The periodic table is a tabular chart that arranges all known chemical elements in order of increasing atomic number, with rows representing periods and columns representing groups that exhibit recurring chemical and physical properties.[1] This organization, which currently includes 118 confirmed elements, allows scientists to predict element behaviors, identify trends such as atomic radius, electronegativity, and ionization energy, and understand the fundamental structure of matter.[1][2]The development of the periodic table began in the 19th century amid efforts to classify elements based on atomic weights and properties.[3] Early contributions included Antoine Lavoisier's 1789 classification of elements into categories like gases, non-metals, and metals, and Johann Döbereiner's 1829 identification of element triads with similar characteristics, such as lithium, sodium, and potassium.[3] The modern periodic table is primarily credited to Dmitri Mendeleev, who published his first version in 1869, arranging the 63 known elements by atomic weight while leaving gaps for undiscovered ones and predicting their properties, which were later verified.[1][3] This breakthrough was refined over time, notably by Henry Moseley in 1913, who reordered elements by atomic number using X-ray spectroscopy, providing the table's current foundation.[3]As a cornerstone of chemistry, the periodic table facilitates research in fields from materials science to nuclear physics and is periodically updated by the International Union of Pure and Applied Chemistry (IUPAC) to incorporate new elements and refined atomic weights.[2] For instance, elements 113, 115, 117, and 118 were officially named in 2016, completing the seventh period.[2] The table's blocks—s, p, d, and f—correspond to electron configurations in atomic orbitals, explaining why elements in the same group share valence electron similarities and reactivity patterns.[1] Today, it remains an indispensable tool for education and innovation, symbolizing the periodicity of the universe's building blocks.[1]
Overview
Definition and organization
The periodic table is a tabular arrangement of the chemical elements, organized by increasing atomic number, which represents the number of protons in each element's atomic nucleus.[2] As of 2025, it encompasses 118 confirmed elements, ranging from hydrogen with atomic number 1 to oganesson with atomic number 118.[4]The structure consists of horizontal rows known as periods, which correspond to the principal quantum numbers or energy levels occupied by the outermost electrons in the atoms of those elements. Vertical columns, referred to as groups, contain elements that possess similar configurations of valence electrons in their outer shells, leading to comparable chemical reactivity and bonding behaviors. This grid-like organization, typically presented in a long form with 7 periods and 18 groups, visually captures the systematic progression of atomic properties across the table.[5]The periodic table serves to systematize the known elements, enabling chemists to predict the characteristics of yet-to-be-discovered elements and to elucidate patterns of chemical similarities and variations. Central to its design is the concept of periodicity, which describes the recurring trends in physical and chemical properties—such as atomic size, ionization energy, and electronegativity—that emerge due to the periodic repetition of similar electron configurations as atomic number increases.[2]
Periods and groups
The periodic table is divided into horizontal rows known as periods and vertical columns known as groups, which organize the elements based on their atomic structure and chemical properties. Periods are numbered from 1 to 7, corresponding to the principal quantum number (n) of the outermost electron shell, ranging from n=1 in period 1 to n=7 in period 7. As the principal quantum number increases, additional subshells (s, p, d, f) become available for electron filling, resulting in a progressive increase in the number of elements per period: period 1 contains 2 elements, periods 2 and 3 have 8 each, periods 4 and 5 have 18 each, and periods 6 and 7 have 32 each when including the f-block elements.[6][7]Groups, also called families, are the vertical columns numbered 1 through 18 according to the standard IUPAC system, which encompasses main groups (1-2 and 13-18), transition metals (3-12), and accommodates inner transition metals. Elements within the same group exhibit similar chemical behaviors because they share comparable outermost electron configurations, leading to analogous valence electron counts and reactivity patterns.[2][8][6]To maintain a compact layout and prevent excessive horizontal width, the lanthanides (elements 58–71, from cerium to lutetium, commonly including lanthanum) and actinides (elements 90–103, from thorium to lawrencium, commonly including actinium) are placed in separate rows below the main body of the table, despite belonging to periods 6 and 7, respectively. This arrangement highlights their role in filling the 4f and 5f subshells while integrating them into group 3.[2][6]
Arrangement and Basis
Group numbering and nomenclature
The numbering and nomenclature of groups in the periodic table have evolved to standardize scientific communication and reflect the electronic structure of elements. Historically, two primary systems were used before the adoption of a unified approach. The older IUPAC system, prevalent in Europe, employed Roman numerals from I to VIII followed by A or B to distinguish main group elements (A subgroups, including groups now numbered 1, 2, and 13–18) from transition metals (B subgroups, now groups 3–12).[9] In contrast, the Chemical Abstracts Service (CAS) system, more common in North America, also used Roman numerals but reversed the A/B designations, assigning A to main groups and B to transition groups, with group VIII encompassing what are now groups 8–10.[9]In 1988, the International Union of Pure and Applied Chemistry (IUPAC) recommended a modern numbering system using Arabic numerals from 1 to 18 for all groups, from left to right across the table, to resolve inconsistencies between regional conventions and provide a clear, continuous sequence.[10] This 18-group format accommodates the s-block (groups 1–2), d-block (groups 3–12), and p-block (groups 13–18) in a single row of 18 columns, while the f-block elements (lanthanoids and actinoids) are conventionally placed below the main body to avoid expanding the table's width beyond 18 units, preserving its compactness without disrupting the primary periodicity trends.[2]IUPAC endorses specific collective names for several groups to highlight shared chemical properties, prioritizing systematic nomenclature over older trivial terms. Group 1 is designated the alkali metals, group 2 the alkaline earth metals, groups 3–12 the transition metals, group 15 the pnictogens, group 16 the chalcogens, group 17 the halogens, and group 18 the noble gases.[11] These names derive from characteristic behaviors, such as the reactivity of alkali metals in water or the inertness of noble gases. While some trivial names persist in informal use—for instance, "coinage metals" for group 11 (copper, silver, gold) due to their historical role in currency—they are not officially recommended by IUPAC, as they lack the precision of systematic designations.[12]
Presentation formats
The periodic table is most commonly presented in a medium-form layout consisting of 18 columns, where the main body includes the s-, d-, and p-blocks, and the f-block elements (lanthanides and actinides) are detached and placed below or to the side for clarity and space efficiency.[2][13] This format, endorsed by organizations like the International Union of Pure and Applied Chemistry (IUPAC), balances comprehensiveness with readability, allowing users to visualize periodic trends across periods and groups without excessive width.[14]Variations in form include the short-form table, which uses 8 columns focused on s- and p-blocks with transition metals condensed, and the long-form table, which expands to 32 columns by integrating the f-block inline between groups 2 and 3.[13] The short form emphasizes valence similarities but omits detailed transition series, while the long form provides a continuous representation of electron shell filling, though it can appear unwieldy in print.[15] These formats cater to different pedagogical needs, with the medium form prevailing in standard references.Thematic variants adapt the table's layout or coloring to highlight specific properties, such as atomic weight, density, or year of discovery, aiding in the study of historical or physical trends. For instance, tables colored by density use gradients from light (e.g., alkali metals like lithium at 0.534 g/cm³) to dark shades for heavy elements (e.g., osmium at 22.59 g/cm³), revealing patterns in metallic character.[16] Similarly, arrangements by discovery year trace elemental chronology, from ancient metals like gold (pre-3000 BCE) to synthetic ones like oganesson (2002).[17] These visualizations, often used in educational materials, prioritize property correlations over the standard atomic number ordering.Specialized representations include spiral forms, such as Theodor Benfey's 1960 design, which arranges elements in a continuous coil to emphasize periodicity without abrupt breaks between periods, placing hydrogen centrally and expanding outward. Three-dimensional models, like those rendering elements as stacked spheres or helical structures, offer spatial insights into orbital overlaps but are typically exploratory rather than standard.[18]Digital and interactive formats have become prevalent, with online tools allowing users to hover over elements for details on properties like electronegativity or isotopes, or to filter by categories such as radioactivity.[19] Platforms like the Royal Society of Chemistry's interactive table integrate videos, podcasts, and data visualizations, enhancing accessibility for research and education.[6]For practical use, periodic tables appear in compact formats, such as pocket-sized cards (e.g., 4x6 inches) listing essentials like atomic numbers and symbols for quick reference, versus expanded wall charts (up to poster size) that include detailed trends, electron configurations, and color-coding for blocks.[20][21] This distinction accommodates portability for students versus comprehensive display in laboratories.
Electron configurations and subshell filling
The electron configuration of an atom describes the distribution of its electrons among the available atomic orbitals, which forms the quantum mechanical foundation for the periodic table's arrangement of elements. According to quantum mechanics, electrons occupy orbitals characterized by quantum numbers: the principal quantum number n (indicating energy level), the azimuthal quantum number l (defining subshell type: l=0 for s, l=1 for p, l=2 for d, l=3 for f), the magnetic quantum number m_l (orbital orientation), and the spin quantum number m_s (\pm 1/2). This configuration determines an element's chemical properties and position in the table, as electrons fill subshells in a specific order to achieve the lowest energy state.[22]The Aufbau principle governs this filling process, stating that electrons occupy the lowest energy orbitals available before moving to higher energy ones in the ground state of an atom or ion. This "building up" approach results in the sequential order of subshells: 1s, 2s, 2p, 3s, 3p, 4s, 3d, 4p, and so on, with each subshell holding a maximum number of electrons (2 for s, 6 for p, 10 for d, 14 for f). The principle relies on the relative energies of orbitals, which are influenced by electron-electron interactions in multi-electron atoms.[22][23]The precise order of subshell filling follows the Madelung rule, also known as the n+l rule, which predicts that orbitals fill in order of increasing sum of the principal quantum number n plus the azimuthal quantum number l (n+l); for orbitals with the same n+l value, the one with lower n fills first. For example, the 4s orbital (n=4, l=0; n+l=4) fills before the 3d orbital (n=3, l=2; n+l=5), leading to the common sequence up to the 4th period. This empirical rule, proposed by Erwin Madelung in 1936, accurately describes the ground-state configurations for most elements up to atomic number 118.[24][25]Filling within subshells is constrained by the Pauli exclusion principle and Hund's rule. The Pauli exclusion principle dictates that no two electrons in an atom can have the same set of four quantum numbers, limiting each orbital to a maximum of two electrons with opposite spins (one with m_s = +1/2, the other m_s = -1/2). This ensures, for instance, that the 1s subshell holds exactly two electrons in helium. Hund's rule complements this by stating that, in degenerate orbitals (those of equal energy within a subshell), electrons occupy each orbital singly with parallel spins before pairing up, maximizing the total spin and minimizing electron repulsion. For example, in the carbon atom's 2p subshell, the three electrons occupy separate p orbitals with parallel spins rather than pairing in one orbital.[26][27]While these rules hold for most elements, exceptions occur in transition metals where the energy difference between the (n+1)s and nd subshells is small, leading to configurations that prioritize subshell stability over strict Aufbau order. Chromium (atomic number 24) has the configuration [Ar] 4s¹ 3d⁵ instead of the expected [Ar] 4s² 3d⁴, as the half-filled 3d subshell provides greater stability through reduced electron repulsion and exchange energy benefits. Similarly, copper (atomic number 29) adopts [Ar] 4s¹ 3d¹⁰ rather than [Ar] 4s² 3d⁹, favoring the fully filled 3d subshell for enhanced stability due to symmetric electron distribution and minimized pairing energy. These anomalies highlight how subtle energy differences can influence configurations in the d-block.[28]The periodic table is divided into blocks based on the subshell being filled by the valence electrons, reflecting the progressive occupation of s, p, d, and f orbitals. The s-block comprises groups 1 and 2 (alkali and alkaline earth metals, plus hydrogen and helium), where valence electrons fill the ns subshell. The p-block includes groups 13 to 18 (main group elements), with valence electrons in the np subshell. The d-block spans groups 3 to 12 (transition metals), characterized by filling of the (n-1)d subshell. The f-block consists of the lanthanides (filling 4f) and actinides (filling 5f), typically placed below the main body of the table. This block classification underscores how electron configurations dictate the table's structure and elemental similarities within blocks.[29]
Anomalies in the Standard Table
Period 1 peculiarities
The first period of the periodic table is unique in comprising only two elements: hydrogen (atomic number 1) and helium (atomic number 2), which occupy the initial row due to the filling of the lowest energy level.[30] Unlike subsequent periods, Period 1 lacks complexity from higher subshells, resulting in simplified electronic structures and behaviors that deviate from broader periodic trends.Hydrogen possesses an electron configuration of [1s¹], with a single electron in the 1s orbital, conferring a valence electron count analogous to alkali metals yet distinct chemical versatility.[31] This leads to dual nonmetallic and metal-like properties: hydrogen predominantly acts as a nonmetal, forming covalent bonds in compounds like H₂O and NH₃, but can also exhibit metallic character by ionizing to H⁺ in acidic solutions or, under specific conditions, accepting an electron to form hydride ions (H⁻) akin to halogens.[32] Consequently, its placement remains debated, traditionally positioned above Group 1 for its ns¹ configuration and reactivity parallels with lithium and sodium, though it shares electronegativity and anion-forming tendencies with Group 17 elements, prompting proposals for alternative positioning or dual placement.[33]Helium, with the configuration [1s²], achieves a closed-shell structure that renders it the quintessential noble gas, despite its isolation in Period 1 without p-block companions.[34] As the smallest atom after hydrogen, helium's extreme inertness stems from this fully occupied 1s orbital, which provides maximal electronic stability and repels interactions with other species under ambient conditions, precluding stable compounds except under extreme pressures.[35] The absence of d and f subshells in Period 1—limited solely to the 1s orbital—eliminates inner core electrons and complex orbital overlaps, fostering atypical bonding: hydrogen engages in straightforward s-s sigma bonds without hybridization possibilities seen in later periods, while helium remains monatomic, defying the diatomic or polyatomic tendencies of other elements.[36]These electronic simplicities manifest in notable physical anomalies with chemical relevance. Hydrogen gas exhibits exceptionally low density (0.0899 g/L at 0°C and 1 atm), far below that of air (1.293 g/L), enabling its historical use in buoyancy applications but also highlighting its minimal intermolecular forces.[37] Helium-4, in liquid form, transitions to a superfluid state below 2.17 K, characterized by zero viscosity and enhanced thermal conductivity due to quantum Bose-Einstein condensation, which facilitates chemical studies by stabilizing reactive species in ultracold, non-interacting environments without solvent interference.[38]
Group 3 composition debate
The composition of Group 3 in the periodic table has been a subject of debate since the early 20th century, centering on whether it should consist of scandium (Sc), yttrium (Y), lanthanum (La), and actinium (Ac), or alternatively Sc, Y, lutetium (Lu), and lawrencium (Lr). The traditional arrangement places La and Ac in Group 3, treating them as the first members of the f-block lanthanides and actinides, respectively, with cerium (Ce) and thorium (Th) following as the start of the 4f and 5f series. This view aligns with historical periodic tables and emphasizes chemical similarities, such as the +3 oxidation state predominant in these elements and their ionic behavior resembling Groups 1 and 2.[39][40]Proponents of the alternative composition argue for Sc, Y, Lu, and Lr based on electronic configurations and structural consistency in extended periodic tables. Lu and Lr exhibit ground-state configurations of [Xe] 4f¹⁴ 5d¹ 6s² and [Rn] 5f¹⁴ 6d¹ 7s², respectively, featuring an incomplete f-subshell and a single d-electron, which better matches the d-block character of Sc ([Ar] 3d¹ 4s²) and Y ([Kr] 4d¹ 5s²). This placement avoids splitting the d-block in 32-column formats and maintains smoother trends in atomic radii and coordination numbers down the group. However, relativistic effects in superheavy elements like Lr complicate this, as its configuration may involve 7p¹ occupancy instead of the expected 6d¹ 7s² due to stabilization of the 7p_{1/2} orbital, potentially aligning its chemistry more closely with p-block elements.[39][41][42]The debate gained renewed attention in 1982 when William B. Jensen proposed the Sc-Y-Lu-Lr arrangement in the Journal of Chemical Education, citing the completion of the 4f subshell in Lu as a key factor for d-block continuity. Despite this, the traditional Sc-Y-La-Ac setup dominates most textbooks and tables due to its adherence to Mendeleev's periodic law, where La represents the first recurrence after Y, preserving isodiagonality (e.g., Ca diagonal to Y and Ce) and the integrity of the f-block contraction. Chemical evidence is mixed: La forms more similar compounds with Sc and Y in some similarity landscapes, while Lu shares hexagonal close-packed structures with them.[39]In 2015, the International Union of Pure and Applied Chemistry (IUPAC) established a task group to resolve the issue, aiming to recommend one composition by 2021. The task group issued a provisional report in 2021 indicating no objective criterion to favor one over the other, allowing flexibility based on context. As of November 2025, no final recommendation has been issued.[40][43][39]
This ongoing uncertainty affects periodic table layouts, with the traditional version suiting 18-column formats and the alternative enhancing symmetry in wider tables, while also influencing the naming of the "scandium group." Implications include potential revisions to educational materials and databases, though no mandatory change has been imposed.
Periodic Trends
Atomic and ionic radii
The atomic radius of an element is a measure of the size of its atom, defined as the distance from the nucleus to the outermost electron shell, though precise measurement varies by context due to the probabilistic nature of electron positions. Common types include the covalent radius, half the distance between nuclei of two identical atoms bonded covalently; the van der Waals radius, half the distance between non-bonded atoms in a molecular solid; and calculated radii, derived from quantum mechanical models such as Hartree-Fock methods. These measures generally align in trends but differ in absolute values, with covalent radii often used for main-group elements and van der Waals for noble gases.[44][45]Across a period, atomic radii decrease from left to right due to increasing effective nuclear charge (Zeff), where protons in the nucleus pull electrons closer without sufficient shielding from inner shells, compressing the electron cloud. Down a group, radii increase as additional electron shells are added, with inner electrons providing shielding that reduces Zeff felt by outer electrons. For instance, in period 2, the atomic radius decreases from lithium (152 pm) to fluorine (72 pm), reflecting the rise in Zeff from 3 to 9 protons while occupying the same 2s/2p subshell. This trend holds broadly but can vary slightly by radius type; covalent radii for period 2 show lithium at 128 pm and fluorine at 64 pm.[44][46]Ionic radii describe the size of ions in ionic compounds, determined from internuclear distances in crystal lattices, and follow similar periodic trends but are influenced by charge. Cations are smaller than their neutral atoms because electron removal reduces electron-electron repulsion, allowing greater nuclear attraction; anions are larger as added electrons increase repulsion. For isoelectronic ions (same electron configuration), radius decreases with increasing nuclear charge, as seen in the series O2– (140 pm), F– (133 pm), Na+ (102 pm), and Mg2+ (72 pm), all with 10 electrons. Across periods, cationic radii decrease, amplified by higher charges in later groups, while anionic radii show less variation; down groups, both increase due to added shells. These values are standardized using effective ionic radii assuming fixed oxide or fluoride references.[44]In transition metals, atomic and ionic radii exhibit irregularities due to poor shielding by d-electrons, leading to smaller-than-expected sizes compared to s- and p-block analogs. For example, in period 4, radii decrease from scandium (161 pm) to zinc (134 pm), but the contraction is less pronounced than in main groups because d-electrons add to Zeff without fully expanding the size. This lanthanide contraction further compresses period 6 transition metal radii due to 4f-electron shielding inefficiencies.[44]
Ionization energies
Ionization energy is defined as the minimum energy required to remove an electron from a gaseous atom or ion in its ground state. The process for the first ionization energy can be represented by the equation:\text{M(g)} \rightarrow \text{M}^+(\text{g}) + \text{e}^- \quad \Delta E = \text{IE}_1where M is a neutral atom, and the energy change corresponds to the ionization energy in units such as kJ/mol.[47] Successive ionization energies refer to the energy needed to remove additional electrons, denoted as IE₂ for the second, IE₃ for the third, and so on, from the resulting cation.[48]In the periodic table, the first ionization energy generally increases from left to right across a period due to the increasing effective nuclear charge (Zeff), which exerts a stronger pull on the valence electrons as the atomic number rises while the principal quantum number remains constant. Conversely, it decreases down a group because the valence electrons are in higher energy shells, farther from the nucleus, and experience greater shielding from inner electrons, reducing Zeff./Descriptive_Chemistry/Periodic_Trends_of_Elemental_Properties/Periodic_Trends) As a result, alkali metals in group 1 exhibit the lowest first ionization energies—for example, cesium at approximately 376 kJ/mol—while noble gases in group 18 have the highest, such as helium at 2372 kJ/mol, owing to their stable, filled valence shells. This trend aligns inversely with atomic radius, as larger atoms facilitate easier electron removal.[48]Successive ionization energies increase progressively for each element, with sharp rises occurring after the removal of all valence electrons, as subsequent electrons are stripped from more stable inner shells closer to the nucleus./Physical_Properties_of_Matter/Atomic_and_Molecular_Properties/Ionization_Energy) For sodium (Na), the first ionization energy is 496 kJ/mol, corresponding to the loss of its single 3s valence electron, but the second jumps to 4562 kJ/mol as it removes an electron from the stable neon-like core.[49] These discontinuities in successive ionization energy plots help identify an element's group number by revealing the number of valence electrons.[48]Key factors influencing ionization energies include the effective nuclear charge, which rises across periods due to poor shielding by electrons in the same shell, and the stability of electron configurations, such as half-filled or fully filled subshells that resist electron removal. Exceptions to the general trend occur, notably a decrease from group 2 to group 13 elements (e.g., beryllium's first IE of 899 kJ/mol exceeds boron's 801 kJ/mol, and magnesium's 738 kJ/mol exceeds aluminum's 577 kJ/mol), attributed to the stability of the ns² configuration in group 2 versus the lower energy required to remove the np¹ electron from group 13./Physical_Properties_of_Matter/Atomic_and_Molecular_Properties/Ionization_Energy)
Electron affinities and electronegativities
Electron affinity (EA) is defined as the energy released when an electron is added to a neutral atom in the gas phase to form a negative ion, corresponding to the process \ce{E + e^- -> E^-}.[50] This quantity is typically expressed in kilojoules per mole (kJ/mol) or electron volts (eV), with negative values indicating an exothermic process where the atom readily accepts the electron./06:_The_Periodic_Table/6.19:Periodic_Trends-_Electron_Affinity)In the periodic table, electron affinities become more negative (higher affinity) from left to right across a period due to increasing effective nuclear charge (Z_\text{eff}), which strengthens the attraction for the added electron, and less negative (lower affinity) down a group as the atomic radius increases and shielding reduces nuclear pull.[51] Halogens exhibit the highest electron affinities among the elements, reflecting their nearly filled valence shells; for example, chlorine has an EA of -349 kJ/mol, facilitating its role in forming stable anions./06:_The_Periodic_Table/6.19:Periodic_Trends-_Electron_Affinity) In contrast, noble gases have positive electron affinities, indicating endothermic addition of an electron to their stable, filled shells; helium, for instance, requires +48 kJ/mol to form \ce{He^-}./06:_The_Periodic_Table/6.19:Periodic_Trends-_Electron_Affinity)Electronegativity (EN) quantifies an atom's ability to attract shared electrons in a chemical bond, distinguishing it from electron affinity by focusing on bonded rather than isolated atoms.[52] The most widely used scale is the Pauling scale, developed by Linus Pauling in 1932 based on differences in bond dissociation energies, where fluorine is assigned the highest value of 4.0 and cesium the lowest at 0.79.[52] On this dimensionless scale, electronegativity increases across a period and decreases down a group, driven by rising Z_\text{eff} that enhances electron attraction without proportional shielding in the valence shell.[51]Another prominent scale is the Mulliken electronegativity, proposed by Robert S. Mulliken in 1934, defined as the average of the first ionization energy (IE) and electron affinity (EA) for the valence state: \chi_M = \frac{IE + EA}{2}, often adjusted to align with Pauling units.[53] This thermodynamic approach captures the balance between an atom's tendencies to lose or gain electrons, yielding trends similar to Pauling's—increasing across periods and decreasing down groups—though it is more sensitive to valence state variations.[53]Exceptions to these trends occur notably in transition metals, where electronegativities show less variation and smoother increases across the first row, from scandium (1.36 on the Pauling scale) to copper (1.90), due to the involvement of d-orbitals and small energy gaps between s and d subshells that alter electron distribution and Z_\text{eff} effects.[54] This variability stems from the partial filling of d subshells, leading to context-dependent electron attraction in bonds.[54]
Valence and oxidation states
Valence electrons are the electrons in an atom's outermost principal energy level, specifically the s and p electrons in the valence shell for main group elements, which largely determine an element's chemical reactivity and bonding behavior.[55] These electrons occupy the highest energy subshells and participate in forming chemical bonds, with the number of valence electrons corresponding to the group number in the periodic table (ranging from 1 in Group 1 to 8 in Group 18).[55] For instance, elements in Group 14, such as carbon, have four valence electrons (2s²2p² configuration), enabling them to form four covalent bonds in compounds like methane (CH₄).[56]Oxidation states, also known as oxidation numbers, represent the hypothetical charge on an atom in a compound assuming complete electron transfer in ionic bonds or formal charge assignment in covalent bonds, as defined by IUPAC guidelines.[57] In main group elements, the maximum positive oxidation state typically equals the group number, reflecting the loss of all valence electrons, while nonmetals often exhibit negative states up to the group number minus 8 (e.g., oxygen in Group 16 commonly shows -2 in oxides like H₂O).[56] Transition metals in the d-block display variable oxidation states due to the involvement of both s and d electrons; for example, iron (Fe) commonly achieves +2 and +3 states in compounds like FeCl₂ and FeCl₃, respectively.[56] Across a period from left to right, the range of stable oxidation states increases up to Group 14 or 15 and then decreases, as valence electrons fill and elements shift from metallic to nonmetallic character.Down a group in the p-block, a key trend is the inert pair effect, where the ns² valence electrons become increasingly reluctant to participate in bonding due to poor shielding by d and f electrons and weakening of metal-ligand bonds, stabilizing lower oxidation states by two units compared to the group maximum.[58] This effect is prominent in heavier elements of Groups 13–16; in Group 13, thallium (Tl) favors +1 over +3 (e.g., stable Tl⁺ in Tl₂O), while in Group 14, lead (Pb) prefers +2 over +4 (e.g., PbO vs. PbO₂).[58] Similarly, in Group 15, bismuth (Bi) exhibits +3 as more stable than +5 (e.g., Bi₂O₃), contrasting with lighter elements like nitrogen, which reaches +5 in HNO₃.[58] These trends arise from relativistic effects and increasing atomic size, which reduce the energy gain from higher oxidation states.[58]
Metallic and nonmetallic character
Metallic character in the periodic table refers to the degree to which elements exhibit properties typical of metals, such as the ability to lose electrons easily, forming positive ions, and displaying physical traits like ductility, malleability, high electrical and thermalconductivity, and luster.[59] These properties arise from the presence of delocalized valence electrons in metallic bonding, where electrons are free to move throughout a lattice of positive metal ions, facilitating conduction and allowing atoms to slide past one another without breaking bonds./Chemical_Bonding/Fundamentals_of_Chemical_Bonding/Metallic_Bonding) Metals occupy the left and lower portions of the periodic table, including alkali metals like sodium and transition metals like iron./CHEM_142%3A_Text_(Brzezinski)/04%3A_Atoms_and_Elements/4.06%3A_Periodic_Trends-_Atomic_Radius/4.6.02%3A_Metallic_and_Nonmetallic_Character)In contrast, nonmetallic character describes elements that tend to gain electrons to form negative ions, exhibiting properties such as brittleness, poor electrical and thermalconductivity, and lack of luster.[59] This behavior stems from localized valence electrons tightly bound to individual atoms, resulting in directional covalent bonds that resist deformation and impede electron flow./Chemical_Bonding/Fundamentals_of_Chemical_Bonding/Metallic_Bonding) Nonmetals are primarily located in the upper right region of the table, including halogens like fluorine and noble gases like helium./CHEM_142%3A_Text_(Brzezinski)/04%3A_Atoms_and_Elements/4.06%3A_Periodic_Trends-_Atomic_Radius/4.6.02%3A_Metallic_and_Nonmetallic_Character)Elements with intermediate properties, known as metalloids, lie along a diagonal "staircase" line that roughly separates metals from nonmetals, typically including boron (B), silicon (Si), germanium (Ge), arsenic (As), antimony (Sb), and tellurium (Te).[60] Metalloids exhibit semiconductor properties, where electrical conductivity is moderate and can be tuned by factors like temperature or doping, due to a small band gap between valence and conduction bands that allows partial electron mobility.[60] For instance, silicon is widely used in electronics for its tunable semiconducting behavior.[60]The trend in metallic character follows a diagonal pattern across the periodic table: it increases from top to bottom within a group, as atomic size grows and valence electrons experience weaker nuclear attraction, making them easier to delocalize./CHEM_142%3A_Text_(Brzezinski)/04%3A_Atoms_and_Elements/4.06%3A_Periodic_Trends-Atomic_Radius/4.6.02%3A_Metallic_and_Nonmetallic_Character) Conversely, metallic character decreases from left to right across a period, as effective nuclear charge rises, pulling electrons closer and favoring localized bonding over delocalization./CHEM_142%3A_Text(Brzezinski)/04%3A_Atoms_and_Elements/4.06%3A_Periodic_Trends-Atomic_Radius/4.6.02%3A_Metallic_and_Nonmetallic_Character) This periodicity is linked to variations in atomic radii, where larger radii correlate with enhanced metallicity./CHEM_142%3A_Text(Brzezinski)/04%3A_Atoms_and_Elements/4.06%3A_Periodic_Trends-_Atomic_Radius/4.6.02%3A_Metallic_and_Nonmetallic_Character)Certain elements demonstrate how structure influences character through allotropes, such as carbon: diamond, with its three-dimensional covalent network, behaves as a nonmetalinsulator, while graphite, featuring layers with delocalized π electrons in hexagonal sheets, acts as a conductor akin to a metal./14%3A_The_Group_14_Elements/14.04%3A_Allotropes_of_Carbon/14.4A%3A_Graphite_and_Diamond_-_Structure_and_Properties)
Element Classification
By chemical properties
Elements in the periodic table are classified by their chemical properties into metals, nonmetals, metalloids, and noble gases, primarily based on their tendencies to form cations or anions, types of bonds in compounds, and overall reactivity.[61] This classification highlights patterns in how elements interact to form ionic, covalent, or other compounds, influencing their roles in chemical reactions and material synthesis.[62]Metals, located predominantly on the left and center of the periodic table, are characterized by their ability to lose electrons and form positive cations in reactions.[62] They typically create ionic compounds with nonmetals, where the metallic element acts as the cation, as seen in salts like sodium chloride (NaCl).[61] Alkali metals in group 1 and alkaline earth metals in group 2 exhibit particularly high reactivity, readily donating one or two electrons to achieve stable electron configurations.[63]Nonmetals, situated on the upper right side, tend to gain electrons to form anions or share electrons in covalent bonds due to their high electronegativity.[62] These elements often produce acidic oxides or compounds, such as sulfur dioxide (SO₂) from sulfur, which reacts with water to form acids.[64] Halogens in group 17 exemplify this by forming diatomic molecules and reacting vigorously with metals to yield ionic halides.[65]Metalloids, or semimetals, occupy a diagonal band separating metals from nonmetals and display hybrid behaviors, forming amphoteric oxides that can act as either acids or bases depending on conditions.[62] Their compounds often involve covalent bonding with partial ionic character, as in the case of silicon, which forms network solids like silicates.[66] Elements such as boron and arsenic illustrate this duality, enabling applications in semiconductors.[61]Noble gases in group 18 are generally inert owing to their completely filled electron shells, which confer stability and minimal reactivity under standard conditions.[62] However, xenon forms compounds like xenon difluoride (XeF₂), first synthesized in 1962 by reacting xenon with fluorine gas, demonstrating that even these elements can participate in covalent bonding under forcing conditions.[67]Reactivity within groups follows predictable trends: for alkali metals, it increases down the group as atomic size grows and ionization energy decreases, making cesium more reactive than lithium.[68] Conversely, in the halogens, reactivity decreases down the group, with fluorine being the most reactive due to its small size and high electronegativity, surpassing chlorine in displacing less reactive halides.[65] These series guide predictions of reaction outcomes and compound stability.[63]
By electron block
The periodic table is divided into four blocks—s, p, d, and f—based on the type of atomic orbital being filled by the valence electrons in the ground state of the atoms. This classification arises from the quantum mechanical description of electron configurations, where the blocks correspond to the subshell (s, p, d, or f) that receives the differentiating electron.[69]The s-block encompasses groups 1 and 2 of the periodic table, consisting of the alkali metals and alkaline earth metals, respectively. These elements have valence electron configurations ending in ns¹ for group 1 or ns² for group 2, where n is the principal quantum number of the outermost shell. As main-group elements, they exhibit properties tied to their single or paired s electrons, such as relatively low densities and luster typical of metals.[69][70]The p-block includes groups 13 through 18 and represents the largest block, featuring a wide diversity of elements ranging from metals (e.g., aluminum in group 13) to metalloids (e.g., silicon in group 14), nonmetals (e.g., oxygen in group 16), and noble gases (group 18). Valence electrons occupy the np¹ to np⁶ subshells, leading to varied bonding behaviors and states of matter across the block. This diversity stems from the progressive filling of the three p orbitals, which allows for multiple valence possibilities.[69][70]The d-block, comprising groups 3 through 12, contains the transition metals, characterized by the filling of (n-1)d subshells alongside ns electrons (typically ns¹ or ns²). These elements often display partially filled d orbitals in their common ions, resulting in unique properties such as the formation of colored compounds due to d-d electronic transitions and variable oxidation states arising from the accessibility of multiple d electrons for bonding. For instance, manganese exhibits oxidation states from +2 to +7, reflecting the flexibility of its 3d electrons. Transition metals are generally hard, high-melting-point solids with good electrical conductivity, and many are paramagnetic owing to unpaired d electrons.[71][72][73]The f-block consists of the lanthanides (elements 58–71, filling the 4f subshell) and actinides (elements 90–103, filling the 5f subshell), typically placed below the main body of the periodic table for compactness. These inner transition elements share similar chemical properties due to the shielding effect of f electrons, which minimally affects outer s and d orbitals; for example, lanthanides often adopt +3 oxidation states and form stable oxides. The actinides, however, are predominantly radioactive, with all elements beyond uranium (Z=92) being synthetic and unstable.[69][73]Across the blocks, there is a general trend of increasing structural and chemical complexity from s to f, as the higher angular momentum orbitals (p, d, f) introduce more electrons into diffuse, shielded regions, leading to greater variability in bonding and electronic transitions. The s-block's simplicity contrasts with the f-block's intricate magnetic and spectroscopic behaviors, reflecting the progressive involvement of inner subshells.[69][72]
Historical Development
Early classifications
Early efforts to classify chemical substances date back to ancient philosophy, where Empedocles of Acragas proposed in the 5th century BCE that all matter consists of four fundamental roots: earth, air, fire, and water, combined and separated by forces of love and strife.[74] This qualitative framework influenced Western thought for centuries but lacked empirical basis or recognition of distinct chemical elements.[74]In the 18th century, Antoine Lavoisier advanced a more systematic approach by compiling a list of 33 substances he considered simple elements in his 1789 treatise Traité élémentaire de chimie, categorizing them into groups such as metals, nonmetals (including those forming acids like oxygen and sulfur), earths, and light (heat).[75][76] Lavoisier's classification emphasized chemical reactivity, particularly acid-base properties, and marked a shift toward experimental verification, though it included some compounds mistaken for elements, like "light."[75]Building on emerging atomic weight measurements, Johann Wolfgang Döbereiner identified patterns in the early 19th century, noting in 1817 and expanding by 1829 that certain groups of three elements, or triads, shared similar chemical properties with the atomic weight of the middle element approximately equaling the average of the other two.[77] For instance, in the triad lithium (atomic weight ~7), sodium (~23), and potassium (~39), sodium's weight closely approximates the mean of lithium and potassium, and all exhibit alkaline reactivity.[77] Döbereiner recognized about five such triads among known elements, suggesting an underlying order based on atomic weights.[78]By the 1850s, John Newlands extended these ideas, arranging the 56 known elements in order of increasing atomic weight and observing in 1865 that their properties repeated every eighth element, akin to the octaves in music.[79][80] He termed this the "law of octaves," proposing a tabular arrangement where elements like lithium, beryllium, boron, carbon, nitrogen, oxygen, and fluorine showed analogous behaviors in successive groups.[79]These early classifications were constrained by the incomplete roster of elements—only about 60 were identified by the mid-19th century—and the absence of atomic numbers as a fundamental ordering principle, which would not be discovered until 1913.[81]Döbereiner's triads covered only a fraction of elements, while Newlands' octaves faltered beyond lighter elements, forcing awkward pairings like cobalt and nickel with halogens.[82]
Mendeleev's contributions
In 1869, Dmitri Mendeleev presented his periodic table to the Russian Chemical Society, arranging the 63 known elements in order of increasing atomic weight across horizontal rows known as periods, while placing elements with similar chemical properties in vertical columns called groups.[83][84] This structure revealed a recurring pattern, or periodicity, in elemental properties such as valence and reactivity, with analogous elements aligning vertically despite variations in atomic weight.[83] Mendeleev left intentional gaps in the table for undiscovered elements, anticipating their existence based on the observed patterns; for instance, he posited a gap below silicon for an element he termed "eka-silicon," which he later predicted would have an atomic weight around 72 and properties intermediate between silicon and tin.[84][85]Mendeleev's predictions extended to several missing elements, including "eka-aluminum" (later identified as gallium, discovered in 1875) and "eka-boron" (scandium, discovered in 1879), with forecasted properties like density, melting point, and valence that closely matched experimental findings upon their isolation.[85] Similarly, germanium, discovered in 1886, aligned remarkably with Mendeleev's eka-silicon predictions, including a specific gravity of about 5.5 and a grayish-white metallic appearance. To maintain chemical coherence, Mendeleev occasionally inverted the order of elements based on atomic weight when properties demanded it, such as placing iodine before tellurium despite tellurium's higher atomic weight of 127.6 compared to iodine's 126.9, prioritizing their halogen and chalcogen affiliations respectively.[86][13]Through this system, Mendeleev formulated the periodic law, stating that "the elements, if arranged according to their atomic weights, exhibit an evident stepwise variation of properties," establishing elemental characteristics as a periodic function of atomic weight.[83] This empirical framework not only organized known elements but also guided future discoveries, demonstrating the table's predictive power and solidifying its role as a cornerstone of chemical classification.[84][85]
Atomic number and modern refinements
In 1911, Antonius van den Broek proposed that the atomic number of an element corresponds to the positive charge of its atomic nucleus, equating it to the number of protons within, which provided an early theoretical basis for ordering elements beyond atomic weight. This idea anticipated experimental confirmation and addressed inconsistencies in earlier periodic arrangements.The breakthrough came in 1913 when Henry Moseley conducted X-ray spectroscopy experiments on elements from aluminum to gold, measuring the frequencies of their characteristic X-ray emissions. He discovered Moseley's law, which states that the square root of the frequency (ν) of these X-rays is linearly proportional to the atomic number Z, expressed as √ν ∝ (Z - b), where b is a constant screening factor. This relationship, approximately ν ∝ Z² for higher Z, allowed precise determination of atomic numbers and resolved ordering anomalies in Mendeleev's table, such as placing cobalt (Z=27) before nickel (Z=28) and tellurium (Z=52) before iodine (Z=53), despite their respective atomic masses suggesting the reverse.[87]Moseley's findings revised the periodic law, establishing that the physical and chemical properties of elements vary periodically with atomic numberZ rather than atomic weight, confirming the fundamental role of nuclear charge in elemental behavior. This ordering predicted gaps for undiscovered elements with Z=43, 61, 72, 75, and others up to uranium (Z=92), guiding subsequent searches and completing the table through natural elements.[87]A key refinement was recognizing the negligible impact of isotopes on chemical properties, as isotopes of an element share the same Z and thus the same number of electrons, determining reactivity, while differences in neutron count primarily affect nuclear stability and mass.[87] This clarified why atomic weight-based ordering had led to discrepancies, as isotopic variations in natural abundances could invert relative masses without altering chemical periodicity.
Integration of quantum mechanics
The integration of quantum mechanics into the understanding of atomic structure in the 1920s provided a theoretical foundation for the periodicity observed in the elements, explaining the arrangement of electrons in discrete energy levels that correspond to the table's groups and periods. Building on Niels Bohr's 1913 model of quantized circular orbits for the hydrogen atom, Arnold Sommerfeld extended this framework in 1916 to include elliptical orbits, introducing the concept of quantized angular momentum. In the Bohr-Sommerfeld model, the angular momentum of an electron's orbit is given by L = k \frac{h}{2\pi}, where k is an integer azimuthal quantum number (now denoted as l, ranging from 1 to n, with n being the principal quantum number), allowing for the explanation of fine spectral lines in alkali metals and the relativistic effects on orbital precession. This model marked a crucial step toward multi-electron atoms, laying the groundwork for associating electron configurations with chemical periodicity by suggesting that orbital shapes and energies influence atomic stability and reactivity.[88]The full incorporation of quantum mechanics came with the identification of four quantum numbers describing electron states: the principal quantum number n (determining the energy shell, n = 1, 2, [3, \dots](/page/3_Dots)), the azimuthal quantum number l (defining subshell shape, [l](/page/L') = 0 to n-1), the magnetic quantum number m_l (specifying orbital orientation, m_l = -[l](/page/L') to +[l](/page/L')), and the spin quantum number m_s (indicating electron spin, m_s = \pm \frac{1}{2}). Edmund C. Stoner proposed in 1924 that electrons fill atomic levels according to these quantum numbers, with the maximum occupancy of subshells following $2(2[l](/page/L') + 1), leading to capacities of 2 for s ([l](/page/L')=0), 6 for p ([l](/page/L')=1), 10 for d ([l](/page/L')=2), and 14 for f ([l](/page/L')=3) subshells. This building-up (Aufbau) principle, formalized in the mid-1920s by Niels Bohr and Wolfgang Pauli, dictates that electrons occupy the lowest available energy states sequentially as atomic number increases, directly accounting for the periodic repetition of properties every 2, 8, 18, or 32 electrons, corresponding to shell completions.[89][90]Wolfgang Pauli's exclusion principle, articulated in 1925, further solidified this framework by stating that no two electrons in an atom can share the same set of four quantum numbers, ensuring that Pauli matrices enforce antisymmetric wavefunctions for fermions and preventing electron collapse into the lowest state. This principle explained the discrete filling of subshells and the stability of noble gas configurations, such as neon's closed 2p subshell. Complementing this, John C. Slater developed empirical shielding rules in 1930 to quantify how inner electrons reduce the effective nuclear charge experienced by outer electrons, expressed as Z_{\text{eff}} = Z - \sigma, where Z is the atomic number and \sigma is the shielding constant derived from electron grouping (e.g., 0.35 for each other electron in the same group, 0.85 for those in the n-1 shell, and 1.00 for deeper shells). These rules provided a practical method to estimate orbital energies and predict trends in ionization potentials across the table.[90][91]By the late 1920s, spectroscopic data confirmed that the periodic table's blocks—s, p, d, and f—align precisely with the sequential filling of these subshells: alkali metals end s blocks, halogens complete p blocks, transition metals fill d blocks, and lanthanides/actinides fill f blocks. This quantum mechanical validation, integrating the atomic number Z as the determinant of electron count, transformed the empirical table of Mendeleev into a predictive tool grounded in wave mechanics, resolving anomalies like the placement of transition series.[89]
Synthetic and superheavy elements
Synthetic elements, also known as artificial or man-made elements, are those produced through nuclear reactions rather than occurring naturally in significant quantities. The first such elements beyond uranium (atomic number 92) are the transuranic elements, beginning with neptunium (93), which was synthesized in 1940 by Edwin McMillan and Philip Abelson at the University of California, Berkeley, via neutron irradiation of uranium-238 in a cyclotron, producing uranium-239 that beta-decayed to neptunium-239.[92]Plutonium (94), the next transuranic element, was discovered in 1941 by Glenn T. Seaborg, Edwin McMillan, Joseph W. Kennedy, and Arthur C. Wahl through deuteron bombardment of uranium, yielding uranium-238 that decayed to neptunium-238 and then plutonium-238; this work laid the foundation for the actinide series concept.[93] Subsequent transuranic elements up to oganesson (118), synthesized in 2002 by a joint Russian-American team using calcium-48 ions accelerated onto californium-249 targets at the Joint Institute for Nuclear Research in Dubna, extended the periodic table through similar nuclear fusion reactions.[94]Synthesis of these elements has evolved from early methods involving neutron capture in nuclear reactors or light particle bombardment in cyclotrons to advanced techniques using heavy-ion accelerators for superheavy elements (atomic numbers 104 and above). For instance, early transuranics like neptunium and plutonium were produced via successive neutron captures and beta decays in uranium targets exposed to neutron fluxes, while superheavies require high-energy collisions of heavy projectiles, such as calcium-48 with actinide targets, to form compound nuclei that de-excite into the desired isotopes; these reactions are detected using gas-filled recoil separators to isolate the short-lived products.[95] Representative examples include the production of technetium (43), a lighter synthetic element, via deuteron bombardment of molybdenum in 1937, illustrating the foundational role of particle accelerators in element synthesis, though transuranics demand more intense neutron or ion fluxes due to the need for crossing the neutron drip line.[96]Naming of synthetic elements follows IUPAC conventions, which assign provisional systematic names based on atomic number using Latin roots—such as "ununoctium" (one-one-eight) for element 118—until permanent names are approved after independent verification of discovery claims. For oganesson (118), the temporary name ununoctium was used from its synthesis in 2002 until IUPAC's official ratification and naming in 2016, honoring Yuri Oganessian for his contributions to superheavy element research.[94] This process ensures international consensus, as seen in the joint credits for elements like flerovium (114) and livermorium (116).Superheavy elements exhibit increasing instability with rising atomic number, characterized by half-lives decreasing from years or days for early actinides like plutonium-239 (24,100 years) to microseconds for isotopes of oganesson, due to fission barriers that weaken as proton and neutron numbers increase, leading to rapid spontaneous fission or alpha decay.[97] Theoretical models predict an "island of stability" around atomic numbers Z=114–126 and neutron numbers near N=184, where closed nuclear shells could enhance stability through higher fission barriers and longer half-lives, potentially allowing for more detailed chemical studies of these elements.[98]
Future Extensions
Beyond the seventh period
The eighth period of the periodic table is predicted to commence with element 119, tentatively named ununennium, and extend theoretically to element 172, incorporating a novel g-block of elements known as superactinides.[99] According to relativistic Dirac-Fock calculations, the electron configuration for element 119 is [Og] 8s¹, followed by [Og] 8s² for element 120, marking the filling of the 8s subshell.[99] Subsequent elements from 121 onward initiate the superactinide series, characterized by the occupation of 5g, 6f, and 7d orbitals, with the g-block spanning elements 121–138 in a 18-electron series, potentially exhibiting compact, non-bonding 5g orbitals akin to superlanthanide behavior.[99]Theoretical models anticipate increased nuclear stability for elements near atomic numbers Z=120 and Z=126 due to predicted proton shell closures, which could enhance fission barriers and extend half-lives in the superheavy regime, forming part of a hypothesized "island of stability."[100] For instance, doubly magic configurations such as Z=120 with N=172 neutrons are forecasted to exhibit particularly robust nuclear binding in relativistic mean-field theories.[101] Chemically, these superactinides may display volatile properties, with high oxidation states enabling the formation of compounds like fluorides or oxides (e.g., potential (E125)F₆ analogs), though relativistic effects are expected to contract s and p orbitals, influencing bonding and reactivity.[99]The extended periodic table reaches up to Z=172 through such calculations, where the 8p¹/₂ subshell fills, but relativistic quantum electrodynamic (QED) effects and orbital collapse impose practical limits beyond Z≈137 without additional screening by electrons.[99] Shell closures, such as the 6f¹⁴ at Z=155, further delineate series within period 8, including a 6f block (Z=141–155) and 7d block (Z=156–164).[99] However, synthesizing and observing these elements presents significant challenges, as heavier nuclei suffer from extreme instability due to fission and alpha decay, with no confirmed detections of period 8 elements to date.[100]
Recent research on new elements
In 2024, researchers at the Lawrence Berkeley National Laboratory (LBNL) advanced efforts toward synthesizing element 120 (unbinilium) by demonstrating the feasibility of using a titanium-50 beam to produce superheavy element 116 (livermorium) as a precursor step. This experiment involved bombarding a plutonium-244 target with titanium-50 ions at LBNL's 88-Inch Cyclotron over 22 days, resulting in the detection of two livermorium-290 atoms, confirming the reaction channel's viability for heavier targets like californium-249 needed for unbinilium.[102] Although international collaborations, including those at the Joint Institute for Nuclear Research (JINR) and GSI Helmholtz Centre for Heavy Ion Research, had explored titanium beams earlier, geopolitical tensions led to the U.S. taking the lead, with plans for a dedicated unbinilium campaign using the same approach starting in 2025.[103]In April 2025, the U.S. Department of Energy (DOE) announced the successful demonstration of livermorium production using the titanium-50 beam, as conducted in the 2024 LBNL experiment. This confirmation highlights the method's potential for pursuing element 120 and the island of stability, where longer-lived isotopes of superheavy elements are predicted, though the experiment yielded two livermorium atoms over 22 days.[104][105]As of November 2025, the International Union of Pure and Applied Chemistry (IUPAC) has not verified any new elements beyond 118, with ongoing reviews of claims for elements 119 and 120 from various labs remaining inconclusive due to insufficient independent confirmations. Verification processes continue, emphasizing the need for replicated syntheses across multiple facilities; notable efforts include RIKEN's ongoing attempts to synthesize element 119 using a vanadium-51 beam on curium-248 and LBNL's planned late-2025 campaign for element 120.[106]
Alternative Tables
Structural variations
Structural variations of the periodic table modify the conventional rectangular layout to better align with electron configurations and atomic properties, often emphasizing the sequential filling of orbitals without the interruptions of traditional rows and columns. These adaptations aim to provide a more intuitive visualization of periodicity, particularly by rearranging blocks to reflect the order in which electrons occupy subshells.[107]One prominent example is Charles Janet's left-step periodic table, proposed in 1928. In this arrangement, the s-block elements are positioned on the left side, with the f-, d-, and p-blocks extending rightward in a stepped manner that follows the Aufbau principle of electron filling. This design highlights the recurrence of ns² electron configurations at the end of each period, creating a continuous sequence that mirrors the Madelung rule for orbital ordering. By placing lutetium (element 71) and lawrencium (element 103) directly below yttrium and scandium in the main body of the table, rather than appending the lanthanides and actinides separately, it resolves longstanding ambiguities in Group 3 placement.[108][107]Another variation is Theodor Benfey's spiral periodic table, introduced in 1964, which arranges elements in a two-dimensional radial spiral starting from hydrogen at the center and proceeding outward by increasing atomic number. The spiral folds around "peninsulas" representing the transition metals, lanthanides, and actinides, integrating these series seamlessly into the overall structure without detached rows. This radial format underscores the continuous nature of element progression and the periodic recurrence of properties as one moves along the coil.[109]These structural modifications offer several advantages over the standard table, including the reduction of anomalies such as the irregular positioning in Group 3 and the clearer depiction of discontinuities between periods and blocks, which aids in understanding electron shell transitions. For instance, Janet's table eliminates the need for footnotes or separate f-block insertions, providing a more streamlined view of superheavy element placements even before their synthesis. Similarly, the spiral design highlights radial similarities in chemical behavior, such as noble gases forming an outer ring. Despite these benefits, such variations are primarily employed as educational tools to illustrate conceptual relationships rather than as replacements for the conventional layout in laboratory or reference use, due to their departure from the familiar grid format.[108][107][109]
Theoretical and visual alternatives
Theoretical and visual alternatives to the traditional tabular periodic table seek to represent the continuity of atomic numbers and periodic trends in more dynamic forms, often incorporating three-dimensional structures or spiral arrangements to highlight relationships that flat grids obscure. One prominent 3D model is the Alexander Arrangement of Elements (AAE), developed by Roy Alexander in 1965 and patented in 1974, which arranges elements in a helical cylinder resembling a globe.[110] This design stacks periods vertically while allowing transition and inner transition elements to form loops that protrude from the main helix, ensuring elements follow atomic number order without artificial breaks between groups.[111] Similarly, pyramidal 3D representations, such as William Jensen's step-pyramid form introduced in 1987, stack periods in ascending layers to emphasize shell and subshell relationships, with each level corresponding to principal quantum numbers and steps reflecting orbital filling sequences.[112]Spiral arrangements offer another visual departure, portraying the periodic system as a continuous coil that radiates outward from hydrogen at the center. The Chemical Galaxy, devised by Philip J. Stewart in 2005 and inspired by earlier spiral concepts like Edgar Longman's 1951 design, positions elements by atomic number (Z) along tightly wound inner spirals for s-block and p-block elements, with looser outer coils for d- and f-blocks, evoking a galactic structure to underscore the expansive nature of the element sequence.[113] This format maintains periodicity through angular repetitions of similar properties while avoiding the tabular separation of noble gases and alkali metals.[114]Quantum-based alternatives reorganize elements according to electronic structure principles, such as orbital energies and effective nuclear charge (Zeff). The ADOMAH periodic table, created by Valery Tsimmerman in 2006 and derived from Charles Janet's 1928 left-step table, divides the system into four blocks aligned with azimuthal quantum numbers (l = 0 for s, 1 for p, 2 for d, 3 for f), ordering elements by the Madelung rule of subshell filling to reflect increasing orbital energy levels.[115] In this arrangement, Zeff—defined as the net positive charge experienced by valence electrons, calculated as Zeff = Z - σ where σ is the shielding constant—influences the progression, as higher Zeff correlates with tighter orbital binding and periodic contractions across blocks.[116] Such models prioritize subshell energies over strict period cuts, briefly referencing the Aufbau principle of sequential filling without detailing derivations.A more recent quantum-based alternative, proposed by Chunhai Lyu in 2025, focuses on highly charged ions and arranges them by the number of electrons rather than protons. This table identifies around 700 such ions suitable for advanced optical atomic clocks, grouping them by electron shells (rows) and subshells (columns) to highlight forbidden transitions for precise timekeeping applications.[117]These alternatives excel in visualizing global trends, such as the smooth increase in atomic radius and electronegativity along the continuous atomic number line, which tabular forms fragment, and they facilitate insights into quantum mechanical periodicity.[118] However, they pose challenges for memorization, as the non-linear layouts complicate quick reference to group or period positions compared to the standard grid.[119] Modern digital implementations, including interactive 3D holograms and rotatable spirals in software like ptable.com, mitigate these drawbacks by allowing user manipulation and layered views.[120]