You are on page 1of 13

Wiki Loves Monuments: Photograph a monument, help Wikipedia and

win!Learn more

Atomic mass
From Wikipedia, the free encyclopedia
Jump to navigationJump to search
Not to be confused with Standard atomic weight, Mass number, or Relative atomic
mass.
For the band Atomic Mass, see Def Leppard §  History.
This article needs to be updated. The reason given is: it needs to reflect
the 2019 redefinition of the SI base units, which came into effect on 20 May
2019. Please update this article to reflect recent events or newly available
information. (January 2020)

Stylized lithium-7 atom: 3 protons, 4 neutrons, and 3 electrons (total electrons are ~ 1⁄4300th of the mass of the
nucleus). It has a mass of 7.016 Da. Rare lithium-6 (mass of 6.015 Da) has only 3 neutrons, reducing the
atomic weight (average) of lithium to 6.941.

The atomic mass (ma or m) is the mass of an atom. Although the the atomic mass is


the SI unit of mass is kilogram (symbol: kg), the atomic mass is often expressed in the
non-SI unit dalton (symbol: Da, or u) where 1 dalton is defined as 1⁄12 of the mass of a
single carbon-12 atom, at rest.[1] The protons and neutrons of the nucleus account for
nearly all of the total mass of atoms, with the electrons and nuclear binding
energy making minor contributions. Thus, the numeric value of the atomic mass when
expressed in daltons has nearly the same value as the mass number. Conversion
between mass in kilograms and mass in daltons can be done using the atomic mass
constant .
The formula used for conversion is:[2][3]
where  is the molar mass constant,  is the Avogadro
constant and  is the experimentally determined molar
mass of carbon-12.
The relative isotopic mass (see section below) can be
obtained by dividing the atomic mass ma of an isotope by
the atomic mass constant mu yielding a dimensionless
value. Thus, the atomic mass of a carbon-12 atom is
12 Da, but the relative isotopic mass of a carbon-12 atom is
simply 12. The sum of relative isotopic masses of all atoms
in a molecule is the relative molecular mass.
The atomic mass of an isotope and the relative isotopic
mass refers to a certain specific isotope of an element.
Because substances are usually not isotopically pure, it is
convenient to use the elemental atomic mass which is the
average (mean) atomic mass of an element, weighted by
the abundance of the isotopes. The dimensionless
(standard) atomic weight is the weighted mean relative
isotopic mass of a (typical naturally-occurring) mixture of
isotopes.
The atomic mass of atoms, ions, or atomic nuclei is slightly
less than the sum of the masses of their constituent
protons, neutrons, and electrons, due to binding energy
mass loss (as per E = mc2).

Contents

 1Relative isotopic mass


 2Similar terms for different quantities
 3Mass defects in atomic masses
 4Measurement of atomic masses
 5Relationship between atomic and molecular masses
 6History
 7See also
 8References
 9External links

Relative isotopic mass[edit]


Relative isotopic mass (a property of a single atom) is not
to be confused with the averaged quantity atomic
weight (see above), that is an average of values for many
atoms in a given sample of a chemical element.
While atomic mass is an absolute mass, relative isotopic
mass is a dimensionless number with no units. This loss of
units results from the use of a scaling ratio with respect to a
carbon-12 standard, and the word "relative" in the term
"relative isotopic mass" refers to this scaling relative to
carbon-12.
The relative isotopic mass, then, is the mass of a given
isotope (specifically, any single nuclide), when this value is
scaled by the mass of carbon-12, where the latter has to be
determined experimentally. Equivalently, the relative
isotopic mass of an isotope or nuclide is the mass of the
isotope relative to 1/12 of the mass of a carbon-12 atom.
For example, the relative isotopic mass of a carbon-12
atom is exactly 12. For comparison, the atomic mass of a
carbon-12 atom is exactly 12 daltons. Alternately, the
atomic mass of a carbon-12 atom may be expressed in any
other mass units: for example, the atomic mass of a
carbon-12 atom is about 1.998467052×10−26 kg.
As is the case for the related atomic mass when expressed
in daltons, the relative isotopic mass numbers of nuclides
other than carbon-12 are not whole numbers, but are
always close to whole numbers. This is discussed more
fully below.

Similar terms for different


quantities[edit]
The atomic mass or relative isotopic mass are sometimes
confused, or incorrectly used, as synonyms of relative
atomic mass (also known as atomic weight) or the standard
atomic weight (a particular variety of atomic weight, in the
sense that it is standardized). However, as noted in the
introduction, atomic mass is an absolute mass while all
other terms are dimensionless. Relative atomic mass and
standard atomic weight represent terms for (abundance-
weighted) averages of relative atomic masses in elemental
samples, not for single nuclides. As such, relative atomic
mass and standard atomic weight often differ numerically
from the relative isotopic mass.
The atomic mass (relative isotopic mass) is defined as the
mass of a single atom, which can only be
one isotope (nuclide) at a time, and is not an abundance-
weighted average, as in the case of relative atomic
mass/atomic weight. The atomic mass or relative isotopic
mass of each isotope and nuclide of a chemical element is,
therefore, a number that can in principle be measured to
high precision, since every specimen of such a nuclide is
expected to be exactly identical to every other specimen,
as all atoms of a given type in the same energy state, and
every specimen of a particular nuclide, are expected to be
exactly identical in mass to every other specimen of that
nuclide. For example, every atom of oxygen-16 is expected
to have exactly the same atomic mass (relative isotopic
mass) as every other atom of oxygen-16.
In the case of many elements that have one naturally
occurring isotope (mononuclidic elements) or one dominant
isotope, the difference between the atomic mass of the
most common isotope, and the (standard) relative atomic
mass or (standard) atomic weight can be small or even nil,
and does not affect most bulk calculations. However, such
an error can exist and even be important when considering
individual atoms for elements that are not mononuclidic.
For non-mononuclidic elements that have more than one
common isotope, the numerical difference in relative atomic
mass (atomic weight) from even the most common relative
isotopic mass, can be half a mass unit or more (e.g. see
the case of chlorine where atomic weight and standard
atomic weight are about 35.45). The atomic mass (relative
isotopic mass) of an uncommon isotope can differ from the
relative atomic mass, atomic weight, or standard atomic
weight, by several mass units.
Relative isotopic masses are always close to whole-
number values, but never (except in the case of carbon-12)
exactly a whole number, for two reasons:

 protons and neutrons have different masses, and


different nuclides have different ratios of protons
and neutrons.
 atomic masses are reduced, to different extents,
by their binding energies.
The ratio of atomic mass to mass number (number of
nucleons) varies from about 0.99884 for 56Fe to 1.00782505
for 1H.
Any mass defect due to nuclear binding energy is
experimentally a small fraction (less than 1%) of the mass
of an equal number of free nucleons. When compared to
the average mass per nucleon in carbon-12, which is
moderately strongly-bound compared with other atoms, the
mass defect of binding for most atoms is an even smaller
fraction of a dalton (unified atomic mass unit, based on
carbon-12). Since free protons and neutrons differ from
each other in mass by a small fraction of a dalton (about
0.0014 Da), rounding the relative isotopic mass, or the
atomic mass of any given nuclide given in daltons to the
nearest whole number always gives the nucleon count, or
mass number. Additionally, the neutron count (neutron
number) may then be derived by subtracting the number of
protons (atomic number) from the mass number (nucleon
count).

Mass defects in atomic masses[edit]

Binding energy per nucleon of common isotopes. A graph of the ratio of


mass number to atomic mass would be similar.

The amount that the ratio of atomic masses to mass


number deviates from 1 is as follows: the deviation starts
positive at hydrogen-1, then decreases until it reaches a
local minimum at helium-4. Isotopes of lithium, beryllium,
and boron are less strongly bound than helium, as shown
by their increasing mass-to-mass number ratios.
At carbon, the ratio of mass (in daltons) to mass number is
defined as 1, and after carbon it becomes less than one
until a minimum is reached at iron-56 (with only slightly
higher values for iron-58 and nickel-62), then increases to
positive values in the heavy isotopes, with increasing
atomic number. This corresponds to the fact that nuclear
fission in an element heavier than zirconium produces
energy, and fission in any element lighter
than niobium requires energy. On the other hand, nuclear
fusion of two atoms of an element lighter
than scandium (except for helium) produces energy,
whereas fusion in elements heavier than calcium requires
energy. The fusion of two atoms of 4He yielding beryllium-
8 would require energy, and the beryllium would quickly fall
apart again. 4He can fuse with tritium (3H) or with 3He; these
processes occurred during Big Bang nucleosynthesis. The
formation of elements with more than seven nucleons
requires the fusion of three atoms of 4He in the triple alpha
process, skipping over lithium, beryllium, and boron to
produce carbon-12.
Here are some values of the ratio of atomic mass to mass
number:

Nuclid
Ratio of atomic mass to mass number
e

1
H 1.00782505

2
H 1.0070508885

3
H 1.0053497592

3
He 1.0053431064

4
He 1.0006508135

6
Li 1.0025204658

12
C 1

14
N 1.0002195718

16
O 0.9996821637

56
Fe 0.9988381696
210
Po 0.9999184462

232
Th 1.0001640315

238
U 1.0002133958

Measurement of atomic masses[edit]


Direct comparison and measurement of the masses of
atoms is achieved with mass spectrometry.

Relationship between atomic and


molecular masses[edit]
Similar definitions apply to molecules. One can compute
the molecular mass of a compound by adding the atomic or
nuclide masses (not the standard atomic weights) of its
constituent atoms (nuclides). Conversely, the molar
mass is usually computed from the standard atomic
weights (not the atomic or nuclide masses). Thus,
molecular mass and molar mass differ slightly in numerical
value and represent different concepts. Molecular mass is
the mass of a molecule, which is the sum of its constituent
atomic masses. Molar mass is an average of the masses of
the constituent molecules in a chemically pure but
isotopically heterogeneous ensemble. In both cases, the
multiplicity of the atoms (the number of times it occurs)
must be taken into account, usually by multiplication of
each unique mass by its multiplicity.

Molar mass of CH4

Numbe Total molar mass (g/mol)


Standard atomic weight
r or molecular weight (Da or g/mol)

C 12.011 1 12.011

H 1.008 4 4.032
CH4 16.043

Molecular mass of 12C1H4

Numbe
Nuclide mass Total molecular mass (Da or u)
r

12
C 12.00 1 12.00

1
H 1.007825 4 4.0313

CH4 16.0313

History[edit]
Main articles: History of chemistry and Unified atomic
mass unit
The first scientists to determine relative atomic masses
were John Dalton and Thomas Thomson between 1803
and 1805 and Jöns Jakob Berzelius between 1808 and
1826. Relative atomic mass (Atomic weight) was originally
defined relative to that of the lightest element, hydrogen,
which was taken as 1.00, and in the 1820s, Prout's
hypothesis stated that atomic masses of all elements would
prove to be exact multiples of that of hydrogen. Berzelius,
however, soon proved that this was not even approximately
true, and for some elements, such as chlorine, relative
atomic mass, at about 35.5, falls almost exactly halfway
between two integral multiples of that of hydrogen. Still
later, this was shown to be largely due to a mix of isotopes,
and that the atomic masses of pure isotopes, or nuclides,
are multiples of the hydrogen mass, to within about 1%.
In the 1860s, Stanislao Cannizzaro refined relative atomic
masses by applying Avogadro's law (notably at
the Karlsruhe Congress of 1860). He formulated a law to
determine relative atomic masses of elements: the different
quantities of the same element contained in different
molecules are all whole multiples of the atomic weight and
determined relative atomic masses and molecular masses
by comparing the vapor density of a collection of gases
with molecules containing one or more of the chemical
element in question.[4]
In the 20th century, until the 1960s, chemists and
physicists used two different atomic-mass scales. The
chemists used a "atomic mass unit" (amu) scale such that
the natural mixture of oxygen isotopes had an atomic mass
16, while the physicists assigned the same number 16 to
only the atomic mass of the most common oxygen isotope
(16O, containing eight protons and eight neutrons).
However, because oxygen-17 and oxygen-18 are also
present in natural oxygen this led to two different tables of
atomic mass. The unified scale based on carbon-12, 12C,
met the physicists' need to base the scale on a pure
isotope, while being numerically close to the chemists'
scale. This was adopted as the 'unified atomic mass unit'.
The current International System of Units (SI) primary
recommendation for the name of this unit is the dalton and
symbol 'Da'. The name 'unified atomic mass unit' and
symbol 'u' are recognized names and symbols for the same
unit.[5]
The term atomic weight is being phased out slowly and
being replaced by relative atomic mass, in most current
usage. This shift in nomenclature reaches back to the
1960s and has been the source of much debate in the
scientific community, which was triggered by the adoption
of the unified atomic mass unit and the realization that
weight was in some ways an inappropriate term. The
argument for keeping the term "atomic weight" was
primarily that it was a well understood term to those in the
field, that the term "atomic mass" was already in use (as it
is currently defined) and that the term "relative atomic
mass" might be easily confused with relative isotopic
mass (the mass of a single atom of a given nuclide,
expressed dimensionlessly relative to 1/12 of the mass of
carbon-12; see section above).
In 1979, as a compromise, the term "relative atomic mass"
was introduced as a secondary synonym for atomic weight.
Twenty years later the primacy of these synonyms was
reversed, and the term "relative atomic mass" is now the
preferred term.
However, the term "standard atomic weights" (referring to
the standardized expectation atomic weights of differing
samples) has not been changed,[6] because simple
replacement of "atomic weight" with "relative atomic mass"
would have resulted in the term "standard relative atomic
mass."

See also[edit]
 Atomic number
 Atomic mass unit
 Isotope
 Isotope geochemistry
 Molecular mass
 Jean Stas

References[edit]
1. ^ IUPAC, Compendium of Chemical Terminology, 2nd
ed. (the "Gold Book") (1997). Online corrected version:
(2006–) "atomic mass". doi:10.1351/goldbook.A00496
2. ^ The International System of Units (SI). v1.06 (9 ed.).
Paris: Bureau International des Poids et Mesures.
2019.  ISBN  978-92-822-2272-0.
3. ^ Peter J. Mohr, Barry N. Taylor (May 20, 2019).  "NIST
Standard Reference Database 121. Fundamental
Physical Constants. atomic mass constant".  The NIST
reference on constants, Units and Uncertainty. National
Institute of Standards and Technology.
Retrieved December 10,  2019.
4. ^ Williams, Andrew (2007). "Origin of the Formulas of
Dihydrogen and Other Simple Molecules".  J. Chem.
Educ.  84  (11):
1779.  Bibcode:2007JChEd..84.1779W. doi:10.1021/ed0
84p1779.
5. ^ Bureau International des Poids et Mesures (2019): The
International System of Units (SI), 9th edition, English
version, page 134. Available at the BIPM website.
6. ^ De Bievre, P.; Peiser, H. S. (1992). "'Atomic weight':
The name, its history, definition, and units"  (PDF).  Pure
Appl. Chem.  64  (10):
1535.  doi:10.1351/pac199264101535.

External links[edit]
 NIST relative atomic masses of all isotopes and
the standard atomic weights of the elements
 AME2003 Atomic Mass Evaluation from
the National Nuclear Data Center
hide
Mole concepts

Avogadro constant

Boltzmann constant

Gas constant

Mass concentration

Molar concentration

Molality

Mass

Volume

Density

Mole fraction

Mass fraction

Amount of substance

Molar mass

Atomic mass

Particle number

Pressure

Thermodynamic temperature

Molar volume

Specific volume

Charles's law

Boyle's law

Gay-Lussac's law

Combined gas law

Avogadro's law

Ideal gas law

BNF: cb11977610z (data)

GND: 4143329-4

LCCN: sh85009318

NDL: 00562390
Categories: 
 Atoms
 Mass
 Chemical properties
 Stoichiometry
Navigation menu
 Not logged in
 Talk
 Contributions
 Create account
 Log in
 Article
 Talk
 Read
 Edit
 View history
Search
Search Go

 Main page
 Contents
 Current events
 Random article
 About Wikipedia
 Contact us
 Donate
Contribute
 Help
 Learn to edit
 Community portal
 Recent changes
 Upload file
Tools
 What links here
 Related changes
 Special pages
 Permanent link
 Page information
 Cite this page
 Wikidata item
Print/export
 Download as PDF
 Printable version
Languages
 ‫العربية‬
 Español
 हिन्दी
 Bahasa Indonesia
 Lumbaart
 Русский
 ‫اردو‬
 Winaray
 中文
82 more
Edit links
 This page was last edited on 29 October 2020, at 09:56 (UTC).
 Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By
using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of
the Wikimedia Foundation, Inc., a non-profit organization.
 Privacy policy

 About Wikipedia

 Disclaimers

 Contact Wikipedia

 Mobile view

 Developers

 Statistics

 Cookie statement

You might also like