fbpx
Wikipedia

Lawrencium

Lawrencium is a synthetic chemical element; it has symbol Lr (formerly Lw) and atomic number 103. It is named in honor of Ernest Lawrence, inventor of the cyclotron, a device that was used to discover many artificial radioactive elements. A radioactive metal, lawrencium is the eleventh transuranic element and the last member of the actinide series. Like all elements with atomic number over 100, lawrencium can only be produced in particle accelerators by bombarding lighter elements with charged particles. Fourteen isotopes of lawrencium are currently known; the most stable is 266Lr with half-life 11 hours, but the shorter-lived 260Lr (half-life 2.7 minutes) is most commonly used in chemistry because it can be produced on a larger scale.

Lawrencium, 103Lr
Lawrencium
Pronunciation/lɒˈrɛnsiəm/ (lo-REN-see-əm)
Appearancesilvery (predicted)[1]
Mass number[266]
Lawrencium in the periodic table
Hydrogen Helium
Lithium Beryllium Boron Carbon Nitrogen Oxygen Fluorine Neon
Sodium Magnesium Aluminium Silicon Phosphorus Sulfur Chlorine Argon
Potassium Calcium Scandium Titanium Vanadium Chromium Manganese Iron Cobalt Nickel Copper Zinc Gallium Germanium Arsenic Selenium Bromine Krypton
Rubidium Strontium Yttrium Zirconium Niobium Molybdenum Technetium Ruthenium Rhodium Palladium Silver Cadmium Indium Tin Antimony Tellurium Iodine Xenon
Caesium Barium Lanthanum Cerium Praseodymium Neodymium Promethium Samarium Europium Gadolinium Terbium Dysprosium Holmium Erbium Thulium Ytterbium Lutetium Hafnium Tantalum Tungsten Rhenium Osmium Iridium Platinum Gold Mercury (element) Thallium Lead Bismuth Polonium Astatine Radon
Francium Radium Actinium Thorium Protactinium Uranium Neptunium Plutonium Americium Curium Berkelium Californium Einsteinium Fermium Mendelevium Nobelium Lawrencium Rutherfordium Dubnium Seaborgium Bohrium Hassium Meitnerium Darmstadtium Roentgenium Copernicium Nihonium Flerovium Moscovium Livermorium Tennessine Oganesson
Lu

Lr

(Ups)
nobeliumlawrenciumrutherfordium
Atomic number (Z)103
Groupgroup 3
Periodperiod 7
Block  d-block
Electron configuration[Rn] 5f14 7s2 7p1
Electrons per shell2, 8, 18, 32, 32, 8, 3
Physical properties
Phase at STPsolid (predicted)
Melting point1900 K ​(1600 °C, ​3000 °F) (predicted)
Density (near r.t.)14.4 g/cm3 (predicted)[2]
Atomic properties
Oxidation states+3
ElectronegativityPauling scale: 1.3 (predicted)[3]
Ionization energies
  • 1st: 479 kJ/mol[4]
  • 2nd: 1428.0 kJ/mol (predicted)
  • 3rd: 2219.1 kJ/mol (predicted)
Other properties
Natural occurrencesynthetic
Crystal structurehexagonal close-packed (hcp)

(predicted)[5]
CAS Number22537-19-5
History
Namingafter Ernest Lawrence
DiscoveryLawrence Berkeley National Laboratory and Joint Institute for Nuclear Research (1961–1971)
Isotopes of lawrencium
Main isotopes[6] Decay
abun­dance half-life (t1/2) mode pro­duct
256Lr synth 27.9 s α 252Md
β+ 256No
260Lr synth 3.0 min α 256Md
β+ 260No
261Lr synth 39 min SF
262Lr synth 4 h β+ 262No
264Lr synth 4.8 h[7] SF
266Lr synth 11 h SF
 Category: Lawrencium
| references

Chemistry experiments confirm that lawrencium behaves as a heavier homolog to lutetium in the periodic table, and is a trivalent element. It thus could also be classified as the first of the 7th-period transition metals. Its electron configuration is anomalous for its position in the periodic table, having an s2p configuration instead of the s2d configuration of its homolog lutetium. However, this does not appear to affect lawrencium's chemistry.

In the 1950s, 1960s, and 1970s, many claims of the synthesis of lawrencium of varying quality were made from laboratories in the Soviet Union and the United States. The priority of the discovery and therefore the name of the element was disputed between Soviet and American scientists. The International Union of Pure and Applied Chemistry (IUPAC) initially established lawrencium as the official name for the element and gave the American team credit for the discovery; this was reevaluated in 1997, giving both teams shared credit for the discovery but not changing the element's name.

Introduction edit

Synthesis of superheavy nuclei edit

 
A graphic depiction of a nuclear fusion reaction. Two nuclei fuse into one, emitting a neutron. Reactions that created new elements to this moment were similar, with the only possible difference that several singular neutrons sometimes were released, or none at all.

A superheavy[a] atomic nucleus is created in a nuclear reaction that combines two other nuclei of unequal size[b] into one; roughly, the more unequal the two nuclei in terms of mass, the greater the possibility that the two react.[13] The material made of the heavier nuclei is made into a target, which is then bombarded by the beam of lighter nuclei. Two nuclei can only fuse into one if they approach each other closely enough; normally, nuclei (all positively charged) repel each other due to electrostatic repulsion. The strong interaction can overcome this repulsion but only within a very short distance from a nucleus; beam nuclei are thus greatly accelerated in order to make such repulsion insignificant compared to the velocity of the beam nucleus.[14] The energy applied to the beam nuclei to accelerate them can cause them to reach speeds as high as one-tenth of the speed of light. However, if too much energy is applied, the beam nucleus can fall apart.[14]

Coming close enough alone is not enough for two nuclei to fuse: when two nuclei approach each other, they usually remain together for approximately 10−20 seconds and then part ways (not necessarily in the same composition as before the reaction) rather than form a single nucleus.[14][15] This happens because during the attempted formation of a single nucleus, electrostatic repulsion tears apart the nucleus that is being formed.[14] Each pair of a target and a beam is characterized by its cross section—the probability that fusion will occur if two nuclei approach one another expressed in terms of the transverse area that the incident particle must hit in order for the fusion to occur.[c] This fusion may occur as a result of the quantum effect in which nuclei can tunnel through electrostatic repulsion. If the two nuclei can stay close for past that phase, multiple nuclear interactions result in redistribution of energy and an energy equilibrium.[14]

External videos
  Visualization of unsuccessful nuclear fusion, based on calculations from the Australian National University[17]

The resulting merger is an excited state[18]—termed a compound nucleus—and thus it is very unstable.[14] To reach a more stable state, the temporary merger may fission without formation of a more stable nucleus.[19] Alternatively, the compound nucleus may eject a few neutrons, which would carry away the excitation energy; if the latter is not sufficient for a neutron expulsion, the merger would produce a gamma ray. This happens in approximately 10−16 seconds after the initial nuclear collision and results in creation of a more stable nucleus.[19] The definition by the IUPAC/IUPAP Joint Working Party (JWP) states that a chemical element can only be recognized as discovered if a nucleus of it has not decayed within 10−14 seconds. This value was chosen as an estimate of how long it takes a nucleus to acquire its outer electrons and thus display its chemical properties.[20][d]

Decay and detection edit

The beam passes through the target and reaches the next chamber, the separator; if a new nucleus is produced, it is carried with this beam.[22] In the separator, the newly produced nucleus is separated from other nuclides (that of the original beam and any other reaction products)[e] and transferred to a surface-barrier detector, which stops the nucleus. The exact location of the upcoming impact on the detector is marked; also marked are its energy and the time of the arrival.[22] The transfer takes about 10−6 seconds; in order to be detected, the nucleus must survive this long.[25] The nucleus is recorded again once its decay is registered, and the location, the energy, and the time of the decay are measured.[22]

Stability of a nucleus is provided by the strong interaction. However, its range is very short; as nuclei become larger, its influence on the outermost nucleons (protons and neutrons) weakens. At the same time, the nucleus is torn apart by electrostatic repulsion between protons, and its range is not limited.[26] Total binding energy provided by the strong interaction increases linearly with the number of nucleons, whereas electrostatic repulsion increases with the square of the atomic number, i.e. the latter grows faster and becomes increasingly important for heavy and superheavy nuclei.[27][28] Superheavy nuclei are thus theoretically predicted[29] and have so far been observed[30] to predominantly decay via decay modes that are caused by such repulsion: alpha decay and spontaneous fission.[f] Almost all alpha emitters have over 210 nucleons,[32] and the lightest nuclide primarily undergoing spontaneous fission has 238.[33] In both decay modes, nuclei are inhibited from decaying by corresponding energy barriers for each mode, but they can be tunnelled through.[27][28]

 
Scheme of an apparatus for creation of superheavy elements, based on the Dubna Gas-Filled Recoil Separator set up in the Flerov Laboratory of Nuclear Reactions in JINR. The trajectory within the detector and the beam focusing apparatus changes because of a dipole magnet in the former and quadrupole magnets in the latter.[34]

Alpha particles are commonly produced in radioactive decays because mass of an alpha particle per nucleon is small enough to leave some energy for the alpha particle to be used as kinetic energy to leave the nucleus.[35] Spontaneous fission is caused by electrostatic repulsion tearing the nucleus apart and produces various nuclei in different instances of identical nuclei fissioning.[28] As the atomic number increases, spontaneous fission rapidly becomes more important: spontaneous fission partial half-lives decrease by 23 orders of magnitude from uranium (element 92) to nobelium (element 102),[36] and by 30 orders of magnitude from thorium (element 90) to fermium (element 100).[37] The earlier liquid drop model thus suggested that spontaneous fission would occur nearly instantly due to disappearance of the fission barrier for nuclei with about 280 nucleons.[28][38] The later nuclear shell model suggested that nuclei with about 300 nucleons would form an island of stability in which nuclei will be more resistant to spontaneous fission and will primarily undergo alpha decay with longer half-lives.[28][38] Subsequent discoveries suggested that the predicted island might be further than originally anticipated; they also showed that nuclei intermediate between the long-lived actinides and the predicted island are deformed, and gain additional stability from shell effects.[39] Experiments on lighter superheavy nuclei,[40] as well as those closer to the expected island,[36] have shown greater than previously anticipated stability against spontaneous fission, showing the importance of shell effects on nuclei.[g]

Alpha decays are registered by the emitted alpha particles, and the decay products are easy to determine before the actual decay; if such a decay or a series of consecutive decays produces a known nucleus, the original product of a reaction can be easily determined.[h] (That all decays within a decay chain were indeed related to each other is established by the location of these decays, which must be in the same place.)[22] The known nucleus can be recognized by the specific characteristics of decay it undergoes such as decay energy (or more specifically, the kinetic energy of the emitted particle).[i] Spontaneous fission, however, produces various nuclei as products, so the original nuclide cannot be determined from its daughters.[j]

The information available to physicists aiming to synthesize a superheavy element is thus the information collected at the detectors: location, energy, and time of arrival of a particle to the detector, and those of its decay. The physicists analyze this data and seek to conclude that it was indeed caused by a new element and could not have been caused by a different nuclide than the one claimed. Often, provided data is insufficient for a conclusion that a new element was definitely created and there is no other explanation for the observed effects; errors in interpreting data have been made.[k]

History edit

 
Albert Ghiorso updating the periodic table in April 1961, writing the symbol "Lw" in as element 103. Codiscoverers Latimer, Sikkeland, and Larsh (left to right) look on.

In 1958, scientists at Lawrence Berkeley National Laboratory claimed the discovery of element 102, now called nobelium. At the same time, they also tried to synthesize element 103 by bombarding the same curium target used with nitrogen-14 ions. Eighteen tracks were noted, with decay energy around 9±MeV and half-life around 0.25 s; the Berkeley team noted that while the cause could be the production of an isotope of element 103, other possibilities could not be ruled out. While the data agrees reasonably with that later discovered for 257Lr (alpha decay energy 8.87 MeV, half-life 0.6 s), the evidence obtained in this experiment fell far short of the strength required to conclusively demonstrate synthesis of element 103. A follow-up on this experiment was not done, as the target was destroyed.[51][52] Later, in 1960, the Lawrence Berkeley Laboratory attempted to synthesize the element by bombarding 252Cf with 10B and 11B. The results of this experiment were not conclusive.[51]

The first important work on element 103 was done at Berkeley by the nuclear-physics team of Albert Ghiorso, Torbjørn Sikkeland, Almon Larsh, Robert M. Latimer, and their co-workers on February 14, 1961.[53] The first atoms of lawrencium were reportedly made by bombarding a three-milligram target consisting of three isotopes of californium with boron-10 and boron-11 nuclei from the Heavy Ion Linear Accelerator (HILAC).[54] The Berkeley team reported that the isotope 257103 was detected in this manner, and that it decayed by emitting an 8.6 MeV alpha particle with a half-life of 8±2 s.[52] This identification was later corrected to 258103,[54] as later work proved that 257Lr did not have the properties detected, but 258Lr did.[52] This was considered at the time to be convincing proof of synthesis of element 103: while the mass assignment was less certain and proved to be mistaken, it did not affect the arguments in favor of element 103 having been synthesized. Scientists at Joint Institute for Nuclear Research in Dubna (then in the Soviet Union) raised several criticisms: all but one were answered adequately. The exception was that 252Cf was the most common isotope in the target, and in the reactions with 10B, 258Lr could only have been produced by emitting four neutrons, and emitting three neutrons was expected to be much less likely than emitting four or five. This would lead to a narrow yield curve, not the broad one reported by the Berkeley team. A possible explanation was that there was a low number of events attributed to element 103.[52] This was an important intermediate step to the unquestioned discovery of element 103, although the evidence was not completely convincing.[52] The Berkeley team proposed the name "lawrencium" with symbol "Lw", after Ernest Lawrence, inventor of the cyclotron. The IUPAC Commission on Nomenclature of Inorganic Chemistry accepted the name, but changed the symbol to "Lr".[55] This acceptance of the discovery was later characterized as being hasty by the Dubna team.[52]

252
98
Cf
+ 11
5
B
263
103
Lr
* → 258
103
Lr
+ 5 1
0
n

The first work at Dubna on element 103 came in 1965, when they reported to have made 256103 in 1965 by bombarding 243Am with 18O, identifying it indirectly from its granddaughter fermium-252. The half-life they reported was somewhat too high, possibly due to background events. Later 1967 work on the same reaction identified two decay energies in the ranges 8.35–8.50 MeV and 8.50–8.60 MeV: these were assigned to 256103 and 257103.[52] Despite repeat attempts, they were unable to confirm assignment of an alpha emitter with a half-life of 8 seconds to 257103.[56][57] The Russians proposed the name "rutherfordium" for the new element in 1967:[51][58] this name was later proposed by Berkeley for element 104.[58]

243
95
Am
+ 18
8
O
261
103
Lr
* → 256
103
Lr
+ 5 1
0
n

Further experiments in 1969 at Dubna and in 1970 at Berkeley demonstrated an actinide chemistry for the new element; so by 1970 it was known that element 103 is the last actinide.[52][59] In 1970, the Dubna group reported the synthesis of 255103 with half-life 20 s and alpha decay energy 8.38 MeV.[52] However, it was not until 1971, when the nuclear physics team at University of California at Berkeley successfully did a whole series of experiments aimed at measuring the nuclear decay properties of the lawrencium isotopes with mass numbers 255 to 260,[60][61] that all previous results from Berkeley and Dubna were confirmed, apart from the Berkeley's group initial erroneous assignment of their first produced isotope to 257103 instead of the probably correct 258103.[52] All final doubts were dispelled in 1976 and 1977 when the energies of X-rays emitted from 258103 were measured.[52]

 
The element was named after Ernest Lawrence.

In 1971, the IUPAC granted the discovery of lawrencium to the Lawrence Berkeley Laboratory, even though they did not have ideal data for the element's existence. But in 1992, the IUPAC Transfermium Working Group (TWG) officially recognized the nuclear physics teams at Dubna and Berkeley as co-discoverers of lawrencium, concluding that while the 1961 Berkeley experiments were an important step to lawrencium's discovery, they were not yet fully convincing; and while the 1965, 1968, and 1970 Dubna experiments came very close to the needed level of confidence taken together, only the 1971 Berkeley experiments, which clarified and confirmed previous observations, finally resulted in complete confidence in the discovery of element 103.[51][55] Because the name "lawrencium" had been in use for a long time by this point, it was retained by IUPAC,[51] and in August 1997, the International Union of Pure and Applied Chemistry (IUPAC) ratified the name lawrencium and the symbol "Lr" during a meeting in Geneva.[55]

Characteristics edit

Physical edit

Lawrencium is the last actinide. Authors considering the subject generally consider it a group 3 element, along with scandium, yttrium, and lutetium, as its filled f-shell is expected to make it resemble the other 7th-period transition metals. In the periodic table, it is to the right of the actinide nobelium, to the left of the 6d transition metal rutherfordium, and under the lanthanide lutetium with which it shares many physical and chemical properties. Lawrencium is expected to be a solid under normal conditions and have a hexagonal close-packed crystal structure (c/a = 1.58), similar to its lighter congener lutetium, though this is not yet known experimentally.[5] The enthalpy of sublimation of lawrencium is estimated at 352 kJ/mol, close to the value of lutetium and strongly suggesting that metallic lawrencium is trivalent with three electrons delocalized, a prediction also supported by a systematic extrapolation of the values of heat of vaporization, bulk modulus, and atomic volume of neighboring elements to lawrencium:[62] this makes it unlike the immediately preceding late actinides which are known to be (fermium and mendelevium) or expected to be (nobelium) divalent.[63] The estimated enthalpies of vaporization show that lawrencium deviates from the trend of the late actinides and instead matches the trend of the succeeding 6d elements rutherfordium and dubnium,[64][65] consistent with lawrencium's interpretation as a group 3 element.[65] Some scientists prefer to end the actinides with nobelium and consider lawrencium to be the first transition metal of the seventh period.[66][67]

Specifically, lawrencium is expected to be a trivalent, silvery metal, easily oxidized by air, steam, and acids,[68] and having an atomic volume similar to that of lutetium and a trivalent metallic radius of 171 pm.[62] It is expected to be a rather heavy metal with a density of around 14.4 g/cm3.[2] It is also predicted to have a melting point of around 1900 K (1600 °C), not far from the value for lutetium (1925 K).[69]

Chemical edit

 
Elution sequence of the late trivalent lanthanides and actinides, with ammonium α-HIB as eluant: the broken curve for lawrencium is a prediction.

In 1949, Glenn T. Seaborg, who devised the actinide concept, predicted that element 103 (lawrencium) should be the last actinide and that the Lr3+ ion should be about as stable as Lu3+ in aqueous solution. It was not until decades later that element 103 was finally conclusively synthesized and this prediction was experimentally confirmed.[70]

1969 studies on the element showed that lawrencium reacts with chlorine to form a product that was most likely the trichloride, LrCl3. Its volatility was found to be similar to the chlorides of curium, fermium, and nobelium and much less than that of rutherfordium chloride. In 1970, chemical studies were performed on 1500 atoms of 256Lr, comparing it with divalent (No, Ba, Ra), trivalent (Fm, Cf, Cm, Am, Ac), and tetravalent (Th, Pu) elements. It was found that lawrencium coextracted with the trivalent ions, but the short half-life of 256Lr precluded a confirmation that it eluted ahead of Md3+ in the elution sequence.[70] Lawrencium occurs as the trivalent Lr3+ ion in aqueous solution and hence its compounds should be similar to those of the other trivalent actinides: for example, lawrencium(III) fluoride (LrF3) and hydroxide (Lr(OH)3) should both be insoluble in water.[70] Due to the actinide contraction, the ionic radius of Lr3+ should be smaller than that of Md3+, and it should elute ahead of Md3+ when ammonium α-hydroxyisobutyrate (ammonium α-HIB) is used as an eluant.[70] Later 1987 experiments on the longer-lived isotope 260Lr confirmed lawrencium's trivalency and that it eluted in roughly the same place as erbium, and found that lawrencium's ionic radius was 88.6±0.3 pm, larger than would be expected from simple extrapolation from periodic trends.[70] Later 1988 experiments with more lawrencium atoms refined this to 88.1±0.1 pm and calculated an enthalpy of hydration value of −3685±13 kJ/mol.[70] It was also found that the actinide contraction at the end of the actinides was larger than the analogous lanthanide contraction, with the exception of the last actinide, lawrencium: the cause was speculated to be relativistic effects.[70]

It has been speculated that the 7s electrons are relativistically stabilized, so that in reducing conditions, only the 7p1/2 electron would be ionized, leading to the monovalent Lr+ ion. However, all experiments to reduce Lr3+ to Lr2+ or Lr+ in aqueous solution were unsuccessful, similarly to lutetium. On the basis of this, the standard electrode potential of the E°(Lr3+ → Lr+) couple was calculated to be less than −1.56 V, indicating that the existence of Lr+ ions in aqueous solution was unlikely. The upper limit for the E°(Lr3+ → Lr2+) couple was predicted to be −0.44 V: the values for E°(Lr3+ → Lr) and E°(Lr4+ → Lr3+) are predicted to be −2.06 V and +7.9 V.[70] The stability of the group oxidation state in the 6d transition series decreases as RfIV > DbV > SgVI, and lawrencium continues the trend with LrIII being more stable than RfIV.[71]

In the molecule lawrencium dihydride (LrH2), which is predicted to be bent, the 6d orbital of lawrencium is not expected to play a role in the bonding, unlike that of lanthanum dihydride (LaH2). LaH2 has La–H bond distances of 2.158 Å, while LrH2 should have shorter Lr–H bond distances of 2.042 Å due to the relativistic contraction and stabilization of the 7s and 7p orbitals involved in the bonding, in contrast to the core-like 5f subshell and the mostly uninvolved 6d subshell. In general, molecular LrH2 and LrH are expected to resemble the corresponding thallium species (thallium having a 6s26p1 valence configuration in the gas phase, like lawrencium's 7s27p1) more than the corresponding lanthanide species.[72] The electron configurations of Lr+ and Lr2+ are expected to be 7s2 and 7s1 respectively. However, in species where all three valence electrons of lawrencium are ionized to give at least formally the Lr3+ cation, lawrencium is expected to behave like a typical actinide and the heavier congener of lutetium, especially because the first three ionization potentials of lawrencium are predicted to be similar to those of lutetium. Hence, unlike thallium but like lutetium, lawrencium would prefer to form LrH3 than LrH, and LrCO is expected to be similar to the also unknown LuCO, both metals having valence configuration σ2π1 in their monocarbonyls. The pπ–dπ bond is expected to be seen in LrCl3 just as it is for LuCl3 and more generally all the LnCl3. The complex anion [Lr(C5H4SiMe3)3] is expected to be stable with a configuration of 6d1 for lawrencium; this 6d orbital would be its highest occupied molecular orbital. This is analogous to the electronic structure of the analogous lutetium compound.[73]

Atomic edit

Lawrencium has three valence electrons: the 5f electrons are in the atomic core.[74] In 1970, it was predicted that the ground-state electron configuration of lawrencium was [Rn]5f146d17s2 (ground state term symbol 2D3/2), per the Aufbau principle and conforming to the [Xe]4f145d16s2 configuration of lawrencium's lighter homolog lutetium.[75] But the next year, calculations were published that questioned this prediction, instead expecting an anomalous [Rn]5f147s27p1 configuration.[75] Though early calculations gave conflicting results,[76] more recent studies and calculations confirm the s2p suggestion.[77][78] 1974 relativistic calculations concluded that the energy difference between the two configurations was small and that it was uncertain which was the ground state.[75] Later 1995 calculations concluded that the s2p configuration should be energetically favored, because the spherical s and p1/2 orbitals are nearest to the atomic nucleus and thus move quickly enough that their relativistic mass increases significantly.[75]

In 1988, a team of scientists led by Eichler calculated that lawrencium's enthalpy of adsorption on metal sources would differ enough depending on its electron configuration that it would be feasible to carry out experiments to exploit this fact to measure lawrencium's electron configuration.[75] The s2p configuration was expected to be more volatile than the s2d configuration, and be more similar to that of the p-block element lead. No evidence for lawrencium being volatile was obtained and the lower limit for the enthalpy of adsorption of lawrencium on quartz or platinum was significantly higher than the estimated value for the s2p configuration.[75]

 
First ionization energy (eV) plotted against atomic number, in units eV. Predicted values are used beyond rutherfordium (element 104). Lawrencium (element 103) has a very low first ionization energy, fitting the start of the d-block trend better than the end of the f-block trend before it.[79]

In 2015, the first ionization energy of lawrencium was measured, using the isotope 256Lr.[4] The measured value, 4.96+0.08
−0.07
eV
, agreed very well with the relativistic theoretical prediction of 4.963(15) eV, and also provided a first step into measuring the first ionization energies of the transactinides.[4] This value is the lowest among all the lanthanides and actinides, and supports the s2p configuration as the 7p1/2 electron is expected to be only weakly bound. As ionisation energies generally increase left to right in the f-block, this low value suggests that lutetium and lawrencium belong in the d-block (whose trend they follow) and not the f-block. That would make them the heavier congeners of scandium and yttrium, rather than lanthanum and actinium.[79] Although some alkali metal-like behaviour has been predicted,[80] adsorption experiments suggest that lawrencium is trivalent like scandium and yttrium, not monovalent like the alkali metals.[64] A lower limit on lawrencium's second ionization energy (>13.3 eV) was experimentally found in 2021.[81]

Even though s2p is now known to be the ground-state configuration of the lawrencium atom, ds2 should be a low-lying excited-state configuration, with an excitation energy variously calculated as 0.156 eV, 0.165 eV, or 0.626 eV.[73] As such lawrencium may still be considered to be a d-block element, albeit with an anomalous electron configuration (like chromium or copper), as its chemical behaviour matches expectations for a heavier analogue of lutetium.[65]

Isotopes edit

Fourteen isotopes of lawrencium are known, with mass number 251–262, 264, and 266; all are radioactive.[82][83][84] Seven nuclear isomers are known. The longest-lived isotope, 266Lr, has a half-life of about ten hours and is one of the longest-lived superheavy isotopes known to date.[85] However, shorter-lived isotopes are usually used in chemical experiments because 266Lr currently can only be produced as a final decay product of even heavier and harder-to-make elements: it was discovered in 2014 in the decay chain of 294Ts.[82][83] 256Lr (half-life 27 seconds) was used in the first chemical studies on lawrencium: currently, the longer-lived 260Lr (half-life 2.7 minutes) is usually used for this purpose.[82] After 266Lr, the longest-lived isotopes are 264Lr (4.8+2.2
−1.3
 h
), 262Lr (3.6 h), and 261Lr (44 min).[82][86][87] All other known lawrencium isotopes have half-lives under 5 minutes, and the shortest-lived of them (251Lr) has a half-life of 24.4 milliseconds.[84][86][87][88] The half-lives of lawrencium isotopes mostly increase smoothly from 251Lr to 266Lr, with a dip from 257Lr to 259Lr.[82][86][87]

Preparation and purification edit

Most isotopes of lawrencium can be produced by bombarding actinide (americium to einsteinium) targets with light ions (from boron to neon). The two most important isotopes, 256Lr and 260Lr, can be respectively produced by bombarding californium-249 with 70 MeV boron-11 ions (producing lawrencium-256 and four neutrons) and by bombarding berkelium-249 with oxygen-18 (producing lawrencium-260, an alpha particle, and three neutrons).[89] The two heaviest and longest-lived known isotopes, 264Lr and 266Lr, can only be produced at much lower yields as decay products of dubnium, whose progenitors are isotopes of moscovium and tennessine.

Both 256Lr and 260Lr have half-lives too short to allow a complete chemical purification process. Early experiments with 256Lr therefore used rapid solvent extraction, with the chelating agent thenoyltrifluoroacetone (TTA) dissolved in methyl isobutyl ketone (MIBK) as the organic phase, and with the aqueous phase being buffered acetate solutions. Ions of different charge (+2, +3, or +4) will then extract into the organic phase under different pH ranges, but this method will not separate the trivalent actinides and thus 256Lr must be identified by its emitted 8.24 MeV alpha particles.[89] More recent methods have allowed rapid selective elution with α-HIB to take place in enough time to separate out the longer-lived isotope 260Lr, which can be removed from the catcher foil with 0.05 M hydrochloric acid.[89]

See also edit

Notes edit

  1. ^ In nuclear physics, an element is called heavy if its atomic number is high; lead (element 82) is one example of such a heavy element. The term "superheavy elements" typically refers to elements with atomic number greater than 103 (although there are other definitions, such as atomic number greater than 100[8] or 112;[9] sometimes, the term is presented an equivalent to the term "transactinide", which puts an upper limit before the beginning of the hypothetical superactinide series).[10] Terms "heavy isotopes" (of a given element) and "heavy nuclei" mean what could be understood in the common language—isotopes of high mass (for the given element) and nuclei of high mass, respectively.
  2. ^ In 2009, a team at the JINR led by Oganessian published results of their attempt to create hassium in a symmetric 136Xe + 136Xe reaction. They failed to observe a single atom in such a reaction, putting the upper limit on the cross section, the measure of probability of a nuclear reaction, as 2.5 pb.[11] In comparison, the reaction that resulted in hassium discovery, 208Pb + 58Fe, had a cross section of ~20 pb (more specifically, 19+19
    -11
     pb), as estimated by the discoverers.[12]
  3. ^ The amount of energy applied to the beam particle to accelerate it can also influence the value of cross section. For example, in the 28
    14
    Si
    + 1
    0
    n
    28
    13
    Al
    + 1
    1
    p
    reaction, cross section changes smoothly from 370 mb at 12.3 MeV to 160 mb at 18.3 MeV, with a broad peak at 13.5 MeV with the maximum value of 380 mb.[16]
  4. ^ This figure also marks the generally accepted upper limit for lifetime of a compound nucleus.[21]
  5. ^ This separation is based on that the resulting nuclei move past the target more slowly then the unreacted beam nuclei. The separator contains electric and magnetic fields whose effects on a moving particle cancel out for a specific velocity of a particle.[23] Such separation can also be aided by a time-of-flight measurement and a recoil energy measurement; a combination of the two may allow to estimate the mass of a nucleus.[24]
  6. ^ Not all decay modes are caused by electrostatic repulsion. For example, beta decay is caused by the weak interaction.[31]
  7. ^ It was already known by the 1960s that ground states of nuclei differed in energy and shape as well as that certain magic numbers of nucleons corresponded to greater stability of a nucleus. However, it was assumed that there was no nuclear structure in superheavy nuclei as they were too deformed to form one.[36]
  8. ^ Since mass of a nucleus is not measured directly but is rather calculated from that of another nucleus, such measurement is called indirect. Direct measurements are also possible, but for the most part they have remained unavailable for superheavy nuclei.[41] The first direct measurement of mass of a superheavy nucleus was reported in 2018 at LBNL.[42] Mass was determined from the location of a nucleus after the transfer (the location helps determine its trajectory, which is linked to the mass-to-charge ratio of the nucleus, since the transfer was done in presence of a magnet).[43]
  9. ^ If the decay occurred in a vacuum, then since total momentum of an isolated system before and after the decay must be preserved, the daughter nucleus would also receive a small velocity. The ratio of the two velocities, and accordingly the ratio of the kinetic energies, would thus be inverse to the ratio of the two masses. The decay energy equals the sum of the known kinetic energy of the alpha particle and that of the daughter nucleus (an exact fraction of the former).[32] The calculations hold for an experiment as well, but the difference is that the nucleus does not move after the decay because it is tied to the detector.
  10. ^ Spontaneous fission was discovered by Soviet physicist Georgy Flerov,[44] a leading scientist at JINR, and thus it was a "hobbyhorse" for the facility.[45] In contrast, the LBL scientists believed fission information was not sufficient for a claim of synthesis of an element. They believed spontaneous fission had not been studied enough to use it for identification of a new element, since there was a difficulty of establishing that a compound nucleus had only ejected neutrons and not charged particles like protons or alpha particles.[21] They thus preferred to link new isotopes to the already known ones by successive alpha decays.[44]
  11. ^ For instance, element 102 was mistakenly identified in 1957 at the Nobel Institute of Physics in Stockholm, Stockholm County, Sweden.[46] There were no earlier definitive claims of creation of this element, and the element was assigned a name by its Swedish, American, and British discoverers, nobelium. It was later shown that the identification was incorrect.[47] The following year, RL was unable to reproduce the Swedish results and announced instead their synthesis of the element; that claim was also disproved later.[47] JINR insisted that they were the first to create the element and suggested a name of their own for the new element, joliotium;[48] the Soviet name was also not accepted (JINR later referred to the naming of the element 102 as "hasty").[49] This name was proposed to IUPAC in a written response to their ruling on priority of discovery claims of elements, signed 29 September 1992.[49] The name "nobelium" remained unchanged on account of its widespread usage.[50]

References edit

  1. ^ Emsley, John (2011). Nature's Building Blocks: An A-Z Guide to the Elements (New ed.). New York, NY: Oxford University Press. p. 278–279. ISBN 978-0-19-960563-7.
  2. ^ a b Gyanchandani, Jyoti; Sikka, S. K. (10 May 2011). "Physical properties of the 6 d -series elements from density functional theory: Close similarity to lighter transition metals". Physical Review B. 83 (17): 172101. Bibcode:2011PhRvB..83q2101G. doi:10.1103/PhysRevB.83.172101.
  3. ^ Brown, Geoffrey (2012). The Inaccessible Earth: An integrated view to its structure and composition. Springer Science & Business Media. p. 88. ISBN 9789401115162.
  4. ^ a b c Sato, T. K.; Asai, M.; Borschevsky, A.; Stora, T.; Sato, N.; Kaneya, Y.; Tsukada, K.; Düllman, Ch. E.; Eberhardt, K.; Eliav, E.; Ichikawa, S.; Kaldor, U.; Kratz, J. V.; Miyashita, S.; Nagame, Y.; Ooe, K.; Osa, A.; Renisch, D.; Runke, J.; Schädel, M.; Thörle-Pospiech, P.; Toyoshima, A.; Trautmann, N. (9 April 2015). "Measurement of the first ionization potential of lawrencium, element 103" (PDF). Nature. 520 (7546): 209–11. Bibcode:2015Natur.520..209S. doi:10.1038/nature14342. PMID 25855457. S2CID 4384213.
  5. ^ a b Östlin, A.; Vitos, L. (2011). "First-principles calculation of the structural stability of 6d transition metals". Physical Review B. 84 (11): 113104. Bibcode:2011PhRvB..84k3104O. doi:10.1103/PhysRevB.84.113104.
  6. ^ Kondev, F. G.; Wang, M.; Huang, W. J.; Naimi, S.; Audi, G. (2021). "The NUBASE2020 evaluation of nuclear properties" (PDF). Chinese Physics C. 45 (3): 030001. doi:10.1088/1674-1137/abddae.
  7. ^ Oganessian, Yu. Ts.; Utyonkov, V. K.; Kovrizhnykh, N. D.; et al. (2022). "New isotope 286Mc produced in the 243Am+48Ca reaction". Physical Review C. 106 (064306). doi:10.1103/PhysRevC.106.064306.
  8. ^ Krämer, K. (2016). "Explainer: superheavy elements". Chemistry World. Retrieved 2020-03-15.
  9. ^ . Lawrence Livermore National Laboratory. Archived from the original on 2015-09-11. Retrieved 2020-03-15.
  10. ^ Eliav, E.; Kaldor, U.; Borschevsky, A. (2018). "Electronic Structure of the Transactinide Atoms". In Scott, R. A. (ed.). Encyclopedia of Inorganic and Bioinorganic Chemistry. John Wiley & Sons. pp. 1–16. doi:10.1002/9781119951438.eibc2632. ISBN 978-1-119-95143-8. S2CID 127060181.
  11. ^ Oganessian, Yu. Ts.; Dmitriev, S. N.; Yeremin, A. V.; et al. (2009). "Attempt to produce the isotopes of element 108 in the fusion reaction 136Xe + 136Xe". Physical Review C. 79 (2): 024608. doi:10.1103/PhysRevC.79.024608. ISSN 0556-2813.
  12. ^ Münzenberg, G.; Armbruster, P.; Folger, H.; et al. (1984). (PDF). Zeitschrift für Physik A. 317 (2): 235–236. Bibcode:1984ZPhyA.317..235M. doi:10.1007/BF01421260. S2CID 123288075. Archived from the original (PDF) on 7 June 2015. Retrieved 20 October 2012.
  13. ^ Subramanian, S. (28 August 2019). "Making New Elements Doesn't Pay. Just Ask This Berkeley Scientist". Bloomberg Businessweek. Retrieved 2020-01-18.
  14. ^ a b c d e f Ivanov, D. (2019). "Сверхтяжелые шаги в неизвестное" [Superheavy steps into the unknown]. nplus1.ru (in Russian). Retrieved 2020-02-02.
  15. ^ Hinde, D. (2017). "Something new and superheavy at the periodic table". The Conversation. Retrieved 2020-01-30.
  16. ^ Kern, B. D.; Thompson, W. E.; Ferguson, J. M. (1959). "Cross sections for some (n, p) and (n, α) reactions". Nuclear Physics. 10: 226–234. Bibcode:1959NucPh..10..226K. doi:10.1016/0029-5582(59)90211-1.
  17. ^ Wakhle, A.; Simenel, C.; Hinde, D. J.; et al. (2015). Simenel, C.; Gomes, P. R. S.; Hinde, D. J.; et al. (eds.). "Comparing Experimental and Theoretical Quasifission Mass Angle Distributions". European Physical Journal Web of Conferences. 86: 00061. Bibcode:2015EPJWC..8600061W. doi:10.1051/epjconf/20158600061. hdl:1885/148847. ISSN 2100-014X.
  18. ^ "Nuclear Reactions" (PDF). pp. 7–8. Retrieved 2020-01-27. Published as Loveland, W. D.; Morrissey, D. J.; Seaborg, G. T. (2005). "Nuclear Reactions". Modern Nuclear Chemistry. John Wiley & Sons, Inc. pp. 249–297. doi:10.1002/0471768626.ch10. ISBN 978-0-471-76862-3.
  19. ^ a b Krása, A. (2010). "Neutron Sources for ADS". Faculty of Nuclear Sciences and Physical Engineering. Czech Technical University in Prague: 4–8. S2CID 28796927.
  20. ^ Wapstra, A. H. (1991). "Criteria that must be satisfied for the discovery of a new chemical element to be recognized" (PDF). Pure and Applied Chemistry. 63 (6): 883. doi:10.1351/pac199163060879. ISSN 1365-3075. S2CID 95737691.
  21. ^ a b Hyde, E. K.; Hoffman, D. C.; Keller, O. L. (1987). "A History and Analysis of the Discovery of Elements 104 and 105". Radiochimica Acta. 42 (2): 67–68. doi:10.1524/ract.1987.42.2.57. ISSN 2193-3405. S2CID 99193729.
  22. ^ a b c d Chemistry World (2016). "How to Make Superheavy Elements and Finish the Periodic Table [Video]". Scientific American. Retrieved 2020-01-27.
  23. ^ Hoffman, Ghiorso & Seaborg 2000, p. 334.
  24. ^ Hoffman, Ghiorso & Seaborg 2000, p. 335.
  25. ^ Zagrebaev, Karpov & Greiner 2013, p. 3.
  26. ^ Beiser 2003, p. 432.
  27. ^ a b Pauli, N. (2019). "Alpha decay" (PDF). Introductory Nuclear, Atomic and Molecular Physics (Nuclear Physics Part). Université libre de Bruxelles. Retrieved 2020-02-16.
  28. ^ a b c d e Pauli, N. (2019). "Nuclear fission" (PDF). Introductory Nuclear, Atomic and Molecular Physics (Nuclear Physics Part). Université libre de Bruxelles. Retrieved 2020-02-16.
  29. ^ Staszczak, A.; Baran, A.; Nazarewicz, W. (2013). "Spontaneous fission modes and lifetimes of superheavy elements in the nuclear density functional theory". Physical Review C. 87 (2): 024320–1. arXiv:1208.1215. Bibcode:2013PhRvC..87b4320S. doi:10.1103/physrevc.87.024320. ISSN 0556-2813.
  30. ^ Audi et al. 2017, pp. 030001-129–030001-138.
  31. ^ Beiser 2003, p. 439.
  32. ^ a b Beiser 2003, p. 433.
  33. ^ Audi et al. 2017, p. 030001-125.
  34. ^ Aksenov, N. V.; Steinegger, P.; Abdullin, F. Sh.; et al. (2017). "On the volatility of nihonium (Nh, Z = 113)". The European Physical Journal A. 53 (7): 158. Bibcode:2017EPJA...53..158A. doi:10.1140/epja/i2017-12348-8. ISSN 1434-6001. S2CID 125849923.
  35. ^ Beiser 2003, p. 432–433.
  36. ^ a b c Oganessian, Yu. (2012). "Nuclei in the "Island of Stability" of Superheavy Elements". Journal of Physics: Conference Series. 337 (1): 012005-1–012005-6. Bibcode:2012JPhCS.337a2005O. doi:10.1088/1742-6596/337/1/012005. ISSN 1742-6596.
  37. ^ Moller, P.; Nix, J. R. (1994). Fission properties of the heaviest elements (PDF). Dai 2 Kai Hadoron Tataikei no Simulation Symposium, Tokai-mura, Ibaraki, Japan. University of North Texas. Retrieved 2020-02-16.
  38. ^ a b Oganessian, Yu. Ts. (2004). "Superheavy elements". Physics World. 17 (7): 25–29. doi:10.1088/2058-7058/17/7/31. Retrieved 2020-02-16.
  39. ^ Schädel, M. (2015). "Chemistry of the superheavy elements". Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 373 (2037): 20140191. Bibcode:2015RSPTA.37340191S. doi:10.1098/rsta.2014.0191. ISSN 1364-503X. PMID 25666065.
  40. ^ Hulet, E. K. (1989). Biomodal spontaneous fission. 50th Anniversary of Nuclear Fission, Leningrad, USSR. Bibcode:1989nufi.rept...16H.
  41. ^ Oganessian, Yu. Ts.; Rykaczewski, K. P. (2015). "A beachhead on the island of stability". Physics Today. 68 (8): 32–38. Bibcode:2015PhT....68h..32O. doi:10.1063/PT.3.2880. ISSN 0031-9228. OSTI 1337838. S2CID 119531411.
  42. ^ Grant, A. (2018). "Weighing the heaviest elements". Physics Today. doi:10.1063/PT.6.1.20181113a. S2CID 239775403.
  43. ^ Howes, L. (2019). "Exploring the superheavy elements at the end of the periodic table". Chemical & Engineering News. Retrieved 2020-01-27.
  44. ^ a b Robinson, A. E. (2019). "The Transfermium Wars: Scientific Brawling and Name-Calling during the Cold War". Distillations. Retrieved 2020-02-22.
  45. ^ "Популярная библиотека химических элементов. Сиборгий (экавольфрам)" [Popular library of chemical elements. Seaborgium (eka-tungsten)]. n-t.ru (in Russian). Retrieved 2020-01-07. Reprinted from "Экавольфрам" [Eka-tungsten]. Популярная библиотека химических элементов. Серебро – Нильсборий и далее [Popular library of chemical elements. Silver through nielsbohrium and beyond] (in Russian). Nauka. 1977.
  46. ^ "Nobelium - Element information, properties and uses | Periodic Table". Royal Society of Chemistry. Retrieved 2020-03-01.
  47. ^ a b Kragh 2018, pp. 38–39.
  48. ^ Kragh 2018, p. 40.
  49. ^ a b Ghiorso, A.; Seaborg, G. T.; Oganessian, Yu. Ts.; et al. (1993). "Responses on the report 'Discovery of the Transfermium elements' followed by reply to the responses by Transfermium Working Group" (PDF). Pure and Applied Chemistry. 65 (8): 1815–1824. doi:10.1351/pac199365081815. S2CID 95069384. (PDF) from the original on 25 November 2013. Retrieved 7 September 2016.
  50. ^ Commission on Nomenclature of Inorganic Chemistry (1997). "Names and symbols of transfermium elements (IUPAC Recommendations 1997)" (PDF). Pure and Applied Chemistry. 69 (12): 2471–2474. doi:10.1351/pac199769122471.
  51. ^ a b c d e Emsley, John (2011). Nature's Building Blocks.
  52. ^ a b c d e f g h i j k Barber, R. C.; Greenwood, N. N.; Hrynkiewicz, A. Z.; Jeannin, Y. P.; Lefort, M.; Sakai, M.; Ulehla, I.; Wapstra, A. P.; Wilkinson, D. H. (1993). "Discovery of the transfermium elements. Part II: Introduction to discovery profiles. Part III: Discovery profiles of the transfermium elements". Pure and Applied Chemistry. 65 (8): 1757. doi:10.1351/pac199365081757. S2CID 195819585. (Note: for Part I see Pure Appl. Chem., Vol. 63, No. 6, pp. 879–886, 1991)
  53. ^ "This Month in Lab History…Lawrencium Added to Periodic Table". today.lbl.gov. Lawrence Berkeley National Laboratory. 9 April 2013. Retrieved 13 February 2021. Lawrencium (Lw) was first synthesized Feb. 14, 1961, by a team led by Ghiorso, who was co-discoverer of a record 12 chemical elements on the periodic table.
  54. ^ a b Ghiorso, Albert; Sikkeland, T.; Larsh, A. E.; Latimer, R. M. (1961). "New Element, Lawrencium, Atomic Number 103". Phys. Rev. Lett. 6 (9): 473. Bibcode:1961PhRvL...6..473G. doi:10.1103/PhysRevLett.6.473.
  55. ^ a b c Greenwood, Norman N. (1997). "Recent developments concerning the discovery of elements 101–111" (PDF). Pure Appl. Chem. 69 (1): 179–184. doi:10.1351/pac199769010179. S2CID 98322292.
  56. ^ Flerov, G. N. (1967). "On the nuclear properties of the isotopes 256103 and 257103". Nucl. Phys. A. 106 (2): 476. Bibcode:1967NuPhA.106..476F. doi:10.1016/0375-9474(67)90892-5.
  57. ^ Donets, E. D.; Shchegolev, V. A.; Ermakov, V. A. (1965). Atomnaya Énergiya (in Russian). 19 (2): 109. {{cite journal}}: Missing or empty |title= (help)
    Translated in Donets, E. D.; Shchegolev, V. A.; Ermakov, V. A. (1965). "Synthesis of the isotope of element 103 (lawrencium) with mass number 256". Soviet Atomic Energy. 19 (2): 109. doi:10.1007/BF01126414. S2CID 97218361.
  58. ^ a b Karpenko, V. (1980). "The Discovery of Supposed New Elements: Two Centuries of Errors". Ambix. 27 (2): 77–102. doi:10.1179/amb.1980.27.2.77.
  59. ^ Kaldor, Uzi & Wilson, Stephen (2005). Theoretical chemistry and physics of heavy and superheavy element. Springer. p. 57. ISBN 1-4020-1371-X.
  60. ^ Silva 2011, pp. 1641–2
  61. ^ Eskola, Kari; Eskola, Pirkko; Nurmia, Matti; Albert Ghiorso (1971). "Studies of Lawrencium Isotopes with Mass Numbers 255 Through 260". Phys. Rev. C. 4 (2): 632–642. Bibcode:1971PhRvC...4..632E. doi:10.1103/PhysRevC.4.632.
  62. ^ a b Silva 2011, p. 1644
  63. ^ Silva 2011, p. 1639
  64. ^ a b Haire, R. G. (11 October 2007). "Insights into the bonding and electronic nature of heavy element materials". Journal of Alloys and Compounds. 444–5: 63–71. doi:10.1016/j.jallcom.2007.01.103.
  65. ^ a b c Jensen, William B. (2015). "The positions of lanthanum (actinium) and lutetium (lawrencium) in the periodic table: an update". Foundations of Chemistry. 17: 23–31. doi:10.1007/s10698-015-9216-1. S2CID 98624395. from the original on 30 January 2021. Retrieved 28 January 2021.
  66. ^ Winter, Mark (1993–2022). "WebElements". The University of Sheffield and WebElements Ltd, UK. Retrieved 5 December 2022.
  67. ^ Cowan, Robert D. (1981). The Theory of Atomic Structure and Spectra. University of California Press. p. 598. ISBN 9780520906150.
  68. ^ John Emsley (2011). Nature's Building Blocks: An A-Z Guide to the Elements. Oxford University Press. pp. 278–9. ISBN 978-0-19-960563-7.
  69. ^ Lide, D. R., ed. (2003). CRC Handbook of Chemistry and Physics (84th ed.). Boca Raton, FL: CRC Press.
  70. ^ a b c d e f g h Silva 2011, pp. 1644–7
  71. ^ Hoffman, Darleane C.; Lee, Diana M.; Pershina, Valeria (2006). "Transactinides and the future elements". In Morss; Edelstein, Norman M.; Fuger, Jean (eds.). The Chemistry of the Actinide and Transactinide Elements (3rd ed.). Dordrecht, The Netherlands: Springer Science+Business Media. p. 1686. ISBN 1-4020-3555-1.
  72. ^ Balasubramanian, K. (4 December 2001). "Potential energy surfaces of Lawrencium and Nobelium dihydrides (LrH2 and NoH2)". Journal of Chemical Physics. 116 (9): 3568–75. Bibcode:2002JChPh.116.3568B. doi:10.1063/1.1446029.
  73. ^ a b Xu, Wen-Hua; Pyykkö, Pekka (8 June 2016). "Is the chemistry of lawrencium peculiar". Phys. Chem. Chem. Phys. 2016 (18): 17351–5. Bibcode:2016PCCP...1817351X. doi:10.1039/c6cp02706g. hdl:10138/224395. PMID 27314425. S2CID 31224634. Retrieved 24 April 2017.
  74. ^ Jensen, William B. (2000). (PDF). Archived from the original (PDF) on 2020-11-10. Retrieved 10 December 2022.
  75. ^ a b c d e f Silva 2011, pp. 1643–4
  76. ^ Nugent, L. J.; Vander Sluis, K. L.; Fricke, Burhard; Mann, J. B. (1974). "Electronic configuration in the ground state of atomic lawrencium" (PDF). Phys. Rev. A. 9 (6): 2270–72. Bibcode:1974PhRvA...9.2270N. doi:10.1103/PhysRevA.9.2270.
  77. ^ Eliav, E.; Kaldor, U.; Ishikawa, Y. (1995). "Transition energies of ytterbium, lutetium, and lawrencium by the relativistic coupled-cluster method". Phys. Rev. A. 52 (1): 291–296. Bibcode:1995PhRvA..52..291E. doi:10.1103/PhysRevA.52.291. PMID 9912247.
  78. ^ Zou, Yu; Froese Fischer C.; Uiterwaal, C.; Wanner, J.; Kompa, K.-L. (2002). "Resonance Transition Energies and Oscillator Strengths in Lutetium and Lawrencium". Phys. Rev. Lett. 88 (2): 183001. Bibcode:2001PhRvL..88b3001M. doi:10.1103/PhysRevLett.88.023001. PMID 12005680. S2CID 18391594.
  79. ^ a b Jensen, W. B. (2015). (PDF). Archived from the original (PDF) on 23 December 2015. Retrieved 20 September 2015.
  80. ^ Gunther, Matthew (9 April 2015). "Lawrencium experiment could shake up periodic table". RSC Chemistry World. Retrieved 21 September 2015.
  81. ^ Kwarsick, Jeffrey T.; Pore, Jennifer L.; Gates, Jacklyn M.; Gregorich, Kenneth E.; Gibson, John K.; Jian, Jiwen; Pang, Gregory K.; Shuh, David K. (2021). "Assessment of the Second-Ionization Potential of Lawrencium: Investigating the End of the Actinide Series with a One-Atom-at-a-Time Gas-Phase Ion Chemistry Technique". The Journal of Physical Chemistry A. 125 (31): 6818–6828. Bibcode:2021JPCA..125.6818K. doi:10.1021/acs.jpca.1c01961. OSTI 1844939. PMID 34242037. S2CID 235785891.
  82. ^ a b c d e Silva 2011, p. 1642
  83. ^ a b Khuyagbaatar, J.; et al. (2014). "48Ca + 249Bk Fusion Reaction Leading to Element Z = 117: Long-Lived α-Decaying 270Db and Discovery of 266Lr" (PDF). Physical Review Letters. 112 (17): 172501. Bibcode:2014PhRvL.112q2501K. doi:10.1103/PhysRevLett.112.172501. hdl:1885/70327. PMID 24836239. S2CID 5949620.
  84. ^ a b Leppänen, A.-P. (2005). Alpha-decay and decay-tagging studies of heavy elements using the RITU separator (PDF) (Thesis). University of Jyväskylä. pp. 83–100. ISBN 978-951-39-3162-9. ISSN 0075-465X.
  85. ^ Clara Moskowitz (May 7, 2014). "Superheavy Element 117 Points to Fabled "Island of Stability" on Periodic Table". Scientific American. Retrieved 2014-05-08.
  86. ^ a b c "Nucleonica :: Web driven nuclear science".
  87. ^ a b c Audi, G.; Kondev, F. G.; Wang, M.; Huang, W. J.; Naimi, S. (2017). "The NUBASE2016 evaluation of nuclear properties" (PDF). Chinese Physics C. 41 (3): 030001. Bibcode:2017ChPhC..41c0001A. doi:10.1088/1674-1137/41/3/030001.
  88. ^ Huang, T.; Seweryniak, D.; Back, B. B.; et al. (2022). "Discovery of the new isotope 251Lr: Impact of the hexacontetrapole deformation on single-proton orbital energies near the Z = 100 deformed shell gap". Physical Review C. 106 (L061301). doi:10.1103/PhysRevC.106.L061301. S2CID 254300224.
  89. ^ a b c Silva 2011, pp. 1642–3

Bibliography edit

  • Audi, G.; Kondev, F. G.; Wang, M.; et al. (2017). "The NUBASE2016 evaluation of nuclear properties". Chinese Physics C. 41 (3): 030001. Bibcode:2017ChPhC..41c0001A. doi:10.1088/1674-1137/41/3/030001.
  • Beiser, A. (2003). Concepts of modern physics (6th ed.). McGraw-Hill. ISBN 978-0-07-244848-1. OCLC 48965418.
  • Hoffman, D. C.; Ghiorso, A.; Seaborg, G. T. (2000). The Transuranium People: The Inside Story. World Scientific. ISBN 978-1-78-326244-1.
  • Kragh, H. (2018). From Transuranic to Superheavy Elements: A Story of Dispute and Creation. Springer. ISBN 978-3-319-75813-8.
  • Silva, Robert J. (2011). "Chapter 13. Fermium, Mendelevium, Nobelium, and Lawrencium". In Morss, Lester R.; Edelstein, Norman M.; Fuger, Jean (eds.). The Chemistry of the Actinide and Transactinide Elements. Netherlands: Springer. doi:10.1007/978-94-007-0211-0_13. ISBN 978-94-007-0210-3.
  • Zagrebaev, V.; Karpov, A.; Greiner, W. (2013). "Future of superheavy element research: Which nuclei could be synthesized within the next few years?". Journal of Physics: Conference Series. 420 (1): 012001. arXiv:1207.5700. Bibcode:2013JPhCS.420a2001Z. doi:10.1088/1742-6596/420/1/012001. ISSN 1742-6588. S2CID 55434734.

External links edit

  • . National Nuclear Data Center (NNDC). Archived from the original on 2018-10-10. Retrieved 2014-08-21.
  • Los Alamos National Laboratory's Chemistry Division: Periodic Table – Lawrencium
  • Lawrencium at The Periodic Table of Videos (University of Nottingham)

lawrencium, synthetic, chemical, element, symbol, formerly, atomic, number, named, honor, ernest, lawrence, inventor, cyclotron, device, that, used, discover, many, artificial, radioactive, elements, radioactive, metal, lawrencium, eleventh, transuranic, eleme. Lawrencium is a synthetic chemical element it has symbol Lr formerly Lw and atomic number 103 It is named in honor of Ernest Lawrence inventor of the cyclotron a device that was used to discover many artificial radioactive elements A radioactive metal lawrencium is the eleventh transuranic element and the last member of the actinide series Like all elements with atomic number over 100 lawrencium can only be produced in particle accelerators by bombarding lighter elements with charged particles Fourteen isotopes of lawrencium are currently known the most stable is 266Lr with half life 11 hours but the shorter lived 260Lr half life 2 7 minutes is most commonly used in chemistry because it can be produced on a larger scale Lawrencium 103LrLawrenciumPronunciation l ɒ ˈ r ɛ n s i e m wbr lo REN see em Appearancesilvery predicted 1 Mass number 266 Lawrencium in the periodic tableHydrogen HeliumLithium Beryllium Boron Carbon Nitrogen Oxygen Fluorine NeonSodium Magnesium Aluminium Silicon Phosphorus Sulfur Chlorine ArgonPotassium Calcium Scandium Titanium Vanadium Chromium Manganese Iron Cobalt Nickel Copper Zinc Gallium Germanium Arsenic Selenium Bromine KryptonRubidium Strontium Yttrium Zirconium Niobium Molybdenum Technetium Ruthenium Rhodium Palladium Silver Cadmium Indium Tin Antimony Tellurium Iodine XenonCaesium Barium Lanthanum Cerium Praseodymium Neodymium Promethium Samarium Europium Gadolinium Terbium Dysprosium Holmium Erbium Thulium Ytterbium Lutetium Hafnium Tantalum Tungsten Rhenium Osmium Iridium Platinum Gold Mercury element Thallium Lead Bismuth Polonium Astatine RadonFrancium Radium Actinium Thorium Protactinium Uranium Neptunium Plutonium Americium Curium Berkelium Californium Einsteinium Fermium Mendelevium Nobelium Lawrencium Rutherfordium Dubnium Seaborgium Bohrium Hassium Meitnerium Darmstadtium Roentgenium Copernicium Nihonium Flerovium Moscovium Livermorium Tennessine Oganesson Lu Lr Ups nobelium lawrencium rutherfordiumAtomic number Z 103Groupgroup 3Periodperiod 7Block d blockElectron configuration Rn 5f14 7s2 7p1Electrons per shell2 8 18 32 32 8 3Physical propertiesPhase at STPsolid predicted Melting point1900 K 1600 C 3000 F predicted Density near r t 14 4 g cm3 predicted 2 Atomic propertiesOxidation states 3ElectronegativityPauling scale 1 3 predicted 3 Ionization energies1st 479 kJ mol 4 2nd 1428 0 kJ mol predicted 3rd 2219 1 kJ mol predicted Other propertiesNatural occurrencesyntheticCrystal structure hexagonal close packed hcp predicted 5 CAS Number22537 19 5HistoryNamingafter Ernest LawrenceDiscoveryLawrence Berkeley National Laboratory and Joint Institute for Nuclear Research 1961 1971 Isotopes of lawrenciumveMain isotopes 6 Decayabun dance half life t1 2 mode pro duct256Lr synth 27 9 s a 252Mdb 256No260Lr synth 3 0 min a 256Mdb 260No261Lr synth 39 min SF 262Lr synth 4 h b 262No264Lr synth 4 8 h 7 SF 266Lr synth 11 h SF Category Lawrenciumviewtalkedit referencesChemistry experiments confirm that lawrencium behaves as a heavier homolog to lutetium in the periodic table and is a trivalent element It thus could also be classified as the first of the 7th period transition metals Its electron configuration is anomalous for its position in the periodic table having an s2p configuration instead of the s2d configuration of its homolog lutetium However this does not appear to affect lawrencium s chemistry In the 1950s 1960s and 1970s many claims of the synthesis of lawrencium of varying quality were made from laboratories in the Soviet Union and the United States The priority of the discovery and therefore the name of the element was disputed between Soviet and American scientists The International Union of Pure and Applied Chemistry IUPAC initially established lawrencium as the official name for the element and gave the American team credit for the discovery this was reevaluated in 1997 giving both teams shared credit for the discovery but not changing the element s name Contents 1 Introduction 1 1 Synthesis of superheavy nuclei 1 2 Decay and detection 2 History 3 Characteristics 3 1 Physical 3 2 Chemical 3 3 Atomic 3 4 Isotopes 4 Preparation and purification 5 See also 6 Notes 7 References 8 Bibliography 9 External linksIntroduction editThis section is an excerpt from Superheavy element Introduction edit Synthesis of superheavy nuclei edit nbsp A graphic depiction of a nuclear fusion reaction Two nuclei fuse into one emitting a neutron Reactions that created new elements to this moment were similar with the only possible difference that several singular neutrons sometimes were released or none at all A superheavy a atomic nucleus is created in a nuclear reaction that combines two other nuclei of unequal size b into one roughly the more unequal the two nuclei in terms of mass the greater the possibility that the two react 13 The material made of the heavier nuclei is made into a target which is then bombarded by the beam of lighter nuclei Two nuclei can only fuse into one if they approach each other closely enough normally nuclei all positively charged repel each other due to electrostatic repulsion The strong interaction can overcome this repulsion but only within a very short distance from a nucleus beam nuclei are thus greatly accelerated in order to make such repulsion insignificant compared to the velocity of the beam nucleus 14 The energy applied to the beam nuclei to accelerate them can cause them to reach speeds as high as one tenth of the speed of light However if too much energy is applied the beam nucleus can fall apart 14 Coming close enough alone is not enough for two nuclei to fuse when two nuclei approach each other they usually remain together for approximately 10 20 seconds and then part ways not necessarily in the same composition as before the reaction rather than form a single nucleus 14 15 This happens because during the attempted formation of a single nucleus electrostatic repulsion tears apart the nucleus that is being formed 14 Each pair of a target and a beam is characterized by its cross section the probability that fusion will occur if two nuclei approach one another expressed in terms of the transverse area that the incident particle must hit in order for the fusion to occur c This fusion may occur as a result of the quantum effect in which nuclei can tunnel through electrostatic repulsion If the two nuclei can stay close for past that phase multiple nuclear interactions result in redistribution of energy and an energy equilibrium 14 External videos nbsp Visualization of unsuccessful nuclear fusion based on calculations from the Australian National University 17 The resulting merger is an excited state 18 termed a compound nucleus and thus it is very unstable 14 To reach a more stable state the temporary merger may fission without formation of a more stable nucleus 19 Alternatively the compound nucleus may eject a few neutrons which would carry away the excitation energy if the latter is not sufficient for a neutron expulsion the merger would produce a gamma ray This happens in approximately 10 16 seconds after the initial nuclear collision and results in creation of a more stable nucleus 19 The definition by the IUPAC IUPAP Joint Working Party JWP states that a chemical element can only be recognized as discovered if a nucleus of it has not decayed within 10 14 seconds This value was chosen as an estimate of how long it takes a nucleus to acquire its outer electrons and thus display its chemical properties 20 d Decay and detection edit The beam passes through the target and reaches the next chamber the separator if a new nucleus is produced it is carried with this beam 22 In the separator the newly produced nucleus is separated from other nuclides that of the original beam and any other reaction products e and transferred to a surface barrier detector which stops the nucleus The exact location of the upcoming impact on the detector is marked also marked are its energy and the time of the arrival 22 The transfer takes about 10 6 seconds in order to be detected the nucleus must survive this long 25 The nucleus is recorded again once its decay is registered and the location the energy and the time of the decay are measured 22 Stability of a nucleus is provided by the strong interaction However its range is very short as nuclei become larger its influence on the outermost nucleons protons and neutrons weakens At the same time the nucleus is torn apart by electrostatic repulsion between protons and its range is not limited 26 Total binding energy provided by the strong interaction increases linearly with the number of nucleons whereas electrostatic repulsion increases with the square of the atomic number i e the latter grows faster and becomes increasingly important for heavy and superheavy nuclei 27 28 Superheavy nuclei are thus theoretically predicted 29 and have so far been observed 30 to predominantly decay via decay modes that are caused by such repulsion alpha decay and spontaneous fission f Almost all alpha emitters have over 210 nucleons 32 and the lightest nuclide primarily undergoing spontaneous fission has 238 33 In both decay modes nuclei are inhibited from decaying by corresponding energy barriers for each mode but they can be tunnelled through 27 28 nbsp Scheme of an apparatus for creation of superheavy elements based on the Dubna Gas Filled Recoil Separator set up in the Flerov Laboratory of Nuclear Reactions in JINR The trajectory within the detector and the beam focusing apparatus changes because of a dipole magnet in the former and quadrupole magnets in the latter 34 Alpha particles are commonly produced in radioactive decays because mass of an alpha particle per nucleon is small enough to leave some energy for the alpha particle to be used as kinetic energy to leave the nucleus 35 Spontaneous fission is caused by electrostatic repulsion tearing the nucleus apart and produces various nuclei in different instances of identical nuclei fissioning 28 As the atomic number increases spontaneous fission rapidly becomes more important spontaneous fission partial half lives decrease by 23 orders of magnitude from uranium element 92 to nobelium element 102 36 and by 30 orders of magnitude from thorium element 90 to fermium element 100 37 The earlier liquid drop model thus suggested that spontaneous fission would occur nearly instantly due to disappearance of the fission barrier for nuclei with about 280 nucleons 28 38 The later nuclear shell model suggested that nuclei with about 300 nucleons would form an island of stability in which nuclei will be more resistant to spontaneous fission and will primarily undergo alpha decay with longer half lives 28 38 Subsequent discoveries suggested that the predicted island might be further than originally anticipated they also showed that nuclei intermediate between the long lived actinides and the predicted island are deformed and gain additional stability from shell effects 39 Experiments on lighter superheavy nuclei 40 as well as those closer to the expected island 36 have shown greater than previously anticipated stability against spontaneous fission showing the importance of shell effects on nuclei g Alpha decays are registered by the emitted alpha particles and the decay products are easy to determine before the actual decay if such a decay or a series of consecutive decays produces a known nucleus the original product of a reaction can be easily determined h That all decays within a decay chain were indeed related to each other is established by the location of these decays which must be in the same place 22 The known nucleus can be recognized by the specific characteristics of decay it undergoes such as decay energy or more specifically the kinetic energy of the emitted particle i Spontaneous fission however produces various nuclei as products so the original nuclide cannot be determined from its daughters j The information available to physicists aiming to synthesize a superheavy element is thus the information collected at the detectors location energy and time of arrival of a particle to the detector and those of its decay The physicists analyze this data and seek to conclude that it was indeed caused by a new element and could not have been caused by a different nuclide than the one claimed Often provided data is insufficient for a conclusion that a new element was definitely created and there is no other explanation for the observed effects errors in interpreting data have been made k History edit nbsp Albert Ghiorso updating the periodic table in April 1961 writing the symbol Lw in as element 103 Codiscoverers Latimer Sikkeland and Larsh left to right look on In 1958 scientists at Lawrence Berkeley National Laboratory claimed the discovery of element 102 now called nobelium At the same time they also tried to synthesize element 103 by bombarding the same curium target used with nitrogen 14 ions Eighteen tracks were noted with decay energy around 9 1 MeV and half life around 0 25 s the Berkeley team noted that while the cause could be the production of an isotope of element 103 other possibilities could not be ruled out While the data agrees reasonably with that later discovered for 257Lr alpha decay energy 8 87 MeV half life 0 6 s the evidence obtained in this experiment fell far short of the strength required to conclusively demonstrate synthesis of element 103 A follow up on this experiment was not done as the target was destroyed 51 52 Later in 1960 the Lawrence Berkeley Laboratory attempted to synthesize the element by bombarding 252Cf with 10B and 11B The results of this experiment were not conclusive 51 The first important work on element 103 was done at Berkeley by the nuclear physics team of Albert Ghiorso Torbjorn Sikkeland Almon Larsh Robert M Latimer and their co workers on February 14 1961 53 The first atoms of lawrencium were reportedly made by bombarding a three milligram target consisting of three isotopes of californium with boron 10 and boron 11 nuclei from the Heavy Ion Linear Accelerator HILAC 54 The Berkeley team reported that the isotope 257103 was detected in this manner and that it decayed by emitting an 8 6 MeV alpha particle with a half life of 8 2 s 52 This identification was later corrected to 258103 54 as later work proved that 257Lr did not have the properties detected but 258Lr did 52 This was considered at the time to be convincing proof of synthesis of element 103 while the mass assignment was less certain and proved to be mistaken it did not affect the arguments in favor of element 103 having been synthesized Scientists at Joint Institute for Nuclear Research in Dubna then in the Soviet Union raised several criticisms all but one were answered adequately The exception was that 252Cf was the most common isotope in the target and in the reactions with 10B 258Lr could only have been produced by emitting four neutrons and emitting three neutrons was expected to be much less likely than emitting four or five This would lead to a narrow yield curve not the broad one reported by the Berkeley team A possible explanation was that there was a low number of events attributed to element 103 52 This was an important intermediate step to the unquestioned discovery of element 103 although the evidence was not completely convincing 52 The Berkeley team proposed the name lawrencium with symbol Lw after Ernest Lawrence inventor of the cyclotron The IUPAC Commission on Nomenclature of Inorganic Chemistry accepted the name but changed the symbol to Lr 55 This acceptance of the discovery was later characterized as being hasty by the Dubna team 52 25298 Cf 115 B 263103 Lr 258103 Lr 5 10 nThe first work at Dubna on element 103 came in 1965 when they reported to have made 256103 in 1965 by bombarding 243Am with 18O identifying it indirectly from its granddaughter fermium 252 The half life they reported was somewhat too high possibly due to background events Later 1967 work on the same reaction identified two decay energies in the ranges 8 35 8 50 MeV and 8 50 8 60 MeV these were assigned to 256103 and 257103 52 Despite repeat attempts they were unable to confirm assignment of an alpha emitter with a half life of 8 seconds to 257103 56 57 The Russians proposed the name rutherfordium for the new element in 1967 51 58 this name was later proposed by Berkeley for element 104 58 24395 Am 188 O 261103 Lr 256103 Lr 5 10 nFurther experiments in 1969 at Dubna and in 1970 at Berkeley demonstrated an actinide chemistry for the new element so by 1970 it was known that element 103 is the last actinide 52 59 In 1970 the Dubna group reported the synthesis of 255103 with half life 20 s and alpha decay energy 8 38 MeV 52 However it was not until 1971 when the nuclear physics team at University of California at Berkeley successfully did a whole series of experiments aimed at measuring the nuclear decay properties of the lawrencium isotopes with mass numbers 255 to 260 60 61 that all previous results from Berkeley and Dubna were confirmed apart from the Berkeley s group initial erroneous assignment of their first produced isotope to 257103 instead of the probably correct 258103 52 All final doubts were dispelled in 1976 and 1977 when the energies of X rays emitted from 258103 were measured 52 nbsp The element was named after Ernest Lawrence In 1971 the IUPAC granted the discovery of lawrencium to the Lawrence Berkeley Laboratory even though they did not have ideal data for the element s existence But in 1992 the IUPAC Transfermium Working Group TWG officially recognized the nuclear physics teams at Dubna and Berkeley as co discoverers of lawrencium concluding that while the 1961 Berkeley experiments were an important step to lawrencium s discovery they were not yet fully convincing and while the 1965 1968 and 1970 Dubna experiments came very close to the needed level of confidence taken together only the 1971 Berkeley experiments which clarified and confirmed previous observations finally resulted in complete confidence in the discovery of element 103 51 55 Because the name lawrencium had been in use for a long time by this point it was retained by IUPAC 51 and in August 1997 the International Union of Pure and Applied Chemistry IUPAC ratified the name lawrencium and the symbol Lr during a meeting in Geneva 55 Characteristics editPhysical edit Lawrencium is the last actinide Authors considering the subject generally consider it a group 3 element along with scandium yttrium and lutetium as its filled f shell is expected to make it resemble the other 7th period transition metals In the periodic table it is to the right of the actinide nobelium to the left of the 6d transition metal rutherfordium and under the lanthanide lutetium with which it shares many physical and chemical properties Lawrencium is expected to be a solid under normal conditions and have a hexagonal close packed crystal structure c a 1 58 similar to its lighter congener lutetium though this is not yet known experimentally 5 The enthalpy of sublimation of lawrencium is estimated at 352 kJ mol close to the value of lutetium and strongly suggesting that metallic lawrencium is trivalent with three electrons delocalized a prediction also supported by a systematic extrapolation of the values of heat of vaporization bulk modulus and atomic volume of neighboring elements to lawrencium 62 this makes it unlike the immediately preceding late actinides which are known to be fermium and mendelevium or expected to be nobelium divalent 63 The estimated enthalpies of vaporization show that lawrencium deviates from the trend of the late actinides and instead matches the trend of the succeeding 6d elements rutherfordium and dubnium 64 65 consistent with lawrencium s interpretation as a group 3 element 65 Some scientists prefer to end the actinides with nobelium and consider lawrencium to be the first transition metal of the seventh period 66 67 Specifically lawrencium is expected to be a trivalent silvery metal easily oxidized by air steam and acids 68 and having an atomic volume similar to that of lutetium and a trivalent metallic radius of 171 pm 62 It is expected to be a rather heavy metal with a density of around 14 4 g cm3 2 It is also predicted to have a melting point of around 1900 K 1600 C not far from the value for lutetium 1925 K 69 Chemical edit nbsp Elution sequence of the late trivalent lanthanides and actinides with ammonium a HIB as eluant the broken curve for lawrencium is a prediction In 1949 Glenn T Seaborg who devised the actinide concept predicted that element 103 lawrencium should be the last actinide and that the Lr3 ion should be about as stable as Lu3 in aqueous solution It was not until decades later that element 103 was finally conclusively synthesized and this prediction was experimentally confirmed 70 1969 studies on the element showed that lawrencium reacts with chlorine to form a product that was most likely the trichloride LrCl3 Its volatility was found to be similar to the chlorides of curium fermium and nobelium and much less than that of rutherfordium chloride In 1970 chemical studies were performed on 1500 atoms of 256Lr comparing it with divalent No Ba Ra trivalent Fm Cf Cm Am Ac and tetravalent Th Pu elements It was found that lawrencium coextracted with the trivalent ions but the short half life of 256Lr precluded a confirmation that it eluted ahead of Md3 in the elution sequence 70 Lawrencium occurs as the trivalent Lr3 ion in aqueous solution and hence its compounds should be similar to those of the other trivalent actinides for example lawrencium III fluoride LrF3 and hydroxide Lr OH 3 should both be insoluble in water 70 Due to the actinide contraction the ionic radius of Lr3 should be smaller than that of Md3 and it should elute ahead of Md3 when ammonium a hydroxyisobutyrate ammonium a HIB is used as an eluant 70 Later 1987 experiments on the longer lived isotope 260Lr confirmed lawrencium s trivalency and that it eluted in roughly the same place as erbium and found that lawrencium s ionic radius was 88 6 0 3 pm larger than would be expected from simple extrapolation from periodic trends 70 Later 1988 experiments with more lawrencium atoms refined this to 88 1 0 1 pm and calculated an enthalpy of hydration value of 3685 13 kJ mol 70 It was also found that the actinide contraction at the end of the actinides was larger than the analogous lanthanide contraction with the exception of the last actinide lawrencium the cause was speculated to be relativistic effects 70 It has been speculated that the 7s electrons are relativistically stabilized so that in reducing conditions only the 7p1 2 electron would be ionized leading to the monovalent Lr ion However all experiments to reduce Lr3 to Lr2 or Lr in aqueous solution were unsuccessful similarly to lutetium On the basis of this the standard electrode potential of the E Lr3 Lr couple was calculated to be less than 1 56 V indicating that the existence of Lr ions in aqueous solution was unlikely The upper limit for the E Lr3 Lr2 couple was predicted to be 0 44 V the values for E Lr3 Lr and E Lr4 Lr3 are predicted to be 2 06 V and 7 9 V 70 The stability of the group oxidation state in the 6d transition series decreases as RfIV gt DbV gt SgVI and lawrencium continues the trend with LrIII being more stable than RfIV 71 In the molecule lawrencium dihydride LrH2 which is predicted to be bent the 6d orbital of lawrencium is not expected to play a role in the bonding unlike that of lanthanum dihydride LaH2 LaH2 has La H bond distances of 2 158 A while LrH2 should have shorter Lr H bond distances of 2 042 A due to the relativistic contraction and stabilization of the 7s and 7p orbitals involved in the bonding in contrast to the core like 5f subshell and the mostly uninvolved 6d subshell In general molecular LrH2 and LrH are expected to resemble the corresponding thallium species thallium having a 6s26p1 valence configuration in the gas phase like lawrencium s 7s27p1 more than the corresponding lanthanide species 72 The electron configurations of Lr and Lr2 are expected to be 7s2 and 7s1 respectively However in species where all three valence electrons of lawrencium are ionized to give at least formally the Lr3 cation lawrencium is expected to behave like a typical actinide and the heavier congener of lutetium especially because the first three ionization potentials of lawrencium are predicted to be similar to those of lutetium Hence unlike thallium but like lutetium lawrencium would prefer to form LrH3 than LrH and LrCO is expected to be similar to the also unknown LuCO both metals having valence configuration s2p1 in their monocarbonyls The pp dp bond is expected to be seen in LrCl3 just as it is for LuCl3 and more generally all the LnCl3 The complex anion Lr C5H4SiMe3 3 is expected to be stable with a configuration of 6d1 for lawrencium this 6d orbital would be its highest occupied molecular orbital This is analogous to the electronic structure of the analogous lutetium compound 73 Atomic edit Lawrencium has three valence electrons the 5f electrons are in the atomic core 74 In 1970 it was predicted that the ground state electron configuration of lawrencium was Rn 5f146d17s2 ground state term symbol 2D3 2 per the Aufbau principle and conforming to the Xe 4f145d16s2 configuration of lawrencium s lighter homolog lutetium 75 But the next year calculations were published that questioned this prediction instead expecting an anomalous Rn 5f147s27p1 configuration 75 Though early calculations gave conflicting results 76 more recent studies and calculations confirm the s2p suggestion 77 78 1974 relativistic calculations concluded that the energy difference between the two configurations was small and that it was uncertain which was the ground state 75 Later 1995 calculations concluded that the s2p configuration should be energetically favored because the spherical s and p1 2 orbitals are nearest to the atomic nucleus and thus move quickly enough that their relativistic mass increases significantly 75 In 1988 a team of scientists led by Eichler calculated that lawrencium s enthalpy of adsorption on metal sources would differ enough depending on its electron configuration that it would be feasible to carry out experiments to exploit this fact to measure lawrencium s electron configuration 75 The s2p configuration was expected to be more volatile than the s2d configuration and be more similar to that of the p block element lead No evidence for lawrencium being volatile was obtained and the lower limit for the enthalpy of adsorption of lawrencium on quartz or platinum was significantly higher than the estimated value for the s2p configuration 75 nbsp First ionization energy eV plotted against atomic number in units eV Predicted values are used beyond rutherfordium element 104 Lawrencium element 103 has a very low first ionization energy fitting the start of the d block trend better than the end of the f block trend before it 79 In 2015 the first ionization energy of lawrencium was measured using the isotope 256Lr 4 The measured value 4 96 0 08 0 07 eV agreed very well with the relativistic theoretical prediction of 4 963 15 eV and also provided a first step into measuring the first ionization energies of the transactinides 4 This value is the lowest among all the lanthanides and actinides and supports the s2p configuration as the 7p1 2 electron is expected to be only weakly bound As ionisation energies generally increase left to right in the f block this low value suggests that lutetium and lawrencium belong in the d block whose trend they follow and not the f block That would make them the heavier congeners of scandium and yttrium rather than lanthanum and actinium 79 Although some alkali metal like behaviour has been predicted 80 adsorption experiments suggest that lawrencium is trivalent like scandium and yttrium not monovalent like the alkali metals 64 A lower limit on lawrencium s second ionization energy gt 13 3 eV was experimentally found in 2021 81 Even though s2p is now known to be the ground state configuration of the lawrencium atom ds2 should be a low lying excited state configuration with an excitation energy variously calculated as 0 156 eV 0 165 eV or 0 626 eV 73 As such lawrencium may still be considered to be a d block element albeit with an anomalous electron configuration like chromium or copper as its chemical behaviour matches expectations for a heavier analogue of lutetium 65 Isotopes edit Main article Isotopes of lawrencium Fourteen isotopes of lawrencium are known with mass number 251 262 264 and 266 all are radioactive 82 83 84 Seven nuclear isomers are known The longest lived isotope 266Lr has a half life of about ten hours and is one of the longest lived superheavy isotopes known to date 85 However shorter lived isotopes are usually used in chemical experiments because 266Lr currently can only be produced as a final decay product of even heavier and harder to make elements it was discovered in 2014 in the decay chain of 294Ts 82 83 256Lr half life 27 seconds was used in the first chemical studies on lawrencium currently the longer lived 260Lr half life 2 7 minutes is usually used for this purpose 82 After 266Lr the longest lived isotopes are 264Lr 4 8 2 2 1 3 h 262Lr 3 6 h and 261Lr 44 min 82 86 87 All other known lawrencium isotopes have half lives under 5 minutes and the shortest lived of them 251Lr has a half life of 24 4 milliseconds 84 86 87 88 The half lives of lawrencium isotopes mostly increase smoothly from 251Lr to 266Lr with a dip from 257Lr to 259Lr 82 86 87 Preparation and purification editMost isotopes of lawrencium can be produced by bombarding actinide americium to einsteinium targets with light ions from boron to neon The two most important isotopes 256Lr and 260Lr can be respectively produced by bombarding californium 249 with 70 MeV boron 11 ions producing lawrencium 256 and four neutrons and by bombarding berkelium 249 with oxygen 18 producing lawrencium 260 an alpha particle and three neutrons 89 The two heaviest and longest lived known isotopes 264Lr and 266Lr can only be produced at much lower yields as decay products of dubnium whose progenitors are isotopes of moscovium and tennessine Both 256Lr and 260Lr have half lives too short to allow a complete chemical purification process Early experiments with 256Lr therefore used rapid solvent extraction with the chelating agent thenoyltrifluoroacetone TTA dissolved in methyl isobutyl ketone MIBK as the organic phase and with the aqueous phase being buffered acetate solutions Ions of different charge 2 3 or 4 will then extract into the organic phase under different pH ranges but this method will not separate the trivalent actinides and thus 256Lr must be identified by its emitted 8 24 MeV alpha particles 89 More recent methods have allowed rapid selective elution with a HIB to take place in enough time to separate out the longer lived isotope 260Lr which can be removed from the catcher foil with 0 05 M hydrochloric acid 89 See also editPortal nbsp ChemistryLawrencium at Wikipedia s sister projects nbsp Definitions from Wiktionary nbsp Media from CommonsNotes edit In nuclear physics an element is called heavy if its atomic number is high lead element 82 is one example of such a heavy element The term superheavy elements typically refers to elements with atomic number greater than 103 although there are other definitions such as atomic number greater than 100 8 or 112 9 sometimes the term is presented an equivalent to the term transactinide which puts an upper limit before the beginning of the hypothetical superactinide series 10 Terms heavy isotopes of a given element and heavy nuclei mean what could be understood in the common language isotopes of high mass for the given element and nuclei of high mass respectively In 2009 a team at the JINR led by Oganessian published results of their attempt to create hassium in a symmetric 136Xe 136Xe reaction They failed to observe a single atom in such a reaction putting the upper limit on the cross section the measure of probability of a nuclear reaction as 2 5 pb 11 In comparison the reaction that resulted in hassium discovery 208Pb 58Fe had a cross section of 20 pb more specifically 19 19 11 pb as estimated by the discoverers 12 The amount of energy applied to the beam particle to accelerate it can also influence the value of cross section For example in the 2814 Si 10 n 2813 Al 11 p reaction cross section changes smoothly from 370 mb at 12 3 MeV to 160 mb at 18 3 MeV with a broad peak at 13 5 MeV with the maximum value of 380 mb 16 This figure also marks the generally accepted upper limit for lifetime of a compound nucleus 21 This separation is based on that the resulting nuclei move past the target more slowly then the unreacted beam nuclei The separator contains electric and magnetic fields whose effects on a moving particle cancel out for a specific velocity of a particle 23 Such separation can also be aided by a time of flight measurement and a recoil energy measurement a combination of the two may allow to estimate the mass of a nucleus 24 Not all decay modes are caused by electrostatic repulsion For example beta decay is caused by the weak interaction 31 It was already known by the 1960s that ground states of nuclei differed in energy and shape as well as that certain magic numbers of nucleons corresponded to greater stability of a nucleus However it was assumed that there was no nuclear structure in superheavy nuclei as they were too deformed to form one 36 Since mass of a nucleus is not measured directly but is rather calculated from that of another nucleus such measurement is called indirect Direct measurements are also possible but for the most part they have remained unavailable for superheavy nuclei 41 The first direct measurement of mass of a superheavy nucleus was reported in 2018 at LBNL 42 Mass was determined from the location of a nucleus after the transfer the location helps determine its trajectory which is linked to the mass to charge ratio of the nucleus since the transfer was done in presence of a magnet 43 If the decay occurred in a vacuum then since total momentum of an isolated system before and after the decay must be preserved the daughter nucleus would also receive a small velocity The ratio of the two velocities and accordingly the ratio of the kinetic energies would thus be inverse to the ratio of the two masses The decay energy equals the sum of the known kinetic energy of the alpha particle and that of the daughter nucleus an exact fraction of the former 32 The calculations hold for an experiment as well but the difference is that the nucleus does not move after the decay because it is tied to the detector Spontaneous fission was discovered by Soviet physicist Georgy Flerov 44 a leading scientist at JINR and thus it was a hobbyhorse for the facility 45 In contrast the LBL scientists believed fission information was not sufficient for a claim of synthesis of an element They believed spontaneous fission had not been studied enough to use it for identification of a new element since there was a difficulty of establishing that a compound nucleus had only ejected neutrons and not charged particles like protons or alpha particles 21 They thus preferred to link new isotopes to the already known ones by successive alpha decays 44 For instance element 102 was mistakenly identified in 1957 at the Nobel Institute of Physics in Stockholm Stockholm County Sweden 46 There were no earlier definitive claims of creation of this element and the element was assigned a name by its Swedish American and British discoverers nobelium It was later shown that the identification was incorrect 47 The following year RL was unable to reproduce the Swedish results and announced instead their synthesis of the element that claim was also disproved later 47 JINR insisted that they were the first to create the element and suggested a name of their own for the new element joliotium 48 the Soviet name was also not accepted JINR later referred to the naming of the element 102 as hasty 49 This name was proposed to IUPAC in a written response to their ruling on priority of discovery claims of elements signed 29 September 1992 49 The name nobelium remained unchanged on account of its widespread usage 50 References edit Emsley John 2011 Nature s Building Blocks An A Z Guide to the Elements New ed New York NY Oxford University Press p 278 279 ISBN 978 0 19 960563 7 a b Gyanchandani Jyoti Sikka S K 10 May 2011 Physical properties of the 6 d series elements from density functional theory Close similarity to lighter transition metals Physical Review B 83 17 172101 Bibcode 2011PhRvB 83q2101G doi 10 1103 PhysRevB 83 172101 Brown Geoffrey 2012 The Inaccessible Earth An integrated view to its structure and composition Springer Science amp Business Media p 88 ISBN 9789401115162 a b c Sato T K Asai M Borschevsky A Stora T Sato N Kaneya Y Tsukada K Dullman Ch E Eberhardt K Eliav E Ichikawa S Kaldor U Kratz J V Miyashita S Nagame Y Ooe K Osa A Renisch D Runke J Schadel M Thorle Pospiech P Toyoshima A Trautmann N 9 April 2015 Measurement of the first ionization potential of lawrencium element 103 PDF Nature 520 7546 209 11 Bibcode 2015Natur 520 209S doi 10 1038 nature14342 PMID 25855457 S2CID 4384213 a b Ostlin A Vitos L 2011 First principles calculation of the structural stability of 6d transition metals Physical Review B 84 11 113104 Bibcode 2011PhRvB 84k3104O doi 10 1103 PhysRevB 84 113104 Kondev F G Wang M Huang W J Naimi S Audi G 2021 The NUBASE2020 evaluation of nuclear properties PDF Chinese Physics C 45 3 030001 doi 10 1088 1674 1137 abddae Oganessian Yu Ts Utyonkov V K Kovrizhnykh N D et al 2022 New isotope 286Mc produced in the 243Am 48Ca reaction Physical Review C 106 064306 doi 10 1103 PhysRevC 106 064306 Kramer K 2016 Explainer superheavy elements Chemistry World Retrieved 2020 03 15 Discovery of Elements 113 and 115 Lawrence Livermore National Laboratory Archived from the original on 2015 09 11 Retrieved 2020 03 15 Eliav E Kaldor U Borschevsky A 2018 Electronic Structure of the Transactinide Atoms In Scott R A ed Encyclopedia of Inorganic and Bioinorganic Chemistry John Wiley amp Sons pp 1 16 doi 10 1002 9781119951438 eibc2632 ISBN 978 1 119 95143 8 S2CID 127060181 Oganessian Yu Ts Dmitriev S N Yeremin A V et al 2009 Attempt to produce the isotopes of element 108 in the fusion reaction 136Xe 136Xe Physical Review C 79 2 024608 doi 10 1103 PhysRevC 79 024608 ISSN 0556 2813 Munzenberg G Armbruster P Folger H et al 1984 The identification of element 108 PDF Zeitschrift fur Physik A 317 2 235 236 Bibcode 1984ZPhyA 317 235M doi 10 1007 BF01421260 S2CID 123288075 Archived from the original PDF on 7 June 2015 Retrieved 20 October 2012 Subramanian S 28 August 2019 Making New Elements Doesn t Pay Just Ask This Berkeley Scientist Bloomberg Businessweek Retrieved 2020 01 18 a b c d e f Ivanov D 2019 Sverhtyazhelye shagi v neizvestnoe Superheavy steps into the unknown nplus1 ru in Russian Retrieved 2020 02 02 Hinde D 2017 Something new and superheavy at the periodic table The Conversation Retrieved 2020 01 30 Kern B D Thompson W E Ferguson J M 1959 Cross sections for some n p and n a reactions Nuclear Physics 10 226 234 Bibcode 1959NucPh 10 226K doi 10 1016 0029 5582 59 90211 1 Wakhle A Simenel C Hinde D J et al 2015 Simenel C Gomes P R S Hinde D J et al eds Comparing Experimental and Theoretical Quasifission Mass Angle Distributions European Physical Journal Web of Conferences 86 00061 Bibcode 2015EPJWC 8600061W doi 10 1051 epjconf 20158600061 hdl 1885 148847 ISSN 2100 014X Nuclear Reactions PDF pp 7 8 Retrieved 2020 01 27 Published as Loveland W D Morrissey D J Seaborg G T 2005 Nuclear Reactions Modern Nuclear Chemistry John Wiley amp Sons Inc pp 249 297 doi 10 1002 0471768626 ch10 ISBN 978 0 471 76862 3 a b Krasa A 2010 Neutron Sources for ADS Faculty of Nuclear Sciences and Physical Engineering Czech Technical University in Prague 4 8 S2CID 28796927 Wapstra A H 1991 Criteria that must be satisfied for the discovery of a new chemical element to be recognized PDF Pure and Applied Chemistry 63 6 883 doi 10 1351 pac199163060879 ISSN 1365 3075 S2CID 95737691 a b Hyde E K Hoffman D C Keller O L 1987 A History and Analysis of the Discovery of Elements 104 and 105 Radiochimica Acta 42 2 67 68 doi 10 1524 ract 1987 42 2 57 ISSN 2193 3405 S2CID 99193729 a b c d Chemistry World 2016 How to Make Superheavy Elements and Finish the Periodic Table Video Scientific American Retrieved 2020 01 27 Hoffman Ghiorso amp Seaborg 2000 p 334 Hoffman Ghiorso amp Seaborg 2000 p 335 Zagrebaev Karpov amp Greiner 2013 p 3 Beiser 2003 p 432 a b Pauli N 2019 Alpha decay PDF Introductory Nuclear Atomic and Molecular Physics Nuclear Physics Part Universite libre de Bruxelles Retrieved 2020 02 16 a b c d e Pauli N 2019 Nuclear fission PDF Introductory Nuclear Atomic and Molecular Physics Nuclear Physics Part Universite libre de Bruxelles Retrieved 2020 02 16 Staszczak A Baran A Nazarewicz W 2013 Spontaneous fission modes and lifetimes of superheavy elements in the nuclear density functional theory Physical Review C 87 2 024320 1 arXiv 1208 1215 Bibcode 2013PhRvC 87b4320S doi 10 1103 physrevc 87 024320 ISSN 0556 2813 Audi et al 2017 pp 030001 129 030001 138 Beiser 2003 p 439 a b Beiser 2003 p 433 Audi et al 2017 p 030001 125 Aksenov N V Steinegger P Abdullin F Sh et al 2017 On the volatility of nihonium Nh Z 113 The European Physical Journal A 53 7 158 Bibcode 2017EPJA 53 158A doi 10 1140 epja i2017 12348 8 ISSN 1434 6001 S2CID 125849923 Beiser 2003 p 432 433 a b c Oganessian Yu 2012 Nuclei in the Island of Stability of Superheavy Elements Journal of Physics Conference Series 337 1 012005 1 012005 6 Bibcode 2012JPhCS 337a2005O doi 10 1088 1742 6596 337 1 012005 ISSN 1742 6596 Moller P Nix J R 1994 Fission properties of the heaviest elements PDF Dai 2 Kai Hadoron Tataikei no Simulation Symposium Tokai mura Ibaraki Japan University of North Texas Retrieved 2020 02 16 a b Oganessian Yu Ts 2004 Superheavy elements Physics World 17 7 25 29 doi 10 1088 2058 7058 17 7 31 Retrieved 2020 02 16 Schadel M 2015 Chemistry of the superheavy elements Philosophical Transactions of the Royal Society A Mathematical Physical and Engineering Sciences 373 2037 20140191 Bibcode 2015RSPTA 37340191S doi 10 1098 rsta 2014 0191 ISSN 1364 503X PMID 25666065 Hulet E K 1989 Biomodal spontaneous fission 50th Anniversary of Nuclear Fission Leningrad USSR Bibcode 1989nufi rept 16H Oganessian Yu Ts Rykaczewski K P 2015 A beachhead on the island of stability Physics Today 68 8 32 38 Bibcode 2015PhT 68h 32O doi 10 1063 PT 3 2880 ISSN 0031 9228 OSTI 1337838 S2CID 119531411 Grant A 2018 Weighing the heaviest elements Physics Today doi 10 1063 PT 6 1 20181113a S2CID 239775403 Howes L 2019 Exploring the superheavy elements at the end of the periodic table Chemical amp Engineering News Retrieved 2020 01 27 a b Robinson A E 2019 The Transfermium Wars Scientific Brawling and Name Calling during the Cold War Distillations Retrieved 2020 02 22 Populyarnaya biblioteka himicheskih elementov Siborgij ekavolfram Popular library of chemical elements Seaborgium eka tungsten n t ru in Russian Retrieved 2020 01 07 Reprinted from Ekavolfram Eka tungsten Populyarnaya biblioteka himicheskih elementov Serebro Nilsborij i dalee Popular library of chemical elements Silver through nielsbohrium and beyond in Russian Nauka 1977 Nobelium Element information properties and uses Periodic Table Royal Society of Chemistry Retrieved 2020 03 01 a b Kragh 2018 pp 38 39 Kragh 2018 p 40 a b Ghiorso A Seaborg G T Oganessian Yu Ts et al 1993 Responses on the report Discovery of the Transfermium elements followed by reply to the responses by Transfermium Working Group PDF Pure and Applied Chemistry 65 8 1815 1824 doi 10 1351 pac199365081815 S2CID 95069384 Archived PDF from the original on 25 November 2013 Retrieved 7 September 2016 Commission on Nomenclature of Inorganic Chemistry 1997 Names and symbols of transfermium elements IUPAC Recommendations 1997 PDF Pure and Applied Chemistry 69 12 2471 2474 doi 10 1351 pac199769122471 a b c d e Emsley John 2011 Nature s Building Blocks a b c d e f g h i j k Barber R C Greenwood N N Hrynkiewicz A Z Jeannin Y P Lefort M Sakai M Ulehla I Wapstra A P Wilkinson D H 1993 Discovery of the transfermium elements Part II Introduction to discovery profiles Part III Discovery profiles of the transfermium elements Pure and Applied Chemistry 65 8 1757 doi 10 1351 pac199365081757 S2CID 195819585 Note for Part I see Pure Appl Chem Vol 63 No 6 pp 879 886 1991 This Month in Lab History Lawrencium Added to Periodic Table today lbl gov Lawrence Berkeley National Laboratory 9 April 2013 Retrieved 13 February 2021 Lawrencium Lw was first synthesized Feb 14 1961 by a team led by Ghiorso who was co discoverer of a record 12 chemical elements on the periodic table a b Ghiorso Albert Sikkeland T Larsh A E Latimer R M 1961 New Element Lawrencium Atomic Number 103 Phys Rev Lett 6 9 473 Bibcode 1961PhRvL 6 473G doi 10 1103 PhysRevLett 6 473 a b c Greenwood Norman N 1997 Recent developments concerning the discovery of elements 101 111 PDF Pure Appl Chem 69 1 179 184 doi 10 1351 pac199769010179 S2CID 98322292 Flerov G N 1967 On the nuclear properties of the isotopes 256103 and 257103 Nucl Phys A 106 2 476 Bibcode 1967NuPhA 106 476F doi 10 1016 0375 9474 67 90892 5 Donets E D Shchegolev V A Ermakov V A 1965 Atomnaya Energiya in Russian 19 2 109 a href Template Cite journal html title Template Cite journal cite journal a Missing or empty title help Translated in Donets E D Shchegolev V A Ermakov V A 1965 Synthesis of the isotope of element 103 lawrencium with mass number 256 Soviet Atomic Energy 19 2 109 doi 10 1007 BF01126414 S2CID 97218361 a b Karpenko V 1980 The Discovery of Supposed New Elements Two Centuries of Errors Ambix 27 2 77 102 doi 10 1179 amb 1980 27 2 77 Kaldor Uzi amp Wilson Stephen 2005 Theoretical chemistry and physics of heavy and superheavy element Springer p 57 ISBN 1 4020 1371 X Silva 2011 pp 1641 2 Eskola Kari Eskola Pirkko Nurmia Matti Albert Ghiorso 1971 Studies of Lawrencium Isotopes with Mass Numbers 255 Through 260 Phys Rev C 4 2 632 642 Bibcode 1971PhRvC 4 632E doi 10 1103 PhysRevC 4 632 a b Silva 2011 p 1644 Silva 2011 p 1639 a b Haire R G 11 October 2007 Insights into the bonding and electronic nature of heavy element materials Journal of Alloys and Compounds 444 5 63 71 doi 10 1016 j jallcom 2007 01 103 a b c Jensen William B 2015 The positions of lanthanum actinium and lutetium lawrencium in the periodic table an update Foundations of Chemistry 17 23 31 doi 10 1007 s10698 015 9216 1 S2CID 98624395 Archived from the original on 30 January 2021 Retrieved 28 January 2021 Winter Mark 1993 2022 WebElements The University of Sheffield and WebElements Ltd UK Retrieved 5 December 2022 Cowan Robert D 1981 The Theory of Atomic Structure and Spectra University of California Press p 598 ISBN 9780520906150 John Emsley 2011 Nature s Building Blocks An A Z Guide to the Elements Oxford University Press pp 278 9 ISBN 978 0 19 960563 7 Lide D R ed 2003 CRC Handbook of Chemistry and Physics 84th ed Boca Raton FL CRC Press a b c d e f g h Silva 2011 pp 1644 7 Hoffman Darleane C Lee Diana M Pershina Valeria 2006 Transactinides and the future elements In Morss Edelstein Norman M Fuger Jean eds The Chemistry of the Actinide and Transactinide Elements 3rd ed Dordrecht The Netherlands Springer Science Business Media p 1686 ISBN 1 4020 3555 1 Balasubramanian K 4 December 2001 Potential energy surfaces of Lawrencium and Nobelium dihydrides LrH2 and NoH2 Journal of Chemical Physics 116 9 3568 75 Bibcode 2002JChPh 116 3568B doi 10 1063 1 1446029 a b Xu Wen Hua Pyykko Pekka 8 June 2016 Is the chemistry of lawrencium peculiar Phys Chem Chem Phys 2016 18 17351 5 Bibcode 2016PCCP 1817351X doi 10 1039 c6cp02706g hdl 10138 224395 PMID 27314425 S2CID 31224634 Retrieved 24 April 2017 Jensen William B 2000 The Periodic Law and Table PDF Archived from the original PDF on 2020 11 10 Retrieved 10 December 2022 a b c d e f Silva 2011 pp 1643 4 Nugent L J Vander Sluis K L Fricke Burhard Mann J B 1974 Electronic configuration in the ground state of atomic lawrencium PDF Phys Rev A 9 6 2270 72 Bibcode 1974PhRvA 9 2270N doi 10 1103 PhysRevA 9 2270 Eliav E Kaldor U Ishikawa Y 1995 Transition energies of ytterbium lutetium and lawrencium by the relativistic coupled cluster method Phys Rev A 52 1 291 296 Bibcode 1995PhRvA 52 291E doi 10 1103 PhysRevA 52 291 PMID 9912247 Zou Yu Froese Fischer C Uiterwaal C Wanner J Kompa K L 2002 Resonance Transition Energies and Oscillator Strengths in Lutetium and Lawrencium Phys Rev Lett 88 2 183001 Bibcode 2001PhRvL 88b3001M doi 10 1103 PhysRevLett 88 023001 PMID 12005680 S2CID 18391594 a b Jensen W B 2015 Some Comments on the Position of Lawrencium in the Periodic Table PDF Archived from the original PDF on 23 December 2015 Retrieved 20 September 2015 Gunther Matthew 9 April 2015 Lawrencium experiment could shake up periodic table RSC Chemistry World Retrieved 21 September 2015 Kwarsick Jeffrey T Pore Jennifer L Gates Jacklyn M Gregorich Kenneth E Gibson John K Jian Jiwen Pang Gregory K Shuh David K 2021 Assessment of the Second Ionization Potential of Lawrencium Investigating the End of the Actinide Series with a One Atom at a Time Gas Phase Ion Chemistry Technique The Journal of Physical Chemistry A 125 31 6818 6828 Bibcode 2021JPCA 125 6818K doi 10 1021 acs jpca 1c01961 OSTI 1844939 PMID 34242037 S2CID 235785891 a b c d e Silva 2011 p 1642 a b Khuyagbaatar J et al 2014 48Ca 249Bk Fusion Reaction Leading to Element Z 117 Long Lived a Decaying 270Db and Discovery of 266Lr PDF Physical Review Letters 112 17 172501 Bibcode 2014PhRvL 112q2501K doi 10 1103 PhysRevLett 112 172501 hdl 1885 70327 PMID 24836239 S2CID 5949620 a b Leppanen A P 2005 Alpha decay and decay tagging studies of heavy elements using the RITU separator PDF Thesis University of Jyvaskyla pp 83 100 ISBN 978 951 39 3162 9 ISSN 0075 465X Clara Moskowitz May 7 2014 Superheavy Element 117 Points to Fabled Island of Stability on Periodic Table Scientific American Retrieved 2014 05 08 a b c Nucleonica Web driven nuclear science a b c Audi G Kondev F G Wang M Huang W J Naimi S 2017 The NUBASE2016 evaluation of nuclear properties PDF Chinese Physics C 41 3 030001 Bibcode 2017ChPhC 41c0001A doi 10 1088 1674 1137 41 3 030001 Huang T Seweryniak D Back B B et al 2022 Discovery of the new isotope 251Lr Impact of the hexacontetrapole deformation on single proton orbital energies near the Z 100 deformed shell gap Physical Review C 106 L061301 doi 10 1103 PhysRevC 106 L061301 S2CID 254300224 a b c Silva 2011 pp 1642 3Bibliography editAudi G Kondev F G Wang M et al 2017 The NUBASE2016 evaluation of nuclear properties Chinese Physics C 41 3 030001 Bibcode 2017ChPhC 41c0001A doi 10 1088 1674 1137 41 3 030001 Beiser A 2003 Concepts of modern physics 6th ed McGraw Hill ISBN 978 0 07 244848 1 OCLC 48965418 Hoffman D C Ghiorso A Seaborg G T 2000 The Transuranium People The Inside Story World Scientific ISBN 978 1 78 326244 1 Kragh H 2018 From Transuranic to Superheavy Elements A Story of Dispute and Creation Springer ISBN 978 3 319 75813 8 Silva Robert J 2011 Chapter 13 Fermium Mendelevium Nobelium and Lawrencium In Morss Lester R Edelstein Norman M Fuger Jean eds The Chemistry of the Actinide and Transactinide Elements Netherlands Springer doi 10 1007 978 94 007 0211 0 13 ISBN 978 94 007 0210 3 Zagrebaev V Karpov A Greiner W 2013 Future of superheavy element research Which nuclei could be synthesized within the next few years Journal of Physics Conference Series 420 1 012001 arXiv 1207 5700 Bibcode 2013JPhCS 420a2001Z doi 10 1088 1742 6596 420 1 012001 ISSN 1742 6588 S2CID 55434734 External links edit Chart of Nuclides National Nuclear Data Center NNDC Archived from the original on 2018 10 10 Retrieved 2014 08 21 Los Alamos National Laboratory s Chemistry Division Periodic Table Lawrencium Lawrencium at The Periodic Table of Videos University of Nottingham Retrieved from https en wikipedia org w index php title Lawrencium amp oldid 1203235448, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.