Вы находитесь на странице: 1из 15

ation Energy Graph

This is a graph of ionization energy versus element atomic number. This graph displays the periodic trend of
ionization energy.

Ionization energy
From Wikipedia, the free encyclopedia

This article includes a list of references, but its sources remain unclear because it
has insufficient inline citations. Please help to improve this article by introducingmore precise
citations. (December 2008)

Periodic trend for ionization energy. Each period begins at a minimum for the alkali metals, and ends at a maximum for the
noble gases.

The ionization energy (IE) of an atomor molecule describes the minimum amount of energy required to
remove an electron (to infinity) from the atom or molecule in the gaseous state.
X + energy X+ + e
The term ionization potential has been used in the past but is not recommended. [1]
The units for ionization energy vary from discipline to discipline. In physics, the ionization energy is
typically specified in electron volts (eV) and refers to the energy required to remove a single electron from
a single atom or molecule. In chemistry, the ionization energy is typically specified as a molar quantity
(molar ionization energy or enthalpy) and is reported in units of kJ/mol or kcal/mol (the amount of energy it
takes for all the atoms in a mole to lose one electron each [2]).
The nth ionization energy refers to the amount of energy required to remove an electron from the species
with a charge of (n-1). For example, the first three ionization energies are defined as follows:
1st ionization energy
X X + + e
2nd ionization energy
X+ X2+ + e
3rd ionization energy
X2+ X3+ + e
Contents
[hide]

1 Values and trends

2 Electrostatic explanation

3 Quantum-mechanical explanation

4 Vertical and adiabatic ionization energy in molecules

5 Analogs of Ionization Energy to Other Systems

6 See also

7 References

Values and trends[edit]

Main articles: Molar ionization energies of the elements andIonization energies of the
elements (data page)
Generally the (n+1)th ionization energy is larger than the nth ionization energy. When the
next ionization energy involves removing an electron from the same electron shell, the
increase in ionization energy is primarily due to the increased net charge of the ion from
which the electron is being removed. Electrons removed from more highly charged ions of a
particular element experience greater forces of electrostatic attraction; thus, their removal
requires more energy. In addition, when the next ionization energy involves removing an
electron from a lower electron shell, the greatly decreased distance between the nucleus and
the electron also increases both the electrostatic force and the distance over which that force
must be overcome to remove the electron. Both of these factors further increase the
ionization energy.
Some values for elements of the third period are given in the following table:

Successive molar ionization energies in kJ/mol


(96.485 kJ/mol = 1 eV/particle)

Element

First

Second

Third

Fourth

Fifth

Sixth

Seventh

Na

496

4,560

Mg

738

1,450

7,730

Al

577

1,816

2,881

11,600

Si

786

1,577

3,228

4,354

16,100

1,060

1,890

2,905

4,950

6,270

21,200

999.6

2,260

3,375

4,565

6,950

8,490

27,107

Cl

1,256

2,295

3,850

5,160

6,560

9,360

11,000

Ar

1,520

2,665

3,945

5,770

7,230

8,780

12,000

Large jumps in the successive molar ionization energies occur when passing noble
gas configurations. For example, as can be seen in the table above, the first two molar
ionization energies of magnesium (stripping the two 3s electrons from a magnesium atom)
are much smaller than the third, which requires stripping off a 2p electron from
the neon configuration of Mg2+. That electron is much closer to the nucleus than the previous
3s electron.
Ionization energy is also a periodic trend within the periodic table organization. Moving left to
right within a period or upward within a group, the first ionization energy generally increases
with a few discrepancies (aluminum and sulphur). As the nuclear charge of the nucleus

increases across the period, the atomic radius decreases and the electron cloud becomes
closer towards the nucleus.
Ionization energy increases from left to right in a period and decreases from top to bottom in
a group.

Electrostatic explanation[edit]
Atomic ionization energy can be predicted by an analysis using electrostatic potential and
the Bohr model of the atom, as follows (note that the derivation uses Gaussian units).
Consider an electron of charge -e and an atomic nucleus with charge +Ze, where Z is the
number of protons in the nucleus. According to the Bohr model, if the electron were to
approach and bond with the atom, it would come to rest at a certain radius a. The
electrostatic potential V at distance a from the ionic nucleus, referenced to a point infinitely
far away, is:

Since the electron is negatively charged, it is drawn inwards by this positive electrostatic
potential. The energy required for the electron to "climb out" and leave the atom is:

This analysis is incomplete, as it leaves the distance a as an unknown variable. It can be


made more rigorous by assigning to each electron of every chemical element a characteristic
distance, chosen so that this relation agrees with experimental data.
It is possible to expand this model considerably by taking a semi-classical approach, in which
momentum is quantized. This approach works very well for the hydrogen atom, which only
has one electron. The magnitude of the angular momentum for a circular orbit is:

The total energy of the atom is the sum of the kinetic and potential energies, that is:

Velocity can be eliminated from the kinetic energy term by setting the Coulomb attraction
equal to the centripetal force, giving:

Solving the angular momentum for v and substituting this into the expression for kinetic
energy, we have:

This establishes the dependence of the radius on n. That is:

Now the energy can be found in terms of Z, e, and r. Using the new value for the kinetic
energy in the total energy equation above, it is found that:

At its smallest value, n is equal to 1 and r is the Bohr radius a0 which equals to

. Now,

the equation for the energy can be established in terms of the Bohr radius. Doing so gives
the result:

Quantum-mechanical explanation[edit]
According to the more complete theory of quantum mechanics, the location of an electron is
best described as a probability distribution. The energy can be calculated by integrating over
this cloud. The cloud's underlying mathematical representation is the wavefunction which is
built from Slater determinants consisting of molecular spin orbitals. These are related
by Pauli's exclusion principle to the antisymmetrized products of the atomic or molecular
orbitals.

In general, calculating the nth ionization energy requires calculating the energies
of

and

electron systems. Calculating these energies exactly is not

possible except for the simplest systems (i.e. hydrogen), primarily because of difficulties in
integrating the electron correlation terms. Therefore, approximation methods are routinely
employed, with different methods varying in complexity (computational time) and in accuracy
compared to empirical data. This has become a well-studied problem and is routinely done
in computational chemistry. At the lowest level of approximation, the ionization energy is
provided by Koopmans' theorem.

Vertical and adiabatic ionization energy in


molecules[edit]

Figure 1. FranckCondon principle energy diagram. For ionization of a diatomic molecule the only
nuclear coordinate is the bond length. The lower curve is the potential energy curve of the neutral
molecule, and the upper curve is for the positive ion with a longer bond length. The blue arrow is vertical
ionization, here from the ground state of the molecule to the v=2 level of the ion.

Ionization of molecules often leads to changes in molecular geometry, and two types of (first)
ionization energy are defined adiabatic and vertical.[3]
Adiabatic ionization energy: The adiabatic ionization energy of a molecule is
the minimum amount of energy required to remove an electron from a neutral molecule, i.e.
the difference between the energy of the vibrationalground state of the neutral species (v" =

0 level) and that of the positive ion (v' = 0). The specific equilibrium geometry of each species
does not affect this value.
Vertical ionization energy: Due to the possible changes in molecular geometry that may
result from ionization, additional transitions may exist between the vibrational ground state of
the neutral species and vibrationalexcited states of the positive ion. In other words, ionization
is accompanied by vibrational excitation. The intensity of such transitions are explained by
the FranckCondon principle, which predicts that the most probable and intense transition
corresponds to the vibrational excited state of the positive ion that has the same geometry as
the neutral molecule. This transition is referred to as the "vertical" ionization energy since it is
represented by a completely vertical line on a potential energy diagram (see Figure).
For a diatomic molecule, the geometry is defined by the length of a singlebond. The removal
of an electron from a bonding molecular orbitalweakens the bond and increases the bond
length. In Figure 1, the lowerpotential energy curve is for the neutral molecule and the upper
surface is for the positive ion. Both curves plot the potential energy as a function of bond
length. The horizontal lines correspond tovibrational levels with their associated vibrational
wave functions. Since the ion has a weaker bond, it will have a longer bond length. This
effect is represented by shifting the minimum of the potential energy curve to the right of the
neutral species. The adiabatic ionization is the diagonal transition to the vibrational ground
state of the ion. Vertical ionization involves vibrational excitation of the ionic state and
therefore requires greater energy.
In many circumstances, the adiabatic ionization energy is often a more interesting physical
quantity since it describes the difference in energy between the two potential energy
surfaces. However, due to experimental limitations, the adiabatic ionization energy is often
difficult to determine, whereas the vertical detachment energy is easily identifiable and
measurable.

Analogs of Ionization Energy to Other Systems[edit]


While the term ionization energy is largely used only for gas-phase atomic or molecular
species, there are a number of analogous quantities that consider the amount of energy
required to remove an electron from other physical systems.
Electron binding energy: A generic term for the ionization energy that can be used for
species with any charge state. For example, the electron binding energy for the chloride ion
is the minimum amount of energy required to remove an electron from the chlorine atom
when it has a charge of -1. In this particular example, the electron binding energy has the
same magnitude as the electron affinity for the neutral chlorine atom. In another example, the

electron binding energy refers the minimum amount of energy required to remove an
electron from the dicarboxylate dianion O2C(CH2)8CO2.
Work function: The minimum amount of energy required to remove an electron from a solid
surface.

See also[edit]

Second, Third, Fourth, and Higher Ionization Energies


By now you know that sodium forms Na+ ions, magnesium forms Mg2+ ions, and
aluminum forms Al3+ ions. But have you ever wondered why sodium doesn't form
Na2+ ions, or even Na3+ ions? The answer can be obtained from data for the second,
third, and higher ionization energies of the element.
The first ionization energy of sodium, for example, is the energy it takes to remove
one electron from a neutral atom.
Na(g) + energy

Na+(g) + e-

The second ionization energy is the energy it takes to remove another electron to form
an Na2+ ion in the gas phase.
Na+(g) + energy

Na2+(g) + e-

The third ionization energy can be represented by the following equation.


Na2+(g) + energy

Na3+(g) + e-

The energy required to form a Na3+ ion in the gas phase is the sum of the first, second,
and third ionization energies of the element.
First, Second, Third, and Fourth Ionization Energies
of Sodium, Magnesium, and Aluminum (kJ/mol)

It doesn't take much energy to remove one electron from a sodium atom to form an
Na+ ion with a filled-shell electron configuration. Once this is done, however, it takes
almost 10 times as much energy to break into this filled-shell configuration to remove
a second electron. Because it takes more energy to remove the second electron than is
given off in any chemical reaction, sodium can react with other elements to form
compounds that contain Na+ ions but not Na2+ or Na3+ ions.
A similar pattern is observed when the ionization energies of magnesium are analyzed.
The first ionization energy of magnesium is larger than sodium because magnesium
has one more proton in its nucleus to hold on to the electrons in the 3s orbital.
Mg: [Ne] 3s2
The second ionization energy of Mg is larger than the first because it always takes
more energy to remove an electron from a positively charged ion than from a neutral
atom. The third ionization energy of magnesium is enormous, however, because the
Mg2+ ion has a filled-shell electron configuration.
The same pattern can be seen in the ionization energies of aluminum. The first
ionization energy of aluminum is smaller than magnesium. The second ionization
energy of aluminum is larger than the first, and the third ionization energy is even
larger. Although it takes a considerable amount of energy to remove three electrons
from an aluminum atom to form an Al3+ ion, the energy needed to break into the
filled-shell configuration of the Al3+ ion is astronomical. Thus, it would be a mistake
to look for an Al4+ ion as the product of a chemical reaction.
Practice Problem 5:
Predict the group in the periodic table in which an element with the following ionization
energies would most likely be found.
1st IE = 786 kJ/mol
2nd IE = 1577
3rd IE = 3232
4th IE = 4355
5th IE = 16,091

6th IE = 19,784
Click here to check your answer to Practice Problem 5

Practice Problem 6:
Use the trends in the ionization energies of the elements to explain the following
observations.
(a) Elements on the left side of the periodic table are more likely than those on the right to
form positive ions.
(b) The maximum positive charge on an ion is equal to the group number of the element
Click here to check your answer to Practice Problem 6

Electron Affinity
Ionization energies measure the tendency of a neutral atom to resist the loss of
electrons. It takes a considerable amount of energy, for example, to remove an
electron from a neutral fluorine atom to form a positively charged ion.
F(g)

F+(g) + e-

Ho = 1681.0 kJ/mol

The electron affinity of an element is the energy given off when a neutral atom in the
gas phase gains an extra electron to form a negatively charged ion. A fluorine atom in
the gas phase, for example, gives off energy when it gains an electron to form a
fluoride ion.
F(g) + e-

F-(g)

Ho = -328.0 kJ/mol

Electron affinities are more difficult to measure than ionization energies and are
usually known to fewer significant figures. The electron affinities of the main group
elements are shown in the figure below.

Several patterns can be found in these data.


Electron affinities generally become smaller as we go down a column of the
periodic table for two reasons. First, the electron being added to the atom is
placed in larger orbitals, where it spends less time near the nucleus of the atom.
Second, the number of electrons on an atom increases as we go down a column,
so the force of repulsion between the electron being added and the electrons
already present on a neutral atom becomes larger.
Electron affinity data are complicated by the fact that the repulsion between the
electron being added to the atom and the electrons already present on the atom
depends on the volume of the atom. Among the nonmetals in Groups VIA and
VIIA, this force of repulsion is largest for the very smallest atoms in these
columns: oxygen and fluorine. As a result, these elements have a smaller
electron affinity than the elements below them in these columns as shown in the
figure below. From that point on, however, the electron affinities decrease as
we continue down these columns.

At first glance, there appears to be no pattern in electron affinity across a row of the
periodic table, as shown in the figure below.

When these data are listed along with the electron configurations of these elements,
however, they make sense. These data can be explained by noting that electron
affinities are much smaller than ionization energies. As a result, elements such as
helium, beryllium, nitrogen, and neon, which have unusually stable electron
configurations, have such small affinities for extra electrons that no energy is given
off when a neutral atom of these elements picks up an electron. These configurations
are so stable that it actually takes energy to force one of these elements to pick up an
extra electron to form a negative ion.
Electron Affinities and Electron Configurations for the First 10 Elements in the
Periodic Table
Element

Electron Affinity (kJ/mol)

Electron Configuration

H
He
Li
Be
B
C
N
O
F
Ne

1s1
1s2
[He] 2s1
[He] 2s2
[He] 2s2 2p1
[He] 2s2 2p2
[He] 2s2 2p3
[He] 2s2 2p4
[He] 2s2 2p5
[He] 2s2 2p6

72.8
<0
59.8
<0
27
122.3
<0
141.1
328.0
<0

Consequences of the Relative Size of Ionization Energies and Electron Affinities


Students often believe that sodium reacts with chlorine to form Na + and Cl- ions
because chlorine atoms "like" electrons more than sodium atoms do. There is no doubt
that sodium reacts vigorously with chlorine to form NaCl.
2 Na(s) + Cl2(g)

2 NaCl(s)

Furthermore, the ease with which solutions of NaCl in water conduct electricity is
evidence for the fact that the product of this reaction is a salt, which contains Na +and
Cl- ions.
NaCl(s)

Na+(aq)

H2O
+ Cl-(aq)

The only question is whether it is legitimate to assume that this reaction occurs
because chlorine atoms "like" electrons more than sodium atoms.
The first ionization energy for sodium is one and one-half times larger than the
electron affinity for chlorine.
Na: 1st IE = 495.8 kJ/mol
Cl: EA = 328.8 kJ/mol

Thus, it takes more energy to remove an electron from a neutral sodium atom than is
given off when the electron is picked up by a neutral chlorine atom. We will obviously
have to find another explanation for why sodium reacts with chlorine to form NaCl.
Before we can do this, however, we need to know more about the chemistry of ionic
compounds.

Вам также может понравиться