Download as pdf or txt
Download as pdf or txt
You are on page 1of 120

Lanthanide

Lanthanides in the periodic table

The lanthanide /ˈlænθənaɪd/ or lanthanoid


/ˈlænθənɔɪd/ series of chemical
elements[1] comprises the fifteen metallic
chemical elements with atomic numbers
57 through 71, from lanthanum through
lutetium.[2][3][4] These fifteen lanthanide
elements, along with the chemically
similar elements scandium and yttrium,
are often collectively known as the rare
earth elements.

The informal chemical symbol Ln is used


in general discussions of lanthanide
chemistry to refer to any lanthanide. All
but one of the lanthanides are f-block
elements, corresponding to the filling of
the 4f electron shell; depending on the
source, either lanthanum or lutetium is
considered a d-block element, but is
included due to its chemical similarities
with the other fourteen.[5] All lanthanide
elements form trivalent cations, Ln3+,
whose chemistry is largely determined by
the ionic radius, which decreases steadily
from lanthanum to lutetium.

They are termed lanthanides because the


elements in the series are chemically
similar to lanthanum. Strictly speaking,
both lanthanum and lutetium have been
labeled as group 3 elements, because they
both have a single valence electron in the
5d shell. However, both elements are often
included in any general discussion of the
chemistry of the lanthanide elements.
Lanthanum is the more often omitted of
the two, because its placement as a group
3 element is somewhat more common in
texts and for semantic reasons: since
"lanthanide" means "like lanthanum", it has
been argued that lanthanum cannot
logically be a lanthanide, even though
IUPAC acknowledges its inclusion based
on common usage.[6]

In presentations of the periodic table, the


lanthanides and the actinides are
customarily shown as two additional rows
below the main body of the table,[2] with
placeholders or else a selected single
element of each series (either lanthanum
and actinium, or lutetium and lawrencium)
shown in a single cell of the main table,
between barium and hafnium, and radium
and rutherfordium, respectively. This
convention is entirely a matter of
aesthetics and formatting practicality; a
rarely used wide-formatted periodic table
inserts the lanthanide and actinide series
in their proper places, as parts of the
table's sixth and seventh rows (periods).

Lanth
Lan‐ Cerium Praseo‐ Neo‐ Prome‐Sama‐Europ
thanum dymiumdymium thium rium ium
57 58 59 60 61 62 63
La Ce Pr Nd Pm Sm Eu

Primordial From decay Synthetic Border shows na

Etymology
Together with scandium and yttrium, the
trivial name "rare earths" is sometimes
used to describe all the lanthanides. This
name arises from the minerals from which
they were isolated, which were uncommon
oxide-type minerals. However, the use of
the name is deprecated by IUPAC, as the
elements are neither rare in abundance nor
"earths" (an obsolete term for water-
insoluble strongly basic oxides of
electropositive metals incapable of being
smelted into metal using late 18th century
technology). Cerium is the 26th most
abundant element in the Earth's crust,
neodymium is more abundant than gold,
thulium (the second least common
naturally occurring lanthanide) is more
abundant than iodine,[7] which is itself
common enough for biology to have
evolved critical usages thereof, and even
the lone radioactive element in the series,
promethium, is more common than the
two rarest naturally occurring elements,
francium and astatine, combined. Despite
their abundance, even the technical term
"lanthanides" could be interpreted to
reflect a sense of elusiveness on the part
of these elements, as it comes from the
Greek λανθανειν (lanthanein), "to lie
hidden". However, if not referring to their
natural abundance, but rather to their
property of "hiding" behind each other in
minerals, this interpretation is in fact
appropriate. The etymology of the term
must be sought in the first discovery of
lanthanum, at that time a so-called new
rare earth element "lying hidden" in a
cerium mineral, and it is an irony that
lanthanum was later identified as the first
in an entire series of chemically similar
elements and could give name to the
whole series. The term "lanthanide" was
introduced by Victor Goldschmidt in
1925.[8]

Physical properties of the


elements
Chemical
La Ce Pr N
element

Atomic number 57 58 59 6

Image

Density (g/cm3) 6.162 6.770 6.77 7

Melting point
920 795 935 10
(°C)

Boiling point (°C) 3464 3443 3520 30

Atomic electron
configuration 5d1 4f15d1 4f3 4
(gas phase)*

Atomic electron
configuration 5d1 4f15d1 4f25d1 4f3
(solid phase)*

Metal lattice (RT) dhcp fcc dhcp dh


Metallic radius
162 181.8 182.4 18
(pm)

57–
Resistivity at
80 73 68 6
25 °C (μOhm·cm)
20 °C

Magnetic
susceptibility +2500 +5
+95.9 +5530(α)
χmol (β) (
/10−6(cm3·mol−1)

* Between initial Xe and final 6s2 electronic


shells

** Sm has a close packed structure like the


other lanthanides but has an unusual 9
layer repeat
Gschneider and Daane (1988) attribute the
trend in melting point which increases
across the series, (lanthanum (920 °C) –
lutetium (1622 °C)) to the extent of
hybridization of the 6s, 5d, and 4f orbitals.
The hybridization is believed to be at its
greatest for cerium, which has the lowest
melting point of all, 795 °C.[9] The
lanthanide metals are soft; their hardness
increases across the series.[6] Europium
stands out, as it has the lowest density in
the series at 5.24 g/cm3 and the largest
metallic radius in the series at 208.4 pm. It
can be compared to barium, which has a
metallic radius of 222 pm. It is believed
that the metal contains the larger Eu2+ ion
and that there are only two electrons in the
conduction band. Ytterbium also has a
large metallic radius, and a similar
explanation is suggested.[6] The
resistivities of the lanthanide metals are
relatively high, ranging from 29 to 134
μOhm·cm. These values can be compared
to a good conductor such as aluminium,
which has a resistivity of 2.655 μOhm·cm.
With the exceptions of La, Yb, and Lu
(which have no unpaired f electrons), the
lanthanides are strongly paramagnetic,
and this is reflected in their magnetic
susceptibilities. Gadolinium becomes
ferromagnetic at below 16 °C (Curie point).
The other heavier lanthanides – terbium,
dysprosium, holmium, erbium, thulium, and
ytterbium – become ferromagnetic at
much lower temperatures.[10]

Chemistry and compounds


Chemical
La Ce Pr
element

Atomic number 57 58 59

Ln3+ electron
4f0 4f1 4f2
configuration*[11]

Ln3+ radius
103 102 99
(pm)[6]

Ln4+ ion color in


Orange-
aqueous — Yellow
yellow
solution[12]

Ln3+ ion color in


aqueous Colorless Colorless Green
solution[11]

Ln2+ ion color in — — —


aqueous
solution[6]

* Not including initial [Xe] core

The colors of lanthanide complexes


originate almost entirely from charge
transfer interactions between the metal
and the ligand. f → f transitions are
symmetry forbidden (or Laporte-
forbidden), which is also true of transition
metals, however t-metals are able to use
vibronic coupling to break this rule. The
valence orbitals in lanthanides are almost
entirely non-bonding and as such little
effective vibronic coupling takes, hence
the spectra from f → f transitions are
much weaker and narrower than those
from d → d transitions. In general this
makes the colors of lanthanide complexes
far fainter than those of transition metal
complexes. f→f transitions are not
possible for the f1 and f13 configurations
of Ce3+ and Yb3+ and thus these ions are
colorless in aqueous solution.[13]

Approximate colors of lanthanide ions in aqueous solution[6][14][15]


Oxidation
57 58 59 60 61 62 63 64 65 66 67 68 69 70 71
state

+2 Sm2+ Eu2+ Tm2+ Yb2+

+3 La3+ Ce3+ Pr3+ Nd3+ Pm3+ Sm3+ Eu3+ Gd3+ Tb3+ Dy3+ Ho3+ Er3+ Tm3+ Yb3+ Lu3+

+4 Ce4+ Pr4+ Nd4+ Tb4+ Dy4+

Effect of 4f orbitals

Going across the lanthanides in the


periodic table, the 4f orbitals are usually
being filled. The effect of the 4f orbitals on
the chemistry of the lanthanides is
profound and is the factor that
distinguishes them from the transition
metals. There are seven 4f orbitals, and
there are two different ways in which they
are depicted: as a "cubic set" or as a
general set. The cubic set is fz3, fxz2, fyz2,
fxyz, fz(x2−y2), fx(x2−3y2) and fy(3x2−y2). The 4f
orbitals penetrate the [Xe] core and are
isolated, and thus they do not participate
in bonding. This explains why crystal field
effects are small and why they do not form
π bonds.[11] As there are seven 4f orbitals,
the number of unpaired electrons can be
as high as 7, which gives rise to the large
magnetic moments observed for
lanthanide compounds. Measuring the
magnetic moment can be used to
investigate the 4f electron configuration,
and this is a useful tool in providing an
insight into the chemical bonding.[16] The
lanthanide contraction, i.e. the reduction in
size of the Ln3+ ion from La3+ (103 pm) to
Lu3+ (86.1 pm), is often explained by the
poor shielding of the 5s and 5p electrons
by the 4f electrons.[11]

Lanthanide oxides: clockwise from top center:


praseodymium, cerium, lanthanum, neodymium,
samarium and gadolinium.

The electronic structure of the lanthanide


elements, with minor exceptions, is
[Xe]6s24fn. The chemistry of the
lanthanides is dominated by the +3
oxidation state, and in LnIII compounds the
6s electrons and (usually) one 4f electron
are lost and the ions have the
configuration [Xe]4fm.[17] All the lanthanide
elements exhibit the oxidation state +3. In
addition, Ce3+ can lose its single f electron
to form Ce4+ with the stable electronic
configuration of xenon. Also, Eu3+ can gain
an electron to form Eu2+ with the f7
configuration that has the extra stability of
a half-filled shell. Other than Ce(IV) and
Eu(II), none of the lanthanides are stable in
oxidation states other than +3 in aqueous
solution. Promethium is effectively a man-
made element, as all its isotopes are
radioactive with half-lives shorter than
20 years.

In terms of reduction potentials, the Ln0/3+


couples are nearly the same for all
lanthanides, ranging from −1.99 (for Eu) to
−2.35 V (for Pr). Thus these metals are
highly reducing, with reducing power
similar to alkaline earth metals such as
Mg (−2.36 V).[6]
Lanthanide oxidation states

All of the lanthanide elements are


commonly known to have the +3 oxidation
state and it was thought that only
samarium, europium, and ytterbium had
the +2 oxidation readily accessible in
solution. Now, it is known that all of the
lanthanides can form +2 complexes in
solution.[18]
Ionization energies and reduction potentials
of the elements
Chemical
La Ce Pr Nd
element

Atomic
57 58 59 60
number

electron
configuration
5d16s2 4f15d16s2 4f36s2 4f46s
above [Xe]
core

E° Ln4+/Ln3+ 1.72

E° Ln3+/Ln2+ −2.6

E° Ln3+/Ln −2.38 −2.34 −2.35 −2.32

1st 538 541 522 530


Ionization
energy
(kJ·mol−1)

2nd
Ionization
1067 1047 1018 1034
energy
(kJ·mol−1)

1st + 2nd
Ionization
1605 1588 1540 1564
energy
(kJ·mol−1)

3rd
Ionization
1850 1940 2090 2128
energy
(kJ·mol−1)

1st + 2nd + 3455 3528 3630 3692


3rd
Ionization
energy
(kJ·mol−1)

4th
Ionization
4819 3547 3761 3900
energy
(kJ·mol−1)

The ionization energies for the lanthanides


can be compared with aluminium. In
aluminium the sum of the first three
ionization energies is 5139 kJ·mol−1,
whereas the lanthanides fall in the range
3455 – 4186 kJ·mol−1. This correlates with
the highly reactive nature of the
lanthanides.
The sum of the first two ionization
energies for europium, 1632 kJ·mol−1 can
be compared with that of barium 1468.1
kJ·mol−1 and europium's third ionization
energy is the highest of the lanthanides.
The sum of the first two ionization
energies for ytterbium are the second
lowest in the series and its third ionization
energy is the second highest. The high
third ionization energy for Eu and Yb
correlate with the half filling 4f7 and
complete filling 4f14 of the 4f subshell, and
the stability afforded by such
configurations due to exchange energy.[11]
Europium and ytterbium form salt like
compounds with Eu2+ and Yb2+, for
example the salt like dihydrides.[19] Both
europium and ytterbium dissolve in liquid
ammonia forming solutions of Ln2+(NH3)x
again demonstrating their similarities to
the alkaline earth metals.[6]

The relative ease with which the 4th


electron can be removed in cerium and (to
a lesser extent praseodymium) indicates
why Ce(IV) and Pr(IV) compounds can be
formed, for example CeO2 is formed rather
than Ce2O3 when cerium reacts with
oxygen.

Separation of lanthanides
The similarity in ionic radius between
adjacent lanthanide elements makes it
difficult to separate them from each other
in naturally occurring ores and other
mixtures. Historically, the very laborious
processes of cascading and fractional
crystallization were used. Because the
lanthanide ions have slightly different radii,
the lattice energy of their salts and
hydration energies of the ions will be
slightly different, leading to a small
difference in solubility. Salts of the
formula Ln(NO3)3·2NH4NO3·4H2O can be
used. Industrially, the elements are
separated from each other by solvent
extraction. Typically an aqueous solution
of nitrates is extracted into kerosene
containing tri-n-butylphosphate. The
strength of the complexes formed
increases as the ionic radius decreases, so
solubility in the organic phase increases.
Complete separation can be achieved
continuously by use of countercurrent
exchange methods. The elements can also
be separated by ion-exchange
chromatography, making use of the fact
that the stability constant for formation of
EDTA complexes increases for log K ≈
15.5 for [La(EDTA)]− to log K ≈ 19.8 for
[Lu(EDTA)]−.[6][20]

Coordination chemistry and


catalysis

When in the form of coordination


complexes, lanthanides exist
overwhelmingly in their +3 oxidation state,
although particularly stable 4f
configurations can also give +4 (Ce, Tb) or
+2 (Eu, Yb) ions. All of these forms are
strongly electropositive and thus
lanthanide ions are hard Lewis acids. The
oxidation states are also very stable and
with the exception of SmI2[21] and
cerium(IV) salts[22] lanthanides are not
used for redox chemistry. 4f electrons
have a high probability of being found
close to the nucleus and are thus strongly
affected as the nuclear charge increases
across the series; this results in a
corresponding decrease in ionic radii
referred to as the lanthanide contraction.

The low probability of the 4f electrons


existing at the outer region of the atom or
ion permits little effective overlap between
the orbitals of a lanthanide ion and any
binding ligand. Thus lanthanide complexes
typically have little or no covalent
character and are not influenced by orbital
geometries. The lack of orbital interaction
also means that varying the metal typically
has little effect on the complex (other than
size), especially when compared to
transition metals. Complexes are held
together by weaker electrostatic forces
which are omni-directional and thus the
ligands alone dictate the symmetry and
coordination of complexes. Steric factors
therefore dominate, with coordinative
saturation of the metal being balanced
against inter-ligand repulsion. This results
in a diverse range of coordination
geometries, many of which are irregular,[23]
and also manifests itself in the highly
fluxional nature of the complexes. As there
is no energetic reason to be locked into a
single geometry, rapid intramolecular and
intermolecular ligand exchange will take
place. This typically results in complexes
that rapidly fluctuate between all possible
configurations.

Many of these features make lanthanide


complexes effective catalysts. Hard Lewis
acids are able to polarise bonds upon
coordination and thus alter the
electrophilicity of compounds, with a
classic example being the Luche
reduction. The large size of the ions
coupled with their labile ionic bonding
allows even bulky coordinating species to
bind and dissociate rapidly, resulting in
very high turnover rates; thus excellent
yields can often be achieved with loadings
of only a few mol%.[24] The lack of orbital
interactions combined with the lanthanide
contraction means that the lanthanides
change in size across the series but that
their chemistry remains much the same.
This allows for easy tuning of the steric
environments and examples exist where
this has been used to improve the catalytic
activity of the complex[25][26][27] and
change the nuclearity of metal
clusters.[28][29]

Despite this, the use of lanthanide


coordination complexes as homogeneous
catalysts is largely restricted to the
laboratory and there are currently few
examples them being used on an
industrial scale.[30] It should be noted
however, that lanthanides exist in many
forms other that coordination complexes
and many of these are industrially useful.
In particular lanthanide metal oxides are
used as heterogeneous catalysts in
various industrial processes.

Ln(III) compounds

The trivalent lanthanides mostly form ionic


salts. The trivalent ions are hard acceptors
and form more stable complexes with
oxygen-donor ligands than with nitrogen-
donor ligands. The larger ions are 9-
coordinate in aqueous solution,
[Ln(H2O)9]3+ but the smaller ions are 8-
coordinate, [Ln(H2O)8]3+. There is some
evidence that the later lanthanides have
more water molecules in the second
coordination sphere.[31] Complexation with
monodentate ligands is generally weak
because it is difficult to displace water
molecules from the first coordination
sphere. Stronger complexes are formed
with chelating ligands because of the
chelate effect, such as the tetra-anion
derived from 1,4,7,10-
tetraazacyclododecane-1,4,7,10-
tetraacetic acid (DOTA).
Samples of lanthanide nitrates in their hexahydrate
form. From left to right: La, Ce, Pr, Nd, Sm, Eu, Gd,
Tb, Dy, Ho, Er, Tm, Yb, Lu.

Ln(II) and Ln(IV) compounds

The most common divalent derivatives of


the lanthanides are for Eu(II), which
achieves a favorable f7 configuration.
Divalent halide derivatives are known for
all of the lanthanides. They are either
conventional salts or are Ln(III) "electride"-
like salts. The simple salts include YbI2,
EuI2, and SmI2. The electride-like salts,
described as Ln3+, 2I−, e−, include LaI2, CeI2
and GdI2. Many of the iodides form soluble
complexes with ethers, e.g.
TmI2(dimethoxyethane)3.[32] Samarium(II)
iodide is a useful reducing agent. Ln(II)
complexes can be synthesized by
transmetalation reactions.

Ce(IV) in ceric ammonium nitrate is a


useful oxidizing agent. Otherwise
tetravalent lanthanides are rare. The Ce(IV)
is the exception owing to the tendency to
form an unfilled f shell.

Hydrides
Chemical
La Ce Pr Nd
element

Atomic
57 58 59 60
number

Metal lattice
dhcp fcc dhcp dhcp
(RT)

Dihydride[19] LaH2+x CeH2+x PrH2+x NdH2+x

Structure CaF2 CaF2 CaF2 CaF2

metal
sub fcc fcc fcc fcc
lattice

Trihydride[19] LaH3−x CeH3−x PrH3−x NdH3−x


metal fcc fcc fcc hcp
sub
lattice

Trihydride
properties
bronze
transparent PrH3−x NdH3−x
red to
insulators fcc hcp
grey[35]
(color where
recorded)

Lanthanide metals react exothermically


with hydrogen to form LnH2, dihydrides.[19]
With the exception of Eu and Yb which
resemble the Ba and Ca hydrides (non
conducting, transparent salt like
compounds) they form black pyrophoric,
conducting compounds[37] where the
metal sub-lattice is face centred cubic and
the H atoms occupy tetrahedral sites.[19]
Further hydrogenation produces a
trihydride which is non-stoichiometric, non-
conducting, more salt like. The formation
of trihydride is associated with and
increase in 8–10% volume and this is
linked to greater localization of charge on
the hydrogen atoms which become more
anionic (H− hydride anion) in character.[19]

Halides
Lanthanide halides[6][38][39][37]
Chemical
La Ce Pr
element

Atomic
57 58 59
number

Tetrafluoride CeF4 PrF4 N


white white
Color m.p. °C
dec dec
Structure C.N. UF4 8 UF4 8

Trifluoride LaF3 CeF3 PrF3 N


white white green
Color m.p. °C
1493[40] 1430 1395
Structure C.N. LaF3 9 LaF3 9 LaF3 9 L

Trichloride LaCl3 CeCl3 PrCl3 N


white white green m
Color m.p. °C
858 817 786
Structure C.N. UCl3 9 UCl3 9 UCl3 9 U

Tribromide LaBr3 CeBr3 PrBr3 N


white white green
Color m.p. °C vio
783 733 691

Structure C.N. UCl3 9 UCl3 9 UCl3 9 P

Triiodide LaI3 CeI3 PrI3


yellow green
Color m.p. °C gre
766 738
Structure C.N. PuBr3 8 PuBr3 8 PuBr3 8 P

Difluoride

Color m.p. °C

Structure C.N.

Dichloride N
Color m.p. °C gre

Structure C.N. P

Dibromide N

Color m.p. °C gre

Structure C.N. P

LaI2 CeI2 PrI2


Diiodide
metallic metallic metallic pr
m
bronze bronze
Color m.p. °C vio
808 758
S
Structure C.N. CuTi2 8 CuTi2 8 CuTi2 8
C
Ln7I12 La7I12 Pr7I12

Sesquichloride La2Cl3
Structure

Sesquibromide
Structure

Monoiodide LaI[41]
NiAs
Structure
type

The only tetrahalides known are the


tetrafluorides of cerium, praseodymium,
terbium, neodymium and dysprosium, the
last two known only under matrix isolation
conditions.[6][42] All of the lanthanides form
trihalides with fluorine, chlorine, bromine
and iodine. They are all high melting and
predominantly ionic in nature.[6] The
fluorides are only slightly soluble in water
and are not sensitive to air, and this
contrasts with the other halides which are
air sensitive, readily soluble in water and
react at high temperature to form
oxohalides.[43] The trihalides were
important as pure metal can be prepared
from them.[6] In the gas phase the
trihalides are planar or approximately
planar, the lighter lanthanides have a
lower % of dimers, the heavier lanthanides
a higher proportion. The dimers have a
similar structure to Al2Cl6.[44]
Some of the dihalides are conducting
while the rest are insulators. The
conducting forms can be considered as
LnIII electride compounds where the
electron is delocalised into a conduction
band, Ln3+ (X−)2(e−). All of the diodides
have relatively short metal-metal
separations.[38] The CuTi2 structure of the
lanthanum, cerium and praseodymium
diodides along with HP-NdI2 contain 44
nets of metal and iodine atoms with short
metal-metal bonds (393-386 La-Pr).[38]
these compounds should be considered to
be two-dimensional metals (two-
dimensional in the same way that graphite
is). The salt-like dihalides include those of
Eu, Dy, Tm, and Yb. The formation of a
relatively stable +2 oxidation state for Eu
and Yb is usually explained by the stability
(exchange energy) of half filled (f7) and
fully filled f14. GdI2 possesses the layered
MoS2 structure, is ferromagnetic and
exhibits colossal magnetoresistance[38]
The sesquihalides Ln2X3 and the Ln7I12
compounds listed in the table contain
metal clusters, discrete Ln6I12 clusters in
Ln7I12 and condensed clusters forming
chains in the sesquihalides. Scandium
forms a similar cluster compound with
chlorine, Sc7Cl12[6] Unlike many transition
metal clusters these lanthanide clusters
do not have strong metal-metal
interactions and this is due to the low
number of valence electrons involved, but
instead are stabilised by the surrounding
halogen atoms.[38]

LaI is the only known monohalide.


Prepared from the reaction of LaI3 and La
metal, it has a NiAs type structure and can
be formulated La3+ (I−)(e−)2.[41]

Oxides and hydroxides

All of the lanthanides form sesquioxides,


Ln2O3. The lighter/larger lanthanides
adopt a hexagonal 7-coordinate structure
while the heavier/smaller ones adopt a
cubic 6-coordinate "C-M2O3" structure.[39]
All of the sesquioxides are basic, and
absorb water and carbon dioxide from air
to form carbonates, hydroxides and
hydroxycarbonates.[45] They dissolve in
acids to form salts.[11]

Cerium forms a stoichiometric dioxide,


CeO2, where cerium has an oxidation state
of +4. CeO2 is basic and dissolves with
difficulty in acid to form Ce4+ solutions,
from which CeIV salts can be isolated, for
example the hydrated nitrate
Ce(NO3)4.5H2O. CeO2 is used as an
oxidation catalyst in catalytic
converters.[11] Praseodymium and terbium
form non-stoichiometric oxides containing
LnIV,[11] although more extreme reaction
conditions can produce stoichiometric (or
near stoichiometric) PrO2 and TbO2.[6]

Europium and ytterbium form salt-like


monoxides, EuO and YbO, which have a
rock salt structure.[11] EuO is
ferromagnetic at low temperatures,[6] and
is a semiconductor with possible
applications in spintronics.[46] A mixed
EuII/EuIII oxide Eu3O4 can be produced by
reducing Eu2O3 in a stream of hydrogen.[45]
Neodymium and samarium also form
monoxides, but these are shiny conducting
solids,[6] although the existence of
samarium monoxide is considered
dubious.[45]

All of the lanthanides form hydroxides,


Ln(OH)3. With the exception of lutetium
hydroxide, which has a cubic structure,
they have the hexagonal UCl3 structure.[45]
The hydroxides can be precipitated from
solutions of LnIII.[11] They can also be
formed by the reaction of the sesquioxide,
Ln2O3, with water, but although this
reaction is thermodynamically favorable it
is kinetically slow for the heavier members
of the series.[45] Fajans' rules indicate that
the smaller Ln3+ ions will be more
polarizing and their salts correspondingly
less ionic. The hydroxides of the heavier
lanthanides become less basic, for
example Yb(OH)3 and Lu(OH)3 are still
basic hydroxides but will dissolve in hot
concentrated NaOH.[6]

Chalcogenides (S, Se, Te)

All of the lanthanides form Ln2Q3 (Q= S,


Se, Te).[11] The sesquisulfides can be
produced by reaction of the elements or
(with the exception of Eu2S3) sulfidizing
the oxide (Ln2O3) with H2S.[11] The
sesquisulfides, Ln2S3 generally lose sulfur
when heated and can form a range of
compositions between Ln2S3 and Ln3S4.
The sesquisulfides are insulators but
some of the Ln3S4 are metallic conductors
(e.g. Ce3S4) formulated (Ln3+)3 (S2−)4 (e−),
while others (e.g. Eu3S4 and Sm3S4) are
semiconductors.[11] Structurally the
sesquisulfides adopt structures that vary
according to the size of the Ln metal. The
lighter and larger lanthanides favoring 7-
coordinate metal atoms, the heaviest and
smallest lanthanides (Yb and Lu) favoring
6 coordination and the rest structures with
a mixture of 6 and 7 coordination.[11]
Polymorphism is common amongst the
sesquisulfides.[47] The colors of the
sesquisulfides vary metal to metal and
depend on the polymorphic form. The
colors of the γ-sesquisulfides are La2S3,
white/yellow; Ce2S3, dark red; Pr2S3, green;
Nd2S3, light green; Gd2S3, sand; Tb2S3,
light yellow and Dy2S3, orange.[48] The
shade of γ-Ce2S3 can be varied by doping
with Na or Ca with hues ranging from dark
red to yellow,[38][48] and Ce2S3 based
pigments are used commercially and are
seen as low toxicity substitutes for
cadmium based pigments.[48]

All of the lanthanides form


monochalcogenides, LnQ, (Q= S, Se,
Te).[11] The majority of the
monochalcogenides are conducting,
indicating a formulation LnIIIQ2−(e-) where
the electron is in conduction bands. The
exceptions are SmQ, EuQ and YbQ which
are semiconductors or insulators but
exhibit a pressure induced transition to a
conducting state.[47] Compounds LnQ2 are
known but these do not contain LnIV but
are LnIII compounds containing
polychalcogenide anions.[49]

Oxysulfides Ln2O2S are well known, they


all have the same structure with 7-
coordinate Ln atoms, and 3 sulfur and 4
oxygen atoms as near neighbours.[50]
Doping these with other lanthanide
elements produces phosphors. As an
example, gadolinium oxysulfide, Gd2O2S
doped with Tb3+ produces visible photons
when irradiated with high energy X-rays
and is used as a scintillator in flat panel
detectors.[51] When mischmetal, an alloy of
lanthanide metals, is added to molten
steel to remove oxygen and sulfur, stable
oxysulfides are produced that form an
immiscible solid.[11]

Pnictides (group 15)

All of the lanthanides form a mononitride,


LnN, with the rock salt structure. The
mononitrides have attracted interest
because of their unusual physical
properties. SmN and EuN are reported as
being "half metals".[38] NdN, GdN, TbN and
DyN are ferromagnetic, SmN is
antiferromagnetic.[52] Applications in the
field of spintronics are being
investigated.[46] CeN is unusual as it is a
metallic conductor, contrasting with the
other nitrides also with the other cerium
pnictides. A simple description is Ce4+N3−
(e–) but the interatomic distances are a
better match for the trivalent state rather
than for the tetravalent state. A number of
different explanations have been
offered.[53] The nitrides can be prepared by
the reaction of lanthanum metals with
nitrogen. Some nitride is produced along
with the oxide, when lanthanum metals are
ignited in air.[11] Alternative methods of
synthesis are a high temperature reaction
of lanthanide metals with ammonia or the
decomposition of lanthanide amides,
Ln(NH2)3. Achieving pure stoichiometric
compounds, and crystals with low defect
density has proved difficult.[46] The
lanthanide nitrides are sensitive to air and
hydrolyse producing ammonia.[37]

The other pnictides phosphorus, arsenic,


antimony and bismuth also react with the
lanthanide metals to form monopnictides,
LnQ. Additionally a range of other
compounds can be produced with varying
stoichiometries, such as LnP2, LnP5, LnP7,
Ln3As, Ln5As3 and LnAs2.[54]

Carbides

Carbides of varying stoichiometries are


known for the lanthanides. Non-
stoichiometry is common. All of the
lanthanides form LnC2 and Ln2C3 which
both contain C2 units. The dicarbides with
exception of EuC2, are metallic conductors
with the calcium carbide structure and can
be formulated as Ln3+C22−(e–). The C-C
bond length is longer than that in CaC2,
which contains the C22− anion, indicating
that the antibonding orbitals of the C22−
anion are involved in the conduction band.
These dicarbides hydrolyse to form
hydrogen and a mixture of
hydrocarbons.[55] EuC2 and to a lesser
extent YbC2 hydrolyse differently
producing a higher percentage of
acetylene (ethyne).[56] The sesquicarbides,
Ln2C3 can be formulated as Ln4(C2)3.
These compounds adopt the Pu2C3
structure[38] which has been described as
having C22− anions in bisphenoid holes
formed by eight near Ln neighbours.[57]
The lengthening of the C-C bond is less
marked in the sesquicarbides than in the
dicarbides, with the exception of Ce2C3.[55]
Other carbon rich stoichiometries are
known for some lanthanides. Ln3C4 (Ho-
Lu) containing C, C2 and C3 units;[58] Ln4C7
(Ho-Lu) contain C atoms and C3 units[59]
and Ln4C5 (Gd-Ho) containing C and C2
units.[60] Metal rich carbides contain
interstitial C atoms and no C2 or C3 units.
These are Ln4C3 (Tb and Lu); Ln2C (Dy, Ho,
Tm)[61][62] and Ln3C[38] (Sm-Lu).

Borides

All of the lanthanides form a number of


borides. The "higher" borides (LnBx where
x > 12) are insulators/semiconductors
whereas the lower borides are typically
conducting. The lower borides have
stoichiometries of LnB2, LnB4, LnB6 and
LnB12.[63] Applications in the field of
spintronics are being investigated.[46] The
range of borides formed by the
lanthanides can be compared to those
formed by the transition metals. The boron
rich borides are typical of the lanthanides
(and groups 1–3) whereas for the
transition metals tend to form metal rich,
"lower" borides.[64] The lanthanide borides
are typically grouped together with the
group 3 metals with which they share
many similarities of reactivity,
stoichiometry and structure. Collectively
these are then termed the rare earth
borides.[63]
Many methods of producing lanthanide
borides have been used, amongst them
are direct reaction of the elements; the
reduction of Ln2O3 with boron; reduction of
boron oxide, B2O3, and Ln2O3 together with
carbon; reduction of metal oxide with
boron carbide, B4C.[63][64][65][66] Producing
high purity samples has proved to be
difficult.[66] Single crystals of the higher
borides have been grown in a low melting
metal (e.g. Sn, Cu, Al).[63]

Diborides, LnB2, have been reported for


Sm, Gd, Tb, Dy, Ho, Er, Tm, Yb and Lu. All
have the same, AlB2, structure containing a
graphitic layer of boron atoms. Low
temperature ferromagnetic transitions for
Tb, Dy, Ho and Er. TmB2 is ferromagnetic
at 7.2 K.[38]

Tetraborides, LnB4 have been reported for


all of the lanthanides except EuB4, all have
the same UB4 structure. The structure has
a boron sub-lattice consists of chains of
octahedral B6 clusters linked by boron
atoms. The unit cell decreases in size
successively from LaB4 to LuB4. The
tetraborides of the lighter lanthanides melt
with decomposition to LnB6.[66] Attempts
to make EuB4 have failed.[65] The LnB4 are
good conductors[63] and typically
antiferromagnetic.[38]
Hexaborides, LnB6 have been reported for
all of the lanthanides. They all have the
CaB6 structure, containing B6 clusters.
They are non-stoichiometric due to cation
defects. The hexaborides of the lighter
lanthanides (La – Sm) melt without
decomposition, EuB6 decomposes to
boron and metal and the heavier
lanthanides decompose to LnB4 with
exception of YbB6 which decomposes
forming YbB12. The stability has in part
been correlated to differences in volatility
between the lanthanide metals.[66] In EuB6
and YbB6 the metals have an oxidation
state of +2 whereas in the rest of the
lanthanide hexaborides it is +3. This
rationalises the differences in conductivity,
the extra electrons in the LnIII hexaborides
entering conduction bands. EuB6 is a
semiconductor and the rest are good
conductors.[38][66] LaB6 and CeB6 are
thermionic emitters, used, for example, in
scanning electron microscopes.[67]

Dodecaborides, LnB12, are formed by the


heavier smaller lanthanides, but not by the
lighter larger metals, La – Eu. With the
exception YbB12 (where Yb takes an
intermediate valence and is a Kondo
insulator), the dodecaborides are all
metallic compounds. They all have the
UB12 structure containing a 3 dimensional
framework of cubooctahedral B12
clusters.[63]

The higher boride LnB66 is known for all


lanthanide metals. The composition is
approximate as the compounds are non-
stoichiometric.[63] They all have similar
complex structure with over 1600 atoms in
the unit cell. The boron cubic sub lattice
contains super icosahedra made up of a
central B12 icosahedra surrounded by 12
others, B12(B12)12.[63] Other complex higher
borides LnB50 (Tb, Dy, Ho Er Tm Lu) and
LnB25 are known (Gd, Tb, Dy, Ho, Er) and
these contain boron icosahedra in the
boron framework.[63]
Organometallic compounds

Lanthanide-carbon σ bonds are well


known; however as the 4f electrons have a
low probability of existing at the outer
region of the atom there is little effective
orbital overlap, resulting in bonds with
significant ionic character. As such
organo-lanthanide compounds exhibit
carbanion-like behavior, unlike the behavior
in transition metal organometallic
compounds. Because of their large size,
lanthanides tend to form more stable
organometallic derivatives with bulky
ligands to give compounds such as
Ln[CH(SiMe3)3].[68] Analogues of
uranocene are derived from
dilithiocyclooctatetraene, Li2C8H8. Organic
lanthanide(II) compounds are also known,
such as Cp*2Eu.[32]

Physical properties
Magnetic and spectroscopic

All the trivalent lanthanide ions, except


lanthanum and lutetium, have unpaired f
electrons. However, the magnetic
moments deviate considerably from the
spin-only values because of strong spin-
orbit coupling. The maximum number of
unpaired electrons is 7, in Gd3+, with a
magnetic moment of 7.94 B.M., but the
largest magnetic moments, at 10.4–10.7
B.M., are exhibited by Dy3+ and Ho3+.
However, in Gd3+ all the electrons have
parallel spin and this property is important
for the use of gadolinium complexes as
contrast reagent in MRI scans.

A solution of 4% holmium oxide in 10% perchloric acid,


permanently fused into a quartz cuvette as a
wavelength calibration standard
Crystal field splitting is rather small for the
lanthanide ions and is less important than
spin-orbit coupling in regard to energy
levels.[6] Transitions of electrons between f
orbitals are forbidden by the Laporte rule.
Furthermore, because of the "buried"
nature of the f orbitals, coupling with
molecular vibrations is weak.
Consequently, the spectra of lanthanide
ions are rather weak and the absorption
bands are similarly narrow. Glass
containing holmium oxide and holmium
oxide solutions (usually in perchloric acid)
have sharp optical absorption peaks in the
spectral range 200–900 nm and can be
used as a wavelength calibration standard
for optical spectrophotometers,[69] and are
available commercially.[70]

As f-f transitions are Laporte-forbidden,


once an electron has been excited, decay
to the ground state will be slow. This
makes them suitable for use in lasers as it
makes the population inversion easy to
achieve. The Nd:YAG laser is one that is
widely used. Europium-doped yttrium
vanadate was the first red phosphor to
enable the development of color television
screens.[71] Lanthanide ions have notable
luminescent properties due to their unique
4f orbitals. Laporte forbidden f-f
transitions can be activated by excitation
of a bound "antenna" ligand. This leads to
sharp emission bands throughout the
visible, NIR, and IR and relatively long
luminescence lifetimes.[72]

Occurrence
The lanthanide contraction is responsible
for the great geochemical divide that splits
the lanthanides into light and heavy-
lanthanide enriched minerals, the latter
being almost inevitably associated with
and dominated by yttrium. This divide is
reflected in the first two "rare earths" that
were discovered: yttria (1794) and ceria
(1803). The geochemical divide has put
more of the light lanthanides in the Earth's
crust, but more of the heavy members in
the Earth's mantle. The result is that
although large rich ore-bodies are found
that are enriched in the light lanthanides,
correspondingly large ore-bodies for the
heavy members are few. The principal ores
are monazite and bastnäsite. Monazite
sands usually contain all the lanthanide
elements, but the heavier elements are
lacking in bastnäsite. The lanthanides
obey the Oddo-Harkins rule – odd-
numbered elements are less abundant
than their even-numbered neighbors.
Three of the lanthanide elements have
radioactive isotopes with long half-lives
(138La, 147Sm and 176Lu) that can be used
to date minerals and rocks from Earth, the
Moon and meteorites.[73]

Applications
Industrial

Lanthanide elements and their compounds


have many uses but the quantities
consumed are relatively small in
comparison to other elements. About
15000 ton/year of the lanthanides are
consumed as catalysts and in the
production of glasses. This 15000 tons
corresponds to about 85% of the
lanthanide production. From the
perspective of value, however, applications
in phosphors and magnets are more
important.[74]

The devices lanthanide elements are used


in include superconductors, samarium-
cobalt and neodymium-iron-boron high-
flux rare-earth magnets, magnesium
alloys, electronic polishers, refining
catalysts and hybrid car components
(primarily batteries and magnets).[75]
Lanthanide ions are used as the active
ions in luminescent materials used in
optoelectronics applications, most notably
the Nd:YAG laser. Erbium-doped fiber
amplifiers are significant devices in
optical-fiber communication systems.
Phosphors with lanthanide dopants are
also widely used in cathode ray tube
technology such as television sets. The
earliest color television CRTs had a poor-
quality red; europium as a phosphor
dopant made good red phosphors
possible. Yttrium iron garnet (YIG) spheres
can act as tunable microwave resonators.
Lanthanide oxides are mixed with tungsten
to improve their high temperature
properties for TIG welding, replacing
thorium, which was mildly hazardous to
work with. Many defense-related products
also use lanthanide elements such as
night vision goggles and rangefinders. The
SPY-1 radar used in some Aegis equipped
warships, and the hybrid propulsion
system of Arleigh Burke-class destroyers
all use rare earth magnets in critical
capacities.[76] The price for lanthanum
oxide used in fluid catalytic cracking has
risen from $5 per kilogram in early 2010 to
$140 per kilogram in June 2011.[77]

Most lanthanides are widely used in lasers,


and as (co-)dopants in doped-fiber optical
amplifiers; for example, in Er-doped fiber
amplifiers, which are used as repeaters in
the terrestrial and submarine fiber-optic
transmission links that carry internet
traffic. These elements deflect ultraviolet
and infrared radiation and are commonly
used in the production of sunglass lenses.
Other applications are summarized in the
following table:[7]

Application Percentage

Catalytic converters 45

Petroleum refining catalysts 25

Permanent magnets 12

Glass polishing and ceramics 7

Metallurgical 7

Phosphors 3

Other 1

The complex Gd(DOTA) is used in


magnetic resonance imaging.
Life science

As mentioned in the industrial applications


section above, lanthanide metals are
particularly useful in technologies that
take advantage of their reactivity to
specific wavelengths of light.[78] Certain
life science applications take advantage of
the unique luminescence properties of
lanthanide ion complexes (Ln(III) chelates
or cryptates). These are well-suited for this
application due to their large Stokes shifts
and extremely long emission lifetimes
(from microseconds to milliseconds)
compared to more traditional fluorophores
(e.g., fluorescein, allophycocyanin,
phycoerythrin, and rhodamine). The
biological fluids or serum commonly used
in these research applications contain
many compounds and proteins which are
naturally fluorescent. Therefore, the use of
conventional, steady-state fluorescence
measurement presents serious limitations
in assay sensitivity. Long-lived
fluorophores, such as lanthanides,
combined with time-resolved detection (a
delay between excitation and emission
detection) minimizes prompt fluorescence
interference.

Time-resolved fluorometry (TRF)


combined with fluorescence resonance
energy transfer (FRET) offers a powerful
tool for drug discovery researchers: Time-
Resolved Fluorescence Resonance Energy
Transfer or TR-FRET. TR-FRET combines
the low background aspect of TRF with the
homogeneous assay format of FRET. The
resulting assay provides an increase in
flexibility, reliability and sensitivity in
addition to higher throughput and fewer
false positive/false negative results.

This method involves two fluorophores: a


donor and an acceptor. Excitation of the
donor fluorophore (in this case, the
lanthanide ion complex) by an energy
source (e.g. flash lamp or laser) produces
an energy transfer to the acceptor
fluorophore if they are within a given
proximity to each other (known as the
Förster’s radius). The acceptor fluorophore
in turn emits light at its characteristic
wavelength.

The two most commonly used lanthanides


in life science assays are shown below
along with their corresponding acceptor
dye as well as their excitation and
emission wavelengths and resultant
Stokes shift (separation of excitation and
emission wavelengths).
Excitation⇒Emission λ Excitation⇒Emission λ Stoke's Shift
Donor Acceptor
(nm) (nm) (nm)

Eu3+ 340⇒615 Allophycocyanin 615⇒660 320

Tb3+ 340⇒545 Phycoerythrin 545⇒575 235

Upcoming Medical Uses

Currently there is research showing that


lanthanides elements can be used as
anticancer agents. The main role of the
lanthanides in these studies is to inhibit
proliferation of the cancer cells.
Specifically cerium and lanthanum have
been studied for their role as anti-cancer
agents.

One of the specific elements from the


lanthanide group that has been tested and
used is cerium (Ce). There have been
studies that use a protein-cerium complex
to observe the effect of cerium on the
cancer cells. The hope was to inhibit cell
proliferation and promote cytotoxicity.[79]
Transferrin receptors in cancer cells, such
as those in breast cancer cells and
epithelial cervical cells, promote the cell
proliferation and malignancy of the
cancer.[79] Transferrin is a protein used to
transport iron into the cells and is needed
to aid the cancer cells in DNA replication.
Transferrin acts as a growth factor for the
cancerous cells and is dependent on iron.
Cancer cells have much higher levels of
transferrin receptors than normal cells and
are very dependent on iron for their
proliferation.[79] Cerium has shown results
as an anti-cancer agent due to its
similarities in structure and biochemistry
to iron. Cerium may bind in the place of
iron on to the transferrin and then be
brought into the cancer cells by transferrin-
receptor mediated endocytosis.[79] The
cerium binding to the transferrin in place
of the iron inhibits the transferrin activity in
the cell. This creates a toxic environment
for the cancer cells and causes a decrease
in cell growth. This is the proposed
mechanism for cerium’s effect on cancer
cells, though the real mechanism may be
more complex in how cerium inhibits
cancer cell proliferation. Specifically in
HeLa cancer cells studied in vitro, cell
viability was decreased after 48 to 72
hours of cerium treatments. Cells treated
with just cerium had decreases in cell
viability, but cells treated with both cerium
and transferrin had more significant
inhibition for cellular activity.[79]

Another specific element that has been


tested and used as an anti-cancer agent is
lanthanum, more specifically lanthanum
chloride (LaCl3 ). The lanthanum ion is
used to affect the levels of let-7a and
microRNAs miR-34a in a cell throughout
the cell cycle. When the lanthanum ion
was introduced to the cell in vivo or in
vitro, it inhibited the rapid growth and
induced apoptosis of the cancer cells
(specifically cervical cancer cells). This
effect was caused by the regulation of the
let-7a and microRNAs by the lanthanum
ions.[80] The mechanism for this effect is
still unclear but it is possible that the
lanthanum is acting in a similar way as the
cerium and binding to a ligand necessary
for cancer cell proliferation.

Biological effects
Due to their sparse distribution in the
earth's crust and low aqueous solubility,
the lanthanides have a low availability in
the biosphere, and for a long time were not
known to naturally form part of any
biological molecules. In 2007 a novel
methanol dehydrogenase that strictly uses
lanthanides as enzymatic cofactors was
discovered in a bacteria from the phylum
Verrucomicrobia, Methylacidiphilum
fumariolicum was found to survive only if
there is presence of lanthanides in the
environment.[81] Compared to most other
nondietary elements, non-radioactive
lanthanides are classified as having low
toxicity.[74]

See also
Actinide
Group 3 element
Lanthanide contraction
Rare earth element
Lanthanide probes

References
1. The current IUPAC recommendation is
that the name lanthanoid be used rather
than lanthanide, as the suffix "-ide" is
preferred for negative ions, whereas the
suffix "-oid" indicates similarity to one of
the members of the containing family of
elements. However, lanthanide is still
favored in most (~90%) scientific articles
and is currently adopted on Wikipedia. In
the older literature, the name "lanthanon"
was often used.
2. Gray, Theodore (2009). The Elements: A
Visual Exploration of Every Known Atom in
the Universe. New York: Black Dog &
Leventhal Publishers. p. 240. ISBN 978-1-
57912-814-2.
3. Lanthanide Archived 2011-09-11 at the
Wayback Machine., Encyclopædia
Britannica on-line
4. Holden, Norman E.; Coplen, Tyler
(January–February 2004). "The Periodic
Table of the Elements" . Chemistry
International. IUPAC. 26 (1): 8.
doi:10.1515/ci.2004.26.1.8 . Retrieved
23 March 2010.
5. F Block Elements, Oxidation States,
Lanthanides and Actinides .
Chemistry.tutorvista.com. Retrieved on
2017-12-14.
6. Greenwood, Norman N.; Earnshaw, Alan
(1997). Chemistry of the Elements (2nd
ed.). Butterworth-Heinemann. pp. 1230–
1242. ISBN 0-08-037941-9.
7. Aspinall, Helen C. (2001). Chemistry of
the f-block elements . CRC Press. p. 8.
ISBN 90-5699-333-X.
8. Hakala, Reino W. (1952). "Letters".
Journal of Chemical Education. 29 (11):
581. Bibcode:1952JChEd..29..581H .
doi:10.1021/ed029p581.2 .
9. Krishnamurthy, Nagaiyar and Gupta,
Chiranjib Kumar (2004) Extractive
Metallurgy of Rare Earths, CRC Press,
ISBN 0-415-33340-7
10. Cullity, B. D. and Graham, C. D. (2011)
Introduction to Magnetic Materials, John
Wiley & Sons, ISBN 9781118211496
11. Cotton, Simon (2006). Lanthanide and
Actinide Chemistry. John Wiley & Sons Ltd.
12. Sroor, Farid M.A.; Edelmann, Frank T.
(2012). "Lanthanides: Tetravalent
Inorganic". Encyclopedia of Inorganic and
Bioinorganic Chemistry.
doi:10.1002/9781119951438.eibc2033 .
ISBN 9781119951438.
13. McGill, Ian (2005), "Rare Earth
Elements", Ullmann's Encyclopedia of
Industrial Chemistry, 31, Weinheim: Wiley-
VCH, p. 191,
doi:10.1002/14356007.a22_607
14. Holleman, p. 1937.
15. dtv-Atlas zur Chemie 1981, Vol. 1,
p. 220.
16. Bochkarev, Mikhail N.; Fedushkin, Igor
L.; Fagin, Anatoly A.; Petrovskaya, Tatyana
V.; Ziller, Joseph W.; Broomhall-Dillard,
Randy N. R.; Evans, William J. (1997).
"Synthesis and Structure of the First
Molecular Thulium(II) Complex:
[TmI2(MeOCH2CH2OMe)3]". Angewandte
Chemie International Edition in English. 36
(12): 133–135.
doi:10.1002/anie.199701331 .
17. Winter, Mark. "Lanthanum ionisation
energies" . WebElements Ltd, UK. Retrieved
2010-09-02.
18. MacDonald, Matthew R.; Bates,
Jefferson E.; Ziller, Joseph W.; Furche,
Filipp; Evans, William J. (3 July 2013).
"Completing the Series of +2 Ions for the
Lanthanide Elements: Synthesis of
Molecular Complexes of Pr, Gd, Tb, and Lu".
Journal of the American Chemical Society.
135 (26): 9857–9868.
doi:10.1021/ja403753j . PMID 23697603 .
19. Fukai, Y. (2005). The Metal-Hydrogen
System, Basic Bulk Properties, 2d edition.
Springer. ISBN 978-3-540-00494-3.
20. Pettit, L. and Powell, K. SC-database .
Acadsoft.co.uk. Retrieved on 2012-01-15.
21. Molander, Gary A.; Harris, Christina R. (1
January 1996). "Sequencing Reactions with
Samarium(II) Iodide". Chemical Reviews. 96
(1): 307–338. doi:10.1021/cr950019y .
PMID 11848755 .
22. Nair, Vijay; Balagopal, Lakshmi; Rajan,
Roshini; Mathew, Jessy (1 January 2004).
"Recent Advances in Synthetic
Transformations Mediated by Cerium(IV)
Ammonium Nitrate". Accounts of Chemical
Research. 37 (1): 21–30.
doi:10.1021/ar030002z . PMID 14730991 .
23. Dehnicke, Kurt; =Greiner, Andreas
(2003). "Unusual Complex Chemistry of
Rare-Earth Elements: Large Ionic Radii—
Small Coordination Numbers". Angewandte
Chemie International Edition. 42 (12):
1340–1354. doi:10.1002/anie.200390346 .
24. Aspinall, Helen C. (2001). Chemistry of
the f-block elements. Amsterdam [u.a.]:
Gordon & Breach. ISBN 905699333X.
25. Kobayashi, Shū; Hamada, Tomoaki;
Nagayama, Satoshi; Manabe, Kei (1
January 2001). "Lanthanide
Trifluoromethanesulfonate-Catalyzed
Asymmetric Aldol Reactions in Aqueous
Media". Organic Letters. 3 (2): 165–167.
doi:10.1021/ol006830z . PMID 11430025 .
26. Aspinall, Helen C.; Dwyer, Jennifer L.;
Greeves, Nicholas; Steiner, Alexander (1
April 1999). "Li3[Ln(binol)3]·6THF: New
Anhydrous Lithium Lanthanide
Binaphtholates and Their Use in
Enantioselective Alkyl Addition to
Aldehydes". Organometallics. 18 (8): 1366–
1368. doi:10.1021/om981011s .
27. Parac-Vogt, Tatjana N.; Pachini, Sophia;
Nockemann, Peter; VanmHecke, Kristof;
Van Meervelt, Luc; Binnemans, Koen (1
November 2004). "Lanthanide(III)
Nitrobenzenesulfonates as New Nitration
Catalysts: The Role of the Metal and of the
Counterion in the Catalytic Efficiency".
European Journal of Organic Chemistry.
2004 (22): 4560–4566.
doi:10.1002/ejoc.200400475 .
28. Lipstman, Sophia; Muniappan, Sankar;
George, Sumod; Goldberg, Israel (1 January
2007). "Framework coordination polymers
of tetra(4-carboxyphenyl)porphyrin and
lanthanide ions in crystalline solids". Dalton
Transactions (30): 3273.
doi:10.1039/B703698A .
29. Bretonnière, Yann; Mazzanti, Marinella;
Pécaut, Jacques; Dunand, Frank A.;
Merbach, André E. (1 December 2001).
"Solid-State and Solution Properties of the
Lanthanide Complexes of a New
Heptadentate Tripodal Ligand: A Route to
Gadolinium Complexes with an Improved
Relaxation Efficiency". Inorganic Chemistry.
40 (26): 6737–6745.
doi:10.1021/ic010591 . PMID 11735486 .
30. Trinadhachari, Ganala Naga; Kamat,
Anand Gopalkrishna; Prabahar, Koilpillai
Joseph; Handa, Vijay Kumar; Srinu,
Kukunuri Naga Venkata Satya; Babu,
Korupolu Raghu; Sanasi, Paul Douglas (15
March 2013). "Commercial Scale Process
of Galanthamine Hydrobromide Involving
Luche Reduction: Galanthamine Process
Involving Regioselective 1,2-Reduction of
α,β-Unsaturated Ketone". Organic Process
Research & Development. 17 (3): 406–412.
doi:10.1021/op300337y .
31. Burgess, J. (1978). Metal ions in
solution. New York: Ellis Horwood. ISBN 0-
85312-027-7.
32. Nief, F. (2010). "Non-classical divalent
lanthanide complexes". Dalton Trans. 39
(29): 6589–6598. doi:10.1039/c001280g .
PMID 20631944 .
33. Kohlmann, H.; Yvon, K. (2000). "The
crystal structures of EuH2 and EuLiH3 by
neutron powder diffraction". Journal of
Alloys and Compounds. 299 (1–2): L16–
L20. doi:10.1016/S0925-8388(99)00818-X .
34. Matsuoka, T.; Fujihisa, H.; Hirao, N.;
Ohishi, Y.; Mitsui, T.; Masuda, R.; Seto, M.;
Yoda, Y.; Shimizu, K.; Machida, A.; Aoki, K.
(2011). "Structural and Valence Changes of
Europium Hydride Induced by Application
of High-Pressure H2". Physical Review
Letters. 107 (2): 025501.
Bibcode:2011PhRvL.107b5501M .
doi:10.1103/PhysRevLett.107.025501 .
PMID 21797616 .
35. Tellefsen, M.; Kaldis, E.; Jilek, E. (1985).
"The phase diagram of the Ce-H2 system
and the CeH2-CeH3 solid solutions". Journal
of the Less Common Metals. 110 (1–2):
107–117. doi:10.1016/0022-
5088(85)90311-X .
36. Kumar, Pushpendra; Philip, Rosen; Mor,
G. K.; Malhotra, L. K. (2002). "Influence of
Palladium Overlayer on Switching
Behaviour of Samarium Hydride Thin
Films". Japanese Journal of Applied
Physics. 41 (Part 1, No. 10): 6023–6027.
Bibcode:2002JaJAP..41.6023K .
doi:10.1143/JJAP.41.6023 .
37. Holleman, p. 1942
38. David A. Atwood, ed. (19 February
2013). The Rare Earth Elements:
Fundamentals and Applications (eBook).
John Wiley & Sons. ISBN 9781118632635.
39. Wells, A. F. (1984). Structural Inorganic
Chemistry (5th ed.). Oxford Science
Publication. ISBN 0-19-855370-6.
40. Perry, Dale L. (2011). Handbook of
Inorganic Compounds, Second Edition .
Boca Raton, Florida: CRC Press. p. 125.
ISBN 978-1-43981462-8. Retrieved
2014-02-17.
41. Ryazanov, Mikhail; Kienle, Lorenz;
Simon, Arndt; Mattausch, Hansjürgen
(2006). "New Synthesis Route to and
Physical Properties of Lanthanum
Monoiodide†". Inorganic Chemistry. 45 (5):
2068–2074. doi:10.1021/ic051834r .
PMID 16499368 .
42. Vent-Schmidt, T.; Fang, Z.; Lee, Z.; Dixon,
D.; Riedel, S. (2016). "Extending the Row of
Lanthanide Tetrafluorides: A Combined
Matrix-Isolation and Quantum-Chemical
Study". Chemistry. 22 (7): 2406–16.
doi:10.1002/chem.201504182 .
PMID 26786900 .
43. Haschke, John. M. (1979). "Chapter
32:Halides". In Gschneider Jr., K. A.
Handbook on the Physics and Chemistry of
Rare Earths vol 4. North Holland Publishing
Company. pp. 100–110. ISBN 0-444-85216-
6.
44. Kovács, Attila (2004). "Structure and
Vibrations of Lanthanide Trihalides: An
Assessment of Experimental and
Theoretical Data". Journal of Physical and
Chemical Reference Data. 33 (1): 377.
Bibcode:2004JPCRD..33..377K .
doi:10.1063/1.1595651 .
45. Adachi, G.; Imanaka, Nobuhito and
Kang, Zhen Chuan (eds.) (2006) Binary Rare
Earth Oxides. Springer. ISBN 1-4020-2568-8
46. Nasirpouri, Farzad and Nogaret, Alain
(eds.) (2011) Nanomagnetism and
Spintronics: Fabrication, Materials,
Characterization and Applications. World
Scientific. ISBN 9789814273053
47. Flahaut, Jean (1979). "Chapter
31:Sulfides, Selenides and Tellurides". In
Gschneider Jr., K. A. Handbook on the
Physics and Chemistry of Rare Earths vol 4.
North Holland Publishing Company.
pp. 100–110. ISBN 0-444-85216-6.
48. Berte, Jean-Noel (2009). "Cerium
pigments". In Smith, Hugh M. High
Performance Pigments. Wiley-VCH. ISBN 3-
527-30204-2.
49. Holleman, p. 1944.
50. Liu, Guokui and Jacquier, Bernard (eds)
(2006) Spectroscopic Properties of Rare
Earths in Optical Materials, Springer
51. Sisniga, Alejandro (2012). "Chapter 15".
In Iniewski, Krzysztof. Integrated
Microsystems: Electronics, Photonics, and
Biotechnology. CRC Press. ISBN 978-3-527-
31405-8.
52. Temmerman, W. M. (2009). "Chapter
241: The Dual, Localized or Band‐Like,
Character of the 4f‐States". In Gschneider
Jr., K. A. Handbook on the Physics and
Chemistry of Rare Earths vol 39. Elsevier.
pp. 100–110. ISBN 978-0-444-53221-3.
53. Dronskowski, R. (2005) Computational
Chemistry of Solid State Materials: A Guide
for Materials Scientists, Chemists,
Physicists and Others, Wiley,
ISBN 9783527314102
54. Hulliger, F. (1979). "Chapter 33: Rare
Earth Pnictides". In Gschneider Jr., K. A.
Handbook on the Physics and Chemistry of
Rare Earths vol 4. North Holland Publishing
Company. pp. 100–110. ISBN 0-444-85216-
6.
55. Greenwood, Norman N.; Earnshaw, Alan
(1997). Chemistry of the Elements (2nd
ed.). Butterworth-Heinemann. pp. 297–299.
ISBN 0-08-037941-9.
56. Spedding, F. H.; Gschneidner, K.; Daane,
A. H. (1958). "The Crystal Structures of
Some of the Rare Earth Carbides". Journal
of the American Chemical Society. 80 (17):
4499–4503. doi:10.1021/ja01550a017 .
57. Wang, X.; Loa, I.; Syassen, K.; Kremer, R.;
Simon, A.; Hanfland, M.; Ahn, K. (2005).
"Structural properties of the sesquicarbide
superconductor La2C3 at high pressure".
Physical Review B. 72 (6): 064520.
arXiv:cond-mat/0503597  .
Bibcode:2005PhRvB..72f4520W .
doi:10.1103/PhysRevB.72.064520 .
58. Poettgen, Rainer.; Jeitschko, Wolfgang.
(1991). "Scandium carbide, Sc3C4, a carbide
with C3 units derived from propadiene".
Inorganic Chemistry. 30 (3): 427–431.
doi:10.1021/ic00003a013 .
59. Czekalla, Ralf; Jeitschko, Wolfgang;
Hoffmann, Rolf-Dieter; Rabeneck, Helmut
(1996). "Preparation, Crystal Structure, and
Properties of the Lanthanoid Carbides
Ln4C7 with Ln: Ho, Er, Tm, and Lu" (PDF). Z.
Naturforsch. B. 51 (5): 646–654.
doi:10.1515/znb-1996-0505 .
60. Czekalla, Ralf; Hüfken, Thomas;
Jeitschko, Wolfgang; Hoffmann, Rolf-Dieter;
Pöttgen, Rainer (1997). "The Rare Earth
Carbides R4C5 with R=Y, Gd, Tb, Dy, and Ho".
Journal of Solid State Chemistry. 132 (2):
294–299. Bibcode:1997JSSCh.132..294C .
doi:10.1006/jssc.1997.7461 .
61. Atoji, Masao (1981). "Neutron-
diffraction study of Ho2C at 4–296 K". The
Journal of Chemical Physics. 74 (3): 1893.
Bibcode:1981JChPh..74.1893A .
doi:10.1063/1.441280 .
62. Atoji, Masao (1981). "Neutron-
diffraction studies of Tb2C and Dy2C in the
temperature range 4–296 K". The Journal
of Chemical Physics. 75 (3): 1434.
Bibcode:1981JChPh..75.1434A .
doi:10.1063/1.442150 .
63. Mori, Takao (2008). "Chapter
238:Higher Borides". In Gschneider Jr., K. A.
Handbook on the Physics and Chemistry of
Rare Earths vol 38. North Holland. pp. 105–
174. ISBN 978-0-444-521439.
64. Greenwood, Norman N.; Earnshaw, Alan
(1997). Chemistry of the Elements (2nd
ed.). Butterworth-Heinemann. p. 147.
ISBN 0-08-037941-9.
65. Refractory Materials, Volume 6-IV:
1976, ed. Allen Alper, Elsevier, ISBN 0-12-
053204-2
66. Zuckerman, J. J. (2009) Inorganic
Reactions and Methods, The Formation of
Bonds to Group-I, -II, and -IIIb Elements, Vol.
13, John Wiley & Sons, ISBN 089573-263-7
67. Reimer, Ludwig (1993). Image
Formation in Low-voltage Scanning
Electron Microscopy . SPIE Press.
ISBN 978-0-8194-1206-5.
68. Cotton, S. A. (1997). "Aspects of the
lanthanide-carbon σ-bond". Coord. Chem.
Rev. 160: 93–127. doi:10.1016/S0010-
8545(96)01340-9 .
69. MacDonald, R. P. (1964). "Uses for a
Holmium Oxide Filter in
Spectrophotometry" (PDF). Clinical
Chemistry. 10 (12): 1117–20.
PMID 14240747 .
70. "Holmium Glass Filter for
Spectrophotometer Calibration" . Archived
from the original on 2010-03-14. Retrieved
2009-06-06.
71. Levine, Albert K.; Palilla, Frank C.
(1964). "A new, highly efficient red-emitting
cathodoluminescent phosphor (YVO4:Eu)
for color television". Applied Physics
Letters. 5 (6): 118.
Bibcode:1964ApPhL...5..118L .
doi:10.1063/1.1723611 .
72. Werts, M. H. V. (2005). "Making Sense
of Lanthanide Luminescence". Science
Progress. 88 (2): 101–131.
doi:10.3184/003685005783238435 .
73. There exist other naturally occurred
radioactive isotopes of lanthanides with
long half-lives (144Nd, 150Nd, 148Sm, 151Eu,
152Gd) but they are not used as
chronometers.
74. McGill, Ian (2005) "Rare Earth Elements"
in Ullmann's Encyclopedia of Industrial
Chemistry, Wiley-VCH, Weinheim.
doi:10.1002/14356007.a22_607 .
75. Haxel G, Hedrick J, Orris J. (2006). Rare
earth elements critical resources for high
technology (PDF). Reston (VA): United
States Geological Survey. USGS Fact Sheet:
087‐02. Retrieved 2008-04-19.
76. Livergood R. (2010). "Rare Earth
Elements: A Wrench in the Supply Chain"
(PDF). Center for Strategic and International
Studies. Retrieved 2010-10-22.
77. Chu, Steven. Critical Materials Strategy
p. 17 United States Department of Energy,
December 2011. Accessed: 23 December
2011.
78. Bunzil, Jean-Claude; Piguet, Claude
(September 2005). "Taking advantage of
luminescent lanthanide ions" (PDF).
Chemical Society Reviews. 34 (12): 1048–
77. doi:10.1039/b406082m .
PMID 16284671 . Retrieved 22 December
2012.
79. Palizban, A. A.; Sadeghi-aliabadi, H.;
Abdollahpour, F. (2010-01-01). "Effect of
cerium lanthanide on Hela and MCF-7
cancer cell growth in the presence of
transferring" . Research in Pharmaceutical
Sciences. 5 (2): 119–125. PMC 3093623  .
PMID 21589800 .
80. Yu, Lingfang; Xiong, Jieqi; Guo, Ling;
Miao, Lifang; Liu, Sisun; Guo, Fei (2015).
"The effects of lanthanum chloride on
proliferation and apoptosis of cervical
cancer cells: involvement of let-7a and miR-
34a microRNAs". BioMetals. 28 (5): 879–
890. doi:10.1007/s10534-015-9872-6 .
81. Pol, A., et al (2014). "Rare Earth Metals
Are Essential for Methanotrophic Life in
Volcanic Mudpots". Environ Microbiol. 16
(1): 255–264. doi:10.1111/1462-
2920.12249 .
Cited sources
Holleman, Arnold F.; Wiberg, Egon;
Wiberg, Nils (2007). Lehrbuch der
Anorganischen Chemie (in German) (102
ed.). Walter de Gruyter. ISBN 978-3-11-
017770-1.

External links
lanthanide Sparkle Model , used in the
computational chemistry of lanthanide
complexes
USGS Rare Earths Statistics and
Information
Ana de Bettencourt-Dias: Chemistry of
the lanthanides and lanthanide-
containing materials

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Lanthanide&oldid=822169744"

Last edited 14 days ago by Jauerback

Content is available under CC BY-SA 3.0 unless


otherwise noted.

You might also like