects the outermost electronic configuration. Precise and accurate experimental determination of IP1 gives information on the binding energy of valence electrons, and also, therefore, on the degree of relativistic stabilization. However, such measurements are hampered by the difficulty in obtaining the heaviest elements on scales of more than one atom at a time. Here we report that the experimentally obtained IP1 of the heaviest actinide, lawrencium (Lr, atomic number 103), is electronvolts. The IP1 of Lr was measured with 256Lr (half-life 27 seconds) using an efficient surface ion-source and a radioisotope detection system coupled to a mass separator. The measured IP1 is in excellent agreement with the value of 4.963(15) electronvolts predicted here by state-of-the-art relativistic calculations. The present work provides a reliable benchmark for theoretical calculations and also opens the way for IP1 measurements of superheavy elements (that is, transactinides) on an atom-at-a-time scale.


Radiochimica Acta | 1997

First Aqueous Chemistry with Seaborgium (Element 106)

M. Schädel; W. Brüchle; B. Schausten; E. Schimpf; E. Jager; G. Wirth; R. Günther; J. V. Kratz; W. Paulus; A. Seibert; P. Thörle; N. Trautmann; S. Zauner; D. Schumann; M. Andrassy; R. Misiak; K. E. Gregorich; Darleane C. Hoffman; D. M. Lee; E. R. Sylwester; Y. Nagame; Y. Oura

For the first time, chemical separations of element 106 (Seaborgium, Sg) were performed in aqueous solutions. The isotopes Sg and Sg were produced in the Cm + Ne reaction at a beam energy of 121 MeV. The reaction products were continuously transported by a He(KCl)-jet to the computer-controlled liquid chromatography system ARCA. In 0.1 M HNO3/5 X ΙΟ -4 M HF, Sg was found to be eluted within 10 s from 1.6X8 mm cation-exchange columns (Aminex A6, 17.5±2 μπι) together with the hexavalent Moand W-ions, while hexavalent U-ions and tetravalent Zr-, Hf-, and element 104 ions were strongly retained on the column. Element 106 was detected by measuring correlated α-decays of the daughter isotopes 78-s 104 and 26-s 102. For the isotope Sg, we have evidence for a spontaneous fission branch. It yields a partial spontaneousfission half-life which is in agreement with recent theoretical predictions. The chemical results show that the most stable oxidation state of Sg in aqueous solution is +6, and that like its homologs Mo and W, Sg forms neutral or anionic oxoor oxohalide-compounds under the present condition. In these first experiments, Sg exhibits properties very characteristic of group 6 elements, and does not show U-like properties.


Journal of Alloys and Compounds | 1998

Trace analysis of plutonium in environmental samples by resonance ionization mass spectroscopy (RIMS)

M. Nunnemann; Nicole Erdmann; H.-U Hasse; G. Huber; J. V. Kratz; Peter Kunz; A. Mansel; G. Passler; O. Stetzer; N. Trautmann; A. Waldek

Abstract Resonance ionization mass spectroscopy (RIMS) is well suited for trace analysis of long-lived radioisotopes in environmental, biological and technical samples. By multiple resonant laser excitation and ionization of the elemental atoms under investigation, an extremely high element selectivity can be achieved. In addition, isotope selectivity is obtained by subsequent mass analysis. The excellent sensitivity results from the large atomic cross-sections in the excitation–ionization process and the good detection efficiency for ions. The element selectivity of RIMS allows a simplified procedure for the chemical preparation of the samples compared to the requirements of thin sources for α-spectroscopy. Various samples have been determined by RIMS with respect to their content and the isotopic composition of plutonium in the ultra-trace regime. A detection limit of 10 6 to 10 7 plutonium atoms has been achieved for all isotopes, independent of their half-life and decay mode. For 239 Pu, this value is distinctly below the radiometric detection limit.

Collaboration


Dive into the J. V. Kratz's collaboration.

Top Co-Authors

Avatar

R. Kulessa

Jagiellonian University

View shared research outputs
Top Co-Authors

Avatar

H. Emling

GSI Helmholtz Centre for Heavy Ion Research

View shared research outputs
Top Co-Authors

Avatar

H. Geissel

GSI Helmholtz Centre for Heavy Ion Research

View shared research outputs
Top Co-Authors

Avatar

Th. W. Elze

Goethe University Frankfurt

View shared research outputs
Top Co-Authors

Avatar

Y. Leifels

Ruhr University Bochum

View shared research outputs
Top Co-Authors

Avatar

M. Schädel

Japan Atomic Energy Agency

View shared research outputs
Top Co-Authors

Avatar

K. Boretzky

GSI Helmholtz Centre for Heavy Ion Research

View shared research outputs
Top Co-Authors

Avatar
Top Co-Authors

Avatar

T. Aumann

Technische Universität Darmstadt

View shared research outputs
Top Co-Authors

Avatar

W. Brüchle

Petersburg Nuclear Physics Institute

View shared research outputs
Researchain Logo
Decentralizing Knowledge