Vous êtes sur la page 1sur 23

Introduction to Solutions

A solution is a particular type of mixture. Mixtures in chemistry are combinations of different substances where each substance retains its
chemical properties. Generally, mixtures can be separated by non-chemical means such as filtration, heating, or centrifugation.
A solution is a homogeneous mixture, but that's not the full definition. Homogeneous means that the mixture is the same all the way through.
You could take two same-sized samples: one from the bottom and one from the top and they would be identical. Homogeneous mixtures do
not settle out if left to sit undisturbed, whereas a heterogeneous mixture would. Blood is a good example of a heterogeneous mixture.
A solution is a homogeneous mixture where all particles exist as individual molecules or ions. This is the definition of a solution.
By the way, there are homogeneous mixtures where the particle size is much larger than individual molecules. However, the particle size is so
small that the mixture never settles out. Terms such as colloid, sol, and gel are used to identify these mixtures. Even though their study is an
important part of chemistry, the ChemTeam will only focus on solutions.

A solution has two components: the solute and the solvent.


The solvent is the substance in greater amount.
It is usually a liquid, although it does not have to be. It is usually water, but it does not have to be. The ChemTeam will focus on water only
and will leave non-aqueous solvents alone.
The solute is the substance in lesser amount.
It is usually a solid, although it does not have to be. The ChemTeam will focus mostly on solid solutes, but the occasional liquid solute will get
mentioned.
Ya know, I can just see that gleam in your eye. "OK, Mr. hot-shot ChemTeam," you say, "what about a situation where both substances are in
equal amount?" That's easy - the water is the solvent. "But," you say, "it's not water, it's, it's ALCOHOL." No problem - the liquid substance is
the solvent. "But...," you start to say as I ....
You finish the story.

The word concentration refers to how much solute is dissolved.


Dilute means that only a little solute is dissolved and concentrated means a lot is dissolved.
These are NOT numerical type numbers, but they are words you should be familiar with.
There are two major concentration words that are numerical in nature: molarity and molality. Usually molarity is introduced first in beginning
classes, so go there first.
Molarity is a unit of concentration measuring the number of moles of a solute per liter of solution.
As should be clear from its name, molarity involves moles. Boy, does it!
The molarity of a solution is calculated by taking the moles of solute and dividing by the liters of solution.

This is probably easiest to explain with examples.

Example #1: Suppose we had 1.00 mole of sucrose (it's about 342.3 grams) and proceeded to mix it into some water. It would dissolve and
make sugar water. We keep adding water, dissolving and stirring until all the solid was gone. We then made sure that when everything was
well-mixed, there was exactly 1.00 liter of solution.
What would be the molarity of this solution?

The answer is 1.00 mol/L. Notice that both the units of mol and L remain. Neither cancels.
A symbol for mol/L is often used. It is a capital M. So, writing 1.00 M for the answer is the correct way to do it.
Some textbooks make the M using italics and some put in a dash, like this: 1.00-M. When you handwrite it; a block capital M is just fine.
When you say it out loud, say this: "one point oh oh molar." You don't have to say the dash (if it's there). By the way, you sometimes see 1.00
M like this: 1.00-molar. A dash is usually used when you write the word 'molar.'
And never forget this: replace the M with mol/L when you do calculations. The M is the symbol for molarity, the mol/L is the unit used in
calculations.

Example #2: Suppose you had 2.00 moles of solute dissolved into 1.00 L of solution. What's the molarity?

The answer is 2.00 M.


Notice that no mention of a specific substance is mentioned at all. The molarity would be the same. It doesn't matter if it is sucrose, sodium
chloride or any other substance. One mole of sucrose or sodium chloride or anything else contains the same number of chemical units. And
that number is 6.022 x 1023 units, called Avogadro's Number.

Example #3: What is the molarity when 0.75 mol is dissolved in 2.50 L of solution?

1
The answer is 0.300 M.

Now, let's change from using moles to grams. This is much more common. After all, chemists use balances to weigh things and balances give
grams, NOT moles.
Example #4: Suppose you had 58.44 grams of NaCl and you dissolved it in exactly 2.00 L of solution. What would be the molarity of the
solution?
The solution to this problem involves two steps which will eventually be merged into one equation.
Step One: convert grams to moles.
Step Two: divide moles by liters to get molality.
In the above problem, 58.44 grams/mol is the molar mass of NaCl. (There is the term "formula weight" and the term "molecular weight."
There is a technical difference between them that isn't important right now. The term "molar mass" is a moe generic term.) To solve the
problem:
Step One: dividing 58.44 grams by 58.44 grams/mol gives 1.00 mol.
Step Two: dividing 1.00 mol by 2.00 L gives 0.500 mol/L (or 0.500 M).
Comment: remember that sometimes, a book will write out the word "molar," as in 0.500-molar.

Example #5: Calculate the molarity of 25.0 grams of KBr dissolved in 750.0 mL.

Example #6: 80.0 grams of glucose (C6H12O6, mol. wt = 180. g/mol) is dissolved in enough water to make 1.00 L of solution. What is its
molarity?

Notice how the phrase "of solution" keeps showing up. The molarity definition is based on the volume of the solution, NOT the volume of
pure water used. For example, to say this:
"A one molar solution is prepared by adding one mole of solute to one liter of water."
is totally incorrect. It is "one liter of solution" not "one liter of water."
Be careful on this, especially when you get to molality.

The most typical molarity problem looks like this:


What is the molarity of "whatever" grams of "whatever" substance dissolved in "whatever" mL of solution.
To solve it, you convert grams to moles, then divide by the volume, like this:

The two steps just mentioned can be combined into one equation. Notice that moles is part of both equations, so one equation can be
substituted into the other. Let's substitute the first into the second and rearrange just a bit to get this:

2
The M stands for molarity, the V for volume. In the second question, GMW has been substituted for molar mass. GMW stands for gram-
molecular weight, which is a very common synonym for molar mass.

Example #7: When 2.00 grams of KMnO4 (molec. wt = 158.0 g/mol) is dissolved into 100.0 mL of solution, what molarity results?

This next example is the most common type you'll see:


Example #8: How many grams of KMnO4 are needed to make 500.0 mL of a 0.200 M solution?

Example #9: 10.0 g of acetic acid (CH3COOH) is dissolved in 500.0 mL of solution. What molarity results?

Example #10: How many mL of solution will result when 15.0 g of H2SO4 is dissolved to make a 0.200 M solution?

Dilution: Definition and Calculations

To dilute a solution means to add more solvent without the addition of more solute. Of course, the resulting solution is thoroughly mixed so as
to ensure that all parts of the solution are identical.
The fact that the solute amount stays constant allows us to develop calculation techniques.
First, we write:
moles solute before dilution = moles solute after dilution
From rearranging the equation that defines molarity, we know that the moles of solute equals the molarity times the volume. (Calculating the
moles of solute from molarity times volume will be very useful in other areas of chemistry, particularly acid base. Remember that the volume
must be in liters.)
So we can substitute MV (molarity times volume) into the above equation, like this:
M1V1 = M2V2
The "sub one" refers to the situation before dilution and the "sub two" refers to after dilution.
This equation does not have an official name like Boyle's Law, so we will just call it the dilution equation.

Example #1: 53.4 mL of a 1.50 M solution of NaCl is on hand, but you need some 0.800 M solution. How many mL of 0.800 M can you
make?
Solution:
Using the dilution equation, we write:
(1.50 mol/L) (53.4 mL) = (0.800 mol/L) (x)
x = 100. mL
Notice that the volumes need not be converted to liters. Any old volume measurement is fine, just so long as the same one is used on each
side. (However, as mentioned above, if you are calculating how many moles of solute are present, you need to have the volume in liters.)

Example #2: 100.0 mL of 2.500 M KBr solution is on hand. You need 0.5500 M. What is the final volume of solution which results?
Placing the proper values into the dilution equation gives:
(2.500 mol/L) (100.0 mL) = (0.5500 mol/L) (x)
x = 454.5454545 mL (oops, my fingers got stuck typing.) (Bad attempt at humor, really bad!)
x = 454.5 mL
Sometimes the problem might ask how much more water must be added. In this last case, the answer is 454.5 - 100.0 = 354.5 mL.
Go ahead and answer the question, if your teacher asks it, but it is bad technique in the lab to just measure out the "proper" amount of water to
add and then add it. This is because the volumes (the soltion and the diluting water) are not necessarily additive. The only volume of
importance is the final solution's volume. You add enough water to get to the final solution volume without caring how much the actual
volume of water you added is.

Example #3: A stock solution of 1.00 M NaCl is available. How many milliliters are needed to make 100.0 mL of 0.750 M
(0.750 mol/L) (100.0 mL) = (1.00 mol/L) (x)
x = 75.0 mL

3
Example #4: What volume of 0.250 M KCl is needed to make 100.0 mL of 0.100 M solution?
(0.100 mol/L) (100.0 mL) = (0.250 mol/L) (x)
Please go ahead and solve for x.

Example #5: 2.00 L of 0.800 M NaNO3 must be prepared from a solution known to be 1.50 M in concentration. How many mL are required?
(1.50 mol/L) (x) = (0.800 mol/L) (2.00 L)
x = 1.067 L
Divide the liters by 1000 to get mL, the answer is 1070 mL (notice that it is rounded off to three sig figs)

These next two are a bit harder and involve slightly more calculation than the discussion above. Here's a summary of the steps:
1) calculate total moles
2) calculate total volume
3) divide moles by volume to get molarity
You can also think of it this way:
M1V1 + M2V2 = M3V3
Where the '1' refers to one starting solution, '2' to the other starting solution and '3' refers to the mixed solution (hints: V 3 is the total volume
after mixing and M3 is almost always the unknown).

Example #6: Calculate the final concentration if 2.00 L of 3.00 M NaCl and 4.00 L of 1.50 M NaCl are mixed. Assume there is no volume
contraction upon mixing.
Here are the two mole calculations:
x = (3.00 mol/L) (2.00 L)
x = (1.50 mol/L) (4.00 L)
I hope it is obvious that you add the two answers to get the total moles.
The total volume calculation is 2.00 + 4.00 = 6.00 L.
Divide total moles by total volume to get the final answer. The answer is 2.00 M.
Using M1V1 + M2V2 = M3V3, we have this:
(3.00 mol/L) (2.00 L) + (1.50 mol/L) (4.00 L) = (x) (6.00 L)
6.00 mol + 6.00 mol = (x) (6.00 L)
x = 12.0 mol / 6.00 L = 2.00 mol/L

Example #7: Calculate the final concentration if 2.00 L of 3.00 M NaCl, 4.00 L of 1.50 M NaCl and 4.00 L of water are mixed. Assume there
is no volume contraction upon mixing.
The solution to this problem is almost exactly the same as 10a. The only "problem child" appears to be the 4.00 L of water. Hint: the water
contributes to the final volume, but NOT to the total moles. The ChemTeam gets a final answer of 1.20 M in this problem.
Using M1V1 + M2V2 = M3V3, we have this:
(3.00 mol/L) (2.00 L) + (1.50 mol/L) (4.00 L) = (x) (10.0 L)
6.00 mol + 6.00 mol = (x) (10.0 L)
General Rules of Solubility

The dividing line between soluble and insoluble is 0.1-molar at 25 C.


Any substance that can form 0.1 M or more concentrated is soluble. Any substance that fails to reach 0.1 M is defined to be insoluble.
This value was picked with a purpose. VERY FEW substances have their maximum solubility near to 0.1 M. Almost every substance of any
importance in chemistry is either much MORE soluble or much LESS soluble.
In the past, some teachers would have a third category: slightly soluble. For the most part, that category has been cast aside.

Solubility rules that apply to water solution:


(1) All alkali metal (lithium, sodium, potassium, rubidium, and cesium) and ammonium compounds are soluble.
(2) All acetate, perchlorate, chlorate, and nitrate compounds are soluble.
(3) Silver, lead, and mercury(I) compounds are insoluble.
(4) Chlorides, bromides, and iodides are soluble
(5) Carbonates, hydroxides, oxides, phosphates, silicates, and sulfides are insoluble. (6) Sulfates are soluble except for calcium and barium.
These rules are to be applied in the order given. For example, PbSO4 is insoluble because rule 3 comes before rule 6. In like manner, AgCl is
insoluble because rule 3 (the smaller) takes precedence over rule 4 (the larger).
Please be aware that these rules are guidelines only. For example, there are some alkali metal compounds that are insoluble. However, these
are rather exotic compounds and can be safely ignored at an introductory level.

Calculations Using Molarity


There are several types of calculations that you need to be able to do with molarity.
First, you should be able to calculate the molarity if you are given the components of the solution.
Second, you should be able to calculate the amount of solute in (or needed to make) a certain volume of solution.
Third, you might need to calculate the volume of a particular solution sample.
Fourth, you might need to calculate the concentration of a solution made by the dilution of another solution. This and
4
related calculations will be covered in a separate page.
In either of the first two cases, the amount of solute might be in moles or grams and the amount of solution might be in liters
or milliliters. Please note that with molarity we are concerned with how much solute there is and with how much solution
there is, but not with how much solvent there is.
Examples (Ex. 3)
The examples that follow are also shown in example 3 in your workbook. You may want to look through the workbook
examples on your own and if they make perfect sense to you, test your understanding by doing exercise 4. Then check your
answers below before continuing on to dilution calculations.
General Relationship (Ex. 3a)
Here is the general relationship that you will be using over and moles of solute
molarity =
over again. The molarity is equal to the number of moles of liter of solution
solute divided by the volume of the solution measured in
y moles
liters. If you like to think of numbers and units instead of xM=
zL
quantities look at the second version of the equation. In this
equation x, y and z represent numbers: 2, 6 and 3 for 6 moles
2M=
example. 3L
Calculating Molarity from Moles and Volume (Ex. 3b)
What is the molarity of a solution containing 0.32 moles of
Here we are given something to figure out. To get the NaCl in 3.4 liters?
molarity we need to divide the number of moles of NaCl by
the volume of the solution. In this case that is 0.32 moles NaCl molarity = 0.32 moles NaCl
divided by 3.4 L, and that gives 0.094 M NaCl. 3.4 L
= 0.094 M NaCl
Calculating Molarity from Mass and Volume (Ex. 3c)
This one is a bit more difficult. To get molarity we still need to
divide moles of solute by volume of solution. But this time What is the molarity of a solution made by dissolving 2.5 g of
we're not given the moles of solute. We have to calculate it NaCl in enough water to make 125 ml of solution?
from the mass of NaCl. We multiply 2.5 g NaCl by the moles of solute
conversion factor of 1 mole NaCl over the formula weight of molarity =
liter of solution
NaCl, 58.5 g. That tells us that we have 0.0427 mole of NaCl. I
kept an extra digit here because we are not done with the 1 mole NaCl
2.5 g NaCl x = 0.0427 mole
calculations. When we are done I'll round off to two digits, the 58.5 g NaCl
same as in the 2.5 g weight of NaCl. Now that we know the 0.0427 mole NaCl
moles we can calculate the molarity. Moles of solute (0.0427) molarity = 0.125 L
divided by the volume of the solution (0.125 L) gives us = 0.34 M NaCl
0.34 M NaCl.
Calculating Mass of Solute from Molarity (Ex. 3d)
This question asks how you would prepare 400. ml of 1.20 M solution of sodium chloride. In this case what you need to find
out is how much NaCl would have to be dissolved in 400 ml to give the concentration that is specified. This amount is going to
have to be in grams because we don't have any balances that weigh in moles. So there is more than one step to this problem.
How would you prepare 400. ml of 1.20 M solution of sodium
chloride?
The approach shown here is a conversion factor approach. It
involves remembering that molarity is a relationship between Remember:
moles and liters. 1.20 M NaCl means there is 1.2 moles of NaCl 1.20 moles NaCl
1.20 M NaCl =
per 1.00 liter of solution. We can use that as a conversion 1.00 L solution
factor to set up the calculation that relates 400. ml (or .400 L)
1.20 moles NaCl
to the appropriate number of moles of NaCl. So we take .400 L 0.400 L solution x
1.00 L solution
and multiply by the conversion factor to get .480 moles NaCl.
= 0.480 moles NaCl
The next step is to find out how many grams that is. We
change from moles of NaCl to grams by using the formula 58.5 g NaCl
weight. It comes out to 28.1 g NaCl. So the answer is that you 0.480 moles NaCl x
1 mole NaCl
would make the solution by dissolving 28.1 g NaCl in enough = 28.1 g NaCl
water to make 400 ml of solution.
Dissolve 28.1 g NaCl in enough water to make 400 mL of
solution.
There is also more than one way to do this problem. If you like Algebraic approach:
the algebraic approach, you would write down the general
equation shown in part a, substitute in the known values, Set up general equation:
solve for moles of NaCl, and then change that into grams. y moles
xM=
zL
Substitute known values:
5
y moles
1.20 M =
0.400 L
Solve for moles:
y moles = 1.20 M x 0.400 L = 0.480 moles
Change to grams:
0.480 mole x 58.5 g/mole = 28.1 g

Calculating Moles of Solute from Molarity (Ex. 3e)


This question is a little easier. We do it the same way as the
first step of the previous problem and then we stop. To find
out how many moles of salt are contained in 300. ml of a How many moles of salt are contained
0.40 M NaCl solution, we start with the volume in liters (0.300 in 300. mL of a 0.40 M NaCl solution?
L) and multiply it by the number of moles per liter of solution,
which is 0.40 moles over 1.00 L. The answer is 0.12 moles of 0.40 moles NaCl
0.300 L x
NaCl. This could also have been done using algebra by writing 1.00 L solution
down the general equation relating molarity, moles and liters, = 0.12 moles NaCl
substituting the known values, and then solving the equation
for moles.
Practice (Ex. 4)
Now you should take some time to review the calculations above (ex. 3). If you have any questions, check with your
instructor. Once you are familiar with how those are done then you should try answering the following questions (exercise 4 in
your workbook). Get help from the instructor if you need it. Check your answers below before you continue with the lesson.
Molarity Calculations: Practice
How would you prepare 100. mL of 0.25 M KNO3 solution?
A chemist dissolves 98.4 g of FeSO4 in enough water to make 2.000 L of solution. What is the molarity of the solution?
How many moles of KBr are in 25.0 mL of a 1.23 M KBr solution?
Battery acid is generally 3 M H2SO4. Roughly how many grams of H2SO4 are in 400. mL of this solution?

Answers (Ex. 4)
Here are the answers to exercise 4.
How would you prepare 100. mL of 0.25 M KNO3 solution? Dissolve 2.53 g of KNO3 in enough water to make 100 ml of
solution.
A chemist dissolves 98.4 g of FeSO4 in enough water to make 2.000 L of solution. What is the molarity of the
solution? 0.324 M
How many moles of KBr are in 25.0 mL of a 1.23 M KBr solution?
0.0308 mol
Battery acid is generally 3 M H2SO4. Roughly how many grams of H2SO4 are in 400. mL of this solution? 120 g

Calculating Molarity

Home

The properties and behavior of many solutions depend not only on the nature of the solute and solvent but also
on the concentration of the solute in the solution. Chemists use many different units when expressing
concentration; however, one of the most common units is molarity. Molarity (M) is the concentration of a solution
expressed as the number of moles of solute per liter of solution:
Molarity (M) = moles solute
liters solution
For example, a 0.25 M NaOH solution (this is read as 0.25 molar) contains 0.25 moles of sodium hydroxide in
every liter of solution. Anytime you see the abbreviation M you should immediately think of it as mol/L.
In order to calculate the molarity of a solution, you need to know the number of moles of solute and the total
volume of the solution.
To calculate molarity:
Calculate the number of moles of solute present.
Calculate the number of liters of solution present.
Divide the number of moles of solute by the number of liters of solution.
Instead of calculating the moles of solute and liters of solution present individually, you can also string all the
calculations together in one problem:
Divide the amount of solute by the volume of solution (regardless of the initial units given).
Use dimensional analysis to convert the amount of solute to moles of solute.
6
Use dimensional analysis to convert the volume of solution to liters of solution (if necessary).
The following example will illustrate both methods.
Example: What is the molarity of a solution prepared by dissolving 15.0 g of sodium hydroxide in enough water to
make a total of 225 mL of solution?
Method 1:
Calculate the number of moles of solute present.
mol NaOH = 15.0g NaOH x 1 mol NaOH
40.0 g NaOH
mol NaOH = 0.375 mol NaOH

Calculate the number of liters of solution present.


L soln = 225 mL x 1L = 0.225 L soln
1000 mL
Divide the number of moles of solute by the number of liters of solution.
M = 0.375 mol NaOH = 1.67 M NaOH
0.225 L soln
Method 2:
Divide the amount of solute by the volume of solution (regardless of the initial units given).
M = 15.0 g NaOH
225 mL soln
Use dimensional analysis to convert the amount of solute to moles of solute.
M = 15.0 g NaOH x 1 mol NaOH
225 mL soln 40.0 g NaOH
Use dimensional analysis to convert the volume of solution to liters of solution (if necessary).
M = 15.0 g NaOH x 1 mol NaOH x 1000 mL soln
225 mL soln 40.0 g NaOH 1 L soln
M
= 1.67 mol NaOH = 1.67 M NaOH
L soln
As you can see, both methods give exactly the same result. Choose whichever method is most comfortable for
you. Just remember that ultimately your units must be mol/L (= M).

Chem 1115
1. What is the molarity of a solution that contains 1.724 moles of H 2SO4 in 2.50 L
of solution?

Calculate the number of moles of solute present.


mol H2SO4 = 1.724 mol (given) Chem 1215

Calculate the number of liters of solution present.


L soln = 2.50 L (given)

Divide the number of moles of solute by the number of liters of solution. Tutorial
List
M = 1.724 mol H2SO4 = 0.690 M H2SO4
2.50 L soln

2. What is the molarity of a solution prepared by dissolving 25.0 g of HCl (g) in


enough water to make 150.0 mL of solution?

M = 25.0 g HCl x 1 mol HCl x 1000 mL soln


150.0 mL soln 36.5 g HCl 1 L soln
M = 4.57 M HCl 7
Molarity is defined as the number of moles of solute per liter of solution. This means that if you
have a 1 M solution of some compound, evaporating one liter will cause one mole of the solute to
precipitate.
Molality is defined as the number of moles of solute per kilogram of solvent. To make a 1 m solution,
you'd take one mole of a substance and add it to 1 L of solvent. As a result, the final volume of a 1
m solution will be somewhat more than 1 L.

Molarity is a unit of concentration measuring the number of moles of a solute per liter of solution. The strategy
to solving molarityproblems is fairly simple. This outlines a straightforward method to calculate molarity of a
solution.

The key to calculating molarityis to remember the units of molarity: moles per liter. Find the number of moles of
the solute dissolved in liters of solution.
Sample Molarity Calculation
Take the following example:
Calculate the molarity of a solution prepared by dissolving 23.7 grams of KMnO 4 into enough water to make 750
mL of solution.

This example has neither moles or liters needed to find molarity. Find the number of moles of the solute first.
To convert grams to moles, the molar mass of the solute is needed. From the periodic table:

Molar mass of K = 39.1 g


Molar mass of Mn = 54.9 g
Molar mass of O = 16.0 g

Molar mass of KMnO4 = 39.1 g + 54.9 g + (16.0 g x 4)


Molar mass of KMnO4 = 158.0 g

Molarity Definition
In chemistry, molarity is a concentration unit, defined to be the number of moles of solute divided by the number
of liters of solution.
Units of Molarity
Molarity is expressed in units of moles per liter (mol/L). It's such a common unit, it has its own symbol, which is a
capital letter M. A solution that has the concentration 5 mol/L would be called a 5 M solution or said to have a
concentration value of 5 molar.
Molarity Examples
There are 6 moles of HCl in one liter of 6 molar HCl or 6 M HCl.
There are 0.05 moles of NaCl in 500 ml of a 0.1 M NaCl solution. (The calculation of moles of ions depends on
their solubility.)
There are 0.1 moles of Na+ ions in one liter of a 0.1 M NaCl solution (aqueous).
Example Problem
Express the concentration of a solution of 1.2 grams of KCl in 250 ml of water.
In order to solve the problem, you need to convert the values into the units of molarity, which are moles and
liters. Start by converting grams of potassium chloride (KCl) into moles.
To do this, look up the atomic masses of the elements on the periodic table. The atomic mass is the mass in
grams of 1 mole of atoms.
mass of K = 39,10 g/mol
mass of Cl = 35.45 g/mol
So, the mass of one mole of KCl is:
mass of KCl = mass of K + mass of Cl
mass of KCl = 39.10 g + 35.45 g
mass of KCl = 74.55 g/mol
You have 1.2 grams of KCl, so you need to find how many moles that is:
moles KCl = (1.2 g KCl)(1 mol/74.55 g)
moles KCl = 0.0161 mol
Now, you know how many moles of solute are present. Next, you need to convert the volume of solvent (water)
from ml to L. Remember, there are 1000 milliliters in 1 liter:
liters of water = (250 ml)(1 L/1000 ml)
liters of water = 0.25 L
Finally, you're ready to determine molarity. Simply express the concentration of KCl in water in terms of moles
solute (KCl) per liters of solute (water):

8
molarity of solution = mol KC/L water
molarity = 0.0161 mol KCl/0.25 L water
molarity of the solution = 0.0644 M (calculator)
Since you were given mass and volume using 2 significant figures, you should report molarity in 2 sig figs also:
molarity of KCl solution = 0.064 M
Advantages and Disadvantages of Using Molarity
There are two big advantages of using molarity to express concentration. The first advantage is that it's easy and
convenient to use because the solute may be measured in grams, converted into moles, and mixed with a
volume.
The second advantage is that the sum of the molar concentrations is the total molar concentration. This permits
calculations of density and ionic strength.
The big disadvantage of molarity is that it changes according to temperature. This is because the volume of a
liquid is affected by temperature. If measurements are all performed at a single temperature (e.g., room
temperature), this is not a problem. However, it's good practice to report the temperature when citing a molarity
value. When making a solution, keep in mind, molarity will change slightly if you use a hot or cold solvent, yet
store the final solution at a different temperature.

Molarity, Molality and Normality


The quantitative relationship between chemical substances in a reaction is known as stoichiometry. Avogadro
was a pioneer in this field of chemistry. Avogadro hypothesized that there was a specific number that would
represent the number of atoms or molecules in a mole of that atom or molecule. The weight of that unit known
as a mole would be equivalent to the atomic or molecular weight of the atom or molecule in grams. According to
this theory, one mole of carbon-12 would have a mass of 12 grams because carbon-12 has an atomic weight of
12 (6 neutrons and 6 protons). One mole of hydrogen would weigh one gram and would contain the same
number of atoms as one mole of carbon. The magical number was, in fact, discovered to be 6.024E23. It was
named the Avogadro Number in honor of the father of stoichiometry even though he did not actually
determine the exact number. The modern definition of a mole is as follows:
Exactly one mole represents the number of carbon atoms in exactly 12 grams of the carbon-12 isotope.
For an atom a mole represents 6.02E23 atoms. For a molecular compound, one mole represents
6.02E23 molecules. One could even take it to extremes and say that one mole of bananas is equal to
6.02E23 bananas.
Now you might ask, "what relevance does the Avogadro number have for me?" The answer is, just about
everything when you are working in a chemistry lab. On the other hand, it is totally useless as a unit of measure
for bananas.
One atom or even ten atoms are too small for an individual to measure out in a lab. A mole of a substance
equals the gram-formula mass or the gram-molar mass. This equals the sum of all of the masses of all the
elements in the formula of the substance. Basically, if one were to count all of the carbon atoms in one one mole
of carbon-12, there would be 6.02E23 atoms and it would weigh 12 grams (remember the atomic weight of
carbon is 12).
Using the information above, it is possible to calculate concentrations of solutions and make up solutions of
desired concentration. It is also possible to use this information to determine how much of a given base would
be needed in order to neutralize a specific acid and reach a pH of 7.
There are five units of concentration that are particularly useful to chemists. The first three: molality, molarity
and normality are dependant upon the mole unit. The last two: percent by volume and percent by weight have
nothing to do with mole, only weight or volume of the soluteor substance to be diluted, versus the weight or
volume of the solvent or substance in which the solute is diluted. Percentages can also be determined for solids
within solids.
Molality: The molal unit is not used nearly as frequently as the molar unit. A molality is the number of
moles of solute dissolved in one kilogram of solvent. Be careful not to confuse molality and molarity.
Molality is represented by a small "m," whereas molarity is represented by an upper case "M." Note that the
solvent must be weighed unless it is water. One liter of water has a specific gravity of 1.0 and weighs one
kilogram; so one can measure out one liter of water and add the solute to it. Most other solvents have a specific
gravity greater than or less than one. Therefore, one liter of anything other than water is not likely to occupy a
liter of space. To make a one molal aqueous (water) solution of sodium chloride (NaCl) , measure out one
kilogram of water and add one mole of the solute, NaCl to it. The atomic weight of sodium is 23 and the atomic
weight of chlorine is 35. Therefore the formula weight for NaCl is 58, and 58 grams of NaCl dissolved in 1kg
water would result in a 1 molal solution of NaCl.
Molarity: The molar unit is probably the most commonly used chemical unit of measurement. Molarity is the
number of moles of a solute dissolved in a liter of solution. A molar solution of sodium chloride is made by
placing 1 mole of a solute into a 1-liter volumetric flask. (Taking data from the example above we will use 58

9
grams of sodium chloride). Water is then added to the volumetric flask up to the one liter line. The result is a
one molar solution of sodium chloride.
Normality: There is a relationship between normality and molarity. Normality can only be calculated when we
deal with reactions, because normality is a function of equivalents.
The example below uses potassium hydroxide (KOH) to neutralize arsenic acid. By studying the reaction it is
possible to determine the proton exchange number to determine the normality of the arsenic acid.
Look at the equation H3AsO4 + 2KOH --> K2HAsO4 + 2H2O:
Equivalent weight = molar mass/(H+ per mole)
Equivalent = mass of compound / Equivalent weight
And Normality = (equivalents of X)/Liter
And the part that is of interest to you is that Normality = molarity x n (where n = the number of protons
exchanged in a reaction).
You probably remember that when a hydrogen atom is ionized and loses its electron, you are left with only a
proton. So a hydrogen ion is basically a proton.
Let's assume that we have a 0.25 M solution of H3AsO4 and want to determine the normality of it if it participates
in the reaction
H3AsO4 + 2KOH --> K2HAsO4 + 2H2O
When H3AsO4 is neutralized by KOH, H3AsO4 provides two protons to form 2H2O.
Note that H3AsO4 has three hydrogens, but K2HAsO4 only has one hydrogen. That means that 2 protons were
exchanged,
Again normality = molarity * n
Remember that normality of the solution is 0.25 mol H3AsO4 and there were two protons exchanged (2
equivalents/mole)

So, in short, while there is a relationship between the normality of a solution and the molarity of a solution, the
normality can only be determined by examining reaction, determining the proton exchange and multiplying
molarity by that number.
Normality is particularly useful in titrations calculations.
Where N = normality, V = volume, a = the substance on the left of the equation involved in proton exchange,
and b=substance on the right of the equation involved in proton exchange:
NaVa = NbVb
Percentages are easy to calculate because they do not require information about the chemical nature of the
substance. Percentages can be determined as percent by weight or percent by volume.
Percent by weight: To make up a solution based on percentage by weight, one would simply determine what
percentage was desired (for example, a 20% by weight aqueous solution of sodium chloride) and the total
quantity to be prepared.
If the total quantity needed is 1 kg, then it would simply be a matter of calculating 20% of 1 kg which, of course
is:
0.20 NaCl * 1000 g/kg = 200 g NaCl/kg.
In order to bring the total quantity to 1 kg, it would be necessary to add 800g water.
Percent by volume: Solutions based on percent by volume are calculated the same as for percent by weight,
except that calculations are based on volume. Thus one would simply determine what percentage was desired
(for example, a 20% by volume aqueous solution of sodium chloride) and the total quantity to be prepared.
If the total quantity needed is 1 liter, then it would simply be a matter of calculating 20% of 1 liter which, of
course is:
0.20 NaCl * 1000 ml/l = 200 ml NaCl/l.
Percentages are used more in the technological fields of chemistry (such as environmental technologies) than
they are in pure chemistry.

Polarity of bonds[edit]

10
In a molecule of hydrogen fluoride(HF), the more electronegative atom (fluoride) is shown in yellow. Because the electrons spend
more time by the fluorine atom in the HF bond, the red represents partially negatively charged regions, while blue represents
partially positively charged regions.
Not all atoms attract electrons with the same force. The amount of "pull" an atom exerts on its electrons is called
its electronegativity. Atoms with high electronegativities such as fluorine, oxygen and nitrogen exert a greater pull on
electrons than atoms with lower electronegativities. In a bond, this leads to unequal sharing of electrons between the atoms, as
electrons will be drawn closer to the atom with the higher electronegativity.
Because electrons have a negative charge, the unequal sharing of electrons within a bond leads to the formation of an electric
dipole: a separation of positive and negative electric charge. Because the amount of charge separated in such dipoles is usually
smaller than a fundamental charge, they are called partial charges, denoted as + (delta plus) and (delta minus). These
symbols were introduced by Christopher Kelk Ingold and Edith Hilda Ingold in 1926.[1][2] The bond dipole moment is calculated by
multiplying the amount of charge separated and the distance between the charges.
These dipoles within molecules can interact with dipoles in other molecules, creating dipole-dipole intermolecular forces.
Classification[edit]
Bonds can fall between one of two extremes being completely nonpolar or completely polar. A completely nonpolar bond occurs
when the electronegativities are identical and therefore possess a difference of zero. A completely polar bond is more correctly
called an ionic bond, and occurs when the difference between electronegativities is large enough that one atom actually takes an
electron from the other. The terms "polar" and "nonpolar" are usually applied to covalent bonds, that is, bonds where the polarity
is not complete. To determine the polarity of a covalent bond using numerical means, the difference between the electronegativity
of the atoms is used.
Bond polarity is typically divided into three groups that are loosely based on the difference in electronegativity between the two
bonded atoms. According to the Pauling scale:
Nonpolar bonds generally occur when the difference in electronegativity between the two atoms is less than 0.4
Polar bonds generally occur when the difference in electronegativity between the two atoms is roughly between 0.4 and 1.7
Ionic bonds generally occur when the difference in electronegativity between the two atoms is greater than 1.7.
Pauling based this classification scheme on the partial ionic character of a bond, which is an approximate function of the
difference in electronegativity between the two bonded atoms. He estimated that a difference of 1.7 corresponds to 50% ionic
character, so that a greater difference corresponds to a bond which is predominantly ionic. [3]
Polarity of molecules[edit]
See also dipole Molecular dipoles.
While the molecules can be described as "polar covalent", "nonpolar covalent", or "ionic", this is often a relative term, with one
molecule simply being more polar or more nonpolar than another. However, the following properties are typical of such
molecules.
A molecule is composed of one or more chemical bonds between molecular orbitals of different atoms. A molecule may be polar
either as a result of polar bonds due to differences in electronegativity as described above, or as a result of an asymmetric
arrangement of nonpolar covalent bonds and non-bonding pairs of electrons known as a full molecular orbital.
Polar molecules[edit]

The water molecule is made up of oxygen and hydrogen, with respective electronegativities of 3.44 and 2.20. The dipoles from
each of the two bonds (red arrows) add together to make the overall molecule polar.
A polar molecule has a net dipole as a result of the opposing charges (i.e. having partial positive and partial negative charges)
from polar bonds arranged asymmetrically. Water (H2O) is an example of a polar molecule since it has a slight positive charge on
one side and a slight negative charge on the other. The dipoles do not cancel out resulting in a net dipole. Due to the polar nature
of the water molecule itself, polar molecules are generally able to dissolve in water. Other examples include sugars
(like sucrose), which have many polar oxygenhydrogen (OH) groups and are overall highly polar.
If the bond dipole moments of the molecule do not cancel, the molecule is polar. For example, the water molecule (H2O) contains
two polar OH bonds in a bent (nonlinear) geometry. The bond dipole moments do not cancel, so that the molecule forms
a molecular dipole with its negative pole at the oxygen and its positive pole midway between the two hydrogen atoms. In the
figure each bond joins the central O atom with a negative charge (red) to an H atom with a positive charge (blue).

The ammonia molecule, NH3, polar as a result of its molecular geometry. The red represents partially negatively charged regions.
The hydrogen fluoride, HF, molecule is polar by virtue of polar covalent bonds in the covalent bond electrons are displaced
toward the more electronegative fluorine atom. Ammonia, NH3, molecule the three NH bonds have only a slight polarity (toward
the more electronegative nitrogen atom). The molecule has two lone electrons in an orbital, that points towards the fourth apex of
the approximate tetrahedron, (VSEPR). This orbital is not participating in covalent bonding; it is electron-rich, which results in a
powerful dipole across the whole ammonia molecule.

11
In ozone (O3) molecules, the two OO bonds are nonpolar (there is no electronegativity difference between atoms of the same
element). However, the distribution of other electrons is uneven since the central atom has to share electrons with two other
atoms, but each of the outer atoms has to share electrons with only one other atom, the central atom is more deprived of
electrons than the others (the central atom has a formal charge of +1, while the outer atoms each have a formal charge of 12).
Since the molecule has a bent geometry, the result is a dipole across the whole ozone molecule.
When comparing a polar and nonpolar molecule with similar molar masses, the polar molecule in general has a higher boiling
point, because the dipoledipole interaction between polar molecules results in stronger intermolecular attractions. One common
form of polar interaction is the hydrogen bond, which is also known as the H-bond. For example, water forms H-bonds and has a
molar mass M = 18 and a boiling point of +100 C, compared to nonpolar methane with M = 16 and a boiling point of 161 C.
Nonpolar molecules[edit]
A molecule may be nonpolar either when there is an equal sharing of electrons between the two atoms of a diatomic molecule or
because of the symmetrical arrangement of polar bonds in a more complex molecule. For example, boron trifluoride (BF3) has a
trigonal planar arrangement of three polar bonds at 120. This results in no overall dipole in the molecule.

In a molecule of boron trifluoride, the trigonal planar arrangement of three polar bonds results in no overall dipole.

Carbon dioxide has two polar C-O bonds in a linear geometry.


Not every molecule with polar bonds is a polar molecule. Carbon dioxide (CO2) has two polar C=O bonds, but the geometry of
CO2 is linear so that the two bond dipole moments cancel and there is no net molecular dipole moment; the molecule is nonpolar.

In methane, the bonds are arranged symmetrically (in a tetrahedral arrangement) so there is no overall dipole.
Examples of household nonpolar compounds include fats, oil, and petrol/gasoline. Therefore, most nonpolar molecules are
water-insoluble (hydrophobic) at room temperature. Many nonpolar organic solvents, such as turpentine, are able to dissolve
polar substances.
In the methane molecule (CH4) the four CH bonds are arranged tetrahedrally around the carbon atom. Each bond has polarity
(though not very strong). However, the bonds are arranged symmetrically so there is no overall dipole in the molecule. The
diatomic oxygenmolecule (O2) does not have polarity in the covalen
Formal charges are theoretical, polarity is a practical value.
Formal charges are the way to describe where electrons should be in a very simple electron shell model. As usual, the truth is
more complex than that, and the actual location of the electrons is often less extreme.
Polarity is a physical property of compounds which relate to other physical properties, such as melting and boiling points or
solubility. Bond polarities arise from bonds between atoms of different electronegativity. A molecule can be ionic, polar or non-
polar
A polar molecule is formed when one end of a molecule has a positive charge and the opposite end has a negative charge, thus
creating electrical poles. A non-polar molecule does not have charges at the ends as the electrons are distributed more
symmetrically and cancel each other out.
When trying to create a solution, a polar molecule does not mix with a non-polar molecule. An example of this is seen with water,
a polar molecule and oil, anon-polar molecule; the two molecules cannot mix to form a solution. However, water and alcohol can
mix to form a solution as they are both polar molecules.
In chemistry, polarity is a separation of electric charge leading to a molecule or its chemical groups having an
electric dipole or multipole moment.
Polar molecules must contain polar bonds due to a difference in electronegativity between the bonded atoms. A polar molecule
with two or more polar bonds must have an asymmetric geometry so that the bond dipolesdo not cancel each other.
Polar molecules interact through dipoledipole intermolecular forces and hydrogen bonds. Polarity underlies a number of
physical properties including surface tension, solubility, and melting and boiling points.
Molecular Polarity
When EXAMPLE Predicting Molecular Polarity:
there are Decide whether the molecules represented by the following formulas are polar or nonpolar.
12
no polar (You may need to draw Lewis structures and geometric sketches to do so.)
bonds in a. CO2 b. OF2 c. CCl4 d. CH2Cl2 e. HCN
a Solution:
molecule a. The Lewis structure for CO2 is
, there is
no
permane The electronegativities of carbon and oxygen are 2.55 and 3.44. The 0.89 difference in
nt electronegativity indicates that the C-O bonds are polar, but the symmetrical arrangement of
charge these bonds makes the molecule nonpolar.
differenc If we put arrows into the geometric sketch for CO2, we see that they exactly balance each
e other, in both direction and magnitude. This shows the symmetry of the bonds.
between
one part
of the b. The Lewis structure for OF2 is
molecule
and The electronegativities of oxygen and fluorine, 3.44 and 3.98, respectively, produce a 0.54
another, difference that leads us to predict that the O-F bonds are polar. The molecular geometry of
and the OF2 is bent. Such an asymmetrical distribution of polar bonds would produce a polar molecule.
molecule
is
nonpolar.
For
example,
the c. The molecular geometry of CCl4 is tetrahedral. Even though the C-Cl bonds are polar, their
Cl2 mole symmetrical arrangement makes the molecule nonpolar.
cule has
no polar
bonds
because
the
electron
charge is
identical d. The Lewis structure for CH2Cl2 is
on both
atoms. It
is
therefore
a
nonpolar
The electronegativities of hydrogen, carbon, and chlorine are 2.20, 2.55, and 3.16. The 0.35
molecule
difference in electronegativity for the H-C bonds tells us that they are essentially nonpolar. The
. None of
0.61 difference in electronegativity for the C-Cl bonds shows that they are polar. The following
the
geometric sketches show that the polar bonds are asymmetrically arranged, so the molecule is
bonds in
polar. (Notice that the Lewis structure above incorrectly suggests that the bonds are
hydrocar
symmetrically arranged. Keep in mind that Lewis structures often give a false impression of the
bon
geometry of the molecules they represent.)
molecule
s, such
as hexan
e,
C6H14,
are
significa
ntly e. The Lewis structure and geometric sketch for HCN are the same:
polar, so
hydrocar
bons are The electronegativities of hydrogen, carbon, and nitrogen are 2.20, 2.55, and 3.04. The 0.35
nonpolar difference in electronegativity for the H-C bond shows that it is essentially nonpolar. The 0.49
molecula difference in electronegativity for the C-N bond tells us that it is polar. Molecules with one polar
13
r bond are always polar.
substanc
es.
A
molecule
can
possess
polar
bonds
and still
be
nonpolar.
If the
polar
bonds
are
evenly
(or
symmetr
ically)
distribut
ed, the
bond
dipoles
cancel
and do
not
create a
molecula
r dipole.
For
example,
the three
bonds in
a
molecule
of
BF3 are
significa
ntly
polar,
but they
are
symmetr
ically
arranged
around
the
central
boron
atom. No
side of
the
molecule
has more
negative
or
positive
14
charge
than
another
side, and
so the
molecule
is
nonpolar
:

A water
molecule
is polar
because
(1) its O-
H bonds
are
significa
ntly
polar,
and (2)
its bent
geometr
y makes
the
distributi
on of
those
polar
bonds
asymmet
rical. The
side of
the
water
molecule
containin
g the
more
electron
egative
oxygen
atom is
partially
negative
, and the
side of
the
molecule
containin
g the
less
electron
15
egative
hydroge
n atoms
is
partially
positive.

Sample
Study
Sheet:
Predictin
g
Molecula
r Polarity
Tip-off
You are
asked to
predict
whether
a
molecule
is polar
or
nonpolar
; or you
are
asked a
question
that
cannot
be
answere
d unless
you
know
whether
a
molecule
is polar
or
nonpolar.
(For
example,
you are
asked to
predict
the type
of
attractio
n holding
the
particles
together

16
in a
given
liquid or
solid.)
General
Steps -
Step 1:
Draw a
reasonab
le Lewis
structure
for the
substanc
e.
Step 2:
Identify
each
bond as
either
polar or
nonpolar.
(If the
differenc
e in
electron
egativity
for the
atoms in
a bond is
greater
than 0.4,
we
consider
the bond
polar. If
the
differenc
e in
electron
egativity
is less
than 0.4,
the bond
is
essential
ly
nonpolar.
)
If there
are no
polar
bonds,
the
molecule
is
nonpolar.
If the
molecule
17
has polar
bonds,
move on
to Step
3.
Step 3: If
there is
only one
central
atom,
examine
the
electron
groups
around
it.
If there
are no
lone
pairs on
the
central
atom,
and if all
the
bonds to
the
central
atom are
the
same,
the
molecule
is
nonpolar.
(This
shortcut
is
describe
d more
fully in
the
Example
that
follows.)
If the
central
atom has
at least
one
polar
bond
and if
the
groups
bonded
to the
central
18
atom are
not all
identical,
the
molecule
is
probably
polar.
Move on
to Step
4.
Step 4:
Draw a
geometri
c sketch
of the
molecule
.
Step 5:
Determin
e the
symmetr
y of the
molecule
using the
following
steps.
Describe
the polar
bonds
with
arrows
pointing
toward
the more
electron
egative
element.
Use the
length of
the
arrow to
show the
relative
polarities
of the
different
bonds.
(A
greater
differenc
e in
electron
egativity
suggests
a more
polar
bond,
19
which is
describe
d with a
longer
arrow.)
Decide
whether
the
arrange
ment of
arrows is
symmetr
ical or
asymmet
rical
If the
arrange
ment is
symmetr
ical and
the
arrows
are of
equal
length,
the
molecule
is
nonpolar.
If the
arrows
are of
different
lengths,
and if
they do
not
balance
each
other,
the
molecule
is polar.
If the
arrange
ment is
asymmet
rical, the
molecule
is polar.

The polarity of molecules


There are three main properties of chemical bonds that must be considerednamely, their strength, length, and
polarity. The polarity of a bond is the distribution of electrical charge over the atoms joined by the bond.
Specifically, it is found that, while bonds between identical atoms (as in H 2) are electrically uniform in the sense
that both hydrogen atoms are electrically neutral, bonds between atoms of different elements are electrically
inequivalent. In hydrogen chloride, for example, the hydrogen atom is slightly positively charged whereas the

20
chlorine atom is slightly negatively charged. The slight electrical charges on dissimilar atoms are called partial
charges, and the presence of partial charges signifies the occurrence of a polar bond.
The polarity of a bond arises from the relative electronegativities of the elements. Electronegativity, it will be
recalled, is the power of an atom of an element to attract electrons toward itself when it is part of a compound.
Thus, although a bond in a compound may consist of a shared pair of electrons, the atom of the more
electronegative element will draw the shared pair toward itself and thereby acquire a partial negative charge.
The atom that has lost its equal share in the bonding electron pair acquires a partial positive charge because its
nuclear charge is no longer fully canceled by its electrons.

The existence of equal but opposite partial charges on the atoms at each end of a heteronuclear bond (i.e., a
bond between atoms of different elements) gives rise to an electric dipole. The magnitude of this dipole is
expressed by the value of its dipole moment, , which is defined as the product of the magnitude of the partial
charges times their separation (essentially, the length of the bond). The dipole moment of a heteronuclear bond
can be estimated from the electronegativities of the atoms A and B, A and B, respectively, by using the simple
relation

where D denotes the unit debye, which is used for reporting molecular dipole moments (1 D = 3.34
1030coulombmetre). Moreover, the negative end of the dipole lies on the more electronegative atom. If the two
bonded atoms are identical, it follows that the dipole moment is zero and the bond is nonpolar.
As the difference in electronegativity between two covalently bonded atoms increases, the dipolar character of
the bond increases as the partial charges increase. When the electronegativities of the atoms are very different,
the attraction of the more electronegative atom for the shared electron pair is so great that it effectively
exercises complete control over them. That is, it has gained possession of the pair, and the bond is best regarded
as ionic. Ionic and covalent bonding therefore can be regarded as constituting a continuum rather than as alternatives.
This continuum can be expressed in terms of resonance by regarding a bond between atoms A and B as a
resonance between a purely covalent form, in which the electrons are shared equally, and a purely ionic form, in
which the more electronegative atom (B) has total control over the electrons:

As the electronegativity difference increases, the resonance lies increasingly in favour of the ionic contribution.
When the electronegativity difference is very large, as between an electropositive atom like sodium and an
electronegative atom like fluorine, the ionic structure dominates the resonance, and the bonding can be regarded
as ionic. Thus, as the electronegativity difference of the two bonded elements increases, a nonpolar bond gives
way to a polar bond, which in turn becomes an ionic bond. There are, in fact, no purely ionic bonds, just as there
are no purely covalent bonds; bonding is a continuum of types.
Even a homonuclear bond, which is a bond between atoms of the same element (as in Cl2), is not purely covalent,
because a more accurate description would be in terms of ionic-covalent resonance:

That the species is nonpolar despite the occurrence of ionic contributions stems from the equal contributions of
the ionic structures ClCl+and Cl+Cl and their canceling dipoles. That Cl2 is commonly regarded as a covalently
bonded species stems from the dominant contribution of the structure ClCl to this resonance mixture. In
contrast, the valence bond theory wavefunction (see below The quantum mechanics of bonding: Valence bond theory)
of hydrogen chloride would be expressed as the resonance hybrid

In this case, the two ionic structures would contribute different amounts (because the elements have different
electronegativities), and the larger contribution of H+Cl is responsible for the presence of partial charges on the
atoms and the polarity of the molecule.
A polyatomic molecule will have polar bonds if its atoms are not identical. However, whether or not the molecule as
a whole is polar (i.e., has a nonzero electric dipole moment) depends on the shape of the molecule. For example,
the carbon-oxygen bonds in carbon dioxide are both polar, with the partial positive charge on the carbon atom and
the partial negative charge on the more electronegative oxygen atom. The molecule as a whole is nonpolar,
however, because the dipole moment of one carbon-oxygen bond cancels the dipole moment of the other, for
the two bond dipole moments point in opposite directions in this linear molecule. In contrast, the water molecule
is polar. Each oxygen-hydrogen bond is polar, with the oxygen atom bearing the partial negative charge and the
hydrogen atom the partial positive charge. Because the molecule is angular rather than linear, the bond dipole
moments do not cancel, and the molecule has a nonzero dipole moment.
The polarity of H2O is of profound importance for the properties of water. It is partly responsible for the existence
of water as a liquid at room temperature and for the ability of water to act as a solvent for many ionic compounds.
The latter ability stems from the fact that the partial negative charge on the oxygen atom can emulate the
21
negative charge of anions that surround each cation in the solid and thus help minimize the energy difference
when the crystal dissolves. The partial positive charge on the hydrogen atoms can likewise emulate that of the
cations surrounding the anions in the solid.

What factors affect solubility?


QUICK ANSWER
The temperature, the polarity of the solutes and solvent, the pressure, and the molecular size affect solubility. All of these factors
play roles in determining which solutes dissolve in which solvents.
FULL ANSWER
Temperature affects solubility by changing the properties of the solvent. As a general rule, as the temperature rises, the solubility
rises. Polarity affects solubility because polar solvents dissolve polar solutes, and nonpolar solvents dissolve nonpolar solutes.
Pressure only affects solubility for gases; the solubility of a gas is proportional to the pressure of the gas solvent. Molecular size
affects solubility because larger molecules are more difficult to dissolve in solvents. Generally, smaller molecules dissolve more
easily and are more soluble.
Electric dipole moment
Electromagnetism

In physics, the electric dipole moment is a measure of the separation of positive and negative electrical charges within a
system, that is, a measure of the system's overall polarity. The electric field strength of the dipole is proportional to the magnitude
of dipole moment. The SI units for electric dipole moment are Coulomb-meter (C m), however the most commonly used unit is
the Debye (D).
Theoretically, an electric dipole is defined by the first-order term of the multipole expansion, and consists of two equal and
opposite charges infinitely close together. This is unrealistic, as real dipoles have separated charge. [1]However, because the
charge separation is very small compared to everyday lengths, the error introduced by treating real dipoles like they are
theoretically perfect is usually negligible. The direction of dipole is usually defined from the negative charge towards the positive
charge.
Intermolecular force
From Wikipedia, the free encyclopedia
Intermolecular forces (IMFs) are forces of attraction or repulsion which act between neighboring particles (atoms, molecules, or
ions). They are weak compared to the intramolecular forces, the forces which keep a molecule together. For example
the covalent bond, involving the sharing of electron pairs between atoms is much stronger than the forces present between the
neighboring molecules. They are an essential part of force fields frequently used in molecular mechanics.
The investigation of intermolecular forces starts from macroscopic observations which point out the existence and action of
forces at a molecular level. These observations include non-ideal-gas thermodynamic behavior reflected by virial
coefficients, vapor pressure, viscosity, superficial tension and absorption data.
The first reference to the nature of microscopic forces is found in Alexis Clairaut's work Theorie de la Figure de la Terre.[1]Other
scientists who have contributed to the investigation of microscopic forces include: Laplace, Gauss, Maxwell and Boltzmann.
Attractive intermolecular forces are considered by the following types:
Ion-induced dipole forces
Ion-dipole forces
van der Waals forces (Keesom force, Debye force, and London dispersion force)
Information on intermolecular force is obtained by macroscopic measurements of properties like viscosity, PVT data. The link to
microscopic aspects is given by virial coefficients and Lennard-Jones potentials.
A hydrogen bond is the electrostatic attraction between two polar groups that occurs when a hydrogen (H) atom covalently
bound to a highly electronegativeatom such as nitrogen (N), oxygen (O), or fluorine (F) experiences the electrostatic field of
another highly electronegative atom nearby.
Hydrogen bonds can occur between molecules (intermolecular) or within different parts of a single molecule (intramolecular).
[4]
Depending on geometry and environment, the hydrogen bond free energy content is between 1 and 5 kcal/mol. This makes it
stronger than a van der Waals interaction, but weaker than covalentor ionic bonds. This type of bond can occur in inorganic
molecules such as water and in organic molecules like DNA and proteins.
Intermolecular hydrogen bonding is responsible for the high boiling point of water(100 C) compared to the other group 16
hydrides that have much weaker hydrogen bonds.[5] Intramolecular hydrogen bonding is partly responsible for
the secondary and tertiary structures of proteins and nucleic acids. It also plays an important role in the structure of polymers,
both synthetic and natural.
In 2011, an IUPAC Task Group recommended a modern evidence-based definition of hydrogen bonding, which was published in
the IUPAC journal Pure and Applied Chemistry. This definition specifies:
The hydrogen bond is an attractive interaction between a hydrogen atom from a molecule or a molecular fragment XH in which
X is more electronegative than H, and an atom or a group of atoms in the same or a different molecule, in which there is
evidence of bond formation.
22
Surface tension is the elastic tendency of a fluid surface which makes it acquire the least surface area possible. Surface tension
allows insects (e.g. water striders), usually denser than water, to float and stride on a water surface.
At liquid-air interfaces, surface tension results from the greater attraction of liquid molecules to each other (due to cohesion) than
to the molecules in the air (due to adhesion). The net effect is an inward force at its surface that causes the liquid to behave as if
its surface were covered with a stretched elastic membrane. Thus, the surface becomes under tension from the imbalanced
forces, which is probably where the term "surface tension" came from. [1] Because of the relatively high attraction of water
molecules for each other through a web of hydrogen bonds, water has a higher surface tension (72.8 millinewtons per meter at
20 C) compared to that of most other liquids. Surface tension is an important factor in the phenomenon of capillarity.
Surface tension has the dimension of force per unit length, or of energy per unit area. The two are equivalent, but when referring
to energy per unit of area, it is common to use the term surface energy, which is a more general term in the sense that it applies
also to solids.
In materials science, surface tension is used for either surface stress or surface free energy.
Solubility is the property of a solid, liquid, or gaseous chemical substance called solute to dissolve in a solid, liquid, or
gaseous solvent. The solubility of a substance fundamentally depends on the physical and chemical properties of the solute and
solvent as well as on temperature, pressure and the pH of the solution. The extent of the solubility of a substance in a specific
solvent is measured as the saturation concentration, where adding more solute does not increase the concentration of the
solution and begins to precipitate the excess amount of solute. The solubility of a substance is an entirely different property from
the rate of solution, which is how fast it dissolves.
Most often, the solvent is a liquid, which can be a pure substance or a mixture. One may also speak of solid solution, but rarely of
solution in a gas (see vaporliquid equilibrium instead).
The extent of solubility ranges widely, from infinitely soluble (without limit) (fully miscible[1]) such as ethanol in water, to poorly
soluble, such as silver chloride in water. The term insoluble is often applied to poorly or very poorly soluble compounds. A
common threshold to describe something as insoluble is less than 0.1 g per 100 mL of solvent. [2]
Under certain conditions, the equilibrium solubility can be exceeded to give a so-called supersaturated solution, which
is metastable.[3] Metastability of crystals can also lead to apparent differences in the amount of a chemical that dissolves
depending on its crystalline form or particle size. A supersaturated solution generally crystallises when 'seed' crystals are
introduced and rapid equilibration occurs. Phenylsalicylate is one such simple observable substance when fully melted and then
cooled below its fusion point.
Solubility is not to be confused with the ability to 'dissolve' a substance, because the solution might also occur because of a
chemical reaction. For example, zinc 'dissolves' (with effervescence) in hydrochloric acid as a result of a chemical reaction
releasing hydrogen gas in a displacement reaction. The zinc ions are soluble in the acid.
The smaller a particle is, the faster it dissolves although there are many factors to add to this generalization.
Crucially solubility applies to all areas of chemistry, geochemistry, inorganic, physical, organic and biochemistry. In all cases it will
depend on the physical conditions (temperature, pressure and concentration) and the enthalpy and entropy directly relating to the
solvents and solutes concerned. By far the most common solvent in chemistry is water which is a solvent for most ionic
compounds as well as a wide range of organic substances. This is a crucial factor in acidity/alkalinity and much environmental
and geochemical work.

23

Vous aimerez peut-être aussi