Isotopes of lead

Last updated
Isotopes of lead  (82Pb)
Main isotopes [1] Decay
abun­dance half-life (t1/2) mode pro­duct
202Pb synth 5.25×104 y ε 202Tl
204Pb1.40% stable
205Pb trace 1.73×107 yε 205Tl
206Pb24.1%stable
207Pb22.1%stable
208Pb52.4%stable
209Pbtrace3.253 h β 209Bi
210Pbtrace22.20 yβ 210Bi
211Pbtrace36.1 minβ 211Bi
212Pbtrace10.64 hβ 212Bi
214Pbtrace26.8 minβ 214Bi
Isotopic abundances vary greatly by sample [2]
Standard atomic weight Ar°(Pb)

Lead (82Pb) has four observationally stable isotopes: 204Pb, 206Pb, 207Pb, 208Pb. Lead-204 is entirely a primordial nuclide and is not a radiogenic nuclide. The three isotopes lead-206, lead-207, and lead-208 represent the ends of three decay chains: the uranium series (or radium series), the actinium series, and the thorium series, respectively; a fourth decay chain, the neptunium series, terminates with the thallium isotope 205Tl. The three series terminating in lead represent the decay chain products of long-lived primordial 238U, 235U, and 232Th. Each isotope also occurs, to some extent, as primordial isotopes that were made in supernovae, rather than radiogenically as daughter products. The fixed ratio of lead-204 to the primordial amounts of the other lead isotopes may be used as the baseline to estimate the extra amounts of radiogenic lead present in rocks as a result of decay from uranium and thorium. (See lead–lead dating and uranium–lead dating.)

Contents

The longest-lived radioisotopes are 205Pb with a half-life of 17.3 million years and 202Pb with a half-life of 52,500 years. A shorter-lived naturally occurring radioisotope, 210Pb with a half-life of 22.2 years, is useful for studying the sedimentation chronology of environmental samples on time scales shorter than 100 years. [5]

The relative abundances of the four stable isotopes are approximately 1.5%, 24%, 22%, and 52.5%, combining to give a standard atomic weight (abundance-weighted average of the stable isotopes) of 207.2(1). Lead is the element with the heaviest stable isotope, 208Pb. (The more massive 209Bi, long considered to be stable, actually has a half-life of 2.01×1019 years.) 208Pb is also a doubly magic isotope, as it has 82 protons and 126 neutrons. [6] It is the heaviest doubly magic nuclide known. A total of 43 lead isotopes are now known, including very unstable synthetic species.

The four primordial isotopes of lead are all observationally stable, meaning that they are predicted to undergo radioactive decay but no decay has been observed yet. These four isotopes are predicted to undergo alpha decay and become isotopes of mercury which are themselves radioactive or observationally stable.

In its fully ionized state, the beta decay of isotope 210Pb does not release a free electron; the generated electron is instead captured by the atom's empty orbitals. [7]

List of isotopes

Nuclide [8]
[n 1]
Historic
name
Z N Isotopic mass (Da) [9]
[n 2] [n 3]
Half-life
Decay
mode

[n 4]
Daughter
isotope

[n 5] [n 6]
Spin and
parity
[n 7] [n 8]
Natural abundance (mole fraction)
Excitation energy [n 8] Normal proportionRange of variation
178Pb8296178.003830(26)0.23(15) ms α 174Hg0+
179Pb8297179.00215(21)#3.9(1.1) msα175Hg(9/2−)
180Pb8298179.997918(22)4.5(11) msα176Hg0+
181Pb8299180.99662(10)45(20) msα (98%)177Hg(9/2−)
β+ (2%)181Tl
182Pb82100181.992672(15)60(40) ms
[55(+40−35) ms]
α (98%)178Hg0+
β+ (2%)182Tl
183Pb82101182.99187(3)535(30) msα (94%)179Hg(3/2−)
β+ (6%)183Tl
183mPb94(8) keV415(20) msα179Hg(13/2+)
β+ (rare)183Tl
184Pb82102183.988142(15)490(25) msα180Hg0+
β+ (rare)184Tl
185Pb82103184.987610(17)6.3(4) sα181Hg3/2−
β+ (rare)185Tl
185mPb60(40)# keV4.07(15) sα181Hg13/2+
β+ (rare)185Tl
186Pb82104185.984239(12)4.82(3) sα (56%)182Hg0+
β+ (44%)186Tl
187Pb82105186.983918(9)15.2(3) sβ+187Tl(3/2−)
α183Hg
187mPb11(11) keV18.3(3) sβ+ (98%)187Tl(13/2+)
α (2%)183Hg
188Pb82106187.980874(11)25.5(1) sβ+ (91.5%)188Tl0+
α (8.5%)184Hg
188m1Pb2578.2(7) keV830(210) ns(8−)
188m2Pb2800(50) keV797(21) ns
189Pb82107188.98081(4)51(3) sβ+189Tl(3/2−)
189m1Pb40(30)# keV50.5(2.1) sβ+ (99.6%)189Tl13/2+
α (.4%)185Hg
189m2Pb2475(30)# keV26(5) μs(10)+
190Pb82108189.978082(13)71(1) sβ+ (99.1%)190Tl0+
α (.9%)186Hg
190m1Pb2614.8(8) keV150 ns(10)+
190m2Pb2618(20) keV25 μs(12+)
190m3Pb2658.2(8) keV7.2(6) μs(11)−
191Pb82109190.97827(4)1.33(8) minβ+ (99.987%)191Tl(3/2−)
α (.013%)187Hg
191mPb20(50) keV2.18(8) minβ+ (99.98%)191Tl13/2(+)
α (.02%)187Hg
192Pb82110191.975785(14)3.5(1) minβ+ (99.99%)192Tl0+
α (.0061%)188Hg
192m1Pb2581.1(1) keV164(7) ns(10)+
192m2Pb2625.1(11) keV1.1(5) μs(12+)
192m3Pb2743.5(4) keV756(21) ns(11)−
193Pb82111192.97617(5)5# minβ+193Tl(3/2−)
193m1Pb130(80)# keV5.8(2) minβ+193Tl13/2(+)
193m2Pb2612.5(5)+X keV135(+25−15) ns(33/2+)
194Pb82112193.974012(19)12.0(5) minβ+ (100%)194Tl0+
α (7.3×10−6%)190Hg
195Pb82113194.974542(25)~15 minβ+195Tl3/2#-
195m1Pb202.9(7) keV15.0(12) minβ+195Tl13/2+
195m2Pb1759.0(7) keV10.0(7) μs21/2−
196Pb82114195.972774(15)37(3) minβ+196Tl0+
α (<3×10−5%)192Hg
196m1Pb1049.20(9) keV<100 ns2+
196m2Pb1738.27(12) keV<1 μs4+
196m3Pb1797.51(14) keV140(14) ns5−
196m4Pb2693.5(5) keV270(4) ns(12+)
197Pb82115196.973431(6)8.1(17) minβ+197Tl3/2−
197m1Pb319.31(11) keV42.9(9) minβ+ (81%)197Tl13/2+
IT (19%)197Pb
α (3×10−4%)193Hg
197m2Pb1914.10(25) keV1.15(20) μs21/2−
198Pb82116197.972034(16)2.4(1) hβ+198Tl0+
198m1Pb2141.4(4) keV4.19(10) μs(7)−
198m2Pb2231.4(5) keV137(10) ns(9)−
198m3Pb2820.5(7) keV212(4) ns(12)+
199Pb82117198.972917(28)90(10) minβ+199Tl3/2−
199m1Pb429.5(27) keV12.2(3) minIT (93%)199Pb(13/2+)
β+ (7%)199Tl
199m2Pb2563.8(27) keV10.1(2) μs(29/2−)
200Pb82118199.971827(12)21.5(4) h EC 200Tl0+
201Pb82119200.972885(24)9.33(3) hEC (99%)201Tl5/2−
β+ (1%)
201m1Pb629.14(17) keV61(2) s13/2+
201m2Pb2718.5+X keV508(5) ns(29/2−)
202Pb82120201.972159(9)5.25(28)×104 yEC202Tl0+
202m1Pb2169.83(7) keV3.54(2) hIT (90.5%)202Pb9−
β+ (9.5%)202Tl
202m2Pb4142.9(11) keV110(5) ns(16+)
202m3Pb5345.9(13) keV107(5) ns(19−)
203Pb82121202.973391(7)51.873(9) hEC203Tl5/2−
203m1Pb825.20(9) keV6.21(8) sIT203Pb13/2+
203m2Pb2949.47(22) keV480(7) ms29/2−
203m3Pb2923.4+X keV122(4) ns(25/2−)
204Pb [n 9] 82122203.9730436(13) Observationally stable [n 10] 0+0.014(1)0.0000–0.0158 [10]
204m1Pb1274.00(4) keV265(10) ns4+
204m2Pb2185.79(5) keV67.2(3) min9−
204m3Pb2264.33(4) keV0.45(+10−3) μs7−
205Pb82123204.9744818(13)1.73(7)×107 yEC205Tl5/2−
205m1Pb2.329(7) keV24.2(4) μs1/2−
205m2Pb1013.839(13) keV5.55(2) ms13/2+
205m3Pb3195.7(5) keV217(5) ns25/2−
206Pb [n 9] [n 11] Radium G [11] 82124205.9744653(13)Observationally stable [n 12] [12] 0+0.241(1)0.0190–0.8673 [10]
206m1Pb2200.14(4) keV125(2) μs7−
206m2Pb4027.3(7) keV202(3) ns12+
207Pb [n 9] [n 13] Actinium D82125206.9758969(13)Observationally stable [n 14] [12] 1/2−0.221(1)0.0035–0.2351 [10]
207mPb1633.368(5) keV806(6) msIT207Pb13/2+
208Pb [n 15] Thorium D82126207.9766521(13)Observationally stable [n 16] [12] 0+0.524(1)0.0338–0.9775 [10]
208mPb4895(2) keV500(10) ns10+
209Pb82127208.9810901(19)3.253(14) hβ209Bi9/2+Trace [n 17]
210PbRadium D
Radiolead
Radio-lead
82128209.9841885(16)22.20(22) yβ (100%)210Bi0+Trace [n 18]
α (1.9×10−6%)206Hg
210mPb1278(5) keV201(17) ns8+
211PbActinium B82129210.9887370(29)36.1(2) minβ211Bi9/2+Trace [n 19]
212PbThorium B82130211.9918975(24)10.64(1) hβ212Bi0+Trace [n 20]
212mPb1335(10) keV6.0(0.8) μsIT212Pb(8+)
213Pb82131212.996581(8)10.2(3) minβ213Bi(9/2+)Trace [n 17]
214PbRadium B82132213.9998054(26)26.8(9) minβ214Bi0+Trace [n 18]
214mPb1420(20) keV6.2(0.3) μsIT212Pb8+#
215Pb82133215.004660(60)2.34(0.19) minβ215Bi9/2+#
216Pb82134216.008030(210)#1.65(0.2) minβ216Bi0+
216mPb1514(20) keV400(40) nsIT216Pb8+#
217Pb82135217.013140(320)#20(5) sβ217Bi9/2+#
218Pb82136218.016590(320)#15(7) sβ218Bi0+
This table header & footer:
  1. mPb  Excited nuclear isomer.
  2. ()  Uncertainty (1σ) is given in concise form in parentheses after the corresponding last digits.
  3. #  Atomic mass marked #: value and uncertainty derived not from purely experimental data, but at least partly from trends from the Mass Surface (TMS).
  4. Modes of decay:
    EC: Electron capture
    IT: Isomeric transition
  5. Bold italics symbol as daughter  Daughter product is nearly stable.
  6. Bold symbol as daughter  Daughter product is stable.
  7. () spin value  Indicates spin with weak assignment arguments.
  8. 1 2 #  Values marked # are not purely derived from experimental data, but at least partly from trends of neighboring nuclides (TNN).
  9. 1 2 3 Used in lead–lead dating
  10. Believed to undergo α decay to 200Hg with a half-life over 1.4×1020 years
  11. Final decay product of 4n+2 decay chain (the Radium or Uranium series)
  12. Experimental lower bound is years; the theoretical lifetime for α decay to 202Hg is years.
  13. Final decay product of 4n+3 decay chain (the Actinium series)
  14. Experimental lower bound is years; the theoretical lifetime for α decay to 203Hg is years.
  15. Heaviest observationally stable nuclide; final decay product of 4n decay chain (the Thorium series)
  16. Experimental lower bound is years; the theoretical lifetime for α decay to 204Hg is years.
  17. 1 2 Intermediate decay product of 237Np
  18. 1 2 Intermediate decay product of 238U
  19. Intermediate decay product of 235U
  20. Intermediate decay product of 232Th

Lead-206

206Pb is the final step in the decay chain of 238U, the "radium series" or "uranium series". In a closed system, over time, a given mass of 238U will decay in a sequence of steps culminating in 206Pb. The production of intermediate products eventually reaches an equilibrium (though this takes a long time, as the half-life of 234U is 245,500 years). Once this stabilized system is reached, the ratio of 238U to 206Pb will steadily decrease, while the ratios of the other intermediate products to each other remain constant.

Like most radioisotopes found in the radium series, 206Pb was initially named as a variation of radium, specifically radium G. It is the decay product of both 210Po (historically called radium F) by alpha decay, and the much rarer 206Tl (radium EII) by beta decay.

Lead-206 has been proposed for use in fast breeder nuclear fission reactor coolant over the use of natural lead mixture (which also includes other stable lead isotopes) as a mechanism to improve neutron economy and greatly suppress unwanted production of highly radioactive byproducts. [13]

Lead-204, -207, and -208

204Pb is entirely primordial, and is thus useful for estimating the fraction of the other lead isotopes in a given sample that are also primordial, since the relative fractions of the various primordial lead isotopes is constant everywhere. [14] Any excess lead-206, -207, and -208 is thus assumed to be radiogenic in origin, [14] allowing various uranium and thorium dating schemes to be used to estimate the age of rocks (time since their formation) based on the relative abundance of lead-204 to other isotopes. 207Pb is the end of the actinium series from 235U.

208Pb is the end of the thorium series from 232Th. While it only makes up approximately half of the composition of lead in most places on Earth, it can be found naturally enriched up to around 90% in thorium ores. [15] 208Pb is the heaviest known stable nuclide and also the heaviest known doubly magic nucleus, as Z = 82 and N = 126 correspond to closed nuclear shells. [16] As a consequence of this particularly stable configuration, its neutron capture cross section is very low (even lower than that of deuterium in the thermal spectrum), making it of interest for lead-cooled fast reactors.

Lead-212

212Pb-containing radiopharmaceuticals have been trialed as therapeutic agents for the experimental cancer treatment targeted alpha-particle therapy. [17]

Related Research Articles

<span class="mw-page-title-main">Decay product</span> The remaining nuclide left over from radioactive decay

In nuclear physics, a decay product is the remaining nuclide left over from radioactive decay. Radioactive decay often proceeds via a sequence of steps. For example, 238U decays to 234Th which decays to 234mPa which decays, and so on, to 206Pb :

<span class="mw-page-title-main">Nuclide</span> Atomic species

A nuclide is a class of atoms characterized by their number of protons, Z, their number of neutrons, N, and their nuclear energy state.

<span class="mw-page-title-main">Decay chain</span> Series of radioactive decays

In nuclear science, the decay chain refers to a series of radioactive decays of different radioactive decay products as a sequential series of transformations. It is also known as a "radioactive cascade". The typical radioisotope does not decay directly to a stable state, but rather it decays to another radioisotope. Thus there is usually a series of decays until the atom has become a stable isotope, meaning that the nucleus of the atom has reached a stable state.

Polonium-210 (210Po, Po-210, historically radium F) is an isotope of polonium. It undergoes alpha decay to stable 206Pb with a half-life of 138.376 days (about 4+12 months), the longest half-life of all naturally occurring polonium isotopes (210–218Po). First identified in 1898, and also marking the discovery of the element polonium, 210Po is generated in the decay chain of uranium-238 and radium-226. 210Po is a prominent contaminant in the environment, mostly affecting seafood and tobacco. Its extreme toxicity is attributed to intense radioactivity, mostly due to alpha particles, which easily cause radiation damage, including cancer in surrounding tissue. The specific activity of 210
Po
is 166 TBq/g, i.e., 1.66 × 1014 Bq/g. At the same time, 210Po is not readily detected by common radiation detectors, because its gamma rays have a very low energy. Therefore, 210
Po
can be considered as a quasi-pure alpha emitter.

Thorium-232 is the main naturally occurring isotope of thorium, with a relative abundance of 99.98%. It has a half life of 14 billion years, which makes it the longest-lived isotope of thorium. It decays by alpha decay to radium-228; its decay chain terminates at stable lead-208.

Uranium (92U) is a naturally occurring radioactive element that has no stable isotope. It has two primordial isotopes, uranium-238 and uranium-235, that have long half-lives and are found in appreciable quantity in the Earth's crust. The decay product uranium-234 is also found. Other isotopes such as uranium-233 have been produced in breeder reactors. In addition to isotopes found in nature or nuclear reactors, many isotopes with far shorter half-lives have been produced, ranging from 214U to 242U. The standard atomic weight of natural uranium is 238.02891(3).

Protactinium (91Pa) has no stable isotopes. The four naturally occurring isotopes allow a standard atomic weight to be given.

Thorium (90Th) has seven naturally occurring isotopes but none are stable. One isotope, 232Th, is relatively stable, with a half-life of 1.405×1010 years, considerably longer than the age of the Earth, and even slightly longer than the generally accepted age of the universe. This isotope makes up nearly all natural thorium, so thorium was considered to be mononuclidic. However, in 2013, IUPAC reclassified thorium as binuclidic, due to large amounts of 230Th in deep seawater. Thorium has a characteristic terrestrial isotopic composition and thus a standard atomic weight can be given.

Radium (88Ra) has no stable or nearly stable isotopes, and thus a standard atomic weight cannot be given. The longest lived, and most common, isotope of radium is 226Ra with a half-life of 1600 years. 226Ra occurs in the decay chain of 238U. Radium has 34 known isotopes from 201Ra to 234Ra.

Bismuth (83Bi) has 41 known isotopes, ranging from 184Bi to 224Bi. Bismuth has no stable isotopes, but does have one very long-lived isotope; thus, the standard atomic weight can be given as 208.98040(1). Although bismuth-209 is now known to be radioactive, it has classically been considered to be a stable isotope because it has a half-life of approximately 2.01×1019 years, which is more than a billion times the age of the universe. Besides 209Bi, the most stable bismuth radioisotopes are 210mBi with a half-life of 3.04 million years, 208Bi with a half-life of 368,000 years and 207Bi, with a half-life of 32.9 years, none of which occurs in nature. All other isotopes have half-lives under 1 year, most under a day. Of naturally occurring radioisotopes, the most stable is radiogenic 210Bi with a half-life of 5.012 days. 210mBi is unusual for being a nuclear isomer with a half-life multiple orders of magnitude longer than that of the ground state.

Tin (50Sn) is the element with the greatest number of stable isotopes. This is probably related to the fact that 50 is a "magic number" of protons. In addition, twenty-nine unstable tin isotopes are known, including tin-100 (100Sn) and tin-132 (132Sn), which are both "doubly magic". The longest-lived tin radioisotope is tin-126 (126Sn), with a half-life of 230,000 years. The other 28 radioisotopes have half-lives of less than a year.

Potassium has 26 known isotopes from 31
K
to 57
K
, with the exception of still-unknown 32
K
, as well as an unconfirmed report of 59
K
. Three of those isotopes occur naturally: the two stable forms 39
K
(93.3%) and 41
K
(6.7%), and a very long-lived radioisotope 40
K
(0.012%)

Neptunium (93Np) is usually considered an artificial element, although trace quantities are found in nature, so a standard atomic weight cannot be given. Like all trace or artificial elements, it has no stable isotopes. The first isotope to be synthesized and identified was 239Np in 1940, produced by bombarding 238
U
with neutrons to produce 239
U
, which then underwent beta decay to 239
Np
.

Plutonium (94Pu) is an artificial element, except for trace quantities resulting from neutron capture by uranium, and thus a standard atomic weight cannot be given. Like all artificial elements, it has no stable isotopes. It was synthesized long before being found in nature, the first isotope synthesized being plutonium-238 in 1940. Twenty plutonium radioisotopes have been characterized. The most stable are plutonium-244 with a half-life of 80.8 million years; plutonium-242 with a half-life of 373,300 years; and plutonium-239 with a half-life of 24,110 years; and plutonium-240 with a half-life of 6,560 years. This element also has eight meta states; all have half-lives of less than one second.

Uranium-236 (236U) is an isotope of uranium that is neither fissile with thermal neutrons, nor very good fertile material, but is generally considered a nuisance and long-lived radioactive waste. It is found in spent nuclear fuel and in the reprocessed uranium made from spent nuclear fuel.

Lead–lead dating is a method for dating geological samples, normally based on 'whole-rock' samples of material such as granite. For most dating requirements it has been superseded by uranium–lead dating, but in certain specialized situations it is more important than U–Pb dating.

<span class="mw-page-title-main">Primordial nuclide</span> Nuclides predating the Earths formation (found on Earth)

In geochemistry, geophysics and nuclear physics, primordial nuclides, also known as primordial isotopes, are nuclides found on Earth that have existed in their current form since before Earth was formed. Primordial nuclides were present in the interstellar medium from which the solar system was formed, and were formed in, or after, the Big Bang, by nucleosynthesis in stars and supernovae followed by mass ejection, by cosmic ray spallation, and potentially from other processes. They are the stable nuclides plus the long-lived fraction of radionuclides surviving in the primordial solar nebula through planet accretion until the present; 286 such nuclides are known.

<span class="mw-page-title-main">Radiogenic nuclide</span>

A radiogenic nuclide is a nuclide that is produced by a process of radioactive decay. It may itself be radioactive or stable.

References

  1. 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.
  2. Meija et al. 2016.
  3. "Standard Atomic Weights: Lead". CIAAW. 2020.
  4. Prohaska, Thomas; Irrgeher, Johanna; Benefield, Jacqueline; Böhlke, John K.; Chesson, Lesley A.; Coplen, Tyler B.; Ding, Tiping; Dunn, Philip J. H.; Gröning, Manfred; Holden, Norman E.; Meijer, Harro A. J. (2022-05-04). "Standard atomic weights of the elements 2021 (IUPAC Technical Report)". Pure and Applied Chemistry. doi:10.1515/pac-2019-0603. ISSN   1365-3075.
  5. Jeter, Hewitt W. (March 2000). "Determining the Ages of Recent Sediments Using Measurements of Trace Radioactivity" (PDF). Terra et Aqua (78): 21–28. Archived from the original (PDF) on March 4, 2016. Retrieved October 23, 2019.
  6. Blank, B.; Regan, P.H. (2000). "Magic and doubly-magic nuclei". Nuclear Physics News. 10 (4): 20–27. doi:10.1080/10506890109411553. S2CID   121966707.
  7. Takahashi, K; Boyd, R. N.; Mathews, G. J.; Yokoi, K. (October 1987). "Bound-state beta decay of highly ionized atoms". Physical Review C. 36 (4): 1522–1528. Bibcode:1987PhRvC..36.1522T. doi:10.1103/PhysRevC.36.1522. ISSN   0556-2813. OCLC   1639677. PMID   9954244 . Retrieved 2016-11-20. As can be seen in Table I (187Re, 210Pb, 227Ac, and 241Pu), some continuum-state decays are energetically forbidden when the atom is fully ionized. This is because the atomic binding energies liberated by ionization, i.e., the total electron binding in the neutral atom, Bn, increases with Z. If [the decay energy] Qn<Bn(Z+1)-Bn(Z), the continuum-state β decay is energetically forbidden.
  8. Half-life, decay mode, nuclear spin, and isotopic composition is sourced in:
    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.
  9. Wang, M.; Audi, G.; Kondev, F. G.; Huang, W. J.; Naimi, S.; Xu, X. (2017). "The AME2016 atomic mass evaluation (II). Tables, graphs, and references" (PDF). Chinese Physics C. 41 (3): 030003-1–030003-442. doi:10.1088/1674-1137/41/3/030003.
  10. 1 2 3 4 "Standard Atomic Weights: Lead". CIAAW. 2020.
  11. Kuhn, W. (1929). "LXVIII. Scattering of thorium C" γ-radiation by radium G and ordinary lead". The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science. 8 (52): 628. doi:10.1080/14786441108564923.
  12. 1 2 3 Beeman, J.W.; et al. (2013). "New experimental limits on the alpha decays of lead isotopes". European Physical Journal A. 49 (4): 50. arXiv: 1212.2422 . Bibcode:2013EPJA...49...50B. doi:10.1140/epja/i2013-13050-7. S2CID   254111888.
  13. Khorasanov, G. L.; Ivanov, A. P.; Blokhin, A. I. (2002). Polonium Issue in Fast Reactor Lead Coolants and One of the Ways of Its Solution. 10th International Conference on Nuclear Engineering. pp. 711–717. doi:10.1115/ICONE10-22330.
  14. 1 2 Woods, G.D. (November 2014). Lead isotope analysis: Removal of 204Hg isobaric interference from 204Pb using ICP-QQQ in MS/MS mode (PDF) (Report). Stockport, UK: Agilent Technologies.
  15. A. Yu. Smirnov; V. D. Borisevich; A. Sulaberidze (July 2012). "Evaluation of specific cost of obtainment of lead-208 isotope by gas centrifuges using various raw materials". Theoretical Foundations of Chemical Engineering. 46 (4): 373–378. doi:10.1134/S0040579512040161. S2CID   98821122.
  16. Blank, B.; Regan, P.H. (2000). "Magic and doubly-magic nuclei". Nuclear Physics News. 10 (4): 20–27. doi:10.1080/10506890109411553. S2CID   121966707.
  17. Kokov, K.V.; Egorova, B.V.; German, M.N.; Klabukov, I.D.; Krasheninnikov, M.E.; Larkin-Kondrov, A.A.; Makoveeva, K.A.; Ovchinnikov, M.V.; Sidorova, M.V.; Chuvilin, D.Y. (2022). "212Pb: Production Approaches and Targeted Therapy Applications". Pharmaceutics. 14 (1): 189. doi: 10.3390/pharmaceutics14010189 . ISSN   1999-4923. PMC   8777968 . PMID   35057083.

Sources

Isotope masses from:

Half-life, spin, and isomer data selected from the following sources.