You are on page 1of 14

ARVIN ENRICO SEC 3 INORGANIC CHEMISTRY

BSED II GENERAL SCIENCE MRS.AILEEN TOLENTINO (MWF 4:30-5:30)

1. Identify the different types of chemical formulas.


2. Write and name binary compounds consisting of metals and nonmetals.
3. Write and name compounds with polyatomic ions or oxyanions, hydrates, binary acids, and
oxyacids. 4. Distinguish by structure an alkane, alkene, alkyne, cycloalkane, halogenated
carbon, and aromatic hydrocarbon.
5. Values Integration: promote cooperation, respect the ideas of others, and practice awareness
and responsibility.
Big Idea: Naming and Writing Formulas of Compounds.
Essential Question: Why is there a need for Chemical Formulas.

1. Identify the different types of chemical formulas.

A chemical formula is a notation used by scientists to show the number and type of atoms
present in a molecule, using atomic symbols and numerical subscripts. A chemical formula is a
simple representation, in writing, of a three-dimensional molecule that exists. A chemical
formula describes a substance, down to the exact atoms which make it up. There are three
basic types of chemical formula, the empirical formula, the molecular formula, and the structural
formula.
Each one of these chemical formula provides slightly different information about the makeup of
a substance, and clues to its three-dimensional shape and how it will interact with other
molecules, atoms, and ions. In a chemical formula, the letters represent the atomic symbol of
each atom. The subscript (lower) represents the number of each atom, while the superscript
(higher) represents the charge on a given atom. A coefficient before a chemical formula
represents that many units of the molecule. Each of the different types of the chemical formula
is read a little differently.
Types of Chemical Formula
Empirical Formula
The empirical chemical formula represents the relative number of atoms of each element in the
compound. Some compounds, like water, have the same empirical and molecular formula
because they are small and have the same ratio of atoms in molecules and the number of
atoms in a molecule. The empirical and molecular formula for water looks like this:
H2O
The empirical formula is determined by the weight of each atom within the molecule. Therefore,
for a slightly bigger molecule like hydrogen peroxide, the empirical formula shows only the ratio
of atoms. In this case:
HO:
However, this empirical chemical formula only shows the basic foundation of the molecule. In
reality, two HO: molecules come together to form a hydrogen peroxide molecule.
Molecular Formula
The molecular formula comes in to show the actual number of atoms within each molecule.
Thus, for hydrogen peroxide the molecular formula is thus:
H2O2
As you can see, this somewhat confuses the actual structure of hydrogen peroxide. While the
empirical chemical formula gives clues that the molecule has two oxygen atoms bonded
together in the middle, the molecular formula does not make that clear at all. However, the
molecular formula is often used to describe molecules, simply because it is convenient and
most molecules can be looked up after their formula is identified.
Structural Formula
The structural formula of a molecule is a chemical formula with a more artistic twist. In these
chemical formulas, the actual bonds between molecules are shown. This helps the reader
understand how the different atoms are connected, and thus how the molecule functions in
space. There are many different structural chemical formulas to consider.
The simplest, the electron dot method, uses colons and periods to show bonds between atoms.
Each colon represents a pair of electrons, shared between the atoms on either side of the colon.
This formula more accurately represents the actual arrangement of atoms within a molecule. In
the case of water, the electron dot formula would look like this:
H:O:H
Another chemical formula, the bond-line formula, also shows the bonds between atoms. Instead
of showing each electron that is shared, a line is used to designate an electron pair shared
between the atoms. Water, in the bond-line formula, looks like this:
H-O-H
Scientists have come up with much more advanced formula and representations of molecules,
including three-dimensional ball-and-stick models, space-filling models, and even models which
consider the electron density of the atoms being modeled. These advanced models consider not
only the atoms present and their number, but the angles, sizes, and distances between atoms
within a molecule. The ball and stick model of water, below, even shows the polarity of the
molecule, as the large oxygen atom tends to attract the most electrons.
Molecular Mass from Chemical Formula
One important skill derived from the chemical formula is calculating the molecular mass. The
molecular mass of a molecule is the sum of all the different atoms within. Each substance has a
particular molecular mass, determined by its particular makeup.
To determine the molecular mass of a substance, consider the chemical formula. The formula
easily displays each atom present. Be sure to multiply by the number of atoms of each
molecule. The subscripts on each atom will indicate how many there are. Some large molecules
with multiple similar groups will show the groups something like the example below:
C(OH)4
In this case, there are four groups of (OH), not 4 hydrogen atoms. Make sure you take this into
account when calculating molecular mass. The molecular mass can be used to identify
substances, properly weigh substances for experiments, and do several calculations involving
the energy involved in chemical reactions. Scientists often use a chemical formula as a store of
much of this information, without having to explain basic chemistry in every paper.

2. Write and name binary compounds consisting of metals and nonmetals.


Generally, two types of inorganic compounds can be formed: ionic compounds and molecular
compounds. Nomenclature is the process of naming chemical compounds with different names
so that they can be easily identified as separate chemicals. Inorganic compounds are
compounds that do not deal with the formation of carbohydrates, or simply all other compounds
that do not fit into the description of an organic compound. For example, organic compounds
include molecules with carbon rings and/or chains with hydrogen atoms (see picture below).
Inorganic compounds, the topic of this section, are every other molecule that does not include
these distinctive carbon and hydrogen structures.
Compounds between Metals and Nonmetals (Cation and Anion)
Compounds made of metal and nonmetal are commonly known as Ionic Compounds, where the
compound name has an ending of –ide. Cations have positive charges while anions have
negative charges. The net charge of an ionic compound must be zero which also means it must
be electrically neutral. For example, one Na+ is paired with one Cl-; one Ca2+ is paired with two
Br-. Two rules must be followed:
The cation (metal) is always named first with its name unchanged
The anion (nonmetal) is written after the cation, modified to end in –ide
example:
Na+ + Cl- = NaCl; Ca2+ + 2Br- = CaBr2
Sodium + Chlorine = Sodium Chloride; Calcium + Bromine = Calcium Bromide

3. Write and name compounds with polyatomic ions or oxyanions, hydrates, binary
acids, and oxyacids.

Many materials exist as simply binary compounds composed of a metal cation and a nonmetal
anion, with each ion consisting of only one type of atom. Other combinations of atoms also
exist, either larger ionic complexes or complete molecules. Some of the most useful materials
we work with contain polyatomic ions.
Polyatomic Ions
A polyatomic ion is an ion composed of more than one atom. The ammonium ion consists of
one nitrogen atom and four hydrogen atoms. Together, they comprise a single ion with
a 1+1+ charge and a formula of NH+4NH4+. The carbonate ion consists of one carbon atom
and three oxygen atoms and carries an overall charge of 2−2−. The formula of the carbonate
ion is CO2−3CO32−. The atoms of a polyatomic ion are tightly bonded together and so the
entire ion behaves as a single unit. The figures below show several models.
(left) The ammonium ion (NH+4)(NH4+) is a nitrogen atom (blue) bonded to four hydrogen
atoms (white). (middle) The hydroxide ion (OH−)(OH−) is an oxygen atom (red) bonded to a
hydrogen atom. (left): The carbonate ion (CO2−3)(CO32−) is a carbon atom (black) bonded to
three oxygen atoms.

The vast majority of polyatomic ions are anions, many of which end in -ate or -ite. Notice that in
some cases such as nitrate (NO−3)(NO3−) and nitrite (NO−2)(NO2−), multiple anions consist of
the same two elements. In these cases, the difference between the ions is in the number of
oxygen atoms present, while the overall charge is the same. As a class, these are called
oxoanions. When there are two oxoanions for a particular element, the one with the greater
number of oxygen atoms gets the -ate suffix, while the one with the fewer number of oxygen
atoms gets the -ite suffix. The four oxoanions of chlorine are shown below.

ClO−ClO−, hypochlorite
ClO−2ClO2−, chlorite
ClO−3ClO3−, chlorate
ClO−4ClO4−, perchlorate

In cases such as this, the ion with one more oxygen atom than the -ate anion is given a per-
prefix. The ion with one fewer oxygen atom than the -ite anion is given a hypo- prefix.
Oxyanions
Some elements can form more than one oxyanion (polyatomic
ions that contain oxygen), each containing a different number of
oxygen atoms. For example, chlorine can combine with oxygen in
four ways to form four different oxyanions: ClO4−, ClO3−, ClO2−,
and ClO−. (Note that in a family of oxyanions, the charge remains
the same; only the number of oxygen atoms varies.)
The most common of the chlorine oxyanions are chlorate, ClO3−.
You will generally find that the most common of an element’s
oxyanions has a name with the form (root)ate. These can be
memorized from See below. The names of the other possible
oxyanions are determined as follows (see below).
The anion with one more oxygen atom than the (root)ate anion is
named by putting per- at the beginning of the root and -ate at the
end. For example, ClO4− is perchlorate.
The anion with one fewer oxygen atom than the (root)ate anion is
named with -ite on the end of the root. ClO2− is chlorite.
The anion with two fewer oxygen atoms than the (root)ate anion is
named by putting hypo- at the beginning of the root and -ite at the
end. ClO− is hypochlorite.
EXAMPLE 1 - Naming Polyatomic Ions:    Write names that
correspond to the following formulas for polyatomic ions:
(1) PO33- (with barium ions in glass)
(2) HSO4− (in perfumes).
SOLUTION:
1. PO33− is phosphite. PO43− is phosphate, and this ion has one
less oxygen.
2. HSO4- is hydrogen sulfate. This could be called by its
nonsystematic name, bisulfate, but it is preferable to use the
systematic name, which shows that one H+  ion has been added
to sulfate, SO42-.

EXAMPLE 2 - Formulas for Polyatomic Ions:    Write formulas that


correspond to the following names for polyatomic ions:
(1) bromite ion, used in the production of cloth
(2) dihydrogen hypophosphite ion, used with manganese(II) ions
as a food additive.
SOLUTION:
1. Bromite ion is BrO2−. Bromate is BrO3−. Bromite has one less
oxygen atom than bromate.
2. Dihydrogen hypophosphite ion is H2PO2−. Hypophosphite has
two fewer oxygen atoms than phosphate, PO43−. The dihydrogen
part of the name indicates that two H+ ions have been added to
hypophosphite, PO23−.  The two H+ ions neutralize two of the
three minus charges, leaving −1.

Common Polyatomic ions that end in -ate


NO3-  nitrate
SO42-  sulfate
PO43-  phosphate
ClO3-  chlorate
BrO3-  bromate
IO3-  iodate

Different number of oxygen atoms

Relationship General Example Example


name name formula

one more oxygen per(root) ate Perchlorate ClO4-


atom than (root)ate

(root)ate Chlorate ClO3-

one less oxygen (root)ite Chlorite ClO2-


atom than
(root)ate

two less oxygen hypo(root)ite Hypochlorite ClO-


atoms than
(root)ate

If you memorize that nitrate is NO3−, you know that NO2-  is


nitrite, because it has one less oxygen atom than nitrate. If you
memorize that iodate is IO3−, you know that IO4-  is periodate,
because it has one more oxygen than iodate.
Some polyatomic ions like HCO3−, HSO3−, and HSO4-  also have
nonsystematic names that are often used (Table 3). You should
avoid using these less accepted names, but because many people
still use them, you should know them.
Systematic and Nonsystematic Names for Some Polyatomic Ions  

Formula Systematic Nonsystematic name


(preferred) name

HCO3- hydrogen carbonate Bicarbonate

HSO4- hydrogen sulfate Bisulfate

HSO3- hydrogen sulfite Bisulfite

Hydrate, any compound containing water in the form of H2O molecules, usually, but not always,


with a definite content of water by weight. The best-known hydrates are crystalline solids that
lose their fundamental structures upon removal of the bound water. Exceptions to this are
the zeolites (aluminum silicate minerals or their synthetic analogs that contain water in indefinite
amounts) as well as similar clay minerals, certain clays, and metallic oxides, which have
variable proportions of water in their hydrated forms; zeolites lose and regain water reversibly
with little or no change in structure.
Substances that spontaneously absorb water from the air to form hydrates are known as
hygroscopic or deliquescent, whereas hydrates that lose so-called water of hydration or water of
crystallization to form the unhydrated (anhydrous) substances are known as efflorescent. In
many cases, the uptake and loss of water (by heating, decreasing pressure, or other means)
are reversible processes, sometimes accompanied by changes in color. For example,
blue vitriol, or copper sulfate pentahydrate (CuSO4∙5H2O), is blue, copper sulfate trihydrate
(CuSO4∙3H2O) is blue, and anhydrous copper sulfate (CuSO4) is white.
Other examples of hydrates are Glauber’s salt (sodium sulfate decahydrate,
Na2SO4∙10H2O); washing soda (sodium carbonate decahydrate,
Na2CO3∙10H2O); borax (sodium tetraborate decahydrate, Na2B4O7∙10H2O); the sulfates
known as vitriols (e.g., Epsom salt, MgSO4∙7H2O); and the double salts are known collectively
as alums (M+2SO4∙M+32(SO4)3∙24H2O, where M+ is a mono positive cation, such as K+or
NH4+, and M3+ is a tripositive cation, such as Al3+ or Cr3+).
Binary Acids
Binary acids are certain molecular compounds in which hydrogen is combined with a second
nonmetallic element; these acids include HF, HCl, HBr, and HI.
HCl, HBr, and HI are all strong acids, whereas HF is a weak acid. The acid strength increases
as the experimental pKa values decrease in the following order:
HF (pKa = 3.1) < HCl (pKa = -6.0) < HBr (pKa = -9.0) < HI (pKa = -9.5).
Why is HF a weak acid, when the rest of the hydrohalic acids are strong? One might correctly
assume that fluorine is very electronegative, so the H-F bond is highly polar and we can expect
HF to dissociate readily in solution; this reasoning is not wrong, but the electronegativity
argument is trumped by considerations of ionic size. Recall the periodic trend that ionic size
increases as we move down the periodic table. Because fluorine is at the top of the halogens,
the F–ion is the smallest halide; therefore, its electrons are concentrated around its nucleus,
and as a result, the H-F bond is relatively short. Shorter bonds are more stable, and thus the H-
F bond is more difficult to break.
Once we move down to chlorine, however, the trend changes. Chlorine is larger and has more
electrons, and therefore the H-Cl bond is longer and weaker. In the presence of water, the
electrostatic attraction between water’s partially negative oxygen and the partially positive
hydrogen on H-Cl is strong enough to break the H-Cl bond, and the ions dissociate in solution.
The same reasoning applies to both HBr and HI. These acids are even stronger than HCl
because the Br– and I– ions are even larger. As such, the H-Br and H-I bonds are even weaker,
and these compounds also readily dissociate in solution.
An oxoacid (sometimes called an oxyacid) is an acid that contains oxygen. To be more specific,
and oxoacid is an acid that:
contains oxygen
contains at least one other element
has at least one hydrogen atom bonded to oxygen
forms an ion by the loss of one or more protons in solution.

Examples of oxoacids include:


Carboxylic acids
Sulfuric acid
Nitric acid
Phosphoric acid

Halogen oxoacids include hypochlorous acid (HOCl); chlorous acid(HOClO); chloric


acid(HOClO2); oerchloric acid(HOClO3); oerbromic acid (HOBrO3)
All oxoacids have the acidic hydrogen bond to an oxygen atom, so bond strength (length) is not
a factor, similar to binary nonmetal acids; instead, the main determining factor for an oxacid’s
relative strength has to do with the central atom’s electronegativity (X), as well as the number of
O atoms around that central atom.
Electronegativity of the Central Atom
Consider the simple oxyacids HOI (hypoiodous acid), HOBr (hypobromous acid), and HOCl
(hypochlorous acid). These acids can be arranged in order of their pKavalues and, by
extension, their relative strengths:
HOCl pKa = 7.5 < HOBr pKa = 8.6 < HOI pKa = 10.6
Recall that smaller values of pKa correspond to greater acid strength. Therefore, HOCl is the
strongest acid and HOI is the weakest, and acid strength decreases as the central halogen
descends on the periodic table.
The strength of the acid is determined by the central atom’s electronegativity relative to the
surrounding atoms in the molecule. Because Cl is the most electronegative, it draws the bulk of
the electrons in the HOCl molecule toward itself; because H and Cl are on opposite ends of the
molecule, Cl pulls at the electrons in the H-O bond, thereby weakening it. The weaker the H-O
bond, the more easily the H+ can ionize in water, and the stronger the acid.
Number of Oxygen Atoms Around the Central Atom
Consider the family of chlorooxoacids, which are arranged below in order of pKa values:
HOClO3 pKa = -8 < HOClO2 pKa = -1.0 < HOClO pKa = 1.92 < HOCl pKa = 7.53
The strongest acid is perchloric acid on the left, and the weakest is hypochlorous acid on the far
right. Notice that the only difference between these acids is the number of oxygens bonded to
chlorine. As the number of oxygens increases, so does the acid strength; again, this has to do
with electronegativity. Oxygen is a highly electronegative element, and the more oxygen atoms
present, the more that the molecule’s electron density will be pulled off the O-H bond,
weakening it and creating a stronger acid.
Carboxylic Acids
Carboxylic acids are an important subclass of organic oxoacids, characterized by the presence
of at least one carboxyl group. The general formula of a carboxylic acid is R-COOH, where R is
some monovalent functional group. A carboxyl group (or carboxy) is a functional group
consisting of a carbonyl (RR’C=O) and a hydroxyl (R-O-H), which has the formula -C(=O)OH,
usually written as -COOH or -CO2H.

A carboxylic acidCarboxylic acids are organic oxoacids characterized by the presence of at


least one carboxyl group, which has the formula -C(=O)OH, usually written as -COOH or -
CO2H.
Carboxylic acids are the most common type of organic acid. Among the simplest examples are
formic acid H-COOH, which occurs in ants, and acetic acid CH3-COOH, which gives vinegar its
sour taste. Acids with two or more carboxyl groups are called dicarboxylic, tricarboxylic, etc. The
simplest dicarboxylic example is oxalic acid (COOH)2, which is just two connected carboxyls.
Mellitic acid is an example of a hexacarboxylic acid. Other important natural examples include
citric acid (in lemons) and tartaric acid (in tamarinds).
Salts and esters of carboxylic acids are called carboxylates. When a carboxyl group is
deprotonated, its conjugate base, a carboxylate anion, forms. Carboxylate ions are resonance
stabilized, and this increased stability makes carboxylic acids more acidic than alcohols.
Carboxylic acids can be seen as reduced or alkylated forms of the Lewis acid carbon dioxide;
under some circumstances, they can be decarboxylated to yield carbon dioxide.

4. Distinguish by structure an alkane, alkene, alkyne, cycloalkane, halogenated carbon,


and aromatic hydrocarbon.
Alkanes
Alkanes, or saturated hydrocarbons, contain only single covalent bonds between carbon atoms.
Each of the carbon atoms in an alkane has sp3 hybrid orbitals and is bonded to four other
atoms, each of which is either carbon or hydrogen. The Lewis structures and models of
methane, ethane, and pentane are illustrated in Figure 22.2.122.2.1. Carbon chains are usually
drawn as straight lines in Lewis structures, but one has to remember that Lewis structures are
not intended to indicate the geometry of molecules. Notice that the carbon atoms in the
structural models (the ball-and-stick and space-filling models) of the pentane molecule do not lie
in a straight line. Because of the sp3 hybridization, the bond angles in carbon chains are close
to 109.5°, giving such chains in an alkane a zigzag shape.
The structures of alkanes and other organic molecules may also be represented in a less
detailed manner by condensed structural formulas (or simply, condensed formulas). Instead of
the usual format for chemical formulas in which each element symbol appears just once, a
condensed formula is written to suggest the bonding in the molecule. These formulas have the
appearance of a Lewis structure from which most or all of the bond symbols have been
removed. Condensed structural formulas for ethane and pentane are shown at the bottom of
Figure 22.2.122.2.1, and several additional examples are provided in the exercises at the end of
this chapter.

Pictured are the Lewis structures, ball-and-stick models, and space-filling models for molecules
of methane, ethane, and pentane.
Alkenes
Organic compounds that contain one or more double or triple bonds between carbon atoms are
described as unsaturated. You have likely heard of unsaturated fats. These are complex organic
molecules with long chains of carbon atoms, which contain at least one double bond between
carbon atoms. Unsaturated hydrocarbon molecules that contain one or more double bonds are
called alkenes. Carbon atoms linked by a double bond are bound together by two bonds, one σ
bond, and one π bond. Double and triple bonds give rise to a different geometry around the
carbon atom that participates in them, leading to important differences in molecular shape and
properties. The differing geometries are responsible for the different properties of unsaturated
versus saturated fats.
Ethene, C2H4, is the simplest alkene. Each carbon atom in ethene, commonly called ethylene,
has a trigonal planar structure. The second member of the series is propene (propylene)
(Figure 22.2.622.2.6); the butene isomers follow in the series. Four carbon atoms in the chain of
butene allow for the formation of isomers based on the position of the double bond, as well as a
new form of isomerism.
Expanded structures, ball-and-stick structures, and space-filling models for the alkenes ethene,
propene, and 1-butene are shown.
Ethylene (the common industrial name for ethene) is a basic raw material in the production of
polyethylene and other important compounds. Over 135 million tons of ethylene were produced
worldwide in 2010 for use in the polymer, petrochemical, and plastic industries. Ethylene is
produced industrially in a process called cracking, in which the long hydrocarbon chains in a
petroleum mixture are broken into smaller molecules.
Alkynes
Hydrocarbon molecules with one or more triple bonds are called alkynes; they make up another
series of unsaturated hydrocarbons. Two carbon atoms joined by a triple bond are bound
together by one σ bond and two π bonds. The sp-hybridized carbons involved in the triple bond
have bond angles of 180°, giving these types of bonds a linear, rod-like shape.
The simplest member of the alkyne series is ethyne, C2H2, commonly called acetylene. The
Lewis structure for ethyne, a linear molecule, is:

The IUPAC nomenclature for alkynes is similar to that for alkenes except that the suffix -one is
used to indicate a triple bond in the chain. For example, CH3CH2C≡CHCH3CH2C≡CH is called
1-butyne.
Aromatic Hydrocarbons
Benzene, C6H6, is the simplest member of a large family of hydrocarbons, called aromatic
hydrocarbons. These compounds contain ring structures and exhibit bonding that must be
described using the resonance hybrid concept of valence bond theory or the delocalization
concept of molecular orbital theory. (To review these concepts, refer to the earlier chapters on
chemical bonding). The resonance structures for benzene, C6H6, are:

Valence bond theory describes the benzene molecule and other planar aromatic hydrocarbon
molecules as hexagonal rings of sp2-hybridized carbon atoms with the unhybridized p orbital of
each carbon atom perpendicular to the plane of the ring. Three valence electrons in
the sp2 hybrid orbitals of each carbon atom and the valence electron of each hydrogen atom
form the framework of σ bonds in the benzene molecule. The fourth valence electron of each
carbon atom is shared with an adjacent carbon atom in their unhybridized p orbitals to yield the
π bonds. Benzene does not, however, exhibit the characteristics typical of an alkene. Each of
the six bonds between its carbon atoms is equivalent and exhibits properties that are
intermediate between those of a C–C single bond and a C=CC=C double bond. To represent
this unique bonding, structural formulas for benzene and its derivatives are typically drawn with
single bonds between the carbon atoms and a circle within the ring as shown below.

This condensed formula shows the unique bonding structure of benzene.


There are many derivatives of benzene. The hydrogen atoms can be replaced by many different
substituents. Aromatic compounds more readily undergo substitution reactions than addition
reactions; replacement of one of the hydrogen atoms with another substituent will leave the
delocalized double bonds intact. The following are typical examples of substituted benzene
derivatives:
Toluene and xylene are important solvents and raw materials in the chemical industry. Styrene
is used to produce the polymer polystyrene.

5. Values Integration: promote cooperation, respect the ideas of others, and practice
awareness and responsibility.
Big Idea: Naming and Writing Formulas of Compounds.
* Importance of naming compounds, you say? That's quite something. To do away with
names. A colorless odorless compound formed by the reaction of two gases, one of
which burns and the other is a supporter of combustion. Got it? Of course, it's water! But
is it? What about hydrogen peroxide? Well! One could argue that we may add further
details to the earlier description to get a better fix on the name. But, then the descriptions
get bigger and bigger and become grossly impractical!

So, we need a name for a compound. Now, we could have named any compound any
fancy name and that's exactly what happened. People named compounds on their
source, uses, or simply on some property of the combining elements or the compound!
However, as the science of chemistry prospered, it became more and more cumbersome
to remember names, particularly in the field of organic chemistry. So, came the practice
of systematic naming! IUPAC recommends the rules for the systematic naming of
compounds. Though we feel comfortable with remembering compounds by their trivial
names, sometimes it is not possible to assign a simple trivial name to a very complicated
compound and this is where systematic naming finds its use.

Essential Question: Why is there a need for Chemical Formulas


* A chemical formula is a way of presenting information about the chemical proportions
of atoms that constitute a particular chemical compound or molecule, using chemical
element symbols, numbers, and sometimes also other symbols, such as parentheses,
dashes, brackets, commas and plus and minus signs.

Writing name and formula wrongly means it is not that element or compound. Zn is the
symbol of Zinc and Sn is the symbol of tin. Exchanging the letter Z with S changes the
story. Similarly, H2O is the formula for water and D20 is the formula for heavy water. So
be careful in learning, remembering, and reproducing chemical names and formulas.sss

You might also like