The extended periodic table is a theoretical construct that extrapolates the organization of chemical elements beyond the currently confirmed limit at atomic number 118 (oganesson), predicting the electronic configurations, chemical properties, and periodic trends for superheavy elements up to atomic numbers as high as 172 or more, guided by relativistic quantum chemistry and nuclear physics models.[1]This concept emerged prominently in the late 20th century amid advances in nuclear synthesis, with Nobel laureate Glenn T. Seaborg proposing in 1969 a major reorganization to accommodate undiscovered superheavy elements from Z=110 to Z=173, introducing the transactinide series (Z=104–121, analogous to transition metals) and the superactinide series (Z=122–153, featuring a new g-block with 5g orbitals).[2] Seaborg's model, detailed in publications like his Scientific American article, emphasized an "island of stability" around Z≈114 and N≈184, where enhanced nuclear shell effects could yield longer-lived isotopes with half-lives potentially exceeding 10^8 years, enabling chemical study. His framework also anticipated further extensions, including a 9th period up to Z=173, though these predictions relied on non-relativistic approximations that later required refinement.[2]Modern extensions incorporate Dirac–Fock calculations accounting for relativistic effects, as in Pekka Pyykkö's 2011 study, which proposed a shell-filling sequence for Z=119–172 as 8s < 5g ≤ 8p_{1/2} < 6f < 7d < 9s < 9p_{1/2} < 8p_{3/2}, placing elements 121–138 in an 8th period g-block and 139–164 in a 9th period, with high oxidation states (up to +6 or more) predicted for early superactinides due to contracted orbitals.[1] These models suggest that superheavy elements would exhibit volatile chemistry, with trends diverging from lighter homologues—such as inertness in group 18 analogs due to relativistic stabilization of s-orbitals—while nuclear instability limits most isotopes to microseconds, except near magic numbers like N=184.[1]As of November 2025, the extended periodic table remains unverified experimentally beyond Z=118, but ongoing efforts at facilities like GSI Helmholtz Centre and Lawrence Berkeley National Laboratory use advanced ion beams (e.g., titanium-50 on berkelium-249 targets for element 119 and on californium-249 for element 120) to pursue elements 119 and 120, with recent breakthroughs in superheavy molecule synthesis hinting at feasible chemical exploration if stability improves.[3][4] These pursuits aim to test theoretical predictions and potentially confirm the island of stability, reshaping our understanding of matter's limits.[5]
Historical Development
Early Predictions and Theoretical Foundations
The periodic law formulated by Dmitri Mendeleev in 1869 provided the foundational framework for the periodic table, enabling predictions of undiscovered elements based on trends in atomic weights and chemical properties. While Mendeleev's predictions in the late 19th century focused on elements up to uranium (Z=92), such as anticipated rare earths and gaps in the table, his concept of periodicity laid the groundwork for later extensions to heavier elements, including transuranics, in the 20th century.Following World War II, Glenn T. Seaborg advanced these ideas with his actinide hypothesis in 1944, proposing that elements beyond uranium (Z=92) form a second inner transition series, the actinides (Z=89–103), homologous to the lanthanides due to the filling of 5f orbitals. This concept, validated by the synthesis of neptunium (Z=93) and plutonium (Z=94), extended the periodic table by recognizing a new f-block series and predicted chemical similarities among these heavy elements, such as a stable +3 oxidation state. Seaborg's work shifted the understanding of heavy element chemistry from d-block analogies to f-block behavior, enabling systematic predictions for further extensions. In the late 1950s and early 1960s, Seaborg began speculating on superheavy elements beyond the actinides, suggesting a transactinide series (Z=104–121) that would continue the pattern of increasing period lengths.The nuclear shell model, independently developed by Maria Goeppert Mayer and J. Hans D. Jensen in 1949, provided a theoretical basis for predicting nuclear stability in superheavy elements. This model treats nucleons as independent particles in a mean-field potential, analogous to atomic electron shells, and explains enhanced stability at "magic numbers" of protons and neutrons (e.g., 2, 8, 20, 28, 50, 82, 126) through closed shells. Mayer and Jensen's framework, detailed in their 1955 book Elementary Theory of Nuclear Shell Structure, forecasted that superheavy nuclei near doubly magic configurations, such as Z=114 and N=184, could exhibit longer half-lives due to shell effects counteracting fission and alpha decay. This nuclear perspective complemented chemical periodicity, highlighting potential "islands of stability" for elements far beyond known actinides. Relativistic effects in heavy atoms, arising from high nuclear charge speeding inner electrons near light speed, were noted to influence orbital contractions but were secondary to shell stability in early predictions.By the 1960s, these foundations converged in explicit proposals for an extended periodic table, including an eighth period with estimated lengths following the aufbau sequence: 2 elements in the s-block (Z=119–120), 18 elements in the g-block (Z=121–138, part of the superactinide series filling 5g orbitals), and further extensions including 6f and 7d orbitals. Seaborg's 1969 proposal outlined this structure, predicting a 32-element superactinide series (Z=122–153) filling 5g and 6f orbitals, thus doubling the table's span and accommodating up to Z=172 or beyond based on quantum mechanical orbital capacities. These early models emphasized conceptual continuity rather than precise properties, setting the stage for later refinements.
Evolution of Extended Models in the 20th and 21st Centuries
In the early 1970s, theoretical models for extending the periodic table began incorporating relativistic effects to predict the structure beyond the actinides. A seminal contribution came from B. Fricke, W. Greiner, and J. T. Waber, who performed relativistic Dirac-Fock calculations on neutral atoms up to Z=172, predicting an eighth period of 32 elements due to the filling of 5g, 6f, and 7d orbitals, influenced by strong relativistic stabilization of inner s and p orbitals.[6] This model highlighted the role of Dirac-Fock methods in accounting for spin-orbit coupling and electron-nucleus interactions in superheavy atoms, marking a shift from non-relativistic approximations.[7]During the 1970s and 1980s, alternative proposals emerged based on detailed orbital filling sequences. V. I. Nefedov and collaborators developed a model using relativistic Dirac–Fock calculations for superheavy elements, suggesting an extended table with eighth, ninth, and tenth periods of 18, 32, and 32 elements, respectively, where the 5g orbitals fill from Z=125 to 144, followed by 6f from Z=145 to 157, and 7d from Z=158 onward.[8] This structure contrasted with Fricke's by incorporating configuration interactions, though later refinements acknowledged partial relativistic influences on orbital energies.[9]By the early 21st century, refinements addressed ongoing debates about period lengths and block assignments, particularly for the superactinide series (Z=121–155). Andrey Kulsha proposed variations building on Nefedov's framework, advocating for a 32-element superactinide block from Z=121 to 152 in some designs, with adjustments to accommodate relativistic effects and maintain group periodicity up to Z=172.[10] Similarly, computational studies by groups including D. F. Smits and Peter Schwerdtfeger explored period lengths through multiconfiguration Dirac-Fock calculations, debating whether the superactinide block spans 32 or 34 elements based on g-orbital contraction and f-block analogies, influencing predictions for chemical homology in elements beyond Z=120.[11]Advancements in the 21st century have been driven by improved computational methods, notably density functional theory (DFT) tailored for relativistic superheavy systems. Relativistic DFT, often using four-component Hamiltonians, has enabled accurate predictions of electronic structures for elements up to Z=126 and beyond, revealing trends in ionization potentials and bonding that refine earlier models; for instance, calculations confirm the stability of 7p_{1/2} orbitals in oganesson (Z=118) and predict similar behavior extending the table.[12] These methods, as reviewed in works by Schwerdtfeger and Pershina, integrate electron correlation and Breit interactions, providing a robust framework for debating the table's endpoint around Z=172 while highlighting divergences from non-relativistic predictions.[13] In 2023, Smits, Düllmann, Indelicato, Nazarewicz, and Schwerdtfeger proposed a revised periodic table placement for elements 119–170, using advanced relativistic coupled-cluster and density functional theory to refine superactinide block assignments and period lengths.[14]
Theoretical Frameworks
Modifications to the Aufbau Principle
The standard Aufbau principle, which orders electron filling based on the n + l rule to minimize energy, begins to fail for elements with atomic numbers Z > 100 due to increasingly dominant relativistic effects that alter orbital energies and configurations.[15] These effects arise from electrons in heavy atoms approaching significant fractions of the speed of light, particularly inner-shell electrons, leading to deviations from non-relativistic predictions and irregularities in the expected filling sequence.[16]Relativistic effects manifest prominently in the actinides and superheavy elements through the stabilization of 7s and 7p orbitals alongside the destabilization of 6d orbitals. The increased effective mass of electrons near the nucleus contracts s and p_{1/2} orbitals, lowering their energies, while d and f orbitals expand due to reduced screening and angular momentum barriers, raising their energies relative to non-relativistic cases.[17] This inversion contributes to ground-state configurations like 6d^q 7s^2 for 6d transition metals in the seventh period, contrasting with the 5d^9 6s^1 and 5d^{10} 6s^1 patterns in lighter analogs such as platinum and gold.[12]To accurately model these phenomena, the Dirac equation is applied, particularly for inner-shell electrons in heavy atoms where velocities approach relativistic speeds. The Dirac-Coulomb Hamiltonian incorporates spin-orbit coupling and relativistic kinematics, yielding binding energies that match experimental data for inner shells far better than Schrödinger-based methods.[18] These inner-shell adjustments propagate to valence shells, influencing overall electron configurations and necessitating fully relativistic treatments for Z > 80.[19]For superheavy elements, relativistic effects predict the onset of g-orbital filling starting at Z = 121, marking the beginning of a superactinide series with 5g orbitals participating in the ground state. Calculations indicate configurations involving 8s^2 5g^1 or similar, driven by near-degeneracies among 5g, 6f, and 7d subshells.[11] Relativistic corrections, including orbital contraction and stabilization of s and p_{1/2} subshells, contribute to these shifts.[18]These modifications impact the periodicity of the extended table by causing contraction of s and p shells, which reduces atomic radii more sharply than in lighter elements and leads to shorter periods in the eighth and subsequent rows.[15] This contraction enhances effective nuclear charge for valence electrons, altering chemical trends and blurring traditional block distinctions beyond Z = 120.[11]
Key Predictive Models (Fricke, Nefedov, Pyykkö, and Others)
One of the early key predictive models for the extended periodic table was proposed by B. Fricke, W. Greiner, and J. T. Waber in 1971, utilizing relativistic Hartree-Fock-Slater calculations to account for the extended nuclear charge distribution in superheavy atoms. This model predicted a distinctive structure for the extended periods starting from the 8th period (Z=119), with relativistic effects leading to a compact arrangement up to an endpoint at Z=172. The approach emphasized altered orbital energies, resulting in shortened blocks such as a 14-element 7d series in the 8th period dominated by 5g and 7d orbitals.[6]In the 1990s and 2000s, V. I. Nefedov and collaborators developed a model based on relativistic group theory, which incorporated spin-orbit splitting and symmetry considerations to predict structures for the extended periods, with the 8th period spanning approximately 50 elements (combining early 18-element and late 32-element subshell fillings) and a 50-element 9th period, extending the table up to Z=168. This framework highlighted larger block sizes in later periods due to the filling of high-angular-momentum orbitals like 6f and 5g, prioritizing group-theoretical degeneracies over simple energy ordering.[9]Pekka Pyykkö's models from the 2000s and 2010s employed pseudopotential and Dirac-Fock methods to refine superheavy electron configurations, proposing an endpoint at Z=172. These calculations demonstrated that relativistic effects cause significant orbital contractions, with the 8th period (Z=119–138) featuring the 8s and 5g blocks (20 elements total, including an 18-element g-block at Z=121–138), the 9th period (Z=139–164) incorporating 6f and 7d (26 elements), and a short 10th period segment up to Z=172. Pyykkö's work built on prior models by incorporating multi-configuration interactions for more accurate shell fillings.[1]Other models, such as those by Andrey Kulsha in the 2000s, introduced variations by refining block sizes for the superactinide region, proposing a 36-element ultransition block (elements 121–156) split into two 18-element series to better accommodate d- and f-orbital intrusions. Recent reviews in the 2020s, including contributions from researchers like Odile R. Smits, explore orbital intrusions and configuration mixing in superheavy atoms using advanced relativistic theories, emphasizing the sensitivity of predictions to computational treatment of electron correlation and nuclear models, with ongoing refinements as of 2025.[20]
Model
Period 8 Length (Z=119+)
Period 9 Length
Period 10 Length
Endpoint
Fricke et al. (1971)
46 (to Z=164)
8
-
Z=172
Nefedov et al. (1990s–2000s)
50
50
-
Z=168
Pyykkö (2000s–2010s)
20 (8s + 5g)
26 (6f + 7d + 9s + 9p_{1/2})
8 (9p_{1/2} + 8p_{3/2})
Z=172
Kulsha (2000s) & recent variations (2020s)
Variable (e.g., 50+ with 36 ultransition)
Variable
Variable
Z=172
Structure of the Extended Table
Eighth-Period Elements and Period Lengths
The eighth period in the extended periodic table is theorized to comprise 50 elements, spanning atomic numbers Z=119 to Z=168, extending the structure beyond the confirmed seventh period that concludes with oganesson (Z=118). This configuration arises from predictions based on relativistic atomic calculations, where the aufbau principle is modified to account for the filling of higher angular momentum orbitals in superheavy atoms. Most theoretical models converge on this 50-element length, reflecting the anticipated completion of the 8s, 5g, 8p_{1/2}, 6f, 7d, and 8p_{3/2} subshells.[21]The period initiates at Z=119 with ununennium, marking the transition from the seventh to the eighth period and initiating the s-block with two elements (Z=119–120) analogous to the alkali and alkaline earth metals. Following this, the g-block spans Z=121–138 (18 elements), filling the 5g orbitals and introducing the superactinide series. A brief p-block insertion follows at Z=139–140 (8p_{1/2}), before the f-block superactinides at Z=141–154 (14 elements), filling the 6f orbitals. The d-block then occupies Z=155–164 (10 elements), representing the filling of 7d orbitals. The period concludes with the 8p_{3/2} subshell at Z=165–168 (4 elements) in the p-block. Relativistic effects, such as spin-orbit splitting of p-orbitals into 1/2 and 3/2 components, cause this irregular structure, with the 8p_{3/2} filling after the 7d due to energy level shifts.[21]Debates on period length persist, with earlier models favoring 32 elements based on a more conservative orbital filling sequence that omitted or shortened the g-block, while relativistic computations support the longer 50-element structure to accommodate the full superactinide and g-block sequences. Relativistic effects, such as orbital contraction and stabilization, play a key role in these placements by altering energy levels for high-Z atoms.[22]Visual representations of the extended table, such as those derived from Dirac-Fock calculations, illustrate the eighth period's blocks integrated into a widened format up to Z=172, highlighting the g-block's position and the overall 50-column layout for periods 8 and 9 to maintain periodicity. These diagrams emphasize the table's expansion while preserving group homologies across rows. The ninth period is anticipated to begin at Z=169 with the 9s orbital.
Superactinides and Elements Beyond Z=120
The superactinides represent a proposed series of hypothetical elements in the extended periodic table, extending beyond the actinides and beginning immediately after elements 119 and 120 in the eighth period. These elements, spanning atomic numbers Z=121 to 154 in Pyykkö's model, are characterized as hybrids involving the filling of 5g, 6f, and 7d orbitals, marking a significant departure from the standard s, p, d, and f blocks. This series arises from the anticipated involvement of higher angular momentum orbitals, particularly the g-block (l=4), which introduces 18 electrons per subshell and extends the f-block analogy into uncharted territory. Theoretical predictions emphasize that the superactinide block would feature an elongated structure, with the 5g subshell filling first, followed by contributions from 6f and 7d, leading to complex electronic configurations influenced by spin-orbit splitting.[21]In Pekka Pyykkö's Dirac-Fock-based model, the superactinides are divided into distinct subshells: a 5g series comprising 18 elements (Z=121–138), a brief 8p_{1/2} insertion (Z=139–140), and a 6f series of 14 elements (Z=141–154), with the 7d orbitals playing a transitional role thereafter (Z=155–164). This configuration predicts a total of 34 superactinides up to Z=154 (18+2+14), though Pyykkö emphasizes the 14-element 6f segment as a core feature analogous to the lanthanides and actinides. In contrast, Viktor Nefedov's model, derived from relativistic Dirac-Fock-Slater calculations, proposes a more unified 34-element superactinide block (Z=122–155), where the 5g (18 electrons) and 6f (14 electrons) subshells fill concurrently, completing at Z=155 before transitioning to further d-block elements. These differences highlight varying interpretations of orbital stability and filling order in superheavy regimes.[16]The predicted block structure for superactinides extends the f-block into a g-involved domain, creating a ~34-element row that disrupts traditional periodicity and potentially requires a wider table format with up to 50 or 54 columns. Relativistic effects, amplified by high nuclear charge, cause severe distortions in these elements, including enhanced s- and p-orbital contraction, destabilization of d and f orbitals, and inverted filling sequences compared to lighter elements. Unlike the actinides, where 5f electrons contribute to variable oxidation states and metallic bonding, superactinides are expected to exhibit non-standard chemistry, such as dominant 8s^2 inert pairs, reduced covalency, and possible volatile or noble-gas-like behaviors due to extreme spin-orbit coupling in the 5g and 6f shells. These distortions could render superactinide compounds highly unstable or exhibit unprecedented coordination geometries, challenging conventional group trends.[16]Beyond the superactinides, elements Z=155–164 fill the 7d block, followed by the 8p_{3/2} subshell (Z=165–168), completing the eighth period. In Pyykkö's extended table, this aligns with a total eighth-period length of 50 elements, where relativistic stabilization of p orbitals allows closure before initiating 9s and higher shells in the ninth period at Z=169. Nefedov's framework similarly positions post-superactinide elements as extensions emphasizing cumulative relativistic contraction. This structure underscores the tentative nature of periodicity at such high Z, where g-orbital intrusion blurs block boundaries and foreshadows even greater deviations in hypothetical ninth-period elements.
Experimental Pursuits
Synthesis Techniques and Facilities
The synthesis of superheavy elements (SHEs) beyond uranium primarily relies on fusion-evaporation reactions, where heavy-ion beams are accelerated to collide with target nuclei, forming a compound nucleus that subsequently evaporates neutrons to produce more stable isotopes. These reactions occur in heavy-ion accelerators, which provide the high energies needed—typically 5-10 MeV per nucleon—to overcome the Coulomb barrier between the projectile and target.[23] Two main approaches have evolved: cold fusion, using neutron-deficient projectiles like calcium-48 on lead or bismuth targets to produce elements up to atomic number Z=112 with fewer neutrons evaporated, and hotfusion, employing more neutron-rich beams such as titanium-50 or vanadium-51 on actinide targets (e.g., plutonium or curium) to access higher Z elements (113-118) with increased neutron evaporation for potential stability.[24] The transition to hotfusion has been crucial for extending the periodic table, as it allows synthesis of isotopes closer to predicted nuclear shell closures.[25]Key facilities worldwide host these experiments, leveraging specialized accelerators and detection systems. The Joint Institute for Nuclear Research (JINR) in Dubna, Russia, operates the Flerov Laboratory's Superheavy Element Factory, featuring the DC280 cyclotron for intense heavy-ion beams up to titanium-54, which has enabled production of elements 113-118 through hot fusion reactions like ^{48}Ca + ^{249}Cf.[26] The GSI Helmholtz Centre in Darmstadt, Germany, uses the UNILAC linear accelerator and SIS synchrotron paired with the SHIP (Separator for Heavy Ion Reaction Products) gas-filled recoil separator, pioneering cold fusion for elements 107-112 and contributing to hot fusion studies.[27]RIKEN's Nishina Center in Japan employs the RIKEN Linear Accelerator (RILAC) and coupled cyclotrons with the GARIS (Gas-filled Recoil Ion Separator) for both cold and hot fusion, notably synthesizing element 113 via ^{70}Zn + ^{209}Bi.[28] In the United States, Lawrence Berkeley National Laboratory's 88-Inch Cyclotron restarted SHE research in 2024, successfully producing livermorium (Z=116) using a novel titanium-50 beam on plutonium targets, signaling renewed efforts to pursue elements beyond Z=118.[29]Beam-target combinations are selected to maximize fusion probability while minimizing fission competition, with cross-sections for Z>118 typically on the order of 1 picobarn (pb) or lower, reflecting the exponential decrease in yield with increasing Z. For instance, proposed reactions for Z=119 include ^{50}Ti + ^{249}Bk or ^{51}V + ^{248}Cm, building on the hot fusionparadigm to incorporate more neutrons.[23] Detection relies on gas-filled separators that exploit the kinematic differences between evaporation residues and scattered beam particles, implanting residues into silicon detectors for observation of correlated alpha decay chains, often analyzed via digital signal processing to distinguish rare events from background.[30]Major challenges include extremely low production rates—often one atom per month or less for the heaviest elements—and half-lives under 1 second, necessitating continuous beam delivery over extended periods and ultra-sensitive instrumentation.[23] These limitations test the limits of current accelerators but are motivated by predictions of enhanced stability in the "island of stability" around Z=114-126.[25]
Recent Attempts for Elements 119–127 (2023–2025 Updates)
Recent experimental efforts to synthesize element 119 have focused on fusion reactions at major facilities, including the Joint Institute for Nuclear Research (JINR) and RIKEN. At JINR, attempts using the ^{50}Ti + ^{249}Bk reaction have been conducted from 2023 through 2025, but no confirmation of the element's production has been reported.[31] Cross-section estimates for this reaction remain below 1 femtobarn, indicating significant challenges in detection.[31] Similarly, RIKEN's ongoing experiments since 2018, involving bombardment of ^{248}Cm targets with ^{51}V beams, have not yielded published confirmations as of mid-2025.[32] These efforts leverage upgraded accelerators to increase beam intensity, yet the low production rates persist due to fission barriers in the compound nuclei.[33]For element 120, planned experiments at the Joint Institute for Nuclear Research (JINR) target chromium beams on curium targets.[3] In the United States, proposals from Lawrence Berkeley National Laboratory and other institutions have advocated for the ^{48}Ca + ^{249}Cf reaction as a viable pathway, building on prior actinide target preparations.[29] A notable advancement occurred with the proposal of a ^{40}Ar fusion method using ^{40}Ar + ^{249}Bk, aimed at producing isotopes in the decay chain toward element 119 more cost-effectively than traditional calcium-based beams.[34] This approach addresses limitations in beam stability and target availability, potentially enabling higher-yield runs in upcoming cycles.[34]Attempts to produce elements 121 through 127 remain sparse and largely preparatory, with emphasis on vanadium and titanium ion beams directed at superactinide targets such as curium or berkelium isotopes.[35] These methods prioritize reactions like ^{51}V + ^{249}Bk for Z=121, though experimental data is limited to feasibility studies without confirmed events.[35] Broader challenges include the need for enhanced detection systems to handle even shorter decay chains.[36]A significant milestone in 2024 was the successful synthesis of two livermorium (element 116) atoms using a titanium-50 beam on a plutonium-244 target at Lawrence Berkeley National Laboratory, marking the first use of this fusion route for superheavies.[37] This breakthrough demonstrates the viability of titanium beams for overcoming neutron deficits in heavier systems, directly informing strategies for elements 119 and beyond by validating higher cross-sections around 10 picobarns.[37] The approach reduces reliance on rare calcium-48 isotopes and opens pathways for scaled-up production at facilities like the Facility for Rare Isotope Beams.[29]Despite these advances, no syntheses of elements 119–127 have been confirmed as of November 2025, with all attempts yielding null results in terms of observable decay signatures. Predicted half-lives for potential isotopes range from 10^{-10} seconds to 1 second, underscoring the transient nature of these nuclei and the need for millisecond-scale detection.[38]
Predicted Properties
Electronic Configurations and Relativistic Effects
The predicted ground-state electronic configurations for elements 119 and 120, the first members of the eighth period, are [Og] 8s¹ and [Og] 8s², respectively, marking a continuation of the s-block filling similar to alkali and alkaline earth metals, though modified by relativistic effects.[9][39] These configurations arise from the relativistic stabilization of the 8s orbital, which lowers its energy relative to non-relativistic expectations, while the preceding 7p subshell exhibits strong spin-orbit splitting, with the 7p_{1/2} level filled and more contracted than the 7p_{3/2}, influencing the overall valence structure.[11][6]For superactinides beginning at Z=121, the configurations involve the filling of the 5g subshell alongside 8s and 8p_{1/2} orbitals, such as [Og] 8s² 8p¹ for element 121 (with 5g filling starting around Z=125 in some models), transitioning into a g-block series up to Z=138.[39][40] Relativistic effects play a critical role here, causing destabilization and spatial expansion of the d and f orbitals through indirect mechanisms, where the contraction of inner s and p_{1/2} orbitals increases nuclear screening, thereby elevating the energies of higher angular momentum orbitals like 6f, 7d, and 5g relative to s/p states. Note that exact configurations vary by model due to near-degeneracies and electron correlation.[41][42] This destabilization alters the expected Aufbau order, promoting mixing between configurations such as 5g, 6f, 7d, and 8p in the superactinide region.Predictions of these configurations rely heavily on multi-configuration Dirac-Fock (MCDF) methods, which incorporate relativistic kinematics and account for electron correlation through variational optimization of multi-electron wave functions.[11] In the Dirac-Fock framework for hydrogen-like atoms, the energy levels are approximated by the Dirac equation solution:E_{n\kappa} = mc^2 \left[ 1 + \left( \frac{Z \alpha}{n - (j + 1/2) + \sqrt{(j + 1/2)^2 - (Z \alpha)^2}} \right)^2 \right]^{-1/2}where m is the electron mass, c the speed of light, Z the atomic number, \alpha the fine-structure constant, n the principal quantum number, and \kappa related to the total angular momentum j; for many-electron systems, MCDF extends this by solving self-consistent field equations.[43][44]There are notable variations across theoretical models regarding the onset of the g-block. Early relativistic Hartree-Fock-Slater calculations by Fricke, Greiner, and Waber placed the 5g filling starting at Z=121 with a distinct g-block up to Z=138, emphasizing period lengths of 32 for the 5g series.[6] In contrast, more recent Dirac-Fock work by Pyykkö refines this by proposing a shell-filling sequence of 8s < 5g ≤ 8p_{1/2} < 6f < 7d, with the g-block formally assigned to Z=121–164 but incorporating greater orbital mixing due to near-degeneracies, leading to potentially different ground states for several elements compared to Fricke's model.[39][9]Relativistic effects also drive lanthanide-like contractions in the superheavy region, where successive filling of the 5g and 6f orbitals causes poor shielding and radial contraction of outer s and p_{1/2} orbitals, resulting in steadily increasing first ionization potentials up to Z=172.[45] MCDF calculations predict these ionization potentials to rise from approximately 4–5 eV for Z=119 to over 10 eV near Z=172, enhancing chemical inertness in later superactinides akin to noble gas behavior.[11][45]
Chemical and Physical Characteristics by Group
Elements 119 and 120, positioned as the initial members of the eighth period in the s-block, are predicted to exhibit metallic characteristics analogous to the alkali and alkaline earth metals, respectively, but with significantly reduced reactivity due to relativistic stabilization of the 8s orbital.[46] The relativistic s-contraction tightens the 8s electrons closer to the nucleus, increasing their binding energy and hindering ionization, which contrasts with the highly reactive nature of lighter homologs like francium (Z=87) and radium (Z=88).[1] This effect is expected to render these elements more inert, potentially forming stable +1 and +2 oxidation states only under extreme conditions, while their predicted volatility arises from weakened metallic bonding influenced by the contracted orbitals.[47]The superactinides, encompassing elements from Z=121 to Z=155, represent a complex series bridging the d-, f-, and g-blocks, where relativistic effects dominate, leading to volatile metallic behaviors distinct from the seventh-period actinides. These elements are forecasted to support high oxidation states ranging from +4 to +8, driven by the availability of g-orbitals that facilitate multiple bonding interactions, analogous to transition metals but with enhanced cluster formation tendencies due to compact atomic sizes.[45] Unlike the more stable, less volatile actinides, superactinides are predicted to form molecular clusters and exhibit semiconductor-like properties in their compounds, with reactivity modulated by strong spin-orbit coupling that alters orbital hybridization.[1] Theoretical models indicate that their chemistry will prioritize covalent bonding over ionic, resulting in volatile oxides and fluorides that evaporate readily at moderate temperatures.[48]In the p-block extension, elements in the range around Z=163 to Z=168 (corresponding to 8p filling in modern models) are anticipated to include analogs to the later main-group elements, with group 17 (halogen-like) showing reduced electronegativity and oxidizing power compared to iodine or astatine, owing to relativistic expansion of p_{3/2} orbitals that weakens bonding, potentially limiting them to +1 or +3 states rather than the typical -1, and group 18 (noble gas-like) expected to be less inert than xenon or radon, with possible compound formation under high pressure due to softened valence shells.[1] These trends mark a departure from seventh-period p-block elements, where relativistic effects already noble-ize chemistry, but intensify further, making superheavy p-block species more gas-like and less interactive. Exact Z assignments vary by model.[45][15]Physically, the extended elements across these groups share compact atomic radii around 100 pm, a consequence of ongoing lanthanide- and actinide-like contractions amplified by relativity, leading to high electron densities near the nucleus.[49] Densities are projected to exceed 20 g/cm³, surpassing even osmium (22.6 g/cm³), due to the heavy nuclear cores packed into small volumes, which enhances interatomic attraction but promotes brittleness in solid phases.[48] A 2025 technique developed at Lawrence Berkeley National Laboratory for identifying heavy element molecules, demonstrated on nobelium (Z=102), employs gas-phase separation and holds promise for probing volatility in superheavy elements like those beyond Z=118 using fluorine-containing gases, aligning with predicted behaviors.[4] Overall, relativistic influences differentiate these elements from their seventh-period counterparts by enhancing nobility and volatility, shifting chemistry toward more discrete, molecular species rather than extended solids.[16]
Nuclear Aspects and Limits
Magic Numbers and Island of Stability
In nuclear physics, magic numbers denote specific counts of protons (Z) or neutrons (N) that fill complete nuclear shells, conferring exceptional stability to the nucleus due to maximized binding energy. For superheavy elements (SHE), extending beyond Z ≈ 100, theoretical calculations predict magic proton numbers at Z = 114, 120, and 126, alongside a prominent magic neutron number at N = 184, which together delineate potential regions of enhanced nuclear stability.[50][51] These predictions arise from extensions of the nuclear shell model to the relativistic regime of heavy nuclei, where shell closures mitigate fission barriers and decay probabilities.The concept of the island of stability refers to a hypothetical cluster of SHE isotopes in the nuclear chart, centered around Z = 114–126 and N = 184, where closed shells are expected to yield dramatically prolonged half-lives relative to typical superheavy isotopes. Whereas observed SHE often decay in microseconds to milliseconds via alpha emission or spontaneous fission, doubly magic or near-doubly magic isotopes in this island could persist for seconds, minutes, days, or potentially years, enabling detailed study of their properties.[52][51] This enhanced longevity stems from higher fission barriers and reduced decay widths near shell closures, contrasting with the rapid instability of off-shell neighbors.The foundational theory underpinning these magic numbers employs the Woods-Saxon potential—a phenomenological form modeling the nuclear mean field—with inclusion of spin-orbit coupling to reproduce observed shell structures in heavy nuclei. This approach, pioneered in calculations for Z > 82 and N > 126, generates single-particle levels that align with experimental magic numbers and predicts analogous closures at higher Z and N for SHE.[53] For the deformed shapes common in superheavy nuclei, the Nilsson model adapts the shell model by incorporating quadrupole deformation into the harmonic oscillator basis plus spin-orbit terms, revealing deformed shell gaps that contribute to stability even in non-spherical configurations.[54]Such shell-stabilized nuclei hold implications for experimental synthesis, as targeting doubly magic candidates like ^{298}\mathrm{Fl} (Z = 114, N = 184) or ^{304}_{120} (Z = 120, N = 184) could yield isotopes with half-lives orders of magnitude longer than current SHE productions, facilitating chemical and spectroscopic investigations.[55] Recent models from the 2020s, incorporating advanced density functional theories and alpha-decay systematics, indicate a refinement of the island toward Z ≈ 120, driven by greater alpha-decay hindrance from elevated proton shell gaps that suppress tunneling probabilities.[56][57]
Relativistic Instability Beyond Z=137 and Table Endpoint
As atomic numbers increase beyond Z=137, relativistic effects in the Dirac equation lead to instability in the electronic structure, where the velocity of the 1s electron approaches the speed of light (c), resulting in pair production and the electron diving into negative energy states as described by the Klein paradox.[42] This barrier arises because the binding energy exceeds 2mc² (approximately 1.022 MeV), allowing electron-positron pair creation from the vacuum, which destabilizes the atom and prevents stable bound states.[58]In the Pyykkö model, based on Dirac-Fock calculations for neutral atoms and ions, the periodic table extends up to Z=172, beyond which no stable electronic configurations with bound states are predicted due to the continued diving of inner-shell orbitals into the negative continuum.[1] This endpoint, around Z=172–173, marks the point where quantum electrodynamic corrections fail to restore binding, effectively limiting the table's extension.[58]Nuclear instabilities further constrain the table for Z>120, as fission barriers decrease significantly in the absence of nearby magic numbers, leading to high Q-values for spontaneous fission and half-lives on the order of microseconds or less.[59] Calculations show that without shell closures like N=184, these barriers drop below 5–6 MeV for Z>126, promoting rapid fission decay even for neutron-rich isotopes.[59]Beyond the primary island of stability, elements with Z>173 remain highly unstable, even at neutron numbers like N=184, due to combined relativistic atomic effects and low fission barriers, though small secondary islands may exist around Z≈174 and N≈410 with marginal enhancements in half-life.[60] Such hyperheavy nuclei could potentially form exotic states, but they do not support periodic chemical behavior owing to their extreme instability.[60]Debates persist on the exact endpoint, with some nuclear models suggesting possibilities up to Z=200+ using exotic electron configurations to mitigate relativistic diving, but the prevailing consensus from atomic structure theory supports Z=172 as the maximum for viable elements.[61][1]