Which two chemical elements comprise approximately 99% of the matter in the universe?

Hydrogen is an element, usually in the form of a gas, that consists of one proton and one electron. Hydrogen is the most abundant element in the universe, accounting for about 75 percent of its normal matter, and was created in the Big Bang. Helium is an element, usually in the form of a gas, that consists of a nucleus of two protons and two neutrons surrounded by two electrons. Helium is the second-most abundant element in the universe, after hydrogen, and accounts for about 25 percent of the atoms in the universe. Most of the helium in the universe was created in the Big Bang, but it also is the product of hydrogen fusion in stars.

Which two chemical elements comprise approximately 99% of the matter in the universe?
Keep up with everything in the sky with a subscription to our daily radio program, hosted by Billy Henry.

Order Now

For more skywatching tips, astronomy news, and much more, read StarDate magazine.

Which two chemical elements comprise approximately 99% of the matter in the universe?

Which two chemical elements comprise approximately 99% of the matter in the universe?

Support StarDate radio by sponsoring our program for a day. You'll get your name on the air (or the name of someone else you would like to honor) and on our web site.

StarDate depends on the support of listeners and astronomy enthusiasts like you. Please consider making a tax-deductible contribution to StarDate.

Support StarDate

Which two chemical elements comprise approximately 99% of the matter in the universe?

Updated February 16, 2020

By Kevin Beck

Reviewed by: Lana Bandoim, B.S.

If asked to list the chemical elements that make up most of the Earth, you might be surprised by how hard it is to guess the right elements without knowing more than the average person does about all of that dirt, rock and metal underfoot. (You would also want to know if your quizzer was inquiring about the composition of the atmosphere above the Earth, also a common geoscience topic.)

Elements are different kinds of atoms, and as of 2020, 118 of them had been identified, 92 of which occur in significant amounts in nature. While it's impossible to know for certain the precise composition of the deeper layers of Earth, four elements alone make up almost 90 percent of the Earth's uppermost portion, or crust; four more account for nine-tenths of the remainder.

The more-or-less-spherical Earth, about 8,000 miles (just under 13,000 kilometers, or km) through the middle, is divided into three geologic layers:

  • A very thin crust extending down from the surface, about 3 miles (5 km) thick; 
  • a mantle about 1,800 miles

    (2,890 km)

    thick, composed mainly of magnesium and iron rocks;  * a core about 2,200 miles (3,400 km) thick and including a solid iron center surrounding a molten (hot liquid) "ring" of iron and nickel (see below).

Earth's crust is made up almost entirely of eight elements, four of them alone claiming almost nine-tenths of this layer's total weight: oxygen comes in at 46.6 percent by mass, followed by silicon at 27.7 percent, aluminum at 8.1 percent and iron at 5 percent.

  • The remaining crust elements are calcium, 3.6 percent; sodium, 2.8 percent, potassium, 2.6 percent, and magnesium, 2.1 percent.

Nearly 100 percent of the mass of Earth's core is composed of the two elements iron and nickel. Scientists have concluded this from the knowledge that the core must be 13 times as dense as water, leaving only some combination of molten iron and nickel as plausible candidates.

Oxygen: This familiar element, is number 6 on the periodic table of elements, makes up about 47 percent of the mass of the crust, and is also abundant (fortunately) in the atmosphere. It is also the main component, by mass, of water.

Since oxygen atoms are light compared to the other predominant elements in Earth's crust, the fact that their total mass accounts for almost half of the crust means that the fraction of its atoms that consists of oxygen is even higher than the mass fraction.

Silicon: This element, number 14 on the periodic table, exists as a crystal solid. It reacts poorly with most other elements, but eluded isolation by chemists for a long time owing to its affinity for oxygen and thus its tendency to remain "disguised" as silicon oxide.

  • Silicon is not to be confused with silicone, which is a polymer made of silicon, oxygen and other elements. It is commonly used in oil, grease and other physical media.

Aluminum: This metal is number 13 on the periodic table, one atomic number and thus one proton shy of silicon. It is non-magnetic and highly reactive, so much so that pure aluminum is rarely found. Instead, it is usually found embedded within rocks and combined with other compounds

Iron: Iron, atomic number 26, is a famous element, vital in both construction and engineering (iron accounts for almost all of the mass of most kinds of steel, for instance) and human physiology (iron is a required component of the oxygen-binding red blood cells, or erythrocytes, in your bloodstream). Is is found in all three of Earth's layers in significant amounts.

The most abundant element in the universe is hydrogen, which makes up about three-quarters of all matter! Helium makes up most of the remaining 25%. Oxygen is the third-most abundant element in the universe. All of the other elements are relatively rare.

The chemical composition of Earth is quite a bit different from that of the universe. The most abundant element in the Earth's crust is oxygen, making up 46.6% of Earth's mass. Silicon is the second most abundant element (27.7%), followed by aluminum (8.1%), iron (5.0%), calcium (3.6%), sodium (2.8%), potassium (2.6%). and magnesium (2.1%). These eight elements account for approximately 98.5% of the total mass of the Earth's crust. Of course, the Earth's crust is only the outer portion of the Earth. Future research will tell us about the composition of the mantle and core.

The most abundant element in the human body is oxygen, making up about 65% of the weight of each person. Carbon is the second-most abundant element, making up 18% of the body. Although you have more hydrogen atoms than any other type of element, the mass of a hydrogen atom is so much less than that of the other elements that its abundance comes in third at 10% by mass.

"Element Distribution in the Earth's Crust"

In order to continue enjoying our site, we ask that you confirm your identity as a human. Thank you very much for your cooperation.

The abundance of the chemical elements is a measure of the occurrence of the chemical elements relative to all other elements in a given environment. Abundance is measured in one of three ways: by the mass-fraction (the same as weight fraction); by the mole-fraction (fraction of atoms by numerical count, or sometimes fraction of molecules in gases); or by the volume-fraction. Volume-fraction is a common abundance measure in mixed gases such as planetary atmospheres, and is similar in value to molecular mole-fraction for gas mixtures at relatively low densities and pressures, and ideal gas mixtures. Most abundance values in this article are given as mass-fractions.

For example, the abundance of oxygen in pure water can be measured in two ways: the mass fraction is about 89%, because that is the fraction of water's mass which is oxygen. However, the mole-fraction is about 33% because only 1 atom of 3 in water, H2O, is oxygen. As another example, looking at the mass-fraction abundance of hydrogen and helium in both the Universe as a whole and in the atmospheres of gas-giant planets such as Jupiter, it is 74% for hydrogen and 23–25% for helium; while the (atomic) mole-fraction for hydrogen is 92%, and for helium is 8%, in these environments. Changing the given environment to Jupiter's outer atmosphere, where hydrogen is diatomic while helium is not, changes the molecular mole-fraction (fraction of total gas molecules), as well as the fraction of atmosphere by volume, of hydrogen to about 86%, and of helium to 13%.[Note 1]

The abundance of chemical elements in the universe is dominated by the large amounts of hydrogen and helium which were produced in the Big Bang. Remaining elements, making up only about 2% of the universe, were largely produced by supernovae and certain red giant stars. Lithium, beryllium, and boron, despite their low atomic number, are rare because, although they are produced by nuclear fusion, they are destroyed by other reactions in the stars.[1][2] The elements from carbon to iron are relatively more abundant in the universe because of the ease of making them in supernova nucleosynthesis. Elements of higher atomic number than iron (element 26) become progressively rarer in the universe, because they increasingly absorb stellar energy in their production. Also, elements with even atomic numbers are generally more common than their neighbors in the periodic table, due to favorable energetics of formation.

The abundance of elements in the Sun and outer planets is similar to that in the universe. Due to solar heating, the elements of Earth and the inner rocky planets of the Solar System have undergone an additional depletion of volatile hydrogen, helium, neon, nitrogen, and carbon (which volatilizes as methane). The crust, mantle, and core of the Earth show evidence of chemical segregation plus some sequestration by density. Lighter silicates of aluminium are found in the crust, with more magnesium silicate in the mantle, while metallic iron and nickel compose the core. The abundance of elements in specialized environments, such as atmospheres, or oceans, or the human body, are primarily a product of chemical interactions with the medium in which they reside.

Ten most common elements in the Milky Way Galaxy estimated spectroscopically[3]
Z Element Mass fraction
(ppm)
1 Hydrogen 739,000
2 Helium 240,000
8 Oxygen 10,400
6 Carbon 4,600
10 Neon 1,340
26 Iron 1,090
7 Nitrogen 960
14 Silicon 650
12 Magnesium 580
16 Sulfur 440
Total 999,060

The elements – that is, ordinary (baryonic) matter made of protons, neutrons, and electrons, are only a small part of the content of the Universe. Cosmological observations suggest that only 4.6% of the universe's energy (including the mass contributed by energy, E = mc2 ↔ m = E / c2) comprises the visible baryonic matter that constitutes stars, planets, and living beings. The rest is thought to be made up of dark energy (68%) and dark matter (27%).[4] These are forms of matter and energy believed to exist on the basis of scientific theory and inductive reasoning based on observations, but they have not been directly observed and their nature is not well understood.

Most standard (baryonic) matter is found in intergalactic gas, stars, and interstellar clouds, in the form of atoms or ions (plasma), although it can be found in degenerate forms in extreme astrophysical settings, such as the high densities inside white dwarfs and neutron stars.

Hydrogen is the most abundant element in the Universe; helium is second. However, after this, the rank of abundance does not continue to correspond to the atomic number; oxygen has abundance rank 3, but atomic number 8. All others are substantially less common.

The abundance of the lightest elements is well predicted by the standard cosmological model, since they were mostly produced shortly (i.e., within a few hundred seconds) after the Big Bang, in a process known as Big Bang nucleosynthesis. Heavier elements were mostly produced much later, inside of stars.

Hydrogen and helium are estimated to make up roughly 74% and 24% of all baryonic matter in the universe respectively. Despite comprising only a very small fraction of the universe, the remaining "heavy elements" can greatly influence astronomical phenomena. Only about 2% (by mass) of the Milky Way galaxy's disk is composed of heavy elements.

These other elements are generated by stellar processes.[5][6][7] In astronomy, a "metal" is any element other than hydrogen or helium. This distinction is significant because hydrogen and helium are the only elements that were produced in significant quantities in the Big Bang. Thus, the metallicity of a galaxy or other object is an indication of stellar activity after the Big Bang.

In general, elements up to iron are made in large stars in the process of becoming supernovae. Iron-56 is particularly common, since it is the most stable nuclide (in that it has the highest nuclear binding energy per nucleon) and can easily be made from alpha particles (being a product of decay of radioactive nickel-56, ultimately made from 14 helium nuclei). Elements heavier than iron are made in energy-absorbing processes in large stars, and their abundance in the universe (and on Earth) generally decreases with increasing atomic number.

The table shows the ten most common elements in our galaxy (estimated spectroscopically), as measured in parts per million, by mass.[3] Nearby galaxies that have evolved along similar lines have a corresponding enrichment of elements heavier than hydrogen and helium. The more distant galaxies are being viewed as they appeared in the past, so their abundances of elements appear closer to the primordial mixture. Since physical laws and processes are uniform throughout the universe, however, it is expected that these galaxies will likewise have evolved similar abundances of elements.

The abundance of elements is in keeping with their origin from the Big Bang and nucleosynthesis in a number of progenitor supernova stars. Very abundant hydrogen and helium are products of the Big Bang, while the next three elements are rare since they had little time to form in the Big Bang and are not made in stars (they are, however, produced in small quantities by breakup of heavier elements in interstellar dust, as a result of impact by cosmic rays).

Beginning with carbon, elements have been produced in stars by buildup from alpha particles (helium nuclei), resulting in an alternatingly larger abundance of elements with even atomic numbers (these are also more stable). The effect of odd-numbered chemical elements generally being more rare in the universe was empirically noticed in 1914, and is known as the Oddo-Harkins rule.

The following graph (note log scale) shows abundance of elements in the Solar System.

 

Estimated abundances of the chemical elements in the Solar System (logarithmic scale)

Most abundant nuclides
in the Solar System[8]
Nuclide A Mass fraction in parts per million Atom fraction in parts per million
Hydrogen-1 1 705,700 909,964
Helium-4 4 275,200 88,714
Oxygen-16 16 9,592 477
Carbon-12 12 3,032 326
Nitrogen-14 14 1,105 102
Neon-20 20 1,548 100
 
Other nuclides: 3,616 172
Silicon-28 28 653 30
Magnesium-24 24 513 28
Iron-56 56 1,169 27
Sulfur-32 32 396 16
Helium-3 3 35 15
Hydrogen-2 2 23 15
Neon-22 22 208 12
Magnesium-26 26 79 4
Carbon-13 13 37 4
Magnesium-25 25 69 4
Aluminium-27 27 58 3
Argon-36 36 77 3
Calcium-40 40 60 2
Sodium-23 23 33 2
Iron-54 54 72 2
Silicon-29 29 34 2
Nickel-58 58 49 1
Silicon-30 30 23 1
Iron-57 57 28 1

 

Periodic table showing the cosmological origin of each element

Relation to nuclear binding energy

Loose correlations have been observed between estimated elemental abundances in the universe and the nuclear binding energy curve. Roughly speaking, the relative stability of various atomic nuclides has exerted a strong influence on the relative abundance of elements formed in the Big Bang, and during the development of the universe thereafter.[9] See the article about nucleosynthesis for an explanation of how certain nuclear fusion processes in stars (such as carbon burning, etc.) create the elements heavier than hydrogen and helium.

A further observed peculiarity is the jagged alternation between relative abundance and scarcity of adjacent atomic numbers in the elemental abundance curve, and a similar pattern of energy levels in the nuclear binding energy curve. This alternation is caused by the higher relative binding energy (corresponding to relative stability) of even atomic numbers compared with odd atomic numbers and is explained by the Pauli Exclusion Principle.[10] The semi-empirical mass formula (SEMF), also called Weizsäcker's formula or the Bethe-Weizsäcker mass formula, gives a theoretical explanation of the overall shape of the curve of nuclear binding energy.[11]

The Earth formed from the same cloud of matter that formed the Sun, but the planets acquired different compositions during the formation and evolution of the solar system. In turn, the natural history of the Earth caused parts of this planet to have differing concentrations of the elements.

The mass of the Earth is approximately 5.97×1024 kg. In bulk, by mass, it is composed mostly of iron (32.1%), oxygen (30.1%), silicon (15.1%), magnesium (13.9%), sulfur (2.9%), nickel (1.8%), calcium (1.5%), and aluminium (1.4%); with the remaining 1.2% consisting of trace amounts of other elements.[12]

The bulk composition of the Earth by elemental-mass is roughly similar to the gross composition of the solar system, with the major differences being that Earth is missing a great deal of the volatile elements hydrogen, helium, neon, and nitrogen, as well as carbon which has been lost as volatile hydrocarbons. The remaining elemental composition is roughly typical of the "rocky" inner planets, which formed in the thermal zone where solar heat drove volatile compounds into space. The Earth retains oxygen as the second-largest component of its mass (and largest atomic-fraction), mainly from this element being retained in silicate minerals which have a very high melting point and low vapor pressure.

Estimated abundances of chemical elements in the Earth.[13] The right two columns give the fraction of the mass in parts per million (ppm) and the fraction by number of atoms in parts per billion (ppb).

Crust

 

Abundance (atom fraction) of the chemical elements in Earth's upper continental crust as a function of atomic number. The rarest elements in the crust (shown in yellow) are rare due to a combination of factors: all but one are the densest siderophiles (iron-loving) elements in the Goldschmidt classification, meaning they have a tendency to mix well with metallic iron, depleting them by being relocated deeper into the Earth's core. Their abundance in meteoroids is higher. Additionally, tellurium has been depleted by preaccretional sorting in the nebula via formation of volatile hydrogen telluride.[14]

The mass-abundance of the nine most abundant elements in the Earth's crust is approximately: oxygen 46%, silicon 28%, aluminium 8.3%, iron 5.6%, calcium 4.2%, sodium 2.5%, magnesium 2.4%, potassium 2.0%, and titanium 0.61%. Other elements occur at less than 0.15%. For a complete list, see abundance of elements in Earth's crust.

The graph at right illustrates the relative atomic-abundance of the chemical elements in Earth's upper continental crust—the part that is relatively accessible for measurements and estimation.

Many of the elements shown in the graph are classified into (partially overlapping) categories:

  1. rock-forming elements (major elements in green field, and minor elements in light green field);
  2. rare earth elements (lanthanides (La–Lu), Sc, and Y; labeled in blue);
  3. major industrial metals (global production >~3×107 kg/year; labeled in red);
  4. precious metals (labeled in purple);
  5. the nine rarest "metals" – the six platinum group elements plus Au, Re, and Te (a metalloid) – in the yellow field. These are rare in the crust from being soluble in iron and thus concentrated in the Earth's core. Tellurium is the single most depleted element in the silicate Earth relative to cosmic abundance, because in addition to being concentrated as dense chalcogenides in the core it was severely depleted by preaccretional sorting in the nebula as volatile hydrogen telluride.[14]

Note that there are two breaks where the unstable (radioactive) elements technetium (atomic number 43) and promethium (atomic number 61) would be. These elements are surrounded by stable elements, yet their most stable isotopes have relatively short half lives (~4 million years and ~18 years respectively). These are thus extremely rare, since any primordial initial fractions of these in pre-Solar System materials have long since decayed. These two elements are now only produced naturally through the spontaneous fission of very heavy radioactive elements (for example, uranium, thorium, or the trace amounts of plutonium that exist in uranium ores), or by the interaction of certain other elements with cosmic rays. Both technetium and promethium have been identified spectroscopically in the atmospheres of stars, where they are produced by ongoing nucleosynthetic processes.

There are also breaks in the abundance graph where the six noble gases would be, since they are not chemically bound in the Earth's crust, and they are only generated in the crust by decay chains from radioactive elements, and are therefore extremely rare there.

The eight naturally occurring very rare, highly radioactive elements (polonium, astatine, francium, radium, actinium, protactinium, neptunium, and plutonium) are not included, since any of these elements that were present at the formation of the Earth have decayed away eons ago, and their quantity today is negligible and is only produced from the radioactive decay of uranium and thorium.

Oxygen and silicon are notably the most common elements in the crust. On Earth and in rocky planets in general, silicon and oxygen are far more common than their cosmic abundance. The reason is that they combine with each other to form silicate minerals.[14] Other cosmically-common elements such as hydrogen, carbon and nitrogen form volatile compounds such as ammonia and methane that easily boil away into space from the heat of planetary formation and/or the Sun's light.

Rare-earth elements

"Rare" earth elements is a historical misnomer. The persistence of the term reflects unfamiliarity rather than true rarity. The more abundant rare earth elements are similarly concentrated in the crust compared to commonplace industrial metals such as chromium, nickel, copper, zinc, molybdenum, tin, tungsten, or lead. The two least abundant rare earth elements (thulium and lutetium) are nearly 200 times more common than gold. However, in contrast to the ordinary base and precious metals, rare earth elements have very little tendency to become concentrated in exploitable ore deposits. Consequently, most of the world's supply of rare earth elements comes from only a handful of sources. Furthermore, the rare earth metals are all quite chemically similar to each other, and they are thus quite difficult to separate into quantities of the pure elements.

Differences in abundances of individual rare earth elements in the upper continental crust of the Earth represent the superposition of two effects, one nuclear and one geochemical. First, the rare earth elements with even atomic numbers (58Ce, 60Nd, ...) have greater cosmic and terrestrial abundances than the adjacent rare earth elements with odd atomic numbers (57La, 59Pr, ...). Second, the lighter rare earth elements are more incompatible (because they have larger ionic radii) and therefore more strongly concentrated in the continental crust than the heavier rare earth elements. In most rare earth ore deposits, the first four rare earth elements – lanthanum, cerium, praseodymium, and neodymium – constitute 80% to 99% of the total amount of rare earth metal that can be found in the ore.

Mantle

The mass-abundance of the seven most abundant elements in the Earth's mantle is approximately: oxygen 44.3%, magnesium 22.3%, silicon 21.3%, iron 6.32%, calcium 2.48%, aluminium 2.29%, nickel 0.19%.[15]

Core

Due to mass segregation, the core of the Earth is believed to be primarily composed of iron (88.8%), with smaller amounts of nickel (5.8%), sulfur (4.5%), and less than 1% trace elements.[12]

Ocean

The most abundant elements in the ocean by proportion of mass in percent are oxygen (85.84%), hydrogen (10.82%), chlorine (1.94%), sodium (1.08%), magnesium (0.13%), sulfur (0.09%), calcium (0.04%), potassium (0.04%), bromine (0.007%), carbon (0.003%), and boron (0.0004%).

Atmosphere

The order of elements by volume-fraction (which is approximately molecular mole-fraction) in the atmosphere is nitrogen (78.1%), oxygen (20.9%),[16] argon (0.96%), followed by (in uncertain order) carbon and hydrogen because water vapor and carbon dioxide, which represent most of these two elements in the air, are variable components. Sulfur, phosphorus, and all other elements are present in significantly lower proportions.

According to the abundance curve graph, argon, a significant if not major component of the atmosphere, does not appear in the crust at all. This is because the atmosphere has a far smaller mass than the crust, so argon remaining in the crust contributes little to mass-fraction there, while at the same time buildup of argon in the atmosphere has become large enough to be significant.

Urban soils

For a complete list of the abundance of elements in urban soils, see Abundances of the elements (data page)#Urban soils.

Elemental abundance in the human body

By mass, human cells consist of 65–90% water (H2O), and a significant portion of the remainder is composed of carbon-containing organic molecules. Oxygen therefore contributes a majority of a human body's mass, followed by carbon. Almost 99% of the mass of the human body is made up of six elements: hydrogen (H), carbon (C), nitrogen (N), oxygen (O), calcium (Ca), and phosphorus (P) . The next 0.75% is made up of the next five elements: potassium (K), sulfur (S), chlorine (Cl), sodium (Na), and magnesium (Mg). Only 17 elements are known for certain to be necessary to human life, with one additional element (fluorine) thought to be helpful for tooth enamel strength. A few more trace elements may play some role in the health of mammals. Boron and silicon are notably necessary for plants but have uncertain roles in animals. The elements aluminium and silicon, although very common in the earth's crust, are conspicuously rare in the human body.[17]

Below is a periodic table highlighting nutritional elements.[18]

Nutritional elements in the periodic table[19]
  • v
  • t
  • e

H   He
Li Be   B C N O F Ne
Na Mg   Al Si P S Cl Ar
K Ca   Sc Ti V Cr Mn Fe Co Ni Cu Zn Ga Ge As Se Br Kr
Rb Sr   Y Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn Sb Te I Xe
Cs Ba * Lu Hf Ta W Re Os Ir Pt Au Hg Tl Pb Bi Po At Rn
Fr Ra ** Lr Rf Db Sg Bh Hs Mt Ds Rg Cn Nh Fl Mc Lv Ts Og
 
  * La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb
  ** Ac Th Pa U Np Pu Am Cm Bk Cf Es Fm Md No

Legend:

  The four basic organic elements

  Quantity elements

  Essential trace elements

  Deemed essential trace element by U.S., not by European Union

  Suggested function from deprivation effects or active metabolic handling, but no clearly-identified biochemical function in humans

  Limited circumstantial evidence for trace benefits or biological action in mammals

  No evidence for biological action in mammals, but essential in some lower organisms.
(In the case of lanthanum, the definition of an essential nutrient as being indispensable and irreplaceable is not completely applicable due to the extreme similarity of the lanthanides. The stable early lanthanides up to Sm are known to stimulate the growth of various lanthanide-using organisms.)[20]

  • Abundances of the elements (data page)
  • Abundance of elements in Earth's crust
  • Natural abundance (isotopic abundance)
  • Goldschmidt classification
  • Primordial nuclide
  • Radiative levitation
  • List of data references for chemical elements

  1. ^ Vangioni-Flam, Elisabeth; Cassé, Michel (2012). Spite, Monique (ed.). Galaxy Evolution: Connecting the Distant Universe with the Local Fossil Record. Springer Science & Business Media. pp. 77–86. ISBN 978-9401142137.
  2. ^ Trimble, Virginia (1996). "The Origin and Evolution of the Chemical Elements". In Malkan, Matthew A.; Zuckerman, Ben (eds.). The origin and evolution of the universe. Sudbury, Mass.: Jones and Bartlett Publishers. p. 101. ISBN 0-7637-0030-4.
  3. ^ a b Croswell, Ken (February 1996). Alchemy of the Heavens. Anchor. ISBN 0-385-47214-5. Archived from the original on 2011-05-13.
  4. ^ What is Dark Energy? Archived 2016-01-15 at the Wayback Machine, Space.com, 1 May 2013.
  5. ^ Suess, Hans; Urey, Harold (1956). "Abundances of the Elements". Reviews of Modern Physics. 28 (1): 53. Bibcode:1956RvMP...28...53S. doi:10.1103/RevModPhys.28.53.
  6. ^ Cameron, A. G. W. (1973). "Abundances of the elements in the solar system". Space Science Reviews. 15 (1): 121. Bibcode:1973SSRv...15..121C. doi:10.1007/BF00172440. S2CID 120201972.
  7. ^ Anders, E.; Ebihara, M. (1982). "Solar-system abundances of the elements". Geochimica et Cosmochimica Acta. 46 (11): 2363. Bibcode:1982GeCoA..46.2363A. doi:10.1016/0016-7037(82)90208-3.
  8. ^ Arnett, David (1996). Supernovae and Nucleosynthesis (First ed.). Princeton, New Jersey: Princeton University Press. p. 11. ISBN 0-691-01147-8. OCLC 33162440.
  9. ^ Bell, Jerry A.; GenChem Editorial/Writing Team (2005). "Chapter 3: Origin of Atoms". Chemistry: a project of the American Chemical Society. New York [u.a.]: Freeman. pp. 191–193. ISBN 978-0-7167-3126-9. Correlations between abundance and nuclear binding energy [Subsection title]
  10. ^ Bell, Jerry A.; GenChem Editorial/Writing Team (2005). "Chapter 3: Origin of Atoms". Chemistry: a project of the American Chemical Society. New York [u.a.]: Freeman. p. 192. ISBN 978-0-7167-3126-9. The higher abundance of elements with even atomic numbers [Subsection title]
  11. ^ Bailey, David. "Semi-empirical Nuclear Mass Formula". PHY357: Strings & Binding Energy. University of Toronto. Archived from the original on 2011-07-24. Retrieved 2011-03-31.
  12. ^ a b Morgan, J. W.; Anders, E. (1980). "Chemical composition of Earth, Venus, and Mercury". Proceedings of the National Academy of Sciences. 77 (12): 6973–6977. Bibcode:1980PNAS...77.6973M. doi:10.1073/pnas.77.12.6973. PMC 350422. PMID 16592930.
  13. ^ a b William F McDonough The composition of the Earth. quake.mit.edu, archived by the Internet Archive Wayback Machine.
  14. ^ a b c Anderson, Don L.; ‘Chemical Composition of the Mantle’ in Theory of the Earth, pp. 147–175 ISBN 0865421234
  15. ^ Wang, Haiyang S.; Lineweaver, Charles H.; Ireland, Trevor R. (2018-01-01). "The elemental abundances (with uncertainties) of the most Earth-like planet". Icarus. 299: 460–474. doi:10.1016/j.icarus.2017.08.024. ISSN 0019-1035.
  16. ^ Zimmer, Carl (3 October 2013). "Earth's Oxygen: A Mystery Easy to Take for Granted". The New York Times. Archived from the original on 3 October 2013. Retrieved 3 October 2013.
  17. ^ Table data from Chang, Raymond (2007). Chemistry (Ninth ed.). McGraw-Hill. p. 52. ISBN 978-0-07-110595-8.
  18. ^ Nielsen, Forrest H. (1998). "Ultratrace minerals.". In Maurice E. Shils; James A. Olsen; Moshe Shine; A. Catharine Ross (eds.). Modern nutrition in health and disease. Baltimore: Lippincott Williams & Wilkins. pp. 283–303. hdl:10113/46493. ISBN 978-0683307696.
  19. ^ Ultratrace minerals. Authors: Nielsen, Forrest H. USDA, ARS Source: Modern nutrition in health and disease / editors, Maurice E. Shils ... et al. Baltimore: Williams & Wilkins, c1999., p. 283-303. Issue Date: 1999 URI: [1]
  20. ^ Daumann, Lena J. (25 April 2019). "Essential and Ubiquitous: The Emergence of Lanthanide Metallobiochemistry". Angewandte Chemie International Edition. doi:10.1002/anie.201904090. Retrieved 15 June 2019.

  1. ^ Below Jupiter's outer atmosphere, volume fractions are significantly different from mole fractions due to high temperatures (ionization and disproportionation) and high density where the Ideal Gas Law is inapplicable.

  • "Rare Earth Elements—Critical Resources for High Technology | USGS Fact Sheet 087-02". geopubs.wr.usgs.gov.
  • "Imagine the Universe! Dictionary". 3 December 2003. Archived from the original on 3 December 2003.
  • List of elements in order of abundance in the Earth's crust (only correct for the twenty most common elements)
  • Cosmic abundance of the elements and nucleosynthesis
  • WebElements.com Lists of elemental abundances for the Universe, Sun, meteorites, Earth, ocean, streamwater, etc.

  Earth sciences  Chemistry  Geography  Astronomy  Stars  Science

Retrieved from "https://en.wikipedia.org/w/index.php?title=Abundance_of_the_chemical_elements&oldid=1109229363"